Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-wzw2p Total loading time: 0 Render date: 2024-05-02T11:51:38.242Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  26 January 2017

Martin P. A. Jackson
Affiliation:
University of Texas, Austin
Michael R. Hudec
Affiliation:
University of Texas, Austin
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
Salt Tectonics
Principles and Practice
, pp. 464 - 494
Publisher: Cambridge University Press
Print publication year: 2017

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Aali, A., Abar, A., Boenke, N., Pollard, M., Ruhli, F., and Stollner, T., 2012, Ancient salt mining and salt men: The interdisciplinary Chehrabad Douzlakh project in north-western Iran: Antiquity, 86, Issue 333, unpaginated.Google Scholar
Aftabi, P., Roustaie, M., Alsop, I. A., and Talbot, C. J., 2010, InSAR mapping and modelling of an active Iranian salt extrusion: Journal of the Geological Society, London, 167, 155170, doi:10.1144/0016–76492008-165.Google Scholar
Ahlborn, O., and Richter-Bernburg, G., 1955, Exkursion zum Salzstock von Benthe (Hannover), mit Befahrung der Kaliwerke Ronnenberg und Hansa: Zeitschrift der Deutschen Geologischen Gesellschaft, 105, 855865.CrossRefGoogle Scholar
Albertz, M., and Ings, S. J., 2012, Some consequences of mechanical stratification in basin-scale numerical models of passive-margin salt tectonics, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 303330, doi:10.1144/SP363.14.Google Scholar
Alexander, R. J., Bjerstedt, T. W., and Moate, S. L., 2005, Edge-Sigsbee folds, Gulf of Mexico, USA, in Shaw, J. H., Connors, C., and Suppe, J., eds., Seismic interpretation of contractional fault-related folds: An AAPG seismic atlas: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 53, 133–134.Google Scholar
Al-Marjeby, A., and Nash, D., 1986, A summary of the geology and oil habitat of the Eastern Flank hydrocarbon province of South Oman: Marine and Petroleum Geology, 3, 306314.Google Scholar
Alsop, G. I., Brown, J. P., Davison, I., and Gibling, M. R., 2000, The geometry of drag zones adjacent to salt diapirs: Journal of the Geological Society, London, 157, 10191029, doi:10.1144/jgs.157.5.1019.Google Scholar
Alsop, G. I., Weinberger, R., Levi, T., and Marco, S., 2015, Deformation within an exposed salt wall: Recumbent folding and extrusion of evaporites in the Dead Sea Basin: Journal of Structural Geology, 70, 95118, doi:10.1016/j.jsg.2014.11.006.Google Scholar
Al-Zoubi, A., Shulman, H., and Ben-Avraham, Z., 2002, Seismic reflection profiles across the southern Dead Sea basin: Tectonophysics, 346, 6169.Google Scholar
Al-Zoubi, A., and ten Brink, U., 2001, Salt diapirs in the Dead Sea basin and their relationship to Quaternary extensional tectonics: Marine and Petroleum Geology, 18, 779797.Google Scholar
Amery, G. B., 1969, Structure of Sigsbee Scarp, Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 53, 24802482.Google Scholar
Anderson, J. E., Cartwright, J., Drysdall, S. J., and Vivian, N., 2000, Controls on turbidite sand deposition during gravity-driven extension of a passive margin: Examples from Miocene sediments in block 4, Angola: Marine and Petroleum Geology, 17, 11651203.Google Scholar
Anderson, R. Y., 1981, Deep-seated salt dissolution in the Delaware Basin, Texas and New Mexico, in Wells, S. G., Lambert, W., and Callender, J. F., eds., Environmental Geology and Hydrology in New Mexico: Socorro, NM, New Mexico Geological Society, Special Publication 10, 133145.Google Scholar
Anderson, R. Y., and Kirkland, D. W., 1980, Dissolution of salt deposits by brine density flow: Geology, 8, 6669.2.0.CO;2>CrossRefGoogle Scholar
Atwater, G. I., and Forman, M. J., 1959, Nature of growth of southern Louisiana salt domes and its effect on petroleum accumulation: American Association of Petroleum Geologists Bulletin, 43, 25922622.Google Scholar
Autin, W. J., 2002, Landscape evolution of the Five Islands of south Louisiana: Scientific policy and salt dome utilization and management: Geomorphology, 47, 227244.Google Scholar
Aydin, A., 2000, Fractures, faults, and hydrocarbon entrapment, migration and flow: Marine and Petroleum Geology, 17, 797814.Google Scholar
Baar, C. A., 1977, Applied salt-rock mechanics; 1, The in-situ behavior of salt rocks: Amsterdam, Elsevier, Developments in Geotechnical Engineering, 16A, 294 p.Google Scholar
Bąbel, M., and Schreiber, B. C., 2014, Geochemistry of evaporites and evolution of seawater, in Mackenzie, F., ed., Treatise on geochemistry (2nd ed.), 9. Sediments, diagenesis, and sedimentary rocks: Amsterdam, Elsevier, Chapter 9.17, 484548.Google Scholar
Bahroudi, A., and Koyi, H. A., 2003, Effect of spatial distribution of Hormuz salt on deformation style in the Zagros fold and thrust belt; An analogue modeling approach: Journal of the Geological Society, London, 160, 719733.CrossRefGoogle Scholar
Bailey, E. B., 1931, Salt plugs: Geological Magazine, 68, 335336.Google Scholar
Baker, A. A., 1933, Geology and oil possibilities of the Moab district, Grand and San Juan Counties, Utah: U.S. Geological Survey Bulletin, 841, 95 p.Google Scholar
Baldschuhn, R., Best, G., and Kockel, F., 1991, Inversion tectonics in the north-west German basin, in Spencer, A. M., ed., Generation, accumulation, and production of Europe’s hydrocarbons: Houten, European Association of Petroleum Geoscientists, Special Publication 1, 149159.Google Scholar
Baldschuhn, R., Binot, F., Frisch, U., and Kockel, F., 2001, Geotektonischer Atlas von Nordwest-Deutschland und dem deutschen Nordsee-Sektor [Tectonic Atlas of Northwest Germany and the German North Sea Sector]: Geologisches Jahrbuch, A 153, 3 CD-ROM, 88 p.Google Scholar
Baldwin, B., and Butler, C. O., 1985, Compaction curves: American Association of Petroleum Geologists Bulletin, 69, 622626.Google Scholar
Balk, R., 1949, Structure of Grand Saline salt dome, Van Zandt County, Texas: American Association of Petroleum Geologists Bulletin, 33, 17911829.Google Scholar
Balk, R., 1953, Salt structure of Jefferson Island salt dome, Iberia and Vermilion Parishes, Louisiana: American Association of Petroleum Geologists Bulletin, 37, 24552474.Google Scholar
Bally, A. W., 1981, Thoughts on the tectonics of folded belts, in McClay, K. R. and Price, N. J., eds., Thrust and nappe tectonics: London, Geological Society, Special Publication 9, 1332.Google Scholar
Bally, A. W., 1982, Musings over sedimentary basin evolution, in Kent, P., Bott, M. H. P., McKenzie, D. P., and Williams, C. A., eds., The evolution of sedimentary basins: London, Royal Society, 325338.Google Scholar
Barberini, V., Burlini, L., Rutter, E. H., and Dapiaggi, M., 2005, High-strain deformation tests on natural gypsum aggregates in torsion, in Bruhn, D. and Burlini, L., eds., High-strain zones: Structure and physical properties: London, Geological Society, Special Publication 245, 277290.Google Scholar
Barnes, H. A., Hutton, J. F., and Walters, K., 1989, An introduction to rheology: Amsterdam, Elsevier, Rheology Series 3, 199 p.Google Scholar
Barnhoorn, A., Bystricky, M., Kunze, K., Burlini, L., and Burg, J.-P., 2005, Strain localisation in bimineralic rocks: Experimental deformation of synthetic calcite–anhydrite aggregates: Earth and Planetary Science Letters, 240, 748763.CrossRefGoogle Scholar
Barron, K., and Toews, N. A., 1964, Deformation around a mine shaft in salt: Proceedings of rock mechanics symposium, Queen’s University, Kingston, Ontario, December 1963, 115–136.Google Scholar
Barton, D. C., 1933, Mechanics of formation of salt domes with special reference to Gulf Coast salt domes of Texas and Louisiana: American Association of Petroleum Geologists Bulletin, 17, 10251083.Google Scholar
Basile, C., and Brun, J. P., 1999, Transtensional faulting patterns ranging from pull-apart basins to transform continental margins: An experimental investigation: Journal of Structural Geology, 21, 2337.CrossRefGoogle Scholar
Baud, R. D., and Haglund, J. L., 1996, Enhanced subsalt exploration utilizing the basal salt shear model: Gulf Coast Association of Geological Societies Transactions, 46, 914.Google Scholar
Baumann, W., 1984, Rheologische Untersuchungen an Gips: Eclogae Geologicae Helvetiae, 77, 301325.Google Scholar
Beloussov, V. V., 1959, Types of folding and their origin: International Geology Review, 1, 121.CrossRefGoogle Scholar
Ben-Avraham, Z., 1978, The structure and tectonic setting of the Levant continental margin, eastern Mediterranean: Tectonophysics, 46, 313331.Google Scholar
Ben-Avraham, Z., Garfunkel, Z., and Lazar, M., 2008, Geology and evolution of the southern Dead Sea fault with emphasis on subsurface structure: Annual Review of Earth and Planetary Sciences, 36, 357387, doi:10.1146/annurev.earth.36.031207.124201.CrossRefGoogle Scholar
Bertoni, C., and Cartwright, J. A., 2005, 3D seismic analysis of circular evaporite dissolution structures, Eastern Mediterranean: Journal of the Geological Society, London, 162, 909926.CrossRefGoogle Scholar
Bertoni, C., and Cartwright, J. A., 2007, Major erosion at the end of the Messinian salinity crisis: Evidence from the Levant Basin, Eastern Mediterranean: Basin Research, 19, 118, doi:10.1111/j.1365–2117.2006.00309.x.Google Scholar
Bertoni, C., and Cartwright, J., 2015, Messinian evaporites and fluid flow: Marine and Petroleum Geology, 1–12, doi:10.1016/j.marpetgeo.2015.02.003.Google Scholar
Billingsley, L. E., 1982, Geometry and mechanisms of folding related to growth faulting in Nordheim Field area (Wilcox), De Witt County, Texas: Gulf Coast Association of Geological Societies Transactions, 32, 263274.Google Scholar
Bishop, D. J., Buchanan, P. G., and Bishop, C. J., 1995, Gravity-driven thin-skinned extension above Zechstein Group evaporites in the western central North Sea: An application of computer-aided section restoration techniques: Marine and Petroleum Geology, 12, 115135.CrossRefGoogle Scholar
Bitterli, T., 1990, The kinematic evolution of a classical Jura fold: A reinterpretation based on 3-dimensional balancing techniques (Weissenstein anticline, Jura Mountains, Switzerland): Eclogae Geologicae Helvetiae, 83, 493511.Google Scholar
Bonini, M., 1998, Chronology of deformation and analogue modelling of the Plio–Pleistocene Tiber Basin: Implications for the evolution of the Northern Apennines (Italy): Tectonophysics, 285, 147165.Google Scholar
Bonini, M., 2003, Detachment folding, fold amplification, and diapirism in thrust wedge experiments: Tectonics, 22, 1065, 28 p., doi:10.1029/2002TC001458.CrossRefGoogle Scholar
Bonini, M., Sani, F., and Antonielli, B., 2012, Basin inversion and contractional reactivation of inherited normal faults: A review based on previous and new experimental models: Tectonophysics, 522–523, 5588, doi:10.1016/j.tecto.2011.11.014.Google Scholar
Booth, J. R., Dean, M. C., DuVernay, A. E. III, and Styzen, M. J., 2003, Paleo-bathymetric controls on the stratigraphic architecture and reservoir development of confined fans in the Auger Basin: Central Gulf of Mexico slope: Marine and Petroleum Geology, 20, 563586.CrossRefGoogle 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, in Deep-water reservoirs of the world: Tulsa, OK, Society of Economic Paleontologists and Mineralogists, Gulf Coast Section, 20th annual research conference, 82103.Google Scholar
Borchert, H., 1977, On the formation of Lower Cretaceous potassium salts and tachyhydrite in the Sergipe Basin (Brazil) with some remarks on similar occurrences in West Africa (Gabon, Angola etc.), in Klemm, D. D. and Schneider, H. J., eds., Time and strata-bound ore deposits: Berlin, Springer, 94111.Google Scholar
Borchert, H., and Muir, R. O., 1964, Salt deposits: The origin, metamorphism and deformation of evaporites: London, Van Nostrand Co., 338 p.Google Scholar
Bornhauser, M., 1969, Geology of Day dome (Madison County, Texas) – A study of salt emplacement: American Association of Petroleum Geologists Bulletin, 53, 14111420.Google Scholar
Bosák, P., Jaroš, J., Spudil, J., Sulovský, P., and Václavek, V., 1998, Salt plugs in the eastern Zagros, Iran: Results of regional geological reconnaissance: Geolines, Praha, GlÚ AV ČR Praha, 7, 3180.Google Scholar
Bourbie, T., Coussy, O., and Zinszner, B., 1987, Acoustics of porous media: Houston, TX, Gulf Publishing Company, 334 p.Google Scholar
Bowie, W., 1927, Isostasy – The science of the equilibrium of the Earth’s crust: New York, E. P. Dutton, 275 p.Google Scholar
Boyer, S. E., and Elliott, D., 1982, Thrust systems: American Association of Petroleum Geologists Bulletin, 66, 11961230.Google Scholar
Boyers, W. C., 2000, Structural style and normal faulting adjacent to the Onion Creek salt diapir, Paradox Basin, Utah: M.S. thesis, Baylor University, Waco, Texas.Google Scholar
Brace, W. F., and Kohlstedt, D. L., 1980, Limits on lithospheric stress imposed by laboratory experiments: Journal of Geophysical Research, 85, 62486252.CrossRefGoogle Scholar
Branney, M. J., 1995, Downsag and extension at calderas: New perspectives on collapse geometries from ice-melt, mining and volcanic subsidence: Bulletin of Volcanology, 57, 303318.CrossRefGoogle Scholar
Brauer, V., Eickemeier, R., Eisenburger, D., Grissemann, C., Hesser, J., Heusermann, S., Kaiser, D., Nipp, H.-S., Nowak, T., Plischke, I., Schinier, H., Schulze, O., Sönnke, J., and Weber, J. R., 2011, Description of the Gorleben site, Part 4: Geotechnical exploration of the Gorleben salt dome: Hannover, Bundesanstalt für Geowissenschaften und Rohstoffe, 1184.Google Scholar
Bredehoeft, J. D., 1988, Will salt repositories be dry?: Eos, Transactions, American Geophysical Union, 69, 121131.Google Scholar
Brinkmann, R., and Lögters, H., 1968, Diapirs in western Pyrenees and foreland, Spain, in Braunstein, J., and O’Brien, G. D., eds., Diapirism and diapirs: Tulsa, OK, American Association of Petroleum Geologists Memoir 8, 275292.Google Scholar
Brown, A., 2011 , Interpretation of three-dimensional seismic data (7th ed.): Tulsa, OK, American Association of Petroleum Geologists Memoir 42, and Tulsa, OK, Society of Exploration Geophysicists Investigations in Geophysics 9, 646 p.Google Scholar
Brown, D., and Spadea, P., 1999, Processes of forearc and accretionary complex formation during arc–continent collision in the southern Ural Mountains: Geology, 27, 649652.2.3.CO;2>CrossRefGoogle Scholar
Brudnik, K., Czop, M., Motyka, J., d’Obyrn, K., Rogoż, M., and Witczak, S., 2010, The complex hydrogeology of the unique Wieliczka salt mine: Przegląd Geologiczny, 58, 787796.Google Scholar
Brun, J.-P., and Choukroune, P., 1983, Normal faulting, block tilting, and décollement in a stretched crust: Tectonics, 2, 345356.Google Scholar
Brun, J. P., and Fort, X., 2004, Compressional salt tectonics (Angolan margin): Tectonophysics, 382, 129150, doi:10.1016/j.tecto.2003.11.014.Google Scholar
Brun, J.-P., and Fort, X., 2011, Salt tectonics at passive margins: Geology versus models: Marine and Petroleum Geology, 28, 11231145, doi:10.1016/j.marpetgeo.2011.03.004.CrossRefGoogle Scholar
Brun, J. P., and Fort, X., 2012, Salt tectonics at passive margins: Geology versus models –Reply: Marine and Petroleum Geology, 37, 195208.Google Scholar
Brun, J.-P., and Mauduit, T. P.-O., 2008, Rollovers in salt tectonics: The inadequacy of the listric fault model: Tectonophysics, 457, 111, doi:10.1016/j.tecto.2007.11.038.Google Scholar
Brun, J. P., and Mauduit, T. P.-O., 2009, Salt rollers: Structure and kinematics from analogue modelling: Marine and Petroleum Geology, 26, 249258, doi:10.1016/j.marpetgeo.2008.02.002.s.Google Scholar
Brun, J. P., and Merle, O., 1985, Strain patterns in models of spreading–gliding nappes: Tectonics, 4, 705719.Google Scholar
Brun, J. P., and Nalpas, T., 1996, Graben inversion in nature and experiments: Tectonics, 15, 677687.CrossRefGoogle Scholar
Bruthans, J., Asadi, N., Filippi, M., Vilhelm, Z., and Zare, M., 2008, A study of erosion rates on salt diapir surfaces in the Zagros Mountains, SE Iran: Environmental Geology, 53, 10791089.Google Scholar
Bruthans, J., Filippi, M., Asadi, N., Zare, M., Šlechta, S., and Churáčková, Z., 2009, Surficial deposits on salt diapirs (Zagros Mountains and Persian Gulf Platform, Iran): Characterization, evolution, erosion and the influence on landscape morphology: Geomorphology, 107, 195209, doi:10.1016/j.geomorph.2008.12.006.Google Scholar
Bruthans, J., Filippi, M., Geršl, M., Zare, M., Melková, J., Pazdur, A., and Bosák, P., 2006, Holocene marine terraces on two salt diapirs in Persian Gulf (Iran): Age, depositional history and uplift rates: Journal of Quaternary Science, 21, 843857, doi:10.1002/jqs.1007.Google Scholar
Bruthans, J., Filippi, M., Zare, M., Churáčková, Z., Asadi, N., Fuchs, M., and Adamovič, J., 2010, Evolution of salt diapir and karst morphology during the last glacial cycle: Effects of sea-level oscillation, diapir and regional uplift, and erosion (Persian Gulf, Iran): Geomorphology, 121, 291304, doi:10.1016/j.geomorph.2010.04.026.CrossRefGoogle Scholar
Bruthans, J., Šmíd, J., Filippi, M., and Zeman, O., 2000, Thickness of cap rock and other important factors affecting morphogenesis of salt karst: Acta Carsologica, 29, 5164.Google Scholar
Bryant, W. R., Antoine, J. W., Ewing, M., and Jones, B. R., 1968, Structure of the Mexican continental shelf and slope, Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 52, 12041228.Google Scholar
Bryant, W. R., and Liu, J. Y., 2000, Bathymetry of the Gulf of Mexico: Texas Sea Grant, College Station, Texas, TAMU-SG-00–606, CD-ROM.Google Scholar
Bucher, W. H., 1956, Role of gravity in orogenesis: Geological Society of America Bulletin, 67, 12951318.Google Scholar
Buchner, E., and Kenkmann, T., 2008, Upheaval Dome, Utah, USA: Impact origin confirmed: Geology, 36, 227230, doi:10.1130/G24287A.1.Google Scholar
Buffler, R. T., Worzel, J. L., and Watkins, J. S., 1978, Deformation and origin of the Sigsbee scarp – Lower continental slope, northern Gulf of Mexico: Offshore Technology Conference, May 8–11, 1978, Houston, Texas, paper OTC 3217.Google Scholar
Burollet, P., 1975, Tectonique en radeaux en Angola: Bulletin de la Société Géologique de France, 17, 503504.Google Scholar
Busch, B., 1907, Etwas über die Expansivkraft des Salzes: Zeitschrift für praktische Geologie, 15, 369371.Google Scholar
Busk, H. G., 1929, Earth flexures; Their geometry and their representation and analysis in geological section with special reference to the problem of oil finding: Cambridge, Cambridge University Press, 106 p.Google Scholar
Cailteux, J. L. H., 1983, Le “Roan” shabien dans la région de Kambove (Shaba, Zaire): Ph.D. thesis, Université de Liège, Belgium.Google Scholar
Callot, J. P., Jahani, S., and Letouzey, J., 2007, The role of pre-existing diapirs in fold and thrust belt development, in Thrust belts and foreland basins: Frontiers in earth sciences, Part V: Berlin, Springer, 309–325, doi:10.1007/978-3-540–69426-7_16.Google Scholar
Callot, J., Trocmé, V., Letouzey, J., Albouy, E., Jahani, S., and Sherkati, S., 2009, Pre-existing salt structures and the folding of the Zagros Mountains: Tehran, Geological Society of Iran, 128 p.Google Scholar
Callot, J., Trocmé, V., Letouzey, J., Albouy, E., Jahani, S., and Sherkati, S., 2012, Pre-existing salt structures and the folding of the Zagros Mountains, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 545561, doi:10.1144/SP363.27.Google Scholar
Camerlo, R. H., and Benson, E. F., 2006, Geometric and seismic interpretation of the Perdido fold belt: Northwestern deep-water Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 90, 363386.Google Scholar
Canérot, J., Hudec, M. R., and Rockenbauch, K., 2005, Mesozoic diapirism in the Pyrenean orogen: Salt tectonics on a transform plate boundary: American Association of Petroleum Geologists Bulletin, 89, 211229.Google Scholar
Carey, S. W., 1962, Folding: Alberta Society of Petroleum Geologists Journal, 10, 95144.Google Scholar
Carruthers, D., Cartwright, J., Jackson, M. P. A., and Schutjens, P., 2013, Origin and timing of layer-bound radial faulting around North Sea salt stocks: New insights into the evolving stress state around rising diapirs: Marine and Petroleum Geology, 48, 130148, doi:10.1016/j.marpetgeo.2013.08.001.Google Scholar
Carter, N. L., and Hansen, F. D., 1983, Creep of rocksalt: Tectonophysics, 92, 275333.Google Scholar
Carter, N. L., Horseman, S. T., Russell, J. E., and Handin, J., 1993, Rheology of rocksalt: Journal of Structural Geology, 15, 12571271.Google Scholar
Carter, N. L., Kronenberg, A. K., Ross, J. V., and Wiltschko, D. V., 1990, Control of fluids on deformation of rocks, in Knipe, R. J., ed., Deformation mechanisms, rheology and tectonics: London, Geological Society, Special Publication 54, 113, doi:10.1144/GSL.SP.1990.054.01.01.Google Scholar
Cartwright, J., Huuse, M., and Aplin, A., 2007, Seal bypass systems: American Association of Petroleum Geologists Bulletin, 91, 11411166.Google Scholar
Cartwright, J. A., and Jackson, M. P. A., 2008, Initiation of gravitational collapse of an evaporite basin margin: The Messinian saline giant, Levant Basin, eastern Mediterranean: Geological Society of America Bulletin, 120, 399413, doi:10.1130/B2608IX.1.Google Scholar
Cartwright, J., Jackson, M., Dooley, T., and Higgins, S., 2012, Strain partitioning in gravity-driven shortening of a thick, multilayered evaporite sequence, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 449470.Google Scholar
Cartwright, J. A., and Mansfield, C. S., 1998, Lateral displacement variation and lateral tip geometry of normal faults in the Canyonlands National Park, Utah: Journal of Structural Geology, 20, 319, doi:10.1016/S0191-8141(97)00079–5.CrossRefGoogle Scholar
Cartwright, J., Stewart, S., and Clark, J., 2001, Salt dissolution and salt-related deformation of the Forth Approaches Basin, UK North Sea: Marine and Petroleum Geology, 18, 757778.CrossRefGoogle 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, 13191326, doi:10.1016/0191–8141(95)00033-A.Google Scholar
Casas, E., and Lowenstein, T. K., 1989, Diagenesis of saline pan halite: Comparison of petrographic features of modern, Quaternary and Permian halites: Journal of Sedimentary Petrology, 59, 724739.Google Scholar
Cater, F. W., 1955a, Geology of the Anderson Mesa Quadrangle, Colorado: U.S. Geological Survey Quadrangle Map GQ-77, scale 1:24,000, 1 sheet.Google Scholar
Cater, F. W., 1955b, Geology of the Gypsum Gap Quadrangle, Colorado: U.S. Geological Survey Quadrangle Map GQ-59, scale 1:24,000, 1 sheet.Google Scholar
Cater, F. W., 1955c, Geology of the Hamm Canyon Quadrangle, Colorado: U.S. Geological Survey Quadrangle Map GQ-69, scale 1:24,000, 1 sheet.Google Scholar
Chapple, W. M., 1978, Mechanics of thin-skinned fold-and-thrust belts: Geological Society of America Bulletin, 89, 11891198.Google Scholar
Chen, S., Tang, L., Jin, Z., Jia, C., and Pi, X., 2004, Thrust and fold tectonics and the role of evaporites in deformation in the western Kuqa foreland of Tarim Basin, northwest China: Marine and Petroleum Geology, 21, 10271042, doi:10.1016/j.marpetgeo.2004.01.008.Google Scholar
Chimney, P. J., 2000, Seismic evidence for “heel–toe” structural rotation, 96-G Prospect, Cabinda, offshore Angola (abs.): American Association of Petroleum Geologists Annual Meeting Program, 9, A26A27.Google Scholar
Chimney, P. J., and Kluth, C., 2002, Evidence for low-angle, subhorizontal hanging faults in rotated fault blocks, Cabinda, offshore Angola: The Leading Edge, November, 1084–1090.Google Scholar
Chivas, A. R., Andrew, A. S., Lyons, W. B., Bird, M. I., and Donnelly, T. H., 1991, Isotopic constraints on the origin of salts in Australian playas: I, Sulphur: Palaeogeography, Palaeoclimatology, Palaeoecology, 84, 309322.Google Scholar
Clabaugh, P. S., 1962, Petrofabric study of deformed salt: Science, 136, 389391.Google Scholar
Clark, J., Cartwright, J., and Stewart, S., 1999, Mesozoic dissolution tectonics on the West Central Shelf, UK Central North Sea: Marine and Petroleum Geology, 16, 283300.Google Scholar
Clark, J., Stewart, S., and Cartwright, J., 1998, Evolution of the NW margin of the North Permian Basin, UK North Sea: Journal of the Geological Society, London, 155, 663676.Google Scholar
Clausen, O. R., Egholm, D. L., Andresen, K. J., and Wesenberg, R., 2014, Fault patterns within sediment layers overlying rising salt structures: A numerical modelling approach: Journal of Structural Geology, 58, 6978, doi:10.1016/j.jsg.2013.11.001.Google Scholar
Cloos, E., 1955, Experimental analysis of fracture patterns: Geological Society of America Bulletin, 66, 241256.CrossRefGoogle Scholar
Cloos, E., 1968, Experimental analysis of Gulf Coast fracture patterns: American Association of Petroleum Geologists Bulletin, 52, 420444.Google Scholar
Cloos, H., 1939, Hebung – Spaltung – Vulkanismus – Elemente einer geometrischen Analyse irdischer Großformen: Geologische Rundschau, 30, 405527.Google Scholar
Cluzel, D., 1985, Géologie et métallogénie de la “Série des mines” au Shaba (ex-Katanga) méridional (Zaire), Métaévaporites et reprises hydrothermales: Comptes rendus de l’Académie des Sciences de Paris, Série II, 301, 12091212.Google Scholar
