Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-vfjqv Total loading time: 0 Render date: 2024-04-26T13:16:40.917Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  04 March 2021

Peter B. Flemings
Affiliation:
University of Texas, Austin
Get access
Type
Chapter
Information
A Concise Guide to Geopressure
Origin, Prediction, and Applications
, pp. 249 - 263
Publisher: Cambridge University Press
Print publication year: 2021

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

Abrams, M. A., & Boettcher, S. S. (2000). Mapping migration pathways using geophysical data, seabed core geochemistry and submersible observations in the central Gulf of Mexico. Paper presented at the AAPG Annual Meeting Expanded Abstracts, New Orleans, Louisiana.Google Scholar
Alberty, M. W., Hafle, M. E., Minge, J. C., Byrd, T. M., & Exploration, BP. (1999). Mechanisms of Shallow Waterflows and Drilling Practices for Intervention. SPE Drilling & Completion, 14(2), 123129.CrossRefGoogle Scholar
Alberty, M. W., & McLean, M. R. (2001). Fracture Gradients in Depleted Reservoirs - Drilling Wells in Late Reservoir Life. Paper presented at the SPE/IADC Drilling Conference, Amsterdam, The Netherlands.Google Scholar
Alberty, M. W., & McLean, M. R. (2004). A Physical Model for Stress Cages. Paper presented at the SPE Annual Technical Conference and Exhibition, 26–29 September, Houston, Texas.Google Scholar
Alberty, M. W., & McLean, M. R. (2014). The Use of Modeling to Enhance the Analysis of Formation Pressure Integrity Tests. Paper presented at the IADC/SPE Drilling Conference and Exhibition, Fort Worth, Texas.Google Scholar
Albertz, M., Beaumont, C., & Ings, S. J. (2011). Geodynamic Modeling of Sedimentation-induced Overpressure, Gravitational Spreading, and Deformation of Passive Margin Mobile Shale Basins. In Shale Tectonics (pp. 29–62).Google Scholar
Alexander, L. L., & Flemings, P. B. (1994). Architecture and Evolution of a Plio-Pleistocene Salt Withdrawal Mini-Basin: Eugene Island, Block 330, Offshore Louisiana. AAPG Annual Meeting. Denver, CO.Google Scholar
Alexander, L. L., & Flemings, P. B. (1995). Geologic Evolution of a Pliocene-Pleistocene Salt-Withdrawal Minibasin: Eugene Island Block 330, Offshore Louisiana. AAPG Bulletin, 79(12), 17371756.Google Scholar
Allen, D. F., Best, D. L., Evans, M., & Holenka, J. M. (1993). The Effect of Wellbore Condition on Wireline and MWD Neutron Density Logs. SPE Formation Evaluation, 8(01), 50–56.Google Scholar
Anderson, E. M. (1951). The dynamics of faulting and dyke formation with applications to Britain. Edinburgh: Oliver and Boyd.Google Scholar
Anderson, R. L., Ratcliffe, I., Greenwell, H. C., Williams, P. A., Cliffe, S., & Coveney, P. V. (2010). Clay swelling – A challenge in the oilfield. Earth-Science Reviews, 98(3–4), 201216.Google Scholar
Aplin, A. C., Yang, Y., & Hansen, S. (1995). Assessment of β the compression coefficient of mudstones and its relationship with detailed lithology. Marine and Petroleum Geology, 12(8), 955963.CrossRefGoogle Scholar
Athy, L. F. (1930). Density, Porosity, and Compaction of Sedimentary Rocks. AAPG Bulletin, 14(1), 124.Google Scholar
Balk, R. (1949). Structure of Grand Saline Salt Dome, Van Zandt County, Texas. AAPG Bulletin, 33(11), 17911829.Google Scholar
Balk, R. (1953). Salt structure of Jefferson Island salt dome, Iberia and Vermilion Parishes, Louisiana. AAPG Bulletin, 37(11), 24552474.Google Scholar
Bangs, N. L., Shipley, T. H., Gulick, S. P. S., Moore, G. F., Kuromoto, S., & Nakamura, Y. (2004). Evolution of the Nankai Trough décollement from the trench into the seismogenic zone: Inferences from three-dimensional seismic reflection imaging. Geology, 32(4).Google Scholar
Barker, J. W., & Meeks, W. R. (2003). Estimating fracture gradient in Gulf of Mexico deepwater, shallow, massive salt sections. Paper presented at the SPE Annual Technical Conference and Exhibition, Denver, Colorado.Google Scholar
Barton, C. A., Zoback, M. D., & Moos, D. (1995). Fluid flow along potentially active faults in crystalline rock. Geology, 23(8), 683686.Google Scholar
Becker, D. E., Crooks, J. H. A., Been, K., & Jefferies, M. G. (1987). Work as a criterion for determining in situ and yield stresses in clays. Canadian Geotechnical Journal, 24(4), 549564.Google Scholar
Beeunas, M. A., Hudson, T. A., Valley, J. A., Clark, W. Y., & Baskin, D. K. (1999). Reservoir continuity and architecture of the Genesis field, Gulf of Mexico (Green Canyon 205); an integration of fluid geochemistry within the geologic and engineering framework. Gulf Coast Association of Geological Societies Transactions, 49, 9095.Google Scholar
Benson, S. M., & Cole, D. R. (2008). CO2 Sequestration in Deep Sedimentary Formations. Elements, 4(5), 325331.Google Scholar
Berg, R. R. (1975). Capillary pressures in stratigraphic traps. AAPG Bulletin, 59(6), 939956.Google Scholar
Best, K. D. (2002). Development of an Integrated Model for Compaction/Water Driven Reservoirs and its Application to the J1 and J2 Sands at Bullwinkle, Green Canyon Block 65, Deepwater Gulf of Mexico. Master’s thesis, The Pennsylvania State University.Google Scholar
Bethke, C. M., Altaner, S. P., Harrison, W. J., & Upson, C. (1988). Supercomputer Analysis of Sedimentary Basins. Science, 239(4837), 261267.Google Scholar
Biot, M. A. (1941). General Theory of Three‐Dimensional Consolidation. Journal of Applied Physics, 12(2), 155164.Google Scholar
Birchwood, R. A., & Turcotte, D. L. (1994). A unified approach to geopressuring, low-permeability zone formation, and secondary porosity generation in sedimentary basins. Journal of Geophysical Research: Solid Earth, 99(B10), 2005120058.CrossRefGoogle Scholar
Bird, D. E., Burke, K., Hall, S. A., & Casey, J. F. (2005). Gulf of Mexico tectonic history: Hotspot tracks, crustal boundaries, and early salt distribution. AAPG Bulletin, 89(3), 311328.Google Scholar
Blunt, M. J. (2017). Multiphase Flow in Permeable Media: A Pore-Scale Perspective. Cambridge, United Kingdom: Cambridge University Press.Google Scholar
Boebert, E., & Blossom, J. M. (2016). Deepwater Horizon. Cambridge, Massachusetts: Harvard University Press.Google Scholar
Boehm, A., & Moore, J. C. (2002). Fluidized sandstone intrusions as an indicator of Paleostress orientation, Santa Cruz, California. Geofluids, 2(2), 147161.Google Scholar
Bowers, G. L. (1995). Pore Pressure Estimation From Velocity Data: Accounting for Overpressure Mechanisms Besides Undercompaction. SPE Drilling & Completion, 10(02), 8995.Google Scholar
Bowers, G. L. (2001). Determining an Appropriate Pore-Pressure Estimation Strategy. Paper presented at the Offshore Technology Conference, Houston, Texas.Google Scholar
