Skip to main content Accessibility help
×
Hostname: page-component-7479d7b7d-68ccn Total loading time: 0 Render date: 2024-07-10T02:43:48.216Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  03 August 2019

Richard A. Schultz
Affiliation:
Orion Geomechanics LLC
Get access
Type
Chapter
Information
Publisher: Cambridge University Press
Print publication year: 2019

Access options

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

References

Aadnøy, B. S. and Looyeh, R. (2010). Petroleum Rock Mechanics: Drilling Operations and Well Design, Elsevier, Amsterdam.Google Scholar
Abercrombie, R. W. and Ekström, G. (2001). Earthquake slip on oceanic transform faults. Nature 410:7476.CrossRefGoogle ScholarPubMed
Abou-Sayed, A. S. (1977). Fracture toughness, KIc, of triaxially-loaded Indiana limestone. Proc. US Symp. Rock Mech. 18:2A3/12A3/8.Google Scholar
Ackermann, R. V., Schlische, R. W. and Withjack, M. O. (2001). The geometric and statistical evolution of normal fault systems: an experimental study of the effects of mechanical layer thickness on scaling laws. J. Struct. Geol. 23:18031819.Google Scholar
Acocella, V., Gudmundsson, A. and Funiciello, R. (2000). Interaction and linkage of extension fractures and normal faults: examples from the rift zone of Iceland. J. Struct. Geol. 22:12331246.CrossRefGoogle Scholar
Aharonov, E. and Katsman, R. (2009). Interaction between pressure solution and clays in stylolite development: insights from modeling. Am. J. Sci. 309:607632.Google Scholar
Ahlgren, S. G. (2001). The nucleation and evolution of Riedel shear zones as deformation bands in porous sandstone. J. Struct. Geol. 23:12031214.Google Scholar
Aki, K. and Richards, P. G. (1980). Quantitative Seismology: Theory and Methods, Vol. I, W. H. Freeman, San Francisco, California.Google Scholar
Albert, R. A., Phillips, R. J., Dombard, A. J. and Brown, C. D. (2002). A test of the validity of yield strength envelopes with an elastoviscoplastic finite element method. Geophys. J. Int. 140:399409.Google Scholar
Al-Chalabi, M. and Huang, C. L. (1974). Stress distribution within circular cylinders in compression. Int. J. Rock Mech. Min. Sci. 11:4556.CrossRefGoogle Scholar
Aliha, M. R. M., Ayatollahi, M. R., Smith, D. J. and Pavier, M. J. (2010). Geometry and size effects on fracture trajectory in a limestone rock under mixed mode loading. Eng. Fracture Mech. 77:22002212.Google Scholar
Allmendinger, R. W. (1998). Inverse and forward numerical modeling of trishear fault-propagation folds. Tectonics 17:640656.Google Scholar
Altman, S. J., Aminzadeh, B., Balhoff, M. T., et al. (2014). Chemical and hydrodynamic mechanisms for long-term geological carbon storage. J. Phys. Chem. 118:15,10315,113.Google Scholar
Ameen, M. S. (1995). Fractography and fracture characterization in the Permo-Triassic sandstones and the Lower Palaeozoic Basement, West Cumbria, UK. In Fractography: Fracture Topography as a Tool in Fracture Mechanics and Stress Analysis (ed. Ameen, M. S.), pp. 97147, Geol. Soc. Spec. Publ. 92.Google Scholar
Anders, M. H. and Schlische, R. W. (1994). Overlapping faults, intrabasin highs, and the growth of normal faults. J. Geol. 102:165179.Google Scholar
Anderson, E. M. (1938). The dynamics of sheet intrusion. Proc. Royal Soc. Edinburgh 58:242251.CrossRefGoogle Scholar
Anderson, E. M. (1951). The Dynamics of Faulting and Dyke Formation, with Applications to Britain, Oliver & Boyd, Edinburgh.Google Scholar
Anderson, O. L. and Grew, P. C. (1977). Stress corrosion theory of crack propagation with applications to geophysics. Rev. Geophys. 15:77104.Google Scholar
Anderson, R. E. (1973). Large-magnitude late Tertiary strike-slip faulting north of Lake Mead, Nevada. US Geol. Surv. Prof. Pap. 794, 18 pp.Google Scholar
Anderson, T. L. (1995). Fracture Mechanics: Fundamentals and Applications (2nd edn), CRC Press, Boca Raton, Florida.Google Scholar
Anderson, T. L. (2017). Fracture Mechanics: Fundamentals and Applications (4th edn), CRC Press, Boca Raton, Florida, 688 pp.CrossRefGoogle Scholar
Andrade, J. E., Avila, C. F., Hall, S. A., Lenoir, N. and Viggiani, G. (2011). Multiscale modeling and characterization of granular matter: from grain kinematics to continuum mechanics. J. Mech. Phys. Solids 59:237250.Google Scholar
Andrews-Hanna, J. C., Zuber, M. T. and Hauck, S. A., II (2008). Strike-slip faults on Mars: Observations and implications for global tectonics and geodynamics. J. Geophys. Res. 113:E08002; doi:10.1029/2007JE002980.Google Scholar
Angelier, J. (1989). From orientation to magnitudes in paleostress determinations using fault slip data. J. Struct. Geol. 11:3750.Google Scholar
Angelier, J. (1994). Fault slip analysis and paleostress reconstruction. In Continental Deformation (ed. Hancock, P. L.), pp. 53100, Pergamon, New York.Google Scholar
Angelier, J., Colletta, B. and Anderson, R. E. (1985). Neogene paleostress changes in the Basin and Range: a case study at Hoover Dam, Nevada-Arizona. Geol. Soc. Am. Bull. 96:347361.Google Scholar
Antonellini, M. and Aydin, A. (1994). Effect of faulting on fluid flow in porous sandstones: geometry and spatial distribution. Am. Assoc. Petrol. Geol. Bull. 78:355377.Google Scholar
Antonellini, M. and Pollard, D. D. (1995). Distinct element modeling of deformation bands in sandstone. J. Struct. Geol. 17:11651182.CrossRefGoogle Scholar
Antonellini, M., Aydin, A. and Pollard, D. D. (1994). Microstructure of deformation bands in porous sandstones at Arches National Park, Utah. J. Struct. Geol. 16:941959.Google Scholar
Aplin, A. C. and Macquaker, J. H. S. (2011). Mudstone diversity: origin and implications for source, seal, and reservoir properties in petroleum systems. Am. Assoc. Petrol. Geol. Bull. 95:20312059.Google Scholar
Argon, A. S., Andrews, R. D., Godrick, J. A. and Whitney, W. (1968). Plastic deformation bands in glassy polystyrene. J. Appl. Physics 39:18991906.Google Scholar
Arthur, J., Dunstan, T., Al-Ani, Q. and Assadi, A. (1977). Plastic deformation and failure in granular media. Géotechnique 27:5374.Google Scholar
Ashby, M. F. (2016). Materials Selection in Mechanical Design (5th edn), Butterworth-Heinemann, 660 pp.Google Scholar
Ashby, M. F. and Sammis, C. G. (1990). The damage mechanics of brittle solids in compression. Pure Appl. Geophys. 133:489521.Google Scholar
Atkinson, B. K. (1984). Subcritical crack growth in geologic materials. J. Geophys. Res. 89:40774114.CrossRefGoogle Scholar
Atkinson, B. K. (1987). Introduction to fracture mechanics and its geophysical applications. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 126, Academic Press, New York.Google Scholar
Atkinson, B. K. and Meredith, P. G. (1987a). The theory of subcritical crack growth with applications to minerals and rocks. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 111166, Academic Press, New York.Google Scholar
Atkinson, B. K. and Meredith, P. G. (1987b). Experimental fracture mechanics data for rocks and minerals. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 477525, Academic Press, New York.Google Scholar
Atkinson, G. M., Eaton, D. W., Ghofrani, H., et al. (2016). Hydraulic fracturing and seismicity in the Western Canada Sedimentary Basin. Seismol. Res. Lett. 87:631647.Google Scholar
Attewell, P. B. and Woodman, J. P. (1971). Stability of discontinuous rock masses under polyaxial stress systems. In 13th Symp. on Rock Mech., Stability of Rock Slopes, pp. 665683, Am. Soc. Civil Eng., New York.Google Scholar
Attewell, P. B. and Farmer, I. W. (1973). Fatigue behaviour of rocks. Int. J. Rock Mech. Min. Sci. 10:19.Google Scholar
Atwater, T. (1970). Implications of plate tectonics for the Cenozoic tectonic evolution of western North America. Geol. Soc. Am. Bull. 81:35133536.CrossRefGoogle Scholar
Awdal, A., Healy, D. and Alsop, G. I. (2014). Geometrical analysis of deformation band lozenges and their scaling relationships to fault lenses. J. Struct. Geol. 66:1123.Google Scholar
Axen, G. J. (1992). Pore pressure, stress increase, and fault weakening in low-angle normal faults. J. Geophys. Res. 97:89798991.CrossRefGoogle Scholar