Cobbold, P. R., and Jackson, M. P. A., 1992, Gum rosin (colophony): A suitable material for thermomechanical modeling of the lithosphere: Tectonophysics, 210, 255271.Google Scholar
Cobbold, P., Rossello, E., and Vendeville, B., 1989, Some experiments on interacting sedimentation and deformation above salt horizons: Bulletin de la Société Géologique de France, 5, 453460.Google Scholar
Cobbold, P. R., and Szatmari, P., 1991, Radial gravitational gliding on passive margins: Tectonophysics, 188, 249289.Google Scholar
Cobbold, P. R., Szatmari, P., Demercian, L. S., Coelho, D., and Rossello, E. A., 1995, Seismic and experimental evidence for thin-skinned horizontal shortening by convergent radial gliding on evaporites, deep-water Santos Basin, Brazil, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists Memoir 65, 305321.Google Scholar
Cohen, H. A., and Hardy, S., 1996, Numerical modelling of stratal architectures resulting from differential loading of a mobile substrate, in Alsop, G. I., Blundell, D. J., and Davison, I., eds., Salt tectonics: London, Geological Society, Special Publication 100 265273, doi:10.1144/GSL.SP.1996.100.01.17.Google Scholar
Colman, S. M., 1983, Influence of the Onion Creek salt diapir on the late Cenozoic history of Fisher Valley, southeastern Utah: Geology, 11, 240243.Google Scholar
Coney, P. J., 1987, The regional tectonic setting and possible causes of Cenozoic extension in the North American Cordillera, in Coward, M. P., Dewey, J. F., and Hancock, P.L., eds., Continental extensional tectonics: London, Geological Society, Special Publication 28, 177186, doi:10.1144/GSL.SP.1987.028.01.13.Google Scholar
Cooper, M., 2007, Structural style and hydrocarbon prospectivity in fold and thrust belts: A global review, in Ries, A. C., Butler, R. W. H., and Graham, R. H., eds., Deformation of the continental crust: The legacy of Mike Coward: London, Geological Society, Special Publication 272, 447472.Google Scholar
Cooper, M. A., Williams, G. D., de Graciansky, P. C., Murphy, R. W., Needham, T., de Paor, D., Stoneley, R., Todd, S. P., Turner, J. P., and Ziegler, P. A., 1989, Inversion tectonics – A discussion, in Cooper, M. A. and Williams, G. D., eds., Inversion tectonics: London, Geological Society, Special Publication 44, 335347.Google Scholar
Correa Perez, I., and Gutierrez y Acosta, J., 1983, Interpretación gravimétrica y magnetométrica del occidente de la Cuenca Salina del Istmo: Revista del Instituto Mexicano del Petróleo, 15, 525.Google Scholar
Cosgrove, J. W., Talbot, C. J., and Aftabi, P., 2009, A train of kink folds in the surficial salt of Qom Kuh, central Iran: Journal of Structural Geology, 31, 12121222.Google Scholar
Costa, E., and Vendeville, B., 2001, Diapirism in convergent settings triggered by hinterland pinch-out of viscous decollement: A hypothesis from modeling, in Koyi, H. A. and Mancktelow, N. S., eds., Tectonic modeling: A volume in honor of Hans Ramberg: Boulder, CO, Geological Society of America Memoir 193, 123130.Google Scholar
Cotton, J. T., and Koyi, H. A., 2000, Modeling of thrust fronts above ductile and frictional detachments: Application to structures in the Salt Range and Potwar Plateau, Pakistan: Geological Society of America Bulletin, 112, 351363.Google Scholar
Couzens-Schultz, B. A., Vendeville, B. C., and Wiltschko, D. V., 2003, Duplex style and triangle zone formation: Insights from physical modeling: Journal of Structural Geology, 25, 16231644.Google Scholar
Coward, M. P., 1996, Balancing sections through inverted basins, in Buchanan, P. G. and Nieuwland, D. A., eds., Modern developments in structural interpretation, validation and modelling: London, Geological Society, Special Publication 99, 5177, doi:10.1144/ GSL.SP.1996.099.01.06.Google Scholar
Coward, M. P., Dewey, J. F., Hempton, M., and Holroyd, J., 2003, Tectonic evolution, in Evans, D., Graham, C., Armour, A., and Bathurst, P., eds., The millennium atlas: Petroleum geology of the central and northern North Sea: London, Geological Society, 1733.Google Scholar
Coward, M., and Dietrich, D., 1989, Alpine tectonics – An overview, in Coward, M. P., Dietrich, D., and Park, R. G., eds., Alpine tectonics: London, Geological Society, Special Publication 45, 129.Google Scholar
Coward, M. P., Gillcrist, R., and Trudgill, B., 1991, Extensional structures and their tectonic inversion in the Western Alps, in Roberts, A. M., Yielding, G., and Freeman, B., eds., The geometry of normal faults: London, Geological Society, Special Publication 56, 93112.Google Scholar
Coward, M., and Stewart, S., 1995, Salt-influenced structures in the Mesozoic–Tertiary cover of the southern North Sea, U.K., in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists Memoir 65, 229250.Google Scholar
Craker, W. E., and Schiller, K. K., 1962, Plastic deformation of gypsum: Nature, 4816, 672673.Google Scholar
Cramez, C., and Jackson, M. P. A., 2000, Superposed deformation straddling the continental–oceanic transition in deep-water Angola: Marine and Petroleum Geology, 17, 10951109.CrossRefGoogle Scholar
Cristescu, N. D., 1998, Evolution of damage in rocksalt, in Aubertin, M. and Hardy, H. R., eds., Mechanical behavior of salt: Clausthal-Zellerfeld, Trans Tech Publications, Series on rock and soil mechanics, 22, 131142,Google Scholar
Cristescu, N. D., and Hunsche, U., 1988, Time effects in rock mechanics: New York, John Wiley & Sons, 342 p.Google Scholar
Cristescu, N., and Hunsche, U., 1992, Determination of nonassociated constitutive equation for rock salt from experiments, in Besdo, D. and Stein, E., eds., Finite inelastic deformations – theory and applications: Berlin, Springer, 511523.CrossRefGoogle Scholar
Crosby, W. O., and Crosby, I. B., 1925, Keystone faults: Geological Society of America Bulletin, 36, 623640.CrossRefGoogle Scholar
Dardeau, G., and De Graciansky, P. C., 1990, Halocinèse et rifting téthysien dans les Alpes-maritimes (France): Bulletin des Centres de Recherches Exploration–Production Elf-Aquitaine, 14, 443464.Google Scholar
Dauteuil, O., Deschamps, F., Bourgeois, O., Mocquet, A., and Guillocheau, F., 2013, Post-breakup evolution and palaeotopography of the North Namibian margin during the Meso-Cenozoic: Tectonophysics, 589, 103115.CrossRefGoogle Scholar
Davis, D. M., and Engelder, T., 1985, The role of salt in fold-and-thrust belts: Tectonophysics, 119, 6788.Google Scholar
Davis, G. H., and Reynolds, S. J., 1996, Structural geology of rocks and regions (2nd ed.): New York, Wiley, 776 p.Google Scholar
Davis, W. M., 1903, An excursion to the plateau province of Utah and Arizona: Bulletin of the Museum of Comparative Zoology, Harvard College, Geological Series 6, 50 p.Google Scholar
Davison, I., 2007, Geology and tectonics of the South Atlantic Brazilian salt basins, in Ries, A. C., Butler, R. W. H., and Graham, R. H., eds., Deformation of the continental crust: The legacy of Mike Coward: London, Geological Society, Special Publication 272, 345359.Google Scholar
Davison, I., 2009, Faulting and fluid flow through salt: Journal of the Geological Society, London, 166, 205216.Google Scholar
Davison, I., Alsop, I., Birch, P., Elders, C., Evans, N., Nicholson, H., Rorison, P., Wade, D., Woodward, J., and Young, M., 2000a, Geometry and late-stage structural evolution of Central Graben salt diapirs, North Sea: Marine and Petroleum Geology, 17, 499522.CrossRefGoogle Scholar
Davison, I., Alsop, I., and Blundell, D., 1996a, Salt tectonics: Some aspects of deformation mechanics, in Alsop, G. I., Blundell, D. J., and Davison, I., eds., Salt tectonics: London, Geological Society, Special Publication 100, 110.Google Scholar
Davison, I., Alsop, G. I., Evans, N. G., and Safaricz, M., 2000b, Overburden deformation patterns and mechanisms of salt diapir penetration in the Central Graben, North Sea: Marine and Petroleum Geology, 17, 601618.CrossRefGoogle Scholar
Davison, I., Anderson, L., and Nuttall, P., 2012, Salt deposition, loading and gravity drainage in the Campos and Santos salt basins, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 159174, doi:10.1144/SP363.8.Google Scholar
Davison, I., Bosence, D. W., Alsop, G. I., and Alawi, M., 1996b, Deformation and sedimentation around active Miocene salt diapirs on the Tihama Plain, northwest Yemen, in Alsop, G. I., Blundell, D. J., and Davison, I., eds., Salt tectonics: London, Geological Society, Special Publication 100, 2239.Google Scholar
Davy, P., and Cobbold, P. R., 1991, Experiments on shortening of a 4-layer model of the continental lithosphere: Tectonophysics, 188, 125, doi:10.1016/0040–1951 (91)90311-F.Google Scholar
Day-Stirrat, R. J., McDonnell, A., and Wood, L. J., 2010, Diagenetic and seismic concerns associated with interpretation of deeply buried mobile shales, in Wood, L., ed., Shale tectonics: Tulsa, OK, American Association of Petroleum Geologists Memoir 93, 527, doi:10.1306/13231306M93730.Google Scholar
Dean, M. C., Booth, J. R., and Mitchell, B. T., 2002, Multiple fields within the sequence stratigraphic framework of the greater Auger basin, Gulf of Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 22nd annual research conference proceedings, 661–680.Google Scholar
Dean, W. E., 1978, Trace and minor elements in evaporites, in Dean, W. and Schreiber, B. C., eds., Marine evaporites: Tulsa, OK, SEPM Society for Sedimentary Geology, short course notes 4, 86104.CrossRefGoogle Scholar
De Böckh, H., Lees, G. M., and Richardson, F. D. S., 1929, Contribution to the stratigraphy and tectonics of the Iranian ranges, in Gregory, J. W., ed., The structure of Asia: London, Methuen, 58176.Google Scholar
Debrand-Passard, S., Courbouleix, S., and Lienhardt, M. J., eds., 1984, Synthèse géologique du sud-est de la France: Orléans, Bureau de Recherches Géologiques et Minières, Mémoires du Bureau de Recherches Géologiques et Minières 125.Google Scholar
DeGolyer, E., 1925, Origin of North American salt domes, in Moore, R. C., ed., Geology of salt dome oil fields: Tulsa, OK, American Association of Petroleum Geologists, 144.Google Scholar
De Graciansky, P. C., Dardeau, G., Lemoine, M., and Tricart, P., 1989, The inverted margin of the French Alps and foreland basin inversion, in Cooper, M. A. and Williams, G. D., eds., Inversion tectonics: London, Geological Society, Special Publication 44, 87104.Google Scholar
De Graciansky, P.-C., Rudkiewicz, J.-L., and Samec, P., 1986, Tectonique salifère d’âge Jurassique dans la zone sub Briançonnaise (Alpes de Savoie, France): Rôle dans le découpage en nappes de charriage et leur progression: Comptes Rendus de l’Académie des Sciences, Paris, Série II, 302, 891896.Google Scholar
De Jong, K. A., and Scholten, R., 1973, Gravity and tectonics: New York, John Wiley & Sons, 502 p.Google Scholar
Del Ventisette, C., Montanari, D., Sani, F., and Bonini, M., 2006, Basin inversion and fault reactivation in laboratory experiments, in Tavarnelli, E., Butler, R., and Grasso, M., eds., Tectonic inversion processes and structural inheritance in mountain belts: Journal of Structural Geology, 28, 20672083.Google Scholar
De Magnée, I., and François, A., 1988, The origin of the Kipushi (Cu, Zn, Pb) deposit in direct relation with a Proterozoic salt diapir: Copperbelt of Central Africa, Shaba, Republic of Zaire, in Friedrich, G. H. and Herzig, P. M., eds., Base metal sulfide deposits in sedimentary and volcanic environments: Berlin, Springer, 5, 7493.Google Scholar
De Meer, S., and Spiers, C. J., 1995, Creep of wet gypsum aggregates under hydrostatic loading conditions: Tectonophysics, 245, 171184.Google Scholar
De Meer, S., and Spiers, C. J., 1997, Uniaxial compaction creep of wet gypsum aggregates: Journal of Geophysical Research, 102, 875891.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
De Ruig, M. J., 1992, Tectono-sedimentary evolution of the Prebetic fold belt of Alicante (SE Spain): A study of stress fluctuations and foreland basin deformation: Ph.D. thesis, Vrije Universiteit, Amsterdam, 207 p.Google Scholar
De Ruig, M. J., 1995, Extensional diapirism in the eastern Prebetic foldbelt, southeastern Spain, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 353367.Google Scholar
Desbois, G., Zavada, P., Schléder, Z., and Urai, J. L., 2010, Deformation and recrystallization mechanisms in actively extruding salt fountain: Microstructural evidence for a switch in deformation mechanisms with increased availability of meteoric water and decreased grain size (Qum Kuh, central Iran): Journal of Structural Geology, 32, 115, doi:10.1016/j.jsg.2010.03.005.Google Scholar
Dewey, J. F., 1982, Plate tectonics and the evolution of the British Isles: Journal of the Geological Society, London, 139, 371412.Google Scholar
Diegel, F. A., Karlo, J. F., Schuster, D. C., Shoup, R. C., and Tauvers, P. R., 1995, Cenozoic structural evolution and tectono-stratigraphic framework of the northern Gulf Coast continental margin, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., 1995, Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 109151.Google Scholar
Diegel, F. A., and Schuster, D. C., 1990, Regional cross sections and palinspastic reconstructions, northern Gulf of Mexico (abs.): Boulder, CO, Geological Society of America, Abstracts with Programs, A66.Google Scholar
Dixon, J. M., 1975, Finite strain and progressive deformation in models of diapiric structures: Tectonophysics, 28, 89124.Google Scholar
Dixon, J. M., and Summers, J. M., 1983, Patterns of total and incremental strain in subsiding troughs: Experimental centrifuge models of interdiapir synclines: Canadian Journal of Earth Sciences, 20, 18431861.Google Scholar
Docherty, C., Sugrue, M., Willacy, C., and Strong, A., 1999, Sub-salt imagery and structure recognition using Pre-SDM: World Oil, 220, 8083.Google Scholar
Doelling, H. H., 1985, Geologic map of Arches National Park and vicinity, Grand County, Utah: Utah Geological and Mineral Survey Map 74, scale 1:50,000, 1 sheet, 15 p.Google Scholar
D’Onfro, P., 1988, Mechanics of salt tongue formation with examples from Louisiana slope (abs.): American Association of Petroleum Geologists Bulletin, 72, 175.Google Scholar
Dooley, T. P., Hudec, M. R., and Jackson, M. P. A., 2012, The structure and evolution of sutures in allochthonous salt: American Association of Petroleum Geologists Bulletin, 96, 10451070, doi:10.1306/09231111036.CrossRefGoogle Scholar
Dooley, T. P., Jackson, M. P. A., and Hudec, M. R., 2007, Initiation and growth of salt-based thrust belts on passive margins: Results from physical models: Basin Research, 19, 165177, doi:10.1111/j.1365–2117.2007.00317.x.Google Scholar
Dooley, T. P., Jackson, M. P. A., and Hudec, M. R., 2009, Inflation and deflation of deeply buried salt stocks during lateral shortening: Journal of Structural Geology, 31, 582600, doi:10.1016/j.jsg.2009.03.013.Google Scholar
Dooley, T. P., Jackson, M. P. A., and Hudec, M. R., 2013, Coeval extension and shortening above and below salt canopies on an uplifted, continental margin: Application to the northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 97, 17371764, doi:10.1306/03271312072.Google Scholar
Dooley, T. P., Jackson, M. P. A., Jackson, C. A.-L., Hudec, M. R., and Rodriguez, C. R., 2015, Enigmatic structures within salt walls of the Santos Basin – Part 2: Mechanical explanation from physical modelling: Journal of Structural Geology, 75, 163187, doi:10.1016/j.jsg.2015.01.009.Google Scholar
Dooley, T., and McClay, K., 1997, Analog modeling of pull-apart basins: American Association of Petroleum Geologists Bulletin, 81, 1804.Google Scholar
Dooley, T., McClay, K. R., Hempton, M., and Smit, D., 2005, Salt tectonics above complex basement extensional fault systems: Results from analogue modelling, in Dore, A. G. and Vining, B. A., eds., Petroleum geology: North-west Europe and global perspectives – Proceedings of the 6th Petroleum Geology Conference: London, Petroleum Geology Conferences Ltd. and the Geological Society, 16311648.Google Scholar
Dooley, T., McClay, K. R., and Pascoe, R., 2003, 3D analogue models of variable displacement extensional faults: Applications to the Revfallet fault system, offshore mid-Norway, in Nieuwland, D. A., ed., New insights into structural interpretation and modelling: London, Geological Society, Special Publication 212, 151167.Google Scholar
Dooley, T., Monastero, F., Hall, B., McClay, K., and Whitehouse, P., 2004, Scaled sandbox modeling of transtensional pull-apart basins – Applications to the Coso geothermal system (ext. abs.): Geothermal Resources Council Transactions, 28, 637641.Google Scholar
Dooley, T. P., and Schreurs, G., 2012, Analogue modelling of intraplate strike–slip tectonics: A review and new experimental results: Tectonophysics, 574–575, 171, doi:10.1016/j.tecto.2012.05.030.Google Scholar
Dow, W. G., 1974, Application of oil correlation and source rock data to exploration in Williston basin: American Association of Petroleum Geologists Bulletin, 58, 12531262.Google Scholar
Dubois, A., Odonne, F., Massonnat, G., Lebourg, T., and Fabre, R., 2002, Analogue modelling of fault reactivation: Tectonic inversion and oblique remobilisation of grabens: Journal of Structural Geology, 24, 17411752.CrossRefGoogle Scholar
Durham, W. B., Abey, A. E., and Trimmer, D. A., 1981, Thermal conductivity, diffusivity, and expansion of Avery Island salt at pressure and temperature: Livermore, CA, Lawrence Livermore Laboratory, report UCRL-83789, 113.Google Scholar
Dutton, D. M., and Trudgill, B. D., 2009, Four-dimensional analysis of the Sembo relay system, offshore Angola: Implications for fault growth in salt-detached settings: American Association of Petroleum Geologists Bulletin, 93, 763794.CrossRefGoogle Scholar
Duval, B., Cramez, C., and Jackson, M. P. A., 1990, Raft tectonics in the Kwanza basin, Angola (abs.): Boulder, CO, Geological Society of America, Abstracts with Programs, 22, A48.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.Google Scholar
Duval, B., Cramez, C., Schultz-Ela, D. D., and Jackson, M. P. A., 1993, Extension, reactive diapirism, salt welding and contraction at Cegonha, Kwanza basin, Angola (ext. abs.): American Association of Petroleum Geologists International Hedberg Research Conference, Abstract Volume: Tulsa, OK, American Association of Petroleum Geologists, 4143.Google Scholar
Dyson, I. A., 1998, The ‘Christmas tree diapir’ and salt glacier at Pinda Springs, central Flinders Ranges: MESA Journal, 10, 4043.Google Scholar
Eckardt, F. D., and Spiro, B., 1999, The origin of sulphur in gypsum and dissolved sulphate in the central Namib Desert, Namibia: Sedimentary Geology, 123, 255273.Google Scholar
Edwards, U., Moore, V., and Odumah, U., 2014, Encased secondary minibasins: An emerging play in the deepwater Gulf of Mexico: American Association of Petroleum Geologists Search and Discovery Article 30359, posted August 29, 2014.Google Scholar
Ehgartner, B., Neal, J., and Hinkelbein, T., 1998, Gas releases from salt: Albuquerque, NM, Sandia National Laboratories, report SAND98-1354, 37 p.Google Scholar
Eichenseer, H. T., Walgenwitz, F. R., and Biondi, P. J., 1999, Stratigraphic control on facies and diagenesis of dolomitized oolitic siliciclastic ramp sequences (Pinda Group, Albian, offshore Angola): American Association of Petroleum Geologists Bulletin, 83, 17291758.Google Scholar
Eisenstadt, G., and Sims, D., 2005, Evaluating sand and clay models: do rheological differences matter?: Journal of Structural Geology, 27, 13991412, doi:10.1016/j.jsg.2005.04.010.Google Scholar
Eisenstadt, G., and Withjack, M. O., 1995, Estimating inversion: Results from clay models: London, Geological Society, Special Publication 88, 119–136.Google Scholar
Ellenblum, R., Marco, S., Agnon, A., Rockwell, T., and Boas, A., 1998, Crusader castle torn apart by earthquake at dawn, 20 May 1202: Geology, 26, 303.Google Scholar
Ellisor, A. C., 1926, Coral reefs in the Oligocene of Texas: American Association of Petroleum Geologists Bulletin, 10, 976985.Google Scholar
Elsley, G. R., and Tieman, H., 2010, A comparison of prestack depth and prestack time imaging of the Paktoa complex, Canadian Beaufort MacKenzie basin, in Wood, L., ed., Shale tectonics: Tulsa, OK, American Association of Petroleum Geologists, Memoir 93, 7990, doi:10.1306/13231309M933419.Google Scholar
Enos, J. S., and Kyle, J. R., 2002, Diagenesis of the Carrizo Sandstone at Butler salt dome, East Texas basin, U.S.A.: Evidence for fluid–sediment interaction near halokinetic structures: Journal of Sedimentary Research, 72, 6881.Google Scholar
Erratt, D., 1993, Relationships between basement faulting, salt withdrawal and Late Jurassic rifting, UK Central North Sea, in Parker, J. R., ed., Petroleum geology of northwest Europe: Proceedings of the 4th conference, London, Geological Society, 12111219.Google Scholar
Escher, B. G., and Kuenen, P. H., 1929, Experiments in connection with salt domes: Leidse Geologische Mededelingen, 3, 151182.Google Scholar
Evans, R., and Kirkland, D. W., 1988, Evaporitic environments as a source of petroleum, in Schrieber, B. C., ed., Evaporites and hydrocarbons: New York, Columbia University Press, 256299.Google Scholar
Ewing, M., and Antoine, J., 1966, New seismic data concerning sediments and diapiric structures in Sigsbee Deep and upper continental slope, Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 50, 479504.Google Scholar
Ewing, M., and Ewing, J., 1962, Rate of salt-dome growth: American Association of Petroleum Geologists Bulletin, 46, 708709.Google Scholar
Faugère, E., and Brun, J.-P., 1984, Modélisation experimentale de la distention continentale: Comptes rendus de l’Académie des Sciences, Série II, 299, 365370.Google Scholar
Faulkner, D. R., Jackson, C. A. L., Lunn, R. J., Schlische, R. W., Shipton, Z. K., Wibberley, C. A. J., and Withjack, M. O., 2010, A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones: Journal of Structural Geology, 32, 15571575.CrossRefGoogle Scholar
Feely, H. W., and Kulp, J. L., 1957, Origin of Gulf Coast salt dome sulfur deposits: American Association of Petroleum Geologists Bulletin, 41, 18021853.Google Scholar
Ferrer, O., Jackson, M. P. A., Roca, E., and Rubinat, M., 2012, Evolution of salt structures during extension and inversion of the offshore Parentis Basin (Eastern Bay of Biscay), in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 359377.Google Scholar
Fiduk, J. C., Anderson, L. E., and Rowan, M. G., 2004, The Wilcox raft: An example of extensional raft tectonics in South Texas, northwestern onshore Gulf of Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 25th annual research conference, 294–314.Google Scholar
Fiduk, J. C., Clippard, M., Power, S., Robertson, V., Rodriguez, L., Ajose, O., Fernandez, D., and Smith, D., 2014, Origin, transportation, and deformation of Mesozoic carbonate rafts in the northern Gulf of Mexico: Gulf Coast Association of Geological Societies Journal, 3, 2032.Google Scholar
Fiduk, J. C., and Rowan, M. G., 2012, Analysis of folding and deformation within layered evaporites in Blocks BM-S-8 & -9, Santos Basin, Brazil, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 471487, doi:10.1144/SP363.22.Google Scholar
Fischer, M., Woodward, N., and Mitchell, M., 1992, The kinematics of break–thrust folds: Journal of Structural Geology, 14, 451460.CrossRefGoogle Scholar
Flemings, P. B., Stump, B. B., Finkbeiner, T., and Zoback, M., 2002, Flow focusing in overpressured sandstones: Theory, observations, and applications: American Journal of Science, 302, 827855, doi:10.2475/ajs.302.10.827.Google Scholar
Fletcher, R. C., Hudec, M. R., and Watson, I. A., 1995, Salt glacier and composite salt–sediment models for the emplacement and early burial of allochthonous salt sheets, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 77108.Google Scholar
Fokker, P. A., Urai, J. L., and Steeneken, P. V., 1995, Production-induced convergence of solution mined caverns in magnesium salts and associated subsidence, in F. B. J. Barends, F. J. J. Brouwer, and F. H. Schröder, eds., Land subsidence: Natural causes, measuring techniques, the Groningen gas field: Proceedings of the Fifth International Conference on Land Subsidence, Den Haag, Netherlands, 281–289.Google Scholar
Forrest, M. C., 1986, Deepwater Gulf of Mexico exploration geology, hydrocarbons and economics: Gulf Coast Association of Geological Societies Transactions, 36, xlvxlvii.Google Scholar
Fort, X., Brun, J.-P., and Chauvel, F., 2004, Salt tectonics on the Angolan margin, synsedimentary deformation processes: American Association of Petroleum Geologists Bulletin, 88, 15231544.Google Scholar
Fossen, H., 2010, Structural geology: Cambridge, Cambridge University Press, 463 p.CrossRefGoogle Scholar
Fossum, A. F., and Fredrich, J. T., 2002, Salt mechanics primer for near-salt and sub-salt deepwater Gulf of Mexico field developments: Albuquerque, NM, Sandia National Laboratories, report SAND2002-2063, 67 p.Google Scholar
Fox, F. G., 1959, Structure and accumulation of hydrocarbons in Southern Foothills, Alberta, Canada: American Association of Petroleum Geologists Bulletin, 43, 9921025.Google Scholar
François, A., 1973, L’extrémité occidentale de l’arc cuprifère shabien, étude géologique: Brussels, Département Géologique de la Gécamines, Likasi, Zaire, 81 pages, 17 figures, 31 plates, 91 photographs, 1:200,000 scale geologic map.Google Scholar
François, A., and Lepersonne, J., 1978, Carte géologique de la région de Kolwezi-Kalukundi (Shaba) République du Zaïre: Tervuren, Musée royal de l’Afrique centrale; authorized by the Département Géologique de la Gécamines, Zaire: 1:100,000 scale.Google Scholar
Fraser, S. I., Robinson, A. M., Johnson, H. D., Underhill, J. R., Kadolsky, D. G. A., Connell, R., Johannessen, P., and Ravnås, R., 2003, Upper Jurassic, in Evans, D., Graham, C., Armour, A., and Bathurst, P., eds., The millennium atlas: Petroleum geology of the central and northern North Sea: London, Geological Society, 157189.Google Scholar
Freitas, J. T. R., 2006, Ciclos deposicionais evaporíticos de Bacia de Santos: Uma análise cicloestrátigrafica a partir de dados de 2 poços e de traços de sísmica: Masters dissertation, Porto Alegre, Universidade Federal do Rio Grande do Sul, 160 p.Google Scholar
Frumkin, A., 1994, Hydrology and denudation rates of halite karst: Journal of Hydrology, 162, 171189.CrossRefGoogle Scholar
Frumkin, A., 1996a, Determining the exposure age of a karst landscape: Quaternary Research, 46, 99106.Google Scholar
Frumkin, A., 1996b, Uplift rate relative to base-levels of a salt diapir (Dead Sea basin, Israel) as indicated by cave levels, in Alsop, G. I., Blundell, D. J., and Davison, I., eds., Salt tectonics: London, Geological Society, Special Publication 100, 4147, doi:10.1144/GSL.SP.1996.100.01.04.Google Scholar
Frumkin, A., 2009, Formation and dating of a salt pillar in Mount Sedom diapir, Israel: Geological Society of America Bulletin, 121, 960, doi:10.1130/B26376.1.Google Scholar
Fuchs, L., Koyi, H., and Schmeling, H., 2015, Numerical modeling of the effect of composite rheology on internal deformation in down-built diapirs: Tectonophysics, 646, 7995.Google Scholar
Furuya, M., Mueller, K., and Wahr, J., 2007, Active salt tectonics in the Needles District, Canyonlands (Utah) as detected by interferometric synthetic aperture radar and point target analysis: 1992–2002: Journal of Geophysical Research, 112, B06418 doi:10.1029/2006JB004302.CrossRefGoogle Scholar
Gaffin, S., 1987, Ridge volume dependence on seafloor generation rate and inversion using long term sealevel change: American Journal of Science, 287, 596611.Google Scholar
Galloway, W. E., Ewing, T. E., Garrett, C. M., Tyler, N., and Bebout, D. G., 1983, Atlas of major Texas oil reservoirs: Austin, TX, Bureau of Economic Geology, The University of Texas at Austin, 139 p.Google Scholar
Gansser, A., 1992, The enigma of the Persian salt dome inclusions: Eclogae Geologicae Helvetiae, 85, 825846.Google Scholar
Garrison, J. M., and McMillan, N. J., 1999, Evidence for Jurassic continental rift magmatism in northeast Mexico: Allogenic meta-igneous blocks in El Papalote diapir, La Popa basin, Nuevo Leon, Mexico, in Bartolini, C., Wilson, J. L., and Lawton, T. F., eds., Mesozoic sedimentary and tectonic history of north-central Mexico: Boulder, CO, Geological Society of America, Special Paper 340, 319332.Google Scholar
Gartrell, A., Hudson, C., and Evans, B., 2005, The influence of basement faults during extension and oblique inversion of the Makassar Straits rift system: Insights from analog models: American Association of Petroleum Geologists Bulletin, 89, 495506.Google Scholar
Gaullier, V., Brun, J. P., Guérin, G., and Lecanu, H., 1993, Raft tectonics: The effects of residual topography below a salt décollement: Tectonophysics, 228, 363381.Google Scholar
Gaullier, V., and Vendeville, B. C., 2005, Salt tectonics driven by sediment progradation: Part II – Radial spreading of sedimentary lobes prograding above salt: American Association of Petroleum Geologists Bulletin, 89, 10811089.Google Scholar
Ge, H., 1996, Kinematics and dynamics of salt tectonics in the Paradox Basin, Utah and Colorado: Field observations and scaled modeling: Ph.D. dissertation, The University of Texas at Austin, Austin, Texas, 317 p.Google Scholar
Ge, H., and Jackson, M. P. A., 1998, Physical modeling of structures formed by salt withdrawal: Implications for deformation caused by salt dissolution: American Association of Petroleum Geologists Bulletin, 82, 228250.Google Scholar
Ge, H., Jackson, M. P. A., and Vendeville, B. C., 1995, Extensional origin of breached Paradox diapirs, Utah and Colorado: Field observations and scaled physical models, in Huffman, A. C. Jr., Lund, W. R., and Godwin, L. H., eds., Geology and resources of the Paradox basin: Salt Lake City, UT, Utah Geological Association, Guidebook 25, 285293.Google Scholar
Ge, H., Jackson, M. P. A., and Vendeville, B. C., 1997, Kinematics and dynamics of salt tectonics driven by progradation: American Association of Petroleum Geologists Bulletin, 81, 398423.Google Scholar
Gealy, B. L., 1955, Topography of the continental slope in northwest Gulf of Mexico: Geological Society of America Bulletin, 66, 203228.Google Scholar
Gemmer, L., Beaumont, C., and Ings, S. J., 2005, Dynamic modeling of passive margin salt tectonics: Effects of water loading, sediment properties, and sedimentation patterns: Basin Research, 17, 383402, doi:10 .1111/j.1365–2117.2005.00274.x.CrossRefGoogle Scholar
Gemmer, L., Ings, S. J., Medvedev, S., and Beaumont, C., 2004, Salt tectonics driven by differential sediment loading; stability analysis and finite-element experiments: Basin Research, 16, 199218.Google Scholar
Gevantman, L. H., ed., 1981, Physical properties data for rock salt: Washington, D.C., National Bureau of Standards, U.S. Department of Commerce, 203 p.Google Scholar
Ghanbarzadeh, S., Hesse, M. A., Prodanovi, M., and Gardner, J. E., 2015, Deformation-assisted fluid percolation in rock salt: Science, 350, 10691072, doi:10.1126/science.aac8747.Google Scholar
Gibbs, A. D., 1984, Structural evolution of extensional basin margins: Journal of the Geological Society, London, 141, 609620.Google Scholar
Giles, K. A., Druke, D. C., Mercer, D. W., and Hunnicutt-Mack, L., 2008, Controls on upper Cretaceous (Maastrichtian) heterozoan carbonate platforms developed on salt diapirs, La Popa basin, northeast Mexico, in Lukasik, J. and Simo, J. A., eds., Controls on carbonate platform development: Tulsa, OK, Society of Economic Paleontologists and Mineralogists, Special Publication 89, 107124.Google Scholar
Giles, K. A., and Lawton, T. F., 1999, Attributes and evolution of an exhumed salt weld, La Popa basin, northeastern Mexico: Geology, 27, 323326, doi:10.1130/0091–7613(1999)027<0323:AAEOAE>2.3.CO;2.2.3.CO;2>CrossRefGoogle Scholar
Giles, K. A., and Lawton, T. F., 2002, Halokinetic sequence stratigraphy adjacent to the El Papalote diapir, northeastern Mexico: American Association of Petroleum Geologists Bulletin, 86, 823840.Google Scholar
Giles, K. A., Lawton, T. F., and Rowan, M. G., 2004, Summary of halokinetic sequence characteristics from outcrop studies of La Popa salt basin, northeastern Mexico, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 1045–1062.Google Scholar
Giles, K. A., and Rowan, M. G., 2012, Concepts in halokinetic-sequence deformation and stratigraphy, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 731, doi:10.1144/SP363.2.Google Scholar
Glennie, K. W., 1990, Outline of North Sea history and structural framework, in Glennie, K. W., ed., Introduction to the petroleum geology of the North Sea (3rd ed.): Oxford, Blackwell Scientific Publications, 3477.Google Scholar
Glennie, K. W., and Boegner, P. L. E., 1981, Sole Pit inversion tectonics, in L. V. Illing and G. D. Hobson, eds., Petroleum geology of the continental shelf of north-west Europe: Proceedings of the second conference, 110–120.Google Scholar
Glennie, K. W., Higham, J., and Semmerik, L., 2003, Permian, in Evans, D., Graham, C., Armour, A., and Bathurst, P., eds., The millennium atlas: Petroleum geology of the central and northern North Sea: London, Geological Society, 91103.Google Scholar
Goetze, C., and Evans, B., 1979, Stress and temperature in the bending lithosphere as constrained by experimental rock mechanics: Geophysical Journal of the Royal Astronomical Society, 59, 463478.Google Scholar
Goldsmith, P. J., Hudson, G., and van Veen, P., 2003, Triassic, in Evans, D., Graham, C., Armour, A., and Bathurst, P., eds., The millennium atlas: Petroleum geology of the central and northern North Sea: London, Geological Society, 105127.Google Scholar
Gomes, P. O., Kilsdonk, B., Grow, T., Minken, J., and Barragan, R., 2012, Tectonic evolution of the outer high of Santos Basin, southern Sao Paulo Plateau, Brazil, and implications for hydrocarbon exploration: American Association of Petroleum Geologists Datapages, doi:10.1306/13351550M1003530.Google Scholar
Goteti, R., Ings, S. J., and Beaumont, C., 2012, Development of salt minibasins under uneven sedimentation: Earth and Planetary Science Letters, 339–340, 156.Google Scholar
Gottschalk, R. R., Anderson, A. V., Walker, J. D., and Da Xilva, J. C., 2004, Modes of contractional salt tectonics in Angola Block 33, Lower Congo basin, West Africa, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 705–734.Google Scholar
Gowers, M., Holtar, E., and Swensson, E., 1993, The structure of the Norwegian Central Trough (Central Graben area): Geological Society, London, Petroleum Geology Conference series, 4, 12451254, doi:10.1144/0041245.Google Scholar
Gradmann, S., and Beaumont, C., 2010, The Eocene salt canopy of the NW Gulf of Mexico explained by the mechanism of squeezed diapirs – A numerical modeling study: CM 2010-Abstracts, 6, 130133.Google Scholar
Gradmann, S., and Beaumont, C., 2012, Coupled fluid flow and sediment deformation in margin-scale salt-tectonic systems: 2. Layered sediment models and application to the northwestern Gulf of Mexico: Tectonics, 31, doi:10.1029/2011TC003035.Google Scholar
Gradmann, S., Beaumont, C., and Albertz, M., 2009, Factors controlling the evolution of the Perdido fold belt, northwestern Gulf of Mexico, determined from numerical models: Tectonics, 28, doi:10.1029/2008TC002326.Google Scholar
Gradmann, S., Beaumont, C., and Ings, S. J., 2012, Coupled fluid flow and sediment deformation in margin-scale salt-tectonic systems: 1. Development and application of simple, single-lithology models: Tectonics, 31, doi:10.1029/2011TC003033.Google Scholar
Graham, R., Jackson, M., Pilcher, R., and Kilsdonk, B., 2012, Allochthonous salt in the sub-Alpine fold-and-thrust belt of Haute Provence, France, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 595615, http://dx.doi.org/10.1144/SP363.30.Google Scholar
Grando, G., and McClay, K., 2004, Structural evolution of the Frampton growth fold system, Atwater Valley–southern Green Canyon area, deep water Gulf of Mexico: Marine and Petroleum Geology, 21, 889910.Google Scholar
Gretener, P. E., 1982, Another look at Alborz Nr. 5 in central Iran: Bulletin der Vereinigung Schweizerische Petroleum-Geologen und -Ingenieure, 48, 18.Google Scholar
Griggs, D. T., 1940, Experimental flow of rocks under conditions favoring recrystallization: Geological Society of America Bulletin, 51, 10011022.CrossRefGoogle Scholar
Groshong, R. H. Jr., 1999, 3-D structural geology: A practical guide to surface and subsurface map interpretation: Berlin, Springer, 324 p.Google Scholar
Grunau, H. R., 1987, A worldwide look at the cap-rock problem: Journal of Petroleum Geology, 10, 243266.Google Scholar
Guardado, L. R., Gamboa, L. A. P., and Lucchesi, C. F., 1989, Petroleum geology of the Campos basin, a model for a producing Atlantic-type basin, in Edwards, J. D. and Santogrossi, P. A., eds., Divergent/passive margin basins: Tulsa, OK, American Association of Petroleum Geologists, Memoir 48, 379.Google Scholar
Guardado, L. R., Spadini, A. R., Brandão, J. S. L., and Mello, M. R., 2000, Petroleum system of the Campos Basin, Brazil, in Mello, M. R. and Katz, B. J., eds., Petroleum systems of: South Atlantic margins: Tulsa, OK, American Association of Petroleum Geologists, Memoir 73, 317324.Google Scholar
Guerra, M. C. M., and Underhill, J. R., 2012, Role of halokinesis in controlling structural styles and sediment dispersal in the Santos Basin, offshore Brazil, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, Sediments and prospectivity: London, Geological Society, Special Publication 363, 159174.Google Scholar
Guglielmo, G., Jackson, M. P. A., and Vendeville, B. C., 1997, Three-dimensional visualization of salt walls and associated fault systems: American Association of Petroleum Geologists Bulletin, 81, 4661.Google Scholar
Guiraud, M., Buta-Neto, A., and Quesne, D., 2010, Segmentation and differential post-rift uplift at the Angola margin as recorded by the transform-rifted Benguela and oblique-to-orthogonal-rifted Kwanza basins: Marine and Petroleum Geology, 27, 10401068.Google Scholar
Gvirtzman, Z., Reshef, M., Buch-Leviatan, O., and Ben Avraham, Z., 2013, Intense salt deformation in the Levant basin in the middle of the Messinian salinity crisis: Earth and Planetary Science Letters, 379, 108119, doi:10.1016/j.epsl.2013.07.018.Google Scholar
Haddou, J., and Tari, G., 2007, Subsalt exploration potential of the Moroccan salt basin: Leading Edge, November, 1454–1460.Google Scholar
Sir Hall, J., 1815, On the vertical position and convolution of certain strata, and their relation with granite: Transactions of the Royal Society, Edinburgh, 7, 79108.Google Scholar
Hall, S. H., 2002, The role of autochthonous salt inflation and deflation in the northern Gulf of Mexico: Marine and Petroleum Geology, 19, 649682.CrossRefGoogle Scholar
Hamblin, W. K., 1965, Origin of reverse drag on the downthrown side of normal faults: Geological Society of America Bulletin, 76, 11451163.Google Scholar
Han, G., Osmond, J., and Zamboni, M., 2010, A USD 100 million “rock”: Bitumen in the deepwater Gulf of Mexico: Society of Petroleum Engineers Drilling and Completion, 25, 290299, doi:10.2118/138228-PA.Google Scholar
Handin, J., and Hager, R. V. Jr., 1957, Experimental deformation of sedimentary rocks under confining pressure: Tests at room temperature on dry samples: American Association of Petroleum Geologists Bulletin, 41, 150.Google Scholar
Handy, M. R., 1990, The solid-state flow of polymineralic rocks: Journal of Geophysical Research, 95, 86478661.Google Scholar
Hanna, M. A., 1953, Fracture porosity in Gulf Coast: American Association of Petroleum Geologists Bulletin, 37, 266281.Google Scholar
Hardie, L. A., 1984, Evaporites: Marine or non-marine?: American Journal of Science, 284, 193240.Google Scholar
Hardie, L. A., 1990, The roles of rifting and hydrothermal CaCl2 brines in the origin of potash evaporites: An hypothesis: American Journal of Science, 290, 43106.Google Scholar
Hardie, L. A., 1996, Secular variation in seawater chemistry: An explanation for the coupled secular variation in the mineralogies of marine limestones and potash evaporites over the past 600 m.y.: Geology, 24, 279283.Google Scholar
Hardin, N. S., 1989, Salt distribution and emplacement processes, northwest lower slope: A suture between two provinces: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 10th annual research conference program and extended abstracts, 54–59.Google Scholar
Harris, G. D., and Veatch, A. C., 1899, A preliminary report on the geology of Louisiana, in Geological Survey of Louisiana report: Baton Rouge: Baton Rouge, LA, Louisiana Geological Survey, Part 5, Geology and Agriculture, 9138.Google Scholar
Harrison, H., Kuhmichel, L., Heppard, P., Milkov, A. V., Turner, J. C., and Greeley, D., 2004, Base of salt structure and stratigraphy – Data and models from Pompano field, VK 989/990, Gulf of Mexico, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 243–270.CrossRefGoogle Scholar
Harrison, H. L., and Patton, B. D., 1995, Translation of salt sheets by basal shear, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference, 99–107.Google Scholar
Harrison, J. C., 1995, Melville Island’s salt-based fold belt, Arctic Canada: Geological Survey of Canada Bulletin, 472, 344 p.Google Scholar
Harrison, J. C., and Jackson, M. P. A., 2008, Bedrock geology, Strand Fiord–Expedition Fiord area, western Axel Heiberg Island, northern Nunavut (parts of NTS 59E, F, G, and H): Ottawa, ON, Geological Survey of Canada, Open File, 5590, CD-ROM.Google Scholar
Harrison, J. C., and Jackson, M. P. A., 2014a, Exposed evaporite diapirs and minibasins above a canopy in central Sverdrup Basin, Axel Heiberg Island, Arctic Canada: Basin Research, 26, 567596, doi:10.1111/bre.12037.Google Scholar
Harrison, J. C., and Jackson, M. P. A., 2014b, Tectonostratigraphy and allochthonous salt tectonics of Axel Heiberg Island, central Sverdrup Basin, Arctic Canada: The University of Texas at Austin, Bureau of Economic Geology Report of Investigations 279, 124 p., ISBN 978-1-100-23499-1, http://doi.org/10.4095/293840.Google Scholar
Harrison, T. S., 1927, Colorado–Utah salt domes: American Association of Petroleum Geologists Bulletin, 11, 111133.Google Scholar
Harrison, W. J., and Summa, L. L., 1991, Paleohydrology of the Gulf of Mexico basin: American Journal of Science, 291, 109176.Google Scholar
Hart, W. H., Jaminski, J., and Albertin, M. L., 2004, Recognition and exploration significance of supra-salt stratal carapaces, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 166–199.Google Scholar
Hay, W. W., 1996, Tectonics and climate: Geologische Rundschau, 85, 409437.Google Scholar
Hayward, A. B., and Graham, R. H., 1989, Some geometrical characteristics of inversion, in Cooper, M. A. and Williams, G. D., eds., Inversion tectonics: London, Geological Society, Special Publication 44, 1739.Google Scholar
Heard, H. C., 1972, Steady state flow in polycrystalline halite at pressure of 2 kilobars, in Heard, B. C., Borg, I. V., Carter, N. L., and Raleigh, C. B., eds., Flow and fracture of rocks: Washington, D.C., American Geophysical Union, Geophysical Monograph Series, 16, 191210.Google Scholar
Hearon, T. E., IV, Rowan, M. G., Giles, K. A., Kernen, R. A., Gannaway, C. E., Lawton, T. F., and Fiduk, J. C., 2015, Allochthonous salt initiation and advance in the northern Flinders and eastern Willouran ranges, South Australia: Using outcrops to test subsurface-based models from the northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 99, 293331, doi:10.1306/08111414022.Google Scholar
Heaton, R. C., Jackson, M. P. A., Bamahoud, M., and Nani, A. S. O., 1995, Superposed Neogene extension, contraction, and salt canopy emplacement in the Yemeni Red Sea, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 333351.Google Scholar
Heidari, M., Nikolinakou, M. A., Flemings, P. B., Hudec, M. R., 2016, A simplified analysis of rising salt domes: Basin Research, 1–14, doi:10.1111/bre.12181.Google Scholar
Henk, A., and Timmerman, M. J., 2005, Europe/Permian basins, in Selley, R. C., Cocks, L. R. M., and Plimer, I. R., eds., Encyclopedia of geology: Amsterdam, Elsevier, 2, 95102.Google Scholar
Hennings, P., 2009, AAPG–SPE–SEG Hedberg research conference on the geologic occurrence and hydraulic significance of fractures in reservoirs: American Association of Petroleum Geologists Bulletin, 93, 14071412.Google Scholar
Hennings, P. H., Olson, J. E., and Thompson, L. B., 2000, A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones: American Association of Petroleum Geologists Bulletin, 84, 830849.Google Scholar
Herrmann, A., and Richter-Bernburg, G., 1955, Frühdiagenetische Störungen der Schichtung und Lagerung im Werra-Anhydrit (Zechstein 1) am Südwestharz: Zeitschrift der Deutschen Geologischen Gesellschaft, 105, 689702.Google Scholar
Herron, D. A., 2011, First steps in seismic interpretation: Tulsa, OK, Society of Exploration Geophysicists, Geophysical Monograph 16, 203 p.Google Scholar
Heward, A. P., 1990, Salt removal and sedimentation in Southern Oman, in Robertson, A. H. F., Searle, M. P., and Ries, A. C., eds., The geology and tectonics of the Oman region: London, Geological Society, Special Publication 49, 637651, doi:10.1144/GSL.SP.1992.049.01.38.Google Scholar
Heybroek, P., 1975, On the structure of the Dutch part of the Central North Sea Graben, in Woodland, A. W., ed., Petroleum and continental shelf of northwest Europe: London, Applied Science Publications, 339351.Google Scholar
Hildenbrand, A., and Urai, J. L., 2003, Investigation of the morphology of pore space in mudstones – First results: Marine and Petroleum Geology, 20, 11851200, doi:10.1016/j.marpetgeo.2003.07.001.Google Scholar
Hite, R. J., Anders, D. E., and Jing, T. G., 1984, Organic-rich source rocks of Pennsylvanian age in the Paradox basin of Utah and Colorado, in Woodward, J., Meissner, F. F., and Clayton, J. L., eds., Hydrocarbon source rocks of the greater Rocky Mountain region: Denver, CO, Rocky Mountain Association of Geologists, 225274.Google Scholar
Hodes, M., Griffiths, P., Smith, K. A., Hurst, W. S., Bowers, W. J., and Sako, K., 2004, Salt solubility and deposition in high temperature and pressure aqueous solutions: American Institute of Chemical Engineers AIChE Journal, 50, 20382049, doi:10.1002/aic.10238.Google Scholar
Hodgson, N. A., Farnsworth, J., and Fraser, A. J., 1992, Salt-related tectonics, sedimentation and hydrocarbon plays in the Central Graben, North Sea, UKCS, in Hardman, R. F. P., ed., Exploration Britain: Geological insights for the next decade: London, Geological Society, Special Publication 67, 3163.Google Scholar
Höfer, K. H., 1964, Results of rheological studies in potash mines: Proceedings, Fourth international conference on strata control and rock mechanics, Columbia University, New York, 6269.Google Scholar
Hofrichter, E., 1980, Probleme der Endlagerung radioaktiver Abfälle in Salzformationen: Zeitschrift der Deutschen Gesellschaft, Hannover, 131, 409430.CrossRefGoogle Scholar
Holland, H. D., Horita, J., and Seyfried, W., 1996, On the secular variations in the composition of Phanerozoic marine potash evaporites: Geology, 24, 993996.Google Scholar
Hollingworth, S. E., Taylor, J. M., and Kellaway, G. A., 1945, Large scale superficial structures in the Northhampton ironstone field: Quarterly Journal of the Geological Society, London, 100, 134.Google Scholar
Holwerda, J. G., and Hutchinson, R. W., 1968, Potash-bearing evaporates in the Danakil area; Ethiopia: Economic Geology, 63, 124150.Google Scholar
Hood, H. C., and Underwood, J. R. Jr., 2001, Geology of Palo Duro Canyon, in Guy, D., ed., The story of Palo Duro Canyon: Lubbock, TX, Texas Tech University Press, 334.Google Scholar
Hood, K. C., Wenger, L. M., Gross, O. P., and Harrison, S. C., 2002, Hydrocarbon systems analysis of the northern Gulf of Mexico: Delineation of hydrocarbon migration pathways using seeps and seismic imaging, in Schumacher, D. and LeSchack, L. A., eds., Surface exploration case histories: Applications of geochemistry, magnetics, and remote sensing: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology 48 and Tulsa, OK, Society of Exploration Geophysicists, Geophysical References Series 11, 2540.Google Scholar
Hooke, R. L., and Hudleston, P. J., 1978, Origin of foliation in glaciers: Journal of Glaciology, 20, 285299.Google Scholar
Hooper, R. J., and More, C., 1995, Evaluation of some salt-related overburden structures in the U.K. Southern North Sea, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective. American Association of Petroleum Geologists, Memoir 65, 251–259.Google Scholar
Hospers, J., and Holthe, J., 1984, Growth of salt-induced structures in the Norwegian–Danish Basin: Tectonophysics, 107, 81104.Google Scholar
Hossack, J. R., and McGuinness, D. B., 1990, Balanced sections and the development of fault and salt structures in the Gulf of Mexico: Boulder, CO, Geological Society of America, Abstracts with Programs, 22, A48.Google Scholar
Hossack, J., 1995, Geometric rules of section balancing for salt structures, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 2940.Google Scholar
House, W. M., and Pritchett, J. A., 1995, Fluid migration and formation pressures associated with allochthonous salt sheets in the northern Gulf of Mexico, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference proceedings, 121–124.Google Scholar
Hovland, M., Fichler, C., Rueslåtten, H., and Johnsen, H. K., 2006a, Deep-rooted piercement structures in deep sedimentary basins – Manifestations of supercritical water generation at depth: Journal of Geochemical Exploration, 89, 157160.Google Scholar
Hovland, M., Kuznetsova, T., Rueslåtten, H., Kvamme, B., Johnsen, H. K., Fladmark, G. E., and Hebach, A., 2006b, Subsurface precipitation of salts in supercritical seawater: Basin Research, 18, 221230.Google Scholar
Hovland, M., MacDonald, I., Rueslåtten, H., Johnsen, H. K., Naehr, T., and Bohrmann, G., 2005, Chapopote asphalt volcano may have been generated by supercritical water: Eos, Transactions, American Geophysical Union, 86, 397402.Google Scholar
Hovland, M., and Rueslåtten, H., 2009, Origin and permeability of deep ocean salts: Geophysical Research Abstracts, 11, EGU2009-6707.Google Scholar
Hovland, M., Rueslåtten, H. G., Johnsen, H. K., Kvamme, B., and Kuznetsova, T., 2006c, Salt formation associated with subsurface boiling and supercritical water: Marine and Petroleum Geology, 23, 855869.Google Scholar
