Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-767nl Total loading time: 0 Render date: 2024-07-11T07:37:59.756Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  13 September 2019

John W. Snedden
Affiliation:
University of Texas, Austin
William E. Galloway
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
The Gulf of Mexico Sedimentary Basin
Depositional Evolution and Petroleum Applications
, pp. 297 - 323
Publisher: Cambridge University Press
Print publication year: 2019

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

Acevedo, J. S., 1980, Giant fields of the southern zone, Mexico, in Halbouty, M. T., ed., Giant Oil and Gas Fields of the Decade 1968–1978: American Association of Petroleum Geologists, Memoir 30, 339385.Google Scholar
Acevedo, J. S., and Dautt, O. M., 1980, Giant fields in the southeast of Mexico: Gulf Coast Association of Geological Societies Transactions, 30, 133.Google Scholar
Achauer, C. A., 1977, Contrasts in cementation, dissolution, and porosity development between two Lower Cretaceous reefs of Texas, in Bebout, D. G. and Loucks, R. G., eds., Cretaceous Carbonates of Texas and Mexico: Applications to Subsurface Exploration: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 89, 127137.Google Scholar
Adams, R. L., and Carr, J. P., 2010, Regional depositional systems of the Woodbine, Eagle Ford, and Tuscaloosa of the U.S. Gulf Coast: Gulf Coast Association of Geological Societies Transactions, 60, 327.Google Scholar
Adams, S., 1985, Lithofacies of the middle Glen Rose reef buildup, Lower Cretaceous shelf margin, east Texas and Louisiana, in Bebout, D. and Ratcliff, D., eds., Lower Cretaceous Depositional Environments from Shoreline to Slope: Society of Economic Paleontologists and Mineralogists, Lower Cretaceous Core Workshop, 1322.Google Scholar
Ahr, W. M., 1973, The carbonate ramp: an alternative to the shelf model: Gulf Coast Association of Geological Societies Transactions, 23, 221225.Google Scholar
Ahr, W. M., 1981, The Gilmer Limestone: oolite tidal bars on the Sabine uplift: Gulf Coast Association of Geological Societies Transactions, 31, 16.Google Scholar
Ajdukiewicz, J. M., Nicholson, P. H., and Esch, W. L., 2010, Prediction of deep reservoir quality from early diagenetic process models in the Jurassic eolian Norphlet formation, Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 94, 11891227.Google Scholar
Alcocer, J. A. E., 2012, Avances y Resultados de la Exploracion en la Porcion Mexicana del Golfo de Mexico Profundo: Pemex Subdirection of Exploration, 30 p.Google Scholar
Alegret, L., Arnillas, I., Arz, J., et al., 2005, Cretaceous–Paleogene boundary deposits at Loma Capiro, central Cuba: evidence for the Chicxulub impact: Geology, 33, 721724, doi:10.1130/G21573.1.Google Scholar
Alexander, C. I., 1951, History of discovery and development of Woodbine oil fields in East Texas, in The Woodbine and Adjacent Strata: History of Discovery and Development of Woodbine Oil Fields in East Texas: Dallas Petroleum Geologists, Fondren Science Series, 1220.Google Scholar
Alnahwi, A., Loucks, R. G., Ruppel, S. C., Scott, R. W., and Tribovillard, N., 2018, Dip-related changes in stratigraphic architecture and associated sedimentological and geochemical variability in the Upper Cretaceous Eagle Ford Group in south Texas: American Association of Petroleum Geologists Bulletin, 102, 25372568.Google Scholar
Alvarez, L. W., Alvarez, W., Asaro, F., and Michel, H. V., 1980, Extraterrestrial cause for the Cretaceous–Tertiary extinction, Science, 208, 10951108.Google Scholar
Alzaga-Ruiz, H., Michel, L., Roure, F., and Seranne, M., 2009a, Interactions between the Laramide Foreland and the passive margin of the Gulf of Mexico: tectonics and sedimentation in the Golden Lane area, Veracruz State, Mexico: Marine and Petroleum Geology 26, 951973.CrossRefGoogle Scholar
Alzaga-Ruiz, H., Granjeon, D., Lopez, M., Seranne, M., and Roure, F., 2009b, Gravitational collapse and Neogene sediment transfer across the western margin of the Gulf of Mexico: insights from numerical models: Tectonophysics, 470, 2141.Google Scholar
Ambrose, W. A., Wawrzyniec, T. F., Fouad, K., et al., 2003, Geologic framework of Upper Miocene and Pliocene gas plays of the Macuspana Basin, southeastern Mexico: American Association of Petroleum Geologists Bulletin, 87, 14111435.Google Scholar
Ambrose, W. A., Jones, R. H., Fouad, K., et al., 2004, Sandstone Architecture of Upper Miocene and Pliocene Shoreface, Deltaic, and Valley-Fill Complexes, Macuspana Basin, Southeastern Mexico: Bureau of Economic Geology, Report of Investigations 270, 37 p.Google Scholar
Ambrose, W. A., Wawrzyniec, T. F., Fouad, K., et al., 2005, Neogene tectonic, stratigraphic, and play framework of the southern Laguna Madre-Tuxpan continental shelf, Gulf of Mexico: American Association of Petroleum Geologists Bulletin 89, 725751.Google Scholar
Ambrose, W. A., Hentz, T. F., Bonnaffee, F., et al., 2009, Sequence-stratigraphic controls on complex reservoir architecture of highstand fluvial-dominated deltaic and lowstand valley-fill deposits in the Upper Cretaceous (Cenomanian) Woodbine Group, East Texas field: regional and local perspectives: American Association of Petroleum Geologists Bulletin, 93, 231269, doi:10.1306/09180808053.Google Scholar
Ambrose, W. A., Loucks, R. G., and Dutton, S. P., 2013, Depositional Systems and Controls on Reservoir Quality (Determined from Core Data) in Deeply Buried Tertiary Strata in the Texas–Louisiana Gulf of Mexico: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 278, 80 p.Google Scholar
Ambrose, W. A., Loucks, R. G., and Dutton, S. P., 2015, Sequence-stratigraphic and depositional controls on reservoir quality in lowstand incised-valley fill and highstand shallow-marine systems in the Upper Cretaceous (Cenomanian) Tuscaloosa Formation, Louisiana, U.S.A.: Gulf Coast Association of Geological Societies Journal, 4, 4366.Google Scholar
Amundsen, L., and Landro, M., 2008, Seismic imaging technology, Part II: lessons from wide azimuth subsalt imaging in deepwater Gulf of Mexico: GEO ExPro, May, 60–62.Google Scholar
Anderson, E. G., 1979, Basic Mesozoic Study in Louisiana: the Northern Coastal Region and the Gulf Basin Province: Louisiana Geological Survey, Folio Series 3, 158.Google Scholar
Anderson, T. H., and Schmidt, V. A., 1983, The evolution of Middle America and the Gulf of Mexico–Caribbean Sea region during Mesozoic time: Geological Society of America Bulletin, 94, 941966.Google Scholar
Angeles-Aquino, F. J., and Cantú-Chapa, A., 2001, Subsurface Upper Jurassic stratigraphy in the Campeche shelf, Gulf of Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 343352.Google Scholar
Antunano, S. E., 2009, The Yegua Formation: gas play in the Burgos Basin, Mexico, in Bartolini, C. and Roman Ramos, J.R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 4978.Google Scholar
Applegate, A. V., 1984, The Brown Dolomite zone of the Lehigh Acres Formation (Aptian) in the South Florida Basin: a potentially prolific producing horizon offshore: Gulf Coast Association of Geological Societies Transactions, 34, 15.Google Scholar
Applegate, A. V., 1987, Part II: The Brow Dolomite Zone of the Lehigh Acres Formation (Aptian) in the South Florida Basin – a potentially prolific producing horizon offshore: State of Florida Department of Natural Resources Information Circular, 104, 4372.Google Scholar
Aquino Lopez, J. A., and Gonzalez, G. O., 2001, The Sihil Reservoir: the fifth block of Cantarell Field: Offshore Technology Conference, 1–15.Google Scholar
Arbouille, D., Andrus, V., Piperi, T., and Xu, T., 2013, Sub-salt and pre-salt plays: how much are left to be discovered?: American Association of Petroleum Geologists, Search and Discovery 10545, 17.Google Scholar
Arce, L. E. P., 2017, Neogene Current-Modified Submarine Fans and Associated Bed Forms in Mexican Deep-water Areas: MS thesis, The University of Texas at Austin, 105 p.Google Scholar
Armentrout, J. M. and Clement, J. F., 1990, Biostratigraphic calibration of depositional cycles: a case study in High Island–Galveston–East Breaks areas, offshore Texas: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, Stratigraphic Analysis Utilizing Advanced Geophysical, Wireline and Borehole Technology for Petroleum Exploration and Production, Seventeenth Annual Research Conference, Program and Abstracts, 21–51.Google Scholar
Armentrout, J. M., Malecek, S. J., Braithwaite, P., and Beeman, C. R., 1996, Intraslope basin reservoirs deposited by gravity-driven processes: south Ship Shoal and Ewing Bank areas, offshore Louisiana: Gulf Coast Association of Geological Societies Transactions, 46, 443448.Google Scholar
Arnold, G. A., Cavalero, S. R., Clifford, P. J., et al., 2010, Thunder Horse takes reservoir management to the next level: Offshore Technology Conference, Paper 20396, 10 p.Google Scholar
Arreguín-Lopez, M. A., and Weimer, P., 2004, Regional sequence stratigraphic setting on Miocene–Pliocene sediments, Veracruz Basin, Mexico: Gulf Coast Association of Geological Societies Transactions, 54, 2540.Google Scholar
Arreguín-Lopez, M. A., Reyna-Martínez, G., Sánchez-Hernández, H., Escamilla-Herrera, A., and Gutiérrez-Araiza, A., 2011, Tertiary turbidite systems in the southwestern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 61, 4553.Google Scholar
Artemieva, N., and Morgan, J., 2009, Modeling the formation of the K–Pg boundary layer, Icarus, 201(2), 768780, doi:10.1016/j.icarus.2009.01.021.CrossRefGoogle Scholar
Artemieva, N., Morgan, J., and Expedition 364 Science Party, 2017, Quantifying the release of climate-active gases by large meteorite impacts with a case study of Chicxulub: Geophysical Research Letters, 44, 180188, doi:10.1002/2017GL074879.CrossRefGoogle Scholar
Arthur, M. A., Schlanger, S. T., and Jenkyns, H. C., 1987, The Cenomanian–Turonian Oceanic Anoxic Event, II: palaeoceanographic controls on organic-matter production and preservation, in Brooks, J. and Fleet, A. J., eds., Marine Petroleum Source Rocks, Geological Society, London, Special Publications 26, 401420.Google Scholar
Aubrey, J., 1984, Recent Jurassic discoveries in Southeastern Cass County, Texas: Gulf Coast Association of Geological Societies Transactions, 34, 445451.Google Scholar
Ayers, W. B. Jr., and Lewis, A. H., 1985, The Wilcox Group and Carrizo Sand (Paleogene) in East-Central Texas: Depositional Systems and Deep-Basin Lignite: The University of Texas at Austin Bureau of Economic Geology, 19 p.Google Scholar
Badalini, G., Kneller, B., and Winker, C. D., 1999, Late Pleistocene Trinity-Brazos turbidite system: depositional processes and architectures in a ponded mini-basin system, Gulf of Mexico continental slope: American Association of Petroleum Geologists Bulletin, 83, 12961346.Google Scholar
Bailey, J., 1978, Black Lake Field Natchitoches Parish, Louisiana: A Review: Gulf Coast Association of Geological Societies Transactions, 28, 1124.Google Scholar
Bailey, J. W., 1983, Stratigraphy, environments of deposition, and petrography of the Cotton Valley Terryville Formation in Eastern Texas: unpublished MS thesis, The University of Texas at Austin, 229 p.Google Scholar
Barboza-Gudiño, J. R., Zavala-Monsiváis, A., Venegas-Rodríguez, G., and Barajas-Nigoche, L. D., 2010, Late Triassic stratigraphy and facies from northeastern Mexico: Tectonic setting and provenance: Geosphere, 6, 621640, doi: 10.1130/GES00545.1.Google Scholar
Baria, L. R., Heydari, E., and Winton, B. G., 2008, Shale layers in the Alabama Smackover Formation and their implications for sea-level change and regional correlation: Gulf Coast Association of Geological Societies Transactions, 58, 6775.Google Scholar
Barrell, K. A., 1997, Sequence stratigraphy and structural trap styles of the Tuscaloosa Trend: Gulf Coast Association of Geological Societies Transactions, 47, 2734.Google Scholar
Barrell, K. A., 2000, Conducting a field study with GIS: Port Hudson field, Tuscaloosa trend, East Baton Rouge Parish, Louisiana, in Coburn, T. C. and Yarus, J. M., eds., Geographic Information Systems in Petroleum Exploration and Development: American Association of Petroleum Geologists, Computer Applications in Geology 4, 187194.Google Scholar
Barros, J. A., 1987, Stratigraphy, Structure, and Paleogeography of the Jurassic–Cretaceous Passive Margin in the Western and Central Cuba: MS thesis, University of Miami.Google Scholar
Bay, A. R., 1985, Carbonate shoaling cycles in the Lower Glen Rose Formation (Lower Cretaceous), South Texas, in Bebout, D. G., ed., Annual Meeting of the Gulf Coast Association of Geological Societies and Society of Economic Paleontologists and Mineralogists Gulf Coast, Section Core Workshop 4: Lower Cretaceous Depositional Environments from Shoreline to Slope, 3746.Google Scholar
Bebout, D. G., 1977, Sligo and Hosston depositional patterns, subsurface of South Texas, in Bebout, D. G. and Loucks, R. G., eds., Cretaceous Carbonates of Texas and Mexico: Applications to Subsurface Exploration: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 89, 7996.Google Scholar
Bebout, D. G. and Gutiérrez, D. R., 1982, Regional Cross Sections, Louisiana Gulf Coast (Western Part): Louisiana Geological Survey, Folio Series 5.Google Scholar
Bebout, D. G. and Gutiérrez, D. R., 1983, Regional Cross Sections, Louisiana Gulf Coast (Eastern Part): Louisiana Geological Survey, Folio Series 6.Google Scholar
Bebout, D. G., Schatzinger, R. A., and Loucks, R. G., 1977, Porosity distribution in the Stuart City Trend, Lower Cretaceous, South Texas in Bebout, D. G. and Loucks, R. G., eds., Cretaceous Carbonates of Texas and Mexico: Applications to Subsurface Exploration: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 89, 234256.Google Scholar
Bebout, D. G., Loucks, R. G., and Gregory, A. R., 1978, Frio Sandstone Reservoirs in the Deep Subsurface along the Texas Gulf Coast: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 91, 92 p.Google Scholar
Bebout, D. G., Wise, B. R., Gregory, A. R., and Edwards, M. B., 1982, Wilcox Sandstone Reservoirs in the Deep Subsurface along the Texas Gulf Coast: Their Potential for Production of Geopressured Geothermal Energy: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 117, 125 p.Google Scholar
Bebout, D. G., White, W. A., Garrett, C. M., and Hentz, T. F., 1992, Atlas of Major Central and Eastern Gulf Coast Gas Reservoirs: Gas Research Institute, Chicago, 88 p.Google Scholar
Bentley, S. J., Blum, M. D., Maloney, J., Pond, L., and Paulsell, R., 2015, The Mississippi River source-to-sink system: perspectives on tectonic, climatic, and anthropogenic influences, Miocene to Anthropocene: Earth-Science Reviews, 153, 135.Google Scholar
Bhattacharya, J. P., Copeland, P., Lawton, T. F., and Holbrook, J., 2016, Estimation of source area, river paleo-discharge, paleoslope, and sediment budgets of linked deep-time depositional systems and implications for hydrocarbon potential: Earth-Science Reviews, 153, 77110, doi:10.1016/j.earscirev.2015.10.013.Google Scholar
Bitter, M. R., 1993, Sedimentation and provenance of Chicontepec sandstones with implications for uplift of the Sierra Madre Oriental and Teziutlan Massif, east-central Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 13th Annual Research Conference, 155–172.CrossRefGoogle Scholar
Blanc, P., and Connan, J., 1993, Crude oils in reservoirs: the factors influencing their composition, in Bordenave, M. L., ed., Applied Petroleum Geochemistry: Editions Technip, 149174.Google Scholar
Blanco, A., Maurrasse, F., Duque-Botero, F., and Delgado, A., 2011, Anoxic–dysoxic–oxic conditions in the Cenomanian Agua Nueva Formation (Upper Cretaceous) in central Mexico and their relation to Oceanic Anoxic Event 2 (OAE 2): Geologic Society of America Annual Meeting, Minneapolis, 172–174.Google Scholar
Blankenship, C. L., Knight, D. H., Kercho, D. A., et al., 2010, Will K: another step in the evolution of the U.S. Gulf of Mexico deep gas play: American Association of Petroleum Geologists, Search and Discovery 90104 (abs.)Google Scholar
Blum, M. D., and Hattier-Womack, J., 2009, Climate change, sea-level change, and fluvial sediment supply to deepwater depositional systems: a review, in Kneller, B., Martinsen, O.J., and McCaffrey, B., eds., External Controls on Deep-Water Depositional Systems: SEPM (Society for Sedimentary Geology), Special Publication, 92, 1539, doi:10.2110/sepmsp.092.015.Google Scholar
Blum, M., and Pecha, M., 2014, Mid-Cretaceous to Paleocene North American drainage reorganization from detrital zircons: Geology, 42(7), 607610, doi:10.1130/G35513.1.Google Scholar
Blum, M. D., Milliken, K. T., Pecha, M. A., et al., 2017, Detrital-zircon records of Cenomanian, Paleocene, and Oligocene Gulf of Mexico drainage integration and sediment routing: implications for scales of basin-floor fans: Geosphere, 13, 137, doi:10.1130/GES01410.1.Google Scholar
BOEM, 2016, Resource Evaluation Program: table of historical assessments, 1 p. www.boem.gov/national-assessment-history.Google Scholar
BOEM, 2017, Assessment of Technically and Economically Recoverable Hydrocarbon Resources of the Gulf of Mexico Outer Continental Shelf as of January 1, 2014: OCS, Report BOEM 2017-005, 50 p. www.boem.gov/BOEM-2017-005.Google Scholar
Boettcher, S. S., and Milliken, K. L., 1994, Mesozoic–Cenozoic unroofing of the southern Appalachian Basin: apatite fission track evidence from Middle Pennsylvanian sandstones: The Journal of Geology, 102, 655668.Google Scholar
Bohor, B. F., Foor, E. E., Modreski, P. J., and Triplehorn, D. M., 1984, Mineralogic evidence for an impact event at the Cretaceous–Tertiary boundary: Science, 224, 867869.Google Scholar
Bolli, H. M., Saunders, J. B., and Perch-Nielsen, K., 1989, Plankton Stratigraphy: Volume 1, Planktic Foraminifera, Calcareous Nannofossils and Calpionellids: CUP Archive, 607 p.Google Scholar
Booth, J. R., Duvernay, A. E. III, Pfeiffer, D. S., and Styzen, M. J., 2000, Sequence stratigraphic framework, depositional models, and stacking patterns of ponded and slope fan systems in the Auger Basin: central Gulf of Mexico slope, in Weimer, P., Slatt, R. M., Coleman, J., et al., eds., Deep-Water Reservoirs of the World: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 20th Annual Bob F. Perkins Research Conference, 82–102.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.Google Scholar
Bovay, A., 2015, New Models of Early Cretaceous Source-To-Sink Pathways in the Eastern Gulf of Mexico: MS thesis, The University of Texas at Austin, 105 p.Google Scholar
Bowman, S. A., 2012, Exploration targets of offshore Western Florida: Gulf Coast Association of Geological Societies Transactions, 62, 1326.Google Scholar
Boyd, D. R., and Dyer, B. F., 1964, Frio barrier bar system of South Texas: Gulf Coast Association of Geological Societies Transactions, 14, 309322.Google Scholar
Bralower, T. J., Arthur, M. A., Leckie, R. M., et al., 1994, Timing and paleoceanography of oceanic dysoxia/anoxia in the Late Barremian to Early Aptian (Early Cretaceous): Palaios, 9, 335369.Google Scholar
Bralower, T. J., Paul, C. K., and Leckie, R. M., 1998, The Cretaceous–Tertiary boundary cocktail: Chicxulub impact triggers margin collapse and extensive sediment gravity flows: Geology, 26, 331334, doi:10.1130/0091-7613(1998)026<0331:TCTBCC>2.3.CO;2.Google Scholar
Bralower, T. J., Cobabe, E., Clement, B., et al., 1999, The record of global change in mid-Cretaceous (Barremian–Albian) sections from the Sierra Madre, northeastern México: Journal of Foraminiferal Research, 29, 418437.Google Scholar
Bramlett, K. W., and Craig, P. A., 2002, Core characterization of slope-channel and channel–levee reservoirs in Ram Powell field, Gulf of Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, Foundation Deep-Water Core Workshop, Northern Gulf of Mexico, Houston, TX, 1–18.Google Scholar
Brand, J. H., 2016, Stratigraphy and mineralogy of the Oxfordian Lower Smackover Formation in the Eastern Gulf of Mexico, in Lowery, C. M., Snedden, J.W., and Blum, M., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: 35th Annual Gulf Coast Section SEPM Foundation Perkins Rosen Research Conference, 14–35.Google Scholar
Braunstein, J., Field, G. W., Garst, J., et al., 1949, Mesozoic cross-section from concordia parish, Louisiana, to Walton County, Florida: Mississippi Geological Society and Mesozoic Committee, 1.Google Scholar
Brown, A., 2011, Interpretation of Three-Dimensional Seismic Data: American Association of Petroleum Geologists, Memoir 42, 646 p, doi:10.1190/1.9781560802884.Google Scholar
Brown, L. B. Jr., Loucks, R. G., Treviño, R. H., and Hammes, U., 2004, Understanding growth-faulted, intraslope sub basins by applying sequence-stratigraphic principles: examples from the south Texas Oligocene Frio Formation: American Association of Petroleum Geologists Bulletin, 88, 15011522.Google Scholar
Brown, L.F., and Fisher, W.L., 1977, Seismic stratigraphic interpretation of depositional systems: examples from Brazilian rift and pull apart basins, in Payton, C., ed., Seismic Stratigraphy: Applications to Hydrocarbon Exploration: American Association of Petroleum Geologists, Memoir 26, 213248.Google Scholar
Bryant, W. R., Lugo, J., Cordova, C., and Salvador, A., 1991, Physiography and bathymetry, in Salvador, A., ed., The Gulf of Mexico Basin: The Geology of North America: Geological Society of America, 1330, doi:10.1130/DNAG-GNA-J.13.Google Scholar
Buffler, R. T., Shaub, F. J., Huerta, R., Ibrahim, A. B. K., and Watkins, J. S., 1981, A model for the early evolution of the Gulf of Mexico basin: Oceanologica Acta, 3, 129136.Google Scholar
Buffler, R. T., Schlager, W., and Pisciotto, K. A., 1984, Introduction and Explanatory Notes: Deep Sea Drilling Project, Volume LXXVII, Leg 77 Glomar Challenger: National Science Foundation, 740 p, doi:10.2973/dsdp.proc.77.101.1984.Google Scholar
Bull, S., Cartwright, J., and Huuse, M., 2009, A review of kinematic indicators from mass-transport complexes using 3D seismic data: Marine and Petroleum Geology, 26, 11321151.Google Scholar
Byerly, G. R., 1991, Igneous activity, in Salvador, A., ed., The Gulf of Mexico Basin: Geological Society of America, 91108.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, doi:10.1306/10120505003.Google Scholar
Cantú-Chapa, A., 1998, Las transgresiones jurásicas en México: Revista Mexicana de Ciencias Geológicas, 15, 2537.Google Scholar
Cantú-Chapa, A., 2009, Upper Jurassic stratigraphy (Oxfordian and Kimmeridgian) in petroleum wells of the Campeche shelf, in Bartolini, C. and Roman Ramos, J. R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum, Memoir 90, 7991, doi:10.1306/13191079M902804.Google Scholar
Cantú-Chapa, A., and Landeros-Flores, R., 2001, The Cretaceous–Paleocene boundary in the subsurface Campeche shelf, southern Gulf of Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologist, Memoir 75, 389395.Google Scholar
Cantú-Chapa, A., and Ortuño-Maldonado, E., 2003, The Tithonian (Upper Jurassic) Edzna Formation, an important hydrocarbon reservoir of the Campeche shelf, Gulf of Mexico, in Bartolini, C., Buffler, R. T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists, Memoir 79, 305311.Google Scholar
Cardona, S., Wood, L. J., Day-Stirrat, R. J., and Moscardelli, L., 2016, Fabric development and pore-throat reduction in a mass-transport deposit in the Jubilee gas field, eastern Gulf of Mexico: consequences for the sealing capacity of MTDs, in, Lamarche, G.Mountjoy, J., Bull, S., Hubble, T., Pecher, I., and Woelz, S., eds., Submarine Mass Movements and Their Consequences, Springer, 2737, doi:10.1007/978-3-319-20979-1.Google Scholar
Carreras, P. E., Johnson, S. G., Turner, S. E., 2006, Tahiti: Assessment of Uncertainty in a Deepwater Reservoir Using Design of Experiments: Society of Petroleum Engineers, SPE 102988, 16 p.Google Scholar
Castillon, M., and Larrios, J.P., 1963, Salt Deposits of the lsthmas of Tehuantepec: Northern Ohio Geological Society, 263280.Google Scholar
Cather, M. C., Chapin, C. E, and Kelley, S. A., 2012, Diachronous episodes of Cenozoic erosion in southwestern North America and their relationship to surface uplift, paleoclimate, paleodrainage, and paleoaltimetry: Geosphere 8, 11771206.Google Scholar
Cepeda, R., Weimer, P., and Dorn, G., 2010, 3D seismic stratigraphic interpretation of the Upper Miocene to Lower Pleistocene deepwater sediments of the Thunder Horse-Mensa area, Southern Mississippi Canyon, northern deep Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 60, 119132.Google Scholar
Chasteen, H. R., 1983, Re-evaluation of the lower Tuscaloosa and Dantzler formations (Mid-Cretaceous) with emphasis on depositional environments and time-stratigraphic relationships: Gulf Coast Association of Geological Societies Transactions, 33, 3140.Google Scholar
Chavez-Valois, V. M. C., Valdes, L. C., Placencia, J. I. J., et al., 2009, A new multidisciplinary focus in the study of tertiary plays in the Sureste Basin, Mexico, in Bartolini, C. and Roman Ramos, J. R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 155190.Google Scholar