Braunsdorf, N., & Kittridge, M. (2003). Overburden Pressure Estimation in Deep Water Settings. Paper presented at the PSU GeoFluids Consortium Annual Meeting (2003.15), Santa Cruz, California.Google Scholar
Breckels, I. M., & van Eekelen, H. A. M. (1982). Relationship Between Horizontal Stress and Depth in Sedimentary Basins. Journal of Petroleum Technology, 34(9), 2191–2199Google Scholar
Brooks, J. M., Kennicutt, M. C., 2nd, Fisher, C. R., Macko, S. A., Cole, K., Childress, J. J., et al. (1987). Deep-sea hydrocarbon seep communities: evidence for energy and nutritional carbon sources. Science, 238(4830), 11381142.Google Scholar
Bryn, P., Berg, K., Forsberg, C. F., Solheim, A., & Kvalstad, T. J. (2005). Explaining the Storegga Slide. Marine and Petroleum Geology, 22(1–2), 1119.Google Scholar
Burland, J. B. (1990). On the compressibility and shear strength of natural clays. Geotechnique, 40(3), 329378.Google Scholar
Burwicz, E., Reichel, T., Wallmann, K., Rottke, W., Haeckel, M., & Hensen, C. (2017). 3-D basin-scale reconstruction of natural gas hydrate system of the Green Canyon, Gulf of Mexico. Geochemistry, Geophysics, Geosystems, 18(5), 19591985.Google Scholar
Butterfield, R. (1979). A natural compression law for soils (an advance on e-logp’). Geotechnique, 29(4), 469480.Google Scholar
Casagrande, A. (1936). The determination of the pre-consolidation load and its practical significance. In Casagrande, A., Rutledge, P. C., & Watson, J. D. (Eds.), Proceedings of the 1st International Conference on Soil Mechanics and Foundation Engineering (Vol. 3, pp. 6064): American Society of Civil Engineers.Google Scholar
Casey, B. (2014). The Consolidation and Strength Behavior of Mechanically Compressed Fine-Grained Sediments. Doctoral thesis, Massachusetts Institute of Technology. Retrieved from http://hdl.handle.net/1721.1/90039Google Scholar
Casey, B., Germaine, J. T., Flemings, P. B., & Fahy, B. P. (2015). Estimating horizontal stresses for mudrocks under one-dimensional compression. Marine and Petroleum Geology, 65, 178186.Google Scholar
Casey, B., Germaine, J. T., Flemings, P. B., & Fahy, B. P. (2016). In situ stress state and strength in mudrocks. Journal of Geophysical Research-Solid Earth, 121(8), 56115623.Google Scholar
Casey, B., Germaine, J. T., Flemings, P. B., Reece, J. S., Gao, B., & Betts, W. (2013). Liquid limit as a predictor of mudrock permeability. Marine and Petroleum Geology, 44, 256263.Google Scholar
Casey, B., Reece, J. S., & Germaine, J. T. (2019). One-dimensional normal compression laws for resedimented mudrocks. Marine and Petroleum Geology, 103, 397403.Google Scholar
Chopra, S., & Huffman, A. R. (2006). Velocity determination for pore-pressure prediction. The Leading Edge, 25(12), 15021515.Google Scholar
Christman, S. A. (1973). Offshore Fracture Gradients. Journal of Petroleum Technology, 25(08), 910–914.Google Scholar
Comisky, J. T. (2002). Petrophysical Analysis and Geologic Model for the Bullwinkle J Sands with Implications for Time-Lapsed Reservoir Monitoring, Green Canyon Block 65, Offshore Louisiana. Master’s thesis, The Pennsylvania State University.Google Scholar
Comisky, J. T., Santiago, M., McCollom, B., Buddhala, A., & Newsham, K. E. (2011). Sample Size Effects on the Application of Mercury Injection Capillary Pressure for Determining the Storage Capacity of Tight Gas and Oil Shales. Paper presented at the Canadian Unconventional Resources Conference, Calgary, Canada.Google Scholar
Cosgrove, J. W. (2001). Hydraulic fracturing during the formation and deformation of a basin: A factor in the dewatering of low-permeability sediments. AAPG Bulletin, 85(4), 737748.Google Scholar
Couzens-Schultz, B. A., Axon, A., Azbel, K., Hansen, K. S., Haugland, M., Sarker, R., et al. (2013). Pore Pressure Prediction in Unconventional Resources. Paper presented at the International Petroleum Technology Conference, Beijing, China.Google Scholar
Couzens-Schultz, B. A., Hedlund, C. A., & Guzman, C. E. (2006). Integrating Geology and Velocity Data to Constrain Pressure Prediction in Foldbelts. Paper presented at the International Oil Conference and Exhibition in Mexico, Cancun, Mexico.Google Scholar
Craig, R. F. (2004). Craig’s Soil Mechanics (7th ed.). London; New York: Spon Press.Google Scholar
Crook, A. J. L., Willson, S. M., Yu, J. G., & Owen, D. R. J. (2006). Predictive modelling of structure evolution in sandbox experiments. Journal of Structural Geology, 28(5), 729744.Google Scholar
Cruz-Atienza, V. M., Villafuerte, C., & Bhat, H. S. (2018). Rapid tremor migration and pore-pressure waves in subduction zones. Nature Communications, 9(1), 2900.Google Scholar
Daines, S. R. (1982). Prediction of fracture pressures for wildcat wells. Journal of Petroleum Technology, 34(4), 863872.Google Scholar
Dake, L. P. (1978a). Chapter 1 Some Basic Concepts in Reservoir Engineering. In Dake, L. P. (Ed.), Fundamentals of Reservoir Engineering (Vol. 8, pp. 143): Elsevier.Google Scholar
Dake, L. P. (1978b). Chapter 10 Immiscible Displacement. In Dake, L. P. (Ed.), Fundamentals of Reservoir Engineering (Vol. 8, pp. 343430): Elsevier.Google Scholar
Darby, D., Haszeldine, R. S., & Couples, G. D. (1996). Pressure cells and pressure seals in the UK Central Graben. Marine and Petroleum Geology, 13(8), 865878.Google Scholar
Darby, D., Haszeldine, R. S., & Couples, G. D. (1998). Central North Sea overpressures: insights into fluid flow from one- and two-dimensional basin modelling. In Duppenbecker, S. J. & Iliffe, J. E. (Eds.), Basin Modelling: Practice and Progress (Vol. 141, pp. 95107). London: Geological Society, Special Publications.Google Scholar
Dawson, W. C., & Almon, W. R. (2006). Shale Facies and Seal Variability in Deepwater Depositional Systems. Poster presented at the AAPG Annual Convention (Search and Discovery Article #40199), Houston, TX.Google Scholar
Day-Stirrat, R. J., Aplin, A. C., Środoń, J., & van der Pluijm, B. A. (2008). Diagenetic Reorientation of Phyllosilicate Minerals in Paleogene Mudstones of the Podhale Basin, Southern Poland. Clays and Clay Minerals, 56(1), 100111.Google Scholar
Day-Stirrat, R. J., Flemings, P. B., You, Y., Aplin, A. C., & van der Pluijm, B. A. (2012). The fabric of consolidation in Gulf of Mexico mudstones. Marine Geology, 295298, 7785.Google Scholar
de Gennes, P.-G., Brochard-Wyart, F., & Quéré, D. (2004). Capillarity and Wetting Phenomena (A. Reisinger, Trans.). New York: Springer Science+Business Media.CrossRefGoogle Scholar
de Marsily, G. (1986). Quantitative Hydrogeology: Groundwater Hydrology for Engineers. Orlando, FL: Academic Press.Google Scholar
Deepwater Horizon Study Group. (2011). Final Report on the Investigation of the Macondo Well Blowout. Retrieved from http://ccrm.berkeley.edu/index.shtmlGoogle Scholar
Dewhurst, D. N., Brown, K. M., Clennell, M. B., & Westbrook, G. K. (1996). A comparison of the fabric and permeability anisotropy of consolidated and sheared silty clay. Engineering Geology, 42(4), 253267.Google Scholar
Dickinson, G. (1953). Geological aspects of abnormal reservoir pressures in Gulf Coast Louisiana. AAPG Bulletin, 37, 410432.Google Scholar
Dugan, B., & Flemings, P. B. (2000a). Overpressure and Fluid Flow in the New Jersey Continental Slope: Implications for Slope Failure and Cold Seeps. Science, 289(5477), 288291.Google Scholar
Dugan, B., & Flemings, P. B. (2000b). Rapid Sediment Loading, Lateral Fluid Flow, and Slope Stability on the US East Coast Margin. Paper presented at the AGU Fall Meeting, San Francisco, CA.Google Scholar
Dusseault, M. B., Maury, V., & Santarelli, F. J. (2004). Drilling Through Salt: Constitutive Behavior and Drilling Strategies. Paper presented at the Gulf Rocks 2004, the 6th North America Rock Mechanics Symposium (NARMS), Houston, Texas.Google Scholar
Dutta, N. C. (1983). Shale compaction and abnormal pore‐pressures: A model of geopressures in the Gulf Coast basin. Paper presented at the SEG Technical Program Expanded Abstracts 1983, Las Vegas.Google Scholar
Dutta, N. C. (2002a). Deepwater geohazard prediction using prestack inversion of large offset P-wave data and rock model. The Leading Edge, 21(2), 193198.Google Scholar
Dutta, N. C. (2002b). Geopressure prediction using seismic data: current status and the road ahead. Geophysics, 67(6), 20122041.Google Scholar
Dutta, N. C. (Ed.) (1987). Geopressure (Vol. 7). Tulsa, Oklahoma: Society of Exploration Geophysicists.Google Scholar
Eaton, B. A. (1969). Fracture gradient prediction and its application in oil field operations. Journal of Petroleum Technology, 21(10), 13531360.Google Scholar
Eaton, B. A. (1975). The Equation for Geopressure Prediction from Well Logs. Paper presented at the Fall Meeting of the Society of Petroleum Engineers of AIME.Google Scholar
Economides, M. J., Nolte, K. G., & Ahmed, U. (Eds.). (1989). Reservoir stimulation (2nd ed.). Englewood Cliffs, New Jersey: Prentice Hall.Google Scholar
Ellis, S., Ghisetti, F., Barnes, P. M., Boulton, C., Fagereng, Å., & Buiter, S. (2019). The contemporary force balance in a wide accretionary wedge: numerical models of the southcentral Hikurangi margin of New Zealand. Geophysical Journal International, 219(2), 776795.Google Scholar
Ellsworth, W. L. (2013). Injection-induced earthquakes. Science, 341(6142), 1225942.Google Scholar
England, A. H., & Green, A. E. (1963). Some two-dimensional punch and crack problems in classical elasticity. Mathematical Proceedings of the Cambridge Philosophical Society, 59(02), 489500.Google Scholar
England, W. A., Mackenzie, A. S., Mann, D. M., & Quigley, T. M. (1987). The movement and entrapment of petroleum fluids in the subsurface. Journal of the Geological Society, 144(2), 327347.Google Scholar
Espinoza, D. N., & Santamarina, J. C. (2017). CO2 breakthrough—Caprock sealing efficiency and integrity for carbon geological storage. International Journal of Greenhouse Gas Control, 66, 218229.Google Scholar
Evans, K. F. (1989). Appalachian Stress Study: 3. Regional Scale Stress Variations and Their Relation to Structure and Contemporary Tectonics. Journal of Geophysical Research:Solid Earth and Planets, 94(B12), 1761917645.Google Scholar
Fang, Y., Flemings, P. B., Daigle, H., Phillips, S. C., Meazell, P. K., & You, K. (2020). Petrophysical properties of the Green Canyon block 955 hydrate reservoir inferred from reconstituted sediments: Implications for hydrate formation and production. AAPG Bulletin, 104(9), 1997–2028.CrossRefGoogle Scholar
Fauria, K. E., & Rempel, A. W. (2011). Gas invasion into water-saturated, unconsolidated porous media: Implications for gas hydrate reservoirs. Earth and Planetary Science Letters, 312(1–2), 188193.Google Scholar
Fertl, W. H. (1976). Abnormal Formation Pressures: Implications to Exploration, Drilling, and Production of Oil and Gas Resources. Amsterdam: Elsevier Scientific Publishing Company.Google Scholar
Finkbeiner, T., Zoback, M., Flemings, P., & Stump, B. (2001). Stress, Pore Pressure, and Dynamically Constrained Hydrocarbon Columns in the South Eugene Island 330 Field, Northern Gulf of Mexico. AAPG Bulletin, 85(6), 10071031.Google Scholar
Flemings, P. B., Comisky, J. T., Liu, X., & Lupa, J. A. (2001). Stress-Controlled Porosity in Overpressured Sands at Bullwinkle (GC65), Deepwater Gulf of Mexico. Paper presented at the Offshore Technology Conference, Houston, Texas.Google Scholar
Flemings, P. B., Long, H., Dugan, B., Germaine, J., John, C. M., Behrmann, J. H., & Sawyer, D. (2008). Erratum to “Pore pressure penetrometers document high overpressure near the seafloor where multiple submarine landslides have occurred on the continental slope, offshore Louisiana, Gulf of Mexico” [Earth and Planetary Science Letters 269/3–4 (2008) 309–32]. Earth and Planetary Science Letters, 274(1–2), 269283.Google Scholar
Flemings, P. B., & Lupa, J. A. (2004). Pressure prediction in the Bullwinkle Basin through petrophysics and flow modeling (Green Canyon 65, Gulf of Mexico). Marine and Petroleum Geology, 21(10), 13111322.Google Scholar
Flemings, P. B., & Saffer, D. M. (2018). Pressure and Stress Prediction in the Nankai Accretionary Prism: A Critical State Soil Mechanics Porosity-Based Approach. Journal of Geophysical Research: Solid Earth, 123(2), 10891115.Google Scholar
Flemings, P. B., Stump, B. B., Finkbeiner, T., & Zoback, M. (2002). Flow focusing in overpressured sandstones: theory, observations, and applications. American Journal of Science, 302(10), 827855.Google Scholar
Fowler, A. C., & Yang, X. (1999). Pressure solution and viscous compaction in sedimentary basins. Journal of Geophysical Research: Solid Earth, 104(B6), 1298912997.Google Scholar
Fredrich, J. T., Fossum, A. F., & Hickman, R. J. (2007). Mineralogy of deepwater Gulf of Mexico salt formations and implications for constitutive behavior. Journal of Petroleum Science and Engineering, 57(3–4), 354374.Google Scholar
Gaarenstroom, L., Tromp, R. A. J., de Jong, M. C., & Brandenburg, A. M. (1993). Overpressures in the Central North Sea: Implications for Trap Integrity and Drilling Safety. Paper presented at the Petroleum Geology of Northwest Europe: 4th Conference, London, UK.Google Scholar
Gao, B. (2018). Stress, Porosity, and Pore Pressure in Fold-and-Thrust Belt Systems. Doctoral thesis, The University of Texas at Austin. Retrieved from https://hdl.handle.net/2152/73911Google Scholar
Gao, B., & Flemings, P. B. (2017). Pore pressure within dipping reservoirs in overpressured basins. Marine and Petroleum Geology, 80, 94111.Google Scholar
Gardner, G. H. F., Gardner, L. W., & Gregory, A. R. (1974). Formation velocity and density; the diagnostic basics for stratigraphic traps. Geophysics, 39(6), 770780.Google Scholar
Gera, F. (1972). Review of Salt Tectonics in Relation to the Disposal of Radioactive Wastes in Salt Formations. Geological Society of America Bulletin, 83(12), 35513574.Google Scholar
Gibson, R. E. (1958). The Progress of Consolidation in a Clay Layer Increasing in Thickness with Time. Geotechnique, 8(4), 171182.Google Scholar
Gordon, D. S., & Flemings, P. B. (1998). Generation of overpressure and compaction-driven fluid flow in a Plio-Pleistocene growth-faulted basin, Eugene Island 330, offshore Louisiana. Basin Research, 10(2), 177196.Google Scholar
Gradmann, S., & Beaumont, C. (2012). Coupled fluid flow and sediment deformation in margin-scale salt-tectonic systems: 2. Layered sediment models and application to the northwestern Gulf of Mexico. Tectonics, 31(4).Google Scholar
Gradmann, S., Beaumont, C., & Ings, S. J. (2012). Coupled fluid flow and sediment deformation in margin-scale salt-tectonic systems: 1. Development and application of simple, single-lithology models. Tectonics, 31(4).Google Scholar
Green, D. H., & Wang, H. F. (1986). Fluid pressure response to undrained compression in saturated sedimentary rock. Geophysics, 51(4), 948956.Google Scholar
Growcock, F. B., Kaageson-Loe, N., Friedheim, J., Sanders, M. W., & Bruton, J. (2009). Wellbore Stability Stabilization and Strengthening. Paper presented at the Offshore Mediterranean Conference (OMC) 2009, Ravenna, Italy.Google Scholar
Gutierrez, M. A., Braunsdor, N. R., & Couzens, B. A. (2006). Calibration and ranking of pore-pressure prediction models. The Leading Edge, 25(12), 15161523.Google Scholar
Hantschel, T., & Kauerauf, A. (2009). Fundamentals of Basin and Petroleum Systems Modeling (1 ed.): Springer-Verlag Berlin Heidelberg.Google Scholar
Harrison, W. J., & Summa, L. L. (1991). Paleohydrology of the Gulf of Mexico basin. American Journal of Science, 291(2), 109176.Google Scholar
Hart, B. S., Flemings, P. B., & Deshpande, A. (1995). Porosity and Pressure – Role of Compaction Disequilibrium in the Development of Geopressures in a Gulf-Coast Pleistocene Basin. Geology, 23(1), 4548.Google Scholar
Hauser, M. R., Couzens-Schultz, B. A., & Chan, A. W. (2014). Estimating the influence of stress state on compaction behavior. Geophysics, 79(6), D389D398.Google Scholar
Hauser, M. R., Petitclerc, T., Braunsdorf, N. R., & Winker, C. D. (2013). Pressure prediction implications of a Miocene pressure regression. The Leading Edge, 32(1), 100109.Google Scholar
Heidari, M., Nikolinakou, M. A., & Flemings, P. B. (2018). Coupling geomechanical modeling with seismic pressure prediction. Geophysics, 83(5), B253B267.Google Scholar
Hickman, S. H., Hsieh, P. A., Mooney, W. D., Enomoto, C. B., Nelson, P. H., Mayer, L. A., et al. (2012). Scientific basis for safely shutting in the Macondo Well after the April 20, 2010 Deepwater Horizon blowout. Proceedings of the National Academy of Sciences of the United States of America, 109(50), 2026820273.Google Scholar
Holman, W. E., & Robertson, S. S. (1994). Field Development, Depositional Model, and Production Performance of the Turbiditic “J” Sands at Prospect Bullwinkle, Green Canyon 65 Field, Outer-Shelf Gulf of Mexico. Paper presented at the GCSSEPM Foundation 15th Annual Research Conference, Submarine Fans and Turbidite Systems, Houston, Texas.Google Scholar
Hubbert, M. K. (1953). Entrapment of Petroleum under Hydrodynamic Conditions. AAPG Bulletin, 37(8), 19542026.Google Scholar
Hubbert, M. K., & Rubey, W. W. (1959). Role of Fluid Pressure in Mechanics of Overthrust Faulting Part I. Mechanics of Fluid-Filled Porous Solids and Its Application to Overthrust Faulting. Geological Society of America Bulletin, 70(2), 115166.Google Scholar
Hubbert, M. K., & Willis, D. G. (1957). Mechanics of Hydraulic Fracturing. Transactions of the AIME, 210(01), 153168.Google Scholar
Hudec, M. R., Jackson, M. P. A., & Schultz-Ela, D. D. (2009). The paradox of minibasin subsidence into salt: Clues to the evolution of crustal basins. Geological Society of America Bulletin, 121(1–2), 201221.Google Scholar
Huffman, A. R., & Bowers, G. L. (Eds.). (2001). Pressure Regimes in Sedimentary Basins and Their Prediction (Vol. 76). Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Iliffe, J. E., Robertson, A. G., Wynn, G. H. F., Pead, S. D. M., & Cameron, N. (1999). The importance of fluid pressures and migration to the hydrocarbon prospectivity of the Faeroe-Shetland White Zone. In Fleet, A. J. & Boldy, S. A. R. (Eds.), Petroleum Geology of Northwest Europe: Proceedings of the 5th Conference (pp. 601611). London: The Geological Society.Google Scholar
Ingebritsen, S. E., Sanford, W. E., & Neuzil, C. E. (2006). Groundwater in Geologic Processes (2nd ed.): Cambridge University Press.Google Scholar
Issler, D. R. (1992). A new approach to shale compaction and stratigraphic restoration, Beaufort-Mackenzie Basin and Mackenzie Corridor, northern Canada. AAPG Bulletin, 76(8), 11701189.Google Scholar
Jaky, J. (1944). The Coefficient of Earth Pressure at Rest. Journal of the Society of Hungarian Architects and Engineers, 355358.Google Scholar
Karig, D. E., & Ask, M. V. S. (2003). Geological perspectives on consolidation of clay-rich marine sediments. Journal of Geophysical Research: Solid Earth, 108(B4), 2197.Google Scholar
Katahara, K. (2006). Overpressure and Shale Properties: Stress Unloading Or Smectite-illite Transformation? In SEG Technical Program Expanded Abstracts 2006 (pp. 1520–1524). Society of Exploration Geophysicists.Google Scholar
Katz, A. J., & Thompson, A. H. (1986). Quantitative prediction of permeability in porous rock. Phys Rev B Condens Matter, 34(11), 81798181.Google Scholar
Katz, A. J., & Thompson, A. H. (1987). Prediction of rock electrical conductivity from mercury injection measurements. Journal of Geophysical Research: Solid Earth, 92(B1), 599607.Google Scholar
Kirsch, G. (1898). Die Theorie der Elastizität und die Bedürfnisse der Festigkeitslehre. Z. Ver. dtsch. Ing, 42, 707.Google Scholar
Kunze, K. R., & Steiger, R. P. (1992). Accurate In-Situ Stress Measurements During Drilling Operations. Paper presented at the SPE Annual Technical Conference and Exhibition 4–7 October, Washington, DC.Google Scholar
Kvalstad, T. J., Andresen, L., Forsberg, C. F., Berg, K., Bryn, P., & Wangen, M. (2005). The Storegga slide: evaluation of triggering sources and slide mechanics. Marine and Petroleum Geology, 22(1–2), 245256.Google Scholar
Lahann, R. (2002). Impact of Smectite Diagenesis on Compaction Modeling and Compaction Equilibrium. In A. R. Huffman & G. L. Bowers (Eds.), Pressure regimes in sedimentary basins and their prediction (pp. 6172).Google Scholar
Lahann, R. W., & Swarbrick, R. E. (2011). Overpressure generation by load transfer following shale framework weakening due to smectite diagenesis. Geofluids, 11(4), 362375.Google Scholar
Lamb, H. (1932). Hydrodynamics (6th ed.). Cambridge: The University Press.Google Scholar
Lambe, T. W., & Whitman, R. V. (1979). Soil Mechanics: SI Version. New York: John Wiley & Sons.Google Scholar
Lerche, I., & Petersen, K. (1995). Salt and Sediment Dynamics: Boca Raton, Florida: CRC Press.Google Scholar
Li, B., & Wong, R. C. K. (2016). Quantifying structural states of soft mudrocks. Journal of Geophysical Research: Solid Earth, 121(5), 33243347.Google Scholar
Liu, X. (2003). Multiphase Centroid Model. Paper presented at the PSU GeoFluids Consortium Annual Meeting (2003–12), Santa Cruz, California.Google Scholar
Long, H., Flemings, P. B., Germaine, J. T., & Saffer, D. M. (2011). Consolidation and overpressure near the seafloor in the Ursa Basin, Deepwater Gulf of Mexico. Earth and Planetary Science Letters, 305(1–2), 1120.Google Scholar
Lopez, J. L., Rappold, P. M., Ugueto, G. A., Wieseneck, J. B., & Vu, C. K. (2004). Integrated shared earth model: 3D pore-pressure prediction and uncertainty analysis. The Leading Edge, 23(1), 5259.Google Scholar
Losh, S., Eglinton, L. B., Schoell, M., & Wood, J. R. (1999). Vertical and lateral fluid flow related to a large growth fault, South Eugene Island Block 330 Field, offshore Louisiana. AAPG Bulletin, 83(2), 244276.Google Scholar
Lupa, J., Flemings, P. B., & Tennant, S. (2002). Pressure and trap integrity in the deepwater Gulf of Mexico. The Leading Edge, 21(2), 184187.Google Scholar
Ma, X., & Zoback, M. D. (2017). Laboratory experiments simulating poroelastic stress changes associated with depletion and injection in low-porosity sedimentary rocks. Journal of Geophysical Research: Solid Earth, 122(4), 24782503.Google Scholar
Mann, D. M., & Mackenzie, A. S. (1990). Prediction of pore fluid pressures in sedimentary basins. Marine and Petroleum Geology, 7(1), 5565.Google Scholar
Matthews, W. R., & Kelly, J. (1967). How to Predict Formation Pressure and Fracture Gradient. Oil and Gas Journal, 20(February), 92106.Google Scholar
Maury, V., & Idelovici, J. L. (1995). Safe Drilling of HP/HT Wells, The Role of the Thermal Regime in Loss and Gain Phenomenon. Paper presented at the SPE/IADC Drilling Conference, Amsterdam, Netherlands.Google Scholar
McKenzie, D. (1984). The Generation and Compaction of Partially Molten Rock. Journal of Petrology, 25(3), 713765.Google Scholar
Merrell, M. P. (2012). Pressure and Stress at Mad Dog Field, Gulf of Mexico. Master’s thesis, The University of Texas at Austin. Retrieved from http://hdl.handle.net/2152/20068Google Scholar
Merrell, M. P., Flemings, P. B., & Bowers, G. L. (2014). Subsalt pressure prediction in the Miocene Mad Dog field, Gulf of Mexico. AAPG Bulletin, 98(2), 315340.Google Scholar
Mesri, G., & Castro, A. (1987). Cα/Cc Concept and K0 During Secondary Compression. Journal of Geotechnical Engineering, 113(3), 230247.Google Scholar
Mesri, G., & Godlewski, P. M. (1977). Time- and stress-compressibility interrelationship. Journal of the Geotechnical Engineering Division, 103(GT5), 417430.Google Scholar
Mesri, G., & Hayat, T. M. (1993). The coefficient of earth pressure at rest. Canadian Geotechnical Journal, 30(4), 647666.Google Scholar
Mikada, H., Becker, K., Moore, J. C., Klaus, A., & Leg 196 Shipboard Scientific Party. (2002). ODP Leg 196: Logging-While-Drilling and Advanced CORKs at the Nankai Trough Accretionary Prism. JOIDES Journal, 28(2), 812.Google Scholar
Milliken, K. L., & Hayman, N. W. (2020). Mudrock Components and the Genesis of Bulk Rock Properties: Review of Current Advances and Challenges. In Dewers, T., Heath, J., & Sánchez., M. (Eds.), Shale: Subsurface Science and Engineering (First Edition ed., Vol. 245, pp. 125). NJ and Washington, D.C.: American Geophysical Union and John Wiley & Sons, Inc.Google Scholar
Mitchell, J. K., & Soga, K. (2005). Fundamentals of Soil Behavior (3rd ed.). Hoboken, New Jersey: John Wiley & Sons.Google Scholar
Mondol, N. H., Bjørlykke, K., Jahren, J., & Høeg, K. (2007). Experimental mechanical compaction of clay mineral aggregates—Changes in physical properties of mudstones during burial. Marine and Petroleum Geology, 24(5), 289311.Google Scholar
Morency, C., Huismans, R. S., Beaumont, C., & Fullsack, P. (2007). A numerical model for coupled fluid flow and matrix deformation with applications to disequilibrium compaction and delta stability. Journal of Geophysical Research, 112(B10).Google Scholar
Morgan, J. K., & Ask, M. V. S. (2004). Consolidation state and strength of underthrust sediments and evolution of the décollement at the Nankai accretionary margin: Results of uniaxial reconsolidation experiments. Journal of Geophysical Research: Solid Earth, 109(B3), B03102.Google Scholar
Morrow, N. R. (1990). Wettability and Its Effect on Oil Recovery. Journal of Petroleum Technology, 42(12), 14761484.Google Scholar
Mouchet, J. P., Mitchell, A., & Boussens, C. d. r. d. (1989). Abnormal Pressures While Drilling: Origins, Prediction, Detection, Evaluation (Vol. 2). Boussens, France: Technip Editions.Google Scholar
Munson, D. E., & Dawson, P. R. (1979). Constitutive model for the low temperature creep of salt (with application to WIPP) (SAND–79–1853). IAEA: Sandia Labs., Albuquerque, NM (USA). Retrieved from: https://inis.iaea.org/search/search.aspx?orig_q=RN:11521637Google Scholar
Nance, R. D., Rovick, J. E., & Wilcox, R. E. (1979). Lithology of the Vacherie dome core. (Topical Report E511-02500-5). Prepared for U.S. Department of Energy. Baton Rouge, Louisiana: Institute for Environmental Studies, Louisiana State University.Google Scholar
Naruk, S. J., Solum, J. G., Brandenburg, J. P., Origo, P., & Wolf, D. E. (2019). Effective stress constraints on vertical flow in fault zones: Learnings from natural CO2 reservoirs. AAPG Bulletin, 103(8), 19792008.Google Scholar
Neuzil, C. E. (1995). Abnormal pressures as hydrodynamic phenomena. American Journal of Science, 295(6), 742786.Google Scholar
Nihei, K. T., Nakagawa, S., Reverdy, F., Myer, L. R., Duranti, L., & Ball, G. (2011). Phased array compaction cell for measurement of the transversely isotropic elastic properties of compacting sediments. Geophysics, 76(3), WA113-WA123.Google Scholar
Nikolinakou, M. A., Gao, B., Flemings, P. B., & Saffer, D. M. (in review). Dramatic fluid overpressuring and megathrust weakening initiate at the trench. Journal of Geophysical Research.Google Scholar
Nikolinakou, M. A., Heidari, M., Flemings, P. B., & Hudec, M. R. (2018). Geomechanical modeling of pore pressure in evolving salt systems. Marine and Petroleum Geology, 93, 272286.Google Scholar
Nur, A., & Byerlee, J. D. (1971). An exact effective stress law for elastic deformation of rock with fluids. Journal of Geophysical Research, 76(26), 6419–6419.Google Scholar
O’Connell, J. K., Kohli, M., & Amos, S. (1993). Bullwinkle: A unique 3-D experiment. Geophysics, 58(1), 167176.Google Scholar
Obradors-Prats, J., Rouainia, M., Aplin, A. C., & Crook, A. J. L. (2017). Hydromechanical Modeling of Stress, Pore Pressure, and Porosity Evolution in Fold-and-Thrust Belt Systems. Journal of Geophysical Research: Solid Earth, 122(11), 93839403.Google Scholar
Osborne, M. J., & Swarbrick, R. E. (1997). Mechanisms for generating overpressure in sedimentary basins: a reevaluation. AAPG Bulletin, 81(6), 10231041.Google Scholar
Ostermeier, R. M., Pelletier, J. H., Winker, C. D., & Nicholson, J. W. (2001). Trends in Shallow Sediment Pore Pressures – Deepwater Gulf of Mexico. Paper presented at the SPE/IADC Drilling Conference, Amsterdam, Netherlands.Google Scholar
Ostermeier, R. M., Pelletier, J. H., Winker, C. D., Nicholson, J. W., Rambow, F. H., & Cowan, K. M. (2002). Dealing with shallow-water flow in the deepwater Gulf of Mexico. The Leading Edge, 21(7), 660668.Google Scholar
Pennebaker, E. S. (1968). An Engineering Interpretation of Seismic Data. Paper presented at the Fall Meeting of the Society of Petroleum Engineers of AIME, Houston, Texas.Google Scholar
Pepin, G., Gonzalez, M., Bloys, J. B., Lofton, J., Schmidt, J., Naquin, C., & Ellis, S. (2004). Effect of Drilling Fluid Temperature on Fracture Gradient: Field Measurements and Model Predictions. Paper presented at the Gulf Rocks 2004, the 6th North America Rock Mechanics Symposium (NARMS): Rock Mechanics Across Borders and Disciplines, Houston, Texas.Google Scholar
Perić, D., & Crook, A. J. L. (2004). Computational strategies for predictive geology with reference to salt tectonics. Computer Methods in Applied Mechanics and Engineering, 193(48–51), 51955222.Google Scholar
Petley, D. N. (1999). Failure envelopes of mudrocks at high confining pressures. Geological Society, London, Special Publications, 158(1), 6171.Google Scholar
Phillips, O. M. (1991). Flow and reactions in permeable rocks. New York: Cambridge University Press.Google Scholar
Pinkston, F. W. M. (2017). Pore Pressure and Stress at the Macondo Well, Mississippi Canyon, Gulf of Mexico. Master’s thesis, The University of Texas at Austin.Google Scholar
Pinkston, F. W. M., & Flemings, P. B. (2019). Overpressure at the Macondo Well and its impact on the Deepwater Horizon blowout. Sci Rep, 9(1), 7047.Google Scholar
Pittman, E. D. (1992). Relationship of Porosity and Permeability to Various Parameters Derived from Mercury Injection-Capillary Pressure Curves for Sandstone. AAPG Bulletin, 76(2), 191198.Google Scholar
Rafalowski, J. W., Regel, B. W., Jordan, D. L., & Lucidi, D. O. (1994). Green Canyon Block 205 lithofacies, seismic facies, and reservoir architecture. In Weimer, P. & Davis, T. L. (Eds.), AAPG Studies in Geology (pp. 133142). Tulsa, OK: AAPG/SEG.Google Scholar
Raleigh, C. B., Healy, J. H., & Bredehoeft, J. D. (1976). An experiment in earthquake control at Rangely, Colorado. Science, 191(4233), 12301237.Google Scholar
Reece, J. S. (2013). Seal Capacity in Mudrocks: the Impact of Silt Fraction. Paper presented at the UT GeoFluids Consortium Annual Meeting (4.21), Austin, TX.Google Scholar
Reece, J. S., Flemings, P. B., & Germaine, J. T. (2013). Data Report: Permeability, compressibility, and microstructure of resedimented mudstone from IODP Expedition 322, Site C0011. In Saito, S., Underwood, M. B., Kubo, Y., & the Expedition 322 Scientists, Proc. IODP, 322: Tokyo (Integrated Ocean Drilling Program Management International, Inc.).Google Scholar
Reilly, M. J. (2008). Deep Pore Pressures and Seafloor Venting in the Auger Basin, Gulf of Mexico. Master’s thesis, Pennsylvania State University. Retrieved from https://etda.libraries.psu.edu/catalog/9148Google Scholar
Reilly, M. J., & Flemings, P. B. (2010). Deep pore pressures and seafloor venting in the Auger Basin, Gulf of Mexico. Basin Research, 22(4), 380397.Google Scholar
Rice, J. R., & Cleary, M. P. (1976). Some basic stress diffusion solutions for fluid-saturated elastic porous media with compressible constituents. Reviews of Geophysics, 14(2), 227241.Google Scholar
Rockfield. (2017). ELFEN Forward Modeling User Manual: Rockfield Software Limited.Google Scholar
Rogers, G., & Dragert, H. (2003). Episodic tremor and slip on the Cascadia subduction zone: the chatter of silent slip. Science, 300(5627), 19421943.Google Scholar
Rohleder, S. A., Sanders, W. W., Williamson, R. N., Faul, G. L., & Dooley, L. B. (2003). Challenges of drilling an ultra-deep well in deepwater – Spa prospect. Paper presented at the SPE/IADC Drilling Conference, Amsterdam, Netherlands.Google Scholar
Roscoe, K. H., & Burland, J. B. (1968). On the generalized stress-strain behaviour of “wet” clay. In J. Heyman & F. A. Leckie (Eds.), Engineering Plasticity (pp. 535609). Cambridge, England: Cambridge University Press.Google Scholar
Roscoe, K. H., Schofield, A. N., & Wroth, C. P. (1958). On the Yielding of Soils. Geotechnique, 8(1), 2253.Google Scholar
Rubey, W. W., & Hubbert, M. K. (1959). Role of Fluid Pressure in Mechanics of Overthrust Faulting Part II. Overthrust Belt in Geosynclinal Area of Western Wyoming in Light of Fluid-Pressure Hypothesis. Geological Society of America Bulletin, 70(2), 167205.Google Scholar
Saffer, D. M. (2003). Pore pressure development and progressive dewatering in underthrust sediments at the Costa Rican subduction margin: Comparison with northern Barbados and Nankai. Journal of Geophysical Research: Solid Earth, 108(B5), 2261.Google Scholar
Saffer, D. M. (2007). Pore Pressure within Underthrust Sediment in Subduction Zones. In Dixon, T. H. & Moore, J. C. (Eds.), The Seismogenic Zone of Subduction Thrust Faults (pp. 171–209): Columbia University Press.Google Scholar
Saffer, D. M., & Tobin, H. J. (2011). Hydrogeology and Mechanics of Subduction Zone Forearcs: Fluid Flow and Pore Pressure. Annual Review of Earth and Planetary Sciences, 39(1), 157186.Google Scholar
Sanford, J., Woomer, J., Miller, J., & Russell, C. (2006). The K2 Project: A Drilling Engineer’s Perspective. Paper presented at the Offshore Technology Conference, Houston, Texas.Google Scholar
Santagata, M. C., & Kang, Y. I. (2007). Effects of geologic time on the initial stiffness of clays. Engineering Geology, 89(1–2), 98111.Google Scholar
Sassen, R., Milkov, A. V., Ozgul, E., Roberts, H. H., Hunt, J. L., Beeunas, M. A., et al. (2003). Gas venting and subsurface charge in the Green Canyon area, Gulf of Mexico continental slope: evidence of a deep bacterial methane source? Organic Geochemistry, 34(10), 14551464.Google Scholar
Sawyer, A. H., Flemings, P., Elsworth, D., & Kinoshita, M. (2008). Response of submarine hydrologic monitoring instruments to formation pressure changes: Theory and application to Nankai advanced CORKs. Journal of Geophysical Research, 113(B1).Google Scholar
Sayers, C. M., Johnson, G. M., & Denyer, G. (2002). Predrill pore-pressure prediction using seismic data. Geophysics, 67(4), 12861292.Google Scholar
Sayers, C. M., Woodward, M. J., & Bartman, R. C. (2002). Seismic pore-pressure prediction using reflection tomography and 4-C seismic data. The Leading Edge, 21(2), 188192.Google Scholar
Schneider, F., Potdevin, J. L., Wolf, S., & Faille, I. (1996). Mechanical and chemical compaction model for sedimentary basin simulators. Tectonophysics, 263(1–4), 307317.Google Scholar
Schneider, J. (2011). Compression and Permeability Behavior of Natural Mudstones. Doctoral thesis, The University of Texas at Austin.Google Scholar
Schneider, J., Flemings, P. B., Day-Stirrat, R. J., & Germaine, J. T. (2011). Insights into pore-scale controls on mudstone permeability through resedimentation experiments. Geology, 39(11), 10111014.Google Scholar
Schowalter, T. T. (1979). Mechanics of secondary hydrocarbon migration and entrapment. AAPG Bulletin, 63(5), 723760.Google Scholar
Scott, A., Hurst, A., & Vigorito, M. (2013). Outcrop-based reservoir characterization of a kilometer-scale sand-injectite complex. AAPG Bulletin, 97(2), 309343.Google Scholar
Screaton, E., Saffer, D., Henry, P., & Hunze, S. (2002). Porosity loss within the underthrust sediments of the Nankai accretionary complex: Implications for overpressures. Geology, 30(1), 1922.Google Scholar
Seldon, B., & Flemings, P. B. (2005). Reservoir pressure and seafloor venting: Predicting trap integrity in a Gulf of Mexico deepwater turbidite minibasin. AAPG Bulletin, 89(2), 193209.Google Scholar
Sheahan, T. C. (1991). An Experimental Study of the Time-Dependent Undrained Shear Behavior of Resedimented Clay Using Automated Stress Path Triaxial Equipment. Doctoral thesis, Massachusetts Institute of Technology.Google Scholar
Shepard, F. P. (1954). Nomenclature based on sand-silt-clay ratios. Journal of Sedimentary Petrology, Vol. 24, 151158.Google Scholar
Shumaker, N., Haymond, D., & Martin, J. (2014). Kinematic linkage between minibasin welds and extreme overpressure in the deepwater Gulf of Mexico. Interpretation, 2(1), SB69-SB77.Google Scholar
Shumaker, N., Lindsay, R., & Ogilvie, J. (2007). Depth-calibrating seismic data in the presence of allochthonous salt. The Leading Edge, 26(11), 14421453.Google Scholar
Shumaker, N., & Vernik, L. (2009). The use of “verticalized” stacking velocities to constrain shale properties in west Africa. The Leading Edge, 28(2), 184188.Google Scholar
Skempton, A. W. (1954). The Pore-Pressure Coefficients A and B. Geotechnique, 4(4), 143147.Google Scholar
Skempton, A. W. (1969). The consolidation of clays by gravitational compaction. Quarterly Journal of the Geological Society, 125(1–4), 373411.Google Scholar
Smith, A. J., Flemings, P. B., & Fulton, P. M. (2014). Hydrocarbon flux from natural deepwater Gulf of Mexico vents. Earth and Planetary Science Letters, 395(0), 241253.Google Scholar
Smith, D. A. (1966). Theoretical Considerations of Sealing and Non-Sealing Faults. AAPG Bulletin, 50(2), 11.Google Scholar
Sone, H., & Zoback, M. D. (2014). Time-dependent deformation of shale gas reservoir rocks and its long-term effect on the in situ state of stress. International Journal of Rock Mechanics and Mining Sciences, 69, 120132.Google Scholar
Stigall, J., & Dugan, B. (2010). Overpressure and earthquake initiated slope failure in the Ursa region, northern Gulf of Mexico. Journal of Geophysical Research: Solid Earth, 115, 11.Google Scholar
Stoffa, P. L., Wood, W. T., Shipley, T. H., Moore, G. F., Nishiyama, E., Botelho, M. A. B., et al. (1992). Deepwater high-resolution expanding spread and split spread seismic profiles in the Nankai Trough. Journal of Geophysical Research: Solid Earth, 97(B2), 16871713.Google Scholar
Stork, C. (1992). Reflection tomography in the postmigrated domain. Geophysics, 57(5), 680692.Google Scholar
Stump, B. B., & Flemings, P. B. (2001). Consolidation State, Permeability, and Stress Ratio as Determined from Uniaxial Strain Experiments on Mudstone Samples from the Eugene Island 330 Area, Offshore Louisiana. In Pressure regimes in sedimentary basins and their prediction (Vol. 76, pp. 131144): AAPG.Google Scholar
Su, K., & Onaisi, A. (2019). Dealing with the Uncertainty in the Prediction of Fracture Gradient. Paper presented at the Second EAGE Workshop on Pore Pressure Prediction, Amsterdam, Netherlands.Google Scholar
Suarez-Rivera, R., & Fjær, E. (2013). Evaluating the Poroelastic Effect on Anisotropic, Organic-Rich, Mudstone Systems. Rock Mechanics and Rock Engineering, 46(3), 569580.Google Scholar
Swanston, A. M., Flemings, P. B., Comisky, J. T., & Best, K. D. (2003). Time-lapse imaging at Bullwinkle Field, Green Canyon 65, offshore Gulf of Mexico. Geophysics, 68(5), 14701484.Google Scholar
Taylor, G. I., & Quinney, H. (1932). The Plastic Distortion of Metals. Philosophical Transactions of the Royal Society of London. Series A, Containing Papers of a Mathematical or Physical Character, 230(681–693), 323362.Google Scholar
Terzaghi, K. (1923). Die Berechnung der Durchlassigkeitzeriffer des Tones aus dem Verlauf der Hydrodymanischen Spannungserscheinungen [The computation of permeability of clays from the progress of hydrodynamic strain]. Akademie der Wissenschaften in Wien, Sitzungsberichte, Mathematisch-Naturwissenschaftliche Klasse, Part IIa, 132, 125138.Google Scholar
Terzaghi, K. (1943). Theoretical Soil Mechanics. London: Chapman and Hall.Google Scholar
Terzaghi, K. (1950). Mechanism of landslides. In Application of Geology to Engineering Practice: Berkey Volume (pp. 83123). New York: Geological Society of America.Google Scholar
Thomeer, J. H. M. (1960). Introduction of a Pore Geometrical Factor Defined by the Capillary Pressure Curve. Journal of Petroleum Technology, 12(03), 7377.Google Scholar
Thomsen, L. (1986). Weak elastic anisotropy. Geophysics, 51(10), 1954 –1966.Google Scholar
Traugott, M. O., & Heppard, P. D. (1994). Prediction of pore pressure before and after drilling – taking the risk out of drilling overpressured prospects. Paper presented at the American Association of Petroleum Geologists Hedberg Research Conference, Denver, Colorado.Google Scholar
Tréhu, A. M., Flemings, P. B., Bangs, N. L., Chevallier, J., Gràcia, E., Johnson, J. E., et al. (2004). Feeding methane vents and gas hydrate deposits at south Hydrate Ridge. Geophysical Research Letters, 31(23), L23310.Google Scholar
Tsuji, T., Tokuyama, H., Costa Pisani, P., & Moore, G. (2008). Effective stress and pore pressure in the Nankai accretionary prism off the Muroto Peninsula, southwestern Japan. Journal of Geophysical Research: Solid Earth, 113(B11).Google Scholar
Velde, B. (1996). Compaction trends of clay-rich deep sea sediments. Marine Geology, 133(3–4), 193201.Google Scholar
Vétel, W., & Cartwright, J. (2010). Emplacement mechanics of sandstone intrusions: insights from the Panoche Giant Injection Complex, California. Basin Research, 22(5), 783807.Google Scholar
Vigorito, M., & Hurst, A. (2010). Regional sand injectite architecture as a record of pore-pressure evolution and sand redistribution in the shallow crust: insights from the Panoche Giant Injection Complex, California. Journal of the Geological Society, 167(5), 889904.Google Scholar
Vigorito, M., Hurst, A., Cartwright, J., & Scott, A. (2008). Regional-scale subsurface sand remobilization: geometry and architecture. Journal of the Geological Society, 165(3), 609612.Google Scholar
Walker, C., Belvedere, P., Petersen, J., Warrior, S., Cunningham, A., Clemenceau, G., et al. (2013). Straining at the Leash: Understanding the Full Potential of the Deep-Water, Subsalt Mad Dog Field, from Appraisal through Early Production. In New Understanding of the Petroleum Systems of Continental Margins of the World: 32nd Annual (Vol. 32, pp. 2564).Google Scholar
Wang, H. (2000). Theory of Linear Poroelasticity: With Applications to Geomechanics and Hydrogeology. Princeton, NJ: Princeton University Press.Google Scholar
Wang, Z. (2002). Seismic anisotropy in sedimentary rocks, part 2: Laboratory data. Geophysics, 67(5), 14231440.Google Scholar
Warpinski, N. R., Branagan, P., & Wilmer, R. (1985). In-Situ Stress Measurements at U.S. DOE’s Multiwell Experiment Site, Mesaverde Group, Rifle, Colorado. Journal of Petroleum Technology, 37(03), 527536.Google Scholar
Watabe, Y., & Leroueil, S. (2015). Modeling and Implementation of the Isotache Concept for Long-Term Consolidation Behavior. International Journal of Geomechanics, 15(5), A4014006.Google Scholar
Watts, N. L. (1987). Theoretical aspects of cap-rock and fault seals for single- and two-phase hydrocarbon columns. Marine and Petroleum Geology, 4(4), 274307.Google Scholar
Weatherl, M. H. (2010). Gulf of Mexico Deepwater Field Development Challenges at Green Canyon 468 Pony. Paper presented at the SPE Deepwater Drilling and Completions Conference, Galveston, Texas.Google Scholar
Wenk, H.-R., Lonardelli, I., Franz, H., Nihei, K., & Nakagawa, S. (2007). Preferred orientation and elastic anisotropy of illite-rich shale. Geophysics, 72(2), E69E75.Google Scholar
White, A. J., Traugott, M. O., & Swarbrick, R. E. (2002). The use of leak-off tests as means of predicting minimum in-situ stress. Petroleum Geoscience, 8(2), 189193.Google Scholar
Wilhelm, R., Franceware, L. B., & Guzman, C. E. (1998). Seismic pressure-prediction method solves problem common in deepwater Gulf of Mexico. Oil and Gas Journal 96(37), 6775.Google Scholar
Williams, K. E., Redhead, R., & Standifird, W. (2008). Geopressure Analysis in the Subsalt Knotty Head Field, Deepwater Gulf of Mexico. Gulf Coast Association of Geological Societies Transactions, 58, 905912.Google Scholar
Winker, C. D., & Stancliffe, R. J. (2007a). Geology of Shallow-Water Flow at Ursa: 1. Setting and Causes. Paper presented at the Offshore Technology Conference, Houston, Texas.Google Scholar
Winker, C. D., & Stancliffe, R. J. (2007b). Geology of Shallow-Water Flow at Ursa: 2. Drilling Principles and Practice. Paper presented at the Offshore Technology Conference, Houston, Texas.Google Scholar
Wood, D. M. (1990). Soil Behaviour and Critical State Soil Mechanics: Cambridge University Press.Google Scholar
Woodward, M., Farmer, P., Nichols, D., & Charles, S. (1998). Automated 3D Tomographic Velocity Analysis of Residual Moveout In Prestack Depth Migrated Common Image Point Gathers. Paper presented at the 1998 SEG Annual Meeting, New Orleans, Louisiana.Google Scholar
Yang, Y., & Aplin, A. C. (1998). Influence of lithology and compaction on the pore size distribution and modelled permeability of some mudstones from the Norwegian margin. Marine and Petroleum Geology, 15(2), 163175.Google Scholar
Yang, Y., & Aplin, A. C. (2004). Definition and practical application of mudstone porosity–effective stress relationships. Petroleum Geoscience, 10(2), 153162.Google Scholar
Yardley, G. S., & Swarbrick, R. E. (2000). Lateral transfer; a source of additional overpressure? Marine and Petroleum Geology, 17(4), 523537.Google Scholar
Yarger, H., DeKay, L., & Hensel, E. G. (2001). Salt canopy modeling with gravity in deepwater Gulf of Mexico. SEG Expanded Abstracts, 20, 4.Google Scholar
Yilmaz, O. (1987). Seismic Data Processing. Tulsa, OK: Society of Exploration Geophysicists.Google Scholar
Yuvancic Strickland, B., Kuhl, E. J., Lee, T. W., Seldon, B. J., Flemings, P. B., & Ertekin, T. (2003). Integration Of Geologic Model And Reservoir Simulation, Popeye Field, Green Canyon 116. Gulf Coast Association of Geological Societies Transactions, 53, 918932.Google Scholar
Zablocki, M. (2019). Lithology Based Prediction of Pressure, Fracture Gradient and Strength. Paper presented at the UT GeoFluids Consortium Annual Meeting (10.03), Austin, TX.Google Scholar
Zhang, J. (2011). Pore pressure prediction from well logs: Methods, modifications, and new approaches. Earth-Science Reviews, 108(1–2), 5063.Google Scholar
Zoback, M. D. (2007). Reservoir Geomechanics (1st ed.). New York, NY: Cambridge University Press.Google Scholar
Zoback, M. D., & Kohli, A. H. (2019). Unconventional Reservoir Geomechanics (1st ed.). New York, NY: Cambridge University Press.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×