Hoy, R. B., Foose, R. M., and O’Neill, J. B. Jr., 1962, Structure of Winnfield salt dome, Winn Parish, Louisiana: American Association of Petroleum Geologists Bulletin, 46, 14441459.Google Scholar
Hsu, K. J., 1972, Origin of saline giants: a critical review after the discovery of the Mediterranean evaporite: Earth-Science Reviews, 8, 371396.Google Scholar
Hubbert, M. K., 1937, Theory of scale models as applied to the study of geological structures: Geological Society of America Bulletin, 48, 14591520.CrossRefGoogle Scholar
Hubbert, M. K., 1945, Strength of the Earth: American Association of Petroleum Geologists Bulletin, 29, 16301653.Google Scholar
Hudec, M. R., 1995, The Onion Creek salt diapir: an exposed diapir fall structure in the Paradox Basin, Utah, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference, 125–134.Google Scholar
Hudec, M. R., 2004, Salt intrusion: time for a comeback?, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 119–132.Google Scholar
Hudec, M. R., and Jackson, M. P. A., 2004, Regional restoration across the Kwanza basin, Angola: Salt tectonics triggered by repeated uplift of a metastable passive margin: American Association of Petroleum Geologists Bulletin, 88, 971990.Google Scholar
Hudec, M. R., and Jackson, M. P. A., 2006, Advance of allochthonous salt sheets in passive margins and orogens: American Association of Petroleum Geologists Bulletin, 90, 15351564, doi:10.1306/05080605143.Google Scholar
Hudec, M. R., and Jackson, M. P. A., 2007, Terra infirma: Understanding salt tectonics: Earth-Science Reviews, 82, 128, doi:10.1016/j.earscirev.2007.01.001.CrossRefGoogle Scholar
Hudec, M. R., and Jackson, M. P. A., 2009, Interaction between spreading salt canopies and their peripheral thrust systems: Journal of Structural Geology, 31, 11141129, doi:10.1016/j.jsg.2009.06.005.Google Scholar
Hudec, M. R., and Jackson, M. P. A., 2011, The salt mine: a digital atlas of salt tectonics: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Udden Book Series 5; Tulsa, OK, American Association of Petroleum Geologists, Memoir 99, 305 p.Google Scholar
Hudec, M. R., Jackson, M. P. A., and Schultz-Ela, D. D., 2009, The paradox of minibasin subsidence into salt: Clues to the evolution of crustal basins: Geological Society of America Bulletin, 121, 201221, doi:10.1130/B26275.1.Google Scholar
Hudec, M. R., Norton, I. O., Jackson, M. P. A., and Peel, F. J., 2013, Jurassic evolution of the Gulf of Mexico salt basin: American Association of Petroleum Geologists Bulletin, 97, 16831710.Google Scholar
Hudleston, P. J., and Hooke, R. L., 1980, Cumulative deformation in the Barnes Ice Cap and implications for the development of foliation: Tectonophysics, 66, 127146.Google Scholar
Humphris, C. C. Jr., 1978, Salt movement on continental slope, northern Gulf of Mexico, in Bouma, A. H., Moore, G. T., and Coleman, J. M., eds., Framework, facies, and oil-trapping characteristics of the upper continental margin: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 7, 6985.Google Scholar
Humphris, C. C. Jr., 1979, Salt movement in continental slope, northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 63, 782798.Google Scholar
Hunsche, U., and Hampel, A., 1999, Rock salt – The mechanical properties of the host rock material for a radioactive waste repository: Engineering Geology, 52, 271291.Google Scholar
Hunsche, U., and Schulze, O., 1994, Das Kriechverhalten von Steinsalz: Kali und Steinsalz, 11, 238255.Google Scholar
Hunsche, U., Schulze, O., and Langer, M., 1994, Creep and failure behaviour of rock salt around underground cavities, in Der Bergbau an der Schwelle des XXI. Jahrhunderts: Proceedings 16th World Mining Congress, 5, Sofia, Bulgaria, 217–230.Google Scholar
Huntoon, P. W., Billingsly, G. H. Jr., and Breed, W. J., 1982, Geologic map of Canyonlands National Park and vicinity, Utah: Moab, UT, Canyonlands Natural History Association, scale 1:62,500, 1 sheet, 1 key, text.Google Scholar
Inderweisen, P. L., 1983, Salt anticline, East Texas Basin, in Bally, A. W., ed., Seismic expression of structural styles – A picture and work atlas: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology Series, 15, 2.3.2-22–2.3.2-26.Google Scholar
Ingram, G. M., and Urai, J. L., 1999, Top-seal leakage through faults and fractures: The role of mudrock properties, in Alpin, A. C., Fleet, A. J., and Macquaker, J. H. S., eds., Muds and mudstones: Physical and fluid flow properties: London, Geological Society, Special Publications, 158, 125135.Google Scholar
Ings, S. J., and Beaumont, C., 2010, Shortening viscous pressure ridges, a solution to the enigma of initiating salt “withdrawal” minibasins: Geology, 38, 339342, doi:10.1130/G30520.1.Google Scholar
Ings, S. J., Beaumont, C., and Gemmer, L., 2004, Numerical modeling of salt tectonics on passive continental margins: Preliminary assessment of the effects of sediment loading, buoyancy, margin tilt, and isostasy, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 36–68.Google Scholar
Israel, R. R., Leavitt, A. D., Shaughnessy, J. M., and Sanclemente, J. F., 2008, Challenges of directional drilling through salt in deepwater Gulf of Mexico: IADC/SPE Drilling Conference, Orlando, Florida, March 4–6, Paper IADC/SPE 112669, 19 p., doi:10.2118/112669-MS.Google Scholar
Jackson, C. A.-L., Jackson, M. P. A., and Hudec, M. R., 2015b, Understanding the kinematics of salt-bearing passive margins: A critical test of competing hypotheses for the origin of the Albian Gap, Santos Basin, offshore Brazil: Geological Society of America Bulletin, 127, 17301751, doi:10.1130/B31290.1.Google Scholar
Jackson, C. A.L., Jackson, M. P. A., Hudec, M. R., and Rodriguez, C., 2014, Internal structure, kinematics, and growth of a salt wall: Insights from 3-D seismic data: Geology, 42, 307310, doi:10.1130/G34865.1.Google Scholar
Jackson, C. A.-L., Jackson, M. P. A., Hudec, M. R., and Rodriguez, C. R., 2015a, Enigmatic structures within salt walls of the Santos Basin – Part 1: Geometry and kinematics from 3D seismic reflection and well data: Journal of Structural Geology, 75, 135162, doi:10.1016/j.jsg.2015.01.010.Google Scholar
Jackson, C. A.-L., Kane, K. E., and Larsen, E., 2010, Structural evolution of minibasins on the Utsira High, northern North Sea; Implications for Jurassic sediment dispersal and reservoir distribution: Petroleum Geoscience, 16, 105120, doi:10.1144/1354-079309-011.Google Scholar
Jackson, M. P. A., 1982, Fault tectonics of the East Texas Basin: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Geological Circular 82-4, 31 p.Google Scholar
Jackson, M. P. A., 1985, Natural strain in diapiric and glacial rock salt, with emphasis on Oakwood dome, East Texas: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 143, 74 p.Google Scholar
Jackson, M. P. A., 1995, Retrospective salt tectonics, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 128.Google Scholar
Jackson, M. P. A., 1997, Conceptual breakthroughs in salt tectonics: a historical review, 1856–1993: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 246, 51 p.Google Scholar
Jackson, M. P. A., and Cornelius, R. R., 1985, Tertiary salt diapirs exposed at different structural levels in the Great Kavir (Dasht-i Kavir) south of Semnan, north-central Iran: A remote-sensing study of their internal structure and shape: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Open File Report OF-WTWI-1985–22, 108 p.Google Scholar
Jackson, M. P. A., and Cornelius, R. R., 1987, Stepwise centrifuge modeling of the effects of differential sediment loading on the deformation of salt structures, in Lerche, I. and O’Brien, J. J., eds., Dynamical geology of salt and related structures: Orlando, FL, Academic Press, 163259.Google Scholar
Jackson, M. P. A., Cornelius, R. R., Craig, C. H., Gansser, A., Stöcklin, J., and Talbot, C. J., 1990, Salt diapirs of the Great Kavir, Central Iran: Boulder, CO, Geological Society of America, Memoir 177, 139 p.Google Scholar
Jackson, M. P. A., Cornelius, R. R., Craig, C. H., and Talbot, C. J., 1987, The Great Kavir salt canopy: A major new class of salt structure (abs.): Geological Society of America, Abstracts with Programs, 19, 714.Google Scholar
Jackson, M. P. A., and Cramez, C., 1989, Seismic recognition of salt welds in salt tectonics regimes, in Gulf of Mexico salt tectonics, associated processes and exploration potential: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 10th annual research conference program and abstracts, 66–71.Google Scholar
Jackson, M. P. A., Cramez, C., and Fonck, J.-M., 2000, Role of subaerial volcanic rocks and mantle plumes in creation of South Atlantic margins: Implications for salt tectonics and source rocks: Marine and Petroleum Geology, 17, 477498.Google Scholar
Jackson, M., Dooley, T., Hudec, M., and McDonnell, A., 2011, The pillow fold belt: A key subsalt structural province in the northern Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery, 10329, 23 p.Google Scholar
Jackson, M. P. A., and Harrison, J. C., 2006, An allochthonous salt canopy on Axel Heiberg Island, Sverdrup basin, Arctic Canada: Geology, 34, 10451048.Google Scholar
Jackson, M. P. A., and Hudec, M. R., 2004, A new mechanism for advance of allochthonous salt sheets, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 220–242.Google Scholar
Jackson, M. P. A., and Hudec, M. R., 2005, Stratigraphic record of translation down ramps in a passive-margin detachment: Journal of Structural Geology, 27, 889911.Google Scholar
Jackson, M. P. A., Hudec, M. R., and Hegarty, K. A., 2005, The great West African Tertiary coastal uplift: Fact or fiction? A perspective from the Angolan divergent margin: Tectonics, 24, TC6014, doi:10.1029/2005TC001836.CrossRefGoogle Scholar
Jackson, M. P. A., Hudec, M. R., Jennette, D. C., and Kilby, R. E., 2008, Evolution of the Cretaceous Astrid thrust belt in the ultradeep-water Lower Congo basin, Gabon: American Association of Petroleum Geologists Bulletin, 92, 487511.Google Scholar
Jackson, M. P. A., and Schultz-Ela, D. D., 2000, Why roho, why counterregional? (abs.): American Association of Petroleum Geologists Annual Meeting Program, 9, A73.Google Scholar
Jackson, M. P. A., Schultz-Ela, D. D., Hudec, M. R., Watson, I. A., and Porter, M. L., 1998, Structure and evolution of Upheaval Dome: A pinched-off salt diapir: Geological Society of America Bulletin, 110, 15471573.Google Scholar
Jackson, M. P. A., Schultz-Ela, D. D., Hudec, M. R., Watson, I. A., and Porter, M. L., 2001, Structure and evolution of Upheaval Dome: Pinched-off salt diapir or meteoritic impact structure?: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 262, 93 p.Google Scholar
Jackson, M. P. A., and Seni, S. J., 1984, Atlas of salt domes in the East Texas Basin: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 140, 102 p.Google Scholar
Jackson, M. P. A., and Talbot, C. J., 1986, External shapes, strain rates, and dynamics of salt structures: Geological Society of America Bulletin, 97, 305323.Google Scholar
Jackson, M. P. A., and Talbot, C. J., 1989a, Anatomy of mushroom-shaped diapirs: Journal of Structural Geology, 11, 211230.Google Scholar
Jackson, M. P. A., and Talbot, C. J., 1989b, Salt canopies: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 10th annual research conference program and extended abstracts, 72–78.Google Scholar
Jackson, M. P. A., and Talbot, C. J., 1991, A glossary of salt tectonics: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Geologic Circular 91-4, 44 p.Google Scholar
Jackson, M. P. A., Talbot, C. J., and Cornelius, R. R., 1988, Centrifuge modeling of the effects of aggradation and progradation on syndepositional salt structures: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 173, 93 p.Google Scholar
Jackson, M. P. A., and Vendeville, B. C., 1990, The rise and fall of diapirs during thin-skinned extension (abs.): American Association of Petroleum Geologists Bulletin, 74, 683.Google Scholar
Jackson, M. P. A., and Vendeville, B. C., 1994, Regional extension as a geologic trigger for diapirism: Geological Society of America Bulletin, 106, 5773.Google Scholar
Jackson, M. P. A., Vendeville, B. C., and Schultz-Ela, D. D., 1994, Structural dynamics of salt systems: Annual Review of Earth and Planetary Sciences, 22, 93117.Google Scholar
Jackson, M. P. A., Warin, O. N., Woad, G. M., and Hudec, M. R., 2003, Neoproterozoic allochthonous salt tectonics during the Lufilian orogeny in the Katangan Copperbelt, central Africa: Geological Society of America Bulletin, 115, 314330.Google Scholar
Jahani, S., Callot, J. P., De Lamotte, D. F., Letouzey, J., and Leturmey, P., 2007, The salt diapirs of the eastern Fars Province (Zagros, Iran): A brief outline of their past and present, in Lacombe, O., Lavé, J., Roure, F., and Vergés, J., eds., Thrust belts and foreland basins from fold kinematics to hydrocarbon systems: Berlin, Springer, 289308.Google Scholar
Jahani, S., Callot, J. P., Letouzey, J., and De Lamotte, D. F., 2009, The eastern termination of the Zagros fold-and-thrust belt, Iran: Structures, evolution, and relationships between salt plugs, folding, and faulting: Tectonics, 28, TC6004, doi:10.1029/2008TC002418.Google Scholar
Jamieson, G. A., 1995, Subsalt and base-of-salt reflector relationships and hydrocarbon trap implications from a large 3D, depth migrated, seismic volume: Offshore Technology Conference, May 1–4, 1995, Houston, Texas, 7638-MS, doi:10.4043/7638-MS.Google Scholar
Jamieson, G., Hannan, A., and Biles, N., 2000, Structural overview of the outer shelf, deepwater, and ultra-deepwater Gulf of Mexico, relationships between salt features and minibasin morphology: Tulsa, OK, American Association of Petroleum Geologists, Datapages Discovery Series (CD-ROM) 1.Google Scholar
Jeanjean, P., Hill, A., and Taylor, S., 2003, The challenges of siting facilities along the Sigsbee escarpment in the southern Green Canyon area of the Gulf of Mexico: Framework for integrated studies: Offshore Technology Conference, May 5–8, 2003, Houston, Texas, paper OTC 15156.Google Scholar
Jenyon, M. K., 1986a, Salt tectonics: Essex, Elsevier Applied Science, 191 p.Google Scholar
Ji, S., Huang, T., Fu, K., and Li, Z., 2010, Dirty salt velocity inversion: The road to a clearer subsalt image: Society of Exploration Geophysicists Annual Meeting, October 17–22, Denver, Colorado, 4103–4107.Google Scholar
Jin, S., and Walraven, D., 2003, Wave equation GSP prestack depth migration and illumination: Leading Edge, 22, 606660, doi:10.1190/1.1599687.Google Scholar
Johnson, H. A., and Bredeson, D. H., 1971, Structural development of some shallow salt domes in Louisiana Miocene productive belt: American Association of Petroleum Geologists Bulletin, 55, 204226.Google Scholar
Jones, I. F., and Davison, I., 2014, Seismic imaging in and around salt bodies: Interpretation, 2, 120, doi:10.1190/INT-2014-0033.1.Google Scholar
Jordan, P. G., 1987, The deformational behaviour of bimineralic limestone–halite aggregates: Tectonophysics, 135, 185197.Google Scholar
Jordan, P., 1988, The rheology of polymineralic rocks – An approach. Geologische Rundschau, 77, 285294.Google Scholar
Jordan, P., 1991, Development of asymmetrical shale pull-aparts in evaporite shear zones: Journal of Structural Geology, 13, 399409.Google Scholar
Kanbur, Z., Louie, J. N., Chávez-Pérez, S., Plank, G., and Morey, D., 2000, Seismic reflection study of Upheaval Dome, Canyonlands National Park, Utah: Journal of Geophysical Research – Planets, 105, 94899505.Google Scholar
Kehle, R. O., 1970, Analysis of gravity sliding and orogenic translation: Geological Society of America Bulletin, 81, 16411664.Google Scholar
Kehle, R. O., 1988, The origin of salt structures, in Schreiber, B. C., ed., Evaporites and hydrocarbons: New York, Columbia University Press, 345404.Google Scholar
Kendall, A. C., 1988, Aspects of evaporite basin stratigraphy, in Schreiber, B. C., ed., Evaporites and hydrocarbons: New York, Columbia University Press, 1165.Google Scholar
Kendall, A. C., 1989, Brine mixing in the Middle Devonian of Western Canada and its possible significance to regional dolomitization: Sedimentary Geology, 64, 271285.Google Scholar
Kendrick, J. W., 2000, Turbidite reservoir architecture in the northern Gulf of Mexico deepwater: Insights from the development of Auger, Tahoe, and Ram/Powell fields, in Deep-water reservoirs of the world: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 20th annual research conference, 450–468.Google Scholar
Kenkmann, T., 2003, Dike formation, cataclastic flow, and rock fluidization during impact cratering: An example from the Upheaval Dome structure, Utah: Earth and Planetary Science Letters, 214, 4358, doi:10.1016/ S0012-821X(03)00359–5.Google Scholar
Kenkmann, T., Jahn, A., Scherler, D., and Ivanov, B. A., 2005, Structure and formation of a central uplift: A case study at the Upheaval Dome impact crater, Utah, in Kenkmann, T., Hörz, F., and Deutsch, A., eds., Large meteorite impacts III: Boulder, CO, Geological Society of America, Special Paper 384, 85115.Google Scholar
Kenkmann, T., and Scherler, D., 2002, New structural constraints on the Upheaval Dome impact crater: Houston, TX, Lunar and Planetary Institute, Lunar and Planetary Science XXXIII, abstract 1037.Google Scholar
Kent, P. E., 1958, Recent studies of south Persian salt plugs: American Association of Petroleum Geologists Bulletin, 42, 29512972.Google Scholar
Kent, P. E., 1970, The salt plugs of the Persian Gulf region: Transactions of the Leicester Literary and Philosophical Society, 64, 5688.Google Scholar
Kent, P. E., 1979, The emergent Hormuz salt plugs of southern Iran: Journal of Petroleum Geology, 2, 117144.Google Scholar
Kent, P. E., 1987, Island salt plugs in the Middle East and their tectonic implications, in Lerche, I., and O’Brien, J. J., eds., Dynamical geology of salt and related structures: Orlando, FL, Academic Press, 337.Google Scholar
Kerr, P. F., and Kopp, O. C., 1958, Salt dome breccia: American Association of Petroleum Geologists Bulletin, 42, 548560.Google Scholar
Kieft, R. L., Jackson, C. A.-L., and Hampson, G. J., 2010, Sedimentology and sequence stratigraphy of the Hugin Formation, quadrant 15, Norwegian sector, South Viking graben: Geological Society Petroleum Geology Conference series 7, 157176, doi:10.1144/0070157.Google Scholar
Kilby, R. E., Diegel, F. A., and Styzen, M. J., 2008, Age of sediments encasing allochthonous salt in the Gulf of Mexico: Clues to emplacement history: American Association of Petroleum Geologists Annual Convention and Exhibition Abstract Volume, p. 109.Google Scholar
Kilsdonk, B., Graham, R., and Pilcher, R., 2010, Deep water Gulf of Mexico sub-salt structural framework: Offshore Technology Conference paper 20937, 15 p.Google Scholar
Kirkland, D. W., Denison, R. E., and Dean, W. E., 2000, Parent brine of the Castile evaporites (Upper Permian), Texas and New Mexico: Journal of Sedimentary Research, Series A: Sedimentary Petrology and Processes, 70, 749761.CrossRefGoogle Scholar
Kirkland, D. W., and Evans, R., 1973, Marine evaporites: Origin, diagenesis and geochemistry: Stroudsburg, PA, Dowden, Hutchinson & Ross.Google Scholar
Kirkland, D. W., and Evans, R., 1981, Source-rock potential of evaporitic environment: American Association of Petroleum Geologists Bulletin, 65, 181190.Google Scholar
Knauth, L. P., and Kumar, M. B., 1983, Isotopic character and origin of brine leaks in the Avery Island salt mine, south Louisiana, U.S.A: Journal of Hydrology, 66, 343350.Google Scholar
Knauth, L. P., Kumar, M. B., and Martinez, J. D., 1980, Isotope geochemistry of water in Gulf Coast saltdomes: Journal of Geophysical Research, 85, 48634871.Google Scholar
Kneller, B., and McCaffrey, B., 1995, Modeling the effects of salt-induced topography on deposition from turbidity currents, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference, 135–145.Google Scholar
Knipping, B., 1989, Basalt intrusions in evaporites: Berlin, Springer, Lecture Notes in Earth Sciences, 24, 132 p.Google Scholar
Kockel, F., 1990, Morphology and genesis of Northwest-German salt structures, in Proceedings, Symposium on diapirism with special reference to Iran, vol. 1: Tehran, Tehran University, Governerate of Hormozgan, and Ministry of Mines and Metals, 225245.Google Scholar
Kockel, F., 1998, Salt problems in Northwest Germany and the German North Sea sector, in Kockel, F. and Marschall, R., eds., Geology and geophysics of salt structures: Journal of Seismic Exploration, 7, 219235.Google Scholar
Kockel, F., 2003, Inversion structures in Central Europe – Expressions and reasons, an open discussion: Netherlands Journal of Geosciences/Geologie en Mijnbouw, 82, 367382.Google Scholar
Koeberl, C., Plescia, J. B., Hayward, C. L., and Reimold, W. U., 1999, A petrographical and geochemical study of quartzose nodules, country rocks, and dike rocks from the Upheaval Dome structure, Utah: Meteoritics and Planetary Science, 34, 861868.Google Scholar
Königsberger, D., and Morath, O., 1913, Theoretische Grundlagen der experimentellen Tektonik: Zeitschrift der Deutschen Geologischen Gesellschaft, Hannover, 65, 6586.Google Scholar
Konstantinovskaya, E. A., Harris, L. B., Poulin, J., and Ivanov, G. M., 2007, Transfer zones and fault reactivation in inverted rift basins: Insights from physical modelling: Tectonophysics, 441, 126.Google Scholar
Kostick, D. S., 2011, 2010 Minerals Yearbook: Salt (advance release): Washington, D.C., U.S. Department of the Interior, U.S. Geological Survey, 63.1–63.23.Google Scholar
Kovalevich, V. M., Peryt, T. M., and Petrichenko, O. I., 1998, Secular variation in seawater chemistry during the Phanerozoic as indicated by brine inclusions in halite: Journal of Geology, 106, 695712.Google Scholar
Koyi, H., 1988, Experimental modeling of role of gravity and lateral shortening in Zagros mountain belt: American Association of Petroleum Geologists Bulletin, 72, 13811394.Google Scholar
Koyi, H., 1998, The shaping of salt diapirs: Journal of Structural Geology, 20, 321338.Google Scholar
Krézsek, C., and Bally, A. W., 2006, The Transylvanian Basin (Romania) and its relation to the Carpathian fold and thrust belt: Insights in gravitational salt tectonics: Marine and Petroleum Geology, 23, 405442, doi:10.1016/j.marpetgeo.2006.03.003.Google Scholar
Kriens, B. J., Shoemaker, E. M., and Herkenhoff, K. E., 1997, Structure and kinematics of a complex impact crater, Upheaval Dome, southeast Utah, in Link, P. K., and Kowallis, B. J., eds., Geological Society of America 1997 Annual Meeting, Salt Lake City, Utah, Field Trip Guidebook: Salt Lake City, UT, Brigham Young University Geology Studies, 42, 1931.Google Scholar
Kriens, B. J., Shoemaker, E. M., and Herkenhoff, K. E., 1999, Geology of the Upheaval Dome impact structure, southeast Utah: Journal of Geophysical Research, 104, 1886718887, doi:10.1029/1998JE000587.Google Scholar
Krijgsman, W., Blanc-Valleron, M. M., Flecker, R., Hilgen, F. J., Kouwenhoven, T. J., Merle, D., Orszag-Sperber, F., and Rouchy, J. M., 2002, The onset of the Messinian salinity crisis in the Eastern Mediterranean (Pissouri basin, Cyprus): Earth and Planetary Science Letters, 194, 299310, doi:10.1016/j.margeo.2008.04.010.Google Scholar
Krijgsman, W., Hilgen, F. J., Raffi, I., Sierro, F. J., and Wilson, D. S., 1999, Chronology, causes and progression of the Messinian salinity crisis: Nature, 400, 652655.Google Scholar
Krueger, S. W., 2010, Dynamics of tear faults in the salt-detached systems of the northern Gulf of Mexico: American Association of Petroleum Geologists Search and Discovery, paper 50346.Google Scholar
Kuhn, K., and Verkerk, B., 1980, Disposal in salt formations, in Simon, R. and Orlowski, S., eds., Radioactive waste management and disposal: Proceedings of the first European Community conference: New York: Harwood Academic Publishers, 385419.Google Scholar
Kukla, P., Reuning, L., Becker, S., Urai, J. L., and Schoenherr, J., 2011, Distribution and mechanisms of overpressure generation and deflation in the late Neoproterozoic to early Cambrian South Oman Salt Basin: Geofluids, doi:10.1111/j.1468–8123.2011.00340.x.Google Scholar
Kunz, T. J., and Jones, T. K., 1994, 3-D geological velocity model construction for 3-D depth imaging: Offshore Technology Conference, May 2–5, 1994, Houston, Texas, 7398-MS, doi:10.4043/7398-MS.Google Scholar
Kupfer, D. H., 1962, Structure of Morton Salt Company mine, Weeks Island salt dome, Louisiana: American Association of Petroleum Geologists Bulletin, 46, 14601467.Google Scholar
Kupfer, D. H., 1968, Relationship of internal to external structure of salt domes, in Braunstein, J., ed., Diapirism and diapirs: Tulsa, OK, American Association of Petroleum Geologists, Memoir 8, 7989.Google Scholar
Kupfer, D. H., 1974, Boundary shear zones in salt stocks, in 4th Symposium on Salt: Cleveland, OH, Northern Ohio Geological Society, 1, 215225.Google Scholar
Kupfer, D. H., 1976, Shear zones inside Gulf Coast stocks help delineate spines of movement: American Association of Petroleum Geologists Bulletin, 60, 14341447.Google Scholar
Kupfer, D. H., 1990, Anomalous features in the Five Islands salt stocks, Louisiana: Gulf Coast Association of Geological Societies Transactions, 40, 425437.Google Scholar