Chen, J., Acosta Aduna, C., Sanchez Lu, F., Patino, J., and Olivella, M., 2001, Petrophysical characteristics, depositional systems, and model of geological evolution in the Golden Lane carbonate sequences: Gulf Coast Association of Geological Societies Transactions, 51, 3144.Google Scholar
Chernikoff, A., Hernandez, J. G., and Schatzinger, R., 2006, Mesozoic extensional tectonics: its impact on oil accumulations in Campeche Sound, Gulf of Mexico: The Leading Edge, 25(10), 12241234, doi:10.1190/1.2360609.Google Scholar
Christeson, G. L., Nakamura, Y., Buffler, R. T., Morgan, J., and, Warner, M., 2001, Deep crustal structure of the Chicxulub impact crater: Journal of Geophysical Research, 106, 2175121769.Google Scholar
Christeson, G. L., Van Avendonk, H. J. A., Norton, I. O., Snedden, J. W., and Eddy, D. R., 2014, Deep crustal structure in the eastern Gulf of Mexico: Journal of Geophysical Research: Solid Earth, 119(9), 67826801, doi:10.1002/2014JB011045.Google Scholar
Christeson, G. L., Gulick, S. P. S., Morgan, J. V., et al., 2018, Extraordinary rocks from the peak ring of the Chicxulub impact crater: P-wave velocity, density, and porosity measurements from IODP/ICDP Expedition 364, Earth and Planetary Science Letters, 495, 111.Google Scholar
Cicero, A. D., Steinhoff, I., McClain, T., Koepke, K. A., and Dezelle, J. D., 2010, Sequence stratigraphy of the Upper Jurassic mixed carbonate/siliciclastic Haynesville and Bossier shale depositional systems in east Texas and northern Louisiana: Gulf Coast Association of Geological Societies Transactions, 60, 133148.Google Scholar
Clara Valdés, M. D. L., Rodriguez, L. V., and Garcia, E. C., 2009, Geochemical integration and interpretation of source rocks, oils, and natural gases in southeastern Mexico, in Bartolini, C., and Roman Ramos, J. R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 337368, doi:10.1306/13191091M903337.Google Scholar
Clark, W. J., Iwere, F. O., Apaydin, O., et al., 2003, Integrated modeling of the Taratunich Field, Bay of Campeche, Southern Mexico: American Association of Petroleum Geologists, Search and Discovery 20014, 1–6.Google Scholar
Claypool, G. E., and Mancini, E. A., 1989, Geochemical relationships of petroleum in Mesozoic reservoirs to carbonate source rocks of Jurassic Smackover Formation, Southwestern Alabama: American Association of Petroleum Geologists Bulletin, 73, 904924.Google Scholar
Cleaves, A. W., 1977, Middle Glen Rose (Cretaceous) facies mosaic, Blanco and Hays Counties, Texas: American Association of Petroleum Geologists, Search and Discovery 90967, 1 p.Google Scholar
CNH, 2014, Southeast offshore basins shallow water area: petroleum geological synthesis: CNH, 1–64.Google Scholar
CNH, 2015a, Saline Basin: petroleum geological synthesis: CNH, 1–47. Available at http://portal.cnih.cnh.gob.mx. June 2017.Google Scholar
CNH, 2015b, Perdido Fold belt, subsalt belt, Mexican Ridges: petroleum geological synthesis, CNH, 1–53. Available at http://portal.cnih.cnh.gob.mx., June 2017.Google Scholar
CNH, 2017a, Pemex to explore Mexico’s pre-salt, CNH. Available at www.upstreamonline.com/hardcopy/1367799/pemex-preparing-pre-salt-test, October 2017.Google Scholar
CNH, 2017b, Geological Atlas Sabina-Burgos Basin, CNH, 1-53. Available at https://portal.cnih.cnh.gob.mx/downloads/en_US/info/Geological_Atlas_Sabinas-Burgos_Basins_V3.pdf, February, 2018.Google Scholar
CNH, 2018, Prospective resources, CNH, 1–2. Available at https://portal.cnih.cnh.gob.mx/downloads/en_US/estadisticas/Prospective%20Resources.pdf, September 2017.Google Scholar
Cobiella-Reguera, J. L., Cruz-Gámez, E. M., Blanco-Bustamante, S., et al., 2015, Cretaceous–Paleogene boundary deposits and paleogeography in western and central Cuba: Revista mexicana de ciencias geológicas, 32(1), 156176.Google Scholar
Coccioni, R., Luciani, V., Marsili, A., 2006, Cretaceous oceanic anoxic events and radially elongated chambered planktonic foraminifera: paleoecological and paleoceanographic implications: Palaeogeography, Palaeoclimatology, Palaeoecology, 235, 6692.Google Scholar
Cole, G. A., Requejo, A. G., Yu, A., et al., 1999, The geochemical and basin modeling aspects of the Jurassic to Lower Cretaceous sourced petroleum system, deepwater to ultra-deepwater Gulf of Mexico, offshore Louisiana: 3rd AMGP/American Association of Petroleum Geologists International Conference Proceedings, 1–8.Google Scholar
Cole, G. A., Yu, A., Peel, F., et al., 2001, Constraining source and charge risk in deepwater areas: World Oil, 222, 6977.Google Scholar
Coleman, J. L. Jr., and Coleman, C. J., 1981, Stratigraphic, sedimentologic and diagenetic framework for the Jurassic Cotton Valley Terryville massive sandstone complex, northern Louisiana: Gulf Coast Association of Geological Societies Transactions, 31, 7180.Google Scholar
Coleman, J. M., Prior, D. B., and Lindsay, J. F., 1983, Deltaic influences on shelfedge instability processes, in Stanley, D. J. and Moore, G. T., eds., The Shelfbreak: Critical Interface on Continental Margins: Society of Economic Paleontologists and Mineralogists, Special Publication 33, 121137.Google Scholar
Colmenares, M., and Hustedt, J., 2015, Overview of the deepwater geology of the Mexican Gulf of Mexico: round one of bidding in the energy reform: American Association of Petroleum Geologists, Search and Discovery 10738, 122.Google Scholar
Combellas-Bigott, R. I., and Galloway, W. E., 2002a, Depositional history and genetic sequence stratigraphic framework of the middle Miocene depositional episode, south Louisiana: Gulf Coast Association of Geological Societies Transactions, 52, 139150.Google Scholar
Combellas-Bigott, R. I., and Galloway, W. E., 2002b, Origin and evolution of the middle Miocene submarine-fan system, east-central Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 52, 151163.Google Scholar
Combellas-Bigott, R. I., and Galloway, W. E., 2006, Depositional and structural evolution of the middle Miocene depositional episode, east-central Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 90, 335362, doi:10.1306/10040504132.Google Scholar
Combes, J. M., 1993, The Vicksburg formation of Texas: depositional systems distribution, sequence stratigraphy, and petroleum geology: American Association of Petroleum Geologists Bulletin, 77, 19421970, doi:10.1306/BDFF8F88-1718-11D7-8645000102C1865D.Google Scholar
Comet, P. A., 1992, Maturity mapping of northern Gulf of Mexico oils using biomarkers: Gulf Coast Association of Geological Societies Transactions, 42, 433448.Google Scholar
Comet, P. A., Rafalska, J. K., and Brooks, J. M., 1993, Sterane and triterpene patterns as diagnostic tools in the mapping of oils, condensates, and source rocks of the Gulf of Mexico region: Organic Geochemistry, 20, 12651296.Google Scholar
Conrad, C. P., 2013, The solid Earth’s influence on sea level: Geological Society of America Bulletin, 125, 10271052, doi:10.1130/B30764.1Google Scholar
Coogan, A. H., Bebout, D. G., and Maggio, C., 1972, Depositional environments and geologic history of Golden Lane and Poza Rica trend, Mexico: an alternative view: American Association of Petroleum Geologists Bulletin, 56, 14191447.Google Scholar
Corbett, K., Friedman, M., and Spang, J., 1987, Fracture development and mechanical stratigraphy of Austin Chalk, Texas: American Association of Petroleum Geologists Bulletin, 71, 1728.Google Scholar
Cordona, S., Wood, L., and Moscardelli, L., 2015, Sealing capacity of mass transport deposits: depositional model for a deepwater reservoir in Jubilee gas field, eastern Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 90216.Google Scholar
Cornelius, S. and Castagna, J. P., 2018, Variation in salt-body interval velocities in the deepwater Gulf of Mexico: Keathley Canyon and Walker Ridge areas: Interpretation, 6, T15T17, doi:10.1190/INT-2017-0069.1.Google Scholar
Cornish, F., 2013, Do Upper Wilcox canyons support Paleogene isolation of the Gulf of Mexico?: Gulf Coast Association of Geological Societies Transactions, 63, 183204.Google Scholar
Cossey, S. P. J., and Bitter, M., 2016, The KPg Impact Deposits in the Tampico–Misantla Basin, Eastern Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Research Conference, 479–486.Google Scholar
Cossey, S., and Jacobs, R., 1992, Oligocene Hackberry Formation of southwest Louisiana: sequence stratigraphy, sedimentology, and hydrocarbon potential: American Association of Petroleum Geologists Bulletin 76, 589606.Google Scholar
Cossey, S. P. J., Pindell, J., and Rosenfeld, J., 2007, Recent geological understanding of the Chicontepec Erosional “paleocanyon,” Tampico–Misantla Basin, Mexico: The Paleogene of the Gulf of Mexico and Caribbean Basins: Processes, Events, and Petroleum Systems, Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 27th Annual Research Conference Program and Extended Abstracts, 273–283, doi:10.5724/gcs.07.27.0273.Google Scholar
Covault, J. A., and Graham, S. A., 2010, Submarine fans at all sea-level stands: tectono-morphologic and climatic controls on terrigenous sediment delivery to the deep sea: Geology, 38, 939942, doi:10.1130/G31081.1.Google Scholar
Cox, R. T., and Van Arsdale, R. B., 2002, The Mississippi Embayment, North America: a first order continental structure generated by the Cretaceous superplume mantle event: Journal of Geodynamics, 34(2), 163176, doi:10.1016/S0264-3707(02)00019-4.Google Scholar
Craddock, W. H., and Kylander-Clark, A. R. C., 2013, U–Pb ages of detrital zircons from the Tertiary Mississippi River Delta in central Louisiana: insights into sediment provenance: Geosphere, 9, 120.Google Scholar
Crane, M. J., 1965, Upper Cretaceous ostracodes of the Gulf Coast area: Micropaleontology, 11, 191254, doi:10.2307/1484517.Google Scholar
Cregg, A. K., and Ahr, W. M., 1983, Depositional framework and reservoir potential of an upper Cotton Valley (Knowles Limestone) patch reef, Milam County, Texas: Gulf Coast Association of Geological Societies Transactions, 33, 5568.Google Scholar
Crevello, P.D., and Harris, P.M., 1984, Depositional models for jurassic reefal buildups, in Ventress, P., ed., The Jurassic of the Gulf Rim: Society of Economic Paleontologists and Mineralogists Gulf Coast Section Foundation 3rd Annual Research Conference, 57–102.Google Scholar
Cruz, M., and Aguilera, R., 2018, Eagle Ford and Pimienta shales in Mexico: a case study: Society of Petroleum Engineers: SPE Canada Unconventional Resources Convention, 1–21, doi:0.2118/189797-MS.Google Scholar
Cruz Luque, M. M., Urban-Rascon, E., Aguilera, R. F., and Aguilera, R., 2018, Mexican unconventional plays: geoscience, endowment, and economic considerations: Society of Petroleum Engineers, 21, 533549, doi:10.2118/189438-PA.Google Scholar
Cruz-Mercado, M. A., Flores-Zamora, J. C., León-Ramírez, R., et al., 2011, Salt provinces in the Mexican portion of the Gulf of Mexico: structural characterization and evolutionary model: Gulf Coast Association of Geological Societies Transactions 61, 93103.Google Scholar
Cumming, E. W., 2002, Core, log, and seismic characteristics of a high-performance turbidite reservoir in a salt-withdrawal minibasin: the Upper Miocene yellow sand, Mars Field, Gulf of Mexico: Deep-Water Core Workshop, Northern Gulf of Mexico Gulf Coast Section, Society of Economic Palentologists and Mineralogists, 19–30.Google Scholar
Cunningham, R. C., Snedden, J. W., Norton, I. O., et al., 2016, Upper Jurassic Tithonian-centered source mapping in the deepwater northern Gulf of Mexico: Interpretation, 4(1), SC97SC123, doi:10.1190/INT-2015-0093.1.Google Scholar
Cunningham, R. C., Phillips, M. P., Snedden, J. W., et al., in prep., The Paleocene–Eocene Thermal Maximum (PETM) in Deepwater Gulf of Mexico: a new Paleogene source rock and basin-scale paleoceanographic model: Paleoceanography.Google Scholar
Curry, M. A. E., Peel, F. J., Hudec, M. R., and Norton, I. O., 2018, Extensional models for the development of passive-margin salt basins, with application to the Gulf of Mexico: Basin Research, 30, 120, doi:10.1111/bre.12299.Google Scholar
Damuth, J., and Olson, H. C., 2015, Latest Quaternary sedimentation in the northern Gulf of Mexico Intraslope Basin Province: I. Sediment facies and depositional processes: Geosphere, 11, 16891718. doi:10.1130/GES01090.1.Google Scholar
Davison, I., and Cunha, T. A., 2017, Allochthonous salt sheet growth: thermal implications for source rock maturation in the deepwater Burgos basin and Perdido fold belt, Mexico: Interpretation, 5, T11T21, doi:10.1190/INT-2016-0035.1.Google Scholar
Davison, I., O’Bierne, E., Faull, T., and Steel, I., 2015, Vast potential, Mexico’s Round Uno: GeoExpro, 12, 6871.Google Scholar
Day, S., and Maslin, M. (2005), Linking large impacts, gas hydrates, and carbon isotope excursions through widespread sediment liquefaction and continental slope failure: the example of the KT boundary event: Geological Society of America, Special Paper 384, 239–258.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: Gulf Coast Section SEPM, 22nd Annual Research Conference, 661–680.Google Scholar
De la Rocha Bascon, L., 2016, Southern Gulf of Mexico Wilcox Source-to-Sink: Investigating Siliciclastic Sedimentation in Mexico Deep-Water: MS thesis, The University of Texas at Austin, 88 p.Google Scholar
Denne, R. A., and Blanchard, R. H., 2013, Regional controls on the formation of the ancestral DeSoto Canyon by the Chicxulub Impact: Gulf Coast Association of Geological Societies Journal, 2, 1728.Google Scholar
Denne, R. A., Scott, E. D., Eickhoff, D. P., et al., 2013, Massive Cretaceous–Paleogene boundary deposit, deep-water Gulf of Mexico: new evidence for widespread Chicxulub-induced slope failure: Geology, 41, 983986, doi:10.1130/G34503.1.Google Scholar
Denne, R. A., Hinote, R. E., Breyer, J. A., et al., 2014, The Cenomanian–Turonian Eagle Ford Group of South Texas: insights on timing and micropaleontological analyses: Paleogeography, Palaeoclimatology, Paleoecology, 413, 228, doi:10.1016/j.palaeo.2014.05.029.Google Scholar
Desselles, R., Wilson, M., and Barminski, J., 2016a, Assessment of Undiscovered Oil and Gas Resources of the Nation’s Outer Continental Shelf: Bureau of Ocean Energy Management, National Assessment of Oil and Gas Resources, 18.Google Scholar
Diaz, J., Weimer, P., Bouroullec, R., and Dorn, G., 2011, 3-D seismic stratigraphic interpretation of Quaternary mass-transport deposits in the Mensa and Thunder Horse intraslope basins, Mississippi Canyon, northern deep Gulf of Mexico, U.S.A., in Shipp, C.R., Weimer, P., and Posamentier, H., eds., Mass Transport Deposits in Deep-Water Settings: Society for Sedimentary Geology, Special Publication, 96, 127149.Google Scholar
Dickey, R. L., and Yancey, T. E., 2010, Palynological age control of sediments bracketing the Paleocene–Eocene Boundary, Bastrop, Texas: Gulf Coast Association of Geological Societies Transactions 60, 717724.Google Scholar
Dickinson, W. R., and Gehrels, G. E., 2008, U–Pb ages of detrital zircons in relation to paleogeography: Triassic paleodrainage networks and sediment dispersal across southwest Laurentia: Journal of Sedimentary Research, 78, 745764, doi:10.2110/jsr.2008.088.CrossRefGoogle Scholar
Dickinson, W. R., and Gehrels, G. E., 2009, Use of U–Pb ages of detrital zircons to infer maximum depositional ages of strata: a test against a Colorado Plateau Mesozoic database: Earth and Planetary Science Letters, 288, 115125, doi:10.1016/j.epsl.2009.09.013.Google Scholar
Dickinson, W. R., Gehrels, G. E., and Stern, R. J., 2010, Late Triassic Texas uplift preceding Jurassic opening of the Gulf of Mexico: evidence from U–Pb ages of detrital zircons: Geosphere, 6, 641662, doi:10.1130/GES00532.1.Google Scholar
Dickson, W., and Schiefelbein, C., 2012, Girassol-Angola’s first deepwater pre-salt discovery?: American Association of Petroleum Geologists, Search and Discovery 20142, 14.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., Salt Tectonics: A Global Perspective: American Association of Petroleum Geologists, Memoir 65, 109151.Google Scholar
Dixon, J. F., Steel, R. J., and Olariu, C., 2012, Shelf-edge delta regime as a predictor of deep-water deposition: Journal of Sedimentary Research, 82, 681687, doi:10.2110/jsr.2012.59.Google Scholar
Dobson, L. M., 1990, Seismic Stratigraphy and Geologic History of Jurassic Rocks, Northeastern Gulf of Mexico, MA thesis; The University of Texas at Austin, 170 p.Google Scholar
Dodge, M. M., and Posey, J. S., 1981, Structural Cross Sections, Tertiary Formations, Texas Gulf Coast: The University of Texas at Austin, Bureau of Economic Geology, 42 p.Google Scholar
Dohmen, T. E., 2002, Age dating of expected MCSB seismic event suggests that it is the K/T Boundary: Gulf Coast Association of Geological Societies Transactions, 52, 177180.Google Scholar
Donovan, A. D., Scott Staerker, T., Pramudito, A., et al., 2012, The Eagle Ford outcrops of West Texas: a laboratory for understanding heterogeneities within unconventional mudstone reservoirs: Gulf Coast Association of Geological Societies Journal, 1, 162185.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
Doughty-Jones, G., Mayall, M., and Lonergan, L., 2017, Stratigraphy, facies, and evolution of deep-water lobe complexes within a salt controlled intraslope minibasin: American Association of Petroleum Geologists Bulletin, 101, 18791904.Google Scholar
Douglas, S. W., 2010, The Jurassic Norphlet Formation of the Deep-Water Eastern Gulf of Mexico: A Sedimentologic Investigation of Aeolian Facies, their Reservoir Characteristics, and their Depositional History: MS thesis, Baylor University, 68 pp.Google Scholar
Dravis, J., 1981, Depositional setting and porosity evolution of the Upper Cretaceous Austin Chalk Formation, South-Central Texas: South Texas Geological Society Bulletin, 24, 414.Google Scholar
Dubiel, R. F., Pitman, J. K., and Steinshouer, D., 2003, Seismic-sequence stratigraphy and petroleum system modeling of the downdip Tuscaloosa-Woodbine, LA and TX: Gulf Coast Association of Geological Societies Transactions, 52, 193203.Google Scholar
Dubiel, R. F., Pitman, J. K., Pearson, O. N., et al., 2012, Assessment of undiscovered oil and gas resources in conventional and continuous petroleum systems in the Upper Cretaceous Eagle Ford Group, U.S. Gulf Coast Region, 2011: US Geological Survey, Fact Sheet 2012-3003, 2 p.Google Scholar
Duckworth, C. J., Dowty, S. G., and Bebout, D. G., 1992, KS-8: Trinity Group Sandstone—Mississippi interior salt basin, in Bebout, D. G., White, W. A., Garrett, C. M. and Hentz, T. F., eds. Atlas of Major Central and Eastern Gulf Coast Gas Reservoirs: Gas Research Institute, Chicago, 88 p.Google Scholar
Durham, C. O. Jr., and Hall, S. B., 1991, The Austin Chalk: bed by bed through Central Texas: in Austin Chalk exploration symposium: geology geophysics and formation evaluation. San Antonio TX, South Texas Geological Society, 25–40.Google Scholar
Dutton, S., and Diggs, T. N., 1992, Evolution of porosity and permeability in the Lower Cretaceous Travis Peak Formation, East Texas: American Association of Petroleum Geologists Bulletin, 76, 252269.Google Scholar
Dutton, S. P., and Loucks, R. G., 2010, Diagenetic controls on evolution of porosity and permeability in lower Tertiary Wilcox sandstones from shallow to ultradeep (200–6700 m) burial, Gulf of Mexico Basin, U.S.A.: Marine and Petroleum Geology, 27, 6981.Google Scholar
Dutton, S. P., Jennette, D. C., Ambrose, W. A., and Martin, M., 2002, Petrography and reservoir quality of Tertiary deepwater sandstones in the Veracruz Basin, Mexico: Gulf Coast Association of Geological Societies Transactions, 52, 229240.Google Scholar
Dutton, S. P., Loucks, R. G., and Day-Stirrat, R. J., 2012, Impact of regional variation in detrital mineral composition on reservoir quality in deep to ultradeep lower Miocene sandstones, western Gulf of Mexico: Marine and Petroleum Geology, 35, 139153.Google Scholar
Dyer, M. J., and Bartolini, C., 2004, Sabinas Basin Lower Cretaceous to Jurassic production: comparison to South Texas equivalents: Gulf Coast Association of Geological Societies Transactions, 54, 169184.Google Scholar
Dyman, T. S., and Condon, S. M., 2006a, Assessment of undiscovered conventional oil and gas resources: Upper Jurassic–Lower Cretaceous Cotton Valley Group, Jurassic Smackover interior salt basins total petroleum system, in the East Texas Basin and Louisiana–Mississippi Salt Basins Provinces: U.S. Geological Survey, Digital Data Series DDS-69-E, chapter 2, 52 p.Google Scholar
Dyman, T. S., and Condon, S. M., 2006b, Assessment of undiscovered conventional oil and gas resources: Lower Cretaceous Travis Peak and Hosston Formations, Jurassic Smackover interior salt basins total petroleum system, in the East Texas Basin and Louisiana–Mississippi Salt Basins Provinces: U.S. Geological Survey, Digital Data Series DDS-69-E, chapter 5, 142 p.Google Scholar
Echols, D. J., and Malkin, D., 1948, Wilcox (Eocene) stratigraphy, a key to production: American Association of Petroleum Geologists Bulletin, 32, 1133.Google Scholar
Echols, R. J., Armentrout, J. M., Root, S. A., et al., 2003, Sequence stratigraphy of the Eocene/Oligocene boundary interval: southeastern Mississippi: in Prothero, D. R., Ivany, L. C., and Nesbitt, E. A. eds., From Greenhouse to Icehouse: The Marine Eocene–Oligocene Transition: Columbia University Press, 189222.Google Scholar
Eddy, D. R., Van Avendonk, J. A., Christeson, G. L., et al., 2014, Deep crustal structure of the northeastern Gulf of Mexico: implications for rift evolution and seafloor spreading: Journal of Geophysical Research: Solid Earth, 119, 68026822, doi:10.1002/2014JB011311.Google Scholar
Eddy, D. R., Avendonk, H. J. A. V., Christeson, G. L., and Norton, I. O., 2018, Structure and origin of the rifted margin of the northern Gulf of Mexico: Geosphere, 14, 114.Google Scholar
Edwards, M. B., 1981, The Live Oak delta complex, an unstable shelf/edge delta in the deep Wilcox of South Texas: American Association of Petroleum Geologists Bulletin, 77, 19421970.Google Scholar
Edwards, M. B., 1991, Control of depositional environments, eustasy, gravity, and salt tectonics in sandstone distribution in an unstable shelf edge delta, Eocene Yegua Formation, Texas and Louisiana: Gulf Coast Association of Geological Societies Transactions, 41, 237252.Google Scholar
Edwards, M. B., 2000, Origin and significance of retrograde failed shelf margins: tertiary northern Gulf Coast basin: Gulf Coast Association of Geological Societies Transactions, 50, 8193.Google Scholar
Egedahl, K., Kinsland, G. L., and Han, D., 2012, Seismic facies study of 3D seismic data, northern Louisiana, Wilcox Formation: Gulf Coast Association of Geological Societies Transactions, 62, 7391.Google Scholar
Eguiluz de Antunano, S., 2001, Geologic evolution and gas resources of the Sabinas in Northeastern Mexico: in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins and Petroleum Systems, American Association of Petroleum Geologists, Memoir 75, 241270.Google Scholar
Eguiluz de Antunano, S., 2007, Laramide deformation in the Burgos Basin, Northeastern Mexico: the Paleogene of the Gulf of Mexico and Caribbean Basins: Processes, Events and Petroleum Systems, Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 27th Annual Bob F. Perkins Research Conference, 688–702.Google Scholar
Eikrem, V., Li, R., Medeiros, M., et al., 2010, Perdido development project: great white WM12 Reservoir and Silvertip M. Frio field development plans and comparison of recent well results with pre-drill models. Offshore Technology Conference, 1–10, doi:10.4043/20879-MS.Google Scholar
Elderbak, K., Leckie, R.M., and Tibert, N.E., 2014, Paleoenvironmental and paleoceanographic changes across the Cenomanian–Turonian Boundary Event (Oceanic Anoxic Event 2) as indicated by foraminiferal assemblages from the eastern margin of the Cretaceous Western Interior Seaway: Palaeogeography, Palaeoclimatology, Palaeoecology, 413, 2948.Google Scholar
Eldrett, J. S., Ma, C., Bergman, S. C., et al., 2015, Origin of limestone–marlstone cycles: astronomic forcing of organic-rich sedimentary rocks from the Cenomanian to early Coniacian of the Cretaceous Western Interior Seaway, USA: Earth and Planetary Science Letters, 423, 98113, doi:10.1016/j.epsl.2015.04.026.Google Scholar
Ellis, B., Johnson, K. R., and Dunn, R. E., 2003, Evidence for an in situ early Paleocene rainforest from Castle Rock, Colorado: Rocky Mountain Geology, 38, 73100, doi:10.2113/gsrocky.38.1.173.Google Scholar
Enomoto, C. B., Scott, K. R., Valentine, B., et al., 2012, Preliminary evaluation of the shale gas prospectivity of the Lower Cretaceous Pearsall Formation in the onshore Gulf Coast region, United States: Gulf Coast Association of Geological Societies Transactions, 62, 93115.Google Scholar
Epstein, S. A., and Clark, D., 2009, Hydrocarbon potential of the Mesozoic carbonates of the Bahamas: Carbonates and Evaporites, 24, 97138.Google Scholar
Escalona, A., and Yang, W., 2013, Subsidence controls on foreland basin development of northwestern offshore Cuba, southeastern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 97, 125, doi:10.1306/06111212002.Google Scholar
Essex, C. W., Robinson, D. M., and Weislogel, A. L., 2016, Regional correlation of lithofacies within the Haynesville Formation from onshore Alabama: analysis and implications for provenance and paleostructure, in Lowery, C. M., Snedden, J. W., and Rosen, N. C., eds., Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Research Conference Program and Extended Abstracts, 309–343.Google Scholar
Ewing, T. E., 1991, Structure framework, in Salvador, A., ed., The Gulf of Mexico Basin: The Geology of North America: Geological Society of America, 3152, doi:10.1130/DNAG-GNA-J.31.Google Scholar
Ewing, T. E., 2007, Fairways in the Downdip and Middip Yegua Trend: a review of 25 years of exploration: Gulf Coast Association of Geological Societies Transactions, 57, 22 p.Google Scholar
Ewing, T. E., 2009, The ups and downs of the Sabine Uplift and the northern Gulf of Mexico Basin: Jurassic basement blocks, Cretaceous thermal uplifts, and Cenozoic flexure: Gulf Coast Association of Geological Societies Transactions, 59, 253269.Google Scholar
Ewing, T. E., 2010, Pre-Pearsall geology and exploration plays in South Texas: Gulf Coast Association of Geological Societies Transactions, 60, 241260.Google Scholar
Ewing, T. E., 2016 Texas Through Time: Lone Star Geology, Landscapes, and Resources, Texas Bureau of Economic Geology, 431 p.Google Scholar
Ewing, T. E., and Fergeson, G., 1991, Stratigraphic framework, structural styles and seismic signatures of Downdip Yegua Gas-condensate fields, Central Wharton County, Texas Gulf Coast: Gulf Coast Association of Geological Societies Transactions, 41, 255275.Google Scholar
Ewing, T. E., and Lopez, R. F., 1991, Principal structural features, Gulf of Mexico Basin, in Salvador, A., ed., The Gulf of Mexico Basin, Plate 2; The Geology of North America, Plate 2: Geological Society of America.Google Scholar
Ewing, T. E., and Reed, R. S., 1984, Depositional Systems and Structural Controls of Hackberry Sandstone Reservoirs in Southeast Texas: The University of Texas at Austin, Bureau of Economic Geology, Geological Circular 84-7, 48 p.Google Scholar
Ewing, T. E., and Vincent, F. S., 1997, Foundered shelf edges: examples from the Yegua and Frio, Texas and Louisiana: Gulf Coast Association of Geological Societies Transactions, 47, 149158.Google Scholar
Fan, M., Brown, E., and Li, L., 2018, Cenozoic drainage evolution of the Rio Grande paleoriver recorded in detrital zircons in South Texas: International Geology Review, doi:10.1080/00206814.2018.1446368.Google Scholar
Fang, G., 2000, Biostratigraphic and Sequence Stratigraphic Analysis of the Yegua Formation, Houston Salt Embayment, Northern Gulf of Mexico: Ph.D. dissertation; The University of Texas at Austin, 284 p.Google Scholar
Faust, M., 1984, Seismic Stratigraphy of the Middle Cretaceous Unconformity (MCU) in the Central Gulf of Mexico Basin: Master’s thesis, The University of Texas at Austin, 171 p.Google Scholar
Fedele, J. J., Hoyal, D., Barnaal, Z., Tulenko, J., and Awalt, S., 2016, Bedforms created by gravity flows, in Budd, D. A., Hajek, E. A., Purkis, S. J., eds., Autogenic Dynamics and Self-Organization in Sedimentary Systems: SEPM Special Publication 106, 95–121, doi:10.2110/sepmsp.106.12.Google Scholar
Feragen, E., Millman, D., Feldman, H., Bierley, R., and Perkins, S., 2007, South Texas sub-regional evaluation: area-wide integrated structural and stratigraphic framework of the Frio and Vicksburg Yields new plays and leads: Gulf Coast Association of Geological Societies Transactions, 57, 249253.Google Scholar
Ferrill, D. A., Smart, K. J., McGinnis, R. N., Morris, A. P., and Gulliver, K. D. H., 2017, Influence of structural position on fracturing in the Austin Chalk: Gulf Coast Association of Geological Societies Journal, 6, 189200.Google Scholar
Ferworn, K., Zumberge, J., and Brown, S., 2003, Integration of geochemistry and reservoir fluid properties: Bureau of Economic Geology PTTC Workshop, First Annual Fluids Symposium Reservoir Fluids 2003-PVT and Beyond, 51 p.Google Scholar
Fiduk, J. C., Weimer, P., Trudgill, B. D., et al., 1999, The Perdido Fold Belt, northwestern deep Gulf of Mexico, part II: seismic stratigraphy and petroleum systems: American Association of Petroleum Geologists Bulletin, 83, 578612.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, in Post, P. J., Olson, D. L., Lyons, K. T., et al., eds., 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 Bob Perkins Research Conference, 293–314.Google Scholar
Fiduk, J. C., Clippard, M., Power, S., et al., 2014, Origin, transportation, and deformation of Mesozoic carbonate rafts in the northern Gulf of Mexico: Gulf Coast Association of Geological Societies Journal, 3, 2023.Google Scholar
Finneran, J. B., 1984, Zuloaga Formation (Upper Jurassic) Shoal Complex, Sierra de Enfrente, Coahuila, Northeast Mexico: American Association of Petroleum Geologists Bulletin, 68, 476 (abst).Google Scholar
Finneran, J. M., Scott, R. W., Taylor, G. A., and Anderson, G. H., 1984, Lowermost Cretaceous ramp reefs: Knowles Limestone, Southwest Flank of the East Texas Basin, in Ventress, W. P. S., Bebout, D. G., Perkins, B. F., and Moore, C. H., eds., The Jurassic of the Gulf Rim: Society of Economic Paleontologists and Mineralogists Gulf Coast Section Foundation, 3rd Annual Research Conference, 3, 125133.Google Scholar
Fisher, W. L., and McGowen, J. H., 1967, Depositional systems in the Wilcox Group of Texas and their relationship to occurrence of oil and gas: Gulf Coast Association of Geological Societies Transactions, 17, 105125.Google Scholar
Fisher, W. L., Proctor, C. V. Jr., Galloway, W. E., and Nagle, J. S., 1970, Depositional Systems in the Jackson Group of Texas: Their Relationship to Oil, Gas and Uranium: The University of Texas at Austin, Bureau of Economic Geology, Geological Circular 70-4, 27 p.Google Scholar
Fitz-Díaz, E., Lawton, T. F., Juárez-Arriaga, E., and Chávez-Cabello, G., 2018, The Cretaceous–Paleogene Mexican orogen: structure, basin development, magmatism and tectonics: Earth Science Reviews, 183, 5684.Google Scholar
Flawn, P., Goldstein, A. Jr., King, P., and Weaver, C. E., 1961, The Ouachita System, The University of Texas at Austin, Bureau of Economic Geology, 401 p.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: American Association of Petroleum Geologists, Memoir 65, 77108.Google Scholar
Föllmi, K. B., Weissert, H., Bisping, M., and Funk, H., 1994, Phosphogenesis, carbon-isotope stratigraphy, an carbonate platform evolution along the Lower Cretaceous northern Tethyan margin: American Association of Petroleum Geologists Bulletin, 106, 729746.Google Scholar
Forgotson, J. M., and Forgotson, J. M. Jr., 1976, Definition of Gilmer Limestone, Upper Jurassic formation, northeast Texas: American Association of Petroleum Geologists Bulletin, 60, 11191123.Google Scholar
Foss, D. C., 1979, Depositional environment of Woodbine sandstones, Polk County, Texas: Gulf Coast Association of Geological Societies Transactions, 29, 8394.Google Scholar
Fouad, K., Brown, L. F. Jr., Ambrose, W. A., et al., 2009, Genetic facies analysis using seismic geomorphology and seismic attributes in the continental shelf of Eastern Mexico: petroleum systems in the southern Gulf of Mexico, in Bartolini, C. and Roman Ramos, J. R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 471476.Google Scholar
Frazier, D. E., 1974, Depositional Episodes: Their Relationship to the Quaternary Stratigraphic Framework in the Northwestern Portion of the Gulf Basin: The University of Texas at Austin, Bureau of Economic Geology, Geological Circular 74-1, 28 p.Google Scholar
Frébourg, G., Ruppel, S. C., Loucks, R. G., and Lambert, J., 2016, Depositional controls on sediment body architecture in the Eagle Ford/Boquillas system: insights from outcrops in west Texas, United States: American Association of Petroleum Geologists Bulletin, 100, 657682.Google Scholar
Frederick, B. C., Blum, M. D. and Snedden, J. W., in review, Pre-Salt Eagle Mills Formation and coeval siliciclastic sources and sinks, Northern Gulf of Mexico basin: Geological Society of America Bulletin.Google Scholar
Fredrich, J. T., Fossum, A. F., and Fossum, R. J., 2007, Mineralogy of deepwater Gulf of Mexico salt formations and implications for constitutive behavior: Journal of Petroleum Science and Engineering, 57, 354374.Google Scholar
Freeman-Lynde, R. P., 1983, Cretaceous and Tertiary samples dredged from the Florida Escarpment: Gulf Coast Association of Geological Societies Transactions, 29, 91100.Google Scholar
Fritz, D. A., Belsher, T. W., Medlin, J. M., et al., 2000, New exploration concepts for the Edwards and Sligo Margins, Cretaceous of Onshore Texas: American Association of Petroleum Geologists Bulletin, 84, 905922.Google Scholar
Frost, W. G., 2010, The somewhat accidental discovery of the Mobile Bay gas field: a story of perseverance and good fortune: American Association of Petroleum Geologists, Search and Discovery 110133, 27 p.Google Scholar
Frush, M. P., and Eicher, D. L., 1975, Cenomanian and Turonian foraminifera and paleoenvironments in the Big Bend region of Texas and Mexico, in Caldwell, W. G. E., ed., The Cretaceous System in the Western Interior of North America, The Geological Society of Canada Special Paper 13, 277301.Google Scholar
Fryberger, S. G., Al-Sari, A. M., and Clisham, T. J., 1983, Eolian dune, interdune, sand sheet, and silica clastic sabkha sediments of an offshore prograding sand sea, Dhahran area, Saudi Arabia: American Association of Petroleum Geologists Bulletin, 67, 280312.Google Scholar
Galloway, W. E., 1968, Depositional systems of the Lower Wilcox Group, north-central Gulf Coast basin: Gulf Coast Association of Geological Societies Transactions, 18, 275289.Google Scholar
Galloway, W. E., 1977, Catahoula Formation of the Texas Coastal Plain: Depositional Systems, Mineralogy, Structural Development, Ground-Water Flow History, and Uranium Distribution: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 87, 59 p.Google Scholar
Galloway, W. E., 1986a, Reservoir facies architecture of microtidal barrier systems: American Association of Petroleum Geologists Bulletin, 70, 787808.Google Scholar
Galloway, W. E., 1986b, Depositional and Structural Framework of the Distal Frio Formation, Texas Coastal Zone and Shelf: The University of Texas at Austin, Bureau of Economic Geology, Geological Circular 86-8, 16 p.Google Scholar
Galloway, W. E., 1989a, Genetic stratigraphic sequences in basin analysis I: architecture and genesis of flooding-surface bounded depositional units: American Association of Petroleum Geologists Bulletin, 73, 125142, doi:10.1306/703C9AF5-1707-11D7-8645000102C1865D.Google Scholar
Galloway, W. E., 1989b, Genetic stratigraphic sequences in basin analysis II: application to Northwest Gulf of Mexico Cenozoic Basin: American Association of Petroleum Geologists Bulletin, 73, 143154, doi:10.1306/703C9AFA-1707-11D7-8645000102C1865D.Google Scholar
Galloway, W. E., 1998, Siliciclastic slope and base-of-slope depositional systems: component facies, stratigraphic architecture, and classification: American Association of Petroleum Geologists Bulletin, 82, 569595.Google Scholar
Galloway, W. E., 2002, Cenozoic evolution of sedimentation accumulation in deltaic and shore-zone depositional systems, northern Gulf of Mexico basin: Marine and Petroleum Geology, 18, 10311040.Google Scholar
Galloway, W. E., 2005a, Cenozoic evolution of the northern Gulf of Mexico continental margin, in Post, P. J., Rosen, N. C., Olson, D. L., et al., eds., Petroleum Systems of Divergent Continental Margins: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 25th Annual Bob F. Perkins Research Conference, 613–633.Google Scholar
Galloway, W. E., 2005b, Gulf of Mexico Basin depositional record of Cenozoic North American drainage basin evolution, in Blum, M. D., Marriott, S. B., and Leclair, S. F., eds., Fluvial Sedimentology VII: International Association of Sedimentologists Special Publication 35, Wiley-Blackwell, 409423.Google Scholar
Galloway, W. E., 2007, Wilcox canyons: distribution, attributes, origins and relationship to basinal sands, 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: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 27th Annual Bob F. Perkins Research Conference, 101–102.Google Scholar
Galloway, W. E., 2008, Depositional evolution of the Gulf of Mexico sedimentary basin, in Hsu, K. J., ed., Sedimentary Basins of the World: Elsevier, 5, 505549, doi:10.1016/S1874-5997(08)00015-4.Google Scholar
Galloway, W. E, 2009, The Gulf of Mexico: geologic foundations of a petroleum mega-province: Geo Expro, 6, 2226.Google Scholar
Galloway, W. E., and Cheng, E. S., 1985, Reservoir Facies Architecture in a Microtidal Barrier System: Frio Formation, Texas Gulf Coast, The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 144, 36 p.Google Scholar
Galloway, W. E., and Hobday, D. K., 1996, Terrigenous clastic depositional systems: Applications to Fossil Fuel and Groundwater Resources, 489, doi:10.1007/978-3-642-61018-9.Google Scholar
Galloway, W. E., and Kaiser, W. R., 1980, Catahoula Formation of the Texas Coastal Plain: Origin, Geochemical Evolution, and Characteristics of Uranium Deposits: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 100, 81 p.Google Scholar
Galloway, W. E., and McGilvery, T. A., 1995, Facies of a Submarine Canyon Fill Reservoir Complex, Lower Wilcox Group (Paleocene), Central Texas Coastal Plain: Society of Economic Paleontologists and Mineralogists, Special Publication 20, 123.Google Scholar
Galloway, W. E., and Morton, R. A., 1989, Geometry, genesis, and reservoir characteristics of shelf sandstone facies, Frio Formation (Oligocene), Texas Coastal Plain, in Morton, R.A. and Nummedal, D., eds., Shelf Sedimentation, Shelf Sequences, and Related Hydrocarbon Accumulation: Society of Economic Paleontologists and Mineralogists Foundation Gulf Coast Section 7th Annual Research Conference, 89–115.Google Scholar
Galloway, W. E., and Williams, T. A., 1991, Sediment accumulation rates in time and space: Paleogene genetic stratigraphic sequences of the northwestern Gulf of Mexico basin: Geology, 19, 986989.2.3.CO;2>CrossRefGoogle Scholar
Galloway, W. E., Finley, R. J. and Henry, C. D., 1979, South Texas Uranium Province: Geologic Perspective: The University of Texas at Austin, Bureau of Economic Geology, Guidebook 18, 81 p.Google Scholar
Galloway, W. E., Hobday, D. K., and Kinji, M., 1982a, Frio Formation of Texas Gulf Coastal Plain: depositional systems, structural framework, and hydrocarbon distribution: American Association of Petroleum Geologists Bulletin, 66, 649688.Google Scholar
Galloway, W. E., Hobday, D. K., and Kinji, M., 1982b, Frio Formation of Texas Gulf Coastal Basin: Depositional Systems, Structural Framework, and Hydrocarbon Origin, Migration, Distribution, and Exploration Potential: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 122, 78 p.Google Scholar
Galloway, W. E., Henry, C. D., and Smith, G. E., 1982c, Depositional Framework, Hydrostratigraphy, and Uranium Mineralization of the Oakville Sandstone (Miocene), Texas Coastal Plain: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 113, 51 p.Google Scholar
Galloway, W. E., Ewing, T. E. Garrett, C. M., Tyler, N., and Bebout, D. G., 1983, Atlas of Major Texas Oil Reservoirs: The University of Texas at Austin, Bureau of Economic Geology, 139 p.Google Scholar
Galloway, W. E., Jirik, L. A., Morton, R. A., and DuBar, J. R., 1986, Lower Miocene (Fleming) Depositional Episode of the Texas Coastal Plain and Continental Shelf: Structural Framework, Facies, and Hydrocarbon Resources: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 150, 50 p.Google Scholar
Galloway, W. E., Bebout, D. G., Fisher, W. L., et al., 1991a, Cenozoic, in Salvador, A., ed., The Gulf of Mexico Basin, The Geology of North America, Geological Society of America, 245324.Google Scholar
Galloway, W. E., Dingus, W. F., and Paige, R. E., 1991b, Seismic and depositional facies of Paleocene–Eocene Wilcox Group submarine canyon fills, northwest Gulf Coast, U.S.A., in Weimer, P. and Link, M. H., eds., Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems: Springer-Verlag, 247271.Google Scholar
Galloway, W. E., Liu, X., Neuberger, D. T., and Xue, L., 1994, Reference High-Resolution Correlation Cross Sections, Paleogene Section, Texas Coastal Plain: The University of Texas at Austin, Bureau of Economic Geology.Google Scholar
Galloway, W. E., Ganey-Curry, P. E., Li, X., and Buffler, R. T., 2000, Cenozoic depositional history of the Gulf of Mexico basin: American Association of Petroleum Geologists Bulletin, 84, 17431774.Google Scholar
Galloway, W. E., Whiteaker, T. L., and Ganey-Curry, P. E., 2011, History of Cenozoic North American drainage basin evolution, sediment yield, and accumulation in the Gulf of Mexico basin: Geosphere, 7, 938973.Google Scholar
Garcia-Molina, G., 1994, Structural Evolution of Southeastern Mexico (Chiapas-Tabasco-Campeche) Offshore and Onshore: Ph.D. dissertation; Rice University, 161 p.Google Scholar
Gawloski, T., 1983, Stratigraphy and environmental significance of the continental Triassic rocks of Texas: Baylor Geological Studies Bulletin, 41, 148.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
Gehrels, G., and Pecha, M., 2014, Detrital zircon U–Pb geochronology and Hf isotope geochemistry of Paleozoic and Triassic passive margin strata of western North America, Geosphere, 10, 4965, doi:10.1130/GES00889.1.Google Scholar
GeoMark Research Ltd. and TDI-Brooks International Inc., 2005a, 1997, 1999, 2005 Western Gulf of Mexico surface geochemical exploration programs: geochemical correlation of seep oils with offshore petroleum systems: Proprietary study.Google Scholar
GeoMark Research Ltd. and TDI-Brooks International Inc., 2005b, 1997–2001 Central/Eastern Gulf of Mexico surface geochemical exploration programs: geochemical correlation of seep oils with offshore petroleum systems: Proprietary study.Google Scholar
Gleason, J. D., Gehrels, G. E., Dickinson, W. R., Patchett, P. J., and Kring, D. A., 2007, Laurentian sources for detrital zircon grains in turbidite and deltaic sandstones of the Pennsylvanian Haymond Formation, Marathon Assemblage, West Texas, U.S.A: Journal of Sedimentary Research, 77, 888900, doi:10.2110/jsr.2007.084.Google Scholar
Glennie, K. W., 1972, Permian Rotliegendes of Northwest Europe interpreted in light of modern desert sedimentation studies: American Association of Petroleum Geologists Bulletin, 56, 10481071.Google Scholar
Godo, T. J., 2006, Identification of stratigraphic traps with subtle seismic amplitude effects in Miocene channel/levee sand systems, NE Gulf of Mexico, in Allen, M. R., Goffey, G. P., Morgan, R. K., and Walker, I. M., eds., The Deliberate Search for the Stratigraphic Trap: Geological Society, Special Publication 254, 127151, doi:10.1144/GSL.SP.2006.254.01.07.Google Scholar
Godo, T., 2017, The Appomattox field: Norphlet aeolian sand dune reservoirs in the deep-water Gulf of Mexico, in Merrill, R. K. and Sternbach, C. A., eds., Giant Fields of the Decade 2000–2010: American Association of Petroleum Geologists, Memoir 113, 2954.Google Scholar
Godo, T. J., Chuparova, E., and McKinney, D. E., 2011, Norphlet aeolian sand fairway established in the deep water Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 90124, 2 p.Google Scholar
Goggin, K., and Rine, J., 2014, Geological Characterization of a Portion of the South Georgia Rift Basin for Source Proximal CO2, Storage Based on Analysis of Samples from Rizer #1 Test Borehole, Colleton County, SC and USGS Clubhouse Crossroads Test Hole no. 3, Dorchester County, SC: Weatherford Laboratories, reports HH-57745 and HH-59364.Google Scholar
Gohrbandt, K. H., 2002, Potential gas resources under the West Florida shelf and slope and their development: Gulf Coast Association of Geological Societies Transactions, 52, 351360.Google Scholar
Goldhammer, R. K., and Johnson, C. A., 2001, Middle Jurassic–Upper Cretaceous paleogeographic evolution and sequence-stratigraphy framework of the Northwest Gulf of Mexico Rim, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 4581.Google Scholar
Gomez, E., Doe, M., Villarroel, S., et al., 2018, Recent Yucatan seismic survey revealing a new frontier exploration in the Gulf of Mexico: American Association Petroleum Geologists, Search and Discovery 30582, 1–6.Google Scholar
Gomez-Cabrera, P. T., and Jackson, M. P. A., 2009a, Regional Neogene salt tectonics in the Offshore Salina del Istmo Basin, Southeastern Mexico, in Bartolini, C. and Roman Ramos, J.R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 128.Google Scholar
Gomez-Cabrera, P. T., and Jackson, M. P. A., 2009b, Neogene stratigraphy and salt tectonics of the Santa Ana area, Offshore Salina del Istmo Basin, Southeastern Mexico, in Bartolini, C. and Roman Ramos, J. R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 237255, doi:10.1306/13191086M9037.Google Scholar
Gong, C., Steel, R. J., Wang, Y., Lin, C., and Olariu, C., 2016, Grain size and transport regime at shelf edge as fundamental controls on delivery of shelf-edge sands to deepwater: Earth-Science Reviews, 157, 3260.Google Scholar
Gonzales, E., and Medrano, M., 2014, Structural slope fans resulting from Paleogene compression in the Veracruz Basin, Mexico: American Association of Petroleum Geologists, Search and Discovery 50963, 17 p.Google Scholar
Gonzalez-Posadas, J. F., Avendano-Lopez, S., Molina, J., 2005, Strike–slip model for the Jacinto and Paredon Fields of the Chiapas-Tabasco Region, South East Basin, Mexico: American Association of Petroleum Geologists, Search and Discovery 20028, 17.Google Scholar
Gordon, M. B., Mann, P., Caceres, D., and Flores, R., 1997, Cenozoic tectonic history of the North America–Caribbean plate boundary zone in western Cuba: Journal of Geophysical Research, 102, 1005510082, doi:10.1029/96JB03177.Google Scholar
Grabowski, G. J., 1981, Source-rock potential of the Austin Chalk, Upper Cretaceous, Southeast Texas: Gulf Coast Association of Geological Societies Transactions, 31, 105113.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
Gradstein, F. M., Ogg, J. G., Schmitz, M., and Ogg, G., 2012, The Geologic Time Scale 2012: Elsevier, 1176 p.Google Scholar
Grajales-Nishimura, J. M., Cedillo-Pardo, E., Rosales-Domínguez, C., et al., 2000, Chicxulub impact: the origin or reservoir and seal facies in the southeastern Mexico oil fields: Geology, 28, 307310, doi:10.1130/0091-7613(2000)28<307:CITOOR>2.0.CO;2.Google Scholar
Gray, G. G., Pottorf, R. J., Yurewicz, D. A., et al., 2001, Thermal and chronological record of syn- to post- Laramide burial and exhumation, Sierra Madre Oriental, Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 159181.Google Scholar
Greene, T. J., and O’Neill, B. E., 2005, Building a facies-based permeability model for deep-water Miocene reservoirs, eastern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 55, 268277.Google Scholar
Gregory, J. L., 1966, A lower Oligocene delta in the subsurface of southeastern Texas: Gulf Coast Association of Geological Societies Transactions, 16, 214227.Google Scholar
Grice, K., Cao, C., Love, G. D., et al. 2005, Photic zone euxinia during the Permia–Triassic superanoxic event: Science, 307, 706709.Google Scholar
Gulick, S. P. S., Barton, P. J., Christeson, G. L., et al., 2008, Importance of pre-impact crustal structure for the asymmetry of the Chicxulub impact crater: Nature Geoscience, 1, 131135, doi:10.1038/ngeo103.Google Scholar
Gulick, S. P. S., Christeson, G. L., Barton, P. J., et al., 2013, Geophysical characterization of the Chicxulub impact crater: Review of Geophysics, 51, 3152, doi:10.1002/rog.20007.Google Scholar
Gutiérrez, M. A., 2018, Systematic Lithologic Characterization of Pleistocene Mass-Transport Deposit, Mississippi Canyon of the Northern Gulf of Mexico, USA: MS thesis, The University of Texas at Austin, 102 p.Google Scholar
Gutiérrez Paredes, H. C., Cantuneanu, O., and Hernandez Romano, U., 2017, Sequence stratigraphy of the Miocene section, southern Gulf of Mexico: Marine and Petroleum Geology, 86, 711732, doi:10.1016/j.marpetgeo.2017.06.022.Google Scholar
Gutiérrez-Puente, N. A., 2006, Estudio Micropaleontologico y Bioestratigrafico de la Columna del Pozo Cupelado – 10 (Cuenca Tampico Misantla): Master’s thesis, Universidad Nacional Autonoma de Mexico, 102 p.Google Scholar
Gutteridge, P., Poprawski, Y., and Horbury, A., 2019, Salt–carbonate interactions in the Sureste Basin, SE Mexico: depositional models and analogs for Cretaceous carbonate breccias: The Geological Society, London, Petroleum Geology of Mexico and the Northern Caribbean Conference (abstract).Google Scholar
Guzmán, , 2001, Exploration and production in México: challenges and opportunities: American Association of Petroleum Geologists, Search and Discovery, 10222, 47 p.Google Scholar
Guzmán, A. E., 2013, Petroleum history of Mexico: how it got to where it is today: American Association of Petroleum Geologists, Search and Discovery 10530, 27 p.Google Scholar