Aydin, A. (1978). Small faults formed as deformation bands in sandstone. Pure Appl. Geophys. 116:913930.Google Scholar
Aydin, A. (1988). Discontinuities along thrust faults and the cleavage duplexes. In Geometries and Mechanisms of Thrusting, with Special Reference to the Appalachians (eds. Mitra, G. and Wojtal, S.), Geol. Soc. Am. Spec. Pap. 222:223232.CrossRefGoogle Scholar
Aydin, A. (2000). Fractures, faults, and hydrocarbon entrapment, migration and flow. Marine Petrol. Geol. 17:797814.Google Scholar
Aydin, A. (2014). Failure modes of shales and their implications for natural and man-made fracture assemblages. Am. Assoc. Petrol. Geol. Bull. 98:23912409.Google Scholar
Aydin, A. and Ahmadov, R. (2009). Bed-parallel compaction bands in aeolian sandstone: their identification, characterization and implications. Tectonophysics 479:277284.Google Scholar
Aydin, A. and Berryman, J. G. (2010). Analysis of the growth of strike-slip faults using effective medium theory. J. Struct. Geol. 32:16291642.Google Scholar
Aydin, A. and DeGraff, J. M. (1988). Evolution of polygonal fracture patterns in lava flows. Science 239:471476.Google Scholar
Aydin, A. and Johnson, A. (1978). Development of faults as zones of deformation bands and as slip surfaces in sandstone. Pure Appl. Geophys. 116:931942.CrossRefGoogle Scholar
Aydin, A. and Johnson, A. (1983). Analysis of faulting in porous sandstones. J. Struct. Geol. 5:1931.Google Scholar
Aydin, A. and Nur, A. (1982). Evolution of pull-apart basins and their scale independence. Tectonics 1:1121.Google Scholar
Aydin, A. and Nur, A. (1985). The types and role of stepovers in strike-slip tectonics. In Strike-Slip Deformation, Basin Formation, and Sedimentation (eds. Biddle, K. T. and Christie-Blick, N.), pp. 3544, Soc. Econ. Paleon. Miner., Spec. Publ. 37.CrossRefGoogle Scholar
Aydin, A. and Page, B. M. (1984). Diverse Pliocene–Quaternary tectonics in a transform environment, San Francisco Bay region, California. Geol. Soc. Am. Bull. 95:13031317.2.0.CO;2>CrossRefGoogle Scholar
Aydin, A. and Reches, Z. (1982). Number and orientation of fault sets in the field and in experiments. Geology 10:107112.Google Scholar
Aydin, A. and Schultz, R. A. (1990). Effect of mechanical interaction on the development of strike-slip faults with echelon patterns. J. Struct. Geol. 12:123129.Google Scholar
Aydin, A., Schultz, R. A. and Campagna, D. (1990). Fault-normal dilation in pull-apart basins: implications for the relationship between strike-slip faults and volcanic activity. Annales Tectonicae 4:4552.Google Scholar
Aydin, A., Borja, R. I. and Eichhubl, P. (2006). Geological and mathematical framework for failure modes in granular rock. J. Struct. Geol. 28:8398.CrossRefGoogle Scholar
Bahat, D. (1979). Theoretical considerations on mechanical parameters of joint surfaces based on studies on ceramics. Geol. Mag. 116:8192.CrossRefGoogle Scholar
Bahat, D. (1991). Tectonofractography, Springer-Verlag, Heidelberg.CrossRefGoogle Scholar
Bahat, D. and Engelder, T. (1984). Surface morphology on cross-fold joints of the Appalachian plateau, New York and Pennsylvania. Tectonophysics 104:299313.Google Scholar
Bahat, D., Bankwitz, P. and Bankwitz, E. (2003). Preuplift joints in granites: evidence for subcritical and postcritical fracture growth. Geol. Soc. Am. Bull. 115:148165.Google Scholar
Bai, T. and Gross, M. R. (1999). Theoretical analysis of cross-joint geometries and their classification. J. Geophys. Res. 104:11631177.Google Scholar
Bai, T. and Pollard, D. D. (1997). Experimental study of joint surface morphology and fracture propagation rate (abstract). Eos, Trans. Am. Geophys. Union 78:F711.Google Scholar
Bai, T. and Pollard, D. D. (2000a). Fracture spacing in layered rocks: a new explanation based on the stress transition. J. Struct. Geol. 22:4357.Google Scholar
Bai, T. and Pollard, D. D. (2000b). Closely spaced fractures in layered rocks: initiation mechanism and propagation kinematics. J. Struct. Geol. 22:14091425.Google Scholar
Bai, T., Maerten, L., Gross, M. R. and Aydin, A. (2002). Orthogonal cross joints: do they imply a regional stress rotation? J. Struct. Geol. 24:7788.Google Scholar
Baig, A. M., Urbancic, T. and Viegas, G. (2012). Do hydraulic fractures induce events large enough to be felt on surface? Can. Soc. Explor. Geophys. Recorder, October 2012: 4046.Google Scholar
Bailey, W. R., Walsh, J. J. and Manzocchi, T. (2005). Fault populations, strain distribution and basement fault reactivation in the East Pennines Coalfield, UK. J. Struct. Geol. 27:913928.CrossRefGoogle Scholar
Ballas, G., Soliva, R., Sizun, J.-P., et al. (2012). The importance of the degree of cataclasis in shear bands for fluid flow in porous sandstone, Provence, France. Am. Assoc. Petrol. Geol. Bull. 96:21672186.Google Scholar
Ballas, G., Soliva, R., Sizun, J.-P., et al. (2013). Shear-enhanced compaction bands formed at shallow burial conditions; implications for fluid flow (Provence, France). J. Struct. Geol. 47:315.Google Scholar
Ballas, G., Soliva, R., Benedicto, A. and Sizun, J.-P. (2014). Control of tectonic setting and large-scale faults on the basin-scale distribution of deformation bands in porous sandstone (Provence, France). Marine Petrol. Geol. 55:142159.Google Scholar
Ballas, G., Fossen, H. and Soliva, R. (2015). Factors controlling permeability of cataclastic deformation bands and faults in porous sandstone reservoirs. J. Struct. Geol. 76:121.CrossRefGoogle Scholar
Banks, C. J. and Warburton, J. (1986). ‘Passive-roof’ duplex geometry in the frontal structures of the Kirthar and Sulaiman mountain belts, Pakistan. J. Struct. Geol. 8:229237.CrossRefGoogle Scholar
Barenblatt, G. I. (1962). The mathematical theory of equilibrium cracks in brittle fracture. Adv. Appl. Mech. 7:55129.Google Scholar
Barka, A. A. and Kadinsky-Cade, K. (1988). Strike-slip fault geometry in Turkey and its influence on earthquake activity. Tectonics 7:663684.Google Scholar
Barnett, J. A. M., Mortimer, J., Rippon, J. H., Walsh, J. J. and Watterson, J. (1987). Displacement geometry in the volume containing a single normal fault. Am. Assoc. Petrol. Geol. Bull. 71:925937.Google Scholar
Barree, R. D., Gilbert, J. V. and Conway, M. W. (2009). Stress and rock property profiling for unconventional reservoir stimulation. Paper SPE 118703 presented at the 2009 SPE Hydraulic Fracturing Technology Conference, The Woodlands, Texas, 19–21 January 2009.Google Scholar
Barrell, D. J. A., Litchfield, N. J., Townsend, D. B., et al. (2011). Strike-slip ground-surface rupture (Greendale Fault) associated with the 4 September 2010 Darfield earthquake, Canterbury, New Zealand. Quart. J. Eng. Geol. Hydrogeol., 44:283291; doi:10.1144/1470-9236/11-034.Google Scholar
Bartlett, W. L., Friedman, M. and Logan, J. M. (1981). Experimental folding and faulting of rocks under confining pressure, Part IX. Wrench faults in limestone layers. Tectonophysics 79:255277.Google 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 measurement. J. Geophys. Res. 99:93739390.Google Scholar
Barton, C. A., Zoback, M. D. and Moos, D. (1995). Fluid flow along potentially active faults in crystalline rock. Geology 23:683686.Google Scholar
Barton, C. C. (1983). Systematic jointing in the Cardium Sandstone along the Bow River, Alberta, Canada. PhD thesis, Yale University, New Haven, CT.Google Scholar
Barton, N. R. (1976). The shear strength of rock and rock joints. Int. J. Rock Mech. Min. Sci. Geomech. Abs. 13:255279.Google Scholar
Barton, N. R. (1990). Scale effects or sampling bias? In Scale Effects in Rock Masses (ed. Cunha, A. P.), pp. 3155. Balkema, Rotterdam.Google Scholar
Barton, N. R. (2013). Shear strength criteria for rock, rock joints, rockfill and rock masses: problems and some solutions. J. Rock Mech. Geotech. Eng. 5:249261.CrossRefGoogle Scholar
Barton, N. R. and Bandis, S. C. (1990). Review of predictive capabilities of JRC-JCS model in engineering practice. In Proc. Int. Symp. Rock Joints (eds. Barton, N. and Stephansson, O.), pp. 603610, Loen, Norway.Google Scholar
Barton, N. R. and Choubey, V. (1977). The shear strength of rock joints in theory and practice. Rock Mech. 10:154.Google Scholar
Barton, N. R., Lien, R. and Lunde, J. (1974). Engineering classification of rock masses for the design of tunnel support. Rock Mech. 6:189236.CrossRefGoogle Scholar
Baud, P., Klein, E. and Wong, T.-f. (2004). Compaction localization in porous sandstones: spatial evolution of damage and acoustic emission activity. J. Struct. Geol. 26:603624.Google Scholar
Baud, P., Vajdova, V. and Wong, T.-f. (2006). Shear-enhanced compaction and strain localization: inelastic deformation and constitutive modeling of four porous sandstones. J. Geophys. Res. 111, B12401; doi:10.1029/2005JB004101.CrossRefGoogle Scholar
Baudon, C. and Cartwright, J. A. (2008). 3D seismic characterization of an array of linked normal faults in the Levant Basin, Eastern Mediterranean. J. Struct. Geol. 30:746760.Google Scholar
Baxevanis, T., Papamichos, E., Flornes, O. and Larsen, I. (2006). Compaction bands and induced permeability reduction in Tuffeau de Maastricht calcarenite. Acta Geotechnica 1:123135.Google Scholar
Bayly, B. (1992). Mechanics in Structural Geology, Springer-Verlag, New York.Google Scholar
Bazant, Z. P. (1976). Instability, ductility and size effect in strain-softening concrete. J. Engng. Mech. 102:331344.Google Scholar
Bazant, Z. P. and Hubler, M. H. (2014). Theory of cyclic creep of concrete based on Paris law for fatigue growth of subcritical microcracks. J. Mech. Phys. Solids 63:187200.Google Scholar
Beach, A. (1975). The geometry of en echelon vein arrays. Tectonophysics 28:245263.Google Scholar
Becker, S. P., Eichhubl, P., Laubach, S. E., et al. (2010). A 48 m.y. history of fracture opening, temperature, and fluid pressure: Cretaceous Travis Peak Formation, East Texas basin. Geol. Soc. Am. Bull. 122:10811093.CrossRefGoogle Scholar
Beeler, N. M. and Lockner, D. A. (2003). Why earthquakes correlate weakly with the solid earth tides: effects of periodic stress on the rate and probability of earthquake occurrence. J. Geophys. Res. 108:2391; doi:10.1029/2001JB001518.Google Scholar
Beeler, N. M., Tullis, T. E. and Weeks, J. D. (1996). Frictional behavior of large displacement experimental faults. J. Geophys. Res. 101:86978715.Google Scholar
Beeler, N. M., Simpson, R. W., Hickman, S. H. and Lockner, D. A. (2000). Pore fluid pressure, apparent friction, and Coulomb failure. J. Geophys. Res. 105:25,53325,542.Google Scholar
Beer, F. P., Johnston, E. R. Jr. and DeWolf, J. T. (2004). Mechanics of Materials (in SI Units) (3rd edn), McGraw-Hill, New York.Google Scholar
Begley, J. A. and Landes, J. D. (1972). The J-integral as a fracture criterion. In Fracture Toughness, pp. 120, Am. Soc. Testing Mat., ASTM STP 514.Google Scholar
Begley, J. A. and Landes, J. D. (1976). Serendipity and the J-integral. Int. J. Fracture 12:764766.Google Scholar
Behn, M. D., Lin, J. and Zuber, M. T. (2002a). Mechanisms of normal fault development at mid-ocean ridges. J. Geophys. Res. 107; doi:10.1029/2001JB000503.Google Scholar
Behn, M. D., Lin, J. and Zuber, M. T. (2002b). Evidence for weak transform faults. Geophys. Res. Lett. 29:2207; doi:10.1029/2002GL015612.Google Scholar
Belfield, W. C. (1998). Incorporating spatial distribution into stochastic modelling of fractures: multifractals and Lévy-stable statistics. J. Struct. Geol. 20:473486.Google Scholar
Bell, F. G. (1993). Engineering Geology, Blackwell, London.Google Scholar
Benedicto, A. and Schultz, R. A. (2010). Stylolites in limestone: magnitude of contractional strain accommodated and scaling relationships. J. Struct. Geol. 32:12501256.Google Scholar
Benedicto, A., Schultz, R. and Soliva, R. (2003). Layer thickness and the shape of faults. Geophys. Res. Lett. 30:2076, doi:10.1029/2003GL018237.Google Scholar
Bense, V. F., Gleeson, T., Loveless, S. E., Bour, O. and Scibek, J. (2013). Fault zone hydrology. Earth-Sci. Rev. 127:171192.Google Scholar
Ben-Zion, Y. (2001). On quantification of the earthquake source. Seismol. Res. Lett. 72:151152.Google Scholar
Bergen, K., and Shaw, J. H. (2010). Displacement profiles and displacement-length scaling relationships of thrust faults constrained by seismic reflection data. Geol. Soc. Am. Bull. 122:12091219; doi:10.1130/B26373.1.Google Scholar
Bergsaker, A. S., Røyne, A., Ougier-Simonin, A. and Renard, F. (2016). The effect of fluid composition, salinity, and acidity on subcritical crack growth in calcite crystals. J. Geophys. Res. 121:16311651; doi:10.1002/2015JB012723.Google Scholar
Bernal, A., Hardy, S. and Gawthorpe, R. L. (2018). Three-dimensional growth of flexural slip fault-bend and fault-propagation folds and their geomorphic expression. Geosciences 8: 110; doi:10.3390/geosciences8040110.Google Scholar
Bertrand, L., Géraud, Y., Le Garzic, E., et al. (2015). A multiscale analysis of a fracture pattern in granite: a case study of the Tamariu granite, Catalunya, Spain. J. Struct. Geol. 78:5266.Google Scholar
Bessenger, B. A., Liu, Z., Cook, N. G. W. and Myer, L. R. (1997). A new fracturing mechanism for granular media. Geophys. Res. Lett. 24:26052608.Google Scholar
Bésuelle, P. (2001a). Compacting and dilating shear bands in porous rock: theoretical and experimental conditions. J. Geophys. Res. 106:13,43513,442.Google Scholar
Bésuelle, P. (2001b). Evolution of strain localisation with stress in a sandstone: brittle and semi-brittle regimes. Phys. Chem. Earth A26:101106.Google Scholar
Bésuelle, P. and Rudnicki, J. W. (2004). Localization: shear bands and compaction bands. In Mechanics of Fluid-Saturated Rocks (eds. Guéguen, Y. and Boutéca, M.), pp. 219321, Elsevier, Amsterdam.Google Scholar
Biddle, K. T. and Christie-Blick, N. (1985). Glossary—Strike-slip deformation, basin formation, and sedimentation. In Strike-Slip Deformation, Basin Formation, and Sedimentation (eds. Biddle, K. T. and Christie-Blick, N.), pp. 375386, Soc. Econ. Paleon. Miner., Spec. Publ. 37.Google Scholar
Biegel, R. L., Sammis, C. G. and Dieterich, J. H. (1989). The frictional properties of a simulated gouge having a fractal particle distribution. J. Struct. Geol. 11:827846.Google Scholar
Bieniawski, Z. T. (1968). The effect of specimen size on compressive strength of coal. Int. J. Rock Mech. Min. Sci. 5:325335.CrossRefGoogle Scholar
Bieniawski, Z. T. (1974). Estimating the strength of rock materials. J. S. Afr. Inst. Min. Metall. 74:312320.Google Scholar
Bieniawski, Z. T. (1978). Determining rock mass deformability—experience from case histories. Int. J. Rock Mech. Min. Sci. 15:237247.Google Scholar
Bieniawski, Z. T. (1989). Engineering Rock Mass Classifications: A Complete Manual for Engineers and Geologists in Mining, Civil, and Petroleum Engineering, Wiley, New York.Google Scholar
Bieniawski, Z. T. and Van Heerden, W. L. (1975). The significance of in-situ tests on large rock specimens. Int. J. Rock Mech. Min. Sci. 12:101113.Google Scholar
Biggs, J., Bastow, I. D., Keir, D. and Lewi, E. (2011). Pulses of deformation reveal frequently recurring shallow magmatic activity beneath the Main Ethiopian Rift. Geochem. Geophys. Geosystems 12:Q0AB10; doi:10.1029/2011GC003662.Google Scholar
Bilby, B. A., Cotterell, A. H. and Swindon, K. H. (1963). The spread of plastic yield from a notch. Proc. Royal Soc. London A, 272:304314.Google Scholar
Bilham, R. and King, G. (1989). The morphology of strike-slip faults: examples from the San Andreas Fault, California. J. Geophys. Res., 94:10,20410,216.Google Scholar
Billi, A., Salvini, F. and Storti, F. (2003). The damage zone-fault core transition in carbonate rocks: implications for fault growth, structure and permeability. J. Struct. Geol. 25:17791794.Google Scholar
Billings, M. P. (1972). Structural Geology (3rd edn), Prentice-Hall, Englewood Cliffs, New Jersey.Google Scholar
Bilotti, F. and Suppe, J. (1999). The global distribution of wrinkle ridges on Venus. Icarus 139:137159.Google Scholar
Bishop, D. G. (1968). The geometric relationships of structural features associated with major strike-slip faults in New Zealand. N.Z. J. Geol. Geophys. 11:405417.Google Scholar
Boettcher, M. S. and Marone, C. (2004). Effects of normal stress variations on the strength and stability of creeping faults. J. Geophys. Res. 109:3406; doi:10.1029/2003JB002824.Google Scholar
Bolton, M. (1986). The strength and dilatancy of sands. Géotechnique 36:6578.Google Scholar
Bonnet, E., Bour, O., Odling, N. E., et al. (2001). Scaling of fracture systems in geological media. Rev. Geophys. 39:347383.Google Scholar
Bons, P. D., Elburg, M. A. and Gomez-Rivas, E. (2012). A review of the formation of tectonic veins and their microstructures. J. Struct. Geol. 43:3362.Google Scholar
Borgia, A., Burr, J., Montero, W., Morales, L. D. and Alvarado, G. A. (1990). Fault propagation folds induced by gravitational failure and slumping of the central Costa Rica volcanic range: implications for large terrestrial and Martian volcanic edifices. J. Geophys. Res. 95:14,357–14,382.Google Scholar
Borgos, H. G., Cowie, P. A. and Dawers, N. H. (2000). Practicalities of extrapolating one-dimensional fault and fracture size–frequency distributions to higher-dimensional samples. J. Geophys. Res. 105:28,37728,391.Google Scholar
Borja, R. I. (2002). Bifurcation of elastoplastic solids to shear band mode at finite strain. Comput. Methods Appl. Mech. Eng. 191:52875314.Google Scholar
Borja, R. I. (2004). Computational modeling of deformation bands in granular media. II. Numerical simulations. Comput. Methods Appl. Mech. Eng. 193:26992718.Google Scholar
Borja, R. I. and Aydin, A. (2004). Computational modeling of deformation bands in granular media. I. Geological and mathematical framework. Comput. Methods Appl. Mech. Eng. 193:26672698.Google Scholar
Bosworth, W. (1985). Geometry of propagating continental rifts. Nature 316:625627.Google Scholar
Bott, M. H. P. (1959). The mechanics of oblique slip faulting. Geol. Mag. 96:109117.Google Scholar
Bowden, P. B. and Raha, S. (1970). The formation of micro shear bands in polystyrene and polymethylmethacrylate. Philosoph. Mag. 22:463482.CrossRefGoogle Scholar
Boyer, S. E. and Elliott, D. (1982). Thrust systems. Am. Assoc. Petrol. Geol. Bull. 66:11961230.Google Scholar
Boyer, S. E. and Mitra, G. (1988). Relations between deformation of crystalline basement and sedimentary cover at the basement/cover transition zone of the Appalachian Blue Ridge province. In Geometries and Mechanisms of Thrusting, with Special Reference to the Appalachians (eds. Mitra, G. and Wojtal, S.), Geol. Soc. Am. Spec. Pap. 222:119136.CrossRefGoogle Scholar
Brace, W. F. (1960). An extension of the Griffith theory of fracture to rocks. J. Geophys. Res. 65:34773480.Google Scholar
Brace, W. F. (1961). Dependence of the fracture strength of rocks on grain size. Penn. State Univ. Min. Ind. Bull. 76:99103.Google Scholar
Brace, W. F. and Bombolakis, E. G. (1963). A note on brittle crack growth in compression. J. Geophys. Res. 68:37093713.Google Scholar
Brace, W. F. and Byerlee, J. D. (1966). Stick-slip as a mechanism for earthquakes. Science 153:990992.Google Scholar
Brace, W. F. and Kohlstedt, D. L. (1980). Limits on lithospheric stress imposed by laboratory experiments. J. Geophys. Res. 85:62486252.Google Scholar
Brace, W. F., Paulding, B. W. and Scholz, C. H. (1966). Dilatancy in the fracture of crystalline rocks. J. Geophys. Res. 71:39393953.Google Scholar
Brady, B. T. (1971a). An exact solution to the radially end-constrained circular cylinder under triaxial loading. Int. J. Rock Mech. Min. Sci. 8:165178.Google Scholar
Brady, B. T. (1971b). The effect of confining pressure on the elastic stress distribution in a radially end-constrained circular cylinder. Int. J. Rock Mech. Min. Sci. 8:153164.Google Scholar
Brady, B. H. G. and Brown, E. T. (1993). Rock Mechanics for Underground Mining, Chapman and Hall, London.Google Scholar
Brandes, C. and Tanner, D. C. (2012). Three-dimensional geometry and fabric of shear deformation-bands in unconsolidated Pleistocene sediments. Tectonophysics 518521:8492.Google Scholar
Brantut, N., Baud, P., Heap, H. J. and Meredith, P. G. (2012). Micromechanics of brittle creep in rocks. J. Geophys. Res. 117, B08412; doi:10.1029/2012JB009299.Google Scholar
Brantut, N., Heap, H. J., Meredith, P. G. and Baud, P. (2013). Time-dependent cracking and brittle creep in crustal rocks: a review. J. Struct. Geol. 52:1743.Google Scholar
Breckels, I. M. and van Eekelen, H. A. M. (1982). Relationship between horizontal stress and depth in sedimentary basins. J. Petrol. Tech. 34:21912199.Google Scholar
Bridgman, P. W. (1936). Shearing phenomena at high pressure of possible importance to geology. J. Geol. 44:653669.Google Scholar
Broberg, K. B. (1999). Cracks and Fracture, Academic, San Diego.Google Scholar
Broch, E. and Franklin, J. A. (1972). The point-load strength test. Int. J. Rock Mech. Min. Sci. 9:669697.Google Scholar
Brodsky, E. E., Roeloffs, E., Woodcock, D., Gall, I. and Manga, M. (2003). A mechanism for sustained groundwater pressure changes induced by distant earthquakes. J. Geophys. Res. 108, 2390; doi:1029/2002/JB002321.Google Scholar
Broek, D. (1986). Elementary Engineering Fracture Mechanics (4th edn), Martinus Nijhoff, Boston.Google Scholar
Brown, C. D. and Phillips, R. J. (1999). Flexural rift flank uplift at the Rio Grande rift, New Mexico. Tectonics 18:12751291.Google Scholar
Brown, E. and Hoek, E. (1978). Trends in relationships between measured in situ stress and depth. Int. J. Rock Mech. Min. Sci. & Geomech. Abs. 15:211215.Google Scholar
Brown, E. and Hoek, E. (1988). Determination of shear failure envelope in rock masses. J. Geotech. Eng. Div. Am. Soc. Civ. Engrs. 114:371376.Google Scholar
Bruhn, R. L. and Schultz, R. A. (1996). Geometry and slip distribution in normal fault systems: implications for mechanics and fault-related hazards. J. Geophys. Res. 101:34013412.Google Scholar
Bruhn, R. L., Yonkee, W. A. and Parry, W. T. (1990). Structural and fluid-chemical properties of seismogenic normal faults. Tectonophysics 175:130157.Google Scholar
Brzesowsky, R. H., Hangx, S. J. T., Brantut, N. and Spiers, C. J. (2014). Compaction creep of sands due to time-dependent grain failure: effects of chemical environment, applied stress, and grain size. J. Geophys. Res. 119:75217541.Google Scholar
Bucher, W. H. (1920). The mechanical interpretation of joints. J. Geol. 28:707730.Google Scholar
Budkewitsch, P. and Robin, P.-Y. (1994). Modelling the evolution of columnar joints. J. Volcanol. Geotherm. Res. 59:219239.Google Scholar
Buiter, S. J. H., Babeyko, A. Y., Ellis, S., et al. (2006). The numerical sandbox: comparison of model results for a shortening and an extension experiment. In Analogue and Numerical Modeling of Crustal-Scale Processes (eds. Buiter, S. J. H. and Schreurs, G.), pp. 2964, Geol. Soc. London Spec. Publ. 253.Google Scholar
Bunger, A. P. (2008). A rigorous tool for evaluating the importance of viscous dissipation in sill formation: it’s in the tip. In Dynamics of Crustal Magma Transfer, Storage and Differentiation (eds. Annen, C. and Zellmer, G. F.), pp. 7181, Geol. Soc. London Spec. Publ. 304.Google Scholar
Burdekin, F. M. and Stone, D. E. W. (1966). The crack opening displacement approach to fracture mechanics in yielding materials. J. Strain Anal. 1:145153.Google Scholar
Burchfiel, B. C. and Stewart, J. H. (1966). “Pull-apart” origin of the central segment of Death Valley, California. Geol. Soc. Am. Bull. 77:439442.Google Scholar
Bürgmann, R. and Pollard, D. D. (1992). Influence of the state of stress on the brittle-ductile transition in granitic rock: evidence from fault steps in the Sierra Nevada, California. Geology 20:645648.Google Scholar
Bürgmann, R. and Pollard, D. D. (1994). Strain accommodation about strike-slip fault discontinuities in granitic rock under brittle-to-ductile conditions. J. Struct. Geol. 16:16551674.Google Scholar
Bürgmann, R., Pollard, D. D. and Martel, S. J. (1994). Slip distributions on faults: effects of stress gradients, inelastic deformation, heterogeneous host-rock stiffness, and fault interaction. J. Struct. Geol. 16:16751690.Google Scholar
Burnley, P. C., Green, H. W., II and Prior, D. (1991). Faulting associated with the olivine to spinel transformation in Mg2GeO4 and its implications for deep-focus earthquakes. J. Geophys. Res. 96:425443.Google Scholar
Busetti, S., Mish, K., Hennings, P. and Reches, Z. (2012). Damage and plastic deformation of reservoir rocks: part 2. Propagation of a hydraulic fracture. Amer. Assoc. Petrol. Geol. Bull. 96:17111732.Google Scholar
Butler, R. W. H. (1982). The terminology of structures in thrust belts. J. Struct. Geol. 4:239245.Google Scholar
Byerlee, J. D. (1970). The mechanics of stick-slip. Tectonophysics 9:475486.Google Scholar
Byerlee, J. D. (1978). Friction of rocks. Pure Appl. Geophys. 116:615626.Google Scholar
Byerlee, J. D. and Brace, W. F. (1968). Stick-slip, stable sliding, and earthquakes: effect of rock type, pressure, strain rate, and stiffness. J. Geophys. Res. 73:60316037.Google Scholar
Caine, J. S., Evans, J. P. and Forster, C. B. (1996). Fault zone architecture and permeability structure. Geology 24:10251028.Google Scholar
Calais, E., d’Oreye, N., Albaric, J., et al. (2008). Aseismic strain accommodation by slow slip and dyking in a youthful continental rift, East Africa. Nature 456; doi:10.1038/nature07478.Google Scholar
Callister, W. D., Jr. (2000). Materials Science and Engineering: An Introduction (5th edn), Wiley, New York.Google Scholar
Campagna, D. J. and Aydin, A. (1991). Tertiary uplift and shortening in the Basin and Range; the Echo Hills, southeastern Nevada. Geology 19:485488.Google Scholar
Campagna, D. J. and Levandowski, D. W. (1991). The recognition of strike-slip fault systems using imagery, gravity, and topographic data sets. Photog. Engng. Rem. Sens. 57:11951201.Google Scholar
Candela, T. and Renard, F. (2012). Segment linkage process at the origin of slip surface roughness: evidence from the Dixie Valley fault. J. Struct. Geol. 45:87100.Google Scholar
Carbotte, S. and Macdonald, K. (1994). Comparison of seafloor tectonic fabric at intermediate, fast, and super fast spreading ridges: influence of spreading rate, plate motions, and ridge segmentation on fault patterns. J. Geophys. Res. 99:13,60913,631.Google Scholar
Carr, M. H. (1974). Tectonism and volcanism of the Tharsis region of Mars. J. Geophys. Res. 79:39433949.Google Scholar
Carter, B. J., Scott Duncan, E. J. and Lajtai, E. Z. (1991). Fitting strength criteria to intact rock. Geotech. Geol. Eng. 9:7381.Google Scholar
Carter, K. E. and Winter, C. L. (1995). Fractal nature and scaling of normal faults in the Española Basin, Rio Grande rift, New Mexico: implications for fault growth and brittle strain. J. Struct. Geol. 17:863873.Google Scholar
Cartwright, J. A. and Lonergan, L. (1996). Volumetric contraction during the compaction of mudrocks: a mechanism for the development of regional-scale polygonal fault systems. Basin Res. 8:183193.Google Scholar
Cartwright, J. A. and Mansfield, J. S. (1998). Lateral tip geometry and displacement gradients on normal faults in the Canyonlands National Park, Utah. J. Struct. Geol. 20:319.Google Scholar
Cartwright, J. A., Trudgill, B. D. and Mansfield, C. S. (1995). Fault growth by segment linkage: an explanation for scatter in maximum displacement and trace length data from the Canyonlands Grabens of SE Utah. J. Struct. Geol. 17:13191326.Google Scholar
Cartwright, J. A., Mansfield, C. and Trudgill, B. (1996). The growth of normal faults by segment linkage. In Modern Developments in Structural Interpretation, Validation and Modelling (eds. Buchanan, P. G. and Nieuwland, D. A.), pp. 163177, Geol. Soc. Spec. Publ. 99.Google Scholar
Cartwright, J. A., Bolton, A. J. and James, D. M. D. (2003). The genesis of polygonal fault systems: a review. In Subsurface Sediment Mobilization (eds. Van Rensbergen, P. et al.), pp. 223243, Geol. Soc. Lond. Spec. Publ. 216; doi:10.1144/GSL.SP.2003.216.01.15.Google Scholar
Cartwright, J. A., Huuse, M. and Aplin, A. (2007). Seal bypass systems. Amer. Assoc. Petrol. Geol. Bull. 91:11411166.Google Scholar
Casagrande, A. (1936). The determination of the pre-consolidation load and its practical significance. Proc. 1st Int. Soil Mech. and Found. Eng. Conf., June 22–26, 1936, Cambridge, Massachusetts, pp. 6064.Google Scholar
Cashman, S., and Cashman, K. (2000). Cataclasis and deformation-band formation in unconsolidated marine terrace sand, Humboldt County, California. Geology 28:111114.Google Scholar
Castaing, C., Halawani, M. A., Gervais, F., et al. (1996). Scaling relationships in intraplate fracture systems related to Red Sea rifting. Tectonophysics 261:291314.Google Scholar
Cates, M. E., Wittmer, J. P., Bouchaud, J.-P. and Claudin, P. (1998). Jamming, force chains, and fragile matter. Phys. Rev. Lett. 81:18411844.Google Scholar
Chadwick, W. W., Jr. and Embley, R. W. (1998). Graben formation associated with recent dike intrusions and volcanic eruptions on the mid-ocean ridge. J. Geophys. Res. 103:98079825.Google Scholar
Challa, V. and Issen, K. A. (2004). Conditions for compaction band formation in porous rock using a two yield surface model. J. Eng. Mech. 130:10891097.Google Scholar
Chambon, G., Schmittbuhl, J., Corfdir, A., et al. (2006). The thickness of faults: from laboratory experiments to field scale observations. Tectonophysics 426:7794.Google Scholar
Champion, J. A., Tate, A., Mueller, K. J. and Guccione, M. (2001). Geometry, numerical modeling and revised slip rate for the Reelfoot blind thrust and trishear fault-propagation fold, New Madrid seismic zone. Eng. Geol. 62:3149.Google Scholar
Chang, M. F. (1991). Interpretation of overconsolidation ratio from in situ tests in Recent clay deposits in Singapore and Malaysia. Can. Geotech. J. 28:210225.Google Scholar
Chapman, C. R. and McKinnon, W. B. (1986). Cratering of planetary satellites. In Satellites, University of Arizona Press, pp. 492580.Google Scholar
Charles, R. J. (1958). Dynamic fatigue of glass. J. Appl. Phys. 29:16571662.Google Scholar
Chell, G. G. (1977). The application of post-yield fracture mechanics to penny-shaped and semi-circular cracks. Engng. Fract. Mech. 9:5563.Google Scholar
Chemenda, A. I. (2009). The formation of tabular compaction-band arrays: theoretical and numerical analysis. J. Mech. Phys. Solids 57:851868.Google Scholar
Chemenda, A. I. (2011). Origin of compaction bands: anti-cracking or constitutive instability? Tectonophysics 499:156164.Google Scholar
Chemenda, A., Deverchere, J. and Calais, E. (2002). Three-dimensional laboratory modeling of rifting: application to the Baikal rift, Russia. Tectonophysics 356:253273.Google Scholar
Chemenda, A. I., Nguyen, S.-H., Petit, J.-P. and Ambre, J. (2011). Mode I cracking versus dilatancy banding: experimental constraints on the mechanisms of extension fracturing. J. Geophys. Res. 116:B04401; doi:10.1029/2010JB008104.Google Scholar
Chemenda, A. I., Ballas, G. and Soliva, R. (2014). Impact of a multilayer structure on initiation and evolution of strain localization in porous rocks: field observations and numerical modeling. Tectonophysics 631:2936.Google Scholar
Chemenda, A. I., Cavalié, O., Vergnolle, M., Bouissou, S. and Delouis, B. (2016). Numerical modeling of formation of a 3-D strike-slip fault system. Comptus Rendus Geoscience 348:6169.Google Scholar
Chen, X., Eichhubl, P. and Olson, J. E. (2017). Effect of water on critical and subcritical fracture properties of Woodford shale. J. Geophys. Res. 122:27362750; doi:10.1002/2016JB013708.Google Scholar
Chester, F. M., Evans, J. P. and Biegel, R. L. (1993). Internal structure and weakening mechanisms of the San Andreas Fault. J. Geophys. Res. 98:771786.Google Scholar
Chester, J. S., Logan, J. M. and Spang, J. H. (1991). Influence of layering and boundary conditions on fault-bend and fault-propagation folding. Geol. Soc. Am. Bull. 103:10591072.Google Scholar
Cheung, C. S. N., Baud, P. and Wong, T.-f. (2012). Effect of grain size distribution on the development of compaction localization in porous sandstone. Geophys. Res. Lett. 39; doi:10.1029/2012GL053739.Google Scholar
Childs, C., Nicol, A., Walsh, J. J., and Watterson, J. (1996). Growth of vertically segmented normal faults. J. Struct. Geol. 18:13891397.Google Scholar
Childs, C., Walsh, J. J. and Watterson, J. (1997). Complexity in fault zone structure and implications for fault seal prediction. In Hydrocarbon Seals: Importance for Exploration and Production. Norwegian Petroleum Society (eds. Møller-Pedersen, P. and Koestler, A. G.), Special Publication 7 (Elsevier), pp. 6172.Google Scholar
Childs, C., Holdsworth, R. E., Jackson, C. A.-L., et al. (2017). Introduction to the geometry and growth of normal faults. In The Geometry and Growth of Normal Faults (eds. Childs, C., Holdsworth, R. E., Jackson, C. A.-L., et al.), Geol. Soc. London Spec. Publ. 439; https://doi.org/10.1144/SP439.24.Google Scholar
Chinnery, M. A. (1961). The deformation of the ground around surface faults. Seismol. Soc. Amer. Bull. 51:355372.Google Scholar
Chinnery, M. A. (1963). The stress changes that accompany strike-slip faulting. Seismol. Soc. Amer. Bull. 53:921932.Google Scholar
Chinnery, M. A. (1965). The vertical displacements associated with transcurrent faulting. J. Geophys. Res. 70:46274632.Google Scholar
Chinnery, M. A. and Petrak, J. A. (1967). The dislocation fault model with a variable discontinuity. Tectonophysics 5:513529.Google Scholar
Choi, J.-H., Edwards, P., Ko, K. and Kim, Y.-S. (2016). Definition and classification of fault damage zones: a review and a new methodological approach. Earth-Sci. Rev. 152:7087.Google Scholar
Christensen, R. M. (2013). The Theory of Materials Failure, Oxford University Press.Google Scholar
Christie, M. A. (1996). Upscaling for reservoir simulation. Paper SPE 37324, in J. Petrol. Technol. 48:10041010.Google Scholar
Christie-Blick, N. and Biddle, K. T. (1985). Deformation and basin formation along strike-slip faults. In Strike-Slip Deformation, Basin Formation, and Sedimentation (eds. Biddle, K. T. and Christie-Blick, N.), pp. 187196, Soc. Econ. Paleon. Miner. Spec. Publ. 37.Google Scholar
Churchill, R. V. and Brown, J. W. (1984). Complex Variables and Applications (4th edn), McGraw-Hill, New York.Google Scholar
Cladouhos, T. T. and Marrett, R. (1996). Are fault growth and linkage models consistent with power-law distributions of fault lengths? J. Struct. Geol. 18:281293.Google Scholar
Clark, D., McPherson, A. and Collins, C. (2011). Australia’s seismogenic neotectonic record: a case for heterogeneous intraplate deformation. Geosci. Australia Record 2011/11, Canberra, 95 pp.Google Scholar
Clark, R. M. and Cox, S. J. D. (1996). A modern regression approach to determining fault displacement–length scaling relationships. J. Struct. Geol. 18:147152.Google Scholar
Clark, R. M., Cox, S. J. D. and Laslett, G. M. (1999). Generalizations of power-law distributions applicable to sampled fault-trace lengths: model choice, parameter estimation and caveats. Geophys. J. Int. 136:357372.Google Scholar
Clark, T. A., Gordon, D., Himwich, W. E., et al. (1987). Determination of relative site motions in the western United States using Mark III very long baseline interferometry. J. Geophys. Res. 92:12,74112,750.Google Scholar
Clayton, L. (1966). Tectonic depressions along the Hope Fault, a transcurrent fault in North Canterbury, New Zealand. N. Z. J. Geol. Geophys. 9:95104.Google Scholar
Cleary, M. P. (1976). Continuously distributed dislocation model for shear-bands in softening materials. Int. J. Num. Methods Engng. 10:679702.Google Scholar
Cleary, M. P., Wright, C. A. and Wright, T. B. (1991). Experimental and modeling evidence for major changes in hydraulic fracturing design and field procedures. Paper presented at SPE Gas Technology Symposium, Soc. of Pet. Eng., Houston, Texas, paper SPE 21494.Google Scholar
Clifton, A. E. and Schlische, R. W. (2001). Nucleation, growth, and linkage of faults in oblique rift zones: results from experimental clay models and implications for maximum fault size. Geology 29:455458.Google Scholar
Cohen, S. C. (1999). Numerical models of crustal deformation in seismic fault zones. Adv. Geophys., 41:133231.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. Int. J. Rock Mech. Min. Sci. 39:695729.Google Scholar
Committee on Fracture Characterization and Fluid Flow (1996). Rock Fractures and Fluid Flow: Contemporary Understanding and Applications, National Academy Press, Washington, DC.Google Scholar
Connolly, P. and Cosgrove, J. (1999). Prediction of fracture-induced permeability and fluid flow in the crust using experimental stress data. Am. Assoc. Petrol. Geol. Bull. 83:757777.Google Scholar
Cook, J., Frederiksen, R. A., Hasbo, K., et al. (2007). Rocks matter: ground truth in geomechanics. Oilfield Rev., Autumn 2007, 3655.Google Scholar
Cook, N. G. W. (1992). Jaeger memorial dedication lecture—Natural joints in rock: mechanical, hydraulic and seismic behaviour and properties under normal stress. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 29:198223.Google Scholar
Cook, T. S. and Erdogan, F. (1972). Stresses in bonded materials with a crack perpendicular to the interface. Int. J. Eng. Sci. 10:677697.Google Scholar
Cooke, M. L. (1997). Fracture localization along faults with spatially varying friction. J. Geophys. Res. 102:22,425–22,434.Google Scholar
Cooke, M. L. and Kameda, A. (2002). Mechanical fault interaction within the Los Angeles Basin: a two-dimensional analysis using mechanical efficiency. J. Geophys. Res. 107; doi:10.1029/2001JB000542.Google Scholar
Cooke, M. L. and Madden, E. H. (2014). Is the Earth lazy? A review of work minimization in fault evolution. J. Struct. Geol. 66:334346.Google Scholar
Cooke, M. L. and Pollard, D. D. (1996). Fracture propagation paths under mixed mode loading within rectangular blocks of polymethyl methacrylate. J. Geophys. Res. 101:33873400.Google Scholar
Cooke, M. L. and Pollard, D. D. (1997). Bedding-plane slip in initial stages of fault-related folding. J. Struct. Geol. 19:567581.Google Scholar
Cooke, M. L. and Underwood, C. A. (2001). Fracture termination and step-over at bedding interfaces due to frictional slip and interface opening. J. Struct. Geol. 23:223238.Google Scholar
Cooke, M. L., Mollema, P. N., Aydin, A. and Pollard, D. D. (2000). Interlayer slip and joint localization in the East Kaibab Monocline, Utah: field evidence and results from numerical modeling. In Forced Folds and Fractures (eds. Cosgrove, J. W. and Ameen, M. S.), pp. 2349, Geol. Soc. London Spec. Publ. 169.Google Scholar
Cosgrove, J. W. (1976). The formation of crenulation cleavage. J. Geol. Soc. London 132:155178.Google Scholar
Cosgrove, J. W. (1995). The expression of hydraulic fracturing in rocks and sediments. In Fractography: Fracture Topography as a Tool in Fracture Mechanics and Stress Analysis (ed. Ameen, M. S.), pp. 97147, Geol. Soc. Spec. Publ. 92.Google Scholar
Costin, L. S. (1987). Time-dependent deformation and failure. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 167215, Academic Press, New York.Google Scholar
Cotterell, B. (1966). Notes on the paths and stability of cracks. Int. J. Fracture Mech. 2:526533.Google Scholar
Cotterell, B. and Rice, J. R. (1980). Slightly curved or kinked cracks. Int. J. Fracture 16:155169.Google Scholar
Cowie, P. A. (1998a). A healing-reloading feedback control on the growth rate of seismogenic faults. J. Struct. Geol. 20:10751087.Google Scholar
Cowie, P. A. (1998b). Normal fault growth in three-dimensions in continental and oceanic crust. In Faulting and Magmatism at Mid-Ocean Ridges (eds. Buck, W. R., Delaney, P. T., Karson, J. A. and Lagabrielle, Y.), pp. 325348, American Geophysical Union Geophys. Mon. 106.Google Scholar
Cowie, P. A. and Roberts, G. P. (2001). Constraining slip rates and spacings for active normal faults. J. Struct. Geol. 23:19011915.Google Scholar
Cowie, P. A. and Scholz, C. H. (1992a). Displacement–length scaling relationship for faults: data synthesis and analysis. J. Struct. Geol. 14:11491156.Google Scholar
Cowie, P. A. and Scholz, C. H. (1992b). Physical explanation for the displacement–length relationship of faults using a post-yield fracture mechanics model. J. Struct. Geol. 14:11331148.Google Scholar
Cowie, P. A. and Scholz, C. H. (1992c). Growth of faults by accumulation of seismic slip. J. Geophys. Res. 97:11,08511,095.Google Scholar
Cowie, P. A. and Shipton, Z. K. (1998). Fault tip displacement gradients and process zone dimensions. J. Struct. Geol. 20:983997.Google Scholar
Cowie, P. A., Scholz, C. H., Edwards, M. and Malinverno, A. (1993a). Fault strain and seismic coupling on mid-ocean ridges. J. Geophys. Res. 98:17,911–17,920.Google Scholar
Cowie, P. A., Sornette, D. and Vanneste, C. (1993b). Statistical physical model for the spatio-temporal evolution of faults. J. Geophys. Res. 98:21,809–21,821.Google Scholar
Cowie, P. A., Malinverno, A., Ryan, W. B. F. and Edwards, M. H. (1994). Quantitative fault studies on the East Pacific Rise: a comparison of sonar imaging techniques. J. Geophys. Res. 99:15,205–15,218.Google Scholar
Cowie, P. A., Sornette, D. and Vanneste, C. (1995). Multifractal scaling properties of a growing fault population. Geophys. J. Int. 122:457469.Google Scholar
Cowie, P. A., Roberts, G. P. and Mortimer, E. (2007). Strain localization within fault arrays over timescales of 100–107 years. In Tectonic Faults: Agents of Change on a Dynamic Earth (eds. Handy, M. R., Hirth, G., and Hovius, N.), pp. 4777, Dahlem Workshop Report 95, MIT Press, Cambridge, MA.Google Scholar
Cox, S. J. D. and Meredith, P. G. (1993). Microcrack formation and material softening in rock measured by monitoring acoustic emissions. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 30:1124.Google Scholar
Cox, S. J. D. and Scholz, C. H. (1988a). Rupture initiation in shear fracture of rocks: an experimental study. J. Geophys. Res. 93:33073320.Google Scholar
Cox, S. J. D. and Scholz, C. H. (1988b). On the formation and growth of faults. J. Struct. Geol. 10:413430.Google Scholar
Crater Analysis Techniques Working Group (1979). Standard techniques for presentation and analysis of crater size-frequency data. Icarus 37:467474.Google Scholar
Crawford, B. R. (1998). Experimental fault sealing: shear band permeability dependency on cataclastic fault gouge characteristics. In Structural Geology in Reservoir Characterization (eds. Coward, M. P., Daltaban, T. S. and Johnson, H.), pp. 27–47, Geol. Soc. London Spec. Publ. 127.Google Scholar
Crider, J. G. (2001). Oblique slip and the geometry of normal-fault linkage: mechanics and a case study from the Basin and Range in Oregon. J. Struct. Geol. 23:19972009.Google Scholar
Crider, J. G. and Peacock, D. C. P. (2004). Initiation of brittle faults in the upper crust: a review of field observations. J. Struct. Geol. 26:691707.Google Scholar
Crider, J. G. and Pollard, D. D. (1998). Fault linkage: three-dimensional mechanical interaction between echelon normal faults. J. Geophys. Res. 103:24,373–24,391.Google Scholar
Crider, J. G., Cooke, M. L., Willemse, E. J. M. and Arrowsmith, J. R. (1996). Linear-elastic crack models of jointing and faulting. In Structural Geology and Personal Computers (ed. De Paor, D. G.), pp. 359388.Google Scholar
Crouch, S. L. (1970). Experimental determination of volumetric strains in failed rock. Int. J. Rock Mech. Min. Sci. Geomech. Abs. 7:589603.Google Scholar
Crouch, S. L. (1976). Analysis of stresses and displacements around underground excavations: an application of the displacement discontinuity method. Geomechanics report to the National Science Foundation, University of Minnesota, Minneapolis, 268 pp.Google Scholar
Crouch, S. L. and Starfield, A. M. (1983). Boundary Element Methods in Solid Mechanics. George Allen & Unwin, London.Google Scholar
Crowell, J. C. (1962). Displacement along the San Andreas fault, California. Geol. Soc. Am. Spec. Pap. 71, 61 pp.Google Scholar
Cruikshank, K. M. and Aydin, A. (1994). Role of fracture localization in arch formation, Arches National Park, Utah. Geol. Soc. Am. Bull. 106:879891.Google Scholar
Cruikshank, K. M. and Aydin, A. (1995). Unweaving the joints in Entrada Sandstone, Arches National Park, Utah, U.S.A. J. Struct. Geol. 17:409421.Google Scholar
Cruikshank, K. M., Zhao, G. and Johnson, A. M. (1991a). Analysis of minor fractures associated with joints and faulted joints. J. Struct. Geol. 13: 865886.Google Scholar
Cruikshank, K. M., Zhao, G. and Johnson, A. M. (1991b). Duplex structures connecting fault segments in Entrada sandstone. J. Struct. Geol. 13:11851196.Google Scholar
Currie, J. B., Patnode, H. W. and Trump, R. P. (1962). Development of folds in sedimentary strata. Geol. Soc. Am. Bull. 73:655674.Google Scholar
Currie, K. L. and Ferguson, J. (1970). The mechanism of intrusion of lamprophyre dikes indicated by “offsetting” of dikes. Tectonophysics 9:525535.Google Scholar
Cuss, R. J., Rutter, E. H. and Holloway, R. F. (2003). The application of critical state soil mechanics to the mechanical behaviour of porous sandstones. Int. J. Rock Mech. Min. Sci. 40:847862.Google Scholar
Dahlen, F. A. (1990). Critical taper model of fold-and-thrust belts and accretionary wedges. Annu. Rev. Earth Planet. Sci. 18:5589.Google Scholar
Dahlen, F. A., Suppe, J. and Davis, D. (1984). Mechanics of fold-and-thrust belts and accretionary wedges: cohesive Coulomb theory. J. Geophys. Res. 89:10,08710,101.Google Scholar
Dahlstrom, C. D. A. (1970). Structural geology in the eastern margin of the Canadian Rocky Mountains. Bull. Can. Petrol. Geol. 18:332406.Google Scholar
d’ Alessio, M. A. and Martel, S. J. (2004). Fault terminations and barriers to fault growth. J. Struct. Geol. 26:18851896.Google Scholar
Das, A., Nguyen, G. D. and Einav, I. (2011). Compaction bands due to grain crushing in porous rocks: a theoretical approach based on breakage mechanics. J. Geophys. Res. 116:B08203; doi:10.1029/2011JB008265.Google Scholar
Das, B. M. (1983). Advanced Soil Mechanics, McGraw-Hill, New York.Google Scholar
Davatzes, N. C. and Aydin, A. (2003). Overprinting faulting mechanisms in high porosity sandstones of SE Utah. J. Struct. Geol. 25:17951813.Google Scholar
Davatzes, N. C., Aydin, A. and Eichhubl, P. (2003). Overprinting faulting mechanisms during the development of multiple fault sets, Chimney Rock fault array, Utah. Tectonophysics 363:118.Google Scholar
Davatzes, N. C., Eichubl, P. and Aydin, A. (2005). Structural evolution of fault zones in sandstone by multiple deformation mechanisms: Moab fault, southeast Utah. Geol. Soc. Am. Bull. 117:135148.Google Scholar
David, C., Menendez, B., Zhu, W. and Wong, T.-f. (2001). Mechanical compaction, microstructures and permeability evolution in sandstones. Phys. Chem. Earth (A) 26:4551.Google Scholar
Davies, R. K. and Pollard, D. D. (1986). Relations between left-lateral strike-slip faults and right-lateral monoclinal kink bands in granodiorite, Mt. Abbot Quadrangle, Sierra Nevada, California. Pure Appl. Geophys. 124:177201.Google Scholar
Davies, R. K., Crawford, M., Dula, W. F., Jr., Cole, M. J. and Dorn, G. A. (1997). Outcrop interpretation of seismic-scale normal faults in southern Oregon: description of structural styles and evaluation of subsurface interpretation methods. Leading Edge 16:11351141.Google Scholar
Davis, D. M. and Engelder, T. (1985). The role of salt in fold-and-thrust belts. In Collision Tectonics: Deformation of Continental Lithosphere (eds. Carter, N. L. and Uyeda, S.), Tectonophysics 119:6788.Google Scholar
Davis, D., Suppe, J. and Dahlen, F. A. (1983). Mechanics of fold-and-thrust belts and accretionary wedges. J. Geophys. Res. 88:11531172.Google Scholar
Davis, G. A. and Burchfiel, B. C. (1973). Garlock fault: an intracontinental transform structure, southern California. Geol. Soc. Am. Bull. 84:14071422.Google Scholar
Davis, G. H. (1997). Field guide to geologic structures in the Bryce Canyon region, Utah. In Am. Assoc. Petrol. Geol., Hedberg Research Conference on “Reservoir-Scale Deformation – Characterization and Prediction” (W. G. Higgs and C. F. Kluth, convenors), June 22–28, 1997, 119 pp.Google Scholar
Davis, G. H. (1998). Fault-fin landscape. Geol. Mag. 135:283286.Google Scholar
Davis, G. H. (1999). Structural geology of the Colorado Plateau region of southern Utah, with special emphasis on deformation bands. Geol. Soc. Am. Spec. Pap. 342, 157 pp.Google Scholar
Davis, G. H., Bump, A. P., García, P. E. and Ahlgren, S. G. (2000). Conjugate Riedel deformation band shear zones. J. Struct. Geol. 22:169190.Google Scholar
Davis, G. H. and Reynolds, S. J. (1996). Structural Geology of Rocks and Regions (2nd edn), Wiley, New York.Google Scholar
Davis, K., Burbank, D. W., Fisher, D., Wallace, S. and Nobes, D. (2005). Thrust-fault growth and segment linkage in the active Ostler fault zone, New Zealand. J. Struct. Geol. 27:15281546.Google Scholar
Davis, R. O. and Selvadurai, A. P. S. (1996). Elasticity and Geomechanics. Cambridge University Press, New York, 201 pp.Google Scholar
Davis, R. O. and Selvadurai, A. P. S. (2002). Plasticity and Geomechanics. Cambridge University Press, New York, 287 pp.Google Scholar
Davison, I. (1994). Linked fault systems; extensional, strike-slip and contractional. In Continental Deformation (ed. Hancock, P. L.), pp. 121142, Pergamon, New York.Google Scholar
Dawers, N. H. and Anders, M. H. (1995). Displacement-length scaling and fault linkage. J. Struct. Geol. 17:607614.Google Scholar
Dawers, N. H., Anders, M. H. and Scholz, C. H. (1993). Growth of normal faults: displacement–length scaling. Geology 21:11071110.Google Scholar
DeGraff, J. M. and Aydin, A. (1987). Surface morphology of columnar joints and its significance to mechanics and direction of joint growth. Geol. Soc. Am. Bull. 99:605617.Google Scholar
DeGraff, J. M. and Aydin, A. (1993). Effect of thermal regime on growth increment and spacing of contraction joints in basaltic lava. J. Geophys. Res. 98:64116430.Google Scholar
Deere, D. U. (1963). Technical description of cores for engineering purposes. Rock Mech. Eng. Geol. 1:1622.Google Scholar
Deere, D. U. and Miller, R. P. (1966). Engineering classification and index properties for intact rock. Tech. Rept. No. AFWL–TR–65–116, Air Force Weapons Laboratory, Kirtland Air Force Base, New Mexico.Google Scholar
de Joussineau, G. and Aydin, A. (2009). Segmentation along strike-slip faults revisited. Pure Appl. Geophys. 166:15751594.Google Scholar
de Joussineau, G., Bazalgette, L., Petit, J.-P., and Lopez, M. (2005). Morphology, intersections, and syn/late-diagenetic origin of vein networks in pelites of the Lodève Permian Basin, Southern France. J. Struct. Geol. 27:6787.Google Scholar
Delaney, P. T. (1982). Rapid intrusion of magma into wet rock: groundwater flow due to pore-pressure increases. J. Geophys. Res. 87:77397756.Google Scholar
Delaney, P. T. and Pollard, D. D. (1981). Deformation of host rocks and flow of magma during growth of minette dikes and breccia-bearing intrusions near Ship Rock, New Mexico. US Geol. Surv. Prof. Pap. 1202, 61 pp.Google Scholar
Delaney, P. T., Pollard, D. D., Ziony, J. I. and McKee, E. H. (1986). Field relations between dikes and joints: emplacement processes and paleostress analysis. J. Geophys. Res. 91:49204938.Google Scholar
DeMets, C., Gordon, R. G. and Argus, D. F. (2010). Geologically current plate motions. Geophys. J. Int. 181:180(erratum: Geophys. J. Int. 187:538).Google Scholar
Deng, S. and Aydin, A. (2012). Distribution of compaction bands in 3D in an aeolian sandstone: the role of cross-bed orientation. Tectonophysics 574575:204218.Google Scholar
Deng, S. and Aydin, A. (2015). The strength anisotropy of localized compaction: a model for the role of the nature and orientation of cross-beds on the orientation and distribution of compaction bands in 3-D. J. Geophys. Res. 120:15231542; doi:10.1002/2014JB011689.Google Scholar
Deng, S., Zuo, L., Aydin, A., Dvorkin, J. and Mukerji, T. (2015). Permeability characterization of natural compaction bands using core flooding experiments and three-dimensional image-based analysis: comparing and contrasting the results from two different methods. Am. Assoc. Petrol. Geol. Bull. 99:2749.Google Scholar
Denkhaus, H. G. (2003). Brittleness and drillability. J. S. Afr. Inst. Min. Metall. 103:523524.Google Scholar
Dershowitz, W. S. (1984). Rock joint systems. Unpublished. PhD dissertation, Massachusetts Inst. Tech., 987 pp.Google Scholar
Dershowitz, W. S. and Herda, H. H. (1992). Interpretation of fracture spacing and intensity. In Proc. 33rd US Symp. Rock Mech. (eds. Tillerson, N. and Wawersik, W.), pp. 757766, Balkema, Rotterdam.Google Scholar
Desai, C. S. and Siriwardane, H. J. (1984). Constitutive Laws for Engineering Materials with Emphasis on Geologic Materials. Prentice-Hall, Englewood Cliffs, NJ.Google Scholar
Desroches, J. and Bratton, T. (2000). Formation characterization: well logs. In Reservoir Stimulation (eds. Economides, M. J. and Nolte, K. G.), pp. 4.14.26, Wiley, New York.Google Scholar
Dewey, J. F., Holdsworth, R. E. and Trachan, R. A. (1998). Transpression and transtension zones. In Continental Transpressional and Transtensional Tectonics (eds. Holdsworth, R.E., Strachan, R. E. and Dewey, J. F.), pp. 114, Geol. Soc. London Spec. Publ. 135.Google Scholar
Dholakia, S. K., Aydin, A., Pollard, D. D. and Zoback, M. D. (1998). Fault-controlled hydrocarbon pathways in the Monterey Formation, California. Am. Assoc. Petrol. Geol. Bull. 82:15511574.Google Scholar
Dibblee, T. W. Jr. (1977). Strike-slip tectonics of the San Andreas fault and its role in Cenozoic basin evolvement. In Late Mesozoic and Cenozoic Sedimentation and Tectonics in California, San Joaquin Geol. Soc. short course, pp. 2638.Google Scholar
Dieterich, J. H. (1972). Time-dependent friction in rocks. J. Geophys. Res. 77:36903697.Google Scholar
Dieterich, J. H. (1978). Time-dependent friction and the mechanics of stick-slip. Pure Appl. Geophys. 116:790806.Google Scholar
Dieterich, J. H. and Linker, M. F. (1992). Fault stability under conditions of variable normal stress. Geophys. Res. Lett. 19:16911694.Google Scholar
Dieterich, J. H., Richards-Dinger, K. B. and Kroll, K. A. (2015). Modeling injection-induced seismicity with the physics-based earthquake simulator RSQSim. Seismol. Res. Lett. 86; doi 10.1785/0220150057.Google Scholar
DiGiovanni, A. A., Fredrich, J. T., Holcomb, D. J. and Olsson, W. A. (2007). Microscale damage evolution in compacting sandstone. In: The Relationship Between Damage and Localization (eds. Lewis, H. and Couples, G. D.), pp. 89103, Geol. Soc. Spec. Publ. 289.Google Scholar
Dimitrova, L. L., Holt, W. E., Haines, A. J. and Schultz, R. A. (2006). Towards understanding the history and mechanisms of Martian faulting: the contribution of gravitational potential energy. Geophys. Res. Lett. 33:L08202; doi: 10.1029/2005GL025307Google Scholar
Dmowska, R. and Rice, J. R. (1986). Fracture theory and its seismological applications. In Continuum Theories in Solid Earth Physics (ed. Teisseyre, R.), pp. 187255, PWN-Polish Scientific Publishers, Warsaw.Google Scholar
Dokka, R. K. and Travis, C. J. (1990). Late Cenozoic strike-slip faulting in the Mojave Desert, California. Tectonics 9:311340.Google Scholar
Dong, P. and Pan, J. (1991). Elastic-plastic analysis of cracks in pressure-sensitive materials. Int. J. Solids Structures 28:11131127.Google Scholar
Douglas, K. J. (2002). The shear strength of rock masses. Unpublished PhD dissertation, University of New South Wales, Sydney, Australia.Google Scholar
Dowling, N. E. (1999). Mechanical Behavior of Materials: Engineering Methods for Deformation, Fracture, and Fatigue (2nd edn), Prentice-Hall, Upper Saddle River, New Jersey.Google Scholar
Dowling, N. E. and Thangjitham, S. (2000). An overview and discussion of basic methodology for fatigue. In Fatigue and Fracture Mechanics: 31st Volume (eds. Halford, G. R. and Gallagher, J. P.), pp. 336, Amer. Soc. Test. Mat. Spec. Publ. 1389, West Conshohocken, Penn.Google Scholar
Downey, M. W. (1984). Evaluating seals for hydrocarbon accumulations. Amer. Assoc. Petrol. Geol. Bull. 68:17521763.Google Scholar
Du, Y. and Aydin, A. (1991). Interaction of multiple cracks and formation of echelon crack arrays. Int. J. Num. Anal. Methods Geomech. 15:205218.Google Scholar
Du, Y. and Aydin, A. (1993). The maximum distortional strain energy density criterion for shear fracture propagation with applications to the growth paths of en échelon faults. Geophys. Res. Lett. 20:10911094.Google Scholar
Du, Y. and Aydin, A. (1995). Shear fracture patterns and connectivity at geometric complexities along strike-slip faults. J. Geophys. Res. 100:18,093–18,102.Google Scholar
Du Bernard, X., Labame, P., Darcel, C., Davy, P. and Bour, O. (2002a). