Skip to main content Accessibility help
×
Hostname: page-component-76fb5796d-x4r87 Total loading time: 0 Render date: 2024-04-28T00:02:51.283Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  10 December 2009

Mark D. Zoback
Affiliation:
Stanford University, California
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2007

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

Aadnoy, B. S. (1990a). “Inversion technique to determine the In situ stress field from fracturing data.”Journal of Petroleum Science and Engineering, 4, 127–141.CrossRefGoogle Scholar
Aadnoy, B. S. (1990b). “In situ stress direction from borehole fracture traces.”Journal of Petroleum Science and Engineering, 4, 143–153.CrossRefGoogle Scholar
Abé, H., Mura, T.et al. (1976). “Growth rate of a penny-shaped crack in hydraulic fracture of rocks.”J. Geophys. Res, 81, 5335.CrossRefGoogle Scholar
Adachi, T. and Oka, F. (1982). “Constitutive equations for normally consolidated clays based on elasto-viscoplasticity.”Soils and Foundations, 22(4), 57–70.CrossRefGoogle Scholar
Addis, M. A., Cauley, M. B. et al. (2001). Brent in-fill drilling programme: Lost circulation associated with drilling depleted reservoirs, Paper number SPE/IADC 67741. SPE/IADC Drilling Conference, Amsterdam, Netherlands, Society of Petroleum Engineers.CrossRef
Alexander, L. L. and Flemings, P. B. (1995). “Geologic evolution of Plio-Pleistocene salt withdrawl minibasin: Eugene Island block 330, offshore Louisiana.”American Association of Petroleum Geologists Bulletin, 79, 1737–1756.Google Scholar
Alnes, J. R. and Lilburn, R. A. (1998). “Mechanisms for generating overpressure in sedimentary basins: A reevaluation. Discussion.”American Association of Petroleum Geologists Bulletin, 82, 2266–2269.Google Scholar
Amadei, B. and Stephansson, O. (1997). Rock Stress and its Measurement. London, Chapman & Hall.CrossRefGoogle Scholar
Anderson, E. M. (1951). The Dynamics of Faulting and Dyke Formation with Applications to Britain. Edinburgh, Oliver and Boyd.Google Scholar
Anderson, R. N., Flemings, P.et al. (1994). “In situ properties of a major Gulf of Mexico growth fault: Implications for behavior as a hydrocarbon migration pathway.”Oil and Gas Journal, 92(23), 97–104.Google Scholar
Angelier, J. (1979). “Determination of the mean principal directions of stresses for a given fault population.” Tectonophysics, 56, T17–T26.CrossRefGoogle Scholar
Angelier, J. (1984). “Tectonic analysis of fault slip data sets.” Journal of Geophysical Research, 89, 5835–5848.CrossRefGoogle Scholar
Antonellini, M. and Aydin, A. (1994). “Effect of faulting and fluid flow in porous sandstones: Geometry and spatial distribution.” American Association of Petroleum Geologists Bulletin, 79(5), 642–671.Google Scholar
Artyushkov, E. V. (1973). “Stresses in the lithosphere caused by crustal thickness inhomogeneities.” Journal of Geophysical Research, 78, 7675–7708.CrossRefGoogle Scholar
Athy, L. F. (1930). “Density, porosity and compaction of sedimentary rocks.” American Association of Petroleum Geologists Bulletin, 14, 1–24.Google Scholar
Atkinson, B. K., Ed. (1987). Fracture Mechanics of Rock. Academic Press Geology Series. London, Academic Press.Google Scholar
Baria, R., Baumgaerdner, J.et al. (1999). “European HDR research programme at Soultz-sous-Forets (France) 1987–1996.” Geothermics, 28, 655–669.CrossRefGoogle Scholar
Barton, C. A., Tessler, L. et al. (1991). Interactive analysis of borehole televiewer data. In Automated Pattern Analysis in Petroleum Exploration, Palaz, I. and Sengupta, S. K. (eds). New York, Springer Verlag.Google Scholar
Barton, C. A. and Zoback, M. D. (1992). “Self-similar distribution and properties of macroscopic fractures at depth in crystalline rock in the Cajon Pass scientific drill hole.” Journal of Geophysical Research, 97, 5181–5200.CrossRefGoogle Scholar
Barton, C. A. and Zoback, M. D. (1994). “Stress perturbations associated with active faults penetrated by boreholes: Possible evidence for near-complete stress drop and a new technique for stress magnitude measurements.” J. Geophys. Res., 99, 9373–9390.CrossRefGoogle Scholar
Barton, C. A. and Zoback, M. D. (2002). Wellbore Imaging Technologies Applied to Reservoir Geomechanics and Environmental Engineering. Geological Applications of Well Logs. M. Lovell and N. Parkinson (eds). AAPG Methods in Exploration, No. 13, 229–239.
Barton, C. A., Zoback, M. D.et al. (1988). “In situ stress orientation and magnitude at the Fenton Geothermal site, New Mexico, determined from wellbore breakouts.” Geophysical Research Letters, 15(5), 467–470.CrossRefGoogle Scholar
Barton, C. A., Zoback, M. D.et al. (1995). “Fluid flow along potentially active faults in crystalline rock.” Geology, 23, 683–686.2.3.CO;2>CrossRefGoogle Scholar
Baumgärtner, J., Carvalho, J. et al. (1989). Fracturing deviated boreholes: An experimental approach. Rock at Great Depth, Proceedings ISRM-SPE International Symposium, Elf Aquitaine, Pau, A.A.Balkema.
Baumgärtner, J., Rummel, F. et al. (1990). Hydraulic fracturing in situ stress measurements to 3 km depth in the KTB pilot hole VB. A summary of a preliminary data evaluation, in KTB Report 90–6a, 353–400.
Baumgärtner, J. and Zoback, M. D. (1989). “Interpretation of hydraulic fracturing pressure – time records using interactive analysis methods.” International Journal of Rock Mechanics and Mining Sciences & Geomechanical Abstracts, 26, 461–469.CrossRefGoogle Scholar
Bell, J. S. (1989). “Investigating stress regimes in sedimentary basins using information from oil industry wireline logs and drilling records.” Geological Applications of Wireline Logs. Special Publication 48(Geological Society of London), 305–325.Google Scholar
Bell, J. S. and Babcock, E. A. (1986). “The stress regime of the Western Canadian Basin and implications for hydrocarbon production.” Bulletin of Canadian Petroleum Geology, 34, 364–378.Google Scholar
Bell, J. S. and Gough, D. I. (1979). “Northeast-southwest compressive stress in Alberta: Evidence from oil wells.” Earth Planet. Sci. Lett., 45, 475–482.CrossRefGoogle Scholar
Bell, J. S. and Gough, D. I. (1983). The use of borehole breakouts in the study of crustal stress, in Hydraulic fracturing stress measurements. D.C, National Academy Press, Washington.Google Scholar
Berry, F. A. F. (1973). “High fluid potentials in California Coast Ranges and their tectonic significance.” American Association of Petroleum Geologists Bulletin, 57, 1219–1245.Google Scholar
Biot, M. A. (1962). “Mechanics of deformation and acoustic propagation in porous media.” Journal of Acoustic Society of America, 28, 168–191.CrossRefGoogle Scholar
Birch, F. (1961). “Velocity of compressional waves in rocks to 10 kilobars, Part 2.” J. Geophys. Res., 66, 2199–2224.CrossRefGoogle Scholar
Boness, N. and Zoback, M. D. (2004). “Stress-induced seismic velocity anisotropy and physical properties in the SAFOD Pilot hole in Parkfield, CA.” Geophysical Research Letters, 31, L15S17.CrossRefGoogle Scholar
Boness, N. and Zoback, M. D. (2006). “A multi-scale study of the mechanisms controlling shear velocity anisotropy in the San Andreas Fault Observatory at Depth.” Geophysics, 71, F131–F136.CrossRefGoogle Scholar
Bourbie, T., Coussy, O. and Zinszner, B. (1987). Acoustics of porous media. Paris, France, Editions Technip.Google Scholar
Bourgoyne, A. T. Jr, Millheim, K. K.et al. (2003). Applied Drilling Engineering. Richardson, Texas, Society of Petroleum Engineers.Google Scholar
Bowers, G. L. (1994). Pore pressure estimation from velocity data: accounting for overpressure mechanisms besides undercompaction. SPE 27488 Dallas, Texas, Society of Petroleum Engineers, 515–589.CrossRef
Brace, W. F. (1980). “Permeability of crystalline and argillaceous rocks.” Int'l. J. Rock Mech. Min. Sci. and Geomech. Abstr., 17, 241–251.CrossRefGoogle Scholar
Brace, W. F. and Kohlstedt, D. L. (1980). “Limits on lithospheric stress imposed by laboratory experiments.” J. Geophys. Res, 85, 6248–6252.CrossRefGoogle Scholar
Brace, W. F., Paulding, B. W.et al. (1966). “Dilatancy in the fracture of crystalline rocks.” Journal of Geophysical Research, 71(16), 3939–3953.CrossRefGoogle Scholar
Bradley, W. B. (1979). “Failure of Inclined Boreholes.” J. Energy Res. Tech., Trans. ASME, 102, 232.CrossRefGoogle Scholar
Bratli, R. K. and Risnes, R. (1981). “Stability and failure of sand arches.” Soc. of Petroleum Engineers Journal, (April), 236–248.CrossRefGoogle Scholar
Bratton, T., Bornemann, T. et al. (1999). Logging-while-drilling images for geomechanical, geological and petrophysical interpretations. SPWLA 40th Annual Logging Symposium, Oslo, Norway, Society of Professional Well Log Analysts.
Breckels, I. M. and Eekelen, H. A. M. (1981). “Relationship between horizontal stress and depth in sedimentary basins: Paper SPE10336, 56th Annual Fall Technical Conference.” Society of Petroleum Engineers of AIME, San Antonio, Texas, October 5–7, 1981.Google Scholar
Bredehoeft, J. D., Wolf, R. G.et al. (1976). “Hydraulic fracturing to determine the regional in situ stress field Piceance Basin colorado.” Geol. Soc. Am. Bull., 87, 250–258.2.0.CO;2>CrossRefGoogle Scholar
Brown, D. (1987). The flow of water and displacement of hydrocarbons in fractured chalk reservoirs. Fluid flow in sedimentary basins and aquifers, Geological Society London Special Publication 34. J. C. Goff and B. P. Williams. London, The Geological Society. 34, 201–218.CrossRefGoogle Scholar
Brown, K. M., Bekins, B.et al. (1994). “Heterogeneous hydrofracture development and accretionary fault dynamics.” Geology, 22, 259–262.2.3.CO;2>CrossRefGoogle Scholar
Brown, S. R. and Scholz, C. H. (1985). “Closure of random elastic surfaces in contact.” J. Geophys. Res., 90, 5531–5545.CrossRefGoogle Scholar
Brown, S. R. and Scholz, C. H. (1986). “Closure of rock joints.” J. Geophy. Res., 91, 4939–4948.CrossRefGoogle Scholar
Bruce, C. H. (1984). “Smectite dehydration: Its relation to structural development and hydrocarbon accumulation in northern Gulf of Mexico basin.” Am. Assoc. Petr. Geol. Bull., 68, 673–683.Google Scholar
Brudy, M. and Zoback, M. D. (1993). “Compressive and tensile failure of boreholes arbitrarily-inclined to principal stress axes: Application to the KTB boreholes, Germany.” International Journal Rock Mechanics Mining Sciences, 30, 1035–1038.CrossRefGoogle Scholar
Brudy, M. and Zoback, M. D. (1999). “Drilling-induced tensile wall-fractures: implications for the determination of in situ stress orientation and magnitude.” International Journal of Rock Mechanics and Mining Sciences, 136, 191–215.CrossRefGoogle Scholar
Brudy, M., Zoback, M. D.et al. (1997). “Estimation of the complete stress tensor to 8 km depth in the KTB scientific drill holes: Implications for crustal strength.” J. Geophys. Res., 102, 18,453–18,475.CrossRefGoogle Scholar
Burrus, J. (1998). Overpressure models for clastic rocks, their relation to hydrocarbon expulsion: A critical reevaluation – AAPG Memoir 70. In Abnormal pressures in hydrocarbon environments. B. E. Law, G. F. Ulmishek and V. I. Slavin (eds). Tulsa, OK, American Association of Petroleum Geologists, 35–63.
Byerlee, J. D. (1978). “Friction of rock.” Pure & Applied Geophysics, 116, 615–626.CrossRefGoogle Scholar
Carmichael, R. S. (1982). Handbook of Physical Properties of Rocks. Boca Raton, FL, CRC Press.Google Scholar
Carman, P. C. (1961). L'écoulement des Gaz á Travers le Milieux Poreux, Bibliothéque des Sciences et Techniques Nucléaires, Presses Universitaires de France, Paris, 198pp.Google Scholar
Castillo, D., Bishop, D. J.et al. (2000). “Trap integrity in the Laminaria high-Nancar trough region, Timor Sea: Prediction of fault seal failure using well-constrained stress tensors and fault surfaces interpreted from 3D seismics.” Appea Journal, 40, 151–173.CrossRefGoogle Scholar
Castillo, D. and Zoback, M. D. (1995). “Systematic stress variations in the southern San Joaquin valley and along the White Wolf fault: Implications for the rupture mechanics of the 1952 Ms 7.8 Kern County earthquake and contemporary seismicity.” Journal of Geophysical Research, 100(B4), 6249–6264.CrossRefGoogle Scholar
Castillo, D. A. and Zoback, M. D. (1994). “Systematic variations in stress state in the Southern San Joaquin Valley: Inferences based on well-bore data and contemporary seismicity.” American Association Petroleum Geologists Bulletin, 78(8), 1257–1275.Google Scholar
Cayley, G. T. (1987). Hydrocarbon migration in the central North Sea. In Petroleum Geology of North West Europe. Brooks, J. and Glennie, K. (eds). London, Graham and Trotman, 549–555.Google Scholar
Chan, A., Hagin, P. et al. (2004). Viscoplastic deformation, stress and strain paths in unconsolidated reservoir sands (Part 2): Field applications using dynamic DARS analysis: ARMA/NARMS 04–568. Gulf Rocks 2004, The Sixth North American Rock Mechanics Symposium, Houston, TX, American Rock Mechanics Association.
Chan, A. and Zoback, M. D. (2002). Deformation analysis in reservoir space (DARS): A simple formalism for prediction of reservoir deformation with depletion – SPE 78174. SPE/ISRM Rock Mechanics Conference, Irving, TX, Society of Petroleum Engineers.CrossRef
Chan, A. W. and Zoback, M. D. (2006). “The role of hydrocarbon production on land subsidence and fault reactivation in the Louisiana coastal zone.” Journal of Coastal Research, submitted.Google Scholar
Chang, C. and Haimson, B. (2000). “True triaxial strength and deformability of the German Continental Deep Drilling Program (KTB) deep hole amphibolite.” Journal of Geophysical Research, 105, 18999–19013.CrossRefGoogle Scholar
Chang, C., Moos, D.et al. (1997). “Anelasticity and dispersion in dry unconsolidated sand.” International Journal of Rock Mechanics and Mining Sciences, 34(3–4), Paper No. 048.Google Scholar
Chang, C., Zoback, M. D.et al. (2006). “Empirical relations between rock strength and physical properties in sedimentary rocks.” Journal of Petroleum Science and Engineering, 51, 223–237.CrossRefGoogle Scholar
Chapple, W. and Forsythe, D. (1979). “Earthquakes and bending of plates at trenches.” Journal of Geophysical Research, 84, 6729–6749.CrossRefGoogle Scholar
Charlez, P. A. (1991). Rock Mechanics: Theoetical Fundamentals. Paris, Editions Technip.Google Scholar
Chen, S. T. (1988). “Shear-wave logging with dipole sources.” Geophysics, 53, 659–667.CrossRefGoogle Scholar
Cheng, C. H., Jinzhong, Z.et al. (1987). “Effects of in-situ permeabilty on the propagation of Stonely (tube) waves in a borehole.” Geophysics, 52, 1279–1289.CrossRefGoogle Scholar
Chester, F. M. and Logan, J. M. (1986). “Implications for mechanical properties of brittle faults from observations of the Punchbowl fault zone, California.” Pure Appl. Geophys, 124, 79–106.CrossRefGoogle Scholar
Chester, J., Chester, F. M.et al. (2005). “Fracture energy of the Punchbowl fault, San Andreas system.” Nature, 437, 133–136.CrossRefGoogle ScholarPubMed
Cloetingh, S. and Wortel, R. (1986). “Stress in the Indo-Australian plate.” Tectonophysics, 132, 49–67.CrossRefGoogle Scholar
Colmenares, L. B. and Zoback, M. D. (2002). “A statistical evaluation of rock failure criteria constrained by polyaxial test data for five different rocks.” International Journal of Rock Mechanics and Mining Sciences, 39, 695–729.CrossRefGoogle Scholar
Colmenares, L. B. and Zoback, M. D. (2003). “Stress field and seismotectonics of northern South America.” Geology, 31, 721–724.CrossRefGoogle Scholar
Coulomb, C. A. (1773). “Sur une application des regles de maximums et minimums a quelques problemes de statistique relatifs a larchitesture, Acad. Roy.” Sci. Mem. Mech. Min Sci., 7, 343–382.Google Scholar
Crampin, S. (1985). “Evaluation of anisotropy by shear wave splitting.” Geophysics, 50, 142–152.CrossRefGoogle Scholar
Crawford, B. R. and Yale, D. P. (2002). Constitutive modeling of deformation and permeability: relationships between critical state and micromechanics: SPE 78189. Society of Petroleum Engineers.CrossRef
Daines, S. R. (1992). “Aquathermal pressuring and geopressure evaluation.” American Association of Petroleum Geologists Bulletin, 66, 931–939.Google Scholar
Daneshy, A. A. (1973). “A Study of Inclined Hydraulic Fractures.” Soc. Pet. Eng. J., 13, 61.CrossRefGoogle Scholar
Davatzes, N. C. and Aydin, A. (2003). “The formation of conjugate normal fault systems in folded sandstone by sequential jointing and shearing, Waterpocket monocline, Utah.” J. Geophys. Res, 108(B10, 2478), ETG 7–1–7–15.CrossRefGoogle Scholar
Davies, R. and Handschy, J. Eds. (2003). Fault Seals. Tulsa, OK, American Association of Petroleum Geologists.Google Scholar
Davis, S. D., Nyffenegger, P. A.et al. (1995). “The April 1993 Earthquake in South Central Texas: Was it Induced by Oil and Gas Production?”Bull. Seismol. Soc. Am., 85, 1888–1895.Google Scholar
Waal, J. A. and Smits, R. M. M. (1988). “Prediction of reservoir compaction and surface subsidence: Field application of a new model.” SPE Formation Evaluation (June), 347–356.CrossRefGoogle Scholar
Desai, C. S. and Siriwardane, H. J. (1984). Constitutive laws for engineering materials with emphasis on geologic materials. Englewood Cliffs, New Jersey, Prentice-Hall.Google Scholar
Dholakia, S. K., Aydin, A.et al. (1998). “Fault-controlled hydrocarbon pathways in the Monterey Formation, California.” Amer. Assoc. Pet. Geol. Bull., 82, 1551–1574.Google Scholar
Dickinson, G. (1953). “Geological aspects of abnormal reservoir pressures in Gulf Coast Lousiana.” American Association of Petroleum Geologists Bulletin, 37, 410–432.Google Scholar
Donath, F. A. (1966). “Experimental study of shear failure in anisotropic rock.” Bulletin of Geological Soc. America, 72, 985–990.CrossRefGoogle Scholar
Dore, A. G. and Jensen, L. N. (1996). “The impact of late Cenozoic uplift and erosion on hydrocarbon exploration: offshore Norway and some other uplifted basins.” Global and Planetary Change, 12, 415–436.CrossRefGoogle Scholar
Doser, D. I., Baker, M. R.et al. (1991). “Seismicity in the War-Wink Gas Field, West Texas, and its Relationship to Petroleum Production.” Bull. Seismol. Soc. Am., 971.Google Scholar
Drucker, D. and Prager, W. (1952). “Soil mechanics and plastic analysis or limit design.” Quantitative and Applied Mathematics, 10, 157–165.CrossRefGoogle Scholar
du Rouchet, J. (1981). “Stress fields, a key to oil migration.” American Association of Petroleum Geologists Bulletin, 74–85.Google Scholar
Dudley, J. W. I., Meyers, M. T. et al. (1994). Measuring compaction and compressibilities in unconsolidated reservoir materials via time-scaling creep. Eurock ′94, Delft, Netherlands, Balkema.CrossRef
Dugan, B. and Flemings, P. B. (1998). Pore pressure prediction from stacking velocities in the Eugene Island 330 Field (Offshore Lousiana). Chicago, Ill., Gas Research Institute, 23.
Dullien, F. A. L. (1992). Porous Media: Fluid Transport and Pore Structure. San Diego, Academic Press.Google Scholar
Dvorkin, J., Mavko, G.et al. (1995). “Squrt flow in fully saturated rocks.” Geophysics, 60, 97–107.CrossRefGoogle Scholar
Eaton, B. A. (1969). “Fracture gradient prediction and its application in oilfield operations.” Journal of Petroleum Technology, 246, 1353–1360.CrossRefGoogle Scholar
Eberhart-Phillips, D., Han, D.-H.et al. (1989). “Empirical relationships among seismic velocity, effective presure, porosity and clay content in sandstone.” Geophysics, 54, 82–89.CrossRefGoogle Scholar
Economides, M. J. and Nolte, K. G. Eds. (2000). Reservoir Simulation. West Sussex, England, John Wiley & Sons, Ltd.
Ekstrom, M. P., Dahan, C. A.et al. (1987). “Formation imaging with microelectrical scanning arrays.” The Log Analyst., 28, 294–306.Google Scholar
Engelder, T. (1987). Joints and shear fractures in rock. In Fracture Mechanics of Rock. B. K. Atkinson. London, Academic Press, 534.Google Scholar
Engelder, T. (1993). Stress regimes in the lithosphere. Princeton, New Jersey, Princeton.Google Scholar
Engelder, T. and Leftwich, J. T. (1997). A pore-pressure limit in overpressured south Texas oil and gas fields. Seals, traps and the petroleum system: AAPG Memoir 67. R. C. Surdam. Tulsa, OK, AAPG, 255–267.
Engelder, T. and Sbar, M. L. (1984). “Near-surface in situ stress: Introduction.” Journal of Geophysical Research, 89, 9321–9322.CrossRefGoogle Scholar
England, W. A., MacKenzie, A. S.et al. (1987). “The movement and entrapment of petroleum fluids in the subsurface.” Journal of the Geological Society, 144, 327–347.CrossRefGoogle Scholar
Eoff, L., Funkhauser, G. P. et al. (1999). High-density monomer system for formation consolidation/water shutoff applications: SPE 50760. International symposium on oilfield chemistry, Houston, TX, Society of Petroleum Engineers.CrossRef
Ewy, R. (1999). “Wellbore-stability predictions by use of a modified Lade criterion.” SPE Drilling and Completion, 14(2), 85–91.CrossRefGoogle Scholar
Ewy, R., Stankowich, R. J.et al. (2003). Mechanical behavior of some clays and shales from 200 m to 3800 m depth, Paper 570. 39th U.S. Rock Mechanics Symposium/12th Panamerican Conference on Soil Mechanics and Geotechnical Engineering, Cambridge, MA.Google Scholar
Færseth, R. B., Sj⊘blom, R. J.et al. (1995). “Sequence Stratigraphy on the Northwest European Margin.” Elsevier, Amsterdam.Google Scholar
Faybishenko, B., Witherspoon, P. A.et al., Eds. (2000). Dynamics of fluids in fractured rock. Geophysical Monograph Series. Washington, D.C., American Geophysical Union.CrossRefGoogle Scholar
Fehler, M., Jupe, A.et al. (2001). “More than a cloud: new techniques for characterizing reservoir structures using induced seismicity.” Leading Edge, 20, 324–328.CrossRefGoogle Scholar
Feignier, B. and Grasso, J.-R. (1990). “Seismicty Induced by Gas ProductionI: Correlation of Focal Mechanism & Dome Structure.” 134 Pure & Applied Geophys., 405.CrossRefGoogle Scholar
Finkbeiner, T. (1998). In situ stress, pore pressure and hydrocarbon migration and accumulation in sedimentary basins. Geophysics. Stanford, CA, Stanford University, 193.Google Scholar
Finkbeiner, T., Barton, C. B.et al. (1997). “Relationship between in-situ stress, fractures and faults, and fluid flow in the Monterey formation, Santa Maria basin, California.” Amer. Assoc. Petrol. Geol. Bull., 81(12), 1975–1999.Google Scholar
Finkbeiner, T., Zoback, M. D.et al. (2001). “Stress, pore pressure and dynamically-constrained hydrocarbon column heights in the south Eugene Island 330 field, Gulf of Mexico.” Amer. Assoc. Petrol. Geol. Bull., 85(June), 1007–1031.Google Scholar
Fisher, N. I., Lewis, T.et al. (1987). Statistical analysis of spherical data. Cambridge, Cambridge University Press.CrossRefGoogle Scholar
Fisher, Q. J., Casey, M.et al. (2003). “Fluid-flow properties of faults in sandstone: The importance of temperature history.” Geology, 31(11), 965–968.CrossRefGoogle Scholar
Fisher, Q. J. and Knipe, R. J. (1998). Fault sealing processes in siliciclastic sediments. In Faulting and fault sealing in hydrocarbon reservoirs, Jones, G.et al. (eds). London, Geological Society (London), 147, 117–134.
Fjaer, E., Holt, R. M.et al. (1992). Petroleum Related Rock Mechanics. Amsterdam, Elsevier.Google Scholar
Fleitout, L. and Froidevaux, C. (1983). “Tectonics and topography for a lithosphere containing density heterogenieties.” Tectonics, 2, 315–324.CrossRefGoogle Scholar
Flemings, P. B., Stump, B. B.et al. (2002). “Flow focusing in overpressured sandstones: theory, observations and applications.” American Journal of Science, 302, 827–855.CrossRefGoogle Scholar
Forsyth, D. and Uyeda, S. (1975). “On the relative importance of the driving forces of plate motion.” Geophys. J. R. Astr. Soc., 43, 163–200.CrossRefGoogle Scholar
Fowler, C. M. R. (1990). The solid earth. Cambridge, U.K., Cambridge University Press.Google Scholar
Fredrich, J. T., Coblentz, D. D. et al. (2003). Stress perturbations adjacent to salt bodies in the deepwater Gulf of Mexico: SPE 84554. SPE Annual Technical Conference and Exhibition, Denver, CO, Society of Petroleum Engineers.CrossRef
Freyburg, D. (1972). “Der Untere und mittlere Buntsandstein SW-Thuringen in seinen gesteinstechnicschen Eigenschaften.” Ber. Dte. Ges. Geol. Wiss. A; Berlin, 17(6), 911–919.Google Scholar
Fuh, G.-A., Morita, N. et al. (1992). A new approach to preventing loss circulation while drilling. SPE 24599, Soc. Petr. Eng. 67th Annual Tech. Conf. and Exhib, Washington, D.C.
Gaarenstroom, L., Tromp, R. A. J.et al. (1993). Overpressures in the Central North Sea: implications for trap integrity and drilling safety. Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference, London.Google Scholar
Geertsma, J. (1973). “A basic theory of subsidence due to reservoir compaction: the homogeneous case.” Trans. Royal Dutch Soc. of Geologists and Mining Eng., 28, 43–62.Google Scholar
Gephart, J. W. (1990). “Stress and the direction of slip on fault planes.” Tectonics, 9, 845–858.CrossRefGoogle Scholar
Gephart, J. W. and Forsyth, D. W. (1984). “An improved method for determining the regional stress tensor using earthquake focal mechanism data: application to the San Fernando earthquake sequence.” Journal of Geophysical Research, 89, 9305–9320.CrossRefGoogle Scholar
Germanovich, L. N. and Dyskin, A. V. (2000). “Fracture mechanisms and instability of openings in compression.” International Journal of Rock Mechanics and Mining Sciences, 37, 263–284.CrossRefGoogle Scholar
Germanovich, L. N., Galybin, A. N.et al. (1996). Borehole stability in laminated rock. Prediction and performance in rock mechanics and rock engineering, Torino, Italy, A. A. Balkema.Google Scholar
Golubev, A. A. and Rabinovich, G. Y. (1976). “Resultay primeneia appartury akusticeskogo karotasa dija predeleina procontych svoistv gornych porod na mestorosdeniaach tverdych isjopaemych.” Prikladnaja GeofizikaMoskva, 73, 109–116.Google Scholar
Gordon, D. S. and Flemings, P. B. (1998). “Generation of overpressure and compaction-driven fluid flow in a Plio-Pleistocene grwoth-faulted basin, Eugene Island 330, offshore Louisiana.” Basin Research, 10, 177–196.CrossRefGoogle Scholar
Grasso, J. R. (1992). “Mechanics of seismic instabilities induced the recovery of hydrocarbons.” Pure & Applied Geophysics, 139, 507–534.CrossRefGoogle Scholar
Grasso, J. R. and Wittlinger, G. (1990). “10 Years of Seismic Monitoring over a Gas Field.” 80 Bull. Seismo. Soc. Am., 450.Google Scholar
Griffith, J. (1936). Thermal Expansion of Typical American Rocks. Iowa State College of Agriculture and Mechanic Arts, Iowa Engineering Experiment, 35(19), 24.
Grollimund, B., Zoback, M. D.et al. (2001). “Regional synthesis of stress orientation, pore pressure and least principal stress data in the Norwegian sector of the North Sea.” Petroleum Geoscience, 7, 173–180.CrossRefGoogle Scholar
Grollimund, B. R. and Zoback, M. D. (2000). “Post glacial lithospheric flexure and induced stresses and pore pressure changes in the northern North Sea.” Tectonophysics, 327, 61–81.CrossRefGoogle Scholar
Grollimund, B. R. and Zoback, M. D. (2001). “Impact of glacially-induced stress changes on hydrocarbon exploration offshore Norway.” American Association of Petroleum Geologists Bulletin, 87(3), 493–506.CrossRefGoogle Scholar
Grollimund, B. R. and Zoback, M. D. (2003). “Impact of glacially induced stress changes on fault-seal integrity offshore Norway.” American Association of Petroleum Geologists Bulletin, 87, 493–506.CrossRefGoogle Scholar
Gudmundsson, A. (2000). “Fracture dimensions, displacements and fluid transport.” Journal of Structural Geology, 22, 1221–1231.CrossRefGoogle Scholar
Guenot, A. (1989). “Borehole breakouts and stress fields.” Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 26, 185–195.CrossRefGoogle Scholar
Guo, G., Morgenstern, N. R.et al. (1993). “Interpretation of hydraulic fracturing pressure: A comparison of eight methods used to identify shut-in pressures.” International Journal of Rock Mechanics and Mining Sciences, 30, 627–631.CrossRefGoogle Scholar
Hagin, P. and Zoback, M. D. (2004a). Viscoplastic deformation in unconsolidated reservoir sands (Part 1): Laboratory observations and time-dependent end cap models ARMA/NARMS 04–567. Gulf Rocks, Houston, TX, American Rock Mechanics Association.Google Scholar
Hagin, P. and Zoback, M. D. (2004b). “Viscous deformation of unconsolidated sands-Part 1: Time-dependent deformation, frequency dispersion and attenuation.” Geophysics, 69, 731–741.CrossRefGoogle Scholar
Hagin, P. and Zoback, M. D. (2004c). “Viscous deformation of unconsolidated reservoir sands-Part 2: Linear viscoelastic models.” Geophysics, 69, 742–751.CrossRefGoogle Scholar
Hagin, P. and Zoback, M. D. (2007). “Characterization of time-dependent deformation of unconsolidated reservoir sands.” Geophysics, in press.Google Scholar
Haimson, B. and Fairhurst, C. (1967). “Initiation and Extension of Hydraulic Fractures in Rocks.” Soc. Petr. Eng. Jour., Sept.: 310–318.CrossRefGoogle Scholar
Haimson, B. and Fairhurst, C. (1970). In situ stress determination at great depth by means of hydraulic fracturing. In 11th Symposium on Rock Mechanics. W. Somerton, Society of Mining Engineers of AIME, 559–584.
Haimson, B. C. (1989). “Hydraulic fracturing stress measurements.” Rock Mech. and Min. Sci. and Geomech. Abstr Special Issue: Inter. Jour., 26.Google Scholar
Haimson, B. C. and Herrick, C. G. (1989). Borehole breakouts and in situ stress. 12th Annual Energy-Sources Technology Conference and Exhibition, Houston, Texas.
Hall, P. L. (1993). Mechanisms of overpressuring-an overview. Geochemistry of clay-pore fluid interactions. Manning, D. A. C., Hall, P. L. and Hughes, C. R.. London, Chapman and Hall, 265–315.Google Scholar
Han, D., Nur, A.et al. (1986). “Effects of porosity and clay content on wave velocities in sandstones.” Geophysics, 51, 2093–2107.CrossRefGoogle Scholar
Handin, J., Hager, R. V.et al. (1963). “Experimental deformation of sedimentary rocks under confining pressure: pore pressure effects.” Bulletin American Assoc. Petrol. Geology, 717–755.Google Scholar
Haneberg, W. C., Mozley, P. S.et al., Eds. (1999). Faults and subsurface fluid flow in the shallow crust. Geophysical Monograph. Washington, D.C., American Geophysical Union.CrossRefGoogle Scholar
Haney, M. M., Snieder, R.et al. (2005). “A fault caught in the act of burping.” Nature, 437, 46.CrossRefGoogle Scholar
Harrison, A. R., Randall, C. J.et al. (1990). Acquisition and analysis of sonic waveforms from a borehole monopole and dipole source for the determination of compressional shear speeds and their relation to rock mechanic propoerties and surface seismic data – SPE 20557. SPE Annual Technical Conference and Exhibition, New Orleans.CrossRefGoogle Scholar
Harrold, T. W., Swarbrick, R. E.et al. (1999). “Pore pressure estimation from mudrock porosities in Tertiary basins, Southeast Asia.” American Association of Petroleum Geologists Bulletin, 83, 1057–1067.Google Scholar
Hart, B. S., Flemings, P. B.et al. (1995). “Porosity and pressure: Role of compaction disequilibrium in the development of geopressures in a Gulf Coast Pleistocene basin.” Geology, 23, 45–48.2.3.CO;2>CrossRefGoogle Scholar
Hayashi, K. and Haimson, B. C. (1991). “Characteristics of shut-in curves in hydraulic fracturing stress measurements and determination of in situ minimum compressive stress.” Journal of Geophysical Research, 96, 18311–18321.CrossRefGoogle Scholar
Healy, J. H., Rubey, W. W.et al. (1968). “The Denver earthquakes.” Science, 161, 1301–1310.CrossRefGoogle ScholarPubMed
Heppard, P. D., Cander, H. S. et al. (1998). Abnormal pressure and the occurrence of hydrocarbons in offshore eastern Trinidad, West Indies. In Abnormal pressures in hydrocarbon environments – AAPG Memoir 70, Law, B. E., Ulmishek, G. F. and Slavin, V. I. (eds). Tulsa, OK, American Association of Petroleum Geologists. Memoir, 70, 215–246.Google Scholar
Hickman, S. (1991). “Stress in the lithosphere and the strength of active faults, U.S. National Report International Union Geodesy and Geophys. 1987–1990.” Geophys., 29, 759–775.Google Scholar
Hickman, S., Sibson, R.et al. (1995a). “Introduction to special section: Mechanical involvement of fluids in faulting.” J. Geophys. Res., 100, 12831–12840.CrossRefGoogle Scholar
Hickman, S. and Zoback, M. D. (2004). “Stress measurements in the SAFOD pilot hole: Implications for the frictional strength of the San Andreas fault.” Geophysical Research Letters, 31, L15S12.Google Scholar
Hickman, S. H., Barton, C. A.et al. (1997). “In situ stress and fracture permeability along the Stillwater fault zone, Dixie Valley, Nevada.” Int. J. Rock Mech. and Min. Sci., 34, 3–4, Paper No. 126.Google Scholar
Hickman, S. H. and Zoback, M. D. (1983). The interpretation of hydraulic fracturing pressure-time data for in situ stress determination. Hydraulic Fracturing Measurements. Washington, D.C, National Academy Press.Google Scholar
Hoak, T. E., Klawitter, A. L.et al., Eds. (1997). Fractured Reservoirs: Characterization and Modeling. Denver, The Rocky Mountain Association of Geologists.Google Scholar
Hoek, E. and Brown, E. (1997). “Practical estimates of rock strength.” International Journal of Rock Mechanics and Mining Sciences, 34(8), 1165–1186.CrossRefGoogle Scholar
Hoek, E. and Brown, E. T. (1980). “Empirical strength criterion for rock masses.” J. Geotechnical Engineering Div., 106, 1013–1035.Google Scholar
Hofmann, R. (2006). Frequency dependent elastic and anelastic properties of clastic rocks. Geophysics. Golden, CO., Colorado School of Mines. Ph.D., 166.
Holbrook, P. W., Maggiori, D. A.et al. (1993). Real-time pore pressure and fracture gradient evaluation in all sedimentary lithologies, SPE 26791. Offshore European Conference, Aberdeen, Scotland, Society of Petroleum Engineersxs.Google Scholar
Holland, D. S., Leedy, J. B. et al. (1990). “Eugene Island Block 330 Field – U.S.A. offshore Louisiana, Structural Traps III: Tectonic Foldand Fault Traps, Atlas of Oil and Gas Fields, E. Beaumont and N. Foster (eds.). American Assoc. of Petroleum Geologists, Tulsa.” 103–143.
Holt, R. M., Flornes, O. et al. (2004). Consequences of depletion-induced stress changes on reservoir compaction and recovery. Gulf Rocks 2004, the 6th North America Rock Mechanics Symposium (NARMS): Rock Mechanics Across Borders and Disciplines – ARMA/NARMS 04–589. Houston.
Horsrud, P. (2001). “Estimating Mechanical Properties of Shale from Empirical Correlations.” SPE Drilling and Completion, 16(2), 68–73.CrossRefGoogle Scholar
Hottman, C. E., Smith, J. E.et al. (1979). “Relationship among earth stresses, pore pressure, and drilling problems offshore Gulf of Alaska.” Journal of Petroleum Technology, November, 1477–1484.CrossRefGoogle Scholar
Hubbert, M. D. and Rubey, W. W. (1959). “Role of fluid pressure in mechanics of overthrust faulting.” Geol. Soc. Am. Bull., 70, 115–205.CrossRefGoogle Scholar
Hubbert, M. K. and Willis, D. G. (1957). “Mechanics of hydraulic fracturing.” Petr. Trans. AIME, 210, 153–163.Google Scholar
Hudson, J. A. (1981). “Wave propagation and attenuation of elastic waves in material containing cracks.” Geophys. J. Roy. Astr. Soc., 64, 122–150.CrossRefGoogle Scholar
Hudson, J. A. and Priest, S. D. (1983). “Discontinuity frequency in rock masses.” Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 20(2), 73–89.CrossRefGoogle Scholar
Huffman, A. and Bowers, G. L., Eds. (2002). Pressure regimes in sedimentary basins and their prediction: AAPG memoir 76. Tulsa, OK, American Association of Petroleum Geologists.Google Scholar
Ingebritsen, S. E., Sanford, W.et al. (2006). Groundwater in Geologic Processes. Cambridge, United Kingdom, Cambridge Press.Google Scholar
Ito, T., Zoback, M. D.et al. (2001). “Utilization of mud weights in excess of the least principal stress to stabilize wellbores: Theory and practical examples.” Soc. of Petroleum Engineers Drilling and Completions, 16, 221–229.Google Scholar
Jaeger, J. C. and Cook, N. G. W. (1971). Fundamentals of Rock Mechanics. London, Chapman and Hall.Google Scholar
Jaeger, J. C. and Cook, N. G. W. (1979). Fundamentals of Rock mechanics, 2nd edn. New York, Chapman and Hall.Google Scholar
Jarosinski, M. (1998). “Contemporary stress field distortion in the Polish part of the western outer Carpathians and their basement.” Tectonophysics, 297, 91–119.CrossRefGoogle Scholar
Jizba, D. (1991). Mechanical and acoustical properties of Sandstones and shales. PhD dissentation, Stanford University.Google Scholar
Jones, G., Fisher, Q. J.et al., Eds. (1998). Faulting, fault sealing and fluid flow in hydrocarbon resrvoirs. London, Geological Society.Google Scholar
Kamb, W. B. (1959). “Ice petrofabric observations from Blue Glacier, Washington, in relation to theory and experiment.” J. Geophys. Res., 64, 1891–1910.CrossRefGoogle Scholar
Kimball, C. V. and Marzetta, T. M. (1984). “Semblance processing of borehole acoustic array data.” Geophysics, 49, 264–281.CrossRefGoogle Scholar
Kirsch, G. (1898). “Die Theorie der Elastizitat und die Bedurfnisse der Festigkeitslehre, Zeitschrift des Verlines Deutscher Ingenieure.” 42, 707.Google Scholar
Klein, R. J. and Barr, M. V. (1986). Regional state of stress in western Europe. In Proc. of International Symposium on Rock Stress and Rock Stress Measurements. Lulea, Sweden, Stockholm, Centek Publ, 694 pp.Google Scholar
Kosloff, D. and Scott, R. F. (1980). “Finite element simulation of Wilmington oilfield subsidence: I, Linear modeling.” Tectonophysics, 65, 339–368.CrossRefGoogle Scholar
Kranz, R. L., Frankel, A. D.et al. (1979). “The permeability of whole and jointed Barre granite.” Int. J. Rock Mech. Min. Sci. Geomech. Abst., 16, 225–234.CrossRefGoogle Scholar
Kristiansen, G. (1998). Geomechanical characterization of the overburden above the compacting chalk reservoir at Valhall. In Eurock ′98, SPE/ISRM Rock Mechanics in Petroleum Engineering. Trondheim, Norway, The Norwegian University of Science and Technology, 193–202.Google Scholar
Kuempel, H. J. (1991). “Poroelasticity: parameters reviewed.” Geophysical Journal International, 105, 783–799.CrossRefGoogle Scholar
Kwasniewski, M. (1989). Laws of brittle failure and of B-D transition in sandstones. Rock at Great Depth, Proceedings ISRM-SPE International Symposium, Elf Aquitaine, Pau, France, A. A. Balkema.Google Scholar
Kwon, O., Kronenberg, A. K.et al. (2001). “Permeability of Wilcox shale and its effective pressure law.” Journal of Geophysical Research, 106, 19339–19353.CrossRefGoogle Scholar
Labenski, F., Reid, P.et al. (2003). Drilling fluids approaches for control of wellbore instability in fractured formations SPE/IADC 85304. SPE/IADC Middle East Drilling Technology Conference and Exhibition, Abu Dhabi, UAE, Society of Petroleum Engineers.Google Scholar
Lachenbruch, A. H. and Sass, J. H. (1992). “Heat flow from Cajon Pass, fault strength and tectonic implications.” J. Geophys. Res., 97, 4995–5015.CrossRefGoogle Scholar
Lade, P. (1977). “Elasto-plasto stress-strain theory for cohesionless soil with curved yield surfaces.” International Journal of Solids and Structures, 13, 1019–1035.CrossRefGoogle Scholar
Lal, M. (1999). Shale stability: drilling fluid interaction and shale strength, SPE 54356. SPE Latin American and Carribean Petroleum Engineering Conference, Caracas, Venezuela, Society of Petroleum Engineering.Google Scholar
Lama, R. and Vutukuri, V. (1978). Handbook on Mechanical Properties of Rock. Clausthal, Germany, Trans Tech Publications.Google Scholar
Lashkaripour, G. R. and Dusseault, M. B. (1993). A statistical study of shale properties; relationship amnog principal shale properties. Proceedings of the Conference on Probabilistic Methods in Geotechnical Engineering, Canberra, Australia.Google Scholar
Laubach, S. E. (1997). “A method to detect natural fracture strike in sandstones.” Amer. Assoc. Petrol. Geol. Bull., 81(4), 604–623.Google Scholar
Law, B. E., Ulmishek, G. F. et al., Eds. (1998). Abnormal pressures in hydrocarbon environments. AAPG Memoir 70, American Association of Petroleum Geologists.
Lekhnitskii, S. G. (1981). Theory of elasticity of an anisotropic body. Moscow, Mir.Google Scholar
Leslie, H. D. and Randall, C. J. (1990). “Eccentric dipole sources in fluid-filled boreholes: Experimental and numerical results.” Journal of Acoustic Society of America, 87, 2405–2421.CrossRefGoogle Scholar
Li, X., Cui, L.et al. (1998). Thermoporoelastic modelling of wellbore stability in non-hydrostatic stress field. 3rd North American Rock Mechanics Symposium.Google Scholar
Ligtenberg, J. H. (2005). “Detection of fluid migration pathways in seismic data: implications for fault seal analysis.” Basin Research, 17, 141–153.CrossRefGoogle Scholar
Lindholm, C. D., Bungum, H. et al. (1995). Crustal stress and tectoncis in Norwegian regions determined from earthquake focal mechanisms. Proceedings of the Workshop on Rock Stresses in the North Sea, Trondheim, Norway.
Lockner, D. A. (1995). Rock Failure. Rock physics and phase relations. Washington, D.C., American Geophysical Union, 127–147.CrossRefGoogle Scholar
Lockner, D. A., Byerlee, J. D.et al. (1991). “Quasi-static fault growth and shear fracture energy in granite.” Nature, 350, 39–42.CrossRefGoogle Scholar
Long, C. S. and a. others (1996). Rock fractures and fluid flow. Washington, D.C., National Academy Press.Google Scholar
Lorenz, J. C., Teufel, L. W.et al. (1991). “Regional fractures I: A mechanism for the formation of regional fractures at depth in flat-lying reservoirs.” Amer. Assoc. Petrol. Geol. Bull., 75(11), 1714–1737.Google Scholar
Losh, S., Eglinton, L.et al. (1999). “Vertical and lateral fluid flow related to a large growth fault, South Eugene Island Block 330 field, Offshore Louisiana.” American Association of Petroleum Geologists Bulletin, 83(2), 244–276.Google Scholar
Lucier, A., Zoback, M. D.et al. (2006). “Geomechanical aspects of CO2 sequestration in a deep saline reservoir in the Ohio River Valley region.” Environmental Geosciences, 13(2), 85–103.CrossRefGoogle Scholar
Lund, B. and Zoback, M. D. (1999). “Orientation and magnitude of in situ stress to 6.5 km depth in the Baltic Shield.” International Journal of Rock Mechanics and Mining Sciences, 36, 169–190.CrossRefGoogle Scholar
Luo, M. and Vasseur, G. (1992). “Contributions of compaction and aquathermal pressuring to geopressure and the influence of environmental conditions.” American Association of Petroleum Geologists Bulletin, 76, 1550–1559.Google Scholar
MacKenzie, D. P. (1969). “The relationship between fault plane solutions for earthquakes and the directions of the principal stresses.” Seismological Society of Amererica Bulletin, 59, 591–601.Google Scholar
Mallman, E. P. and Zoback, M. D. (2007). “Subsidence in the Louisiana Coastal Zone due to hydrocarbon production.” Journal of Coastal Research, in press.Google Scholar
Mastin, L. (1988). “Effect of borehole deviation on breakout orientations.” J. Geophys. Res., 93(B8), 9187–9195.CrossRefGoogle Scholar
Matthews, W. R. and Kelly, J. (1967). “How to predict formation pressure and fracture gradient.” Oil and Gas Journal, February, 92–106.Google Scholar
Maury, V. and Zurdo, C. (1996). “Drilling-induced lateral shifts along pre-existing fractures: A common cause of drilling problems.” SPE Drilling and Completion (March), 17–23.CrossRefGoogle Scholar
Mavko, G., Mukerjii, T.et al. (1998). Rock Physics Handbook. Cambridge, United Kingdom (GBR), Cambridge University Press.Google Scholar
Mavko, G. and Nur, A. (1997). “The effect of percolation threshold in the Kozeny-Carman relation.” Geophysics, 622, 1480–1482.CrossRefGoogle Scholar
Maxwell, S. C. (2000). Comparison of production-induced microseismicity from Valhall and Ekofisk. Passive Seismic Method in E&P of Oil and Gas Workshop, 62nd EAGE Conference.Google Scholar
Maxwell, S. C., Urbancic, T. I.et al. (2002). Microseismic imaging of hydraulkic fracture complexity in the Barnett shale, Paper 77440. Society Petroleum Engineering Annual Technical Conference, San Antonio, TX, Society of Petroleum Engineers.Google Scholar
McGarr, A. (1991). “On a Possible Connection Between Three Major Earthquakes in California and Oil Production.” Bull. of Seismol. Soc. of Am., 948, 81.Google Scholar
McGarr, A. and Gay, N. C. (1978). “State of stress in the earth's crust.” Ann. Rev. Earth Planet. Sci., 6, 405–436.CrossRefGoogle Scholar
McLean, M. and Addis, M. A. (1990). Wellbore stability: the effect of strength criteria on mud weight recommendations. 65th Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, New Orleans, Society of Petroleum Engineers.Google Scholar
McNally, G. H. N. (1987). “Estimation of coal measures rock strength using sonic and neutron logs.” Geoexploration, 24, 381–395.CrossRefGoogle Scholar
McNutt, M. K. and Menard, H. W. (1982). “Constraints on yield strength in the oceanic lithosphere derived from observations of flexure.” Geophys. J. R. Astron. Soc., 71, 363–394.CrossRefGoogle Scholar
Mereu, R. F., Brunet, J.et al. (1986). “A study of the Microearthquakes of the Gobbles Oil Field Ara of Southwestern Ontario.” Bulle. Seismol. Soc. Am., 1215, 76.Google Scholar
Michael, A. (1987). “Use of focal mechanisms to determine stress: A control study.” Journal of Geophysical Research, 92, 357–368.CrossRefGoogle Scholar
Militzer, M.a. S., P. (1973). “Einige Beitrageder geophysics zur primadatenerfassung im Bergbau.” Neue Bergbautechnik, Lipzig, 3(1), 21–25.Google Scholar
Mitchell, A. and Grauls, D. Eds. (1998). Overpressure in petroleum exploration. Pau, France, Elf-Editions.Google Scholar
Mody, F. K. and Hale, A. H. (1993). A borehole stability model to couple the mechanics and chemistry of drilling-fluid/shale interactions. SPE/IADC Drilling Conference, Amsterdam.Google Scholar
Mogi, K. (1971). “Effect of the triaxial stress system on the failure of dolomite and limestone.” Tectonophysics, 11, 111–127.CrossRefGoogle Scholar
Mollema, P. N. and Antonellini, M. A. (1996). “Compaction bands; a structural analog for anti-mode I cracks in aeolian sandstone.” Tectonophysics, 267, 209–228.CrossRefGoogle Scholar
Moore, D. E. and Lockner, D. A. (2006). Friction of the smectite clay montmorillonite: A review and interpretation of data. The Seismogenic Zone of Subduction Thrust Faults, MARGINS Theoretical and Experimental Science Series. C. M. and T. Dixon. New York, Columbia University Press. 2.Google Scholar
Moos, D., Peska, P. et al. (2003). Comprehensive wellbore stability analysis using quantitative risk assessment. Jour. Petrol. Sci. and Eng., Spec. Issue on Wellbore Stability, 38. B. S. Aadnoy and S. Ong, 97–109.
Moos, D. and Zoback, M. D. (1990). “Utilization of Observations of Well Bore Failure to Constrain the Orientation and Magnitude of Crustal Stresses: Application to Continental Deep Sea Drilling Project and Ocean Drilling Program Boreholes.” J. Geophys. Res., 95, 9305–9325.CrossRefGoogle Scholar
Moos, D. and Zoback, M. D. (1993). “State of stress in the Long Valley caldera, California.” Geology, 21, 837–840.2.3.CO;2>CrossRefGoogle Scholar
Moos, D., Zoback, M. D.et al. (1999). Feasibility study of the stability of openhole multilaterals, Cook Inlet, Alaska. 1999 SPE Mid-continent Operations Symposium, Oklahoma City, OK, Society of Petroleum Engineers.CrossRefGoogle Scholar
Morita, N., Black, A. D.et al. (1996). “Borehole Breakdown Pressure with Drilling Fluids – I. Empirical Results.” Int. J. Rock Mech. and Min. Sci., 33, 39.CrossRefGoogle Scholar
Morita, N. and McLeod, H. (1995). “Oriented perforation to prevent casing collapse for highly inclined wells.” SPE Drilling and Completion, (September), 139–145.CrossRefGoogle Scholar
Morrow, C., Radney, B.et al. (1992). Frictional strength and the effective pressure law of montmorillonite and illite clays. Fault Mechanics and Transport Properties of Rocks, Academic, San Diego, Calif.Google Scholar
Mouchet, J. P. and Mitchell, A. (1989). Abnormal pressures while drilling. Manuels techniques Elf Aquitaine, 2. Boussens, France, Elf Aquitaine.Google Scholar
Mount, V. S. and Suppe, J. (1987). “State of stress near the San Andreas fault: Implications for wrench tectonics.” Geology, 15, 1143–1146.2.0.CO;2>CrossRefGoogle Scholar
Mueller, M. C. (1991). “Prediction of lateral variability in fracture intensity using multicomponent shear-wave seismic as a precursor to horizontal drilling.” Geophysical Journal International, 107, 409–415.CrossRefGoogle Scholar
Munns, J. W. (1985). “The Valhall field: a geological overview.” Marine and Petroleum Geology, 2, 23–43.CrossRefGoogle Scholar
Murrell, S. A. F. (1965). “Effect of triaxial systems on the strength of rocks at atmospheric temperatures.” Geophysical Journal Royal Astron. Soc., 106, 231–281.CrossRefGoogle Scholar
Nakamura, K., Jacob, K. H.et al. (1977). “Volcanoes as possible indicators of tectonic stress orientation – Aleutians and Alaska.” Pure and Applied Geophysics, 115, 87–112.CrossRefGoogle Scholar
Nashaat, M. (1998). Abnormally high formation pressure and seal impacts on hydrocarbon accumulations in the Nile Delta and North Sinai basins, Egypt. Abnormal pressure in hydrocarbon environments: AAPG Memoir 70. B. E. Law, G. F. Ulmishek and V. I. Slavin. Tulsa, OK, AAPG, 161–180.
Nolte, K. G. and M. J. Economides (1989). Fracturing diagnosis using pressure analysis. Reservoir Simulation. Economides, M. J. and Nolte, K. G.. Englewood Cliffs, N.J., Prentice Hall.
Nur, A. and Byerlee, J. D. (1971). “An exact effective sress law for elastic deformation of rock with fluids.” J. Geophys. Res., 6414–6419.CrossRefGoogle Scholar
Nur, A. and Simmons, G. (1969). “Stress induced velocity anisotropy in rock: An experimental study.” J. Geophys. Res., 74, 6667–6674.CrossRefGoogle Scholar
Nur, A. and Walder, J. (1990). Time-Dependent Hydraulics of the Earth's Crust. The role of fluids in crustal processes. Washington D.C., National Research Council, 113–127.Google Scholar
Okada, Y. (1992). “Internal deformation due to shear and tensile faults in a half space.” Bulletin of Seismological Society of America, 82, 1018–1040.Google Scholar
Ortoleva, P., Ed. (1994). Basin Compartments and Seals. Tulsa, American Association of Petroleum Geologists.Google Scholar
Ostermeier, R. M. (1995). “Deepwater Gulf of Mexico turbidites – compaction effects on porosity and permeabilty.” Soc. of Petroleum Engineers Formation Evaluation, 79–85.CrossRefGoogle Scholar
Ostermeier, R. M. (2001). “Compaction effects on porosity and permeability: deepwater Gulf of Mexico turbidites.” Journal of Petroleum Technology, 53(Feb. 2001), 68–74.CrossRefGoogle Scholar
Ott, W. K. and Woods, J. D. (2003). Modern Sandface Completion Practises. Houston, Texas, World Oil.Google Scholar
Ottesen, S., Zheng, R. H.et al. (1999). Wellbore Stability Assessment Using Quantitative Risk Analysis, SPE/IADC 52864. SPE/IADC Drilling Conference, Amsterdam, The Netherlands, Society of Petroleum Engineers.Google Scholar
Paterson, M. S. and Wong, T.-f . (2005). Experimental rock deformation – The brittle field. Berlin, Springer.Google Scholar
Paul, P. and Zoback, M. D. (2006). Wellbore Stability Study for the SAFOD Borehole through the San Andreas Fault: SPE 102781. SPE Annual Technical Conference, San Antonio, TX.CrossRefGoogle Scholar
Pennington, W. D., Davis, S. D.et al. (1986). “The Evolution of Seismic Barriers and Asperities Caused by the Depressuring of Fault Planes in Oil & Gas Fields of South Texas.” Bull. Seism. Soc. Am., 188, 78.Google Scholar
Pepin, G., Gonzalez, M.et al. (2004). “Effect of drilling fluid temperature on fracture gradient.” World Oil, October, 39–48.Google Scholar
Perzyna, P. (1967). Fundamental Problems in Viscoplasticity. Advances in Applied Mechanics, 9, 244–368.Google Scholar
Peska, P. and Zoback, M. D. (1995). “Compressive and tensile failure of inclined wellbores and determination of in situ stress and rock strength.” Journal of Geophysical Research, 100(B7), 12791–12811.CrossRefGoogle Scholar
Peska, P. and Zoback, M. D. (1996). “Stress and failure of inclined boreholes SFIB v2.0: Stanford Rock Physics and Borehole Geophysics Annual Report, 57, Paper H3.” Stanford University Department of Geophysics.Google Scholar
Pine, R. J. and Batchelor, A. S. (1984). “Downward migration of shearing in jointed rock during hydraulic injections.” Int. Journ. Rock Mech. Min. Sci. & Geomech. Abst., 21(5), 249–263.CrossRefGoogle Scholar
Pine, R. J., Jupe, A.et al. (1990). An evaluation of in situ stress measurements affecting different volumes of rock in the Carnmenellis granite. Scale Effects in Rock Masses. P. d. Cunha. Rotterdam, Balkema, 269–277.Google Scholar
Plumb, R. A. and Cox, J. W. (1987). “Stress directions in eastern North America determined to 4.5 km from borehole elongation measurements.” Journal of Geophysical Research, 92, 4805–4816.CrossRefGoogle Scholar
Plumb, R. A. and Hickman, S. H. (1985). “Stress-induced borehole elongation: A comparison between the four-arm dipmeter and the borehole televiewer in the Auburn geothermal well.” Journal of Geophysical Research, 90, 5513–5521.CrossRefGoogle Scholar
Pollard, D. and Aydin, A. (1988). “Progress in understanding jointing over the past century.” Geological Society of America Bulletin, 100, 1181–1204.2.3.CO;2>CrossRefGoogle Scholar
Pollard, D. and Fletcher, R. C. (2005). Fundamentals of Structural Geology. Cambridge, United Kingdom, Cambridge University Press.Google Scholar
Pollard, D. and Segall, P. (1987). Theoretical displacements and stresses near fractures in rock: With applications to faults, joints, veins, dikes and solution surfaces. Fracture Mechanics of Rock. B. K. Atkinson, Academic Press.Google Scholar
Powley, D. E. (1990). “Pressures and hydrogeology in petroleum basins.” Earth Sci. Rev., 29, 215–226.CrossRefGoogle Scholar
Qian, W., Crossing, K. S.et al. (1994). “Corrections to “Inversion of borehole breakout orientation data by Wei Qian and L.B. Pedersen.” Journal of Geophysical Research, 99, 707–710.CrossRefGoogle Scholar
Qian, W. and Pedersen, L. B. (1991). “Inversion of borehole breakout orientation data.” Journal of Geophysical Research, 96, 20093–20107.CrossRefGoogle Scholar
Raaen, A. M. and Brudy, M. (2001). Pump in/Flowback tests reduce the estimate of horizontal in situ stress significantly, SPE 71367. SPE Annual Technical Conference, New Orleans, LA, Society of Petroleum Engineers.Google Scholar
Raleigh, C. B., Healy, J. H.et al. (1972). Faulting and crustal stress at Rangely, Colorado., Flow and Fracture of Rocks. J. C. Heard. Washington, D.C., American Geophysical Union, 275–284.Google Scholar
Raleigh, C. B., Healy, J. H.et al. (1976). “An experiment in earthquake control at Rangely, Colorado.” Science, 191, 1230–1237.CrossRefGoogle ScholarPubMed
Reid, P. and Santos, H. (2003). Novel drilling, completion and workover fluids for depleted zones: Avoiding losses, formation damage and stuck pipe: SPE/IADC 85326. SPE/IADC Middle East Drilling Technology Conference and Exhibition, Abu Dhabi, UAE, Society of Petroleum Engineers.Google Scholar
Richardson, R. (1992). “Ridge forces, absolute plate motions and the intraplate stress field.” Journal of Geophysical Research, 97(B8), 11739–11748.CrossRefGoogle Scholar
Richardson, R. M. (1981). Hydraulic fracture in arbitrarily oriented boreholes: An analytic approach. Workshop on Hydraulic Fracturing Stress Measurements, Monterey, California, National Academy Press.Google Scholar
Riis, F. (1992). “Dating and measuring of erosion, uplift and subsidence in Norway and the Norwegian shelf in glacial periods.” Norsk Geologisk Tidsskrift, 72, 325–331.Google Scholar
Ritchie, R. H. and Sakakura, A. Y. (1956). “Asymptotic expansions of solutions of the heat conduction equation in internally bounded cylindrical geometry.” J. Appl. Phys., 27, 1453–1459.CrossRefGoogle Scholar
Roegiers, J. C. and Detournay, E. (1988). Considerations on failure initiation in inclined boreholes. Key Questions in Rock Mechanics, Balkeema, Brookfield, Vermont.Google Scholar
Rogers, S. (2002). Critical stress-related permeability in fractured rocks. Fracture and in situ stress characterization of hydrocarbon reservoirs. M. Ameen. London, The Geological Society, 209, 7–16.Google Scholar
Rojas, J. C., Clark, D. E.et al. (2006). Optimized salinity delivers improved drilling performance: AADE-06-DF-HO-11. AADE 2006 Fluids Conference, Houston, Texas, American Association of Drilling Engineers.Google Scholar
Rudnicki, J. W. (1999). Alteration of regional stress by reservoirs and other inhomogeneities: Stabilizing or destabilizing? Proceedings International Congress on Rock Mechanics. G. Vouille and P. Berest, International Society of Rock Mechanics, 3, 1629–1637.Google Scholar
Rummel, F. and Hansen, J. (1989). “Interpretation of hydrofrac pressure recordings using a simple fracture mechanics simulation model, Inter. Jour. Rock Mech. and Min.” Sci. and Geomech. Abstr., 26, 483–488.Google Scholar
Rummel, F. and Winter, R. B. (1983). “Fracture mechanics as applied to hydraulic fracturing stress measurements.” Earthq. Predict. Res., 2, 33–45.Google Scholar
Rutledge, J. T., Phillips, W. S.et al. (2004). “Faulting induced by forced fluid injection and fluid flow forced by faulting: An interpretation of hydraulic-fracture microseismicity, Carthage Cotten Valley Gas Field, Texas.” Bulletin of the Seismological Society of America, 94(5), 1817–1830.CrossRefGoogle Scholar
Rzhevsky, V. and Novick, G. (1971). The physics of rocks. Moscow, Russia, Mir.Google Scholar
Sayers, C. M. (1994). “The elastic anisotropy of shales.” J. Geophys. Res., 99, 767–774.CrossRefGoogle Scholar
Schmitt, D. R. and Zoback, M. D. (1992). “Diminished pore pressure in low-porosity crystalline rock under tensional failure: Apparent strengthening by dilatancy.” J. Geophys. Res., 97, 273–288.CrossRefGoogle Scholar
Schowalter, T. T. (1979). “Mechanics of Secondary Hydrocarbon Migration and Entrapment.” American Association of Petroleum Geologists Bulletin, 63(5), 723–760.Google Scholar
Schutjens, P. M. T. M., Hanssen, T. H. et al. (2001). Compaction-induced porosity/permeability reduction in sandstone reservoirs: Data and model for elasticity-dominated deformation, SPE 71337. SPE Annual Technical Conference and Exhibition, New Orleans, LA, Society of Petroleum Engineers.CrossRef
Secor, D. T. (1965). “Role of fluid pressure in jointing.” American Journal of Science, 263, 633–646.CrossRefGoogle Scholar
Segall, P. (1985). “Stress and Subsidence from Subsurface Fluid Withdrawal in the Epicentral Region of the 1983 Coalinga Earthquake.” J. Geophys. Res., 6801, 90.Google Scholar
Segall, P. (1989). “Earthquakes Triggered by Fluid Extraction.” Geology, 17, 942–946.2.3.CO;2>CrossRefGoogle Scholar
Segall, P. and Fitzgerald, S. D. (1996). “A note on induced stress changes in hydrocarbon and geothermal reservoirs.” Tectonophysics, 289, 117–128.CrossRefGoogle Scholar
Seldon, B. and Flemings, P. B. (2005). “Reservoir pressure and seafloor venting: Predicting trap integrity in a Gulf of Mexico deepwater turbidite minibasin.” American Association of Petroleum Geologists Bulletin, 89(2), 193–209.CrossRefGoogle Scholar
Shamir, G. and Zoback, M. D. (1992). “Stress orientation profile to 3.5 km depth near the San Andreas Fault at Cajon Pass California.” Jour. Geophys Res., 97, 5059–5080.CrossRefGoogle Scholar
Sibson, R. H. (1992). “Implications of fault valve behavior for rupture nucleation and recurrence.” Tectonophysics, 211, 283–293.CrossRefGoogle Scholar
Sinha, B. K. and Kostek, S. (1996). “Stress-induced azimuthal anisotropy in borehole flexural waves.” Geophysics, 61, 1899–1907.CrossRefGoogle Scholar
Sinha, B. K., Norris, A. N.et al. (1994). “Borehole flexural modes in anisotropic formations.” Geophysics, 59, 1037–1052.CrossRefGoogle Scholar
Sonder, L. (1990). “Effects of density contrasts on the orientation of stresses in the lithosphere: Relation to principal stress directions in the Transverse Ranges, California.” Tectonics, 9, 761–771.CrossRefGoogle Scholar
Stein, R. S., King, G. C.et al. (1992). “Change in failure stress on the southern San Andreas fault system caused by the 1992 magnitude 7.4 Landers earthquake.” Science, 258, 1328–1332.CrossRefGoogle ScholarPubMed
Stein, S. and Klosko, E. (2002). Earthquake mechanisms and plate tectonics. International Handbook of Earthquake and Engineering Seismology Part A. Lee, W. H. K., Kanamori, H., Jennings, P. C. and Kisslinger, K.. Amsterdam, Academic Press, 933.Google Scholar
Stephens, G. and Voight, B. (1982). “Hydraulic fracturing theory for conditions of thermal stress.” International Journal of Rock Mechanics and Mining Sciences, 19, 279–284.CrossRefGoogle Scholar
Sternlof, K. R., Karimi-Fard, M., Pollard, D. D. and Durlofsky, L. J. (2006). “Flow and transport effects of compaction bands in sandstone at scales relevant to aquifer and reservoir management.” Water Resources Research, 42, Wo7425.CrossRefGoogle Scholar
Sternlof, K. R., Rudnicki, J. W.et al. (2005). “Anticrack inclusion model for compaction bands in sandstone.” J. Geophys. Res, 110(B11403), 1–16.CrossRefGoogle Scholar
Stock, J. M., Healy, J. H.et al. (1985). “Hydraulic fracturing stress measurements at Yucca Mountain, Nevada, and relationship to the regional stress field.” Journal of Geophysical Research, 90(B10), 8691–8706.CrossRefGoogle Scholar
Stump, B. B. (1998). Illuminating basinal fluid flow in Eugene Island 330 (Gulf of Mexico) through in situ observations, deformation experiments, and hydrodynamic modeling. Geosciences, Pennsylvania State, 121.
Sulak, R. M. (1991). “Ekofisk field: The first 20 years.” Journal of Petroleum Technology, 33, 1265–1271.CrossRefGoogle Scholar
Swarbrick, R. E. and Osborne, M. J. (1998). Mechanisms that generate abnormal pressures: An overview. Abnormal pressures in hydrocarbon environments, AAPG Memoir 70. B. E. Law, G. F. Ulmishek and V. I. Slavin. Tulsa, OK, American Association of Petroleum Geologists, 13–34.
Takahashi, M. and Koide, H. (1989). Effect of the intermediate principal stress on strength and deformation behavior of sedimentary rock at the depth shallower than 2000 m. Rock at Great Depth, Pau, France, Balkema, Rotterdam.
Tang, X. M. and Cheng, C. H. (1996). “Fast inversion of formation permeability from borehole Stonely wave logs.” Geophysics, 61, 639–645.CrossRefGoogle Scholar
Terzaghi, K. (1923). Theoretical Soil Mechanics. John Wiley, New York.Google Scholar
Teufel, L. W. (1992). Production-induced changes in reservoir stress state: Applications to reservoir management. Society of Exploration Geophysicists 62nd Annual International Meeting, New Orleans, SEG, Tulsa.Google Scholar
Teufel, L. W., Rhett, D. W.et al. (1991). Effect of reservoir depletion and pore pressure drawdown on In situ stress and deformation in the Ekofisk field, North Sea. Rock Mechanics as a Multidisciplinary Science. J. C. Roegiers. Rotterdam, Balkema.Google Scholar
Tezuka, K. (2006). Hydraulic injection and microseismic monitoring in the basement gas reservoir in Japan. 2006 SPE Forum Series in Asia Pacific – Hydraulic Fracturing Beyond 2010, Macau, China, Society of Petroleum Engineers.
Thompson, A. L. (1993). Poly3D: A three-dimensional, polygonal element, displacement discontinuity boundary element computer program iwth applications to fractures, faults and cavities in the earth's crust. Geology. Stanford, CA, Stanford University.Google Scholar
Thomsen, L. (1986). “Weak elastic anisotropy.” Geophysics, 51, 1954–1966.CrossRefGoogle Scholar
Toublanc, A., Renaud, S.et al. (2005). “Ekofisk field: fracture permeability evaluation and implementation in the flow model.” Petroleum Geoscience, 11, 321–330.CrossRefGoogle Scholar
Townend, J. (2003). Mechanical constraints on the strength of the lithosphere and plate-bounding faults. Geophysics. Stanford, CA, Stanford University. PhD, 135.Google Scholar
Townend, J. and Zoback, M. D. (2000). “How faulting keeps the crust strong.” Geology, 28(5), 399–402.2.0.CO;2>CrossRefGoogle Scholar
Townend, J. and Zoback, M. D. (2001). Implications of earthquake focal mechanisms for the frictional strength of the San Andreas fault system. The Nature and Tectonic Significance of Fault Zone Weakening. R. E. Holdsworth, R. A. Strachan, J. J. Macloughlin and R. J. Knipe. London, Special Publication of the Geological Society of London, 186, 13–21.Google Scholar
Townend, J. and Zoback, M. D. (2004). “Regional tectonic stress near the San Andreas fault in Central and Northern California.” Geophysical Research Letters, 31, L15–18.CrossRefGoogle Scholar
Traugott, M. O. and Heppard, P. D. (1994). Prediction of pore pressure before and after drilling- taking the risk out of drilling overpressured prospects. AAPG Hedberg Research Conference, American Association of Petroleum Geologists.Google Scholar
Tsvankin, I. (2001). Seismic Signatures and Analysis of Reflection Data in Anisotropic Media. Cambridge, MA, Elsevier Science.Google Scholar
Turcotte, D. L. and Schubert, G. (2002). Geodynamics. Cambridge, Cambridge.CrossRefGoogle Scholar
Tutuncu, A. N., Podio, A. L.et al. (1998). “Nonlinear viscoelastic behavior of sedimentary rocks: Part I, Effect of frequency and strain amplitude.” Geophysics, 63(1), 184–194.CrossRefGoogle Scholar
Tutuncu, A. N., Podio, A. L.et al. (1998). “Nonlinear viscoelastic behavior of sedimentary rocks: Part II, Hysteresis effects and influence of type of fluid on elastic moduli.” Geophysics, 63(1), 195–203.CrossRefGoogle Scholar
Twiss, R. J. and Moores, E. M. (1992). Structural Geology. New York, W. H. Freeman and Company.Google Scholar
Balen, R. T. and Cloetingh, S. A. (1993). Stress-induced fluid flow in rifted basins. Diagenesis and basin development. A. D. Horbury and A. G. Robinson. Tulsa, American Association of Petroleum Geologists, 36, 87–98.Google Scholar
Oort, E., Gradisher, J.et al. (2003). Accessing deep reservoirs by drilling severly depleted formations: SPE 79861. SPE/IADC Drilling Conference, Amsterdam, Society of Petroleum Engineers.Google Scholar
Oort, E., Hale, A. H.et al. (1995). Manipulation of coupled osmotic flows for stabilization of shales exposed to water-based drilling fluids: SPE 30499. SPE Annual Technical Conf. and Exhibition, Dallas, Texas, Society of Petroleum Engineers.Google Scholar
Vardulakis, I., , S. J.et al. (1988). “Borehole instabilities as bifurcation phenomena.” Intl. J. Rock Mech. Min. & Geomech. Abstr., 25, 159–170.CrossRefGoogle Scholar
Veeken, C., Walters, J. et al. (1989). Use of plasticity models for predicting borehole stability. Rock at Great Depth, Vol. 2.Maury, V. and Fourmaintraux, D.. Rotterdam, Balkema, 835–844.Google Scholar
Vernik, L., Bruno, M.et al. (1993). “Empirical relations between compressive strength and porosity of siliciclastic rocks. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 30, 7, 677–680.CrossRefGoogle Scholar
Vernik, L., Lockner, D.et al. (1992). “Anisotropic strength of some typical metamorphic rocks from the KTB pilot hole, Germany.” Scientific Drilling, 3, 153–160.Google Scholar
Vernik, L. and Zoback, M. D. (1990). Strength anisotropy of crystalline rock: Implications for assessment of In situ stresses from wellbore breakouts. Rock Mechanics Contributions and Challenges. Balkema, Rotterdam.Google Scholar
Vigneresse, J., Ed. (2001). Fluids and fractures in the lithosphere. Tectonophysics. Amsterdam, Elsevier.Google Scholar
Walls, J. and Nur, A. (1979). Pore pressure and confining pressure dependence of permeability in sandstone. 7th Formation Evaluation Symposium, Calgary, Canada, Canadian Well Logging Society.Google Scholar
Wang, H. F. (2000). Theory of linear poroelasticity with applications to geomechanics and hydrogeology. Princeton, NJ, Princeton University Press.Google Scholar
Ward, C. D. and Beique, M. (1999). “How to identify Lost Circulation Problems with Real-time Pressure Measurement: Downhole Pressure Sensing heads off Deepwater Challenge.” Offshore, August.Google Scholar
Ward, C. D. and Clark, R. (1998). Bore hole ballooning diagnosis with PWD. Workshop on Overpressure, Pau, France., Elf EP-Editions.Google Scholar
Warren, W. E. and Smith, C. W. (1985). “In situ stress estimates from hydraulic fractruring and direct observation of crack orientation.” Journal of Geophysical Research, 90, 6829–6839.CrossRefGoogle Scholar
Webb, S., Anderson, T. et al. (2001). New treatments substantially increase LOT/FIT pressures to solve deep HTHP drilling challenges: SPE 71390. Annual Technical Conference and Exhibition, New Orleans, LA, Society of Petroleum Engineers.CrossRef
Weng, X. (1993). Fracture Initiation and Propagation from Deviated Wellbores. paper SPE 26597 presented at the 1993 Annual Technical Conference and Exhibition, Houston, 3–6 October.CrossRef
Whitehead, W., Hunt, E. R. et al. (1986). In-Situ Stresses: A comparison between log-derived values and actual field-measured values in the Travis Peak formation of east Texas: SPE 15209. Unconventional Gas Technology Symposium, Louisville, Kentucky, Society of Petroleum Engineers.CrossRef
Wiebols, G. A. and Cook, N. G. W. (1968). “An energy criterion for the strength of rock in polyaxial compression.” International Journal of Rock Mechanics and Mining Sciences, 5, 529–549.CrossRefGoogle Scholar
Willson, S. M., Last, N. C.et al. (1999). Drilling in South America: A wellbore stability approach for complex geologic condtions, SPE 53940. 6th LACPEC Conference, Caracas, Venezuela, Society of Petroleum Engineers.Google Scholar
Winterstein, D. F. and Meadows, M. A. (1995). “Analysis of shear-wave polarization in VSP data: A tool for reservoir development SPE 234543.” Dec., SPE Formation Evaluation, 10, No. 4, 223–231.CrossRefGoogle Scholar
Wiprut, D., Zoback, M.et al. (2000). “Constraining the full stress tensor from observations of drilling-induced tensile fractures and leak-off tests: Application to borehole stability and sand production on the Norwegian margin.” Int. J. Rock Mech. & Min. Sci., 37, 317–336.CrossRefGoogle Scholar
Wiprut, D. and Zoback, M. D. (2000). “Fault reactivation and fluid flow along a previously dormant normal fault in the northern North Sea.” Geology, 28, 595–598.2.0.CO;2>CrossRefGoogle Scholar
Wiprut, D. and Zoback, M. D. (2002). Fault reactivation, leakage potential and hydrocarbon column heights in the northern North Sea. Hydrocarbon Seal Quantification, Stavanger, Norway, Elsevier.CrossRefGoogle Scholar
Wolhart, S. L., Berumen, S. et al. (2000). Use of hydraulic fracture diagnostics to optimize fracturing jobs in the Arcabuz-Calebra field, SPE 60314. 2000 SPE Rocky Mountain Region/Low Permeability Reservoirs Symposium, Denver, CO, Society of Petroleum Engineers.
Wong, T.-f., David, C.et al. (1997). “The transition from brittle faulting to cataclastic flow in porous sandstones: Mechanical deformation.” Journal of Geophysical Research, 102(B2), 3009–3025.CrossRefGoogle Scholar
Wood, D. M. (1990). Soil behaviour and critical state soil mechanics. Cambridge, England, Cambridge University.Google Scholar
Wright, C. A. and Conant, R. A. (1995). Hydraulic fracture reorientation in primary and secondary recovery from low permeability reservoirs, SPE 30484. 1995 SPe Technical Conference and Exhibition, Dallas, TX, Society of Petroleum Engineers.Google Scholar
Wright, C. A., Stewart, D. W. et al. (1994). Reorientation of propped refracture treatments in the Lost Hills field, SPE 27896. 1994 SPE Western Regional Meeting, Long Beach, California.
Yale, D. P. (2002). Coupled geomechanics-fluid flow modeling: effects of plasticity and permeability alteration: SPE 78202. Society of Petroleum Engineers.Google Scholar
Yale, D. P. (2003). Fault and stress magnitude controls on variations in the orientation of in situ stress. Fracture and in-situ stress characterization of hydrocarbon reservoirs. M. Ameen. London, Geological Society, 209, 55–64.Google Scholar
Yale, D. P., Nabor, G. W.et al. (1993). Application of variable formation compressibility for improved reservoir analysis: SPE 26647, Society of Petroleum Engineers.CrossRefGoogle Scholar
Yale, D. P., Rodriguez, J. M., et al. (1994). In-Situ stress orientation and the effects of local structure – Scott Field, North Sea. Eurock ′94, Delft, Netherlands, Balkema.CrossRefGoogle Scholar
Yassir, N. A. and Bell, J. S. (1994). “Relationships between pore pressure, stresses and present-day geodynamics in the Scotian shelf, offshore eastern Canada.” American Association of Petroleum Geology Bulletin, 78(12), 1863–1880.Google Scholar
Yassir, N. A. and Zerwer, A. (1997). “Stress regimes in the Gulf Coast, offshore Lousiana from wellbore breakout analysis.” American Association of Petroleum Geologists Bulletin, 81(2), 293–307.Google Scholar
Yew, C. H. and Li, Y. (1988). “Fracturing of a deviated well.” SPE Production Engineering, 3, 429–437.CrossRefGoogle Scholar
Zajac, B. and Stock, J. M. (1992). “Using borehole breakouts to constrain complete stress tensor.” AGU 1992 Fall Meeting Abstract Supplement to EOS, Dec. 1992, 559.Google Scholar
Zemanek, J., Glenn, E. E.et al. (1970). “Formation evaluation by inspection with the borehole televiewer.” Geophysics, 35, 254–269.CrossRefGoogle Scholar
Zheng, Z., Kemeny, J.et al. (1989). “Analysis of borehole breakouts.” Journal of Geophysical Research, 94(B6), 7171–7182.CrossRefGoogle Scholar
Zhou, S. (1994). “A program to model initial shape and extent of borehole breakout.” Computers and Geosciences, 20, 7/8, 1143–1160.CrossRefGoogle Scholar
Zimmer, M. (2004). Controls on the seismic velocities of unconsolidated sands: Measurements of pressure, porosity and compaction effects. Geophysics. Stanford, CA., Stanford University.Google Scholar
Zinke, J. C. and Zoback, M. D. (2000). “Structure-related and stress-induced shear-wave velocity anisotropy: Observations from microearthquakes near the Calaveras fault in central California.” Bulletin of Seismological Society of America, 90, 1305–1312.CrossRefGoogle Scholar
Zoback, M. D., Apel, R.et al. (1993). “Upper crustal strength inferred from stress measurements to 6 km depth in the KTB borehole.” Nature, 365, 633–635.CrossRefGoogle Scholar
Zoback, M. D., Barton, C. B.et al. (2003). “Determination of stress orientation and magnitude in deep wells.” International Journal of Rock Mechanics and Mining Sciences, 40, 1049–1076.CrossRefGoogle Scholar
Zoback, M. D. and Byerlee, J. D. (1975). “Permeability and effective stress.” Am. Assoc.Petr. Geol. Bull., 59, 154–158.Google Scholar
Zoback, M. D. and Byerlee, J. D. (1976). “A note on the deformational behavior and permeability of crushed granite.” Int'l. J. Roch Mech., 13, 291–294.CrossRefGoogle Scholar
Zoback, M. D., Day-Lewis, A. and Kim, S.-M. (2007). Predicting changes in hydrofrac orientation in depleting oil and gas reservoirs, Patent Application Pending.Google Scholar
Zoback, M. D. and Haimson, B. C. (1982). Status of the hydraulic fracturing method for in-situ stress measurements. 23rd Symposium on Rock Mechanics, Soc. Mining Engineers, New York.
Zoback, M. D. and Haimson, B. C. (1983). “Workshop on Hydraulic Fracturing Stress Measurements.” National Academy Press, 44–54.Google Scholar
Zoback, M. D. and Harjes, H. P. (1997). “Injection induced earthquakes and crustal stress at 9 km depth at the KTB deep drilling site, Germany.” J. Geophys. Res., 102, 18477–18491.CrossRefGoogle Scholar
Zoback, M. D. and Healy, J. H. (1984). “Friction, faulting, and “in situ” stresses.” Annales Geophysicae, 2, 689–698.Google Scholar
Zoback, M. D. and Healy, J. H. (1992). “In situ stress measurements to 3.5 km depth in the Cajon Pass Scientific Research Borehole: Implications for the mechanics of crustal faulting.” J. Geophys. Res., 97, 5039–5057.CrossRefGoogle Scholar
Zoback, M. D., Mastin, L.et al. (1987). In situ stress measurements in deep boreholes using hydraulic fracturing, wellbore breakouts and Stonely wave polarization. In Rock Stress and Rock Stress Measurements. Stockholm, Sweden, Centrek Publ., Lulea.Google Scholar
Zoback, M. D., Moos, D.et al. (1985). “Well bore breakouts and In situ stress.” Journal of Geophysical Research, 90(B7), 5523–5530.CrossRefGoogle Scholar
Zoback, M. D. and Peska, P. (1995). “In situ stress and rock strength in the GBRN/DOE ‘Pathfinder’ well, South Eugene Island, Gulf of Mexico.” Jour. Petrol Tech., 37, 582–585.CrossRefGoogle Scholar
Zoback, M. D. and Pollard, D. D. (1978). Hydraulic fracture propagation and the interpretation of pressure-time records for in-situ stress determinations. 19th U.S. Symposium on Rock Mechanics, MacKay School of Mines, Univ. of Nevada, Reno, Nevada.Google Scholar
Zoback, M. D. and Townend, J. (2001). “Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere.” Tectonophysics, 336, 19–30.CrossRefGoogle Scholar
Zoback, M. D., Townend, J.et al. (2002). “Steady-state failure equilibrium and deformation of intraplate lithosphere.” International Geology Review, 44, 383–401.CrossRefGoogle Scholar
Zoback, M. D. and Zinke, J. C. (2002). “Production-induced normal faulting in the Valhall and Ekofisk oil fields.” Pure & Applied Geophysics, 159, 403–420.CrossRefGoogle Scholar
Zoback, M. D. and Zoback, M. L. (1991). Tectonic stress field of North America and relative plate motions. In The Geology of North America. Neotectonics of North America. D. B. a. o. Slemmons. Boulder, Colo, Geological Society of America, 339–366.CrossRef
Zoback, M. D., Zoback, M. L.et al. (1987). “New evidence on the state of stress of the San Andreas fault system.” Science, 238, 1105–1111.CrossRefGoogle ScholarPubMed
Zoback, M. L. (1992). “First and second order patterns of tectonic stress: The World Stress Map Project.” Journal of Geophysical Research, 97, 11,703–11,728.CrossRefGoogle Scholar
Zoback, M. L. and Mooney, W. D. (2003). “Lithospheric buoyancy and continental intraplate stress.” International Geology Review, 7, 367–390.Google Scholar
Zoback, M. L. and others, a. (1989). “Global patterns of intraplate stresses; a status report on the world stress map project of the International Lithosphere Program.” Nature, 341, 291–298.CrossRefGoogle Scholar
Zoback, M. L. and Zoback, M. D. (1980). “State of stress in the conterminous United States.” J. Geophys. Res., 85, 6113–6156.CrossRefGoogle Scholar
Zoback, M. L. and Zoback, M. D. (1989). “Tectonic stress field of the conterminous United States.” Geol. Soc. Am. Memoir., 172, 523–539.CrossRefGoogle Scholar
Zoback, M. L., Zoback, M. D.et al. (1989). “Global patterns of tectonic stress.” Nature, 341, 291–298. (28 September 1989).CrossRefGoogle Scholar
Aadnoy, B. S. (1990a). “Inversion technique to determine the In situ stress field from fracturing data.”Journal of Petroleum Science and Engineering, 4, 127–141.CrossRefGoogle Scholar
Aadnoy, B. S. (1990b). “In situ stress direction from borehole fracture traces.”Journal of Petroleum Science and Engineering, 4, 143–153.CrossRefGoogle Scholar
Abé, H., Mura, T.et al. (1976). “Growth rate of a penny-shaped crack in hydraulic fracture of rocks.”J. Geophys. Res, 81, 5335.CrossRefGoogle Scholar
Adachi, T. and Oka, F. (1982). “Constitutive equations for normally consolidated clays based on elasto-viscoplasticity.”Soils and Foundations, 22(4), 57–70.CrossRefGoogle Scholar
Addis, M. A., Cauley, M. B. et al. (2001). Brent in-fill drilling programme: Lost circulation associated with drilling depleted reservoirs, Paper number SPE/IADC 67741. SPE/IADC Drilling Conference, Amsterdam, Netherlands, Society of Petroleum Engineers.CrossRef
Alexander, L. L. and Flemings, P. B. (1995). “Geologic evolution of Plio-Pleistocene salt withdrawl minibasin: Eugene Island block 330, offshore Louisiana.”American Association of Petroleum Geologists Bulletin, 79, 1737–1756.Google Scholar
Alnes, J. R. and Lilburn, R. A. (1998). “Mechanisms for generating overpressure in sedimentary basins: A reevaluation. Discussion.”American Association of Petroleum Geologists Bulletin, 82, 2266–2269.Google Scholar
Amadei, B. and Stephansson, O. (1997). Rock Stress and its Measurement. London, Chapman & Hall.CrossRefGoogle Scholar
Anderson, E. M. (1951). The Dynamics of Faulting and Dyke Formation with Applications to Britain. Edinburgh, Oliver and Boyd.Google Scholar
Anderson, R. N., Flemings, P.et al. (1994). “In situ properties of a major Gulf of Mexico growth fault: Implications for behavior as a hydrocarbon migration pathway.”Oil and Gas Journal, 92(23), 97–104.Google Scholar
Angelier, J. (1979). “Determination of the mean principal directions of stresses for a given fault population.” Tectonophysics, 56, T17–T26.CrossRefGoogle Scholar
Angelier, J. (1984). “Tectonic analysis of fault slip data sets.” Journal of Geophysical Research, 89, 5835–5848.CrossRefGoogle Scholar
Antonellini, M. and Aydin, A. (1994). “Effect of faulting and fluid flow in porous sandstones: Geometry and spatial distribution.” American Association of Petroleum Geologists Bulletin, 79(5), 642–671.Google Scholar
Artyushkov, E. V. (1973). “Stresses in the lithosphere caused by crustal thickness inhomogeneities.” Journal of Geophysical Research, 78, 7675–7708.CrossRefGoogle Scholar
Athy, L. F. (1930). “Density, porosity and compaction of sedimentary rocks.” American Association of Petroleum Geologists Bulletin, 14, 1–24.Google Scholar
Atkinson, B. K., Ed. (1987). Fracture Mechanics of Rock. Academic Press Geology Series. London, Academic Press.Google Scholar
Baria, R., Baumgaerdner, J.et al. (1999). “European HDR research programme at Soultz-sous-Forets (France) 1987–1996.” Geothermics, 28, 655–669.CrossRefGoogle Scholar
Barton, C. A., Tessler, L. et al. (1991). Interactive analysis of borehole televiewer data. In Automated Pattern Analysis in Petroleum Exploration, Palaz, I. and Sengupta, S. K. (eds). New York, Springer Verlag.Google Scholar
Barton, C. A. and Zoback, M. D. (1992). “Self-similar distribution and properties of macroscopic fractures at depth in crystalline rock in the Cajon Pass scientific drill hole.” Journal of Geophysical Research, 97, 5181–5200.CrossRefGoogle Scholar
Barton, C. A. and Zoback, M. D. (1994). “Stress perturbations associated with active faults penetrated by boreholes: Possible evidence for near-complete stress drop and a new technique for stress magnitude measurements.” J. Geophys. Res., 99, 9373–9390.CrossRefGoogle Scholar
Barton, C. A. and Zoback, M. D. (2002). Wellbore Imaging Technologies Applied to Reservoir Geomechanics and Environmental Engineering. Geological Applications of Well Logs. M. Lovell and N. Parkinson (eds). AAPG Methods in Exploration, No. 13, 229–239.
Barton, C. A., Zoback, M. D.et al. (1988). “In situ stress orientation and magnitude at the Fenton Geothermal site, New Mexico, determined from wellbore breakouts.” Geophysical Research Letters, 15(5), 467–470.CrossRefGoogle Scholar
Barton, C. A., Zoback, M. D.et al. (1995). “Fluid flow along potentially active faults in crystalline rock.” Geology, 23, 683–686.2.3.CO;2>CrossRefGoogle Scholar
Baumgärtner, J., Carvalho, J. et al. (1989). Fracturing deviated boreholes: An experimental approach. Rock at Great Depth, Proceedings ISRM-SPE International Symposium, Elf Aquitaine, Pau, A.A.Balkema.
Baumgärtner, J., Rummel, F. et al. (1990). Hydraulic fracturing in situ stress measurements to 3 km depth in the KTB pilot hole VB. A summary of a preliminary data evaluation, in KTB Report 90–6a, 353–400.
Baumgärtner, J. and Zoback, M. D. (1989). “Interpretation of hydraulic fracturing pressure – time records using interactive analysis methods.” International Journal of Rock Mechanics and Mining Sciences & Geomechanical Abstracts, 26, 461–469.CrossRefGoogle Scholar
Bell, J. S. (1989). “Investigating stress regimes in sedimentary basins using information from oil industry wireline logs and drilling records.” Geological Applications of Wireline Logs. Special Publication 48(Geological Society of London), 305–325.Google Scholar
Bell, J. S. and Babcock, E. A. (1986). “The stress regime of the Western Canadian Basin and implications for hydrocarbon production.” Bulletin of Canadian Petroleum Geology, 34, 364–378.Google Scholar
Bell, J. S. and Gough, D. I. (1979). “Northeast-southwest compressive stress in Alberta: Evidence from oil wells.” Earth Planet. Sci. Lett., 45, 475–482.CrossRefGoogle Scholar
Bell, J. S. and Gough, D. I. (1983). The use of borehole breakouts in the study of crustal stress, in Hydraulic fracturing stress measurements. D.C, National Academy Press, Washington.Google Scholar
Berry, F. A. F. (1973). “High fluid potentials in California Coast Ranges and their tectonic significance.” American Association of Petroleum Geologists Bulletin, 57, 1219–1245.Google Scholar
Biot, M. A. (1962). “Mechanics of deformation and acoustic propagation in porous media.” Journal of Acoustic Society of America, 28, 168–191.CrossRefGoogle Scholar
Birch, F. (1961). “Velocity of compressional waves in rocks to 10 kilobars, Part 2.” J. Geophys. Res., 66, 2199–2224.CrossRefGoogle Scholar
Boness, N. and Zoback, M. D. (2004). “Stress-induced seismic velocity anisotropy and physical properties in the SAFOD Pilot hole in Parkfield, CA.” Geophysical Research Letters, 31, L15S17.CrossRefGoogle Scholar
Boness, N. and Zoback, M. D. (2006). “A multi-scale study of the mechanisms controlling shear velocity anisotropy in the San Andreas Fault Observatory at Depth.” Geophysics, 71, F131–F136.CrossRefGoogle Scholar
Bourbie, T., Coussy, O. and Zinszner, B. (1987). Acoustics of porous media. Paris, France, Editions Technip.Google Scholar
Bourgoyne, A. T. Jr, Millheim, K. K.et al. (2003). Applied Drilling Engineering. Richardson, Texas, Society of Petroleum Engineers.Google Scholar
Bowers, G. L. (1994). Pore pressure estimation from velocity data: accounting for overpressure mechanisms besides undercompaction. SPE 27488 Dallas, Texas, Society of Petroleum Engineers, 515–589.CrossRef
Brace, W. F. (1980). “Permeability of crystalline and argillaceous rocks.” Int'l. J. Rock Mech. Min. Sci. and Geomech. Abstr., 17, 241–251.CrossRefGoogle Scholar
Brace, W. F. and Kohlstedt, D. L. (1980). “Limits on lithospheric stress imposed by laboratory experiments.” J. Geophys. Res, 85, 6248–6252.CrossRefGoogle Scholar
Brace, W. F., Paulding, B. W.et al. (1966). “Dilatancy in the fracture of crystalline rocks.” Journal of Geophysical Research, 71(16), 3939–3953.CrossRefGoogle Scholar
Bradley, W. B. (1979). “Failure of Inclined Boreholes.” J. Energy Res. Tech., Trans. ASME, 102, 232.CrossRefGoogle Scholar
Bratli, R. K. and Risnes, R. (1981). “Stability and failure of sand arches.” Soc. of Petroleum Engineers Journal, (April), 236–248.CrossRefGoogle Scholar
Bratton, T., Bornemann, T. et al. (1999). Logging-while-drilling images for geomechanical, geological and petrophysical interpretations. SPWLA 40th Annual Logging Symposium, Oslo, Norway, Society of Professional Well Log Analysts.
Breckels, I. M. and Eekelen, H. A. M. (1981). “Relationship between horizontal stress and depth in sedimentary basins: Paper SPE10336, 56th Annual Fall Technical Conference.” Society of Petroleum Engineers of AIME, San Antonio, Texas, October 5–7, 1981.Google Scholar
Bredehoeft, J. D., Wolf, R. G.et al. (1976). “Hydraulic fracturing to determine the regional in situ stress field Piceance Basin colorado.” Geol. Soc. Am. Bull., 87, 250–258.2.0.CO;2>CrossRefGoogle Scholar
Brown, D. (1987). The flow of water and displacement of hydrocarbons in fractured chalk reservoirs. Fluid flow in sedimentary basins and aquifers, Geological Society London Special Publication 34. J. C. Goff and B. P. Williams. London, The Geological Society. 34, 201–218.CrossRefGoogle Scholar
Brown, K. M., Bekins, B.et al. (1994). “Heterogeneous hydrofracture development and accretionary fault dynamics.” Geology, 22, 259–262.2.3.CO;2>CrossRefGoogle Scholar
Brown, S. R. and Scholz, C. H. (1985). “Closure of random elastic surfaces in contact.” J. Geophys. Res., 90, 5531–5545.CrossRefGoogle Scholar
Brown, S. R. and Scholz, C. H. (1986). “Closure of rock joints.” J. Geophy. Res., 91, 4939–4948.CrossRefGoogle Scholar
Bruce, C. H. (1984). “Smectite dehydration: Its relation to structural development and hydrocarbon accumulation in northern Gulf of Mexico basin.” Am. Assoc. Petr. Geol. Bull., 68, 673–683.Google Scholar
Brudy, M. and Zoback, M. D. (1993). “Compressive and tensile failure of boreholes arbitrarily-inclined to principal stress axes: Application to the KTB boreholes, Germany.” International Journal Rock Mechanics Mining Sciences, 30, 1035–1038.CrossRefGoogle Scholar
Brudy, M. and Zoback, M. D. (1999). “Drilling-induced tensile wall-fractures: implications for the determination of in situ stress orientation and magnitude.” International Journal of Rock Mechanics and Mining Sciences, 136, 191–215.CrossRefGoogle Scholar
Brudy, M., Zoback, M. D.et al. (1997). “Estimation of the complete stress tensor to 8 km depth in the KTB scientific drill holes: Implications for crustal strength.” J. Geophys. Res., 102, 18,453–18,475.CrossRefGoogle Scholar
Burrus, J. (1998). Overpressure models for clastic rocks, their relation to hydrocarbon expulsion: A critical reevaluation – AAPG Memoir 70. In Abnormal pressures in hydrocarbon environments. B. E. Law, G. F. Ulmishek and V. I. Slavin (eds). Tulsa, OK, American Association of Petroleum Geologists, 35–63.
Byerlee, J. D. (1978). “Friction of rock.” Pure & Applied Geophysics, 116, 615–626.CrossRefGoogle Scholar
Carmichael, R. S. (1982). Handbook of Physical Properties of Rocks. Boca Raton, FL, CRC Press.Google Scholar
Carman, P. C. (1961). L'écoulement des Gaz á Travers le Milieux Poreux, Bibliothéque des Sciences et Techniques Nucléaires, Presses Universitaires de France, Paris, 198pp.Google Scholar
Castillo, D., Bishop, D. J.et al. (2000). “Trap integrity in the Laminaria high-Nancar trough region, Timor Sea: Prediction of fault seal failure using well-constrained stress tensors and fault surfaces interpreted from 3D seismics.” Appea Journal, 40, 151–173.CrossRefGoogle Scholar
Castillo, D. and Zoback, M. D. (1995). “Systematic stress variations in the southern San Joaquin valley and along the White Wolf fault: Implications for the rupture mechanics of the 1952 Ms 7.8 Kern County earthquake and contemporary seismicity.” Journal of Geophysical Research, 100(B4), 6249–6264.CrossRefGoogle Scholar
Castillo, D. A. and Zoback, M. D. (1994). “Systematic variations in stress state in the Southern San Joaquin Valley: Inferences based on well-bore data and contemporary seismicity.” American Association Petroleum Geologists Bulletin, 78(8), 1257–1275.Google Scholar
Cayley, G. T. (1987). Hydrocarbon migration in the central North Sea. In Petroleum Geology of North West Europe. Brooks, J. and Glennie, K. (eds). London, Graham and Trotman, 549–555.Google Scholar
Chan, A., Hagin, P. et al. (2004). Viscoplastic deformation, stress and strain paths in unconsolidated reservoir sands (Part 2): Field applications using dynamic DARS analysis: ARMA/NARMS 04–568. Gulf Rocks 2004, The Sixth North American Rock Mechanics Symposium, Houston, TX, American Rock Mechanics Association.
Chan, A. and Zoback, M. D. (2002). Deformation analysis in reservoir space (DARS): A simple formalism for prediction of reservoir deformation with depletion – SPE 78174. SPE/ISRM Rock Mechanics Conference, Irving, TX, Society of Petroleum Engineers.CrossRef
Chan, A. W. and Zoback, M. D. (2006). “The role of hydrocarbon production on land subsidence and fault reactivation in the Louisiana coastal zone.” Journal of Coastal Research, submitted.Google Scholar
Chang, C. and Haimson, B. (2000). “True triaxial strength and deformability of the German Continental Deep Drilling Program (KTB) deep hole amphibolite.” Journal of Geophysical Research, 105, 18999–19013.CrossRefGoogle Scholar
Chang, C., Moos, D.et al. (1997). “Anelasticity and dispersion in dry unconsolidated sand.” International Journal of Rock Mechanics and Mining Sciences, 34(3–4), Paper No. 048.Google Scholar
Chang, C., Zoback, M. D.et al. (2006). “Empirical relations between rock strength and physical properties in sedimentary rocks.” Journal of Petroleum Science and Engineering, 51, 223–237.CrossRefGoogle Scholar
Chapple, W. and Forsythe, D. (1979). “Earthquakes and bending of plates at trenches.” Journal of Geophysical Research, 84, 6729–6749.CrossRefGoogle Scholar
Charlez, P. A. (1991). Rock Mechanics: Theoetical Fundamentals. Paris, Editions Technip.Google Scholar
Chen, S. T. (1988). “Shear-wave logging with dipole sources.” Geophysics, 53, 659–667.CrossRefGoogle Scholar
Cheng, C. H., Jinzhong, Z.et al. (1987). “Effects of in-situ permeabilty on the propagation of Stonely (tube) waves in a borehole.” Geophysics, 52, 1279–1289.CrossRefGoogle Scholar
Chester, F. M. and Logan, J. M. (1986). “Implications for mechanical properties of brittle faults from observations of the Punchbowl fault zone, California.” Pure Appl. Geophys, 124, 79–106.CrossRefGoogle Scholar
Chester, J., Chester, F. M.et al. (2005). “Fracture energy of the Punchbowl fault, San Andreas system.” Nature, 437, 133–136.CrossRefGoogle ScholarPubMed
Cloetingh, S. and Wortel, R. (1986). “Stress in the Indo-Australian plate.” Tectonophysics, 132, 49–67.CrossRefGoogle Scholar
Colmenares, L. B. and Zoback, M. D. (2002). “A statistical evaluation of rock failure criteria constrained by polyaxial test data for five different rocks.” International Journal of Rock Mechanics and Mining Sciences, 39, 695–729.CrossRefGoogle Scholar
Colmenares, L. B. and Zoback, M. D. (2003). “Stress field and seismotectonics of northern South America.” Geology, 31, 721–724.CrossRefGoogle Scholar
Coulomb, C. A. (1773). “Sur une application des regles de maximums et minimums a quelques problemes de statistique relatifs a larchitesture, Acad. Roy.” Sci. Mem. Mech. Min Sci., 7, 343–382.Google Scholar
Crampin, S. (1985). “Evaluation of anisotropy by shear wave splitting.” Geophysics, 50, 142–152.CrossRefGoogle Scholar
Crawford, B. R. and Yale, D. P. (2002). Constitutive modeling of deformation and permeability: relationships between critical state and micromechanics: SPE 78189. Society of Petroleum Engineers.CrossRef
Daines, S. R. (1992). “Aquathermal pressuring and geopressure evaluation.” American Association of Petroleum Geologists Bulletin, 66, 931–939.Google Scholar
Daneshy, A. A. (1973). “A Study of Inclined Hydraulic Fractures.” Soc. Pet. Eng. J., 13, 61.CrossRefGoogle Scholar
Davatzes, N. C. and Aydin, A. (2003). “The formation of conjugate normal fault systems in folded sandstone by sequential jointing and shearing, Waterpocket monocline, Utah.” J. Geophys. Res, 108(B10, 2478), ETG 7–1–7–15.CrossRefGoogle Scholar
Davies, R. and Handschy, J. Eds. (2003). Fault Seals. Tulsa, OK, American Association of Petroleum Geologists.Google Scholar
Davis, S. D., Nyffenegger, P. A.et al. (1995). “The April 1993 Earthquake in South Central Texas: Was it Induced by Oil and Gas Production?”Bull. Seismol. Soc. Am., 85, 1888–1895.Google Scholar
Waal, J. A. and Smits, R. M. M. (1988). “Prediction of reservoir compaction and surface subsidence: Field application of a new model.” SPE Formation Evaluation (June), 347–356.CrossRefGoogle Scholar
Desai, C. S. and Siriwardane, H. J. (1984). Constitutive laws for engineering materials with emphasis on geologic materials. Englewood Cliffs, New Jersey, Prentice-Hall.Google Scholar
Dholakia, S. K., Aydin, A.et al. (1998). “Fault-controlled hydrocarbon pathways in the Monterey Formation, California.” Amer. Assoc. Pet. Geol. Bull., 82, 1551–1574.Google Scholar
Dickinson, G. (1953). “Geological aspects of abnormal reservoir pressures in Gulf Coast Lousiana.” American Association of Petroleum Geologists Bulletin, 37, 410–432.Google Scholar
Donath, F. A. (1966). “Experimental study of shear failure in anisotropic rock.” Bulletin of Geological Soc. America, 72, 985–990.CrossRefGoogle Scholar
Dore, A. G. and Jensen, L. N. (1996). “The impact of late Cenozoic uplift and erosion on hydrocarbon exploration: offshore Norway and some other uplifted basins.” Global and Planetary Change, 12, 415–436.CrossRefGoogle Scholar
Doser, D. I., Baker, M. R.et al. (1991). “Seismicity in the War-Wink Gas Field, West Texas, and its Relationship to Petroleum Production.” Bull. Seismol. Soc. Am., 971.Google Scholar
Drucker, D. and Prager, W. (1952). “Soil mechanics and plastic analysis or limit design.” Quantitative and Applied Mathematics, 10, 157–165.CrossRefGoogle Scholar
du Rouchet, J. (1981). “Stress fields, a key to oil migration.” American Association of Petroleum Geologists Bulletin, 74–85.Google Scholar
Dudley, J. W. I., Meyers, M. T. et al. (1994). Measuring compaction and compressibilities in unconsolidated reservoir materials via time-scaling creep. Eurock ′94, Delft, Netherlands, Balkema.CrossRef
Dugan, B. and Flemings, P. B. (1998). Pore pressure prediction from stacking velocities in the Eugene Island 330 Field (Offshore Lousiana). Chicago, Ill., Gas Research Institute, 23.
Dullien, F. A. L. (1992). Porous Media: Fluid Transport and Pore Structure. San Diego, Academic Press.Google Scholar
Dvorkin, J., Mavko, G.et al. (1995). “Squrt flow in fully saturated rocks.” Geophysics, 60, 97–107.CrossRefGoogle Scholar
Eaton, B. A. (1969). “Fracture gradient prediction and its application in oilfield operations.” Journal of Petroleum Technology, 246, 1353–1360.CrossRefGoogle Scholar
Eberhart-Phillips, D., Han, D.-H.et al. (1989). “Empirical relationships among seismic velocity, effective presure, porosity and clay content in sandstone.” Geophysics, 54, 82–89.CrossRefGoogle Scholar
Economides, M. J. and Nolte, K. G. Eds. (2000). Reservoir Simulation. West Sussex, England, John Wiley & Sons, Ltd.
Ekstrom, M. P., Dahan, C. A.et al. (1987). “Formation imaging with microelectrical scanning arrays.” The Log Analyst., 28, 294–306.Google Scholar
Engelder, T. (1987). Joints and shear fractures in rock. In Fracture Mechanics of Rock. B. K. Atkinson. London, Academic Press, 534.Google Scholar
Engelder, T. (1993). Stress regimes in the lithosphere. Princeton, New Jersey, Princeton.Google Scholar
Engelder, T. and Leftwich, J. T. (1997). A pore-pressure limit in overpressured south Texas oil and gas fields. Seals, traps and the petroleum system: AAPG Memoir 67. R. C. Surdam. Tulsa, OK, AAPG, 255–267.
Engelder, T. and Sbar, M. L. (1984). “Near-surface in situ stress: Introduction.” Journal of Geophysical Research, 89, 9321–9322.CrossRefGoogle Scholar
England, W. A., MacKenzie, A. S.et al. (1987). “The movement and entrapment of petroleum fluids in the subsurface.” Journal of the Geological Society, 144, 327–347.CrossRefGoogle Scholar
Eoff, L., Funkhauser, G. P. et al. (1999). High-density monomer system for formation consolidation/water shutoff applications: SPE 50760. International symposium on oilfield chemistry, Houston, TX, Society of Petroleum Engineers.CrossRef
Ewy, R. (1999). “Wellbore-stability predictions by use of a modified Lade criterion.” SPE Drilling and Completion, 14(2), 85–91.CrossRefGoogle Scholar
Ewy, R., Stankowich, R. J.et al. (2003). Mechanical behavior of some clays and shales from 200 m to 3800 m depth, Paper 570. 39th U.S. Rock Mechanics Symposium/12th Panamerican Conference on Soil Mechanics and Geotechnical Engineering, Cambridge, MA.Google Scholar
Færseth, R. B., Sj⊘blom, R. J.et al. (1995). “Sequence Stratigraphy on the Northwest European Margin.” Elsevier, Amsterdam.Google Scholar
Faybishenko, B., Witherspoon, P. A.et al., Eds. (2000). Dynamics of fluids in fractured rock. Geophysical Monograph Series. Washington, D.C., American Geophysical Union.CrossRefGoogle Scholar
Fehler, M., Jupe, A.et al. (2001). “More than a cloud: new techniques for characterizing reservoir structures using induced seismicity.” Leading Edge, 20, 324–328.CrossRefGoogle Scholar
Feignier, B. and Grasso, J.-R. (1990). “Seismicty Induced by Gas ProductionI: Correlation of Focal Mechanism & Dome Structure.” 134 Pure & Applied Geophys., 405.CrossRefGoogle Scholar
Finkbeiner, T. (1998). In situ stress, pore pressure and hydrocarbon migration and accumulation in sedimentary basins. Geophysics. Stanford, CA, Stanford University, 193.Google Scholar
Finkbeiner, T., Barton, C. B.et al. (1997). “Relationship between in-situ stress, fractures and faults, and fluid flow in the Monterey formation, Santa Maria basin, California.” Amer. Assoc. Petrol. Geol. Bull., 81(12), 1975–1999.Google Scholar
Finkbeiner, T., Zoback, M. D.et al. (2001). “Stress, pore pressure and dynamically-constrained hydrocarbon column heights in the south Eugene Island 330 field, Gulf of Mexico.” Amer. Assoc. Petrol. Geol. Bull., 85(June), 1007–1031.Google Scholar
Fisher, N. I., Lewis, T.et al. (1987). Statistical analysis of spherical data. Cambridge, Cambridge University Press.CrossRefGoogle Scholar
Fisher, Q. J., Casey, M.et al. (2003). “Fluid-flow properties of faults in sandstone: The importance of temperature history.” Geology, 31(11), 965–968.CrossRefGoogle Scholar
Fisher, Q. J. and Knipe, R. J. (1998). Fault sealing processes in siliciclastic sediments. In Faulting and fault sealing in hydrocarbon reservoirs, Jones, G.et al. (eds). London, Geological Society (London), 147, 117–134.
Fjaer, E., Holt, R. M.et al. (1992). Petroleum Related Rock Mechanics. Amsterdam, Elsevier.Google Scholar
Fleitout, L. and Froidevaux, C. (1983). “Tectonics and topography for a lithosphere containing density heterogenieties.” Tectonics, 2, 315–324.CrossRefGoogle Scholar
Flemings, P. B., Stump, B. B.et al. (2002). “Flow focusing in overpressured sandstones: theory, observations and applications.” American Journal of Science, 302, 827–855.CrossRefGoogle Scholar
Forsyth, D. and Uyeda, S. (1975). “On the relative importance of the driving forces of plate motion.” Geophys. J. R. Astr. Soc., 43, 163–200.CrossRefGoogle Scholar
Fowler, C. M. R. (1990). The solid earth. Cambridge, U.K., Cambridge University Press.Google Scholar
Fredrich, J. T., Coblentz, D. D. et al. (2003). Stress perturbations adjacent to salt bodies in the deepwater Gulf of Mexico: SPE 84554. SPE Annual Technical Conference and Exhibition, Denver, CO, Society of Petroleum Engineers.CrossRef
Freyburg, D. (1972). “Der Untere und mittlere Buntsandstein SW-Thuringen in seinen gesteinstechnicschen Eigenschaften.” Ber. Dte. Ges. Geol. Wiss. A; Berlin, 17(6), 911–919.Google Scholar
Fuh, G.-A., Morita, N. et al. (1992). A new approach to preventing loss circulation while drilling. SPE 24599, Soc. Petr. Eng. 67th Annual Tech. Conf. and Exhib, Washington, D.C.
Gaarenstroom, L., Tromp, R. A. J.et al. (1993). Overpressures in the Central North Sea: implications for trap integrity and drilling safety. Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference, London.Google Scholar
Geertsma, J. (1973). “A basic theory of subsidence due to reservoir compaction: the homogeneous case.” Trans. Royal Dutch Soc. of Geologists and Mining Eng., 28, 43–62.Google Scholar
Gephart, J. W. (1990). “Stress and the direction of slip on fault planes.” Tectonics, 9, 845–858.CrossRefGoogle Scholar
Gephart, J. W. and Forsyth, D. W. (1984). “An improved method for determining the regional stress tensor using earthquake focal mechanism data: application to the San Fernando earthquake sequence.” Journal of Geophysical Research, 89, 9305–9320.CrossRefGoogle Scholar
Germanovich, L. N. and Dyskin, A. V. (2000). “Fracture mechanisms and instability of openings in compression.” International Journal of Rock Mechanics and Mining Sciences, 37, 263–284.CrossRefGoogle Scholar
Germanovich, L. N., Galybin, A. N.et al. (1996). Borehole stability in laminated rock. Prediction and performance in rock mechanics and rock engineering, Torino, Italy, A. A. Balkema.Google Scholar
Golubev, A. A. and Rabinovich, G. Y. (1976). “Resultay primeneia appartury akusticeskogo karotasa dija predeleina procontych svoistv gornych porod na mestorosdeniaach tverdych isjopaemych.” Prikladnaja GeofizikaMoskva, 73, 109–116.Google Scholar
Gordon, D. S. and Flemings, P. B. (1998). “Generation of overpressure and compaction-driven fluid flow in a Plio-Pleistocene grwoth-faulted basin, Eugene Island 330, offshore Louisiana.” Basin Research, 10, 177–196.CrossRefGoogle Scholar
Grasso, J. R. (1992). “Mechanics of seismic instabilities induced the recovery of hydrocarbons.” Pure & Applied Geophysics, 139, 507–534.CrossRefGoogle Scholar
Grasso, J. R. and Wittlinger, G. (1990). “10 Years of Seismic Monitoring over a Gas Field.” 80 Bull. Seismo. Soc. Am., 450.Google Scholar
Griffith, J. (1936). Thermal Expansion of Typical American Rocks. Iowa State College of Agriculture and Mechanic Arts, Iowa Engineering Experiment, 35(19), 24.
Grollimund, B., Zoback, M. D.et al. (2001). “Regional synthesis of stress orientation, pore pressure and least principal stress data in the Norwegian sector of the North Sea.” Petroleum Geoscience, 7, 173–180.CrossRefGoogle Scholar
Grollimund, B. R. and Zoback, M. D. (2000). “Post glacial lithospheric flexure and induced stresses and pore pressure changes in the northern North Sea.” Tectonophysics, 327, 61–81.CrossRefGoogle Scholar
Grollimund, B. R. and Zoback, M. D. (2001). “Impact of glacially-induced stress changes on hydrocarbon exploration offshore Norway.” American Association of Petroleum Geologists Bulletin, 87(3), 493–506.CrossRefGoogle Scholar
Grollimund, B. R. and Zoback, M. D. (2003). “Impact of glacially induced stress changes on fault-seal integrity offshore Norway.” American Association of Petroleum Geologists Bulletin, 87, 493–506.CrossRefGoogle Scholar
Gudmundsson, A. (2000). “Fracture dimensions, displacements and fluid transport.” Journal of Structural Geology, 22, 1221–1231.CrossRefGoogle Scholar
Guenot, A. (1989). “Borehole breakouts and stress fields.” Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 26, 185–195.CrossRefGoogle Scholar
Guo, G., Morgenstern, N. R.et al. (1993). “Interpretation of hydraulic fracturing pressure: A comparison of eight methods used to identify shut-in pressures.” International Journal of Rock Mechanics and Mining Sciences, 30, 627–631.CrossRefGoogle Scholar
Hagin, P. and Zoback, M. D. (2004a). Viscoplastic deformation in unconsolidated reservoir sands (Part 1): Laboratory observations and time-dependent end cap models ARMA/NARMS 04–567. Gulf Rocks, Houston, TX, American Rock Mechanics Association.Google Scholar
Hagin, P. and Zoback, M. D. (2004b). “Viscous deformation of unconsolidated sands-Part 1: Time-dependent deformation, frequency dispersion and attenuation.” Geophysics, 69, 731–741.CrossRefGoogle Scholar
Hagin, P. and Zoback, M. D. (2004c). “Viscous deformation of unconsolidated reservoir sands-Part 2: Linear viscoelastic models.” Geophysics, 69, 742–751.CrossRefGoogle Scholar
Hagin, P. and Zoback, M. D. (2007). “Characterization of time-dependent deformation of unconsolidated reservoir sands.” Geophysics, in press.Google Scholar
Haimson, B. and Fairhurst, C. (1967). “Initiation and Extension of Hydraulic Fractures in Rocks.” Soc. Petr. Eng. Jour., Sept.: 310–318.CrossRefGoogle Scholar
Haimson, B. and Fairhurst, C. (1970). In situ stress determination at great depth by means of hydraulic fracturing. In 11th Symposium on Rock Mechanics. W. Somerton, Society of Mining Engineers of AIME, 559–584.
Haimson, B. C. (1989). “Hydraulic fracturing stress measurements.” Rock Mech. and Min. Sci. and Geomech. Abstr Special Issue: Inter. Jour., 26.Google Scholar
Haimson, B. C. and Herrick, C. G. (1989). Borehole breakouts and in situ stress. 12th Annual Energy-Sources Technology Conference and Exhibition, Houston, Texas.
Hall, P. L. (1993). Mechanisms of overpressuring-an overview. Geochemistry of clay-pore fluid interactions. Manning, D. A. C., Hall, P. L. and Hughes, C. R.. London, Chapman and Hall, 265–315.Google Scholar
Han, D., Nur, A.et al. (1986). “Effects of porosity and clay content on wave velocities in sandstones.” Geophysics, 51, 2093–2107.CrossRefGoogle Scholar
Handin, J., Hager, R. V.et al. (1963). “Experimental deformation of sedimentary rocks under confining pressure: pore pressure effects.” Bulletin American Assoc. Petrol. Geology, 717–755.Google Scholar
Haneberg, W. C., Mozley, P. S.et al., Eds. (1999). Faults and subsurface fluid flow in the shallow crust. Geophysical Monograph. Washington, D.C., American Geophysical Union.CrossRefGoogle Scholar
Haney, M. M., Snieder, R.et al. (2005). “A fault caught in the act of burping.” Nature, 437, 46.CrossRefGoogle Scholar
Harrison, A. R., Randall, C. J.et al. (1990). Acquisition and analysis of sonic waveforms from a borehole monopole and dipole source for the determination of compressional shear speeds and their relation to rock mechanic propoerties and surface seismic data – SPE 20557. SPE Annual Technical Conference and Exhibition, New Orleans.CrossRefGoogle Scholar
Harrold, T. W., Swarbrick, R. E.et al. (1999). “Pore pressure estimation from mudrock porosities in Tertiary basins, Southeast Asia.” American Association of Petroleum Geologists Bulletin, 83, 1057–1067.Google Scholar
Hart, B. S., Flemings, P. B.et al. (1995). “Porosity and pressure: Role of compaction disequilibrium in the development of geopressures in a Gulf Coast Pleistocene basin.” Geology, 23, 45–48.2.3.CO;2>CrossRefGoogle Scholar
Hayashi, K. and Haimson, B. C. (1991). “Characteristics of shut-in curves in hydraulic fracturing stress measurements and determination of in situ minimum compressive stress.” Journal of Geophysical Research, 96, 18311–18321.CrossRefGoogle Scholar
Healy, J. H., Rubey, W. W.et al. (1968). “The Denver earthquakes.” Science, 161, 1301–1310.CrossRefGoogle ScholarPubMed
Heppard, P. D., Cander, H. S. et al. (1998). Abnormal pressure and the occurrence of hydrocarbons in offshore eastern Trinidad, West Indies. In Abnormal pressures in hydrocarbon environments – AAPG Memoir 70, Law, B. E., Ulmishek, G. F. and Slavin, V. I. (eds). Tulsa, OK, American Association of Petroleum Geologists. Memoir, 70, 215–246.Google Scholar
Hickman, S. (1991). “Stress in the lithosphere and the strength of active faults, U.S. National Report International Union Geodesy and Geophys. 1987–1990.” Geophys., 29, 759–775.Google Scholar
Hickman, S., Sibson, R.et al. (1995a). “Introduction to special section: Mechanical involvement of fluids in faulting.” J. Geophys. Res., 100, 12831–12840.CrossRefGoogle Scholar
Hickman, S. and Zoback, M. D. (2004). “Stress measurements in the SAFOD pilot hole: Implications for the frictional strength of the San Andreas fault.” Geophysical Research Letters, 31, L15S12.Google Scholar
Hickman, S. H., Barton, C. A.et al. (1997). “In situ stress and fracture permeability along the Stillwater fault zone, Dixie Valley, Nevada.” Int. J. Rock Mech. and Min. Sci., 34, 3–4, Paper No. 126.Google Scholar
Hickman, S. H. and Zoback, M. D. (1983). The interpretation of hydraulic fracturing pressure-time data for in situ stress determination. Hydraulic Fracturing Measurements. Washington, D.C, National Academy Press.Google Scholar
Hoak, T. E., Klawitter, A. L.et al., Eds. (1997). Fractured Reservoirs: Characterization and Modeling. Denver, The Rocky Mountain Association of Geologists.Google Scholar
Hoek, E. and Brown, E. (1997). “Practical estimates of rock strength.” International Journal of Rock Mechanics and Mining Sciences, 34(8), 1165–1186.CrossRefGoogle Scholar
Hoek, E. and Brown, E. T. (1980). “Empirical strength criterion for rock masses.” J. Geotechnical Engineering Div., 106, 1013–1035.Google Scholar
Hofmann, R. (2006). Frequency dependent elastic and anelastic properties of clastic rocks. Geophysics. Golden, CO., Colorado School of Mines. Ph.D., 166.
Holbrook, P. W., Maggiori, D. A.et al. (1993). Real-time pore pressure and fracture gradient evaluation in all sedimentary lithologies, SPE 26791. Offshore European Conference, Aberdeen, Scotland, Society of Petroleum Engineersxs.Google Scholar
Holland, D. S., Leedy, J. B. et al. (1990). “Eugene Island Block 330 Field – U.S.A. offshore Louisiana, Structural Traps III: Tectonic Foldand Fault Traps, Atlas of Oil and Gas Fields, E. Beaumont and N. Foster (eds.). American Assoc. of Petroleum Geologists, Tulsa.” 103–143.
Holt, R. M., Flornes, O. et al. (2004). Consequences of depletion-induced stress changes on reservoir compaction and recovery. Gulf Rocks 2004, the 6th North America Rock Mechanics Symposium (NARMS): Rock Mechanics Across Borders and Disciplines – ARMA/NARMS 04–589. Houston.
Horsrud, P. (2001). “Estimating Mechanical Properties of Shale from Empirical Correlations.” SPE Drilling and Completion, 16(2), 68–73.CrossRefGoogle Scholar
Hottman, C. E., Smith, J. E.et al. (1979). “Relationship among earth stresses, pore pressure, and drilling problems offshore Gulf of Alaska.” Journal of Petroleum Technology, November, 1477–1484.CrossRefGoogle Scholar
Hubbert, M. D. and Rubey, W. W. (1959). “Role of fluid pressure in mechanics of overthrust faulting.” Geol. Soc. Am. Bull., 70, 115–205.CrossRefGoogle Scholar
Hubbert, M. K. and Willis, D. G. (1957). “Mechanics of hydraulic fracturing.” Petr. Trans. AIME, 210, 153–163.Google Scholar
Hudson, J. A. (1981). “Wave propagation and attenuation of elastic waves in material containing cracks.” Geophys. J. Roy. Astr. Soc., 64, 122–150.CrossRefGoogle Scholar
Hudson, J. A. and Priest, S. D. (1983). “Discontinuity frequency in rock masses.” Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 20(2), 73–89.CrossRefGoogle Scholar
Huffman, A. and Bowers, G. L., Eds. (2002). Pressure regimes in sedimentary basins and their prediction: AAPG memoir 76. Tulsa, OK, American Association of Petroleum Geologists.Google Scholar
Ingebritsen, S. E., Sanford, W.et al. (2006). Groundwater in Geologic Processes. Cambridge, United Kingdom, Cambridge Press.Google Scholar
Ito, T., Zoback, M. D.et al. (2001). “Utilization of mud weights in excess of the least principal stress to stabilize wellbores: Theory and practical examples.” Soc. of Petroleum Engineers Drilling and Completions, 16, 221–229.Google Scholar
Jaeger, J. C. and Cook, N. G. W. (1971). Fundamentals of Rock Mechanics. London, Chapman and Hall.Google Scholar
Jaeger, J. C. and Cook, N. G. W. (1979). Fundamentals of Rock mechanics, 2nd edn. New York, Chapman and Hall.Google Scholar
Jarosinski, M. (1998). “Contemporary stress field distortion in the Polish part of the western outer Carpathians and their basement.” Tectonophysics, 297, 91–119.CrossRefGoogle Scholar
Jizba, D. (1991). Mechanical and acoustical properties of Sandstones and shales. PhD dissentation, Stanford University.Google Scholar
Jones, G., Fisher, Q. J.et al., Eds. (1998). Faulting, fault sealing and fluid flow in hydrocarbon resrvoirs. London, Geological Society.Google Scholar
Kamb, W. B. (1959). “Ice petrofabric observations from Blue Glacier, Washington, in relation to theory and experiment.” J. Geophys. Res., 64, 1891–1910.CrossRefGoogle Scholar
Kimball, C. V. and Marzetta, T. M. (1984). “Semblance processing of borehole acoustic array data.” Geophysics, 49, 264–281.CrossRefGoogle Scholar
Kirsch, G. (1898). “Die Theorie der Elastizitat und die Bedurfnisse der Festigkeitslehre, Zeitschrift des Verlines Deutscher Ingenieure.” 42, 707.Google Scholar
Klein, R. J. and Barr, M. V. (1986). Regional state of stress in western Europe. In Proc. of International Symposium on Rock Stress and Rock Stress Measurements. Lulea, Sweden, Stockholm, Centek Publ, 694 pp.Google Scholar
Kosloff, D. and Scott, R. F. (1980). “Finite element simulation of Wilmington oilfield subsidence: I, Linear modeling.” Tectonophysics, 65, 339–368.CrossRefGoogle Scholar
Kranz, R. L., Frankel, A. D.et al. (1979). “The permeability of whole and jointed Barre granite.” Int. J. Rock Mech. Min. Sci. Geomech. Abst., 16, 225–234.CrossRefGoogle Scholar
Kristiansen, G. (1998). Geomechanical characterization of the overburden above the compacting chalk reservoir at Valhall. In Eurock ′98, SPE/ISRM Rock Mechanics in Petroleum Engineering. Trondheim, Norway, The Norwegian University of Science and Technology, 193–202.Google Scholar
Kuempel, H. J. (1991). “Poroelasticity: parameters reviewed.” Geophysical Journal International, 105, 783–799.CrossRefGoogle Scholar
Kwasniewski, M. (1989). Laws of brittle failure and of B-D transition in sandstones. Rock at Great Depth, Proceedings ISRM-SPE International Symposium, Elf Aquitaine, Pau, France, A. A. Balkema.Google Scholar
Kwon, O., Kronenberg, A. K.et al. (2001). “Permeability of Wilcox shale and its effective pressure law.” Journal of Geophysical Research, 106, 19339–19353.CrossRefGoogle Scholar
Labenski, F., Reid, P.et al. (2003). Drilling fluids approaches for control of wellbore instability in fractured formations SPE/IADC 85304. SPE/IADC Middle East Drilling Technology Conference and Exhibition, Abu Dhabi, UAE, Society of Petroleum Engineers.Google Scholar
Lachenbruch, A. H. and Sass, J. H. (1992). “Heat flow from Cajon Pass, fault strength and tectonic implications.” J. Geophys. Res., 97, 4995–5015.CrossRefGoogle Scholar
Lade, P. (1977). “Elasto-plasto stress-strain theory for cohesionless soil with curved yield surfaces.” International Journal of Solids and Structures, 13, 1019–1035.CrossRefGoogle Scholar
Lal, M. (1999). Shale stability: drilling fluid interaction and shale strength, SPE 54356. SPE Latin American and Carribean Petroleum Engineering Conference, Caracas, Venezuela, Society of Petroleum Engineering.Google Scholar
Lama, R. and Vutukuri, V. (1978). Handbook on Mechanical Properties of Rock. Clausthal, Germany, Trans Tech Publications.Google Scholar
Lashkaripour, G. R. and Dusseault, M. B. (1993). A statistical study of shale properties; relationship amnog principal shale properties. Proceedings of the Conference on Probabilistic Methods in Geotechnical Engineering, Canberra, Australia.Google Scholar
Laubach, S. E. (1997). “A method to detect natural fracture strike in sandstones.” Amer. Assoc. Petrol. Geol. Bull., 81(4), 604–623.Google Scholar
Law, B. E., Ulmishek, G. F. et al., Eds. (1998). Abnormal pressures in hydrocarbon environments. AAPG Memoir 70, American Association of Petroleum Geologists.
Lekhnitskii, S. G. (1981). Theory of elasticity of an anisotropic body. Moscow, Mir.Google Scholar
Leslie, H. D. and Randall, C. J. (1990). “Eccentric dipole sources in fluid-filled boreholes: Experimental and numerical results.” Journal of Acoustic Society of America, 87, 2405–2421.CrossRefGoogle Scholar
Li, X., Cui, L.et al. (1998). Thermoporoelastic modelling of wellbore stability in non-hydrostatic stress field. 3rd North American Rock Mechanics Symposium.Google Scholar
Ligtenberg, J. H. (2005). “Detection of fluid migration pathways in seismic data: implications for fault seal analysis.” Basin Research, 17, 141–153.CrossRefGoogle Scholar
Lindholm, C. D., Bungum, H. et al. (1995). Crustal stress and tectoncis in Norwegian regions determined from earthquake focal mechanisms. Proceedings of the Workshop on Rock Stresses in the North Sea, Trondheim, Norway.
Lockner, D. A. (1995). Rock Failure. Rock physics and phase relations. Washington, D.C., American Geophysical Union, 127–147.CrossRefGoogle Scholar
Lockner, D. A., Byerlee, J. D.et al. (1991). “Quasi-static fault growth and shear fracture energy in granite.” Nature, 350, 39–42.CrossRefGoogle Scholar
Long, C. S. and a. others (1996). Rock fractures and fluid flow. Washington, D.C., National Academy Press.Google Scholar
Lorenz, J. C., Teufel, L. W.et al. (1991). “Regional fractures I: A mechanism for the formation of regional fractures at depth in flat-lying reservoirs.” Amer. Assoc. Petrol. Geol. Bull., 75(11), 1714–1737.Google Scholar
Losh, S., Eglinton, L.et al. (1999). “Vertical and lateral fluid flow related to a large growth fault, South Eugene Island Block 330 field, Offshore Louisiana.” American Association of Petroleum Geologists Bulletin, 83(2), 244–276.Google Scholar
Lucier, A., Zoback, M. D.et al. (2006). “Geomechanical aspects of CO2 sequestration in a deep saline reservoir in the Ohio River Valley region.” Environmental Geosciences, 13(2), 85–103.CrossRefGoogle Scholar
Lund, B. and Zoback, M. D. (1999). “Orientation and magnitude of in situ stress to 6.5 km depth in the Baltic Shield.” International Journal of Rock Mechanics and Mining Sciences, 36, 169–190.CrossRefGoogle Scholar
Luo, M. and Vasseur, G. (1992). “Contributions of compaction and aquathermal pressuring to geopressure and the influence of environmental conditions.” American Association of Petroleum Geologists Bulletin, 76, 1550–1559.Google Scholar
MacKenzie, D. P. (1969). “The relationship between fault plane solutions for earthquakes and the directions of the principal stresses.” Seismological Society of Amererica Bulletin, 59, 591–601.Google Scholar
Mallman, E. P. and Zoback, M. D. (2007). “Subsidence in the Louisiana Coastal Zone due to hydrocarbon production.” Journal of Coastal Research, in press.Google Scholar
Mastin, L. (1988). “Effect of borehole deviation on breakout orientations.” J. Geophys. Res., 93(B8), 9187–9195.CrossRefGoogle Scholar
Matthews, W. R. and Kelly, J. (1967). “How to predict formation pressure and fracture gradient.” Oil and Gas Journal, February, 92–106.Google Scholar
Maury, V. and Zurdo, C. (1996). “Drilling-induced lateral shifts along pre-existing fractures: A common cause of drilling problems.” SPE Drilling and Completion (March), 17–23.CrossRefGoogle Scholar
Mavko, G., Mukerjii, T.et al. (1998). Rock Physics Handbook. Cambridge, United Kingdom (GBR), Cambridge University Press.Google Scholar
Mavko, G. and Nur, A. (1997). “The effect of percolation threshold in the Kozeny-Carman relation.” Geophysics, 622, 1480–1482.CrossRefGoogle Scholar
Maxwell, S. C. (2000). Comparison of production-induced microseismicity from Valhall and Ekofisk. Passive Seismic Method in E&P of Oil and Gas Workshop, 62nd EAGE Conference.Google Scholar
Maxwell, S. C., Urbancic, T. I.et al. (2002). Microseismic imaging of hydraulkic fracture complexity in the Barnett shale, Paper 77440. Society Petroleum Engineering Annual Technical Conference, San Antonio, TX, Society of Petroleum Engineers.Google Scholar
McGarr, A. (1991). “On a Possible Connection Between Three Major Earthquakes in California and Oil Production.” Bull. of Seismol. Soc. of Am., 948, 81.Google Scholar
McGarr, A. and Gay, N. C. (1978). “State of stress in the earth's crust.” Ann. Rev. Earth Planet. Sci., 6, 405–436.CrossRefGoogle Scholar
McLean, M. and Addis, M. A. (1990). Wellbore stability: the effect of strength criteria on mud weight recommendations. 65th Annual Technical Conference and Exhibition of the Society of Petroleum Engineers, New Orleans, Society of Petroleum Engineers.Google Scholar
McNally, G. H. N. (1987). “Estimation of coal measures rock strength using sonic and neutron logs.” Geoexploration, 24, 381–395.CrossRefGoogle Scholar
McNutt, M. K. and Menard, H. W. (1982). “Constraints on yield strength in the oceanic lithosphere derived from observations of flexure.” Geophys. J. R. Astron. Soc., 71, 363–394.CrossRefGoogle Scholar
Mereu, R. F., Brunet, J.et al. (1986). “A study of the Microearthquakes of the Gobbles Oil Field Ara of Southwestern Ontario.” Bulle. Seismol. Soc. Am., 1215, 76.Google Scholar
Michael, A. (1987). “Use of focal mechanisms to determine stress: A control study.” Journal of Geophysical Research, 92, 357–368.CrossRefGoogle Scholar
Militzer, M.a. S., P. (1973). “Einige Beitrageder geophysics zur primadatenerfassung im Bergbau.” Neue Bergbautechnik, Lipzig, 3(1), 21–25.Google Scholar
Mitchell, A. and Grauls, D. Eds. (1998). Overpressure in petroleum exploration. Pau, France, Elf-Editions.Google Scholar
Mody, F. K. and Hale, A. H. (1993). A borehole stability model to couple the mechanics and chemistry of drilling-fluid/shale interactions. SPE/IADC Drilling Conference, Amsterdam.Google Scholar
Mogi, K. (1971). “Effect of the triaxial stress system on the failure of dolomite and limestone.” Tectonophysics, 11, 111–127.CrossRefGoogle Scholar
Mollema, P. N. and Antonellini, M. A. (1996). “Compaction bands; a structural analog for anti-mode I cracks in aeolian sandstone.” Tectonophysics, 267, 209–228.CrossRefGoogle Scholar
Moore, D. E. and Lockner, D. A. (2006). Friction of the smectite clay montmorillonite: A review and interpretation of data. The Seismogenic Zone of Subduction Thrust Faults, MARGINS Theoretical and Experimental Science Series. C. M. and T. Dixon. New York, Columbia University Press. 2.Google Scholar
Moos, D., Peska, P. et al. (2003). Comprehensive wellbore stability analysis using quantitative risk assessment. Jour. Petrol. Sci. and Eng., Spec. Issue on Wellbore Stability, 38. B. S. Aadnoy and S. Ong, 97–109.
Moos, D. and Zoback, M. D. (1990). “Utilization of Observations of Well Bore Failure to Constrain the Orientation and Magnitude of Crustal Stresses: Application to Continental Deep Sea Drilling Project and Ocean Drilling Program Boreholes.” J. Geophys. Res., 95, 9305–9325.CrossRefGoogle Scholar
Moos, D. and Zoback, M. D. (1993). “State of stress in the Long Valley caldera, California.” Geology, 21, 837–840.2.3.CO;2>CrossRefGoogle Scholar
Moos, D., Zoback, M. D.et al. (1999). Feasibility study of the stability of openhole multilaterals, Cook Inlet, Alaska. 1999 SPE Mid-continent Operations Symposium, Oklahoma City, OK, Society of Petroleum Engineers.CrossRefGoogle Scholar
Morita, N., Black, A. D.et al. (1996). “Borehole Breakdown Pressure with Drilling Fluids – I. Empirical Results.” Int. J. Rock Mech. and Min. Sci., 33, 39.CrossRefGoogle Scholar
Morita, N. and McLeod, H. (1995). “Oriented perforation to prevent casing collapse for highly inclined wells.” SPE Drilling and Completion, (September), 139–145.CrossRefGoogle Scholar
Morrow, C., Radney, B.et al. (1992). Frictional strength and the effective pressure law of montmorillonite and illite clays. Fault Mechanics and Transport Properties of Rocks, Academic, San Diego, Calif.Google Scholar
Mouchet, J. P. and Mitchell, A. (1989). Abnormal pressures while drilling. Manuels techniques Elf Aquitaine, 2. Boussens, France, Elf Aquitaine.Google Scholar
Mount, V. S. and Suppe, J. (1987). “State of stress near the San Andreas fault: Implications for wrench tectonics.” Geology, 15, 1143–1146.2.0.CO;2>CrossRefGoogle Scholar
Mueller, M. C. (1991). “Prediction of lateral variability in fracture intensity using multicomponent shear-wave seismic as a precursor to horizontal drilling.” Geophysical Journal International, 107, 409–415.CrossRefGoogle Scholar
Munns, J. W. (1985). “The Valhall field: a geological overview.” Marine and Petroleum Geology, 2, 23–43.CrossRefGoogle Scholar
Murrell, S. A. F. (1965). “Effect of triaxial systems on the strength of rocks at atmospheric temperatures.” Geophysical Journal Royal Astron. Soc., 106, 231–281.CrossRefGoogle Scholar
Nakamura, K., Jacob, K. H.et al. (1977). “Volcanoes as possible indicators of tectonic stress orientation – Aleutians and Alaska.” Pure and Applied Geophysics, 115, 87–112.CrossRefGoogle Scholar
Nashaat, M. (1998). Abnormally high formation pressure and seal impacts on hydrocarbon accumulations in the Nile Delta and North Sinai basins, Egypt. Abnormal pressure in hydrocarbon environments: AAPG Memoir 70. B. E. Law, G. F. Ulmishek and V. I. Slavin. Tulsa, OK, AAPG, 161–180.
Nolte, K. G. and M. J. Economides (1989). Fracturing diagnosis using pressure analysis. Reservoir Simulation. Economides, M. J. and Nolte, K. G.. Englewood Cliffs, N.J., Prentice Hall.
Nur, A. and Byerlee, J. D. (1971). “An exact effective sress law for elastic deformation of rock with fluids.” J. Geophys. Res., 6414–6419.CrossRefGoogle Scholar
Nur, A. and Simmons, G. (1969). “Stress induced velocity anisotropy in rock: An experimental study.” J. Geophys. Res., 74, 6667–6674.CrossRefGoogle Scholar
Nur, A. and Walder, J. (1990). Time-Dependent Hydraulics of the Earth's Crust. The role of fluids in crustal processes. Washington D.C., National Research Council, 113–127.Google Scholar
Okada, Y. (1992). “Internal deformation due to shear and tensile faults in a half space.” Bulletin of Seismological Society of America, 82, 1018–1040.Google Scholar
Ortoleva, P., Ed. (1994). Basin Compartments and Seals. Tulsa, American Association of Petroleum Geologists.Google Scholar
Ostermeier, R. M. (1995). “Deepwater Gulf of Mexico turbidites – compaction effects on porosity and permeabilty.” Soc. of Petroleum Engineers Formation Evaluation, 79–85.CrossRefGoogle Scholar
Ostermeier, R. M. (2001). “Compaction effects on porosity and permeability: deepwater Gulf of Mexico turbidites.” Journal of Petroleum Technology, 53(Feb. 2001), 68–74.CrossRefGoogle Scholar
Ott, W. K. and Woods, J. D. (2003). Modern Sandface Completion Practises. Houston, Texas, World Oil.Google Scholar
Ottesen, S., Zheng, R. H.et al. (1999). Wellbore Stability Assessment Using Quantitative Risk Analysis, SPE/IADC 52864. SPE/IADC Drilling Conference, Amsterdam, The Netherlands, Society of Petroleum Engineers.Google Scholar
Paterson, M. S. and Wong, T.-f . (2005). Experimental rock deformation – The brittle field. Berlin, Springer.Google Scholar
Paul, P. and Zoback, M. D. (2006). Wellbore Stability Study for the SAFOD Borehole through the San Andreas Fault: SPE 102781. SPE Annual Technical Conference, San Antonio, TX.CrossRefGoogle Scholar
Pennington, W. D., Davis, S. D.et al. (1986). “The Evolution of Seismic Barriers and Asperities Caused by the Depressuring of Fault Planes in Oil & Gas Fields of South Texas.” Bull. Seism. Soc. Am., 188, 78.Google Scholar
Pepin, G., Gonzalez, M.et al. (2004). “Effect of drilling fluid temperature on fracture gradient.” World Oil, October, 39–48.Google Scholar
Perzyna, P. (1967). Fundamental Problems in Viscoplasticity. Advances in Applied Mechanics, 9, 244–368.Google Scholar
Peska, P. and Zoback, M. D. (1995). “Compressive and tensile failure of inclined wellbores and determination of in situ stress and rock strength.” Journal of Geophysical Research, 100(B7), 12791–12811.CrossRefGoogle Scholar
Peska, P. and Zoback, M. D. (1996). “Stress and failure of inclined boreholes SFIB v2.0: Stanford Rock Physics and Borehole Geophysics Annual Report, 57, Paper H3.” Stanford University Department of Geophysics.Google Scholar
Pine, R. J. and Batchelor, A. S. (1984). “Downward migration of shearing in jointed rock during hydraulic injections.” Int. Journ. Rock Mech. Min. Sci. & Geomech. Abst., 21(5), 249–263.CrossRefGoogle Scholar
Pine, R. J., Jupe, A.et al. (1990). An evaluation of in situ stress measurements affecting different volumes of rock in the Carnmenellis granite. Scale Effects in Rock Masses. P. d. Cunha. Rotterdam, Balkema, 269–277.Google Scholar
Plumb, R. A. and Cox, J. W. (1987). “Stress directions in eastern North America determined to 4.5 km from borehole elongation measurements.” Journal of Geophysical Research, 92, 4805–4816.CrossRefGoogle Scholar
Plumb, R. A. and Hickman, S. H. (1985). “Stress-induced borehole elongation: A comparison between the four-arm dipmeter and the borehole televiewer in the Auburn geothermal well.” Journal of Geophysical Research, 90, 5513–5521.CrossRefGoogle Scholar
Pollard, D. and Aydin, A. (1988). “Progress in understanding jointing over the past century.” Geological Society of America Bulletin, 100, 1181–1204.2.3.CO;2>CrossRefGoogle Scholar
Pollard, D. and Fletcher, R. C. (2005). Fundamentals of Structural Geology. Cambridge, United Kingdom, Cambridge University Press.Google Scholar
Pollard, D. and Segall, P. (1987). Theoretical displacements and stresses near fractures in rock: With applications to faults, joints, veins, dikes and solution surfaces. Fracture Mechanics of Rock. B. K. Atkinson, Academic Press.Google Scholar
Powley, D. E. (1990). “Pressures and hydrogeology in petroleum basins.” Earth Sci. Rev., 29, 215–226.CrossRefGoogle Scholar
Qian, W., Crossing, K. S.et al. (1994). “Corrections to “Inversion of borehole breakout orientation data by Wei Qian and L.B. Pedersen.” Journal of Geophysical Research, 99, 707–710.CrossRefGoogle Scholar
Qian, W. and Pedersen, L. B. (1991). “Inversion of borehole breakout orientation data.” Journal of Geophysical Research, 96, 20093–20107.CrossRefGoogle Scholar
Raaen, A. M. and Brudy, M. (2001). Pump in/Flowback tests reduce the estimate of horizontal in situ stress significantly, SPE 71367. SPE Annual Technical Conference, New Orleans, LA, Society of Petroleum Engineers.Google Scholar
Raleigh, C. B., Healy, J. H.et al. (1972). Faulting and crustal stress at Rangely, Colorado., Flow and Fracture of Rocks. J. C. Heard. Washington, D.C., American Geophysical Union, 275–284.Google Scholar
Raleigh, C. B., Healy, J. H.et al. (1976). “An experiment in earthquake control at Rangely, Colorado.” Science, 191, 1230–1237.CrossRefGoogle ScholarPubMed
Reid, P. and Santos, H. (2003). Novel drilling, completion and workover fluids for depleted zones: Avoiding losses, formation damage and stuck pipe: SPE/IADC 85326. SPE/IADC Middle East Drilling Technology Conference and Exhibition, Abu Dhabi, UAE, Society of Petroleum Engineers.Google Scholar
Richardson, R. (1992). “Ridge forces, absolute plate motions and the intraplate stress field.” Journal of Geophysical Research, 97(B8), 11739–11748.CrossRefGoogle Scholar
Richardson, R. M. (1981). Hydraulic fracture in arbitrarily oriented boreholes: An analytic approach. Workshop on Hydraulic Fracturing Stress Measurements, Monterey, California, National Academy Press.Google Scholar
Riis, F. (1992). “Dating and measuring of erosion, uplift and subsidence in Norway and the Norwegian shelf in glacial periods.” Norsk Geologisk Tidsskrift, 72, 325–331.Google Scholar
Ritchie, R. H. and Sakakura, A. Y. (1956). “Asymptotic expansions of solutions of the heat conduction equation in internally bounded cylindrical geometry.” J. Appl. Phys., 27, 1453–1459.CrossRefGoogle Scholar
Roegiers, J. C. and Detournay, E. (1988). Considerations on failure initiation in inclined boreholes. Key Questions in Rock Mechanics, Balkeema, Brookfield, Vermont.Google Scholar
Rogers, S. (2002). Critical stress-related permeability in fractured rocks. Fracture and in situ stress characterization of hydrocarbon reservoirs. M. Ameen. London, The Geological Society, 209, 7–16.Google Scholar
Rojas, J. C., Clark, D. E.et al. (2006). Optimized salinity delivers improved drilling performance: AADE-06-DF-HO-11. AADE 2006 Fluids Conference, Houston, Texas, American Association of Drilling Engineers.Google Scholar
Rudnicki, J. W. (1999). Alteration of regional stress by reservoirs and other inhomogeneities: Stabilizing or destabilizing? Proceedings International Congress on Rock Mechanics. G. Vouille and P. Berest, International Society of Rock Mechanics, 3, 1629–1637.Google Scholar
Rummel, F. and Hansen, J. (1989). “Interpretation of hydrofrac pressure recordings using a simple fracture mechanics simulation model, Inter. Jour. Rock Mech. and Min.” Sci. and Geomech. Abstr., 26, 483–488.Google Scholar
Rummel, F. and Winter, R. B. (1983). “Fracture mechanics as applied to hydraulic fracturing stress measurements.” Earthq. Predict. Res., 2, 33–45.Google Scholar
Rutledge, J. T., Phillips, W. S.et al. (2004). “Faulting induced by forced fluid injection and fluid flow forced by faulting: An interpretation of hydraulic-fracture microseismicity, Carthage Cotten Valley Gas Field, Texas.” Bulletin of the Seismological Society of America, 94(5), 1817–1830.CrossRefGoogle Scholar
Rzhevsky, V. and Novick, G. (1971). The physics of rocks. Moscow, Russia, Mir.Google Scholar
Sayers, C. M. (1994). “The elastic anisotropy of shales.” J. Geophys. Res., 99, 767–774.CrossRefGoogle Scholar
Schmitt, D. R. and Zoback, M. D. (1992). “Diminished pore pressure in low-porosity crystalline rock under tensional failure: Apparent strengthening by dilatancy.” J. Geophys. Res., 97, 273–288.CrossRefGoogle Scholar
Schowalter, T. T. (1979). “Mechanics of Secondary Hydrocarbon Migration and Entrapment.” American Association of Petroleum Geologists Bulletin, 63(5), 723–760.Google Scholar
Schutjens, P. M. T. M., Hanssen, T. H. et al. (2001). Compaction-induced porosity/permeability reduction in sandstone reservoirs: Data and model for elasticity-dominated deformation, SPE 71337. SPE Annual Technical Conference and Exhibition, New Orleans, LA, Society of Petroleum Engineers.CrossRef
Secor, D. T. (1965). “Role of fluid pressure in jointing.” American Journal of Science, 263, 633–646.CrossRefGoogle Scholar
Segall, P. (1985). “Stress and Subsidence from Subsurface Fluid Withdrawal in the Epicentral Region of the 1983 Coalinga Earthquake.” J. Geophys. Res., 6801, 90.Google Scholar
Segall, P. (1989). “Earthquakes Triggered by Fluid Extraction.” Geology, 17, 942–946.2.3.CO;2>CrossRefGoogle Scholar
Segall, P. and Fitzgerald, S. D. (1996). “A note on induced stress changes in hydrocarbon and geothermal reservoirs.” Tectonophysics, 289, 117–128.CrossRefGoogle Scholar
Seldon, B. and Flemings, P. B. (2005). “Reservoir pressure and seafloor venting: Predicting trap integrity in a Gulf of Mexico deepwater turbidite minibasin.” American Association of Petroleum Geologists Bulletin, 89(2), 193–209.CrossRefGoogle Scholar
Shamir, G. and Zoback, M. D. (1992). “Stress orientation profile to 3.5 km depth near the San Andreas Fault at Cajon Pass California.” Jour. Geophys Res., 97, 5059–5080.CrossRefGoogle Scholar
Sibson, R. H. (1992). “Implications of fault valve behavior for rupture nucleation and recurrence.” Tectonophysics, 211, 283–293.CrossRefGoogle Scholar
Sinha, B. K. and Kostek, S. (1996). “Stress-induced azimuthal anisotropy in borehole flexural waves.” Geophysics, 61, 1899–1907.CrossRefGoogle Scholar
Sinha, B. K., Norris, A. N.et al. (1994). “Borehole flexural modes in anisotropic formations.” Geophysics, 59, 1037–1052.CrossRefGoogle Scholar
Sonder, L. (1990). “Effects of density contrasts on the orientation of stresses in the lithosphere: Relation to principal stress directions in the Transverse Ranges, California.” Tectonics, 9, 761–771.CrossRefGoogle Scholar
Stein, R. S., King, G. C.et al. (1992). “Change in failure stress on the southern San Andreas fault system caused by the 1992 magnitude 7.4 Landers earthquake.” Science, 258, 1328–1332.CrossRefGoogle ScholarPubMed
Stein, S. and Klosko, E. (2002). Earthquake mechanisms and plate tectonics. International Handbook of Earthquake and Engineering Seismology Part A. Lee, W. H. K., Kanamori, H., Jennings, P. C. and Kisslinger, K.. Amsterdam, Academic Press, 933.Google Scholar
Stephens, G. and Voight, B. (1982). “Hydraulic fracturing theory for conditions of thermal stress.” International Journal of Rock Mechanics and Mining Sciences, 19, 279–284.CrossRefGoogle Scholar
Sternlof, K. R., Karimi-Fard, M., Pollard, D. D. and Durlofsky, L. J. (2006). “Flow and transport effects of compaction bands in sandstone at scales relevant to aquifer and reservoir management.” Water Resources Research, 42, Wo7425.CrossRefGoogle Scholar
Sternlof, K. R., Rudnicki, J. W.et al. (2005). “Anticrack inclusion model for compaction bands in sandstone.” J. Geophys. Res, 110(B11403), 1–16.CrossRefGoogle Scholar
Stock, J. M., Healy, J. H.et al. (1985). “Hydraulic fracturing stress measurements at Yucca Mountain, Nevada, and relationship to the regional stress field.” Journal of Geophysical Research, 90(B10), 8691–8706.CrossRefGoogle Scholar
Stump, B. B. (1998). Illuminating basinal fluid flow in Eugene Island 330 (Gulf of Mexico) through in situ observations, deformation experiments, and hydrodynamic modeling. Geosciences, Pennsylvania State, 121.
Sulak, R. M. (1991). “Ekofisk field: The first 20 years.” Journal of Petroleum Technology, 33, 1265–1271.CrossRefGoogle Scholar
Swarbrick, R. E. and Osborne, M. J. (1998). Mechanisms that generate abnormal pressures: An overview. Abnormal pressures in hydrocarbon environments, AAPG Memoir 70. B. E. Law, G. F. Ulmishek and V. I. Slavin. Tulsa, OK, American Association of Petroleum Geologists, 13–34.
Takahashi, M. and Koide, H. (1989). Effect of the intermediate principal stress on strength and deformation behavior of sedimentary rock at the depth shallower than 2000 m. Rock at Great Depth, Pau, France, Balkema, Rotterdam.
Tang, X. M. and Cheng, C. H. (1996). “Fast inversion of formation permeability from borehole Stonely wave logs.” Geophysics, 61, 639–645.CrossRefGoogle Scholar
Terzaghi, K. (1923). Theoretical Soil Mechanics. John Wiley, New York.Google Scholar
Teufel, L. W. (1992). Production-induced changes in reservoir stress state: Applications to reservoir management. Society of Exploration Geophysicists 62nd Annual International Meeting, New Orleans, SEG, Tulsa.Google Scholar
Teufel, L. W., Rhett, D. W.et al. (1991). Effect of reservoir depletion and pore pressure drawdown on In situ stress and deformation in the Ekofisk field, North Sea. Rock Mechanics as a Multidisciplinary Science. J. C. Roegiers. Rotterdam, Balkema.Google Scholar
Tezuka, K. (2006). Hydraulic injection and microseismic monitoring in the basement gas reservoir in Japan. 2006 SPE Forum Series in Asia Pacific – Hydraulic Fracturing Beyond 2010, Macau, China, Society of Petroleum Engineers.
Thompson, A. L. (1993). Poly3D: A three-dimensional, polygonal element, displacement discontinuity boundary element computer program iwth applications to fractures, faults and cavities in the earth's crust. Geology. Stanford, CA, Stanford University.Google Scholar
Thomsen, L. (1986). “Weak elastic anisotropy.” Geophysics, 51, 1954–1966.CrossRefGoogle Scholar
Toublanc, A., Renaud, S.et al. (2005). “Ekofisk field: fracture permeability evaluation and implementation in the flow model.” Petroleum Geoscience, 11, 321–330.CrossRefGoogle Scholar
Townend, J. (2003). Mechanical constraints on the strength of the lithosphere and plate-bounding faults. Geophysics. Stanford, CA, Stanford University. PhD, 135.Google Scholar
Townend, J. and Zoback, M. D. (2000). “How faulting keeps the crust strong.” Geology, 28(5), 399–402.2.0.CO;2>CrossRefGoogle Scholar
Townend, J. and Zoback, M. D. (2001). Implications of earthquake focal mechanisms for the frictional strength of the San Andreas fault system. The Nature and Tectonic Significance of Fault Zone Weakening. R. E. Holdsworth, R. A. Strachan, J. J. Macloughlin and R. J. Knipe. London, Special Publication of the Geological Society of London, 186, 13–21.Google Scholar
Townend, J. and Zoback, M. D. (2004). “Regional tectonic stress near the San Andreas fault in Central and Northern California.” Geophysical Research Letters, 31, L15–18.CrossRefGoogle Scholar
Traugott, M. O. and Heppard, P. D. (1994). Prediction of pore pressure before and after drilling- taking the risk out of drilling overpressured prospects. AAPG Hedberg Research Conference, American Association of Petroleum Geologists.Google Scholar
Tsvankin, I. (2001). Seismic Signatures and Analysis of Reflection Data in Anisotropic Media. Cambridge, MA, Elsevier Science.Google Scholar
Turcotte, D. L. and Schubert, G. (2002). Geodynamics. Cambridge, Cambridge.CrossRefGoogle Scholar
Tutuncu, A. N., Podio, A. L.et al. (1998). “Nonlinear viscoelastic behavior of sedimentary rocks: Part I, Effect of frequency and strain amplitude.” Geophysics, 63(1), 184–194.CrossRefGoogle Scholar
Tutuncu, A. N., Podio, A. L.et al. (1998). “Nonlinear viscoelastic behavior of sedimentary rocks: Part II, Hysteresis effects and influence of type of fluid on elastic moduli.” Geophysics, 63(1), 195–203.CrossRefGoogle Scholar
Twiss, R. J. and Moores, E. M. (1992). Structural Geology. New York, W. H. Freeman and Company.Google Scholar
Balen, R. T. and Cloetingh, S. A. (1993). Stress-induced fluid flow in rifted basins. Diagenesis and basin development. A. D. Horbury and A. G. Robinson. Tulsa, American Association of Petroleum Geologists, 36, 87–98.Google Scholar
Oort, E., Gradisher, J.et al. (2003). Accessing deep reservoirs by drilling severly depleted formations: SPE 79861. SPE/IADC Drilling Conference, Amsterdam, Society of Petroleum Engineers.Google Scholar
Oort, E., Hale, A. H.et al. (1995). Manipulation of coupled osmotic flows for stabilization of shales exposed to water-based drilling fluids: SPE 30499. SPE Annual Technical Conf. and Exhibition, Dallas, Texas, Society of Petroleum Engineers.Google Scholar
Vardulakis, I., , S. J.et al. (1988). “Borehole instabilities as bifurcation phenomena.” Intl. J. Rock Mech. Min. & Geomech. Abstr., 25, 159–170.CrossRefGoogle Scholar
Veeken, C., Walters, J. et al. (1989). Use of plasticity models for predicting borehole stability. Rock at Great Depth, Vol. 2.Maury, V. and Fourmaintraux, D.. Rotterdam, Balkema, 835–844.Google Scholar
Vernik, L., Bruno, M.et al. (1993). “Empirical relations between compressive strength and porosity of siliciclastic rocks. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr., 30, 7, 677–680.CrossRefGoogle Scholar
Vernik, L., Lockner, D.et al. (1992). “Anisotropic strength of some typical metamorphic rocks from the KTB pilot hole, Germany.” Scientific Drilling, 3, 153–160.Google Scholar
Vernik, L. and Zoback, M. D. (1990). Strength anisotropy of crystalline rock: Implications for assessment of In situ stresses from wellbore breakouts. Rock Mechanics Contributions and Challenges. Balkema, Rotterdam.Google Scholar
Vigneresse, J., Ed. (2001). Fluids and fractures in the lithosphere. Tectonophysics. Amsterdam, Elsevier.Google Scholar
Walls, J. and Nur, A. (1979). Pore pressure and confining pressure dependence of permeability in sandstone. 7th Formation Evaluation Symposium, Calgary, Canada, Canadian Well Logging Society.Google Scholar
Wang, H. F. (2000). Theory of linear poroelasticity with applications to geomechanics and hydrogeology. Princeton, NJ, Princeton University Press.Google Scholar
Ward, C. D. and Beique, M. (1999). “How to identify Lost Circulation Problems with Real-time Pressure Measurement: Downhole Pressure Sensing heads off Deepwater Challenge.” Offshore, August.Google Scholar
Ward, C. D. and Clark, R. (1998). Bore hole ballooning diagnosis with PWD. Workshop on Overpressure, Pau, France., Elf EP-Editions.Google Scholar
Warren, W. E. and Smith, C. W. (1985). “In situ stress estimates from hydraulic fractruring and direct observation of crack orientation.” Journal of Geophysical Research, 90, 6829–6839.CrossRefGoogle Scholar
Webb, S., Anderson, T. et al. (2001). New treatments substantially increase LOT/FIT pressures to solve deep HTHP drilling challenges: SPE 71390. Annual Technical Conference and Exhibition, New Orleans, LA, Society of Petroleum Engineers.CrossRef
Weng, X. (1993). Fracture Initiation and Propagation from Deviated Wellbores. paper SPE 26597 presented at the 1993 Annual Technical Conference and Exhibition, Houston, 3–6 October.CrossRef
Whitehead, W., Hunt, E. R. et al. (1986). In-Situ Stresses: A comparison between log-derived values and actual field-measured values in the Travis Peak formation of east Texas: SPE 15209. Unconventional Gas Technology Symposium, Louisville, Kentucky, Society of Petroleum Engineers.CrossRef
Wiebols, G. A. and Cook, N. G. W. (1968). “An energy criterion for the strength of rock in polyaxial compression.” International Journal of Rock Mechanics and Mining Sciences, 5, 529–549.CrossRefGoogle Scholar
Willson, S. M., Last, N. C.et al. (1999). Drilling in South America: A wellbore stability approach for complex geologic condtions, SPE 53940. 6th LACPEC Conference, Caracas, Venezuela, Society of Petroleum Engineers.Google Scholar
Winterstein, D. F. and Meadows, M. A. (1995). “Analysis of shear-wave polarization in VSP data: A tool for reservoir development SPE 234543.” Dec., SPE Formation Evaluation, 10, No. 4, 223–231.CrossRefGoogle Scholar
Wiprut, D., Zoback, M.et al. (2000). “Constraining the full stress tensor from observations of drilling-induced tensile fractures and leak-off tests: Application to borehole stability and sand production on the Norwegian margin.” Int. J. Rock Mech. & Min. Sci., 37, 317–336.CrossRefGoogle Scholar
Wiprut, D. and Zoback, M. D. (2000). “Fault reactivation and fluid flow along a previously dormant normal fault in the northern North Sea.” Geology, 28, 595–598.2.0.CO;2>CrossRefGoogle Scholar
Wiprut, D. and Zoback, M. D. (2002). Fault reactivation, leakage potential and hydrocarbon column heights in the northern North Sea. Hydrocarbon Seal Quantification, Stavanger, Norway, Elsevier.CrossRefGoogle Scholar
Wolhart, S. L., Berumen, S. et al. (2000). Use of hydraulic fracture diagnostics to optimize fracturing jobs in the Arcabuz-Calebra field, SPE 60314. 2000 SPE Rocky Mountain Region/Low Permeability Reservoirs Symposium, Denver, CO, Society of Petroleum Engineers.
Wong, T.-f., David, C.et al. (1997). “The transition from brittle faulting to cataclastic flow in porous sandstones: Mechanical deformation.” Journal of Geophysical Research, 102(B2), 3009–3025.CrossRefGoogle Scholar
Wood, D. M. (1990). Soil behaviour and critical state soil mechanics. Cambridge, England, Cambridge University.Google Scholar
Wright, C. A. and Conant, R. A. (1995). Hydraulic fracture reorientation in primary and secondary recovery from low permeability reservoirs, SPE 30484. 1995 SPe Technical Conference and Exhibition, Dallas, TX, Society of Petroleum Engineers.Google Scholar
Wright, C. A., Stewart, D. W. et al. (1994). Reorientation of propped refracture treatments in the Lost Hills field, SPE 27896. 1994 SPE Western Regional Meeting, Long Beach, California.
Yale, D. P. (2002). Coupled geomechanics-fluid flow modeling: effects of plasticity and permeability alteration: SPE 78202. Society of Petroleum Engineers.Google Scholar
Yale, D. P. (2003). Fault and stress magnitude controls on variations in the orientation of in situ stress. Fracture and in-situ stress characterization of hydrocarbon reservoirs. M. Ameen. London, Geological Society, 209, 55–64.Google Scholar
Yale, D. P., Nabor, G. W.et al. (1993). Application of variable formation compressibility for improved reservoir analysis: SPE 26647, Society of Petroleum Engineers.CrossRefGoogle Scholar
Yale, D. P., Rodriguez, J. M., et al. (1994). In-Situ stress orientation and the effects of local structure – Scott Field, North Sea. Eurock ′94, Delft, Netherlands, Balkema.CrossRefGoogle Scholar
Yassir, N. A. and Bell, J. S. (1994). “Relationships between pore pressure, stresses and present-day geodynamics in the Scotian shelf, offshore eastern Canada.” American Association of Petroleum Geology Bulletin, 78(12), 1863–1880.Google Scholar
Yassir, N. A. and Zerwer, A. (1997). “Stress regimes in the Gulf Coast, offshore Lousiana from wellbore breakout analysis.” American Association of Petroleum Geologists Bulletin, 81(2), 293–307.Google Scholar
Yew, C. H. and Li, Y. (1988). “Fracturing of a deviated well.” SPE Production Engineering, 3, 429–437.CrossRefGoogle Scholar
Zajac, B. and Stock, J. M. (1992). “Using borehole breakouts to constrain complete stress tensor.” AGU 1992 Fall Meeting Abstract Supplement to EOS, Dec. 1992, 559.Google Scholar
Zemanek, J., Glenn, E. E.et al. (1970). “Formation evaluation by inspection with the borehole televiewer.” Geophysics, 35, 254–269.CrossRefGoogle Scholar
Zheng, Z., Kemeny, J.et al. (1989). “Analysis of borehole breakouts.” Journal of Geophysical Research, 94(B6), 7171–7182.CrossRefGoogle Scholar
Zhou, S. (1994). “A program to model initial shape and extent of borehole breakout.” Computers and Geosciences, 20, 7/8, 1143–1160.CrossRefGoogle Scholar
Zimmer, M. (2004). Controls on the seismic velocities of unconsolidated sands: Measurements of pressure, porosity and compaction effects. Geophysics. Stanford, CA., Stanford University.Google Scholar
Zinke, J. C. and Zoback, M. D. (2000). “Structure-related and stress-induced shear-wave velocity anisotropy: Observations from microearthquakes near the Calaveras fault in central California.” Bulletin of Seismological Society of America, 90, 1305–1312.CrossRefGoogle Scholar
Zoback, M. D., Apel, R.et al. (1993). “Upper crustal strength inferred from stress measurements to 6 km depth in the KTB borehole.” Nature, 365, 633–635.CrossRefGoogle Scholar
Zoback, M. D., Barton, C. B.et al. (2003). “Determination of stress orientation and magnitude in deep wells.” International Journal of Rock Mechanics and Mining Sciences, 40, 1049–1076.CrossRefGoogle Scholar
Zoback, M. D. and Byerlee, J. D. (1975). “Permeability and effective stress.” Am. Assoc.Petr. Geol. Bull., 59, 154–158.Google Scholar
Zoback, M. D. and Byerlee, J. D. (1976). “A note on the deformational behavior and permeability of crushed granite.” Int'l. J. Roch Mech., 13, 291–294.CrossRefGoogle Scholar
Zoback, M. D., Day-Lewis, A. and Kim, S.-M. (2007). Predicting changes in hydrofrac orientation in depleting oil and gas reservoirs, Patent Application Pending.Google Scholar
Zoback, M. D. and Haimson, B. C. (1982). Status of the hydraulic fracturing method for in-situ stress measurements. 23rd Symposium on Rock Mechanics, Soc. Mining Engineers, New York.
Zoback, M. D. and Haimson, B. C. (1983). “Workshop on Hydraulic Fracturing Stress Measurements.” National Academy Press, 44–54.Google Scholar
Zoback, M. D. and Harjes, H. P. (1997). “Injection induced earthquakes and crustal stress at 9 km depth at the KTB deep drilling site, Germany.” J. Geophys. Res., 102, 18477–18491.CrossRefGoogle Scholar
Zoback, M. D. and Healy, J. H. (1984). “Friction, faulting, and “in situ” stresses.” Annales Geophysicae, 2, 689–698.Google Scholar
Zoback, M. D. and Healy, J. H. (1992). “In situ stress measurements to 3.5 km depth in the Cajon Pass Scientific Research Borehole: Implications for the mechanics of crustal faulting.” J. Geophys. Res., 97, 5039–5057.CrossRefGoogle Scholar
Zoback, M. D., Mastin, L.et al. (1987). In situ stress measurements in deep boreholes using hydraulic fracturing, wellbore breakouts and Stonely wave polarization. In Rock Stress and Rock Stress Measurements. Stockholm, Sweden, Centrek Publ., Lulea.Google Scholar
Zoback, M. D., Moos, D.et al. (1985). “Well bore breakouts and In situ stress.” Journal of Geophysical Research, 90(B7), 5523–5530.CrossRefGoogle Scholar
Zoback, M. D. and Peska, P. (1995). “In situ stress and rock strength in the GBRN/DOE ‘Pathfinder’ well, South Eugene Island, Gulf of Mexico.” Jour. Petrol Tech., 37, 582–585.CrossRefGoogle Scholar
Zoback, M. D. and Pollard, D. D. (1978). Hydraulic fracture propagation and the interpretation of pressure-time records for in-situ stress determinations. 19th U.S. Symposium on Rock Mechanics, MacKay School of Mines, Univ. of Nevada, Reno, Nevada.Google Scholar
Zoback, M. D. and Townend, J. (2001). “Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere.” Tectonophysics, 336, 19–30.CrossRefGoogle Scholar
Zoback, M. D., Townend, J.et al. (2002). “Steady-state failure equilibrium and deformation of intraplate lithosphere.” International Geology Review, 44, 383–401.CrossRefGoogle Scholar
Zoback, M. D. and Zinke, J. C. (2002). “Production-induced normal faulting in the Valhall and Ekofisk oil fields.” Pure & Applied Geophysics, 159, 403–420.CrossRefGoogle Scholar
Zoback, M. D. and Zoback, M. L. (1991). Tectonic stress field of North America and relative plate motions. In The Geology of North America. Neotectonics of North America. D. B. a. o. Slemmons. Boulder, Colo, Geological Society of America, 339–366.CrossRef
Zoback, M. D., Zoback, M. L.et al. (1987). “New evidence on the state of stress of the San Andreas fault system.” Science, 238, 1105–1111.CrossRefGoogle ScholarPubMed
Zoback, M. L. (1992). “First and second order patterns of tectonic stress: The World Stress Map Project.” Journal of Geophysical Research, 97, 11,703–11,728.CrossRefGoogle Scholar
Zoback, M. L. and Mooney, W. D. (2003). “Lithospheric buoyancy and continental intraplate stress.” International Geology Review, 7, 367–390.Google Scholar
Zoback, M. L. and others, a. (1989). “Global patterns of intraplate stresses; a status report on the world stress map project of the International Lithosphere Program.” Nature, 341, 291–298.CrossRefGoogle Scholar
Zoback, M. L. and Zoback, M. D. (1980). “State of stress in the conterminous United States.” J. Geophys. Res., 85, 6113–6156.CrossRefGoogle Scholar
Zoback, M. L. and Zoback, M. D. (1989). “Tectonic stress field of the conterminous United States.” Geol. Soc. Am. Memoir., 172, 523–539.CrossRefGoogle Scholar
Zoback, M. L., Zoback, M. D.et al. (1989). “Global patterns of tectonic stress.” Nature, 341, 291–298. (28 September 1989).CrossRefGoogle 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
  • Mark D. Zoback, Stanford University, California
  • Book: Reservoir Geomechanics
  • Online publication: 10 December 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511586477.014
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
  • Mark D. Zoback, Stanford University, California
  • Book: Reservoir Geomechanics
  • Online publication: 10 December 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511586477.014
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
  • Mark D. Zoback, Stanford University, California
  • Book: Reservoir Geomechanics
  • Online publication: 10 December 2009
  • Chapter DOI: https://doi.org/10.1017/CBO9780511586477.014
Available formats
×