Guzmán, A. E., 2018, The upstream in México under the new energy reform: American Association of Petroleum Geologists, Search and Discovery 703, 19 p.Google Scholar
Guzmán-Vega, M. A., and Mello, M. R., 1999, Origin of oil in the Sureste Basin, Mexico: American Association of Petroleum Geologists Bulletin, 83, 10681095.Google Scholar
Guzmán-Vega, M. A., Castro Ortíz, L., Román-Ramos, J. R., et al., 2001, Classification and origin of petroleum in the Mexican Gulf Coast Basin: an overview, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems, American Association of Petroleum Geologists, Memoir 75, 127142.Google Scholar
Hackley, P. C., 2012, Geological and geochemical characterization of the Lower Cretaceous Pearsall Formation, Maverick Basin, south Texas: a future shale gas resource?: American Association of Petroleum Geologists Bulletin, 96, 14491482, doi:10.1306/11221111071.Google Scholar
Hackley, P. C., and Ewing, T. E., 2010, Assessment of undiscovered conventional oil and gas resources, onshore Claiborne Group, United States part of the northern Gulf of Mexico Basin: American Association of Petroleum Geologists Bulletin, 94,16071636, doi:10.1306/04061009139.Google Scholar
Haczewski, G., 1976, Sedimentological reconnaissance of the San Cayetano Formation: an accumulative continental margin in the Jurassic of western Cuba: Acta Geologica Polonica, 26.2, 331353.Google Scholar
Haddad, S. A., and Mancini, E. A., 2013, Reservoir characterization modeling, and evaluation of Upper Jurassic Smackover microbial carbonate and associated facies in Little Cedar Creek Field, Southwest Alabama, Eastern Gulf Coastal Plain of the United States: American Association of Petroleum Geologists Bulletin, 97, 20592083.Google Scholar
Halbouty, M. T., and Hardin, G. C., 1956, Genesis of salt domes of Gulf Coastal Plain: American Association of Petroleum Geologists Bulletin, 40, 737746.Google Scholar
Hall, D. J., Bowen, B. E., Rosen, R. N., Wu, S., and Bally, A. W., 1993, Mesozoic and Early Cenozoic development of the Texas Margin: a new integrated cross-section from the Cretaceous Shelf Edge to the Perdido Fold Belt: Society of Economic Paleontologists and Mineralogists Gulf Coast Section 13th Annual Research Conference, 21–31.Google Scholar
Hamlin, H. S., 1988, Depositional and Ground-Water Flow Systems of the Carrizo-Upper Wilcox, South Texas: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 175, 61 p.Google Scholar
Hammes, U., and Frébourg, G., 2012, Haynesville and Bossier mudrocks: a facies and sequence stratigraphic investigation, East Texas and Louisiana, USA: Marine and Petroleum Geology, 31, 826, doi:10.1016/j.marpetgeo.2011.10.001.Google Scholar
Hammes, U., Hamlin, H. S., and Ewing, T. E., 2011, Geologic analysis of the Upper Jurassic Haynesville Shale in east Texas and west Louisiana: American Association of Petroleum Geologists Bulletin, 95, 16431666, doi:10.1306/02141110128.Google Scholar
Hammes, U., Eastwood, R., McDaid, G., et al., 2016, Regional assessment of the Eagle Ford Group of South Texas, USA: insights from lithology, pore volume, water saturation, organic richness, and productivity correlations: Interpretation, 4, 125150, doi:10.1190/int-2015-0099.1.Google Scholar
Handford, C. R., and Loucks, R. G., 1993, Carbonate depositional sequences and systems tracts: responses of carbonate platforms to relative sea-level changes, in Loucks, R. G., and Sarg, J. F., eds., Carbonate Sequence Stratigraphy: Recent Developments and Applications: American Association of Petroleum Geologists, Memoir 57, 342.Google Scholar
Haq, B. U., 2014, Cretaceous eustasy revisited: Global and Planetary Change, 113, 4458, doi:10.1016/j.gloplacha.2013.12.007.Google Scholar
Haq, B. U., 2017, Jurassic sea-level variations: a reappraisal: GSA Today, 28, 7, doi: 10.1130/GSATG359A.1.Google Scholar
Haq, B. U., Hardenbol, J., and Vail, P. R., 1987, Chronology of fluctuating sea levels since the Triassic: Science, 235, 11561167.Google Scholar
Harbour, J. L., and Mathis, R. L., 1984, Sedimentation, diagenesis and porosity evolution of carbonate sands in the Black Lake Field of Central Louisiana, in Harris, P. M., ed., Society of Economic Paleontologists and Mineralogists Core Workshop 5: Carbonate Sands, 306–333.Google Scholar
Hardenbol, J., Thierry, J., Farley, M. B., et al., 1998, Mesozoic and Cenozoic sequence chronostratigraphic framework of European basins, in de Graciansky, P. C., Hardenbol, J., Jacquin, T., and Vail, P., eds., Mesozoic and Cenozoic Sequence Stratigraphy of European Basins: Society of Economic Paleontologists and Mineralogists, Special Publication 60, 314.Google Scholar
Harding, A., Walker, L., Ehlinger, S., and Chapman, T., 2016, The siliciclastic Upper Cretaceous play of Eastern Mississippi Canyon, in Lowery, C. M. and Snedden, J. W., eds. Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Research Conference Program and Extended Abstracts.Google Scholar
Harris, P. M., 2008, Stratigraphic framework and new exploration concepts for the Lower Cretaceous Shelf Margin carbonates of Texas: American Association of Petroleum Geologists, Search and Discovery 40303, 41 p.Google Scholar
Harris, P. M., and Kowalik, W. S., 2005, Satellite Images of Carbonate Depositional Settings, American Association of Petroleum Geologists, Methods in Exploration 11, digital reprint.Google Scholar
Hart, W., Jacek, J., and Martin, A., 2004, Recognition and exploration significance of supra-salt stratal carapaces, in Post, P. J., Olson, D. L., Lyons, K. T., et al., eds., Salt–Sediment Interactions and Hydrocarbon Prospectivity: Concepts, Applications, and Case Studies for the 21st Century: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 21st Annual Research Conference Program and Extended Abstracts, 166–199.Google Scholar
Hartley, A. J., Weissman, G., and Scuderi, L., 2015, Controls on the apex location of large deltas: Journal of the Geological Society, 174, 1013.Google Scholar
Harwood, G., and Fontana, C., 1984, Smackover deposition and diagenesis and structural history of the Bryan’s Mill Area, Cass and Bowie Counties, Texas: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 3rd Annual Research Conference Proceedings, 135–147.Google Scholar
Haymond, D., 1991, The Austin Chalk: an overview: Houston Geological Society Bulletin, 24, 2127.Google Scholar
Heatherington, A. L., and Mueller, P. A., 2003, Mesozoic igneous activity in the Suwannee Terrane, Southeastern USA: petrogenesis and Gondwanan Affinities: Gondwana Research, 6, 296311, doi:10.1016/S1342-937X(05)70979-5.Google Scholar
Heffner, D. M., 2013, Tectonics of the South Georgia Rift: Ph.D. dissertation, The University of South Carolina, 178 p.Google Scholar
Heintz, M. L., Yancey, T. E., Miller, B. V., and Heizler, M. T., 2015, Tephrochronology and geochemistry of Eocene and Oligocene volcanic ashes of east and central Texas: Geological Society of America Bulletin, 127, 770780.Google Scholar
Helland-Hansen, W., Sømme, T. O., Martinsen, O. J., Lunt, I., and Thurmond, J., 2016, Deciphering Earth’s natural hourglasses: perspectives on source-to-sink analysis: Journal of Sedimentary Research, 86, 10081033, doi:10.2110/jsr.2016.56.Google Scholar
Henry, L. C., Wadsworth, J. A., Hansen, B., 2017, Visualizing a sub-salt field with image logs: image facies, mass transport complexes, and reservoir implications from Thunder Horse, Mississippi Canyon, Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 10938, 21 p.Google Scholar
Hentz, T. F., and Ruppel, S. C., 2011, Regional stratigraphic and rock characteristics of Eagle Ford Shale in its play area: Maverick Basin to East Texas Basin: American Association of Petroleum Geologists, Search and Discovery 10325, 20 p.Google Scholar
Hermann, L. A., 1971, Lower Cretaceous Sligo reef trends in central Louisiana: Gulf Coast Association of Geological Societies Transactions, 21, 187198.Google Scholar
Hernandez Calvento, L., Jackson, D. W. T., Cooper, A., and Perez-Chacon, E., 2017, Island-encapsulating eolian sedimentary systems of the Canary and Cape Verde archipelagos: Journal of Sedimentary Research, 87, 117125, doi:10.2110/jsr.2017.6.Google Scholar
Hernandez-Mendoza, H. S., 2013, Stratigraphic Characterization and Evolution of a Mid-Tertiary Age Deep Water System, Holok Area, SW Gulf of Mexico: Ph.D. dissertation: University of Aberdeen, 358 p.Google Scholar
Hernandez-Mendoza, J. J., 2000, Interpretation and Distribution of Depositional Systems: Oligocene Frio Depisode in the Subsurface of Burgos Basin, Northeastern Mexico: MS thesis, The University of Texas at Austin, 102 p.Google Scholar
Hernandez-Mendoza, J. J., DeAngelo, M. V., Wawrzyniec, T. F., and Hentz, T. F., 2008a, Miocene chronostratigraphy, paleogeography, and play framework of the Burgos Basin, southern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 92, 15011535.Google Scholar
Hernandez-Mendoza, J. J., DeAngelo, M. V., Wawrzyniec, T. F., and Hentz, T. F., 2008b, Major structural elements of the Miocene section, Burgos Basin, northeastern Mexico: American Association of Petroleum Geologists Bulletin, 92, 14791499.Google Scholar
Herries, R. D., 1993, Contrasting styles of fluvial-aeolian interaction at a downwind erg margin: Jurassic Kayenta–Navajo transition, northeastern Arizona US, in North, C. P. and Prosser, D. J., eds., Characterization of Fluvial and Aeolian Reservoirs, London Geological Society, Special Publication 73, 199218.Google Scholar
Herron, D. A., 2011, First Steps in Seismic Interpretation: Society of Exploration Geophysicists, Geophysical Monograph 16, 217 p., doi:10.1190/1.9781560802938.Google Scholar
Herron, D. A., 2014, Thoughts and observation on interpreting depth-imaged data in the Jurassic Norphlet Play, Deepwater Eastern Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 41342, 16 p.Google Scholar
Hessler, A. M., Zhang, J., Covault, J., and Ambrose, W., 2017, Continental weathering coupled to Paleogene climate changes in North America: Geology, 45, 911914.Google Scholar
Hessler, A. M., Covault, J. A., Stockli, D., and Fildani, A., 2018, Late Cenozoic cooling favored glacial over tectonic controls on sediment supply to the western Gulf of Mexico: Geology, 46, 955998, doi:10.1130/G45528.1.Google Scholar
Hildebrand, A. R., Penfield, G. T., Kring, D. A., et al., 1991, Chicxulub crater: a possible Cretaceous/Tertiary boundary impact crater in the Yucatán Peninsula, Mexico. Geology, 19, 867871, doi:10.1130/0091-7613(1991)019<0867:CCAPCT>2.3.CO;2.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: American Association of Petroleum Geologists, 2548.Google Scholar
Horbury, A. D., 2000, Report detailing the carbonate sequence stratigraphy and reservoir geology of the Cantarell Field, Cuidad Carmen, Campeche: Pemex internal report.Google Scholar
Horbury, A. D., Celestino, J., Oxtoby, N., Soto, A., and Johnson, S., 1996, Diagénesis y evolución de la porosidad en el campo Petrolifero Arenque, costa afuera de Tajypjco, Tamaulipas, Mexico: Boletín de la Asociación Mexicana de Geólogos Petroleros, 45, 5880.Google Scholar
Horbury, A. D, Fall, S. I., Gonzalez-I, F.., et al., 2003, Tectonic sequence stratigraphy of the western margin of the Gulf of Mexico in the late Mesozoic and Cenozoic: less passive than previously imagined, in Bartolini, C., Bufflcr, R. T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists, Memoir 79, 184245.Google Scholar
Horn, B. W., 2012, Identifying new exploration fairways in the Gulf of Mexico: Deepwater Tuscaloosa/Woodbine play: Gulf Coast Association of Geological Societies Transactions, 61, 245256.Google Scholar
Hovorka, S. D., and Nance, H. S., 1994, Dynamic depositional and early diagenetic processes in a deep-water shelf setting, Upper Cretaceous Austin Chalk, North Texas: Gulf Coast Association of Geological Societies Transactions, 44, 269276.Google Scholar
Hudec, M. R., and Jackson, M. P. A., 2011, The Salt Mine: a Digital Atlas of Salt Tectonics: American Association of Petroleum Geologists, Memoir 99, 305 p.Google Scholar
Hudec, M. R., and Norton, I. O., 2018, Upper Jurassic structure and evolution of the Yucatán and Campeche subbasins, southern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, doi: 10.1306/11151817405.Google Scholar
Hudec, M. R., Norton, I. O., Jackson, M. P. A., and Peel, F. J., 2013a, Jurassic evolution of the Gulf of Mexico salt basin: American Association of Petroleum Geologists Bulletin, 97, 16831710, doi:10.1306/04011312073.Google Scholar
Hudec, M. R., Jackson, M. P. A., and Peel, F. J., 2013b, Influence of deep Louann structure on the evolution of the northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 97, 17111735, doi:10.1306/04011312074.Google Scholar
Hudec, M.R., Dooley, T.P., Peel, F.J., and Soto, J.I., accepted, Structure and evolution of the Salina del Bravo region, northeastern Mexico continental slope: Geological Society of America Bulletin.Google Scholar
Hull, D. C., 2011, Stratigraphic Architecture, Depositional Systems, and Reservoir Characteristics of the Pearsall Shale-Gas System, Lower Cretaceous, South Texas: Master’s thesis, The University of Texas at Austin, 192 p.Google Scholar
Hull, D., and Loucks, R., 2010, Depositional systems and stratal architecture of the Lower Cretaceous (Aptian) Pearsall Formation in South Texas: Gulf Coast Association of Geological Societies Transactions, 60, 901906.Google Scholar
Humphrey, W. E., and Diaz, T.,, 2003, Jurassic and Lower Cretaceous Stratigraphy and Tectonics of Northeast Mexico, The University of Texas at Austin, Bureau of Economic Geology, 167 p.Google Scholar
Hunt, B. W., 2013, Regional Norphlet Facies Correlation, Analysis and Implications for Paleostructure and Provenance, Eastern Gulf of Mexico: Master’s thesis, The University of Alabama, 112 p.Google Scholar
Hunt, B., Robinson, D. M., Weislogel, A. L., and Ewing, R. C., 2017, Sediment source regions and paleotransport of the Upper Jurassic Norphlet Formation, eastern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 101, 15191542.Google Scholar
Iannello, C., 2001, Regional Characteristics, Timing, and Significance of Dissolution and Collapse Features in Lower Cretaceous Carbonate Platform Strata, DeSoto Canyon Area, Offshore Alabama–Florida: Master’s thesis, Texas A&M University, 73 p.Google Scholar
Ice, R. G., and McNulty, C. L., 1980, Foraminifers and calcispheres from the Cuesta del Cura and Lower Agua Nueva(?) Formations (Cretaceous) in east-central Mexico: Gulf Coast Association of Geological Societies Transactions, 30, 403425.Google Scholar
Imbert, P., and Phillippe, Y., 2005, The Mesozoic opening of the Gulf of Mexico: Part 2, integrating seismic and magnetic data into a general opening model: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 25th Annual Research Conference Program and Extended Abstracts, 1151–1189, doi:10.5724/gcs.05.25.1151.Google Scholar
IUGS International Commission on Stratigraphy, 2011, Global chronostratigraphical correlation table for the last 2.7 million years: stratigraphic charts for the Quaternary, 2011.Google Scholar
Jackson, M. P. A., and Cramez, C., 1989, Seismic recognition of salt welds in salt tectonics regimes in Gulf Coast Section of the Society of Economic Paleontologists and Mineralogists Tenth Annual Research Conference Program and Abstracts, 66–71.Google Scholar
Jackson, M. P. A., and Hudec, M. R., 2017, Salt Tectonics: Principles and Practice: Cambridge University Press, 498 p.Google Scholar
Jackson, M. P. A., and Seni, S. J., 1984, Atlas of Salt Domes in the East Texas Basin: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 140, 102 p.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 Science, 22, 93117, doi:10.1146/annurev.ea.22.050194.000521.Google Scholar
Jackson, M. P. A., 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
Jacobo Albarabn, J., Garduño, M., Innocenti, F., Pasquare, M. G., and Tonarini, S., 1992. Datos sobre el vulcanismo neogénico-reciente del complejo volcánico de Los Tuxtlas, Edo. de Veracruz, México: Evolución petrológica y geovulcanológica, 11.o Convención Geológica Nacional, Veracruz, Libro de Resúmenes, 97–98.Google Scholar
Jacques, J. M., and Clegg, H., 2002, Late Jurassic source rock distribution and quality in the Gulf of Mexico: inferences from plate tectonic modelling: Gulf Coast Association of Geological Societies Transactions, 52, 429440.Google Scholar
James, A. T., Wenger, L. M., Melia, M. B., Ross, A. H., and Kuminez, C. P., 1993, Recognition of a new hydrocarbon play in a mature exploration area through integration of geochemical, palynologic, geologic, and seismic interpretations (onshore northern Gulf of Mexico) (abs.): American Association of Petroleum Geologists, Search and Discovery 90987, 123 p.Google Scholar
Janson, X., 2004, Golden Lane Platform Northeastern Margin: large scale margin erosion and karst features: American Association of Petroleum Geologists Annual Convention and Exposition, Dallas, Texas (Abstract).Google Scholar
Janson, X., Loucks, R., Kerans, C., Marhx, A., and Reyes, C., 2004, Karstification of the Lower Cretaceous Tuxpan Detached Platform (abs.): American Association of Petroleum Geologists, Search and Discovery 90026, 1 p.Google Scholar
Janson, X., Kerans, C., Loucks, R., et al., 2011, Seismic architecture of a Lower Cretaceous platform-to-slope system, Santa Agueda and Poza Rica fields, Mexico: American Association of Petroleum Geologists Bulletin, 95, 105146, doi:10.1306/06301009107.Google Scholar
Jenkyns, H.C., 2010, Geochemistry of oceanic anoxic events: Geochemistry, Geophysics, Geosystems, 11, Q03004.Google Scholar
Jennette, D., Wawrzyniec, T., Fouad, K., et al., 2003, Traps and turbidite reservoir characteristics from a complex and evolving tectonic setting, Veracruz Basin, southeastern Mexico: American Association of Petroleum Geologists Bulletin, 87, 15991622.Google Scholar
John, C. J., Maciasz, G., and Harder, B. J., 1998, Gulf Coast Geopressured-Geothermal Program Summary Report Compilation: Volumes I, II A, II B, III, and IV: Basin Research Institute, Louisiana State University, DOE contract DE-FG07-95ID13366.Google Scholar
Jordan, L., Applin, E., Caldwell, E., et al., 1949, Mesozoic cross sections: Southeastern Geological Society, cross section, 4 p.Google Scholar
Joyce, J. E., Tjalsma, L. R. C., and Prutzman, J. M., 1993, North American glacial meltwater history for the past 2.3 m.y.: Oxygen isotope evidence from the Gulf of Mexico: Geology, 21, 483486.Google Scholar
Juarez, M.A.G., 2001, Evaluacion del Sistema Petrolifero en la Parte Oriental de los Campos Ek-Balam, Chac y Takin, Thesis, Universidad Nacional Autonoma de Mexico, 377 p.Google Scholar
Judice, P. C., and Mazzullo, S. J., 1982, The gray sandstones (Jurassic) in Terryville Field, Louisiana: basinal deposition and exploration model: Gulf Coast Association of Geological Societies Transactions, 32, 2443.Google Scholar
Kaiser, W. R., Ayers, W. B. Jr., and LaBrie, L. W., 1980, Lignite Resources in Texas: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 104, 52 p.Google Scholar
Kane, I. A., and Ponten, A. S. M., 2012, Submarine transitional flow deposits in the Paleogene Gulf of Mexico: Geology, 40, 11191122.Google Scholar
Kennard, J. M., and James, N. P., 1986, Thrombolites and stromatolites: two distinct types of microbial structure: Palaios, 1, 492503.Google Scholar
Kidwell, A. L., 1951, Mesozoic igneous activity in the northern Gulf Coastal Plain: GCAGS Transactions, 1, 182199.Google Scholar
Kiessling, W., Flufel, E., and Golonka, J., 1999, Paleoreef maps: evaluation of a comprehensive database on Phanerozoic reefs: American Association of Petroleum Geologists Bulletin, 83, 15521587.Google Scholar
Kinsland, G. L., and Snedden, J. W., 2016, Comparison of a portion of the K/Pg boundary deposits in two locations: Webb County, Texas, and LaSalle Parish, Louisiana: Gulf Coast Association of Geological Societies Transactions, 66, 789797.Google Scholar
Kinsland, G. L., Shellhouse, K., Muchiri, E., Snedden, J. W., and Virdell, J. W., 2017, Midway Shale: Post-Cretaceous/Paleogene boundary deposition: Gulf Coast Association of Geological Societies 67, 177185.Google Scholar
Kiyokawa, S., Tada, R., Iturralde-Vinent, M., et al., 2002, Cretaceous–Tertiary boundary sequence in the Cacarajicara formation, western Cuba, in Koeberl, C. and MacLeod, K. G., eds., Catastrophic Events and Mass Extinctions: Impacts and Beyond: Geological Society of America, Special Paper 356, 125144.Google Scholar
Klein, G. D., and Chaivre, K. R., 2002, Sequence and seismic stratigraphy of the Bossier Formation (Tithonian), Western East Texas Basin: Gulf Coast Association of Geological Societies Transactions, 52, 551561.Google Scholar
Klemme, H. D., and Ulmishek, G. F., 1991, Effective petroleum source rocks of the world: stratigraphic distribution and controlling depositional factors: American Association of Petroleum Geologists Bulletin, 75, 18091851.Google Scholar
Kneller, E. A., and Johnson, C. A., 2011, Plate kinematics of the Gulf of Mexico based on integrated observations from the Central and South Atlantic: Gulf Coast Association of Geological Societies Transactions, 61, 283299.Google Scholar
Kocurek, G., and Havholm, K. G., 1993, Eolian sequence stratigraphy: a conceptual framework, in Weimer, P. and Posamentier, H., eds., Siliciclastic Sequence Stratigraphy: Recent Developments and Applications: American Association of Petroleum Geologists, Memoir 58, 393409.Google Scholar
Kosters, E., Bebout, D., Seni, S., et al., 1989, Atlas of Major Texas Gas Reservoirs: The University of Texas at Austin, Bureau of Economic Geology, 175 p.Google Scholar
Krafve, A., 1980, Field Study: Black Lake Field – Pettet, Natchitoches Parish, Louisiana, Report on Selected Oil and Gas Fields – North Louisiana and South Arkansas, Shreveport Geological Society, 7 p.Google Scholar
Krapf, C. B. E., Stollhofen, H., Stanistreet, I. G., 2003, Contrasting styles of ephemeral river systems and their interaction with dunes of the Skeleton Coast erg (Namibia): Quaternary International, 104, 4152.Google Scholar
Ladd, J. W., and Sheridan, R. E., 1987, Seismic stratigraphy of the Bahamas: American Association of Petroleum Bulletin, 71, 719736.Google Scholar
Land, L. S., Eustice, R. A., Mack, L. E., and Horita, J., 1995, Reactivity of evaporites during burial: an example from the Jurassic of Alabama: Geochimica et Cosmochimica Acta, 59, 37653778.Google Scholar
Langford, R. P., and Combes, J. M., 1994, Depositional Environments of Unstable Shelf-Margin Deltas of the Oligocene Vicksburg Formation, McAllen Ranch Field, South Texas: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 219, 60 p.Google Scholar
Laubach, S. E., and Jackson, M. L. W., 1990, Origin of arches in northwestern Gulf of Mexico basin: Geology, 18, 595598, doi:10.1130/0091-7613(1990)018<0595:OOAITN>2.3.CO;2.Google Scholar
Lawless, P. N., Fillon, R. H., and Lytton, R. G. III, 1997, Gulf of Mexico Cenozoic biostratigraphic, lithostratigraphic, and sequence stratigraphic event chronology: Gulf Coast Association of Geological Societies Transactions, 47, 275282.Google Scholar
Lawton, T. F., and Molina-Garza, R. S., 2014, U–Pb geochronology of the type Nazas Formation and superjacent strata, northeastern Durango, Mexico: implications of a Jurassic age for continental-arc magmatism in north-central Mexico, Bull: Geological Society of American Bulletin, 126, 11811199.Google Scholar
Lawton, T. F., and Pindell, J., 2017, Upper Triassic–Middle Jurassic strata of Plomosas Uplift and Sierra Samalayuca, Chihuahua, Mexico: onshore record of Syn-rift Gulf of Mexico fault history: American Association of Petroleum Geologists, Search and Discovery 90291, 16.Google Scholar
Lawton, T. F., Vega, F. J., Giles, K. A., and Rosales-Domıngues, C., 2001, Stratigraphy and the origin of the La Popa Basin, Nuevo León and Coahuila, Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, U., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 219240.Google Scholar
Lawton, T. F., Bradford, I. A., Vega, F. J., Gehrels, G. E., and Amato, J. M., 2009, Provenance of Upper Cretaceous–Paleogene sandstones in the foreland basin system of the Sierra Madre Oriental, northeastern Mexico, and its bearing on fluvial dispersal systems of the Mexican Laramide province: Geological Society of America Bulletin, 121, 820836, doi:10.1130/B26450.1.Google Scholar
Lawton, T. F., Pindell, J., Beltran-Triviño, A., et al., 2015, Late Cretaceous–Paleogene foreland sediment-dispersal systems in Northern and Eastern Mexico: interpretations from preliminary detrital-zircon analysis: American Association of Petroleum Geologists, Search and Discovery 30423, 137.Google Scholar
Lawton, T. F., Ruiz Urueña, J. E., Solari, L. A., et al., 2018, Provenance of Upper Triassic–Middle Jurassic strata of the Plomosas uplift, east-central Chihuahua, Mexico, and possible sedimentologic connections with Colorado Plateau depositional systems, in Ingersoll, R. V., Lawton, T. F., and Graham, S. A., eds., Tectonics, Sedimentary Basins, and Provenance: A Celebration of William R. Dickinson’s Career: Geological Society of America, Special Paper 540, 481507.Google Scholar