Kurlansky, M., 2002, Salt: A world history: New York, Penguin, 484 p.Google Scholar
Kusznir, N. J., and Park, R. G., 1986, Continental lithosphere strength: The critical role of lower crustal deformation, in Dawson, J. B., Carswell, D. A., Hall, J., and Wedepohl, K. H., eds., The nature of the lower continental crust: London, Geological Society, Special Publication 24, 7993.Google Scholar
Kyle, J. R., and Price, P. E., 1986, Metallic sulphide mineralization in salt-dome cap rocks, Gulf Coast, U.S.A.: Transactions of the Institution of Mining and Metallurgy, 95, 616.Google Scholar
Lacazette, A., Thomas, A. R., and Kuhfal, D., 2007, Fracture development in salt dome caprock, Hardin County, Texas: American Association of Petroleum Geologists, Search and Discovery paper 50058.Google Scholar
Ladzekpo, D. H., Sekharan, K. K., and Gardner, G. H. F., 1988, Physical modeling for hydrocarbon exploration: Society of Exploration Geophysicists, 58th annual meeting, SEG Abstracts 1, 586588.Google Scholar
Lamplugh, G. W., 1920, Structure of the Weald and analogous tracts: Quarterly Journal of the Geological Society, London, 75, 7395 (Anniversary Address of the President).Google Scholar
Land, L. S., 1995, Na–Ca–Cl saline formation waters, Frio Formation (Oligocene) south Texas, USA: Products of diagenesis: Geochimica et Cosmochimica Acta, 59, 21632174.Google Scholar
Land, L. S., Kupecz, J. A., and Mack, L. E., 1988, Louann Salt geochemistry (Gulf of Mexico sedimentary basin, USA): A preliminary synthesis: Chemical Geology, 74, 2535.Google Scholar
Larsen, B., Ben-Avraham, Z., and Shulman, H., 2002, Fault and salt tectonics in the southern Dead Sea basin: Tectonophysics, 346, 7190.Google Scholar
Larsen, J. G., 1984, Textural and petrofabric analyses of rock salt related to mechanical test data – A quantitative approach: Zechstein Salt Denmark, Salt Research Project EFP-81: Copenhagen, Geological Survey of Denmark and Greenland, Series C, 1, 772.Google Scholar
Laubach, S. E., Olson, J. E., and Gross, M. R., 2009, Mechanical and fracture stratigraphy: American Association of Petroleum Geologists Bulletin, 93, 14131426, doi:10.1306/07270909094.Google Scholar
Laubscher, H. P., 1977, Fold development in the Jura: Tectonophysics, 37, 337362.Google Scholar
Laudon, R. C., 1984, Evaporite diapirs in the La Popa basin, Nuevo Leon, Mexico: Geological Society of America Bulletin, 95, 12191225.Google Scholar
Lawton, T. F., and Giles, K. A., 1997, El Papalote diapir. La Popa basin: Structure, stratigraphy and paleontology of Late Cretaceous–Early Tertiary Parras–La Popa foreland basin near Monterrey, northeast Mexico: American Association of Petroleum Geologists, field trip guidebook 10, 5574.Google Scholar
Lawton, T. F., Vega, F. J., Giles, K. A., and Rosales-Domínguez, C., 2001, Stratigraphy and origin of the La Popa basin, Nuevo Léon and Coahuila, Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The western Gulf of Mexico basin: Tectonics, sedimentary basins, and petroleum systems: Tulsa, OK, American Association of Petroleum Geologists, Memoir 75, 219240.Google Scholar
Lee, G. H., Bryant, W. R., and Watkins, J. S., 1989, Salt structures and sedimentary basins in the Keathley Canyon area, northwestern Gulf of Mexico: Their development and tectonic implications: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 10th annual research conference program and extended abstracts, 90–93.Google Scholar
Lee, G. H., Watkins, J. S., and Bryant, W. R., 1996, Bryant Canyon fan system: An unconfined, large river-sourced system in the northwestern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 80, 340358.Google Scholar
Lees, G. M., 1927, Salzgletcher in Persien: Mitteilungen der Geologischen Gesellschaft in Wien, 20, 2934.Google Scholar
Leitner, C., Neubauer, F., Urai, J. L., and Schoenherr, J., 2011, Structure and evolution of a rocksalt–mudrock–tectonite: The haselgebirge in the northern Calcareous Alps: Journal of Structural Geology, 33, 970984.Google Scholar
Lentini, M. R., Fraser, S. I., Sumner, H. S., and Davies, R. J., 2010, Geodynamics of the central South Atlantic conjugate margins: Implications for hydrocarbon potential: Petroleum Geoscience, 16, 217229.Google Scholar
Letouzey, J., 1990, Fault reactivation, inversion and fold-thrust belt, in Letouzey, J., ed., Petroleum and tectonics in mobile belts: Paris, Technip, 101128.Google Scholar
Letouzey, J., Baghbani, D., Rudkiewicz, J. L., Cuilhe, L., and Kazemi, K., 2006, Structural evolution and petroleum system of the Qum region (Central Iran): European Association of Geoscientists and Engineers, 68th EAGE Conference and Exhibition, Vienna.Google Scholar
Letouzey, J., Colletta, B., Vially, R., and Chermette, J. C., 1995, Evolution of salt-related structures in compressional settings, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective. Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 4160.Google Scholar
Letouzey, J., and Sherkati, S., 2004, Salt movement, tectonic events, and structural style in the central Zagros fold and thrust belt (Iran), in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 753–778.Google Scholar
Letouzey, J., Werner, P., and Marty, A., 1990, Fault reactivation and structural inversion; Backarc and intraplate compressive deformations; Example of the eastern Sunda shelf (Indonesia): Tectonophysics, 183, 341362.Google Scholar
Lewis, S., and Holness, M., 1996, Equilibrium halite-H2O dihedral angles; high rock-salt permeability in the shallow crust?: Geology, 24, 431434.Google Scholar
Li, J., Webb, A. A. G., Mao, X., Eckhoff, I., Colón, C., Zhang, K., Wang, H., Li, A., and He, D., 2014, Active surface salt structures of the western Kuqa fold–thrust belt, northwestern China: Geosphere, 10, 12191234.Google Scholar
Liang, W., Yang, C., Zhao, Y., Dusseault, M. B., and Liu, J., 2007, Experimental investigation of mechanical properties of bedded salt rock: International Journal of Rock Mechanics and Mining Sciences, 44, 400411, doi:10.1016/j.ijrmms.2006.09.007.Google Scholar
Lickorish, W. H., and Ford, M., 1998, Sequential restoration of the external Alpine Digne thrust system, SE France, constrained by kinematic data and synorogenic sediments, in Mascle, A., Puigdefabregas, C., Luterbacker, H. P., and Fernandez, M., eds., Cenozoic foreland basins of western Europe: London, Geological Society, Special Publication 134, 189211.Google Scholar
Lickorish, W. H., Ford, M., Bürgisser, J., and Cobbold, P. R., 2002, Arcuate thrust systems in sandbox experiments: A comparison to the external arcs of the Western Alps: Geological Society of America Bulletin, 114, 10891107.Google Scholar
Lin, S. T., 1992, Experimental study of syndepositional and postdepositional gravity spreading of a brittle overburden and viscous substratum: Master’s thesis, The University of Texas at Austin, Austin, Texas, 196 p.Google Scholar
Link, T. A., 1930, Experiments relating to salt-dome structures: American Association of Petroleum Geologists Bulletin, 14, 483508.Google Scholar
Liro, L. M., Murillas, J., Villalobos, L., Gatenby, G., and Mathur, V., 2004, Salt sutures in single- and multi-tiered allochthons, Green Canyon and Walker Ridge areas, deep-water Gulf of Mexico, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 200–219.Google Scholar
Liro, L. M., Weber, N., O’Hara, S., Eternadi, M., Lock-Williams, S., Cubsanski, M., Kadri, M., and Montecchi, P., 2001, Subsalt exploration trap styles, Walker Ridge and Keathley Canyon areas, deep-water Gulf of Mexico, in Petroleum systems of deep-water basins: Global and Gulf of Mexico experience: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 21st annual research conference, 21 p.Google Scholar
Lister, J. R., and Kerr, R. C., 1989, The propagation of two-dimensional and axisymmetric viscous gravity currents at a fluid interface: Journal of Fluid Mechanics, 203, 215249.Google Scholar
Lobao, J. L., and Pilger, R. H. Jr., 1985, Early evolution of salt structures in the north Louisiana salt basin: Gulf Coast Association of Geological Societies Transactions, 35, 189198.Google Scholar
Lock, B. E., and Duex, T. W., 1996, Xenolithic inclusions within the salt at Weeks Island, Louisiana, and their significance: Gulf Coast Association of Geological Societies Transactions, 46, 229234, doi:10.1306/2DC40B23 -0E47-11D7-8643000102C1865D.Google Scholar
Lock, B. E., Walker Fife, A., and Anderson, E., 2004, Bacteria–petroleum reactions; salt dome cap rock genesis compared with similar processes from Permian outcrops in West Texas: Gulf Coast Association of Geological Societies Transactions, 54, 361367.Google Scholar
Lohest, M., 1921, A propos des plis diapirs. Rappel de quelques principes de tectonique: Société Géologique de Belgique Annales, 44, B94B107.Google Scholar
Lohmann, H. H., 1972, Salt dissolution in subsurface of British North Sea as interpreted from seismograms: American Association of Petroleum Geologists Bulletin, 56, 472479.Google Scholar
Lowenstein, T. K., Hardie, L. A., Timofeeff, M. N., and Demicco, R. V., 2003, Secular variation in seawater chemistry and the origin of calcium chloride basinal brines: Geology, 31, 857860.Google Scholar
Lowenstein, T. K., Timofeeff, M. N., Brennan, S. T., Demicco, H. L. A., and Demicco, R. V., 2001, Oscillations in Phanerozoic seawater chemistry: Evidence from fluid inclusions: Science, 294, 10861088.Google Scholar
Lowrie, A., Yu, Z., and Lerche, I., 1991, Hydrocarbon trap types and deformation styles modeled using quantified rates of salt movement, Louisiana margin: Gulf Coast Association of Geological Societies Transactions, 41, 445459.Google Scholar
Lucia, F. J., 1972, Recognition of evaporite-carbonate shoreline sedimentation, in Rigby, J. K. and Hamblin, W. K., eds., Recognition of ancient sedimentary environments: Tulsa, OK, Society of Economic Paleontologists and Mineralogists, Special Publication 16, 160191.Google Scholar
Lundin, E. R., 1992, Thin-skinned extensional tectonics on a salt detachment, northern Kwanza Basin, Angola: Marine and Petroleum Geology, 9, 405411.Google Scholar
Luo, G., Flemings, P. B., Hudec, M. R., and Nikolinakou, M. A., 2015, The role of pore fluid overpressure in the substrates of advancing salt sheets, ice glaciers, and critical-state wedges: Journal of Geophysical Research: Solid Earth, 120, 87105, doi:10.1002/2014JB011326.Google Scholar
Luo, G., Nikolinakou, M. A., Flemings, P. B., and Hudec, M. R., 2012, Geomechanical modeling of stresses adjacent to salt bodies: Part 1 – Uncoupled models: American Association of Petroleum Geologists Bulletin, 96, 4364.Google Scholar
Lux, K.-H., 2005, Long-term behaviour of sealed liquid-filled salt cavities – A new approach for physical modelling and numerical simulation – Basics from theory and lab investigations: Erdöl Erdgas Kohle, 121, 414422.Google Scholar
Lyell, C., 1830, Principles of geology, being an attempt to explain the former changes of the earth’s surface, by reference to causes now in operation: Vol. 1, London, Murray.Google Scholar
Magoon, L. B., and Dow, W. G., 1994, The petroleum system, in Magoon, L. B. and Dow, W. G., eds., The petroleum system – From source to trap: Tulsa, OK, American Association of Petroleum Geologists, Memoir 60, 324.Google Scholar
Maione, S. J., 2001, Discovery of ring faults associated with salt withdrawal basins of early Cretaceous age in the East Texas basin: Leading Edge, 20, 818829.Google Scholar
Malek-Aslani, M., 1980, Environmental and diagenetic controls of carbonate and evaporite source rocks: Gulf Coast Association of Geological Societies Transactions, 30, 445456.Google Scholar
Malthe-Sørenssen, A., Walmann, T., Jamtveit, B., Feder, J., Jøssang, T., 1999, Simulation and characterization of fracture patterns in glaciers: Journal of Geophysical Research – Solid Earth, 104, 2315723174, doi:10.1029/1999JB900219Google Scholar
Mann, P., 2007, Global catalogue, classification and tectonic origins of restraining- and releasing bends on active and ancient strike–slip fault systems, in Cunningham, W. D. and Mann, P., eds., Tectonics of strike–slip restraining and releasing bends: London, Geological Society, Special Publication 290, 13142.Google Scholar
Mann, P., Gahagan, L., and Gordon, M., 2001, Tectonic setting of the world’s giant oil fields: World Oil, September, 42–50; October, 78–84; November, 56–60.Google Scholar
Marco, S., Weinberger, R., and Agnon, A., 2002, Radial clastic dykes formed by a salt diapir in the Dead Sea rift, Israel: Terra Nova, 14, 288294.Google Scholar
Martin, R. G., 1978, Northern and eastern Gulf of Mexico continental margin: Stratigraphic and structural framework, in Bouma, A. H., Moore, G. T., and Coleman, J. M., eds., Framework, facies, and oil-trapping characteristics of the upper continental margin: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 7, 2142.Google Scholar
Martinez, J. D., 1978, An investigation of the utility of Gulf Coast salt domes for the storage or disposal of radioactive wastes, Vol. 1: Baton Rouge, LA, Louisiana State University Institute for Environmental Studies, Contract Report EW-78-C-05–5941/53, 390 p.Google Scholar
Marton, G., Schoenborn, G., Carpenter, D., and Cool, T., 2004, Salt tectonics of the continent–ocean transition, deep-water Angola, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 588–613.Google Scholar
Marton, G., Tari, C., and Lehmann, C. T., 1998, Evolution of salt-related structures and their impact on the post-salt petroleum systems of the Lower Congo Basin, offshore Angola: American Association of Petroleum Geologists International Conference and Exhibition, Rio de Janeiro, Extended Abstracts Volume, 834–367.Google Scholar
Masson, M. P., 1972, L’exploration pétrolière en Angola: Revue de l’Association Française des Techniciens du Pétrole, 212, 2134.Google Scholar
Master, S., Rainaud, C., Armstrong, R. A., Phillips, D., and Robb, L. J., 2005, Provenance ages of the Neoproterozoic Katanga Supergroup (Central African Copperbelt), with implications for basin evolution: Journal of African Earth Sciences, 42, 4160.Google Scholar
Mauduit, T., Gaullier, V., and Brun, J.-P., 1998, On the asymmetry of turtle-back growth anticlines: Marine and Petroleum Geology, 19, 12191230.Google Scholar
Mauduit, T., Guerin, G., Brun, J.-P., and Lecanu, H., 1997, Raft tectonics: the effects of basal slope angle and sedimentation rate on progressive extension: Journal of Structural Geology, 14, 763771.Google Scholar
Mazurov, M. P., Grishina, S. N., Istomin, V. E., and Titov, A. T., 2007, Metasomatism and ore formation at contacts of dolerite with saliferous rocks in the sedimentary cover of the southern Siberian platform: Geology of Ore Deposits, 49, 271284.Google Scholar
McBride, B. C., 1998, The evolution of allochthonous salt along a megaregional profile across the northern Gulf of Mexico Basin: American Association of Petroleum Geologists Bulletin, 82, 10371054.Google Scholar
McBride, B. C., Rowan, M. G., and Weimer, P., 1998, The evolution of allochthonous salt systems, northern Green Canyon and Ewing Bank (offshore Louisiana), northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 82, 10131036.Google Scholar
McBride, E. F., Weidie, A. E., Wolleben, J. A., and Laudon, R. C., 1974, Stratigraphy and structure of the Parras and La Popa basins, northeastern Mexico: Geological Society of America Bulletin, 85, 16031622, doi:10.1130/0016–7606(1974)85<1603:SASOTP>2.0.CO;2.Google Scholar
McClay, K. R., 1990, Extensional fault systems in sedimentary basins: A review of analogue model studies: Marine and Petroleum Geology, 7, 206233.Google Scholar
McClay, K. R., 1995, Recent advances in analogue modeling: Uses in section interpretation and validation, in Buchanan, P. G. and Nieuwland, D. A., eds., Modern developments in structural interpretations, validation and modelling: London, Geological Society, Special Publication 99, 201225.Google Scholar
McClay, K. R., Dooley, T., Ferguson, A., and Poblet, J., 2000, Tectonic evolution of the Sanga Sanga block, Mahakam Delta, Kalimantan, Indonesia: American Association of Petroleum Geologists Bulletin, 84, 765786.Google Scholar
McClay, K. R., and Scott, A. D., 1991, Experimental models of hangingwall deformation in ramp-flat listric extensional fault systems: Tectonophysics, 188, 8596.Google Scholar
McDonnell, A., Hudec, M. R., and Jackson, M. P. A., 2009, Distinguishing salt welds from shale detachments on the inner Texas shelf, western Gulf of Mexico: Basin Research, 21, 4759, doi:10.1111/j.1365–2117.2008.00375.x.Google Scholar
McDonnell, A., Jackson, M. P. A., and Hudec, M. R., 2010, Origin of transverse folds in an extensional growth-fault setting: Evidence from an extensive seismic volume in the western Gulf of Mexico: Marine and Petroleum Geology, 27, 14941507.Google Scholar
McDowell, A. N., 1951, The origin of the structural depression above Gulf Coast salt domes with particular reference to Clay Creek dome, Washington County, Texas: M.S. thesis, Texas A&M University, College Station, Texas, 26 p.Google Scholar
McGill, G. E., Schultz, R. A., and Moore, J. M., 2000, Fault growth by segment linkage: An explanation for scatter in maximum displacement and trace length data from the Canyonlands grabens of SE Utah: Discussion: Journal of Structural Geology, 22, 135140.Google Scholar
McGill, G. E., and Stromquist, A. W., 1979, The Grabens of Canyonlands National Park, Utah: Geometry, mechanics, and kinematics: Journal of Geophysical Research, 84, 45474563.Google Scholar
McGuinness, D. B, and Hossack, J. R., 1993, The development of allochthonous salt sheets as controlled by the rates of extension, sedimentation, and salt supply: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 14th annual research conference, extended abstracts, 127–139.Google Scholar
Means, W. D., 1990, Kinematics, stress, deformation and material behavior: Journal of Structural Geology, 12, 953971.Google Scholar
Meert, J. G., and Torsvik, T. H., 2003, The making and unmaking of a supercontinent: Rodinia revisited: Tectonophysics, 375, 261288.Google Scholar
Meier, B., 1977, Zur Geologie der Klus von Balsthal–Oensingen (Ostseite): unpublished Diploma thesis, Universität Basel, Switzerland.Google Scholar
Mello, M. R., Bender, A. A., Azambuja Filho, N. C., and de Mio, E., 2011, Giant sub-salt hydrocarbon province of the greater Campos basin, Brazil: Offshore Technology Conference Brasil, Rio de Janeiro, Brazil, October 4–6, 2011, Paper OTC 22818, 9 p.Google Scholar
Mello, U. T., Karner, G. D., and Anderson, R. N., 1995, Role of salt in restraining the maturation of subsalt source rocks: Marine and Petroleum Geology, 12, 697716.Google Scholar
Melvin, J. L., ed., 1991, Evaporites, petroleum and mineral resources: Amsterdam, Elsevier, Developments in Sedimentology, 50, 556 p.Google Scholar
Merle, O., 1986, Patterns of stretch trajectories and strain rates within spreading–gliding nappes: Tectonophysics, 124, 211222.Google Scholar
Miralles, L., Sans, M., Pueyo, J. J., and Santanach, P., 2000, Recrystallization salt fabric in a shear zone (Cardona diapir, southern Pyrenees, Spain), in Vendeville, B. C., Mart, Y., and Vigneresse, J.-L., eds., Salt, shale and igneous diapirs in and around Europe: London, Geological Society, Special Publication 174, 149167.Google Scholar
Mitchell, N. C., Ligi, M., Ferrante, V., Bonatti, E., and Rutter, E., 2010, Submarine salt flows in the central Red Sea: Geological Society of America Bulletin, 122, 701713, doi:10.1130/B26518.1.Google Scholar
Modica, C. J., and Brush, E. R., 2004, Postrift sequence stratigraphy, paleogeography, and fill history of the deep-water Santos Basin, offshore southeast Brazil: American Association of Petroleum Geologists Bulletin, 88, 923945, doi:10.1306/01220403043.Google Scholar
Mohr, M., Kukla, P. A., Urai, J. L., and Bresser, G., 2005, Multiphase salt tectonic evolution in NW Germany: Seismic interpretation and retro-deformation: International Journal of Earth Sciences, 94, 917940, doi:10.1007/s00531-005-0039-5.Google Scholar
Mohr, M., Warren, J. K., Kukla, P. A., Urai, J. L., and Irmen, A., 2007, Subsurface seismic record of salt glaciers in an extensional intracontinental setting (Late Triassic of northwestern Germany): Geology, 35, 963966, doi:10.1130/G23378A.1.Google Scholar
Mohriak, W. U., Macedo, J. M., Castellani, R. T., Rangel, H. D., Barros, A. Z. N., Latgé, M. A. L., Ricci, J. A., Mizusaki, A. M. P., Szatmari, P., Demercian, L. S., Rizzo, J. G., and Aires, J. R., 1995, Salt tectonics and structural styles in the deep-water province of the Cabo Frio region, Rio de Janeiro, Brazil, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 273304.Google Scholar
Mohriak, W. U., and Szatmari, P., 2008, Tectônica de sal, in Mohriak, W. U., Szatmari, P., and Couto Anjos, S. M. (eds.), Sal geologia e tectônica: Exemplos nas bacias Brasileiras: São Paulo, Beca Edições, 91163.Google Scholar
Mohriak, W. U., Szatmari, P., and Couto Anjos, S. M., eds., 2008, Sal geologia e tectônica: Exemplos nas bacias Brasileiras: São Paulo, Beca Edições, 448 p.Google Scholar
Molinda, G. M., 1988, Investigation of methane occurrence and outbursts in the Cote Blanche domal salt mine, Louisiana: Washington, D.C., Bureau of Mines, U.S. Department of the Interior, Report of Investigations 9186, 21 p.Google Scholar
Morley, C. K., King, R., Hillis, R., Tingay, M., and Backe, G., 2010, Deepwater fold and thrust belt classification, tectonics, structure and hydrocarbon prospectivity: A review: Earth-Science Reviews, 104, 4191, doi:10.1016/j.earscirev.2010.09.010.Google Scholar
Morley, C. K., Kongwung, B., Julapour, A. A., Abdolghafourian, M., Hajian, M., Waples, D., Warren, J., Otterdoom, H., Srisuriyon, K., and Kazemi, H., 2009, Structural development of a major late Cenozoic basin and transpressional belt in central Iran: The central basin in the Qom–Saveh area: Geosphere, 5, 325362, doi:10.1130/GES00223.1.Google Scholar
Mostofi, B., and Gansser, A., 1957, The story behind the 5 Alborz: Oil and Gas Journal, 55, 7884.Google Scholar
Moulin, M., Aslanian, D., Rabineau, M., Patriat, M., and Matias, L., 2013, Kinematic keys of the Santos–Namibe basins, in Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. C., Nemčok, M., and Sinha, S. T., eds., Conjugate divergent margins: London, Geological Society, Special Publication 369, 91107.Google 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.Google Scholar
Mount, V., Dull, K., and Mentemeier, S., 2007, Structural style and evolution of traps in the Paleogene play, deepwater Gulf of Mexico, in Kennan, L., Pindell, J., and Rosen, N. C., eds., The Paleogene of the Gulf of Mexico and Caribbean basins: Processes, events, and petroleum systems: Tulsa, OK, Society of Economic Paleontologists and Mineralogists, Gulf Coast Section, 5480.Google Scholar
Mraz, D. Z. F., 1973, Behavior of rooms and pillars in deep potash mines: Canadian Institute of Mining Transactions, 76, 138143.Google Scholar
Mrazec, L., 1907, Despre cute cu simbure de străpungere [On folds with piercing cores]: Buletinul Societății de Științe din Bucureșci, 16, 68.Google Scholar
Muehlberger, W. R., 1959, Internal structure of the Grand Saline salt dome, Van Zandt County, Texas: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 38, 22 p.Google Scholar
Muehlberger, W. R., and Clabaugh, P. S., 1968, Internal structure and petrofabrics of Gulf Coast salt domes, in Braunstein, J. and O’Brien, G. D., eds., Diapirism and diapirs: Tulsa, OK, American Association of Petroleum Geologists, Memoir 8, 9098.Google Scholar
Mukherjee, S., Talbot, C. J., and Koyi, H. A., 2010, Viscosity estimates of salt in the Hormuz and Namakdan salt diapirs, Persian Gulf: Geological Magazine, 147, 497507, doi:10.1017/S001675680999077X.Google Scholar
Müller, W. H., Schmid, S. M., and Briegel, U., 1981, Deformation experiments on anhydrite rocks of different grain sizes: Rheology and microfabric: Tectonophysics, 78, 527543.Google Scholar
Nalpas, T., and Brun, J. P., 1993, Salt flow and diapirism related to extension at crustal scale: Tectonophysics, 228, 349362.Google Scholar
Nance, D., Rovik, J., and Wilcox, R. E., 1979, Lithology of the Vacherie salt dome core: A topical report: Baton Rouge, LA, Louisiana State University, Institute for Environmental Studies, 343 p.Google Scholar
Nance, D., and Wilcox, R. E., 1979, Lithology of the Rayburn’s Dome salt core: A topical report: Baton Rouge, LA, Louisiana State University, Institute for Environmental Studies, 307 p.Google Scholar
Naumann, C. F., 1858, Lehrbuch der Geognosie, Vol. 1 (2nd ed.): Leipzig, Wilhelm Engelmann.Google Scholar
Neal, J. T., Mogorian, T. R., Thoms, R. L., Autin, W. J., and McCulloh, R. P., 1993, Anomalous zones in Gulf Coast salt domes with special reference to Big Hill, TX, and Weeks Island, LA: Albuquerque, NM, Sandia National Laboratories, report SAND92-2283, 65 p.Google Scholar
Nelson, R. A., 2001, Geologic analysis of naturally fractured reservoirs (2nd ed.): Boston, Gulf Professional Publishing, 332 p.Google Scholar
Nelson, T. H., 1989, Style of salt diapirs as a function of the stage of evolution and the nature of the encasing sediments (abs.): Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 10th annual research conference program and abstracts, 109–110.Google Scholar
Nelson, T. H., 1991, Salt tectonics and listric-normal faulting, in Salvador, A., ed., The Gulf of Mexico basin: Boulder, CO, Geological Society of America, 7389.Google Scholar
Nelson, T. H., and Fairchild, L. H., 1989, Emplacement and evolution of salt sills in northern Gulf of Mexico (abs.): American Association of Petroleum Geologists Bulletin, 73, 395.Google Scholar
Nely, G., 1989, Les séries A évaporites en exploration pétrolière, 2, Méthodes géophysiques: Paris, Technip, 259 p.Google Scholar
Nettleton, L. L., and Elkins, T. A., 1947, Geologic models made from granular materials: American Geophysical Union Transactions, 28, 451466.Google Scholar
Netzeband, G. L., Hubscher, C. P., and Gajewski, D., 2006, The structural evolution of the Messinian evaporites in the Levantine Basin: Marine Geology, 230, 249273.Google Scholar
Niedoroda, A. W., Reed, C. W., Hatchett, L., Young, A., Lanier, D., Kasch, V., Jeanjean, P., Orange, D., and Bryant, W., 2003a, Analysis of past and future debris flows and turbidity currents generated by slope failures along the Sigsbee escarpment in the deep Gulf of Mexico: Offshore Technology Conference, May 5–8, 2003, Houston, Texas, paper OTC 15162.Google Scholar
Niedoroda, A. W., Reed, C. W., Hatchett, L., Jeanjean, P., Driver, D., Briaud, J.-L., and Bryant, W., 2003b, Bottom currents, deep sea furrows, erosion rates, and dating slope failure-induced debris flows along the Sigsbee escarpment in the deep Gulf of Mexico: Offshore Technology Conference, May 5–8, 2003, Houston, Texas, paper OTC 15199.Google Scholar
Nieuwland, D. A., 2003, Introduction: New insights into structural interpretation and modelling, in Nieuwland, D. A., ed., New insights into structural interpretation and modelling: London, Geological Society, Special Publication 212, 15.Google Scholar
Nikolinakou, M. A., Flemings, P. B., and Hudec, M. R., 2014, Modeling stress evolution around a rising salt diapir: Marine and Petroleum Geology, 51, 230238.Google Scholar
Nikolinakou, M. A., Luo, G., Hudec, M. R., and Flemings, P. B., 2012, Geomechanical modeling of stresses adjacent to salt bodies: Part 2 – Poroelastoplasticity and coupled overpressures: American Association of Petroleum Geologists Bulletin, 96, 6585.Google Scholar
Nilsen, K. T., Vendeville, B. C., and Johansen, J.-T., 1995, Influence of regional tectonics on halokinesis in the Nordkapp basin, Barents Sea, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 413436.Google Scholar
Norton, I. O., 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, 44, 55–58.Google Scholar
Norton, I. O., Jackson, M. P. A., and Hudec, M. R., 2009, Tectonics of passive margin salt basins: Crustal structure of the Gulf of Mexico and South Atlantic during salt deposition: American Association of Petroleum Geologists Annual Convention and Exhibition, paper 4271.Google Scholar
Nye, J. F., 1952, The mechanics of glacial flow: Journal of Glaciology, 2, 8293.Google Scholar
O’Brien, J. J., and Lerche, I., 1988, Heat flow through and around salt sheets (abs.): American Association of Petroleum Geologists Bulletin, 72, 230.Google Scholar
Ochsenius, C., 1888, On the formation of rock-salt beds and mother-liquor salts: Proceedings of the Academy of Natural Sciences of Philadelphia, 181–187.Google Scholar
Olson, J. E., Laubach, S. E., and Lander, R. H., 2009, Natural fracture characterization in tight gas sandstones: Integrating mechanics and diagenesis: American Association of Petroleum Geologists Bulletin, 93, 15351549.Google Scholar
Owen, E. W., 1964, An historical sketch of petroleum geology in the Gulf Coast region: Gulf Coast Association of Geological Societies Transactions, 14, 5165.Google Scholar
Owen, P. F., and Taylor, N. G., 1983, A salt pillow structure in the southern North Sea, in Bally, A. W., ed., Seismic expression of structural styles; A picture and work atlas: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 15, 2.3.2-7–2.3.2-10.Google Scholar
Paine, W. R., Mitchell, M. W., Copeland, R. R. Jr., and Gimbrede, L. de A., 1965, Frio and Anahuac sediment inclusions, Belle Isle salt dome, St. Mary’s Parish, Louisiana: American Association of Petroleum Geologists Bulletin, 49, 616620.Google Scholar
Parea, G. C., and Ricci-Lucchi, F., 1972, Resedimented evaporites in the Periadriatic trough (Upper Miocene, Italy): Israel Journal Earth Sciences, 21, 125141.Google Scholar
Parker, T. J., and McDowell, A. N., 1951, Scale models as guide to interpretation of salt-dome faulting: American Association of Petroleum Geologists Bulletin, 35, 20762086.Google Scholar
Parker, T. J., and McDowell, A. N., 1955, Model studies of salt-dome tectonics: American Association of Petroleum Geologists Bulletin, 39, 23842470.Google Scholar
Pascoe, R., Hooper, R., Storhaug, K., and Harper, H., 1999, Evolution of extensional styles at the southern termination of the Nordland ridge, mid-Norway: A response to variations in coupling above Triassic salt, in Fleet, A. J. and Boldy, S. A. R., eds., Petroleum geology of northwest Europe: Proceedings of the 5th Conference: London, Geological Society, 8390.Google Scholar
Passchier, C. W., and Trouw, R. A. J., 1998, Microtectonics: Berlin, Springer, 289 p.Google Scholar
Paterson, M. S., 2001, Relating experimental and geological rheology: International Journal of Earth Sciences, 90, 157167, doi:10.1007/s005310000158.Google Scholar
Pautot, G., Renard, V., Daniel, J., and Dupont, J., 1973, Morphology, limits, origin, and age of salt layer along South Atlantic African margin: American Association of Petroleum Geologists Bulletin, 57, 16581671.Google Scholar
Peach, C. J., and Spiers, C. J., 1996, Influence of crystal plastic deformation on dilatancy and permeability development in synthetic salt rock: Tectonophysics, 256, 101128.Google Scholar
Peach, C. J., Spiers, C. J., and Trimby, P. W., 2001, Effect of confining pressure on dilatation, recrystallization and flow of rocksalt at 150 °C: Journal of Geophysical Research, 106, 1331513328.Google Scholar
Peel, F. J., 2014, The engines of gravity-driven movement on passive margins: Quantifying the relative contribution of spreading vs. gravity sliding mechanisms: Tectonophysics, 633, 126142, doi:10.1016/j.tecto.2014.06.023.Google Scholar
Peel, F., Jackson, M., and Ormerod, D., 1998, Influence of major steps in the base of salt on the structural style of overlying thin-skinned structures in deep water Angola: American Association of Petroleum Geologists International Conference and Exhibition, Extended Abstracts Volume, Rio de Janeiro, Brazil, November, 366–367.Google Scholar
Peel, F. J., Travis, C. J., and Hossack, J. R., 1995, Genetic structural provinces and salt tectonics of the Cenozoic offshore U.S. Gulf of Mexico: A preliminary analysis, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective. Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 153175.Google Scholar
Pe’eri, S., Zebker, H. A., Ben-Avraham, Z., Frumkin, A., and Hall, J. K., 2004, Spatially-resolved uplift rate of the Mount Sedom (Dead Sea) salt diapir from InSAR observations: Israel Journal of Earth Sciences, 53, 99106.Google Scholar
Penge, J., Taylor, B., Huckerby, J. A., and Munns, J. W., 1993, Extension and salt tectonics in the East Central graben, in Geological Society, London, Petroleum Geology Conference series, 4, 11971209, doi:10.114/0041197.Google Scholar
Penge, J., Taylor, B., and Munns, J., 1999, Rift–raft tectonics: examples of gravitational sliding structures from the Zechstein basins of northwest Europe, in Fleet, A. J. and Boldy, S. A. R., eds., Petroleum geology of northwest Europe: Proceedings of the 5th Conference: London, Geological Society, 201213.Google Scholar
Pepper, A. S., and Yu, Z., 1995, Influence of an inclined salt sheet on petroleum emplacement in the Pompano field area, offshore Gulf of Mexico, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference proceedings, 197–205.Google Scholar
Pequeno, M. A., 2009, Albian/Maastrichtian tectono-stratigraphic evolution of central Santos Basin, offshore Brazil: Ph.D. dissertation, The University of Texas at Austin, Austin, Texas, 219 p.Google Scholar
Perez, M. A., Clyde, R., D’Ambrosio, P., Israel, R., Leavitt, T., Nutt, L., Johnson, C., and Williamson, D., 2008, Meeting the subsalt challenge: Oilfield Review, 20, 3245.Google Scholar
Peron-Pinvidic, G., Manatschal, G., and Osmundsen, P. T., 2013, Structural comparison of archetypal Atlantic rifted margins: A review of observations and concepts: Marine and Petroleum Geology, 43, 2147, doi:10.1016/j.marpetgeo.2013.02.002.Google Scholar
Pilcher, R. S., and Blumstein, R. D., 2007, Brine volume and salt dissolution rates in Orca Basin northeast Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 91, 823833, doi:10.1306/12180606049.Google Scholar
Pilcher, R. S., Kilsdonk, B., and Trude, J., 2011, Primary basins and their boundaries in the deep-water northern Gulf of Mexico: Origin, trap types, and petroleum system implications: American Association of Petroleum Geologists Bulletin, 95, 219240, doi:10.1306/06301010004.Google Scholar
Pinto, L., Muñoz, C., Nalpas, T., and Charrier, R., 2010, Role of sedimentation during basin inversion in analogue modelling: Journal of Structural Geology, 32, 554565.Google Scholar
Pirazzoli, P., Reyss, J., Fontugne, M., Haghipour, A., Hilgers, A., Kasper, H., Nazari, H., Preusser, F., and Radtke, U., 2004, Quaternary coral-reef terraces from Kish and Qeshm islands, Persian Gulf: New radiometric ages and tectonic implications: Quaternary International, 120, 1527.Google Scholar
Player, R. A., 1969, The Hormuz salt plugs of southern Iran: Ph.D. thesis, University of Reading.Google Scholar
Poblet, J., and McClay, K., 1996, Geometry and kinematics of single-layer detachment folds: American Association of Petroleum Geologists Bulletin, 80, 10851109.Google Scholar
Poblet, J., McClay, K., Storti, F., and Munoz, J. A., 1997, Geometries of syntectonic sediments associated with single-layer detachment folds: Journal of Structural Geology, 19, 369381.Google Scholar
Pollard, A. M., Brothwell, D. R., Aali, A., Buckley, S., Fazeli, H., Dehkordi, M. H., Holden, T., Jones, A. K. G., Shokouhi, J. J., Vatandoust, R., and Wilson, A. S., 2008, Below the salt: A preliminary study of the dating and biology of five preserved bodies from Zanjan Province, Iran: British Institute of Persian Studies, 46, 135150.Google Scholar
Pollard, D. D., and Fletcher, R. C., 2005, Fundamentals of structural geology: Cambridge, Cambridge University Press, 500 p.Google Scholar
Popp, T., Kern, H., and Schulze, O., 2001, Evolution of dilatancy and permeability in rock salt during hydrostatic compaction and triaxial deformation: Journal of Geophysical Research, 106, 40614078.Google Scholar
Pošepný, F., 1871, Studien aus dem Salinargebiete Siebenbürgens: Kaiserlich-Königlichen Geologischen Reichsanstalt Jahrbuch, 21, 123186.Google Scholar
Posey, H. H., and Kyle, J. R., 1988, Fluid–rock interactions in the salt dome environment: An introduction and review: Chemical Geology, 74, 124.Google Scholar
Prather, B. E., 2000, Calibration and visualization of depositional process models for above-grade slopes: a case study from the Gulf of Mexico: Marine and Petroleum Geology, 17, 619638.Google Scholar
Prather, B. E., 2003, Controls on reservoir distribution, architecture and stratigraphic trapping in slope settings: Marine and Petroleum Geology, 20, 529545, doi:10.1016/j.marpetgeo.2003.03.009.Google Scholar
Prather, B. E., Booth, J. R., Steffens, G. S., Craig, P. A., 1998, Classification, lithologic calibration, and stratigraphic succession of seismic facies of intraslope basins, deep-water Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 82, 701728.Google Scholar
Pratson, L. F., and Ryan, W. B. F., 1994, Pliocene to recent infilling and subsidence of intraslope basins offshore Louisiana: American Association of Petroleum Geologists Bulletin, 78, 14831506.Google Scholar
Pray, L. C., and Murray, R. C., eds., 1965, Dolomitization and limestone diagenesis: A symposium: Tulsa, OK, Society of Economic Paleontologists and Mineralogists, Special Publication 13, 180 p.Google Scholar
Preiss, W. V., 1985, Stratigraphy and tectonics of the Worumba anticline and associated intrusive breccias: Geological Survey of South Australia Bulletin, 52, 85 p.Google Scholar
Price, N. J., and Cosgrove, J. W., 1990, Analysis of geological structures: Cambridge, Cambridge University Press, 502 p.Google Scholar
Price, R. H., 1982, Effects of anhydrite and pressure on the mechanical behavior of synthetic rocksalt: Geophysical Research Letters, 9, 10291032.Google Scholar
Prouvost, P., 1930, Sédimentation et subsidence, reprinted in Centenaire de la Société Géologique de France Livre Jubilaire, 1820–1990: Paris, Société Géologique de France, 2, 545564.Google Scholar
Puigdefabregas, C., Muñoz, J. A., and Vergés, J., 1992, Thrusting and foreland basin evolution in the Southern Pyrenees, in McClay, K. R., ed., Thrust tectonics: London, Chapman & Hall, 247254.Google Scholar
Purser, B. H., 1973, Sedimentation around bathymetric highs in the southern Persian Gulf, in Purser, B. H., ed., The Persian Gulf: Holocene carbonate sedimentation and diagenesis in a shallow epicontinental sea: Berlin, Springer, 157177.Google Scholar
Quirk, D. G., Hertle, M., Jeppesen, J. W., Raven, M., Mohriak, W. U., Kann, D. J., Norgaard, M., Howe, M. J., Hsu, D., Coffey, B., and Mendes, M. P., 2013, Rifting, subsidence and continental break-up above a mantle plume in the central South Atlantic, in Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. C., Nemčok, M., and Sinha, S. T., eds., Conjugate divergent margins: London, Geological Society, Special Publication 369, doi:10.1144/SP369.20.Google Scholar
Quirk, D. G., and Pilcher, R. S., 2012, Flip-flop salt tectonics, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 245264.Google Scholar
Quirk, D. G., Schodt, N., Lassen, B., Ings, S. J., Hsu, D., Hirsch, K. K., and Von Nicolai, C., 2012, Salt tectonics on passive margins: Examples from the Santos, Campos and Kwanza Basins, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 207244.Google Scholar
Ramberg, H., 1981a, Gravity, deformation and the Earth’s crust in theory, experiments and geological application (2nd ed.): London, Academic Press, 452 p.Google Scholar
Ramberg, H., 1981b, The role of gravity in orogenic belts, in McClay, K. R. and Price, N. J., eds., Thrust and nappe tectonics: London, Geological Society, Special Publication 9, 125140.Google Scholar
Ramsay, J. G., 1967, Folding and fracturing of rocks: New York, McGraw-Hill, 568 p.Google Scholar
Ramsay, J. G., and Graham, R. H., 1970, Strain variation in shear belts: Canadian Journal of Earth Sciences, 7, 786813.Google Scholar
Ranalli, G., and Murphy, D. C., 1987, Rheological stratification of the lithosphere: Tectonophysics, 132, 281295.Google Scholar
Raup, O. B., 1982, Gypsum precipitation by mixing seawater brines: American Association of Petroleum Geologists Bulletin, 66, 363367.Google Scholar
Reilly, M. J., and Flemings, P. B., 2010, Deep pore pressures and seafloor venting in the Auger Basin, Gulf of Mexico: Basin Research, 22, 380397, doi:10.1111/j.1365–2117.2010.00481.x.Google Scholar
Rettger, R. E., 1935, Experiments on soft-rock deformation: American Association of Petroleum Geologists Bulletin, 19, 271292.Google Scholar
Reyer, E., 1888, Theoretische Geologie: Stuttgart, E. Schweizerbart’sche Verlagshandlung.Google Scholar
Reynolds, T. D., and Gloyna, E. F., 1960, Reactor fuel waste disposal project: Permeability of rock salt and creep of underground salt cavities: Austin, TX, The University of Texas, Department of Civil Engineering, Sanitary Engineering Research Laboratory, Report, 121 p.Google Scholar
Reyss, J. L., Pirazzoli, P. A., Haghipour, A., Hatté, C., and Fontugne, M., 1998, Quaternary marine terraces and tectonic uplift rates on the south coast of Iran, in Stewart, I. S. and Vita-Finzi, C., eds., Coastal tectonics: London, Geological Society, Special Publication 146, 225237.Google Scholar
Richard, P., 1991, Experiments on faulting in a two-layer cover sequence overlying a reactivated basement fault with oblique-slip: Journal of Structural Geology, 13, 459469.Google Scholar
Richard, P., Mocquet, B., and Cobbold, P., 1991, Experiments on simultaneous faulting and folding above a basement wrench fault: Tectonophysics, 188, 133141.Google Scholar
Richter-Bernburg, G., 1950, Zur Frage der absoluten Geschwindigkeit geologischer Vorgänge: Naturwissenschaften, 37, 18.Google Scholar
Richter-Bernburg, G., 1955, Über salinare Sedimentation: Zeitschrift der Deutschen Geologischen Gesellschaft, 105, 593645.Google Scholar
Richter-Bernburg, G., 1970, Geologische Voraussetzungen für die Anlage von Rohöl-Speichern in Salz-Kavernen: Zeitschrift OEL, 8, 209212.Google Scholar
Richter-Bernburg, G., 1980, Salt tectonics, interior structures of salt bodies: Bulletin des Centres de Recherches Exploration–Production Elf-Aquitaine, 4, 373393.Google Scholar
Ringenbach, J.-C., Salel, J. F., Kergaravat, C., and Ribes, C., 2013, Salt tectonics in the Sivas Basin, Turkey: Outstanding seismic analogues from outcrops: First Break, 31, 93101.Google Scholar
Roberts, A. M., and Yielding, G., 1991, Deformation around basin–margin faults in the North Sea/mid-Norway rift, in Roberts, A. M., Yielding, G., and Freeman, B., eds., The geometry of normal faults: London, Geological Society, Special Publication 56, 6178.Google Scholar
Roca, E., Anadon, P., Utrilla, R., and Vazquez, A., 1996, Rise, closure and reactivation of the Bicorb–Quesa diapir, eastern Prebetics, Spain: Journal of the Geological Society, London, 153, 311321.Google Scholar
Roca, E., Beamud, E., Rubinat, M., Soto, R., and Ferrer, O., 2013, Paleomagnetic and inner diapiric structural constraints on the kinematic evolution of a salt-wall: The Bicorb–Quesa and northern Navarrés salt-wall segments case (Prebetic Zone, SE Iberia): Journal of Structural Geology, 52, 8095.Google Scholar
Roca, E., Sans, M., and Koyi, H. A., 2006, Polyphase deformation of diapiric areas in models and in the eastern Prebetics (Spain): American Association of Petroleum Geologists Bulletin, 90, 115136.Google Scholar
Roedder, E., 1984, The fluids in salt: American Mineralogist, 69, 413439.Google Scholar
Roedder, E., and Belkin, H. E., 1979, Fluid inclusions in salt from the Rayburn and Vacherie domes, Louisiana: Washington, D.C., U.S. Geological Survey, Open-File Report 79-1675, 25 p.Google Scholar
Rogers, G. S., 1918, Intrusive origin of the Gulf Coast salt domes: Economic Geology, 13, 447485.Google Scholar
Ross, D. A., and Uchupi, E., 1973, Shallow structure and geologic development of the southern Red Sea: Geological Society of America Bulletin, 84, 38273848.Google Scholar
Ross, J. V., and Bauer, S. J., 1992, Semi-brittle deformation of anhydrite–halite shear zones simulating mylonite formation: Tectonophysics, 213, 303320Google Scholar
Rouby, D., Cobbold, P. R., Szatmari, P., Demercian, S., Coelho, D., and Rici, J. A., l993a, Least-squares palinspastic restoration of regions of normal faulting: Application to the Campos basin (Brazil): Tectonophysics, 221, 439452.Google Scholar
Rouby, D., Cobbold, P. R., Szatmari, P., Demercian, S., Coelho, D., and Rici, J. A., 1993b, Restoration in plan view of faulted Upper Cretaceous and Oligocene horizons and its bearing on the history of salt tectonics in the Campos basin (Brazil): Tectonophysics, 228, 435445.Google Scholar
Rouby, D., Guillocheau, N. F., Robin, N. C., Bouroullec, W. R., Raillard, Z. S., Castelltort, S., and Nalpas, T., 2003, Rates of deformation of an extensional growth fault/raft system (offshore Congo, West African margin) from combined accommodation measurements and 3-D restoration: Basin Research, 15, 183200.Google Scholar
Rouby, D., Raillard, S., Guillocheau, F., Bouroullec, R., and Nalpas, T., 2002, Kinematics of a growth/raft system on the West African margin using 3-D restoration: Journal of Structural Geology, 24, 783796.Google Scholar
Rouby, D., Suppe, J., and Xiao, H., 2000, 3-D restoration of complexly faulted and folded surfaces using multiple unfolding mechanisms: American Association of Petroleum Geologists Bulletin, 84, 805829.Google Scholar
Rouvier, H., Perthuisot, V., and Mansouri, A., 1985, Pb–Zn deposits and salt-bearing diapirs in southern Europe and north Africa: Economic Geology, 80, 666687.Google Scholar
Roveri, M., Manzi, V., Lucci, F. R., and Rogledi, S., 2003, Sedimentary and tectonic evolution of the Vena del Gesso basin (northern Appenines, Italy): Implications for the onset of the Messinian salinity crisis: Geological Society of America Bulletin, 115, 387405.Google Scholar
Rowan, M. G., 1995, Structural styles and evolution of allochthonous salt, central Louisiana outer shelf and upper slope, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 199228.Google Scholar
Rowan, M. G., 2002, Salt-related accommodation in the Gulf of Mexico deepwater: withdrawal or inflation, autochthonous or allochthonous?: Gulf Coast Association of Geological Societies Transactions, 52, 861869.Google Scholar
Rowan, M. G., 2004, Do salt welds seal?, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 390–403.Google Scholar
Rowan, M. G., 2014, Passive-margin salt basins: Hyperextension, evaporite deposition, and salt tectonics: Basin Research, 26, 154182, doi:10.1111/bre.12043.Google Scholar
Rowan, M. G., and Inman, K. F., 2011, Salt-related deformation recorded by allochthonous salt rather than growth strata: Gulf Coast Association of Geological Societies Transactions, 61, 379390.Google Scholar
Rowan, M. G., Jackson, M. P. A., and Trudgill, B. D., 1999, Salt-related fault families and fault welds in the northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 83, 14541484.Google Scholar
Rowan, M. G., Lawton, T. F., and Giles, K. A., 2012, Anatomy of an exposed vertical salt weld and flanking strata, La Popa basin, Mexico, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodgkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 3357, doi:10.1144/SP363.3.Google Scholar
Rowan, M. G., Lawton, T. F., Giles, K. A., and Ratliff, R. A., 2003, Near-salt deformation in La Popa basin, Mexico, and the northern Gulf of Mexico: A general model for passive diapirism: American Association of Petroleum Geologists Bulletin, 87, 733756.Google Scholar
Rowan, M. G., Peel, F. J., and Vendeville, B. C., 2004, Gravity-driven fold belts on passive margins, in McClay, K. R., ed., Thrust tectonics and hydrocarbon systems: Tulsa, OK, American Association of Petroleum Geologists, Memoir 82, 157182.Google Scholar
Rowan, M. G., and Ratliff, R. A., 2012, Cross-section restoration of salt-related deformation: Best practices and potential pitfalls: Journal of Structural Geology, 41, 2437, doi:10.1016/j.jsg.2011.12.012.Google Scholar
Rowan, M. G., Ratliff, R. A., Trudgill, B. D., and Duarte, J. B., 2001, Emplacement and evolution of the Mahogany salt body, central Louisiana outer shelf, northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 85, 947969.Google Scholar
Rowan, M. G., Trudgill, B. D., and Fiduk, J. C., 2000, Deep-water, salt-cored foldbelts: Lessons from the Mississippi Fan and Perdido foldbelts, northern Gulf of Mexico, in Mohriak, W. and Talwani, M., eds., Atlantic rifts and continental margins: Washington, D.C., American Geophysical Union, Geophysical Monograph 115, 173191.Google Scholar
Rowan, M. G., and Vendeville, B. C., 2006, Foldbelts with early salt withdrawal and diapirism: Physical model and examples from the northern Gulf of Mexico and the Flinders Ranges, Australia: Marine and Petroleum Geology, 23, 871891.Google Scholar
Roy, A., and Chazalnoel, N., 2011, RTM technology for improved salt imaging in the Santos Basin, Brazil: Society of Exploration Geophysicists Annual Meeting, San Antonio, 5 p.Google Scholar
Rubinat, M., Ledo, J., Roca, E., Rosell, O., and Queralt, P., 2010, Magnetotelluric characterization of a salt diapir: A case study on Bicorb–Quesa Diapir (Prebetic Zone, SE Spain): Journal of the Geological Society, London, 167, 145153.Google Scholar
Sadler, P. M., 1993, Time scale dependence of the rates of unsteady geologic processes, in Rates of geologic processes: Tectonics, sedimentation, eustasy, and climate: Implications for hydrocarbon exploration: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 14th annual research conference, 221–228.Google Scholar
Salvador, A., 1987, Late Triassic–Jurassic paleogeography and origin of Gulf of Mexico basin: American Association of Petroleum Geologists Bulletin, 71, 419451.Google Scholar
Samec, P., de Graciansky, P.-C., and Rudkiewicz, J.-L., 1988, Tectonique distensive et halocinèse d’âge jurassique: la zone subbriançonnaise en Maurienne (Savoie): Bulletin de la Société Géologique de France, Paris, 4, 659667.Google Scholar
Sanderson, D. J., and Marchini, W., 1984, Transpression: Journal of Structural Geology, 6, 449458.Google Scholar
Sani, F., Del Ventisette, C., Montanari, D., Bendkik, A., and Chenakeb, M., 2007, Structural evolution of the Rides Prerifaines (Morocco): Structural and seismic interpretation and analogue modelling experiments: International Journal of Earth Sciences, 96, 685706.Google Scholar
Sannemann, D., 1968, Salt-stock families in northwestern Germany, in Braunstein, J. and O’Brien, G. D., eds., Diapirism and diapirs: Tulsa, OK, American Association of Petroleum Geologists, Memoir 8, 261270.Google Scholar
Sans, M., 2003, From thrust tectonics to diapirism; The role of evaporites in the kinematic evolution of the eastern South Pyrenean front: Geologica Acta, 1, 239259.Google Scholar
Sans, M., and Koyi, H. A., 2001, Modeling the role of erosion in diapir development in contractional settings, in Koyi, H. A. and Mancktelow, N. S., eds., Tectonic modeling: A volume in honor of Hans Ramberg: Boulder, CO, Geological Society of America, Memoir 193, 111122.Google Scholar
Sans, M., Sánchez, A. L., and Santanach, P., 1996, Internal structure of a detachment horizon in the most external part of the Pyrenean fold and thrust belt (northern Spain), in Alsop, G. I., Blundell, D. J., and Davison, I., eds., Salt tectonics: London, Geological Society, Special Publication 100, 6576.Google Scholar
Sans, M., and Vergés, J., 1995, Fold development related to contractional salt tectonics: southeastern Pyrenean thrust front, Spain, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 369378.Google Scholar
Sarkar, A., Nunn, J. A., and Hanor, J. S., 1995, Free thermohaline convection beneath allochthonous salt sheets: An agent for salt tectonics and fluid flow in Gulf Coast sediments, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference, 245–252.Google Scholar
Sawtelle, G., 1936, Salt-dome statistics: American Association of Petroleum Geologists Bulletin, 20, 726735.Google Scholar
Schachl, E., 1968, Mine Mariaglück, Höfer, Excursion A21, in International symposium on geology of saline deposits, Hannover.Google Scholar
Schardt, H., 1893, Sur l’origine des préalpes romandes: Eclogae Geologicae Helvetiae, 4, 129142.Google Scholar
Schauberger, O., 1986, Bau und Bildung der Salzlagerstätten des ostalpinen Salinars: Archiv für Lagerstättenforschung der Geologischen Bundesanstalt, 7, 217254.Google Scholar
Schenk, O., and Urai, J. L., 2005, The migration of fluid-filled grain boundaries in recrystallizing synthetic bischofite: First results of in-situ high-pressure, high-temperature deformation experiments in transmitted light: Journal of Metamorphic Geology, 23, 695709.Google Scholar
Schenk, O., Urai, J. L., and Piazolo, S., 2006, Structure of grain boundaries in wet, synthetic polycrystalline, statically recrystallizing halite – Evidence from cryo-SEM observations: Geofluids, 6, 93104.Google Scholar
Schléder, Z., Burliga, S., and Urai, J. L., 2007, Dynamic and static recrystallization-related microstructures in halite samples from the Kodawa salt wall (central Poland) as revealed by gamma-irradiation: Neues Jahrbuch für Mineralogie-Abhandlungen, 184, 1728.Google Scholar
Schléder, Z., and Urai, J. L., 2005, Microstructural evolution of deformation-modified primary halite from the Middle Triassic Röt Formation at Hengelo, The Netherlands: International Journal of Earth Sciences, 94, 941956, doi:10.1007/s00531-005–0503-2.Google Scholar
Schléder, Z., and Urai, J. L., 2007, Deformation and recrystallization mechanisms in mylonitic shear zones in naturally deformed extrusive Eocene–Oligocene rocksalt from Eyvanekey plateau and Garmsar hills (central Iran): Journal of Structural Geology, 29, 241255, doi:10.1016/j.jsg.2006.08.014.Google Scholar
Schléder, Z., Urai, J. L., Nollet, S., and Hilgers, C., 2008, Solution–precipitation creep and fluid flow in halite: A case study in Zechstein (Z1) rocksalt from Neuhof salt mine (Germany): International Journal of Earth Science, 97, 10451056.Google Scholar
Schlische, R. W., 1995, Geometry and origin of fault-related folds in extensional settings: American Association of Petroleum Geologists Bulletin, 79, 16611678.Google Scholar
Schlumberger and TGS-NOPEC, 2000, Advertisement in AAPG Explorer, American Association of Petroleum Geologists, 21, no. 3, 89.Google Scholar
Schmeling, H., Cruden, A., and Marquart, G., 1988, Finite deformation in and around a fluid sphere moving through a viscous medium: Implications for diapiric ascent: Tectonophysics, 149, 1734.Google Scholar
Schmid, S. M., Boland, J. N., and Paterson, M. S., 1977, Superplastic flow in fine-grained limestone: Tectonophysics, 43, 257291.Google Scholar
Schoell, M., Backer, H., and Baumann, Q., 1974, The Red Sea geothermal system: New aspects on their brines and associated sediments, in Problems of ore deposition: 4th Symposium of International Association on the Genesis of Ore Deposits, Varna, Vol. 1, 303–314.Google Scholar
Schoenherr, J., Reuning, L., Kukla, P. A., Littke, R., Urai, J. L., Siemann, M., and Rawahi, Z., 2009, Halite cementation and carbonate diagenesis of intra-salt reservoirs from the late Neoproterozoic to Early Cambrian Ara Group (South Oman Salt Basin): Sedimentology, 56, 567589, doi:10.1111/j.1365–3091.2008.00986.x.Google Scholar
Schoenherr, J., Schléder, Z., Urai, J. L., Fokker, P. A., and Schulze, O., 2007b, Deformation mechanisms and rheology of Pre-cambrian rocksalt from the south Oman salt basin, in Wallner, M., Lux, K.-H., Minkley, W., and Hardy, H., eds., The mechanical behavior of salt – Understanding of THMC processes in salt: Basingstoke, Taylor & Francis Group, 167173.Google Scholar
Schoenherr, J., Urai, J. L., Kukla, P. A., Littke, R., Schléder, Z., Larroque, J.-M., Newall, M. J., Al-Abry, N., Al-Siyabi, H. A., and Rawahi, Z., 2007a, Limits to the sealing capacity of halite: A case study of the infra-Cambrian Ara Salt from the south Oman salt basin: American Association of Petroleum Geologists Bulletin, 91, 15411557, doi:10.1306/06200706122Google Scholar
Schowalter, T. T., 1979, Mechanics of secondary hydrocarbon migration and entrapment: American Association of Petroleum Geologists Bulletin, 63, 723760.Google Scholar
Schreiber, B. C., ed., 1988, Evaporites and hydrocarbons: New York, Columbia University Press, 475 p.Google Scholar
Schreiber, B. C., Friedman, G. M., Decima, A., and Schreiber, E., 1976, Depositional environments of Upper Miocene (Messinian) evaporite deposits of the Sicilian basin: Sedimentology, 23, 729760.Google Scholar
Schreiber, B. C., and Helman, M. L., 2005, Criteria for distinguishing primary evaporite features from deformation features in sulfate evaporates: Journal of Sedimentary Research, 75, 525533, doi:10.2110/jsr.2005.043.Google Scholar
Schreiber, B. C., and Hsu, K. J., 1980, Evaporites, in Hobson, G. D., ed., Developments in petroleum geology: London, Applied Science Publishers, 2, 87138.Google Scholar
Schreiber, B. C., Lugli, S., and Bąbel, M., eds., 2007, Evaporites through space and time: London, Geological Society, Special Publication 285, 384 p.Google Scholar
Schuh, W., Leveille, R., Fay, I., and North, R., 2012, Geology of the Tenke-Fungurume sediment-hosted strata-bound copper–cobalt district, Katanga, Democratic Republic of Congo, in Hedenquist, J. W., Harris, M., and Camus, F., eds., Geology and genesis of major copper deposits and districts of the world: Littleton, CO, Society of Economic Geologists, Special Publication 16, 269301.Google Scholar
Schultz-Ela, D. D., 1992, Restoration of cross-sections to constrain deformation processes of extensional terranes: Marine and Petroleum Geology, 9, 372388.Google Scholar
Schultz-Ela, D. D., 2001, Excursus on gravity gliding and gravity spreading: Journal of Structural Geology, 23, 725731.Google Scholar
Schultz-Ela, D., 2003, Origin of drag folds bordering salt diapirs: American Association of Petroleum Geologists Bulletin, 87, 757780.Google Scholar
Schultz-Ela, D. D., Jackson, M. P. A., and Vendeville, B. C., 1993, Mechanics of active salt diapirism: Tectonophysics, 228, 275312.Google Scholar
Schultz-Ela, D. D., and Walsh, P., 2002, Modeling of grabens extending above evaporites in Canyonlands National Park, Utah: Journal of Structural Geology, 24, 247275.Google Scholar
Schuster, D. C., 1995, Deformation of allochthonous salt and evolution of related salt-structural systems, eastern Louisiana Gulf Coast, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 177198.Google Scholar
Schutjens, P. M. T. M., and Spiers, C. J., 1999, Intergranular pressure solution in NaCl: Grain-to-grain contact experiments under the optical microscope: Oil & Gas Science and Technology – Revue de l’IFP, 54, 729750.Google Scholar
Schwerdtner, W. M., 1967, Intragranular gliding in domal salt: Tectonophysics, 5, 353380.Google Scholar
Schwerdtner, W. M., 1982, Salt stocks as natural analogues of Archaean gneiss diapirs: Geologische Rundschau, 71, 370379.Google Scholar
Schwerdtner, W. M., and Clark, A. R., 1967, Structural analysis of Mokka Fiord and South Fiord domes, Axel Heiberg Island, Canadian Arctic: Canadian Journal of Earth Sciences, 4, 12291245.Google Scholar
Schwerdtner, W. M., Torrance, J. G., and van Berkel, J. T., 1989, Pattern of apparent total strain in the bedded anhydrite cap of a folded salt wall: Canadian Journal of Earth Sciences, 26, 983992.Google Scholar
Schwerdtner, W. M., and Troëng, B., 1978, Strain distribution within arcuate diapiric ridges of silicone putty: Tectonophysics, 50, 1328.Google Scholar
Scrope, G. P., 1825, Considerations on volcanoes: London, W. Phillips, 270 p.Google Scholar
Seni, S. J., and Jackson, M. P. A., 1983a, Evolution of salt structures, East Texas diapir province, Part 1: sedimentary record of halokinesis: American Association of Petroleum Geologists Bulletin, 67, 12191244.Google Scholar
Seni, S. J., and Jackson, M. P. A., 1983b, Evolution of salt structures, East Texas diapir province, Part 2: patterns and rates of halokinesis: American Association of Petroleum Geologists Bulletin, 67, 12451274.Google Scholar
Seni, S. J., and Jackson, M. P. A., 1984, Sedimentary record of Cretaceous and Tertiary salt movement, East Texas basin: Times, rates, and volumes of salt flow and their implications for nuclear waste isolation and petroleum exploration: Austin, TX, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 139, 89 p.Google Scholar
Senseny, P. E., Hansen, F. D., Russell, J. E., Carter, N. L., and Handin, J. W., 1992, Mechanical behaviour of rock salt: Phenomenology and micromechanisms: International Journal of Rock Mechanics and Mining Sciences and Geomechanical Abstracts, 29, 363378.Google Scholar
Sepehr, M., and Cosgrove, J., 2005, Role of the Kazerun fault zone in the formation and deformation of the Zagros fold–thrust belt, Iran: Tectonics, 24, TC5005, doi:10.1029/2004TC001725.Google Scholar
Serata, S., and Gloyna, E. F., 1959, Reactor fuel waste disposal project, development of design principle for disposal of reactor fuel waste into underground salt cavities: Austin, TX, The University of Texas, Department of Civil Engineering, Sanitary Engineering Research Laboratory, Technical Report to Atomic Energy Commission, 173 p.Google Scholar
Shaw, J. H., Connors, C., and Suppe, J., eds., 2005, Seismic interpretation of contractional fault-related folds: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 53, 156 p.Google Scholar
Shen, H., Chen, G., and Li, T., 2012, Salt emplacement-induced azimuthal anisotropy in Garden Banks, Gulf of Mexico, in Society of Exploration Geophysicists Annual Meeting, November 4–9, Las Vegas, Nevada, 5 p., doi:10.1190/segam2012-0533.1.Google Scholar
Sheriff, R. E., 1984, Encyclopedic dictionary of exploration geophysics (2nd ed.): Tulsa, OK, Society of Exploration Geophysicists, 323 p.Google Scholar
Sherkati, S., and Letouzey, J., 2004, Variation of structural style and basin evolution in the central Zagros (Izeh zone and Dezful embayment), Iran: Marine and Petroleum Geology, 21, 535554.Google Scholar
Sherkati, S., Letouzey, J., and Frizon de Lamotte, D., 2006, Central Zagros fold–thrust belt (Iran): New insights from seismic data, field observations and sandbox modeling: Tectonics, 25, TC4007, doi:10.1029/2004TC001766.Google Scholar
Sherkati, S., Molinaro, M., Frizon de Lamotte, D., and Letouzey, J., 2005, Detachment folding in the central and eastern Zagros fold-belt (Iran): Salt mobility, multiple detachments and late basement control: Journal of Structural Geology, 27, 16801696, doi:10.1016/j.jsg.2005.05.010.Google Scholar
Sherwood, J. W. C., Sherwood, K., Tieman, H., and Schleicher, K., 2008, 3D beam prestack depth migration with examples from around the world, in Society of Exploration Geophysicists Annual Meeting, November 9–14, Las Vegas, Nevada, 438–442.Google Scholar
Shimeld, J., 2004, A comparison of salt tectonic subprovinces beneath the Scotian slope and Laurentian fan, in Salt–sediment interactions and hydrocarbon prospectivity: Concepts, applications, and case studies for the 21st century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 24th annual research conference, 502–532.Google Scholar
Shoemaker, E. M., 1956, Geology of the Roc Creek quadrangle, Colorado: Reston, VA, U.S. Geological Survey, Quadrangle Map GQ-83, scale 1:24,000, 1 sheet.Google Scholar
Shokes, R. F., Trabant, P. K., Presley, B. J., and Reid, D. F., 1977, Anoxic, hypersaline basin in the northern Gulf of Mexico: Science, 196, 14431446.Google Scholar
Sibson, R. H., 1981, Fluid flow accompanying faulting: field evidence and models, in Simpson, D. W. and Richards, P. G., eds., Earthquake prediction: Washington, D.C., American Geophysical Union, 593603, doi:10.1029/ME004p0593.Google Scholar
Simmons, G. R., 1992, The regional distribution of salt in the northwestern Gulf of Mexico: Styles of emplacement and implications for early tectonic history: Ph.D. dissertation, Texas A&M University, College Station, Texas, 180 p.Google Scholar
Sims, D. W., Morris, A. P., Wyrick, D. Y., Ferrill, D. A., Waiting, D. J., Franklin, N. M., Colton, S. L., Umezawa, Y. T., Takanashi, M., and Beverly, E. J., 2013, Analog modeling of normal faulting above Middle East domes during regional extension: American Association of Petroleum Geologists Bulletin, 97, 877898, doi:10.1306/02101209136.Google Scholar
Smit, J., Brun, J.-P., Cloetingh, S., and Ben-Avraham, Z., 2008a, Pull-apart basin formation and development in narrow transform zones with application to the Dead Sea basin: Tectonics, 27, TC6018, doi:10.1029/2007TC002119.Google Scholar
Smit, J., Brun, J. P., Fort, X., Cloetingh, S., and Ben-Avraham, Z., 2008b, Salt tectonics in pull-apart basins with application to the Dead Sea basin: Tectonophysics, 449, 116, doi:10.1016/j.tecto.2007.12.004.Google Scholar
Smith, A. P., Fischer, M. P., and Evans, M. A., 2012, Fracture-controlled palaeohydrology of a secondary salt weld, La Popa basin, NE Mexico, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodhkinson, R., eds., Salt tectonics, sediments and prospectivity: London, Geological Society, Special Publication 363, 107130, doi:10.1144/SP363.6.Google Scholar
Smith, R., 2004, Silled sub-basins to connected tortuous corridors: Sediment distribution systems on topographically complex sub-aqueous slopes, in Lomas, S. A. and Joseph, P., eds., Confined turbidite systems: London, Geological Society, Special Publication 222, 2343, doi:10.1144/GSL.SP.2004.222.01.03.Google Scholar
Smith, R. I., Hodgson, N., and Fulton, M., 1993, Salt control on Triassic reservoir distribution, UKCS central North Sea, in Parker, J. R., ed., Petroleum geology of northwest Europe: Proceedings of the 4th Conference: London, Geological Society, 547557.Google Scholar
Smith, S. W., 1988, Tiger Shoal field, offshore Louisiana, in Studlick, J. R. J., Bryant, J. G., Hartman, J. A., and Shew, R. D., eds., Offshore Louisiana oil and gas fields: New Orleans, LA, New Orleans Geological Society, 2, 173178.Google Scholar
Smoluchowski, M. S., 1909, Some remarks on the mechanics of overthrusts: Geological Magazine, 6, 204205.Google Scholar
Snider-Pellegrini, A., 1858: La Création et ses mystères dévoilés: Paris, A. Franck et E. Dentu, 48 p.Google Scholar
Snyder, F. C., and Nugent, J. A., 1995, Teak – Testing a subsalt hydrocarbon trap geometry, South Timbalier block 260, Gulf of Mexico, in Salt, sediment and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference program with papers, 257–267.Google Scholar
Snyder, J. D., and Dellwig, L. F., 1961, Plastic flowage of salt in mines at Hutchinson and Lyons: Kansas State Geological Survey Bulletin, 152, 3146.Google Scholar
Sommaruga, A., 1999, Décollement tectonics in the Jura foreland fold-and-thrust belt: Marine and Petroleum Geology, 15, 111134.Google Scholar
Sonnenfeld, P., 1985, Evaporites as oil and gas source rocks: Journal of Petroleum Geology, 8, 253271.Google Scholar
Spencer, J., Jeronimo, P., Tari, G., and Vendeville, B., 1998, Protracted salt deformation history of the Lower Congo and Kwanza Basins of offshore Angola (abs.). American Association of Petroleum Geologists Annual Meeting Abstracts CD, unpaginated.Google Scholar
Spencer, J. A., and Sharpe, C. L., 1993, Deep-seated salt sheet in eastern Calcasieu Parish, Louisiana: Gulf Coast Association of Geological Societies Transactions, 43, 363371.Google Scholar
Spiers, C. J., and Brzesowsky, R. H., 1993, Densification behaviour of wet granular salt: Theory versus experiment, in Kakihana, H., Hardy, H. R. Jr., Hoshi, T., and Toyokura, K., eds., Seventh symposium on salt, Vol. 1: Amsterdam, Elsevier Science, 8392.Google Scholar
Spiers, C. J., and Carter, N. L., 1998, Microphysics of rocksalt flow in nature, in Aubertin, M. and Hardy, H. R., eds., The mechanical behavior of salt: Proceedings of the 4th conference: Clausthal-Zellerfeld, Trans Tech Publications, Series on Rock and Soil Mechanics, 22, 115128.Google Scholar
Spiers, C. J., de Meer, S., Niemeijer, A., and Zhang, X., 2004, Kinetics of rock deformation by pressure solution and the role of thin aqueous films, in Nakashima, A., ed., Proceedings of the international symposium on physiochemistry of water and dynamics of materials and the Earth – Structures and behaviours of the thin film water: September 2003: Tokyo, Universal Academy Press, Frontier Science Series, 44.Google Scholar
Spiers, C. J., Schutjens, P. M. T. M., Brzesowsky, R. H., Peach, C. J., Liezenberg, J. L., and Zwart, H. J., 1990, Experimental determination of constitutive parameters governing creep of rocksalt by pressure solution, in Knipe, R. J. and Rutter, E. H., eds., Deformation mechanisms, rheology and tectonics: London, Geological Society, Special Publication 54, 215227, doi:10.1144/GSL.SP.1990.054.01.21.Google Scholar
Staupendahl, G., and Schmidt, M. W., 1984, Field investigation of the long-term deformational behavior of a 10,000 m3 cavity at the Asse salt mine, in Hardy, H. R. Jr., and Langer, M., eds., The mechanical behavior of salt: Proceedings of the First Conference: Clausthal-Zellerfeld, Trans Tech Publications, 511525.Google Scholar
Ștefănescu, M., Dicea, O., and Tari, G., 2000, Influence of extension and compression on salt diapirism in its type area, East Carpathians Bend area, Romania, in Vendeville, B. C., Mart, Y., and Vigneresse, J.-L., eds., Salt, shale and igneous diapirs in and around Europe: London, Geological Society, Special Publication 174, 131147.Google Scholar
Steffens, G. S., Biegert, E. K., Sumner, H. S., and Bird, D., 2003, Quantitative bathymetric analyses of selected deepwater siliciclastic margins: Receiving basin configurations for deepwater fan systems: Marine and Petroleum Geology, 20, 547561, doi:10.1016/j.marpetgeo.2003.03.007.Google Scholar
Stewart, R. D., and Unterberger, R. R., 1976, Seeing through rock salt with radar: Geophysics, 41, 123132.Google Scholar
Stewart, S. A., 2006, Implications of passive salt diapir kinematics for reservoir segmentation by radial and concentric faults: Marine and Petroleum Geology, 23, 843853, doi:10.1016/j.marpetgeo.2006.04.001.Google Scholar
Stewart, S. A., 2007, Salt tectonics in the North Sea basin: a structural style template for seismic interpreters, in Ries, A. C., Butler, R. W. H., and Graham, R. H., eds., Deformation of the continental crust: The legacy of Mike Coward: London, Geological Society, Special Publication 272, 361396.Google Scholar
Stewart, S. A., and Argent, J. D., 2000, Relationship between polarity of extensional fault arrays and presence of detachments: Journal of Structural Geology, 22, 693711.Google Scholar
Stewart, S. A., and Clark, A. J., 1999, Impact of salt on the structure of the central North Sea hydrocarbon fairways, in Fleet, A. J. and Boldy, S. A. R., eds., Petroleum geology of northwest Europe: Proceedings of the 5th conference: London, Geological Society, 179200.Google Scholar
Stewart, S. A., and Coward, M. P., 1995, Synthesis of salt tectonics in the southern North Sea, UK: Marine and Petroleum Geology, 12, 457475.Google Scholar
Stewart, S. A., and Coward, M. P., 1996, Genetic interpretation and mapping of salt structures: First Break, 14, 135141.Google Scholar
Stewart, S. A., Fraser, S. I., Cartwright, J. A., Clark, J. A., and Johnson, H. D., 1999, Controls on Upper Jurassic sediment distribution in the Durward–Dauntless area, UK blocks 21/1 l, 21/16, in Fleet, A. J. and Boldy, S. A. R., eds., Petroleum geology of northwest Europe: Proceedings of the 5th conference: London, Geological Society, 879896.Google Scholar
Stier, K., 1915, Strukturbild des Benther Salzgebirges: Jahresbericht des Niedersächsischen Geologischen Vereins, 8, 115.Google Scholar
Stille, H., 1910, Faltung des deutschen Bodens und des Salzgebirges: Kali Zeitschrift, 5, 17.Google Scholar
Stille, H., 1924, Grundfragen der vergleichenden Tektonik: Berlin, Borntraeger, 443 p.Google Scholar
Stille, H., 1925, The upthrust of the salt masses of Germany: American Association of Petroleum Geologists Bulletin, 9, 417441.Google Scholar
Storti, F., Soto Marin, R., Rossetti, F., and Casas Sainz, A. M., 2007, Evolution of experimental thrust wedges accreted from along-strike tapered, silicone-floored multilayers: Journal of the Geological Society, London, 164, 7385.Google Scholar
Strozyk, F., van Gent, H., Urai, J. L., and Kukla, P. A., 2012, 3D seismic study of complex intra-salt deformation: An example from the Upper Permian Zechstein 3 stringer, western Dutch offshore, in Alsop, G. I., Archer, S. G., Hartley, A. J., Grant, N. T., and Hodhkinson, R., eds., Salt tectonics, sediments and prospectivity : London, Geological Society, Special Publication 363, 489581.Google Scholar
Sumner, H. S., Robison, B. A., Dirks, W. K., and Holiday, J. C., 1990, Morphology and evolution of salt/mini-basin systems: Lower shelf and upper slope, central offshore Louisiana (abs.): Geological Society of America, Abstracts with Programs, A48.Google Scholar
Suppe, J., 1985, Principles of structural geology: Englewood Cliffs, NJ, Prentice-Hall, 537 p.Google Scholar
Suppe, J., Chou, G. T., and Hook, S. C., 1992, Rates of folding and faulting determined from growth strata, in McClay, K. R., ed., Thrust tectonics: London, Chapman & Hall, 105121.Google Scholar
Suter, J. R., and Berryhill, H. L. Jr., 1985, Late Quaternary shelf-margin deltas, northwest Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 69, 7791.Google Scholar
Szatmari, P., Tibana, P., de Araújo Simões, I., Senna de Carvalho, R., Cezar Leite, D., 2008, Atlas petrográfico dos evaporitos, in Mohriak, W., Szatmari, P., and Couto Anjos, S. M., eds., Sal geologia e tectônica: Exemplos nas bacias Brasileiras: São Paulo, Beca Edições-Petrobras, 4363.Google Scholar
Talbot, C. J., 1978, Halokinesis and thermal convection: Nature, 273, 739741.Google Scholar
Talbot, C. J., 1979, Fold trains in a glacier of salt in southern Iran: Journal of Structural Geology, 1, 518.Google Scholar
Talbot, C. J., 1981, Sliding and other deformation mechanisms in a glacier of salt in S. Iran, in McClay, K. R., and Price, N. J., eds., Thrust and nappe tectonics: London, Geological Society, Special Publication 9, 173183.Google Scholar
Talbot, C. J., 1992, Quo vadis tectonophysics? With a pinch of salt!: Journal of Geodynamics, 16, 120.Google Scholar
Talbot, C. J., 1993, Spreading of salt structures in the Gulf of Mexico: Tectonophysics, 228, 151166.Google Scholar
Talbot, C. J., 1995, Molding of salt diapirs by stiff overburden, in Jackson, M. P. A., Roberts, D. G., and Snelson, S., eds., Salt tectonics: A global perspective: Tulsa, OK, American Association of Petroleum Geologists, Memoir 65, 6175.Google Scholar
Talbot, C. J., 1998, Extrusions of Hormuz salt in Iran, in Blundell, D. J., and Scott, A. C., eds., Lyell: The past is the key to the present: London, Geological Society, Special Publication 143, 315334.Google Scholar
Talbot, C. J., 2005, Discussion of “Evidence for Triassic salt domes in the Tunisian Atlas from gravity and geological data”: Tectonophysics, 406, 249254.Google Scholar
Talbot, C. J., 2008, Hydrothermal salt – but how much? Discussion: Marine and Petroleum Geology, 25, 191202.Google Scholar
Talbot, C. J., and Aftabi, P., 2004, Geology and models of salt extrusion at Qum Kuh, central Iran: Journal of the Geological Society, London, 161, 114.Google Scholar
Talbot, C. J., Aftabi, P., and Chemia, Z., 2009a, Potash in a salt mushroom at Hormoz Island, Hormoz Strait, Iran: Ore Geology Reviews, 35, 317332, doi:10.1016/j.oregeorev.2008.11.005.Google Scholar
Talbot, C. J., and Alavi, M., 1996, The past of a future syntaxis across the Zagros, in Alsop, G. I., Blundell, D. J., and Davison, I., eds., Salt tectonics: London, Geological Society, Special Publication 100, 89110.Google Scholar
Talbot, C. J., Farhadi, R., and Aftabi, P., 2009b, Potash in salt extruded at Sar Pohl diapir, southern Iran: Ore Geology Reviews, 35, 352366, doi:10.1016/j.oregeorev.2008.11.002.Google Scholar
Talbot, C. J., and Jackson, M. P. A., 1987a, Internal kinematics of salt diapirs: American Association of Petroleum Geologists Bulletin, 71, 10681093.Google Scholar
Talbot, C. J., and Jackson, M. P. A., 1987b, Salt tectonics: Scientific American, 255, 7079.Google Scholar
Talbot, C. J., and Jarvis, R. J., 1984, Age, budget and dynamics of an active salt extrusion in Iran: Journal of Structural Geology, 6, 521533.Google Scholar
Talbot, C. J., Medvedev, S., Alavi, M., Shahrivar, H., and Heidari, E., 2000, Salt extrusion rates at Kuh-e-Jahani, Iran: June 1994 to November 1997, in Vendeville, B. C., Mart, Y., and Vigneresse, J. L., eds., Salt, shale and igneous diapirs in and around Europe: London, Geological Society, Special Publication 174, 93110.Google Scholar
Talbot, C. J., and Pohjola, V., 2009, Subaerial salt extrusions in Iran as analogues of ice sheets, streams and glaciers: Earth-Science Reviews, 97, 167195.Google Scholar
Talbot, C. J., and Rogers, E. A., 1980, Seasonal movements in a salt glacier in Iran: Science, 208, 395397.Google Scholar
Talbot, C. J., Rönnlund, P., Schmeling, H., Koyi, H., and Jackson, M. P. A., 1991, Diapiric spoke patterns: Tectonophysics, 188, 187201.Google Scholar
Talebian, M., and Jackson, J., 2004, A reappraisal of earthquake focal mechanisms and active shortening in the Zagros mountains of Iran: Geophysical Journal International, 156, 506526, doi:10.1111/j.1365–246X.2004.02092.x.Google Scholar
Tankard, A. J., and Balkwill, H. R., eds., 1989, Extensional tectonics and stratigraphy of the North Atlantic margins: Tulsa, OK, American Association of Petroleum Geologists, Memoir 46, 641 p.Google Scholar
Tari, G., Coterill, K., Molnar, J., Valasek, D., Walters, G., and Alvarez, Y., 2004, Salt tectonics and sedimentation in the offshore Majunga basin, Madagascar: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 614–639.Google Scholar
Tari, G., and Jabour, H., 2013, Salt tectonics along the Atlantic margin of Morocco, in Mohriak, W. U., Danforth, A., Post, P. J., Brown, D. E., Tari, G. C., Nemčok, M., and Sinha, S. T., eds., Conjugate divergent margins: London, Geological Society, Special Publication 369, 337353, doi:10.1144/SP369.23.Google Scholar
Tchalenko, J. S., 1970, Similarities between shear zones of different magnitudes: Geological Society of America Bulletin, 81, 16251640.Google Scholar
Teichmüller, R., 1948, Das Oberflachenbild des Salzdomes von Segeberg in Holstein: Zeitschrift der Deutschen Geologischen Gesellschaft, 98, 729.Google Scholar
Ter Heege, J. H., De Bresser, J. H. P., and Spiers, C. J., 2005a, Dynamic recrystallization of wet synthetic polycrystalline halite: Dependence of grain size distribution on flow stress, temperature and strain: Tectonophysics, 396, 3557, doi:10.1016/j.tecto.2004.10.002.Google Scholar
Ter Heege, J. H., De Bresser, J. H. P., and Spiers, C. J., 2005b, Rheological behaviour of synthetic rocksalt: The interplay between water, dynamic recrystallization and deformation mechanisms: Journal of Structural Geology, 27, 948963, doi:10.1016/j.jsg.2005.04.008.Google Scholar
Tesoriero, A. J., and Knauth, L. P., 1988, The distribution of trace water around brine leaks in the Avery Island salt mine: Implications for the natural migration of water in salt: Nuclear and Chemical Waste Management, 8, 189197.Google Scholar
Tester, J., Holgate, H. R., Armellini, F. J., Webley, P. A., Killilea, W. R., Hong, G. T., and Berner, H. E., 1993, Supercritical water oxidation technology, in Emerging technologies in hazardous waste management: Washington, D.C., American Chemical Society, 3, 3576.Google Scholar
Thoms, R. L., and Gehle, R. M., 2000, A brief history of salt cavern use, in Geertman, R. M., ed., The 8th World Salt Symposium: Amsterdam, Elsevier, 2, 207214.Google Scholar
Thoms, R. L., Mogharrebi, M., and Gehle, R. M., 1982, Geomechanics of borehole closure in salt domes, in Gas Processors Association, 61st annual convention, Dallas, Texas, 228–230.Google Scholar
Torrey, P. D., and Fralich, C. E., 1926, An experimental study of the origin of salt domes: Journal of Geology, 34, 224234.Google 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, 13151333.Google Scholar
Trabant, P. K., and Presley, B. J., 1978, Orca Basin, anoxic depression on the continental slope, northwest Gulf of Mexico, in Bouma, A. H., Moore, G. T., and Coleman, J. M., eds., Framework, facies and oil-trapping characteristics of the upper continental margin: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 7, 303312.Google Scholar
Trudgill, B., 2011, Evolution of salt structures in the northern Paradox Basin: Controls on evaporite deposition, salt wall growth and supra-salt stratigraphic architecture: Basin Research, 23, 208238.Google Scholar
Trusheim, F., 1957, Über Halokinese und ihre Bedeutung für die strukturelle Entwicklung Norddeutschlands: Zeitschrift der Deutschen Geologischen Gesellschaft, 109, 111151.Google Scholar
Trusheim, F., 1960, Mechanism of salt migration in northern Germany: American Association of Petroleum Geologists Bulletin, 44, 15191540.Google Scholar
Tsvankin, I., 1997, Anisotropic parameters and P-wave velocity for orthorhombic media: Geophysics, 62, 12921309.Google Scholar
Tucker, R. M., 1981, Giant polygons in the Triassic salt of Cheshire, England: A thermal contraction model for their origin: Journal of Sedimentary Petrology, 51, 779786.Google Scholar
Turcotte, D. L., and Schubert, G., 2002, Geodynamics (2nd ed.): Cambridge, Cambridge University Press, 456 p.Google Scholar
Twiss, R. J., and Moores, E. M., 1992, Structural geology: New York, W. H. Freeman, 532 p.Google Scholar
Ulmishek, G. F., 2001, Petroleum geology and resources of the North Caspian basin, Kazakhstan and Russia: Washington, D.C., U.S. Geological Survey, Bulletin 2201-B, 25 p.Google Scholar
Unternehr, P., Péron-Pinvidic, G., Manatschal, G., and Sutra, E., 2010, Hyper-extended crust in the South Atlantic: In search of a model: Petroleum Geoscience, 16, 207215, doi:10.1144/1354-079309-904.Google Scholar
Urai, J. L., 1983, Water-assisted dynamic recrystallization and weakening in polycrystalline bischofite: Tectonophysics, 96, 125157.Google Scholar
Urai, J. L., 1985, Water-enhanced dynamic recrystallization and solution transfer in experimentally deformed carnallite: Tectonophysics, 120, 285317.Google Scholar
Urai, J. L., and Boland, J. N., 1985, Development of microstructures and the origin of hematite in naturally deformed carnallite: Neues Jahrbuch für Mineralogie-Abhandlungen, 2, 5872.Google Scholar
Urai, J. L., Means, W. D., and Lister, G. S., 1986a, Dynamic recrystallization of minerals, in Hobbs, B. E. and Heard, H. C., eds., Mineral and rock deformation: Laboratory studies; Paterson volume: Washington, D.C., American Geophysical Union, Geophysical Monograph 36, 161199.Google Scholar
Urai, J. L., Schléder, Z., Spiers, C. J., and Kukla, P. A., 2008, Flow and transport properties of salt rocks, in Littke, R., Bayer, U., Gajewski, D., and Nelskamp, S., eds., Dynamics of complex intracontinental basins: The central European basin system: Berlin, Springer, 277290.Google Scholar
Urai, J. L., and Spiers, C. J., 2007, The effect of grain boundary water on deformation mechanisms and rheology of rocksalt during long-term deformation, in Wallner, M., Lux, K., Minkley, W., and Hardy, H. Jr., eds., Proceedings of the 6th conference on the mechanical behavior of salt, Hannover, Germany, May 22–25, 2007: Basingstoke: Taylor and Francis, 19.Google Scholar
Urai, J. L., Spiers, C. J. Zwart, H. J., and Lister, G. S., 1986b, Weakening of rock salt by water during long-term creep: Nature, 324, 554557.Google Scholar
Usiglio, M. J., 1849, Etudes sur la composition de l’eau de la Méditerranée et sur l’exploitation des sels qu’elle contient: Annales de Chimie et de Physique, Série 3, 27, 172191.Google Scholar
Valle, P. J., Gjelberg, J. G., and Helland-Hansen, W., 2001, Tectonostratigraphic development in the eastern Lower Congo basin, offshore Angola, West Africa: Marine and Petroleum Geology, 18, 909927.Google Scholar
Van Berkel, J. T., 1986, A structural study of evaporite diapirs, folds and faults, Axel Heiberg Island, Canadian Arctic Islands: Amsterdam, University of Amsterdam, GUA Papers of Geology, Series 1, 26-1986, 149 p.Google Scholar
Van Beukel, J., Handschy, J., Ge, H., and Diegel, F., 2000, Salt dome geometries, Gulf of Mexico shelf, in American Association of Petroleum Geologists Hedberg Meeting, Integration of geologic models for understanding risk in the Gulf of Mexico, September 20–24, 1998, Galveston, Texas, 1–2.Google Scholar
Van de Fliert, J. R., 1953, Tectonique d’écoulement et Trias diapir au Chetthaabas, sud-ouest de la ville de Constantine, Algérie, in 19th International Geological Congress, Algeria, 3, 7192.Google Scholar
Van der Pluijm, B. A., and Marshak, S., 2004, Earth structure (2nd ed.): New York, W. W. Norton, 656 p.Google Scholar
Van Gent, H., Urai, J. L., and De Keijzer, M., 2011, The internal geometry of salt structures – A first look using 3D seismic data from the Zechstein of the Netherlands: Journal of Structural Geology, 33, 292311, doi:10.1016/j.jsg.2010.07.005.Google Scholar
Van Keken, P. E., Spiers, C. J., van den Berg, A. P., and Muyzert, E. J., 1993, The effective viscosity of rocksalt: implementation of steady-state creep laws in numerical models of salt diapirism: Tectonophysics, 225, 457476.Google Scholar
Velaj, T., Davison, I., Serjani, A., and Alsop, I., 1999, Thrust tectonics and the role of evaporites in the Ionian zone of the Albanides: American Association of Petroleum Geologists Bulletin, 83, 14081425.Google Scholar
Vendeville, B. C., 2002, A new interpretation of Trusheim’s classic model of salt-diapir growth: Gulf Coast Association of Geological Societies Transactions, 52, 943952.Google Scholar
Vendeville, B., and Cobbold, P. R., 1987, Glissements gravitaires synsédimentaires et failles normales listriques: modèles expérimentaux: Comptes Rendus de l’Académie des Sciences de Paris, Série II, 35, 13131320.Google Scholar
Vendeville, B., Ge, H., and Jackson, M. P. A., 1995, Scale models of salt tectonics during basement-involved extension: Petroleum Geoscience, 1, 179183.Google Scholar
Vendeville, B. C., and Jackson, M. P. A., 1991, Deposition, extension, and the shape of downbuilding salt diapirs (abs.): American Association of Petroleum Geologists Bulletin, 75, 687688.Google Scholar
Vendeville, B. C., and Jackson, M. P. A., 1992a, The rise of diapirs during thin-skinned extension: Marine and Petroleum Geology, 9, 331353.Google Scholar
Vendeville, B. C., and Jackson, M. P. A., 1992b. The fall of diapirs during thin-skinned extension: Marine and Petroleum Geology, 9, 354371.Google Scholar
Vendeville, B. C., and Nilsen, K. T., 1995, Episodic growth of salt diapirs driven by horizontal shortening, in Salt, sediment, and hydrocarbons: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 16th annual research conference program and extended abstracts, 285–295.Google Scholar
Vengosh, A., Chivas, A. R., Starinsky, A., Kolodny, Y., Zhang, B., and Zhang, P., 1995, Chemical and boron isotope compositions on nonmarine brines from the Qaidam basin, Qinhai, China: Chemical Geology, 120, 135154.Google Scholar
Vergés, J., Goodarzi, M. G. H., Emami, H., Karpuz, R., Efstathiou, J., and Gillespie, P., 2011, Multiple detachment folding in Pusht-e Kuh arc, Zagros: Role of mechanical stratigraphy, in McClay, K., Shaw, J., and Suppe, J., eds., Thrust-fault-related folding: Tulsa, OK, American Association of Petroleum Geologists, Memoir 94, 6994, doi:10.1306/13251333M942899.Google Scholar
Veritel, 2008, Advertisement in AAPG Explorer, American Association of Petroleum Geologists, 29, no. 2.Google Scholar
Vially, R., Letouzey, J., Bénard, F., Haddadi, N., Desforges, G., Askri, H., and Boudjema, A., 1994, Basin inversion along the north African Margin, the Saharan Atlas (Algeria), in Roure, F., ed., Peritethyan platforms: Paris, Technip, 79118.Google Scholar
Ville, L., 1859, Notice géologique sur les salines des Zahrez et les gîtes de sel gemme du Rang-el-Melah et d’Aïn Hadjera (Algérie): Annales des Mines, Paris, 15, 351410.Google Scholar
Visser, H. J. M., and Spiers, C. J., 1992, Compaction experiments on NaNO3 aggregates plus saturated solution: An analogue for partially molten systems (abs.): Eos, Transactions, American Geophysical Union, 73, 528.Google Scholar
Voigt, W. A., 1962, Über Randtröge vor Schollenrändern und ihre Bedeutung im Gebiet der mitteleuropäischen Senke und angrenzender Gebiete: Zeitschrift der Deutschen Geologischen Gesellschaft, 114, 378418.Google Scholar
Volozh, Y., Talbot, C., and Ismail-Zadeh, A., 2003, Salt structures and hydrocarbons in the Pricaspian basin: American Association of Petroleum Geologists Bulletin, 87, 313334.Google Scholar
Wade, A., 1931, Intrusive salt bodies in coastal Asir, south western Arabia: Institute of Petroleum Technologists Journal, 17, 321–330 and 357–361.Google Scholar
Wagner, B. H., III, 2010, An analysis of salt welding: Ph.D. dissertation, The University of Texas at Austin, 218 p.Google Scholar
Wagner, B. H., III, and Jackson, M. P. A., 2011, Viscous flow during salt welding: Tectonophysics, 510, 309326, doi:10.1016/j.tecto.2011.07.012.Google Scholar
Wakefield, L. L., Droste, H., Giles, M. R., and Janssen, R., 1993, Late Jurassic plays along the western margin of the Central Graben, in Parker, J. R., ed., Petroleum Geology of Northwest Europe: Proceedings of the 4th conference: London, Geological Society, 459468.Google Scholar
Wallner, M., 1983, Stability calculations concerning a room and pillar design in rock salt, in 5th Congress, International Society for Rock Mechanics, Melbourne, Australia, Section D, 9–15.Google Scholar
Waltham, D., 1997, Why does salt start to move?: Tectonophysics, 282, 117128.Google Scholar
Warren, J., 1986, Shallow water evaporitic environments and their source rock potential: Journal of Sedimentary Petrology, 56, 442454.Google Scholar
Warren, J. K., 1991, Sulfate dominated sea-marginal and platform evaporative settings, in Melvin, J. L., ed., Evaporites, petroleum, and mineral resources: Amsterdam, Elsevier, Developments in Sedimentology, 50, 477533.Google Scholar
Warren, J., 1999, Evaporites: Their evolution and economics: Oxford, Blackwell Science, 438 p.Google Scholar
Warren, J., 2000a, Evaporites, brines and base metals: Fluids, flow and ‘the evaporite that was’: Australian Journal of Earth Sciences, 44, 149183, doi:10.1080/08120099708728302.Google Scholar
Warren, J., 2000b, Dolomite: Occurrence, evolution and economically important associations: Earth-Science Reviews, 52, 181.Google Scholar
Warren, J., 2006, Evaporites: Sediments, resources, and hydrocarbons: Berlin, Springer, 1035 p.Google Scholar
Watkins, J. S., Ladd, J. W., Buffler, R. T., Schaub, F. J., Houston, M. H., and Worzel, J. L., 1978, Occurrence and evolution of salt in deep Gulf of Mexico, in Bouma, A. H., Moore, G. T., and Coleman, J. M., eds., Framework, facies, and oil-trapping characteristics of the upper continental margin: Tulsa, OK, American Association of Petroleum Geologists, Studies in Geology, 7, 4365.Google Scholar
Wawersik, W. R., and Zeuch, D. H., 1986, Modelling and mechanistic interpretation of creep of rock salt below 200 °C: Tectonophysics, 121, 125152.Google Scholar
Weijermars, R., 1988a, Neogene tectonics in the western Mediterranean may have caused the Messinian salinity crisis and an associated glacial event: Tectonophysics, 148, 211219.Google Scholar
Weijermars, R., 1988b, Convection experiments in high Prandtl number silicones; Part 1. Rheology, equipment, nomograms and dynamic scaling of stress- and temperature-dependent convection in a centrifuge: Tectonophysics, 154, 7196.Google Scholar
Weijermars, R., Hudec, M. R., Dooley, T. P., and Jackson, M. P. A., 2015, Downbuilding salt stocks and sheets quantified in 3-D analytical models: Journal of Geophysical Research: Solid Earth, 120, 46164644, doi:10.1002/2014JB011704.Google Scholar
Weijermars, R., Jackson, M. P. A., and Dooley, T. P., 2014, Quantifying drag on wellbore casings in moving salt sheets: Geophysical Journal International, 198, 965977, doi:10.1093/gji/ggu174.Google Scholar
Weijermars, R., Jackson, M. P. A., and van Harmelen, A., 2013, Closure of open wellbores in creeping salt sheets: Geophysical Journal International, 196, 279290, doi:10.1093/gji/ggt346.Google Scholar
Weijermars, R., Jackson, M. P. A., and Vendeville, B. C., 1993, Rheological and tectonic modeling of salt provinces: Tectonophysics, 217, 143174.Google Scholar
Weijermars, R., and Schmeling, H., 1986, Scaling of Newtonian and non-Newtonian fluid dynamics without inertia for quantitative modelling of rock flow due to gravity (including the concept of rheological similarity): Physics of the Earth and Planetary Interiors, 43, 316330.Google Scholar
Weinberger, R., Lyakhovsky, V., Baer, G., and Begin, Z. B., 2006, Mechanical modeling and InSAR measurements of Mount Sedom uplift, Dead Sea basin: Implications for effective viscosity of rock salt: Geochemistry Geophysics Geosystems, 7, Q05014, doi:10.1029/2005GC001185.Google Scholar
Weir, G. W., Kennedy, V. C., Puffett, W. P., and Dodson, C. L., 1961a, Preliminary geologic map of the Mount Peale 2 NW quadrangle, San Juan County, Utah: Washington, D.C., U.S. Geological Survey, Mineral Investigations Field Studies Map MF-152, scale 1:24,000, 1 sheet.Google Scholar
Wendlandt, E. A., and Knebel, G. M., 1929, Lower Claiborne of East Texas, with special reference to Mount Sylvan dome and salt movements: American Association of Petroleum Geologists Bulletin, 13, 13471375.Google Scholar
Wendlandt, E. A., Shelby, T. H. Jr., and Bell, J. S., 1946, Hawkins field, Wood County, Texas: American Association of Petroleum Geologists Bulletin, 30, 18301856.Google Scholar
Wenkert, D. D., 1979, The flow of salt glaciers: Geophysical Research Letters, 6, 523526.Google Scholar
Wiener, R. W., Mann, M. G., Angelich, M. T., and Molyneux, J. B., 2010, Mobile shale in the Niger Delta: Characteristics, structure, and evolution, in Wood, L., ed., Shale tectonics: Tulsa, OK, American Association of Petroleum Geologists, Memoir 93, 145161, doi:10.1306/13231313M933423.Google Scholar
Wilckens, O., 1912, Grundzüge der tektonischen Geologie: Jena, Fischer.Google Scholar
Williams-Stroud, S. C., and Paul, J., 1997, Initiation and growth of gypsum piercement structures in the Zechstein basin: Journal of Structural Geology, 19, 897907.Google Scholar
Willis, B., 1893, Mechanics of Appalachian structure: U.S. Geological Survey Annual Report 13, part 2, 217281.Google Scholar
Willson, S. M., and Fredrich, J. T., 2005, Geomechanics considerations for through- and near-salt well design: Society of Petroleum Engineers Annual Technical Conference and Exhibition, Dallas, Texas, October 9–12, SPE 95621, 17 p.Google Scholar
Wilson, T. P., and Long, D. T., 1993, Geochemistry and isotope chemistry of Ca–Na–Cl brines in Silurian strata, Michigan basin, USA: Applied Geochemistry, 8, 507524.Google Scholar
Winker, C. D., 1993, Pleistocene “lowstand” deposits of the Mississippi, Mobile, and Trinity rivers: Model for turbidite sedimentation in the Gulf of Mexico (abs.): American Association of Petroleum Geologists Annual Convention Program, 202.Google Scholar
Winker, C. D., and Booth, J. R., 2000, Sedimentary dynamics of the salt-dominated continental slope, Gulf of Mexico: Integration of observations from the seafloor, near-surface, and deep subsurface, in Deep-water reservoirs of the world: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 20th annual research conference, 1059–1086.Google Scholar
Withjack, M. O., and Callaway, S., 2000, Active normal faulting beneath a salt layer: An experimental study of deformation patterns in the cover sequence: American Association of Petroleum Geologists Bulletin, 84, 627651.Google Scholar
Withjack, M., Olson, J., and Peterson, E., 1990, Experimental models of extensional forced folds: American Association of Petroleum Geologists Bulletin, 74, 10381054.Google Scholar
Withjack, M. O., and Scheiner, C., 1982, Fault patterns associated with domes – An experimental and analytical study: American Association of Petroleum Geologists Bulletin, 66, 302316.Google Scholar
Withjack, M. O., and Schlische, R. W., 2006, Geometric and experimental models of extensional fault-bend folds, in Buiter, S. J. H. and Schreurs, G., eds., Analogue and numerical modelling of crustal-scale processes: London, Geological Society, Special Publications, 253, 285305.Google Scholar
Wood, L. J., 2010, Shale tectonics: A preface, in Wood, L., ed., Shale tectonics: Tulsa, OK, American Association of Petroleum Geologists, Memoir 93, 14, doi:10.1306/13231305M93730.Google Scholar
Woodcock, N., and Fischer, M., 1986, Strike–slip duplexes: Journal of Structural Geology, 8, 725735.Google Scholar
Woolnough, W. G., 1937, Sedimentation in barred basins, and source rocks of oil: American Association of Petroleum Geologists Bulletin, 21, 11011157.Google Scholar
Worrall, D. M., and Snelson, S., 1989, Evolution of the northern Gulf of Mexico, with emphasis on Cenozoic growth faulting and the role of salt, in Bally, A. W. and Palmer, A. R., eds., The geology of North America – An overview: Boulder, CO, Geological Society of America, 97138.Google Scholar
Wu, J. E., McClay, K., Whitehouse, P., and Dooley, T., 2009, 4D analogue modelling of transtensional pull-apart basins: Marine and Petroleum Geology, 26, 16081623, doi:10.1016/j.marpetgeo.2008.06.007.Google Scholar
Yaramanci, U., 1994, Relation of in situ resistivity to water content in salt rocks: Geophysical Prospecting, 41, 229239.Google Scholar
Yeilding, C. A., and Travis, C. J., 1997, Nature and significance of irregular geometries at the salt–sediment interface: examples from the deepwater Gulf of Mexico (abs.): American Association of Petroleum Geologists Annual Convention Official Program, A128.Google Scholar
Yin, H., and Groshong, R. H. Jr., 2007, A three-dimensional kinematic model for the deformation above an active diapir: American Association of Petroleum Geologists Bulletin, 91, 343363, doi:10.1306/10240606034.Google Scholar
Yovanovitch, B., 1922, La géologie du pétrole au Maroc: Bulletin de la Société Géologique de France, 22, 234245.Google Scholar
Zak, I., and Freund, R., 1980, Strain measurements in eastern marginal shear zone of Mount Sedom salt diapir, Israel: American Association of Petroleum Geologists Bulletin, 64, 568581.Google Scholar
Zalán, P. V., do Carmo G. Severino, M., Rigoti, C. A., Portugal Magnavita, L., Bach de Oliveira, J. A., and Roessler Vianna, A., 2011, An entirely new 3D-view of the crustal and mantle structure of a South Atlantic passive margin – Santos, Campos and Espírito Santo basins, Brazil: Tulsa, OK, American Association of Petroleum Geologists, Search and Discovery Article #30177.Google 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: London, Geological Society, 4559.Google Scholar
Zhang, Y., 2015, Promises and superiority of reverse time migration: Paris, CGGVeritas, www.cgg.com/technicalDocuments/cggv_0000008633.pdf, 71 p.Google Scholar
Zharkov, M. A., 1981, History of Paleozoic salt accumulation: Berlin, Springer, 308 p.Google Scholar
Zharkov, M. A., 1984, Paleozoic salt bearing formations of the world: Berlin, Springer, 427 p.Google Scholar
Ziegler, P. A., 1975, Geologic evolution of North Sea and its tectonic framework: American Association of Petroleum Geologists Bulletin, 59, 10731097.Google Scholar
Ziegler, P. A., 1978, Northwestern Europe: tectonics and basin development: Geologie en Mijnbouw, 57, 589626.Google Scholar
Ziegler, P. A., 1989, Geodynamic model for Alpine intra-plate compressional deformation in western and central Europe, in Cooper, M. A. and Williams, G. D., eds., Inversion tectonics: London, Geological Society, Special Publication 44, 6385.Google Scholar
Ziegler, P. A., 1990, Geologic atlas of western and central Europe: The Hague, Shell International Petroleum, Maastschappij B.V., 239 p.Google Scholar
Ziegler, P. A., 2005, Europe: Permian to recent evolution, in Selley, R. C., Cocks, L. R. M., and Plimer, I. R., eds., Encyclopedia of geology: Amsterdam, Elsevier, 2, 102125.Google Scholar
Ziegler, P. A., Cloetingh, S., and Van Wees, J.-D., 1995, Dynamics of intra-plate compressional deformation: The Alpine foreland and other examples: Tectonophysics, 252, 759.Google Scholar
Ziegler, P. A., and van Hoorn, B., 1989, Evolution of North Sea rift system, in Tankard, A. J. and Balkwill, H. R., eds., Extensional tectonics and stratigraphy of the North Atlantic margins: Tulsa, OK, American Association of Petroleum Geologists, Memoir 46, 471500.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

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
×