Le Roy, C., and Rangin, C., 2008, Cenozoic crustal deformation of the offshore Burgos basin region (NE Gulf of Mexico): a new interpretation of deep penetration multichannel seismic reflection lines: Bulletin de la Société Géologique de France, 178, 161174.Google Scholar
Le Roy, C., Rangin, C., Le Pichon, X., et al., 2007, Neogene crustal shear zone along the western Gulf of Mexico margin and its implications for gravity sliding processes: evidences from 2D and 3D multichannel seismic data: Bulletin de la Société Géologique de France, 178(2), 175185.Google Scholar
Leckie, R. M., Bralower, T. J., and Cashman, R., 2002, Oceanic anoxic events and plankton evolution: biotic response to tectonic forcing during the mid-Cretaceous: Paleoceanography, 17, 129, doi:10.1029/2001PA000623.Google Scholar
Lehmann, C., Olseger, D. A., Montanez, I. P., et al., 1999, Evolution of Cupido and Coahuila carbonate platforms, Early Cretaceous, northeastern Mexico: Geological Society of America Bulletin, 111, 7.Google Scholar
Lehmann, C., Osleger, D. A., and Montanez, I., 2000, Sequence stratigraphy of Lower Cretaceous (Barremian–Albian) carbonate platforms of northeastern Mexico: regional and global correlations: Journal of Sedimentary Research, 70, 373391, doi:10.1306/2DC40917-0E47-11D7-8643000102C1865.Google Scholar
Leinfelder, R. R., Schmid, D. U., Nose, M., and Werner, W., 2002, Jurassic reef patterns: expression of a changing globe, in Kiessling, W., Flugel, E., and Golonka, J., eds., Phanerozoic Reef Patterns: Society of Economic Paleontologists and Mineralogists, Special Publication 72, 465520.Google Scholar
Lewis, J., Clinch, S., Meyer, D., et al., 2007, Exploration and Appraisal Challenges in the Gulf of Mexico Deep-Water Wilcox: Part 1 – Exploration Overview, Reservoir Quality, and Seismic Imaging, Society of Economic Paleontologists and Mineralogists Gulf Coast Section Foundation, 398414.Google Scholar
Leyendecker, E. A., 2014, The Gulf of Mexico advantage: American Association of Petroleum Geologists, Search and Discovery 110175, 21 p.Google Scholar
Li, Y., and Ayers, W. B., 2008, Hydrocarbon potential of the deep Travis Peak Formation and underlying strata, western margin of the East Texas Basin: Gulf Coast Association of Geological Societies Transactions, 58, 607621.Google Scholar
Lin, P., Bird, D.E., and Mann, P., 2019, Crustal structure of an extinct, late Jurassic-to-earliest Cretaceous spreading center and its adjacent oceanic crust in the eastern Gulf of Mexico: Marine Geophysical Research, 1–24. DOI: 10.1007/s11001-019-09379-5.Google Scholar
Lisi, A. F., 2013, Provenance of the Upper Jurassic Norphlet and Surrounding Formations from U–Pb Detrital Zircon Geochronology: Master’s thesis, West Virginia University, 148 p.Google Scholar
Liu, C., 2015, Stratigraphy, Depositional History, and Pore Network of the Lower Cretaceous Sunniland Carbonates in the South Florida Basin: Master’s thesis, The University of Texas at Austin, 69 p.Google Scholar
Liu, L., 2014, Rejuvenation of Appalachian topography caused by subsidence-induced differential erosion: Nature Geoscience, 7(7), 518523.Google Scholar
Liu, L., 2015, The ups and downs of North America: evaluating the role of mantle dynamic topography since the Mesozoic, Reviews of Geophysics, 53, 10221049, doi:10.1002/2015RG000489.Google Scholar
Locklin, J. A., 1985, A Rodessa stratigraphic trap (Ingram-Trinity Field), Lower Cretaceous, East Texas Basin: Annual Meeting of the Gulf Coast Association of Geological Societies and Society of Economic Paleontologists and Mineralogists Gulf Coast Section: Lower Cretaceous Depositional Environments From Shoreline to Slope – A Core Workshop, 85–91.Google Scholar
Long, J., 1985, The Eocene Lobo gravity slide, Webb and Zapata counties, Texas, in Contributions to the Geology of South Texas: South Texas Geological Society, 270293.Google Scholar
Longman, M. W., Luneau, B. A., and Landon, S. M., 1998, Nature and distribution of Niobrara lithologies in the Cretaceous Western interior seaway of the Rocky Mountain region: Rocky Mountain Association of Geologists: The Mountain Geologist, 35, 137170.Google Scholar
Longoria, J. F., and Gamper, M. A., 1977, Albian planktonic foraminifera from the Sabinas Basin of northern Mexico: Journal of Foraminiferal Research, 7, 196215, doi:10.2113/gsjfr.7.3.196.Google Scholar
Lopez Ramos, E., 1982, Geología de México: Consejo Nacional de Ciencia y Technología, México, D. F., 2, 454 p.Google Scholar
Lore, G. L., Marin, D. A., Batchelder, E. C., et al., 2001, 2000 Assessment of Conventionally Recoverable Hydrocarbon Resources of the Gulf of Mexico and Atlantic Outer Continental Shelf as of January 1, 1999, Outer Continental Shelf Report: U.S. Department of the Interior Minerals Management Service, 1525.Google Scholar
Loucks, R. G., 1977. Porosity Development and Distribution in Shoal-Water Carbonate Complexes: Subsurface Pearsall Formation (Lower Cretaceous) South Texas. Texas Bureau of Economic Geology.Google Scholar
Loucks, R. G., 1978, Sandstone distribution and potential for geopressured geothermal energy production in the Vicksburg Formation along the Texas Gulf Coast: Gulf Coast Association of Geological Societies Transactions, 28, 239271.Google Scholar
Loucks, R. G., and Crump, J. O., 1985, Vertical facies sequences of the Sunniland and Punta Gorda Formations in the Lower Cretaceous South Florida embayment: Natural Resource Management Corporation No. 31-2 Alico Core, in Bebout, D., and Ratcliff, D., eds., Annual Meeting of the Gulf Coast Association of Geological Societies and Society of Economic Paleontologists and Mineralogists Gulf Coast Section: Lower Cretaceous Depositional Environments From Shoreline to Slope – A Core Workshop, 111–117.Google Scholar
Loucks, R. G., and Longman, M. W., 1987. Lower Cretaceous Ferry Lake Anhydrite, Fairway Field, east Texas: product of shallow-subtidal deposition: Society of Economic Paleontologists and Mineralogists Core Workshop 3: Depositional and Diagenetic Spectra of Evaporites, 130–173.Google Scholar
Loucks, R. G., Dodge, M. M., and Galloway, W. E., 1986, Controls on Porosity and Permeability of Hydrocarbon Reservoirs in Lower Tertiary Sandstones along the Texas Gulf Coast: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 149, 78 p.Google Scholar
Loucks, R. G., Lucia, J. F., and Waite, L. E., 2013, Origin and description of the micropore network within the Lower Cretaceous Stuart City trend tight-gas limestone reservoir in Pawnee Field in South Texas: Gulf Coast Association of Geological Societies Journal, 2, 2941.Google Scholar
Loucks, R. G., Frébourg, G., and Rowe, H. D., 2017a, Upper Cretaceous (Campanian) Ozan and Annona Chalks in Caddo-Pine Island Field, Northwestern Louisiana: depositional setting, lithofacies, and nanopore/micropore network: Gulf Coast Association of Geological Societies Journal, 6, 7391.Google Scholar
Loucks, R. G., Kerans, C., Zeng, H., and Sullivan, P. A., 2017b, Documentation and characterization of the Lower Cretaceous (Valanginian) Calvin and Winn carbonate shelves and shelf margins, onshore north-central Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 101, 119142, doi:10.1306/06281615248.Google Scholar
Lovell, T. R., 2013, Detrital Zircon U–Pb Age Constraints on the Provenance of the Late Jurassic Norphlet Formation, Eastern Gulf of Mexico: Implications for Paleogeography: MS thesis, The University of Alabama, 179 p.Google Scholar
Lovell, T., and Weislogel, A. L., 2010, Detrital zircon U–Pb age constraints on the provenance of the Upper Jurassic Norphlet Formation, eastern Gulf of Mexico: implications for paleogeography: Gulf Coast Association of Geological Societies Transactions, 60, 443460.Google Scholar
Lowery, C. M., Corbett, M. J., Leckie, R. M, et al., 2014, Foraminiferal and nannofossil paleoecology and paleoceanography of the Cenomanian – Turonian Eagle Ford Shale of southern Texas: Paleogeography, Paleoclimatology, Palaeoecology, 413, 4965, doi:10.1016/j.palaeo.2014.07.025.Google Scholar
Lowery, C. M., Cunningham, R., Barrie, C. D., Bralower, T. J., and Snedden, J. W., 2017, The northern Gulf of Mexico during OAE2 and the relationship between water depth and black shale development: Paleoceanography, 32, 13161335, doi:10.1002/2017PA003180.Google Scholar
Lowery, C. M., Bralower, T. J., Owens, J. D., et al. 2018, Rapid recovery of life at ground zero of the end- Cretaceous mass extinction: Nature, 558, 288291, doi:10.1038/s41586-018-0163-6.Google Scholar
Lundquist, J. J., 2000, Foraminiferal Biostratigraphic and Paleoceanographic Analysis of the Eagle Ford, Austin, and Lower Taylor Groups (Middle Cenomanian Through Lower Campanian) of Central Texas: Ph.D. dissertation, The University of Texas at Austin, 545 p.Google Scholar
Lundquist, J. J., 2015, Austin Chalk (UK): stratigraphic, geophysical, hydrogeological, and hydrocarbon exploration/production characteristics: Austin Geological Society, 50th Anniversary Field Symposium, 140 p.Google Scholar
Luneau, B., Doe, M., Leif Colson, J., et al., 2003, Constructing a static model of a fractured reservoir with disparate data sets: Antonio J. Bermudez Complex, Reforma District, Southern Mexico: American Association of Petroleum Geologists, Search and Discovery 90013.Google Scholar
Machel, H. G., 2001, Bacterial and thermochemical sulfate reduction in diagenetic settings, old and new insights: Sedimentary Geology, 140, 143175.Google Scholar
Mackey, G. N., Horton, B. K., and Milliken, K. L., 2012, Provenance of the Paleocene–Eocene Wilcox Group, western Gulf of Mexico basin: evidence for integrated drainage of the southern Laramide Rocky Mountains and Cordilleran arc: Geological Society of America Bulletin, 124, 10071024.Google Scholar
MacRae, G., 1994, Mesozoic Development of the DeSoto Canyon Salt Basin in the Framework of the Early Evolution of the Gulf of Mexico: Ph.D. dissertation, Texas A&M University, 152 p.Google Scholar
Magoon, L. B., Hudson, T. L., and Cook, H. E., 2001, Pimienta-Tamabra(!): a giant supercharged petroleum system in the southern Gulf of Mexico, onshore and offshore Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 83125.Google Scholar
Mancini, E. A., 2010, Jurassic depositional systems, facies, and reservoirs of the northern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 60, 481486.Google Scholar
Mancini, E. A., and Puckett, T. M., 1995, Upper Cretaceous sequence stratigraphy of the Mississippi–Alabama area: Gulf Coast Association of Geological Societies Transactions, 45, 377384.Google Scholar
Mancini, E. A., Mink, R. M., Bearden, B. L., and Wilkerson, R. P., 1985, Norphlet Formation (Upper Jurassic) of Southwestern and Offshore Alabama: environments of deposition and petroleum geology: American Association of Petroleum Geologists Bulletin, 69, 881898.Google Scholar
Mancini, E. A., Aurell, M., Llinas, J. C., et al., 2004, Upper Jurassic thrombolite reservoir play, northeastern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 88, 15731602.Google Scholar
Mancini, E. A., Parcell, W. C., and Ahr, W. M., 2006, Upper Jurassic Smackover thrombolite buildups and associated nearshore facies, southwest Alabama: Gulf Coast Association of Geological Transactions, 56, 551563.Google Scholar
Mander, J., d’Ablaing, J., Wells, K., et al., 2012, 21st Century Atlantis: Incremental Knowledge from a Staged Approach to Development, Illustrated by a Complex, Deepwater Field: Gulf Coast Section of the Society for Sedimentary Geology, 118.Google Scholar
Mandujano-Velazquez, J. J., and Keppie, J. D., 2009. Middle Miocene Chiapas fold and thrust belt of Mexico: a result of collision of the Tehuantepec Transform/Ridge with the Middle America, in Murphy, J. B., Keppie, J. D., and Hynes, A. J. eds., Ancient Orogens and Modern Analogues: London Geological Society, Special Publication 327, 1, 5569.Google Scholar
Mankiewicz, P. J., Pottorf, R. J., Kozar, M. G., and Vrolijk, P., 2009, Gas geochemistry of the Mobile Bay Jurassic Norphlet Formation: thermal controls and implications for reservoir connectivity: American Association of Petroleum Geologists Bulletin, 93, 13191346, doi:10.1306/05220908171.Google Scholar
Manzano, B. K., Fowler, M. G., and Machel, H. G., 1997, The influence of thermochemical sulfate reduction on hydrocarbon composition in Nisku Reservoirs, Brazeau River Area, Alberta, Canada: Organic Geochemistry, 27, 507521.Google Scholar
Marchand, M. E., Apps, G., Li, W., and Rotzien, J. R., 2015, Depositional processes and impact on reservoir quality in deepwater Paleogene reservoirs, US Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 99, 16351648.Google Scholar
Martin, J., Weimer, P., and Bouroullec, R., 2004, Sequence stratigraphy of upper Miocene to upper Pliocene sediments of west-central Mississippi Canyon and northern Atwater Valley, northern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 54, 425441.Google Scholar
Martin, R. G., and Case, J. E., 1975, Geophysical studies in the Gulf of Mexico, in Nairn, A. and Stehli, F., eds., The Gulf of Mexico and Caribbean, Ocean Basins and Margins, 3, Plenum Press, 65106.Google Scholar
Martinez-Medrano, M., Vega-Escobar, R., Flores-Cruz, F., Angeles-Marin, D., and Lopez-Martinez, C., 2009, Integrated seismic and petrographic analysis of the sandstone reservoirs of the Tertiary Veracruz Basin, Mexico, in Bartolini, C. and Roman Ramos, J. R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 217235.Google Scholar
Martinez-Medrano, M., Gonzalez-Mercado, E., and Fernandez-Avendano, E., 2011, Stratigraphy, petrology, and provenance of Neogene sandstones of the Veracruz basin, Mexico: Gulf Coast Association of Geological Societies Transactions, 61, 607620.Google Scholar
Martini, M., and Ortega-Gutiérrez, F., 2016, Tectono-stratigraphic evolution of eastern Mexico during the break-up of Pangea: Earth-Science Reviews, 183, 3855, doi:10.1016/j.earscirev.2016.06.013.Google Scholar
Martini, M., Solari, L., López-Martínez, M., 2014, Correlation of the Arperos Basin from Guanajuato, central Mexico, to Santo Tomás, southern Mexico: Implications for the paleogeography and origin of the Guerrero terrane: Geosphere, 10, 13851401, doi:10.1130/GES01055.1.Google Scholar
Marton, G. L., and Buffler, R. T., 1999, Jurassic–Early Cretaceous tectono-paleogeographic evolution of the southeastern Gulf of Mexico Basin, in Mann, P., ed., Caribbean Basins: Sedimentary Basins of the World: Elsevier Science, 6391.Google Scholar
Marzano, M. S., Pense, G. M., and Andronaco, P., 1988, A comparison of the Jurassic Norphlet Formation in Mary Ann Field, Mobile Bay, Alabama to onshore regional Norphlet trends: Gulf Coast Association of Geological Societies Transactions, 38, 85100.Google Scholar
Maxwell, W. G. H., and Swinchatt, J. P., 1970, Great Barrier Reef: regional variation in a terrigenous-carbonate province: Geological Society of America Bulletin, 81, 691724, doi:10.1130/0016-7606(1970)81[691:GBRRVI]2.0.CO;2.Google Scholar
McArthur, J. M., Howarth, R. J., and Bailey, T. R., 2001, Strontium isotope stratigraphy: LOWESS Version 3 – best fit to the marine Sr-isotope curve for 0–509 Ma and accompanying look-up table for deriving numerical age: The Journal of Geology, 109, 155170.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
McDonald, K. C., 1982, Mid-ocean ridges: fine-scale tectonic, volcanic and hydrothermal processes within the plate tectonic boundary zone: Annual Reviews of the Earth and Planetary Sciences, 10, 155190.Google Scholar
McDonnell, A., Loucks, R. G., and Galloway, W. E., 2008, Paleocene to Eocene deep-water slope canyons, western Gulf of Mexico: further insights for the provenance of deep-water offshore Wilcox Group plays: American Association of Petroleum Geologists Bulletin, 92, 11691189, doi:10.1306/05150808014.Google Scholar
McDonnell, A., Jackson, M., and Hudec, M., 2010, Salt–sediment Interactions during the Jurassic to Miocene of Mississippi Canyon Area, Northern Gulf of Mexico: The University of Texas at Austin, Bureau of Economic Geology, Applied Geodynamics Laboratory, 34 p.Google Scholar
McFarlan, E., 1977, Lower Cretaceous sedimentary facies and sea level changes, U.S. Gulf Coast, in Bebout, D. G., and Loucks, R. G., eds., Cretaceous Carbonates of Texas and Mexico: Applications to Subsurface Exploration: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 89, 512.Google Scholar
McFarlan, E. Jr., and Menes, L. S., 1991, Lower Cretaceous, in Salvador, A., ed., The Gulf of Mexico Basin: The Geology of North America: Geological Society of America, 181204, doi:10.1130/DNAG-GNA-J.181.Google Scholar
McGowen, M. K., and Lopez, C. M., 1983, Depositional Systems in the Nacatoch Formation (Upper Cretaceous), Northeast Texas and Southwest Arkansas: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 137, 63 p.Google Scholar
McKinney, M. L., 1984, Suwannee Channel of the Paleogene coastal plain: support for the “carbonate suppression” model of basin formation, Geology, 12, 343345.Google Scholar
McMillan, M. E., Heller, P. L., and Wing, S. L., 2006, History and causes of post-Laramide relief in the Rocky Mountain orogenic plateau: Geological Society of America Bulletin, 118(3–4), 393405.Google Scholar
Meckel, L. D. III, 2002, Core, log, and seismic characteristics of a high rate amalgamated channel reservoir in a salt-withdrawal minibasin: the upper Miocene “above Magenta” sand, Ursa field, northern Gulf of Mexico, in Weimer, P., Sweet, M., Sullivan, M., et al., eds., Deep-Water Core Workshop, Northern Gulf of Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 6174.Google Scholar
Meckel, L. D. III, and Galloway, W. E., 1996, Formation of high-frequency sequences and their bounding surfaces: case study of the Eocene Yegua Formation, Texas Gulf Coast, USA: Sedimentary Geology, 102, 155186.Google Scholar
Meckel, L. D. III, Ugueto, G. A., Lynch, D. H., et al., 2002, Genetic stratigraphy, stratigraphic architecture, and reservoir stacking patterns of the Upper Miocene–Lower Pliocene Greater Mars–Ursa intraslope basin, Mississippi Canyon, Gulf of Mexico, in Armentrout, J.M. and Rosen, N.C., eds., Sequence Stratigraphic Models for Exploration and Production: Evolving Methodologies, Emerging Models, and Application Histories: Society of Economic Paleontologists and Mineralogists Gulf Coast Section 22nd Annual Bob F. Perkins Research Conference, 113–147.Google Scholar
Meckel, T., 2010, Classifying and characterizing sand-prone mass-transport deposits: American Association of Petroleum Geologists, Search and Discovery 50270.Google Scholar
Melbana Energy, 2017, Petroleum potential of Block 9 PSC: Havana, Cuba, Geosciencias Conference, 1–18.Google Scholar
Meneses-Rocha, J. J., 2001, Tectonic evolution of the Ixtapa graben, an example of a strike–slip basin in southeastern Mexico: implications for regional petroleum systems, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 183216.Google Scholar
Meneses-Scherrer, E. J., Gülen, G., and Tinker, S. W., 2017, Non-geologic factors necessary to develop a shale industry in Mexico: American Association of Petroleum Geologists, Search and Discovery 70000, 1–3.Google Scholar
Meyer, D., Zarra, L., and Yun, J., 2007, From BAHA to Jack, evolution of the Lower Tertiary Wilcox Trend in the deepwater Gulf of Mexico: The Sedimentary Record, 5, 49.Google Scholar
Meyerhoff, A. A., and Hatten, C. W., 1968, Diapiric Structures in Central Cuba, American Association of Petroleum Geologists, Memoir A153, 315357.Google Scholar
Mickus, K., Stern, R. J., Keller, G. R., and Anthony, E. Y., 2009, Potential field evidence for a volcanic rifted margin along the Texas gulf coast: Geology, 37, 387390, doi:10.1130/G25465A.1.Google Scholar
Miller, K. G., Kominz, M. A., Browning, J. V., et al., 2005, The Phanerozoic record of global sea-level change: Science, 310, 12931298, doi:10.1126/science.1116412.Google Scholar
Milliken, J. V., 1988, Late Paleozoic and Early Mesozoic Geologic Evolution of the Arklatex Area: Master’s thesis, Rice University, 286 p.Google Scholar
Milliken, K. T., Blum, M. D., Snedden, J. W., and Galloway, W. E., 2018, Application of fluvial scaling relationships to reconstruct drainage-basin evolution and sediment routing for the Cretaceous and Paleocene of the Gulf of Mexico: Geosphere, 14,119, doi:10.1130/GES01374.1.Google Scholar
Mink, R. M., Bearden, B. L., and Mancini, E. A., 1985, Regional Jurassic Geological Framework of Alabama Coastal Waters Area and Adjacent Federal Waters Area: Geological Survey of Alabama and State Oil and Gas Board, Final Interim Report, 83 p., doi:10.1016/0025-3227(89)90112-6.Google Scholar
Miranda Peralta, I. L. R., Alvarado, A. C., Villalón, R. M., et al., 2014, Play hipotético pre-sal en aguas profundas del Golfo de México: Ingeniería Petrolera, 5, 256266.Google Scholar
Mitchell-Tapping, H. J., 1986, Exploration petrology of the Sunoco Felda trend of South Florida: Gulf Coast Association of Geological Societies Transactions, 36, 241256.Google Scholar
Mitchell-Tapping, H. J., 2002, Exploration analysis of basin maturity in the South Florida Sub-Basin: Gulf Coast Association of Geological Societies Transactions, 52, 753764.Google Scholar
Mitchum, R. M. Jr., and Van Wagoner, J. C., 1991, High-frequency sequences and their stacking patterns: sequence-stratigraphic evidence of high-frequency eustatic cycles: Sedimentary Geology, 70, 131160, doi:10.1016/0037-0738(91)90139-5.Google Scholar
Mitra, S., Figueroa, G. C., Garcia, J. H., and Alvarado, A. M., 2005, Three-dimensional structural model of the Cantarell and Sihil structures, Campeche Bay province, Mexico: American Association of Petroleum Geologists Bulletin, 89, 126.Google Scholar
Mitra, S., Gonzalez, J. D., Garcia, J. H., Hernandez, S., and Banerjee, S., 2006, Structural geometry and evolution of the Ku, Zaap, and Maloob structures, Campeche Bay, Mexico: American Association of Petroleum Geologists Bulletin, 90, 15651584.Google Scholar
Mitra, S., Gonzalez, J., Garcia, J., and Ghosh, K., 2007, Ek-Balam field: a structure related to multiple stages of salt tectonics and extension: American Association of Petroleum Geologists Bulletin, 91, 16191636, doi:10.1306/06260706112.Google Scholar
Mixon, R. B., 1963, Geology of the Huizachal Redbeds, Sierra Madre Oriental, Mexico: Ph.D. dissertation, Louisiana State University, 128 p., https://digitalcommons.lsu.edu/gradschool_disstheses/819Google Scholar
Moffett, J. R., 2015, Discovering the missing piece of the Gulf of Mexico geologic puzzle: American Association of Petroleum Geologists, Search and Discovery 110198, 22 p.Google Scholar
Mohn, K., and Bowen, B., 2012, Florida: the next US frontier – revisiting an old exploration region of the Gulf of Mexico with Modern 3D Data: GeoExpro, 9, 7478.Google Scholar
Moldowan, J. M., Seifert, W. K., and Gallegos, E. J., 1985, Relationship between petroleum composition and depositional environment of petroleum source rocks: American Association of Petroleum Geologists Bulletin, 69, 12551268.Google Scholar
Moldowan, J. M., Dahl, J., Huizinga, B. J., et al., 1994, The molecular fossil record of oleanane and its relation to angiosperms, Science, 265, 768771.Google Scholar
Molina, E., Alegret, L., Arenillas, I., et al., 2006, The global boundary stratotype section and point for the base of the Danian Stage (Paleocene, Paleogene, “Tertiary”, Cenozoic) at El Kef, Tunisia – original definition and revision: Episodes, 29, 263272.Google Scholar
Montgomery, S., 1996, Cotton Valley Lime Pinnacle Reef Play: Branton Field: American Association of Petroleum Geologists Bulletin, 80, 617629.Google Scholar
Moore, C. H., 1984, The Upper Smackover of the Gulf Rim: depositional systems, diagenesis, porosity evolution and hydrocarbon production, the Jurassic of the Gulf Rim: Society of Economic Paleontologists and Mineralogists Gulf Coast Section Foundation 3rd Annual Research Conference, 283–307.Google Scholar
Moore, G. T., Hayashida, D. N., Ross, C. A., and Jacobson, S. R., 1992, Paleoclimate of the Kimmeridgian/Tithonian (Late Jurassic) world: I. Results using a general circulation model: Palaeogeography, Palaeoclimatology, Palaeoecology, 93, 113150, doi:10.1016/0031-0182(92)90186-9.Google Scholar
Moore, M. G., 2010, Exploration, appraisal, and development of turbidite reservoirs in the Western Atwater Foldbelt, Deep Water Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 90104.Google Scholar
Moore, V., and Hinton, D., 2013, Secondary basins and sediment pathways in Green Canyon, Deepwater Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 10499, 4 p.Google Scholar
Moretti, I., Tenreyro, R., Linares, E., et al., 2003, Petroleum system of the Cuban northwest offshore zone, in Bartolini, C., Buffler, R. T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists, Memoir 79, 675696.Google Scholar
Morgan, J. V., Gulick, S., Bralower, T., et al., 2016, The formation of peak rings in large impact craters: Science, 354, 878882, doi:10.1126/science.aah6561.Google Scholar
Morton, C. H., and Weimer, P., 2000, Sequence stratigraphy of the Alaminos Fan (Upper Miocene–Pleistocene), northwestern deep Gulf of Mexico, in Weimer, P., Slatt, R.M., Coleman, J., et al., eds., Deep-Water Reservoirs of the World: Society of Economic Paleontologists and Mineralogists Gulf Coast Section 20th Annual Bob F. Perkins Research Conference, 667–685.Google Scholar
Morton, R. A., 1993, Attributes and origins of ancient submarine slides and filled embayments: examples from the Gulf Coast basin: American Association of Petroleum Geologists Bulletin, 77, 10641081.Google Scholar
Morton, R. A., and Ayers, W. B. Jr., 1992, Plio-Pleistocene Genetic Sequences of the Southwestern Louisiana Continental Shelf and Slope: Geologic Framework, Sedimentary Facies, and Hydrocarbon Distribution: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations, 210, 77 p.Google Scholar
Morton, R. A., and Jirik, L. A., 1989, Structural Cross Sections, Plio-Pleistocene Series, Southeastern Texas Continental Shelf: The University of Texas at Austin, Bureau of Economic Geology.Google Scholar
Morton, R. A., Jirik, L. A., and Foote, R. Q., 1985, Structural Cross Sections, Miocene Series, Texas Continental Shelf: The University of Texas at Austin, Bureau of Economic Geology, 26 p.Google Scholar
Morton, R. A., Jirik, L. A., and Galloway, W. E., 1988, Middle–Upper Miocene Depositional Sequences of the Texas Coastal Plain and Continental Shelf: Geologic Framework, Sedimentary Facies, and Hydrocarbon Plays: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 174, 40 p.Google Scholar
Morton, R. A., Sams, R. H., and Jirik, L. A., 1991, Plio-Pleistocene Depositional Sequences of the Southeastern Texas Continental Shelf and Slope: Geologic Framework, Sedimentary Facies, and Hydrocarbon Distribution: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 200, 80 p.Google Scholar
Moscardelli, L., and Wood, L. J.. 2016. Morphometry of mass-transport deposits as a predictive tool: Geological Society of America Bulletin, 128, 4780.Google Scholar
Moscardelli, L., Wood, L., Torres-Vargas, R., Bermudez, J., and Lopez-Leyva, G., 2008, Processes of Late Tertiary-Age mass transport and associated deposits along the eastern Mexico margin, southern Gulf of Mexico: American Association of Petroleum Geologists, Search and Discovery 90078.Google Scholar
Mullins, H. T., Gardulski, A. F., Wise, S. W. Jr., and Applegate, J., 1983, Middle Miocene oceanographic event in the eastern Gulf of Mexico: implications for seismic stratigraphic succession and loop current/Gulf stream circulation: Geological Society of America Bulletin, 98, 702713.Google Scholar
Mullins, H. T., Gardulski, A. F., and Hine, A. C., 1986, Catastrophic collapse of the west Florida carbonate platform margin: Geology, 14, 167170, doi:10.1130/0091-7613(1986)14<167:CCOTWF>2.0.CO;2.Google Scholar
Murillo-Muneton, G., Grajales-Nishimura, J., Cedillo-Pardon, E., et al., 2002, Stratigraphic architecture and sedimentology of the main oil-producing interval at the Cantarell Oil Field: the K-T boundary sedimentary succession: Society of Petroleum Engineers International Petroleum Conference and Exhibition, Villahermosa, Mexico, 7 p., doi:10.2118/74431-MS.Google Scholar
Murray, G., 1961, Geology of the Atlantic and Gulf Coastal Province of North America: Harper and Brothers, 692 p.Google Scholar
Myczynski, R., Federico, O., and Villaseñor, A. B., 1998, Revised biostratigraphy and correlations of the Middle–Upper Oxfordian in the Americas (southern USA, Mexico, Cuba, and northern Chile): Neues Jahrbuch für Geologie und Paläontologie, 207, 185206.Google Scholar
Naehr, T. H., MacDonald, I. R., Bohrmann, G., and Briones, E. E., 2007, Biogeochemistry of hydrocarbon seeps on the Campeche Escarpment, southern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 57, 599604.Google Scholar
Nelson, T. H., 1991, Salt tectonics and listric-normal faulting, in Salvador, A., ed., The Gulf of Mexico Basin: The Geology of North America: Geological Society of America, 7389, doi:10.1130/DNAG-GNA-J.73.Google Scholar
Nguyen, L. C., and Mann, P.,, 2016, Gravity and magnetic constraints on the Jurassic opening of the oceanic Gulf of Mexico and the location and tectonic history of the Western Main transform fault along the eastern continental margin of Mexico: Interpretation, 4, SC23SC33, doi:10.1190/INT-2015-0110.1.Google Scholar
Nicholas, R. L., and Waddell, D. E., 1989, The Ouachita system in the subsurface of Texas, Arkansas, and Louisiana: The Geological Society of America 1989, 112.Google Scholar
Nieto, J. O., 2010, Analisis estratigrafico de la secuencia sedimentaria en el grupo Chicontepec: unpublished dissertation, Universidad Nacional Autonoma de Mexico, 200 p.Google Scholar
Norton, I. O., Carruthers, D. T., and Hudec, M. R., 2015, Rift to drift transition in the south Atlantic salt basins: a new flavor of oceanic crust: Geology, 44, 5558, doi:10.1130/G37265.1.Google Scholar
Norton, I. O., Lawver, L. A., and Snedden, J. W., 2016, Gulf of Mexico tectonic evolution from Mexico deformation to oceanic crust, in Lowery, C. M. and Snedden, J. W., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Research Conference, 1–12.Google Scholar
Norton, I. O., Lawver, L., Snedden, J., 2018, Rift to drift transition in the Gulf of Mexico: AGU Fall Meeting, https://agu.confex.com/agu/fm18/meetingapp.cgi/Paper/358043.Google Scholar
Núñez-Useche, F., Barragán, R., Moreno-Bedmar, J. A., and Canet, C., 2014, Mexican archives for the major Cretaceous oceanic anoxic events: Boletín de la Sociedad Geológica Mexicana, 66, 491505.Google Scholar
Núñez-Useche, F., Canet, C., Barragan, R., and Alfonso, P., 2016, Bioevents and redox conditions around the Cenomanian–Turonian anoxic event in Central Mexico: Palaeogeography, Palaeoclimatology, Palaeoecology, 449, 205226.Google Scholar
Nyberg, B., Helland-Hansen, W., Gawthorpe, R. L., et al., 2018, Revisiting morphological relationships of modern source-to-sink segments as a first-order approach to scale ancient sedimentary systems: Sedimentary Geology, 373, 111133.Google Scholar
Nyberg, J., and Howell, J. A., 2016, Global distribution of modern shallow marine shorelines: implications for exploration and reservoir analogue studies: Marine and Petroleum Geology, 71, 83104.Google Scholar
Oehler, J. H., 1984, Carbonate source rocks in the Jurassic Smackover Trend of Mississippi, Alabama and Florida, in Palacas, J. G., ed., Petroleum Geochemistry and Source Rock Potential of Carbonate Rocks: American Association of Petroleum Geologists, Studies in Geology 18, 6369.Google Scholar
Officer, C. B., and Drake, C. L., 1985, Terminal Cretaceous environmental events. Science, 227, 11611167.Google Scholar
Ogg, J. G., Ogg, G. M., and Gradstein, F. M., 2016, A Concise Geologic Time Scale: Elsevier, 230 p.Google Scholar
Ogiesoba, O. C., Ambrose, W. A., and Loucks, R. G., 2018, Application of instantaneous-frequency attribute and gamma-ray wireline logs in the delineation of lithology in Serbin field, Southeast Texas: a case study: Interpretation, 6, T1023T1043, doi:10.1190/int-2018-0041.1.Google Scholar
Oivanki, S. M., 1974, Paleodepositional environments in the Upper Jurassic Zuloaga Formation (Smackover), northeastern Mexico: Gulf Coast Association of Geological Societies Transactions, 34, 258278.Google Scholar
Olariu, M. I., Hammes, U., Ambrose, W. A., 2013, Depositional architecture of growth-fault related wave-dominated shelf edge deltas of the Oligocene Frio Formation in Corpus Christi Bay: Marine and Petroleum Geology, 48, 428440.Google Scholar
Oloriz, F., Villasenor, A. B., Gonzalez-Arreola, C., 2003, Major lithostratigraphic units in land-outcrops of north-central Mexico and the subsurface along the northern rim of Gulf of Mexico Basin (Upper Jurassic–lowermost Cretaceous): a proposal for correlation of tectono-eustatic sequences: Journal of South American Earth Sciences, 16, 119142, doi:10.1016/S0895-9811(03)00049-X.Google Scholar
Olson, H. C., Snedden, J. W., and Cunningham, R., 2015, Development and application of a robust chronostratigraphic framework in Gulf of Mexico Mesozoic exploration: Interpretation, 3, SN39SN58, doi:10.1190/INT-2014-0179.1.Google Scholar
Olsen, P. E., 1997, Stratigraphic record of the Early Mesozoic breakup of Pangea in the Laurasia–Gondwana Rift System: Annual Review of Earth and Planetary Sciences, 25, 337401, doi:10.1146/annurev.earth.25.1.337.Google Scholar
Ortega-Flores, B., Solari, L., Lawton, T. F., and Ortega-Obregón, C., 2014, Detrital-zircon record of major Middle Triassic–Early Cretaceous provenance shift, central Mexico: demise of Gondwanan continental fluvial systems and onset of back-arc volcanism and sedimentation: International Geology Review, 56, 237261, doi:10.1080/00206814.2013.844313.Google Scholar
Ortuno-Arzate, S., Ferket, H., Cacas, M.-C., Swennen, R., and Roure, F., 2003, Late Cretaceous carbonate reservoirs in the Cordoba Platform and Veracruz Basin, eastern Mexico, in Bartolini, C., Buffler, R. T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists, Memoir 79, 476514.Google Scholar
Oxley, M. L., and Minihan, E. D., 1969, Alabama exploration underway. Pt. 1: Oil and Gas Journal, 67, 207212.Google Scholar
Padilla y Sánchez, R. J., 2007, Evolucion Geologica del Sureste Mexicano desde el Mesozoico al presente en el contexto regional del Golfo de Mexico: Boletín de la Sociedad Geologica Mexicana, 59, 1942.Google Scholar
Padilla y Sánchez, R. J., 2014, Tectonics of Eastern Mexico: Gulf of Mexico and its hydrocarbon potential, American Association of Petroleum Geologists, Search and Discovery 10622, 54 p.Google Scholar
Padilla y Sánchez, R. J., and Jose, R., 2016, Late Triassic–Late Cretaceous paleography of Mexico and the Gulf of Mexico, in Lowery, C. M., and Snedden, J. W., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists, Gulf Coast Section, 35th Annual Research Conference, 1–30.Google Scholar
Padilla y Sánchez, R. J., Dominguez Trejo, I., Lopez Azcarraga, A. G., et al., 2013, Tectonic map of Mexico: Division de Ingenieria en Ciencias de la Tierra Facultad de Ingenieria Universidad Nacional Autonoma de Mexico [map].Google Scholar
Paine, W. R., 1971, Petrology and sedimentation of the Hackberry sequence of southwest Louisiana: Gulf Coast Association of Geological Societies Transactions, 21, 3755.Google Scholar
Paleo-Data Inc., 2017, PDI Neogene biostrat chart of the Gulf Basin US.Google Scholar
Pardo, G., 1975, Geology of Cuba, in Nairn, A. E. M. and Stehli, F. G., eds., The Ocean Basins and Margins: The Gulf of Mexico and the Caribbean: Plenum, 3, 553615.Google Scholar
Pardo, G., 2009, The Geology of Cuba: American Association of Petroleum Geologists, Studies in Geology 58, 311341.Google Scholar
Paredes, H. C. G., Medrano, M. M., and Sessarego, H. L., 2009, Provenance for the Middle and Upper Miocene sandstones of the Veracruz Basin, Mexico, in Bartolini, C. and Roman Ramos, J.R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 397407.Google Scholar
Parra, P. A., Rubio, N., Ramirez, C., et al. 2013, Unconventional reservoir development in Mexico: lessons learned from the first exploratory wells: Society of Petroleum Engineers Unconventional Resources Conference, 1–14, doi:10.2118/164545-MS.Google Scholar
Parrish, J. T., and Peterson, F., 1988, Wind directions predicted from global circulation models and wind directions determined from eolian sandstones of the western United States – a comparison: Sedimentary Geology, 56, 261282.Google Scholar
Pashin, J. C., Guohai, J., and Hills, D. J., 2016, Mesozoic structure and petroleum systems in the DeSoto Canyon salt basin in the Mobile, Pensacola, Destine Dome, and Viosca Knoll Areas of the MAFLA Shelf: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Research Conference, 416–449.Google Scholar
Passey, Q. R., Creany, S., Kulla, J. B., Moretti, F. J., and Stroud, J. D., 1990, A practical model for organic richness from porosity and resistivity logs: American Association of Petroleum Geologists Bulletin, 74, 17771794.Google Scholar
Passey, Q. R., Bohacs, K., Esc, W. L., Klimentidis, R., and Sinha, S., 2010, From oil-prone source rock to gas-producing shale reservoir-geologic and petrophysical characterization of unconventional shale gas reservoirs: International Oil and Gas Conference and Exhibition in China: Society of Petroleum Engineers, 1–29.Google Scholar
Paxton, S. T., 2017a, Assessment of undiscovered oil and gas resources in the Haynesville Formation, U.S. Gulf Coast, 2016: United States Geological Survey: Fact Sheet 2017-3016, 1–2, doi:10.3133/fs20173015.Google Scholar
Paxton, S. T., 2017b, Assessment of undiscovered oil and gas resources in the Bossier Formation, U.S. Gulf Coast, 2016: United States Geological Survey, Fact Sheet 2017-3015, 1–2, doi:10.3133/fs20173015.Google Scholar
Pearson, K., Dubiel, R. F., Person, O. N., and Pitman, J. K., 2011, Assessment of undiscovered oil and gas resources of the Upper Cretaceous Austin Chalk and Tokio and Eutaw Formations, Gulf Coast, 2010: U.S. Geological Survey National Assessment of Oil and Gas Fact Sheet 3046, 1–2.Google Scholar
Peel, F. J., Hossack, J. R., and Travis, C. J., 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: American Association of Petroleum Geologists, Memoir 65, 153175.Google Scholar
Pemex Exploración y Producción, 2013, Provincia Petrolera Golfo de Mexico Profundo: Pemex, 1–26.Google Scholar
Penfield, G. T., and Camargo-Zanoguera, A., 1981, Definition of a major igneous zone in the central Yucatan platform with aeromagnetics and gravity, in Technical Program, Abstracts and Bibliographies: Society of Exploration Geophysicists: 51st Annual Meeting (Abs.), 37.Google Scholar
Pepper, F., 1982, Depositional environments of the Norphlet Formation (Jurassic) for Southwestern Alabama: Gulf Coast Association of Geological Societies Transactions, 32, 1722.Google Scholar
Perez, L. E., 2017, Neogene Current-Modified Submarine Fans and Associated Bed Forms in Mexican Deep-Water Areas: MS thesis, University of Texas at Austin, 91 p.Google Scholar
Perfit, M. R., and Chadwick, W. W. Jr., 1998, Magmatism at mid-ocean ridges: constraints from volcanological and geochemical investigations, in Buck, W. R., Delaney, P. T., Karson, J. A., and Lagabrielle, Y., eds., Faulting and Magmatism at Mid-Ocean Ridges: American Geophysical Union, 59115, doi:10.1029/GM106p0059.Google Scholar
Perkins, B. F., 1985, Caprinid reefs and related facies in the Comanche Cretaceous Glen Rose Limestone of Central Texas: Annual Meeting of the Gulf Coast Association of Geological Societies and Society of Economic Paleontologists and Mineralogists Gulf Coast Section: Lower Cretaceous Depositional Environments From Shoreline to Slope – A Core Workshop, 129–140.Google Scholar
Peters, K. E., Walters, C. C., and Moldowan, J. M., 2005, The Biomarker Guide, Volume 2: Biomarkers and Isotopes in Petroleum Exploration and Earth History, Cambridge University Press, 300 p.Google Scholar
Pettigrew, R., Priddy, C., Elson, A., et al., 2017, Fluvial–aeolian-evaporitic interactions in arid continental basins: implications for basin-scale migration and reservoir characterisation [poster]: American Association of Petroleum Geologists, Search and Discovery 70291.Google Scholar
Petty, A. J., 1995, Ferry Lake, Rodessa, and Punta Gorda anhydrite bed correlation, Lower Cretaceous, offshore eastern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 45, 487493.Google Scholar
Petty, A. J., 1999, Petroleum exploration and stratigraphy of the Lower Cretaceous James Limestone (Aptian) and Andrew Formation (Albian): Main Pass, Viosca Knoll, and Mobile area, northeastern Gulf of Mexico: Gulf Coast Association of Geological Societies, 49, 440450, doi:10.1306/E4FD3F45-1732-11D7-8645000102C1865D.Google Scholar
Petty, A. J., 2008, Stratigraphy and petroleum exploration history of the Cotton Valley Group (Lower Cretaceous to Upper Jurassic) and Haynesville Group (Upper Jurassic), offshore northeastern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 58, 713728.Google Scholar
Phelps, R. M., 2011, Middle-Hauterivian to Lower-Campanian Sequence Stratigraphy and Stable Isotope Geochemistry of the Comanche Platform, South Texas: Ph.D. dissertation, The University of Texas at Austin, 227 p.Google Scholar
Phelps, R. M., Kerans, C., Loucks, R. G., et al., 2014, Oceanographic and eustatic control of carbonate platform evolution and sequence stratigraphy on the Cretaceous (Valanginian–Campanian) passive margin, northern Gulf of Mexico: Sedimentology, 61, 461496, doi:10.1111/sed.12062.Google Scholar
Phelps, R. M., Kerans, C., Da-Gama, R., et al., 2015, Response and recovery of the Comanche carbonate platform surrounding multiple Cretaceous oceanic anoxic events, northern Gulf of Mexico: Cretaceous Research, 54, 117144, doi:10.1016/j.cretres.2014.09.002.Google Scholar
Phillips, S., 1987, Shelf Sedimentation and Depositional Sequence Stratigraphy of the Upper Cretaceous Woodbine-Eagle Ford Groups, East Texas: Ph.D. dissertation, Cornell University, 507 p.Google Scholar
Pierce, J. D., 2014, U–Pb Geochronology of the Late Cretaceous Eagle Ford Shale: Defining Chronostratigraphic Boundaries and Volcanic Ash Source: MS thesis, The University of Texas at Austin, 144 p.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 implication: American Association of Petroleum Geologists Bulletin, 95, 219240, doi:10.1306/06301010004.Google Scholar
Pilcher, R. S., Murphy, R. T., and Ciosek, J. M., 2014, Jurassic raft tectonics in the northeastern Gulf of Mexico: Interpretation, 2, 3955, doi:10.1190/INT-2014-0058.1.Google Scholar
Pindell, J., and Kennan, L., 2001, Kinematic evolution of the Gulf of Mexico and Caribbean, in Fillon, R. H., Rosen, N. C., Weimer, P., et al., eds., Petroleum Systems of Deep-Water Basins: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 21st Annual Research Conference, 193–220.Google Scholar
Pindell, J., Miranda, E. C., Ceron, A., and Hernandez, L., 2016, Aeromagnetic map constrains Jurassic–Early Cretaceous syn-rift, break up, and rotational seafloor spreading history in the Gulf of Mexico, in Lowry, C. D. and Snedden, J. W., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists, Gulf Coast Section, 35th Annual Research Conference, 1–24.Google Scholar
Pitman, J. K., 2014, Reservoirs and Petroleum Systems of the Gulf Coast: U. S. Geological Survey, 4 p.Google Scholar
Pittman, J. G., 1989, Stratigraphy of the Glen Rose Formation, western Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 39, 247264.Google Scholar
Pogge von Strandmann, P. A. E., Jenykyns, H. C., and Woodfine, R., 2013, Lithium isotope evidence for 13C enhanced weathering during Oceanic Anoxic Event 2: Nature Geoscience, 6, 668672.Google Scholar
Pollastro, R. M., 2001, 1995 USGS National Oil and Gas Play-Based Assessment of the South Florida Basin, Florida Peninsula Province: National Assessment of Oil and Gas Project: Petroleum Systems and Assessment of South Florida Basin, U.S. Geological Survey, Digital Data Series 69-A, 117.Google Scholar
Pollastro, R. M., Schenk, C. J., and Charpentier, R. R., 2001, Assessment of Undiscovered Oil and Gas in the Onshore and State Waters Portion of the South Florida Basin, Florida – USGS Province 50: National Assessment of Oil and Gas Project: Petroleum Systems and Assessment of South Florida Basin, U.S. Geological Survey, Digital Data Series 69-A, 117.Google Scholar
Popenoe, P., Henry, V. J., and Idris, F. M. 1987, Gulf trough: the Atlantic connection: Geology, 15, 327332.Google Scholar
Porres Luna, A. A., 2018, Plenaria 5 “Desarrollo de Yacimientos No Convencionales y en Aguas Profundas” Situacion Actual de los Yacimientos No Convencionales y en Aguas Profundas en Mexico y Tecnologias necesarias para desarrollarlos, Comision Nacional de Hidrocarburos, 31.Google Scholar
Potter-McIntyre, S. L., Breeden, J. R., and Malone, D. H., 2018, A Maastrichtian birth of the ancestral Mississippi River system: evidence from the U–Pb detrital zircon geochronology of the McNairy Sandstone, Illinois, USA: Cretaceous Research: 91, 7179, doi:10.1016/j.cretres.2018.05.010.Google Scholar
Power, B., Covault, J., Fildani, A., et al., 2013, Facies analysis and interpretation of argillaceous sandstone beds in the Paleogene Wilcox Formation, deepwater Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 63, 575578.Google Scholar
Prather, B. E., 1992, Evolution of a late Jurassic carbonate/evaporate platform, Conecuh Embayment, Northeastern Gulf Coast, USA: American Association of Petroleum Geologists Bulletin, 76, 164190.Google Scholar
Prather, B. E., 2000, Calibration and visualization of depositional process models for above-ground slopes: a case study from the Gulf of Mexico: Marine and Petroleum Geology, 17, 619638.Google Scholar
Prather, B. E., Booth, J. R., Steffens, G. S., and Craig, P. A., 1998, Classification, lithologic calibration, and stratigraphic succession of the seismic facies of intraslope basins, deep-water Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 82, 701728.Google Scholar
Prather, B. E., O’Byrne, C., Pirmez, C., and Sylvester, Z., 2017, Sediment partitioning, continental slopes and base-of-slope systems: Basin Research, 29: 394416. doi: 10.1111/bre.12190.Google Scholar
Presley, M. W., and Reed, C. H., 1984, Jurassic exploration trends of East Texas, in Presley, M. W., ed., The Jurassic of East Texas, East Texas Geological Society, 1122.Google Scholar
Prost, G., and Aranda-Garcia, M., 2001, Tectonics and hydrocarbon systems of the Veracruz Basin, Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 271291.Google Scholar
Puga-Bernabéu, A., Webster, J. M., Beaman, R. J., and Guilbaud, V., 2013, Variation in canyon morphology on the Great Barrier Reef margin, northeastern Australia: The influence of slope and barrier reefs: Geomorphology, 191, 3550, doi:10.1016/j.geomorph.2013.03.001.Google Scholar
Pulham, A. J. 1993, Variations in slope deposition, Pliocene–Pleistocene, offshore Louisiana, northeast Gulf of Mexico, in Weimer, P. and Posamentier, H. W., eds., Siliciclastic Sequence Stratigraphy: Recent Developments and Applications: American Association of Petroleum Geologists, Memoir 58, 199234.Google Scholar
Purkey Phillips, M., in prep., The first regional biostratigraphic characterization of the Paleocene/Eocene Thermal Maximum (PETM) in the Gulf of Mexico.Google Scholar
Radovich, B. J., Moon, J., Connors, C. D., and Bird, D., 2007, Insights into structure and stratigraphy of the northern Gulf of Mexico from 2D pre-stack depth migration imaging of mega-regional onshore to deep water, long-offset seismic data: Gulf Coast Association of Geological Societies Transactions, 57, 633637.Google Scholar
Radovich, B., Horn, B., Nuttall, P., and McGrail, A., 2011, The only complete regional perspective: GEO ExPro, 8(2), 3638.Google Scholar
Rahl, J. M., Reiners, P. W., Campbell, I. H., Nicolescu, S., and Allen, C. M., 2003. Combined single-grain (U–Th)/He and U/Pb dating of detrital zircons from the Navajo Sandstone, Utah: Geology, 31, 761764, doi:10.1130/G19653.1.Google Scholar
Rainwater, E. H., 1964, Regional stratigraphy of the Gulf Coast Miocene: Gulf Coast Association of Geological Societies Transactions, 14, 81124.Google Scholar
Ramos, A., and Galloway, W. E., 1990, Facies and sand-body geometry of the Queen City (Eocene) tide-dominated delta-margin embayment, NW Gulf of Mexico Basin: Sedimentology, 37, 10791098.Google Scholar
Randazzo, A. F., 1997, The sedimentary platform of Florida: Mesozoic to Cenozoic, in Randazzo, A. F. and Jones, D. S., eds., The Geology of Florida, University Press of Florida, 3956.Google Scholar
Raymond, D. E., Osborne, W. E., Copeland, C. W., and Neathery, T. L., 1988, Alabama Stratigraphy. Geological Survey of Alabama, Circular, 140, 1102.Google Scholar
Reading, H. G., and Richards, M., 1994, Turbidite systems in deep-water basin margins classified by grain size and feeder system: American Association of Petroleum Geologists Bulletin, 78, 792822.Google Scholar
Reed, J. C., Leyendecker, C. L., Khan, A. S., et al., 1987, Correlation of Cenozoic Sediments, Gulf of Mexico Outer Continental Shelf. Part I: Galveston Area, Offshore Texas Through Vermilion Area, Offshore Louisiana: Gulf of Mexico OCS Regional Office, Report MMS 87-0026.Google Scholar
Reese, D., 1976, Pre-ferry lake lower Cretaceous deltas of south Mississippi and producing trends: Gulf Coast Association of Geological Societies Transactions, 26, 5963.Google Scholar
Reiners, P. W., Campbell, I. H., Nicolescu, S., et al., 2005, (U–Th)/(He–Pb) double dating of detrital zircons: American Journal of Science, 305, 259311.Google Scholar
Renne, P. R., Deino, A. L., Hilgen, F. J., et al., 2013, Time scales of critical events around the Cretaceous–Paleogene boundary: Science, 339, 684687.Google Scholar
Reynolds, T., 2000, Reservoir architecture in the Mars field, deepwater Gulf of Mexico, USA: the implications of production, seismic, core and well-log data: Society of Economic Paleontologists and Mineralogists Gulf Coast Section 20th Annual Research Conference, 877–892.Google Scholar
Ricoy-Paramo, V., 2005, 3D Seismic Characterisation of the Cantarell Field, Campeche Basin, Mexico: Master’s thesis, Cardiff University, 418 p.Google Scholar
Riggs, N. R., Lehman, T. M., Gehrels, G. E., and Dickinson, W. R., 1996, Detrital zircon link between headwaters and terminus of the Upper Triassic Chinle–Dockum paleoriver system: Science, 273, 97100.Google Scholar
Riller, U., Poelchau, M. H., Rae, A. S. P., et al., 2018, Rock fluidization during peak-ring formation of large impact structures: Nature, 11, 511518, doi:10.1038/s41586–018-0607-z.Google Scholar
Rine, J. M., 2014, Reconstruction of diagenetic and burial history of South Georgia Rift Basin: analysis of sandstones from the Rizer #1 South Carolina: Houston Geological Society Bulletin, 57, 1114.Google Scholar
Rine, J. M., Hollon, B. E., Fu, R., Houghton, N., and Waddell, M., 2014, Diagenetic and burial history of a portion of the Late Triassic South Georgia Rift Basin based on petrologic and isotopic (d18O) analyses of sandstones from test borehole rizer #1, Colleton County, SC: American Association of Petroleum Geologists, Search and Discovery 51016, 31 p.Google Scholar
Rios Lopez, J. J., 1996, Disposicion de Los Cuerpos de Areniscas y su Relacion en el Mantemiento de La Produccion del Campo Ek-Balam, Campeche: Asociacion Mexicana de Geologos Petroleros, 65, 138.Google Scholar
Rios Lopez, J. J., and Cantú-Chapa, A., 2009, Stratigraphy and sedimentology of Middle Eocene Kumaza calcarenites “member” in the Ku, Maloob, and Zaap oil fields, offshore Campeche, Mexico, in Bartolini, C. and Roman Ramos, J.R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 257277.Google Scholar
Rivas, L. F., Sanclemente, J., and Rickets, W. K., 2009, Tahiti subsurface: drilling and completion technology challenges and accomplishments: Offshore Technology Conference, 1–15. doi:10.4043/19861-MS.Google Scholar
Roca-Ramisa, L., and Arnabar, D., 1994, Geological and geomechanical reservoir characterization: The Ek-Balam Field Study in Mexico: Congresso Brasileiro de Petroleo, 1–28.Google Scholar
Rodriguez Hernandez, M. L. A., Oritz Gomez, P. L., Sanchez, C. B., et al., 2003, Distribution of depositional facies and reservoir properties from middle Cretaceous carbonates of the Cordoba Platform: American Association of Petroleum Geologists, Search and Discovery 90020, 1 p.Google Scholar
Roesink, J. G., Weimer, P., and Bouroullec, R., 2004, Sequence stratigraphy of Miocene to Pleistocene sediments of east-central Mississippi canyon, northern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 54, 587601.Google Scholar
Rogers, R., 1987, A palynological age determination for the Dorcheat and Hosston Formations: the Jurassic–Cretaceous boundary in northern Louisiana: Gulf Coast Association of Geological Societies Transactions, 37, 447456.Google Scholar
Rosenfeld, J. and Pindell, J., 2003, Early Paleogene isolation of the Gulf of Mexico from the world’s oceans? Implications for hydrocarbon exploration and eustasy, in Bartolini, C., Buffler, R. T., and Blickwed, J.F., eds., The Circum-Gulf of Mexico and Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists, Memoir 79, 89103.Google Scholar
Roure, F., Alzaga-Ruiz, H., Callot, J.-P., et al., 2009, Long lasting interactions between tectonic loading, unroofing, post-rift thermal subsidence and sedimentary transfers along the western margin of the Gulf of Mexico: some insights from integrated quantitative studies: Tectonophysics, 475, 119189.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: American Association of Petroleum Geologists, Memoir 65, 199228.Google Scholar
Rowan, M. G., 2018, The South Atlantic and Gulf of Mexico salt basins: crustal thinning, subsidence and accommodation for salt and presalt strata, in McClay, K. R. and Hammerstein, J. A., eds., Passive Margins: Tectonics, Sedimentation and Magmatism, London, Geological Society, Special Publication 476, 144. doi:10.1144/SP476.6.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. and Weimer, P., 1998, Salt–sediment interaction, Northern Green Canyon and Ewing Bank (Offshore Louisiana), Northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 82, 10551082.Google Scholar
Rowan, M. G., Trudgill, B. D., and Fiduk, J. C., 2000, Deepwater, salt-cored fold belts: lessons from the Mississippi Fan and Perdido fold belts, northern Gulf of Mexico, in Mohriak, W. and Talwani, M., eds., Atlantic Rifts and Continental Margins: American Geophysical Union, Geophysical Monograph, 115, 173191, doi:10.1029/GM115p0173.Google Scholar
Rowan, M. G., Peel, F. J., and Venderville, B. C., 2004, Gravity-driven fold belts on passive margins, in McClay, K. R., ed., Thrust Tectonics and Hydrocarbon Systems: American Association of Petroleum Geologists, Memoir 82, 152182.Google Scholar
Rowan, M. G., Giles, K. A., Hearon, T. E. IV, and Fiduk, J. C., 2016, Megaflaps adjacent to salt diapirs: American Association of Petroleum Geologists Bulletin, 100, 17231747, doi:10.1306/05241616009.Google Scholar
Rozendal, R. A., and Erskine, W. S., 1971, Deep test in Ouachita Structural Belt of Central Texas, American Association of Petroleum Geologists Bulletin, 55, 20082017.Google Scholar
Ryan, P. C. and Reynolds, R. C. Jr., 1997, The chemical composition of serpentine/chlorite in the Tuscaloosa Formation, United States Gulf Coast: EDX vs. XRD determinations, implications for mineralogic reactions and the origin of anatase: Clays and Clay Minerals 45(3), 339352.Google Scholar
Rynott, T., 2015, Gulf of Mexico Inboard Lower Tertiary and Cretaceous: plays and potential (abs.): American Association of Petroleum Geologists, Search and Discovery 90219.Google Scholar
Salomon-Mora, L. E., Aranda-Garcia, M., and Roman Ramos, J. R., 2009, Contractional growth faulting in the Mexican Ridges, Gulf of Mexico, in Bartolini, C. and Roman Ramos, J.R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 93116.Google Scholar
Salter, R., Beller, M., Shelander, D., et al., 2005, Increasing accuracy and resolution in porosity prediction for carbonate reservoirs using high-resolution seismic technology in the Arenqu-Lobina Area, Offshore Mexico: American Association of Petroleum Geologists, Search and Discovery 90039.Google Scholar
Salvador, A., 1987, Late Triassic–Jurassic paleogeography and origin of the Gulf of Mexico Basin: American Association of Petroleum Geologists Bulletin, 71, 419451.Google Scholar
Salvador, A., ed., 1991a, The Gulf of Mexico basin, in The Geology of North America, V. Geological Society of America, 568 p.Google Scholar
Salvador, A., 1991b, Origin and development of the Gulf of Mexico basin, in Salvador, A., ed., The Geology of North America V: The Gulf of Mexico Basin: Geological Society of America, 389344.Google Scholar
Salvador, A. and Quezada-Muneton, J. M., 1989, Stratigraphic correlation chart, the Gulf of Mexico Basin, in Salvador, A., ed., The Geology of North America, V: The Gulf of Mexico Basin, Geological Society of America, plate 5.Google Scholar
Sánchez-Hernández, H., 2013, Stratigraphic Characterization and Evolution of a Mid-Tertiary Age Deep Water System, Holok area, SW Gulf of Mexico: Ph.D. dissertation, The University of Aberdeen, 358 p.Google Scholar
Sanchez-Montes de Oca, R., 1980, Geología petrolera de la Sierra de Chiapas: Boletin de la Asociacion Mexicana de Geologos Petroleros, 31, 6797.Google Scholar
Sandrea, I., Sandrea, R., Limon, M., Vazquez, K., Horbury, A., and Shann, M., 2018, Mexico History of Oil Exploration, its Amazing Carbonates, and Untapped Oil Potential: Pennwell, 144p.Google Scholar
Sandwell, D. T., Müller, R. D., Smith, W. H. F., Garcia, E., and Francis, R., 2014, New global marine gravity model from CryoSat-2 and Jason-1 reveals buried tectonic structure: Science, 346, 6567, doi:10.1126/science.1258213.Google Scholar
Sanford, J. C., Snedden, J. W., and Gulick, S. P. S., 2016, The Cretaceous–Paleogene boundary deposit in the Gulf of Mexico: large-scale oceanic basin response to the Chicxulub impact: Journal of Geophysical Research: Solid Earth, 121, 22 p., doi:10.1002/2015JB012615.Google Scholar
Santamaria Orozco, D. M., 2000, Organic Geochemistry of Tithonian Source Rocks and Associated Oils from the Sonda De Campeche, Mexico: Ph.D. dissertation, Rhine-Westphalia Institute of Technology, 201 p.Google Scholar
Sassen, R., 1988, Geochemical and carbon isotopic studies of crude oil destruction, bitumen precipitation, and sulfate reduction in the Deep Smackover Formation: Organic Geochemistry, 12, 351361.Google Scholar
Sassen, R., 1990, Geochemistry of carbonate source rocks and crude oils in Jurassic salt basins of the Gulf Coast, in Brooks, J., ed., Classic Petroleum Provinces, London Geological Society, Special Publications 50, 265277.Google Scholar
Sassen, R., Moore, C. H., and Meendsen, F. C., 1987, Distribution of hydrocarbon source potential in the Jurassic Smackover Formation: Organic Geochemistry, 11.5, 379383.Google Scholar
Saucier, A. E., 1985, Geologic framework of the Travis Peak (Hosston) Formation of east Texas and northern Louisiana, in Finley, R. J., Dutton, S. P., Lin, Z. S., and Saucier, A. E., eds., The Travis Peak (Hosston) Formation: Geologic Framework, Core Studies, and Engineering Field Analysis: The University of Texas at Austin, Bureau of Economic Geology, report prepared for the Gas Research Institute under contract no. 5082-211-0708, 233 p.Google Scholar
Saunders, M., Gieger, L., Rodriguez, K., and Hargreaves, P., 2016, The delineation of pre-salt license blocks in the Deep Offshore Campeche-Yucatan Basin: American Association of Petroleum Geologists, Search and Discovery 10667, 8 p.Google Scholar
Sawyer, D. S., Buffler, R. T., and Pilger, R. H. Jr., 1991, The crust under the Gulf of Mexico basin, in Salvador, A., ed., The Gulf of Mexico Basin: The Geology of North America: Geological Society of America, 5372, doi:10.1130/DNAG-GNA-J.53.Google Scholar
Schenk, C. J., 2008, Jurassic–Cretaceous composite total petroleum system and geologic models for oil and gas assessment of the North Cuba Basin, Cuba, in U.S. Geological Survey North Cuba Basin Assessment Team, Jurassic–Cretaceous Composite Total Petroleum System and Geologic Assessment of Oil and Gas Resources of the North Cuba Basin, Cuba: U.S. Geological Survey, Digital Data Series DDS-69-M, 94 p.Google Scholar
Schlager, W., Buffler, R. T., Angstadt, D., and Phair, R. L., 1984, Geologic history of the southeastern Gulf of Mexico, in Buffler, R. T. and Schlager, W., eds., Deep Sea Drilling Project Initial Reports 77, US GPO, 715738.Google Scholar
Schlanger, S. O., and Jenkyns, H. C., 1976, Cretaceous oceanic anoxic events: causes and consequences: Geologie en Mijnbouw, 55, 179184.Google Scholar
Schlische, R. W., 2003, Progress in understanding the structural geology, basin evolution, and tectonic history of the Eastern North American Rift System, in LeTourneau, P. M. and Olsen, P. E., eds., The Great Rift Valleys of Pangea in Eastern North America: Columbia University Press, 2164.Google Scholar
Schulte, P. J., Alegret, L., Arenillas, I., et al., 2010, The Chicxulub asteroid impact and mass extinction at the Cretaceous–Paleogene boundary: Science, 327, 12141218, doi:10.1126/science.1177265.Google Scholar
Schulte, P., Smit, J., Deutsch, A., et al., 2012, Tsunami backwash deposits with Chicxulub impact ejecta and dinosaur remains from the Cretaceous–Palaeogene boundary in the La Popa Basin, Mexico: Sedimentology, 59, 737765, doi:10.1111/j.1365-3091.2011.01274.x.Google Scholar
Schumacher, D., and Parker, R. M., 1990, Possible pre-Jurassic origin for some Jurassic-reservoired oil, Cass Co., Texas, in Gulf Coast Oils and Gases, Their Characteristics, Origin, Distribution, and Exploration and Production Significance: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 9th Annual Research Conference, 59–68.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: American Association of Petroleum Geologists, Memoir 65, 177198.Google Scholar
Scotese, C. R., 2017: Paleomap project, www.scotese.com/Default.htm.Google Scholar
Scott, E. D., Denne, R. A., Kaiser, J. S., and Eickhoff, D. P., 2014, Impact on sedimentation into the north-central deepwater Gulf of Mexico as a result of the Chicxulub event: Gulf Coast Association of Geological Societies Journal, 3, 4150.Google Scholar
Scott, E., Denne, R., Kaiser, J., and Eichkoff, D., 2016, Immediate and post-event effects of the K/Pg boundary Chicxulub impact on the northern Gulf of Mexico: New Understanding of the Petroleum Systems of the Continental Margins of the World: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 32nd Annual Research Conference Proceedings, 96–109, doi:10.5724/gcs.12.32.Google Scholar
Scott, K. R., Hayes, W. E., and Fietz, R. P., 1961, Geology of the Eagle Mills Formation: Gulf Coast Association of Geological Societies Transactions, 11, 114.Google Scholar
Scott, R. W., 1984, Significant fossils of the Knowles Limestone, Lower Cretaceous, Texas, in Ventress, W. P. S., Bebout, D. G., Perkins, B. F., and Moore, C. H., eds., The Jurassic of the Gulf Rim: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 3rd Annual Research Conference Proceedings, 333–346.Google Scholar
Scott, R. W., 1990, Models and stratigraphy of mid-Cretaceous reef communities, Gulf of Mexico: in Lidz, B. H., ed., Concepts in Sedimentology and Paleontology, 2, Society of Economic Paleontologists and Mineralogists, 1102.Google Scholar
Scott, R. W., Benson, D. G., Morin, R. W., Shaffer, B. L., and Oboh-Ikuenobe, F. E., 2002, Integrated Albian-Lower Cenomanian chronostratigraphy standard, Trinity River section Texas, in Scott, R. W., ed., U. S. Gulf Coast Cretaceous Stratigraphy and Paleoecology: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 23rd Bob F. Perkins Memorial Conference, 277–334.Google Scholar
Scott, R., Whitney, W., Hojnacki, R., Lin, X., and Wang, Y., 2016, Albian stratigraphy of the San Marcos Platform, Texas: why the person formation correlates with Upper Fredericksburg group not Washita Group, in Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Gulf Section SEPM Foundation Perkins Rosen Research Conference, 536–545.Google Scholar
Sempere, J. C., Lin, J., Brown, H. S., Schouten, H., and Purdy, G. M., 1993, Segmentation and morphotectonic variations along a slow-spreading center: the Mid-Atlantic Ridge (24°00'N–30°40'N): Marine Geophysical Researches, 15, 153200.Google Scholar
Seni, S. J., and Jackson, M. P. A., 1983, Evolution of salt structures, East Texas diapir province, part 1: sedimentary record of halokinesis: American Association of Petroleum Geologists Bulletin, 67, 12191244.Google Scholar
Sharman, G. R., Covault, J. A., Stockli, D. F., Wroblewski, A F.-J., and Bush, M. A., 2016, Early Cenozoic drainage reorganization of the United States Western Interior: Gulf of Mexico routing system: Geology, 44, 187190. doi10.1130/G38765.1.Google Scholar
Shellhouse, K., 2017, The Cretaceous–Paleogene Boundary Deposit in LaSalle Parish, Louisiana: Master’s thesis, The University of Louisiana at Lafayette, 175 p.Google Scholar
Shideler, G. L., 1986, Regional Geologic Cross-Section Series of Neogene-Quaternary Deposits, Louisiana Continental Shelf: Geological Society of America, Map and Chart Series MC-54, 8 p.Google Scholar
Shultz, A. W., 2010: Facies and environments of the Lobo trend, Webb and Zapata counties, Texas: examples from core, in Long, J., Stapp, W. L., Debus, R. W., and Smith, A. N., eds., Contributions to the Geology of South Texas: South Texas Geological Society, 229240.Google Scholar
Sluijs, A., van Roij, L., Harrington, G. J., et al., 2014, Warming, euxinia and sea level rise during the Paleocene–Eocene Thermal Maximum on the Gulf Coastal Plain: implications for ocean oxygenation and nutrient cycling: Climate of the Past, 10, 14211439.Google Scholar
Smit, J., and Hertogen, J., 1980, An extraterrestrial event at the Cretaceous–Tertiary boundary: Nature, 285, 198.Google Scholar
Smith, T., 2013, Unleasing the mad dog: GeoExPro, 10, 2227.Google Scholar
Smith, T., 2015, Puzzling salt structures: GeoExPro, 12, 2024.Google Scholar
Smith, W. H. F., and Sandwell, D. T., 1997, Global seafloor topography from satellite altimetry and ship depth soundings, Science, 277, 19571962.Google Scholar
Snedden, J. W., and Jumper, R. S., 1990, Shelf and shoreface reservoirs, Tom Walsh-Owen Field, Texas, in Barwis, J. H., McPherson, J. G., and Studlick, J. R. J., eds., Sandstone Petroleum Reservoirs: Casebooks in Earth Sciences, Springer, 415436.Google Scholar
Snedden, J. W., and Kersey, D. G., 1982, Depositional environments and gas reproduction trends Olmos Sandstone, Upper Cretaceous, Webb County, Texas: Gulf Coast Association of Geological Societies Transactions, 32, 497518.Google Scholar
Snedden, J. W., and Liu, C., 2011, Recommendations for a uniform chronostratigraphic designation system for Phanerozoic depositional sequences: American Association of Petroleum Geologists Bulletin 95, 10951122, doi:10.1306/01031110138.Google Scholar
Snedden, J. W., and Stockli, D. F., 2019, Paleogeographic and depositional reconstruction of Oxfordian Aeolian sandstone reservoirs in Mexico offshore areas: comparison to the Norphlet Aeolian play of the Northern Gulf of Mexico: The Geological Society, London, Petroleum Geology of Mexico and the Northern Caribbean Conference (abstract).Google Scholar
Snedden, J. W., Tillman, R. W., and Culver, S. J., 2011, Genesis and evolution of a mid-shelf, storm-built sand ridge, New Jersey continental shelf, U.S.A.: Journal of Sedimentary Research, 81, 534552, doi:10.2110/jsr.2011.26.Google Scholar
Snedden, J. W., Galloway, W. E., Whiteaker, T. L., and Ganey-Curry, P. E., 2012, Eastward shift of deep-water fan axes during the Miocene in the Gulf of Mexico: possible causes and models: Gulf Coast Association of Geological Societies Journal, 1, 131144.Google Scholar
Snedden, J., Eddy, D., Christeson, G., et al., 2013, A new temporal model for eastern Gulf of Mexico Mesozoic deposition: Gulf Coast Association of Geological Societies Transactions, 63, 609612.Google Scholar
Snedden, J. W., Norton, I. O., Christeson, G. L., and Sanford, J. C., 2014, Interaction of deepwater deposition and a mid-ocean spreading center, eastern Gulf of Mexico basin, USA: Gulf Coast Association of Geological Societies Transactions, 64, 371383.Google Scholar
Snedden, J., Galloway, W. E., Ganey-Curry, P., and Blum, M., 2015, The geologic history of submarine fans in the deepwater Gulf of Mexico: Mesozoic to modern: Gulf Coast Association of Geological Societies Transactions, 65, 521527.Google Scholar
Snedden, J. W., Bovay, A. C., and Xu, J., 2016a, New models of Early Cretaceous source-to-sink pathways in the Eastern Gulf of Mexico, in Lowery, C. M., Snedden, J. W., and Blum, M. D., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Gulf Coast Section SEPM, Perkins Rosen Research Conference, 380–415.Google Scholar
Snedden, J. W., Virdell, J., Whiteaker, T. L., and Ganey-Curry, P., 2016b, A basin-scale perspective on Cenomanian–Turonian (Cretaceous) depositional systems, greater Gulf of Mexico (USA): Interpretation, 4, SC1SC22, doi:10.1190/INT-2015-0082.1.Google Scholar
Snedden, J. W., Galloway, W. E., Milliken, K. T., et al., 2018a, Validation of empirical source-to-sink scaling relationships in a continental-scale system: the Gulf of Mexico basin Cenozoic record: Geosphere, 14, 117, doi:10.1130/GES01452.1.Google Scholar
Snedden, J. W., Tinker, L. D. and Virdell, J., 2018b, Southern Gulf of Mexico Wilcox source-to-sink: investigating and predicting Paleogene Wilcox reservoirs in Eastern Mexico deep-water areas: American Association of Petroleum Geologists Bulletin, 102, 20452074.Google Scholar
Snedden, J. W., Norton, I., Hudec, M., Eljalafi, A., and Peel, F., 2018c, Paleogeographic reconstruction of the Louann salt basin in the Gulf of Mexico: American Association of Petroleum Geologists, Annual Convention and Exhibition (abs).Google Scholar
Snyder, F., and Ysaccis, R., 2018, New offshore exploration opportunities within the Salina Del Istmo Basin, Mexico: American Association of Petroleum Geologists, Search and Discovery 11143, 41 p, . doi:10.1306/11143snyder.Google Scholar
Sømme, T. O., Helland-Hansen, W., and Granjeon, D., 2009a, Impact of eustatic amplitude variations on shelf morphology, sediment dispersal, and sequence stratigraphic interpretation: icehouse versus greenhouse systems: Geology, 37, 587590, doi:10.1130/G31134Y.1.Google Scholar
Sømme, T. O., Helland-Hansen, W., Martinsen, O. J., and Thurmond, J. B., 2009b, Relationships between morphological and sedimentological parameters in source-to-sink systems: a basis for predicting semi-quantitative characteristics in subsurface systems: Basin Research, 21, 361387, doi:10.1111/j.1365-2117.2009.00397.x.Google Scholar
Sosa Patron, A. A., Cardenas Lopez, J. G., Lara, C. C., et al., 2009, Integrated geological interpretation and impact on the definition of Neogene Plays in the Isthmus Saline Basin, Mexico, in Bartolini, C. and Roman Ramos, J.R., eds., Petroleum Systems in the Southern Gulf of Mexico: American Association of Petroleum Geologists, Memoir 90, 2948.Google Scholar
Stabler, C., 2016, Triple-porosity diagram proposed to characterize complex carbonate reservoirs – examples from Mexico, in Lowery, C. M., Snedden, J. W., and Rosen, N. C., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 35th Annual Research Conference.Google Scholar
Steffens, G. S., Biegert, E. K., Sumner, H. S., and Bird, D., 2003, Quantitative bathymetric analysis of selected 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
Steier, A., and Mann, P., 2019, Late Mesozoic gravity sliding and Oxfordian hydrocarbon reservoir potential of the northern Yucatan margin: Marine and Petroleum Geology, 103, 681701.Google Scholar
Stephens, B. P., 2009, Basement controls on subsurface geologic patterns and coastal geomorphology across the northern Gulf of Mexico: implications for subsidence studies and coastal restoration: Gulf Coast Association of Geological Societies Transactions, 59, 729751.Google Scholar
Stephens, B. P., 2010, Basement controls on subsurface geologic patterns and near-surface geology across the northern Gulf of Mexico: a deeper perspective on coastal Louisiana: American Association of Petroleum Geologists, Search and Discovery 30129.Google Scholar
Stephenson, S. N., Roberts, G. G., Hoggard, M. J., and Whittaker, A. C., 2014, A Cenozoic uplift history of Mexico and its surroundings from longitudinal river profiles: Geochemistry, Geophysics, Geosystems, 15, 47344758.Google Scholar
Stern, R. J., and Dickinson, W. R., 2010, The Gulf of Mexico is a Jurassic backarc basin: Geosphere, 6, 739754, doi:10.1130/GES00585.1.Google Scholar
Stern, R. J., Anthony, E. Y., Ren, M., et al., 2011, Southern Louisiana salt dome xenoliths: first glimpse of Jurassic (ca. 160 Ma) Gulf of Mexico crust: Geology, 39, 315318.Google Scholar
Stoneburner, R. K., 2015, The discovery, reservoir attributes, and significance of the Hawkville Field and the Eagle Ford Trend: implications for future development: Gulf Coast Association of Geological Societies Transactions, 65, 377387.Google Scholar
Strong, M. A., 2013, Investigation and Characterization of Features on a Cretaceous–Paleogene Seismic Horizon in Northern Louisiana: Master’s thesis, University of Louisiana at Lafayette, 94 p.Google Scholar
Strong, M. A., and Kinsland, G. L., 2014, Chicxulub impact tsunami deposits at the K–Pg boundary in northern Louisiana?: American Association of Petroleum Geologists, Search and Discovery 330379.Google Scholar
Styzen, M. J., 1996, A chart in two sheets of the Late Cenozoic chronostratigraphy of the Gulf of Mexico. The Gulf Coast Section of the Society of Economic Paleontologists and Mineralogists, GCSSEPM Foundation 16th Annual Research Conference.Google Scholar
Suter, J. R., and Berryhill, H. L., 1985. Late Quaternary shelf-margin deltas, northwest Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 69, 7791.Google Scholar
Swanson, S. M., Enomoto, C. B., Dennen, K. O., Valentine, B. J., and Lohr, C. D., 2013, Geologic model for the assessment of undiscovered hydrocarbons in Lower to Upper Cretaceous carbonate rocks of the Fredericksburg and Washita groups, U.S. Gulf Coast Region: Gulf Coast Association of Geological Societies Transactions, 63, 423437.Google Scholar
Sweet, M. L., and Blum, M. D., 2012, Paleocene–Eocene Wilcox submarine canyons and thick deepwater sands of the Gulf of Mexico: very large systems in a greenhouse world, not a Messinian-like crisis: Gulf Coast Association of Geological Societies Transactions, 61, 443450.Google Scholar
Sweet, M. L., and Blum, M. D., 2016, Connections between fluvial to shallow marine environments and submarine canyons: implications for sediment transfer to deep water: Journal of Sedimentary Research, 86, 11471162, doi: 10.2110/jsr.2016.64.Google Scholar
Swift, T. E., and Mladenka, P., 1997, Technology tackles low-permeability sand in South Texas: Oil & Gas Journal, 95, 6872.Google Scholar
Syvitski, J. P. M., and Milliman, J. D., 2007, Geology, geography, and humans battle for dominance over the delivery of fluvial sediment to the coastal ocean: Journal of Geology, 115, 119.Google Scholar
Tada, R., Iturralde-Vinent, M. A., Matsui, T., et al., 2003, K/T boundary deposits in the Paleo-western Caribbean basin, in Bartolini, C., Buffler, R. T., and Blickwede, J., eds., The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation, and Plate Tectonics: American Association of Petroleum Geologists Memoir 79, 582604.Google Scholar
Talling, P. J., 2013, Hybrid submarine flows comprising turbidity current and cohesive debris flow: deposits, theoretical and experimental analyses, and generalized models: Geosphere, 9, 460488, doi:10.1130/GES00793.1.Google Scholar
Taylor, T. R., Giles, M. R., Hathon, L. A., et al., 2010, Sandstone diagenesis and reservoir quality prediction: models, myths, and reality: American Association of Petroleum Geologists Bulletin, 94, 10931132. doi:10.1306/04211009123.Google Scholar
Tew, B. H., Mink, R. M., Mancini, E. A., Mann, S. D., and Kopaska-Merkel, D. C., 1991, Regional Geologic Framework and Petroleum Geology of the Smackover Formation, Alabama Coastal Waters and Adjacent Federal Waters Area, Geological Survey of Alabama and State Oil and Gas Board, Energy and Coastal Geology Division, Report, 84 p.Google Scholar
Thacher, C., Stefani, J., Wu, C., et al., 2013, Subsalt imaging and 4D reservoir monitoring evaluation of Tahiti field, Gulf of Mexico: Society of Petroleum Engineers, Annual Meeting, 4875–4879.Google Scholar
Thomas, W. A., 2011, The Iapetan rifted margin of southern Laurentia: Geosphere, 7, 97120, doi: 10.1130/GES00574.1.Google Scholar
Tian, Y., Ayers, W. B., and McCain, W. D. Jr., 2012, Regional analysis of stratigraphy, reservoir characteristics, and fluid phases in the Eagle Ford Shale, South Texas: Gulf Coast Association of Geological Societies Transactions, 62, 471483.Google Scholar
Treviño García, F. J., 2012, Caracterización inicial de un campo marino en aguas someras del Sur del Golfo de México: Master’s thesis, Instituto Politecnico National, 193 p.Google Scholar
Trudgill, B. D., Rowan, M. G., Weimer, P., et al., 1995, The structural geometry and evolution of the salt-related Perdido Fold Belt, Alaminos Canyon, Northwestern Deep Gulf of Mexico: Society of Economic Paleontologists and Mineralogists Gulf Coast Section 16th Annual Research Conference, 275–284.Google Scholar
Trudgill, B. D., Rowan, M. G., Fiduk, J. C., et al., 1999, The Perdido fold belt, northwestern deep Gulf of Mexico, part 1: structural geometry, evolution and regional implications: American Association of Petroleum Geologists Bulletin, 83, 88113.Google Scholar
Tsikos, H., Jenkyns, H.C., Walsworth-Bell, B., et al., 2004, Carbon-isotope stratigraphy recorded by the Cenomanian–Turonian oceanic anoxic event: correlation and implications based on three key localities: Journal of the Geological Society of London, 161, 711719.Google Scholar
Tye, R. S., 1992, Fluvial–sandstone reservoirs of the Travis Peak Formation, East Texas Basin, in Miall, A. D., and Tyler, N., eds., The Three-Dimensional Facies Architecture of Terrigenous Clastic Sediments and Its Implications for Hydrocarbon Discovery and Recovery : Society of Economic Paleontologists and Mineralogists, Concepts in Sedimentology and Paleontology 3, 172178.Google Scholar
Tyler, N., and Ambrose, W. A., 1984, Facies Architecture and Production Characteristics of Strandplain Reservoirs in the Frio Formation: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 146, 42 p.Google Scholar
Tyler, N., and Ambrose, W. A., 1986, Depositional Systems and Oil and Gas Plays in the Cretaceous Olmos Formation, South Texas: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 152, 49 p.Google Scholar
Umbarger, K. F., 2018, Late Triassic North American Paleodrainage networks and sediment dispersal of the Chinle Formation: a quantitative approach using detrital zircons: MS thesis, University of Kansas, 131 p.Google Scholar
Umbarger, K. F., and Snedden, J. W., 2016, Delineation of post-KPg carbonate slope deposits as a sedimentary record of the Paleogene linkage of De Soto canyon and the Suwannee strait, northern Gulf of Mexico: Interpretation, 4, SC51SC60.Google Scholar
United States Department of Energy, Minerals Management Service, 1987, Correlation of Cenozoic Sediments, Gulf of Mexico Outer Continental Shelf, Part 1: Galveston Area Offshore Texas Through Vermilion Area Offshore Louisiana, OCS Report MMS 87-0026.Google Scholar
USGS, 2005, Assessment of undiscovered oil and gas resources of the North Cuba Basin, Cuba, 2004: United States Geological Survey, World Assessment of Oil and Gas Fact Sheet, 2 p, https://pubs.usgs.gov/fs/2005/3009/pdf/fs2005_3009.pdf.Google Scholar
USGS, 2014, Geology and assessment of unconventional resources of northeastern Mexico: US Geological Survey, Open-File Report 2015-1112, 1 p., doi:10.3133/ofr20151112.Google Scholar
Valdés, M. D., Rodriguez, L. V., and Garcia, E. C., 2009, Geochemical Integration and Interpretation of Source Rocks, Oils, and Natural Gases in Southeastern Mexico: American Association of Petroleum Geologists, Memoir 90, 337368.Google Scholar
Vallejo, V. V. A., Solis, E. F., Olivares, A., et al., 2012, Drilling a deep-water well in a subsalt structure in Mexico: Deep Offshore Technology International, paper 145, 1–17.Google Scholar
Van Avendonk, H. J., Eddy, D. R., Christeson, G., et al., 2013, Structure and early evolution of the Northwestern Gulf of Mexico: new constraints from marine seismic refraction data: American Association of Petroleum Geologists, Search and Discovery 90163.Google Scholar
Van Avendonk, H., Christeson, G., Norton, I., and Eddy, D., 2015, Continental rifting and sediment infill in the northwestern Gulf of Mexico: Geology, 43, 631634, doi:10.1130/G36798.1.Google Scholar
Vásquez, R. O., Cossey, S. P. J., van Nieuwenhuise, D. S., et al., 2014, New insights into the stratigraphic framework and depositional history of the Paleocene and Eocene Chicontepec Formation, onshore Eastern Mexico: American Association of Petroleum Geologists, Search and Discovery 30334.Google Scholar
Vega, F. J., and Lawton, T. F., 2011, Upper Jurassic (Lower Kimmeridgian-Olvido) carbonate strata from the La Popa Basin diapirs, NE Mexico: Boletín de la Sociedad Geológica Mexicana, 63(2), 313321.Google Scholar
Veltman, W., Velez, E., and Lujan, V., 2012, A fresh look for natural fracture characterization using advance borehole acoustics techniques: American Association of Petroleum Geologists, Search and Discovery 40915, 17 p.Google Scholar
Vernon, R. C., 1971, Possible Future Petroleum Potential of Pre-Jurassic, Western Gulf Basin: American Association of Petroleum Geologists, Memoir 15, 954979.Google Scholar
Vila-Concejo, A., Harris, D. L., Power, H. E., Shannon, A. M., and Webster, J. M., 2013, Sediment transport and mixing depth on a coral reef sand apron: Geomorphology, 222, 143150, doi:10.1016/j.geomorph.2013.09.034.Google Scholar
Waite, L. E., 2009, Edwards (Stuart City) shelf margin of South Texas: new data, new concepts: American Association of Petroleum Geologists, Search and Discovery 10177.Google Scholar
Walles, F. E., 1993, Tectonic and diagenetically induced seal failure within the south-western Great Bahamas Bank: Marine and Petroleum Geology, 10, 1428, doi:10.1016/0264-8172(93)90096-B.Google Scholar
Walsh, J. P., Wiberg, P. L., Aalto, R., Nittrouer, C. A. and Kuehl, S. A., 2016, Source-to-sink research: economy of the Earth’s surface and its strata: Earth-Science Reviews, 153, 16. doi:10.1016/j.earscirev.2015.11.010.Google Scholar
Wang, F. P., Hammes, U., and Qinghui, L., 2013, Overview of Haynesville shale properties and production, in Hammes, U. and Gale, J., eds., Geology of the Haynesville Gas Shale in East Texas and West Louisiana: American Association of Petroleum Geologists Bulletin, Memoir 105, 155177, doi:10.1306/13441848M1053527.Google Scholar
Ward, W. C., Keller, G., Stinnesbeck, W., and Adatte, T., 1995, Yucatan subsurface stratigraphy: implications and constraints for the Chicxulub impact: Geology, 23 873876, doi:10.1130/0091-7613(1995)023<0873:YNSSIA>2.3.CO;2.Google Scholar
Warwick, P. E., 2017, Geologic Assessment of Undiscovered Conventional Oil and Gas Resources in the Lower Paleocene Midway and Wilcox Groups, and the Carrizo Sand of the Claiborne Group of the Northern Gulf Coast Region: U.S. Geological Survey, Open-File Report 2017-111, 60 p.Google Scholar
Warzeski, E. R., 1987, Revised stratigraphy of the Mural Limestone: a lower Cretaceous carbonate shelf in Arizona and Sonora, in Dickinson, W. R. and Klute, M. F., eds., Mesozoic Rocks of Southern Arizona and Adjacent Areas: Arizona Geological Society, Digest 18, 335363.Google Scholar
Watkins, J. S., Bradshaw, B. E., Huh, S., Li, R., and Zhang, J., 1996a, Structure and distribution of growth faults in the northern Gulf of Mexico OCS: Gulf Coast Association of Geological Societies Transactions, 46, 6377.Google Scholar
Watkins, J. S., Bryant, W. R., and Buffler, R. T., 1996b, Structural Framework Map of the Northern Gulf of Mexico, Gulf Coast Association of Geological Societies, Special Publication 80, 9598.Google Scholar
Weber, R. D., and Parker, B. W., 2016, Pre-Albian biostratigraphical and paleoecological observations from the De Soto Canyon area; Gulf of Mexico, USA, in Lowery, C. M., Snedden, J. W., and Blum, M. D., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Gulf Coast Section SEPM, Perkins Rosen Research Conference paper, 154–172.Google Scholar
Webster, R. E., Luttner, D., and Liu, L., 2008, Fairway James Lime Field, East Texas: still developing after 48 years: American Association of Petroleum Geologists, Search and Discovery 110061.Google Scholar
Weeks, L. G., 1958, Habitat of oil and some factors that control it, in Weeks, L., ed., Habitat of Oil: A Symposium: American Association of Petroleum Geologists, 161.Google Scholar
Weeks, W. B., 1938, South Arkansas stratigraphy with emphasis on the older Coastal Plain beds: American Association of Petroleum Geologists Bulletin, 22, 958964.Google Scholar
Weimer, P., 1990, Sequence stratigraphy, facies geometries, and depositional history of the Mississippi Fan, Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 74, 425453.Google Scholar
Weimer, P., and Buffler, R. T., 1992, Structural geology and evolution of the Mississippi Fan Fold Belt, Deep Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 76, 225251.Google Scholar
Weimer, P., Varnai, P., Budhijanto, F. M., et al., 1998, Sequence stratigraphy of Pliocene and Pleistocene turbidite systems, northern Green Canyon and Ewing Bank (offshore Louisiana), northern Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 82, 918960.Google Scholar
Weimer, P., Zimmerman, E., Bouroullec, R., et al., 2016, Temporal and spatial evolution of reservoirs, Northern Deepwater Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 66, 539555.Google Scholar
Weimer, P., Bouroullec, R., and Serrano, O., 2017, Petroleum geology of the Mississippi Canyon, Atwater Valley, western DeSoto Canyon, and western Lloyd Ridge protraction areas, northern deep-water Gulf of Mexico: traps, reservoirs, and tectono-stratigraphic evolution: American Association of Petroleum Geologists Bulletin, 101, 10731108.Google Scholar
Weise, B. R., 1980, Wave-Dominated Delta Systems of the Upper Cretaceous San Miguel Formation, Maverick Basin, South Texas: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 107, 39 p.Google Scholar
Weislogel, A. L., Hunt, B., Lisi, A., Lovell, T., and Robinson, D. M., 2015, Detrital zircon provenance of the eastern Gulf of Mexico subsurface: constraints on Late Jurassic paleogeography and sediment dispersal of North America, in Anderson, T. H., Didenko, A. N., Johnson, C. L., Khanchuk, A. I., and MacDonald, J. H. Jr., eds., Late Jurassic Margin of Laurasia: A Record of Faulting Accommodating Plate Rotation: Geological Society, Special Publication 513, 89105, doi:10.1130/2015.2513(02).Google Scholar
Weislogel, A. L., Wiley, K., Bowman, S., and Robinson, D., 2016, Triassic–Jurassic provenance signatures in the nascent Eastern Gulf of Mexico region from detrital zircon geochronology, in Lowery, C. M., Snedden, J. W., and Blum, M. D., eds., Mesozoic of the Gulf Rim and Beyond: New Progress in Science and Exploration of the Gulf of Mexico Basin: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, Perkins Rosen Research Conference, 252–270.Google Scholar
Weissert, H., Lini, A., Föllmi, K. B., and Kuhn, O., 1998, Correlation of Early Cretaceous carbon isotope stratigraphy and platform drowning events: a possible link?: Palaeogeography, Palaeoclimatology, Palaeoecology, 137, 189203.Google Scholar
Wenger, L. M., Goodoff, L. R., Gross, O. P., Harrison, S. C., and Hood, K. C., 1994, Northern Gulf of Mexico: an integrated approach to source, maturation, and migration, in Scheidermann, N., Cruz, P., and Sanchez, R., eds., Geologic Aspects of Petroleum Systems: 1st Joint AAPG-AMGP Hedberg Research Conference, 5 p.Google Scholar
Whidden, K. J., Pitman, J., Pearson, O., et al., 2018, The 2017 USGS Assessment of undiscovered oil and gas resources in the Eagle Ford Shale and associated Cenomanian Strata, Texas (abs.): American Association of Petroleum Geologists, Annual Convention & Exhibition.Google Scholar
White, B. R., and Sawyer, J. R., 1966, Black Lake Field: before and after: Gulf Coast Association of Geological Societies Transactions, 16, 219225.Google Scholar
White, C., and Snedden, J., 2016, Seismic stratigraphic analysis of the Yoakum/Lavaca Canyon System, South Texas, USA (abs.): American Geophysical Union, Fall Meeting 2016, abstract #EP43B-0956.Google Scholar
White, C., and Snedden, J.,in prep., Seismic stratigraphic analysis of the Yoakum/Lavaca Canyon System, South Texas, USA.Google Scholar
White, G., Blanke, S., and Clawson, C. II, 1999, Evolutionary model of the Jurassic sequences of the East Texas Basin: implications for hydrocarbon exploration: Gulf Coast Association of Geological Societies Transactions, 49, 488498.Google Scholar
Wiley, K. S., 2017, Provenance of Syn-rift Clastics in the Eastern Gulf of Mexico: Insight from U–Pb Detrital Zircon Geochronology and Thin Sections: MS thesis, West Virginia University, 194 p.Google Scholar
Williams-Rojas, C. T., and Hurley, N. F., 2001, Geologic controls on reservoir performance in Muspac and Catedral Gas Fields, Southeast Mexico, in Bartolini, C., Buffler, R. T., and Cantú-Chapa, A., eds., The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists, Memoir 75, 443472.Google Scholar
Williams-Rojas, C., Reyey-Tovar, E., Miranda Peralta, L., et al., 2012, Hydrocarbon potential of the deepwater portion of the “Salina del Istmo” province, southeastern Gulf of Mexico, Mexico: Gulf Coast Association of Geological Societies Transactions, 62, 641644.Google Scholar
Wilson, H. H., 1993, The age of salt in the Gulf of Mexico Basin: Journal of Petroleum Geology, 16, 125152.Google Scholar
Wilson, J. L., 1975, Carbonate Facies in Geologic History: Springer-Verlag, 471 p.Google Scholar
Winker, C. D., 1982, Cenozoic shelf margins, northwestern Gulf of Mexico: Gulf Coast Association of Geological Societies Transactions, 32, 427448.Google Scholar
Winker, C. D., 1984, Clastic shelf margins of the post-Comanchean Gulf of Mexico: implications for deep-water sedimentation: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 5th Annual Research Conference, 109–120.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 Weimer, P., Slatt, R. M., Coleman, J., et al., eds., Deep-Water Reservoirs of the World: Society of Economic Paleontologists and Mineralogists Gulf Coast Section 20th Annual Bob F. Perkins Research Conference, 1059–1086.Google Scholar
Winker, C. D., and Buffler, R. T., 1988, Paleogeographic evolution of early deep-water Gulf of Mexico and margins, Jurassic to Middle Cretaceous (Comanchean): American Association of Petroleum Geologists Bulletin, 72, 318346.Google Scholar
Winker, C. D., Morton, R. A., Ewing, T. E., and Garcia, D. D., 1983, Depositional Setting, Structural Style, and Sandstone Distribution in Three Geopressured Geothermal Areas, Texas Gulf Coast: The University of Texas at Austin, Bureau of Economic Geology, Report of Investigations 134, 60 p.Google Scholar
Winston, G. O., 1976, Florida’s Ocala Uplift is not an uplift: American Association of Petroleum Geologists Bulletin, 60(6), 992994.Google Scholar
Winter, R. R., 2018, Coarse-grained deep water, slope and basin-floor systems: Influence of tectonic processes on internal and external architecture: unpublished PhD dissertation, The University of Texas at Austin, 163 p.Google Scholar
Withjack, M. O., Schlische, R. W., and Olsen, P. E., 1998, Diachronous rifting, drifting, and inversion on the passive margin of central eastern North America: an analog for other passive margins: American Association of Petroleum Geologists Bulletin, 82, 817835.Google Scholar
Withjack, M. O., Schlische, R. W., and Olsen, P. E., 2002, Rift-basin structure and its influence on sedimentary systems, in Renault, R. and Ashley, G., eds., Sedimentation in Continental Rifts: Society of Economic Paleontologists and Mineralogists, Special Publication, 73, 5781, doi:10.2110/pec.02.73.0057.Google Scholar
Witt, C., Brichau, S., and Carter, A., 2012, New constraints on the origin of the Sierra Madre de Chiapas (south Mexico) from sediment provenance and apatite thermochronometry: Tectonics, 31(6), TC6001, doi:10.1029/2012TC003141.Google Scholar
Wood, G. D., and Benson, D. G. Jr., 2000, The North American occurrence of the algal coenobium plaesiodictyon: paleographic, paleoecologic, and biostratigraphic importance in the Triassic: American Association of Stratigraphic Palynologists, 24, 820.Google Scholar
Woods, R. D., and Addington, J. W., 1973, Pre-Jurassic geologic framework Northern Gulf Basin: Gulf Coast Association of Geological Societies Transactions, 23, 92108.Google Scholar
Woolf, K., 2012, Regional Character of the Lower Tuscaloosa Formation Depositional Systems and Trends in Reservoir Quality: Ph.D. dissertation, The University of Texas at Austin, 241 p.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: Geological Society of America, 97138.Google Scholar
Wu, X., and Galloway, W. E., 2002, Upper Miocene depositional history of the central Gulf of Mexico basin: Gulf Coast Association of Geological Societies Transactions, 52, 10191030.Google Scholar
Wu, X., and Galloway, W. E., 2003, Upper Miocene depositional history and paleogeographic evolution of central Gulf of Mexico basin, in Roberts, H. H., Rosen, N. C., Fillon, R. H., and Anderson, J. B., eds., Shelf Margin Deltas and Linked Downslope Petroleum Systems: Global Significance and Future Exploration Potential: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 23rd Annual Gulf Coast Section SEPM Bob. F. Perkins Research Conference Proceedings, 139–165.Google Scholar
Xaio, L., Zhao, J. W., Liu, H. S., et al., 2017, Ages and geochemistry of the basement granites of the Chicxulub impact crater: implications for peak ring formation: Lunar and Planetary Science, 48, 12.Google Scholar
Xu, J., Snedden, J. W. Stockli, D. F., Fulthorpe, C. S., and Galloway, W. E., 2016a, Early Miocene continental-scale sediment supply to the Gulf of Mexico Basin based on detrital zircon analysis: Geological Society of America Bulletin, 129, 322, doi:10.1130/B31465.1.Google Scholar
Xu, J., Snedden, J. W., Galloway, W. E., Milliken, K. T., and Blum, M. D., 2016b, Channel-belt scaling relationship and application to early Miocene source-to-sink systems in the Gulf of Mexico basin: Geosphere, 13, 179200.Google Scholar
Xu, J., Stockli, D. F., and Snedden, J. W., 2017, Enhanced provenance interpretation using combined U–Pb and (U–Th)/He double dating of detrital zircon grains from lower Miocene strata, proximal Gulf of Mexico Basin, North America: Earth and Planetary Science Letters, 475, 4457.Google Scholar
Xue, L., 1997, Depositional cycles and evolution of the Paleogene Wilcox strata, Gulf of Mexico basin, Texas: American Association of Petroleum Geologists Bulletin, 81, 937953.Google Scholar
Xue, L., and Galloway, W. E. 1993, Sequence stratigraphic and depositional framework of the Paleocene lower Wilcox strata, northwest Gulf of Mexico Basin: Gulf Coast Association of Geological Societies Transactions, 43, 453464.Google Scholar
Xue, L., and Galloway, W. E., 1995, High-resolution depositional framework of the Paleocene Middle Wilcox strata, Texas coastal plain: American Association of Petroleum Geologists Bulletin, 79, 205230.Google Scholar
Yancey, T. E., 1996, Stratigraphy and depositional environments of the Cretaceous–Tertiary boundary complex and basal Paleocene section, Brazos River, Texas: Gulf Coast Association of Geological Societies Transactions, 46, 433442.Google Scholar
Yancey, T. E., Elsik, W. C., and Sancay, R. H., 2003, The palynological record of Late Eocene climate change, northwest Gulf of Mexico, in Prothero, D.R., Ivany, L.C., and Nesbitt, E.A., eds., From Greenhouse to Icehouse: The Marine Eocene–Oligocene Transition: Columbia University Press, 252268.Google Scholar
Yancey, T. E., Heizler, M. T., Miller, B. V., and Guillemette, R. N., 2018, Eocene–Oligocene chronostratigraphy of ignimbrite flareup volcanic ash beds on the Gulf of Mexico coastal plains: Geosphere, 14, 12321252. doi.org/10.1130/GES01621.1.Google Scholar
Young, K., 1977, Rocks of the Austin area, in Young, K., ed., Guidebook to the Geology of Travis County, University of Texas Student Geology Society, 1625.Google Scholar
Young, K., 1985, The Austin Division of central Texas, in Young, K. and Woodruff, C. M., Austin Chalk in Its Type Area, Stratigraphy and Structure, Austin Geological Society, Guidebook, 7, 352.Google Scholar
Ysaccis, R., Hernandez, G., Villa, R., et al., 2006, Structural styles of the Canela and its surrounded areas, Tabasco– Southern Mexico: American Association of Petroleum Geologists Search and Discovery 90052.Google Scholar
Yurewicz, D. A., Marler, T. B., Meyerholtz, K. A., and Siroky, F. X., 1993, Early Cretaceous carbonate platform, north rim of the Gulf of Mexico, Mississippi and Louisiana, in Simo, J. A. T., Scott, R. W., and Masse, J.-P., eds., Cretaceous Carbonate Platforms: American Association of Petroleum Geologists Memoir 56, 8196.Google Scholar
Yurewicz, D. A., Chuchla, R. J., Richardson, M., et al., 1997, Hydrocarbon generation and migration in the Tampico–Misantla Basin and Sierra Madre Oriental, east-central Mexico: evidence from an exhumed oil field in the Sierra de el Abra, in Sedimentation and Diagenesis of Middle Cretaceous Platform Margins, East Central Mexico: American Association of Petroleum Geologists Bulletin, Field Trip Guidebook, Dallas Geological Society and Society of Economic Paleontologists and Mineralogists, 24 p.Google Scholar
Zarra, L., 2007, Chronostratigraphic framework for the Wilcox Formation (Upper Paleocene–Lower Eocene) in the deep-water Gulf of Mexico: biostratigraphy, sequences, and depositional systems, 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: Society of Economic Paleontologists and Mineralogists Gulf Coast Section, 26th Annual Research Conference, 81–146.Google Scholar
Zeng, H., and Hentz, T. F., 2002, High-frequency sequence stratigraphy from seismic sedimentology: applied to Miocene, Vermilion Block 50, Tiger Shoal area, offshore Louisiana: American Association of Petroleum Geologists Bulletin, 88, 153174.Google Scholar
Zhang, J., Steel, R., and Ambrose, W. A., 2016, Greenhouse shoreline migration: Wilcox deltas: American Association of Petroleum Geologists Bulletin, 100, 18031831.Google Scholar
Zhang, J., Covault, J., Pyrcz, M., et al., 2018, Quantifying sediment supply to continental margins: applications to the Paleogene Wilcox Group, Gulf of Mexico: American Association of Petroleum Geologists Bulletin, 100, 16851702, doi: 10.1306/01081817308.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • John W. Snedden, University of Texas, Austin, William E. Galloway, University of Texas, Austin
  • Book: The Gulf of Mexico Sedimentary Basin
  • Online publication: 13 September 2019
  • Chapter DOI: https://doi.org/10.1017/9781108292795.016
Available formats
×

Save book to Dropbox

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

  • References
  • John W. Snedden, University of Texas, Austin, William E. Galloway, University of Texas, Austin
  • Book: The Gulf of Mexico Sedimentary Basin
  • Online publication: 13 September 2019
  • Chapter DOI: https://doi.org/10.1017/9781108292795.016
Available formats
×

Save book to Google Drive

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

  • References
  • John W. Snedden, University of Texas, Austin, William E. Galloway, University of Texas, Austin
  • Book: The Gulf of Mexico Sedimentary Basin
  • Online publication: 13 September 2019
  • Chapter DOI: https://doi.org/10.1017/9781108292795.016
Available formats
×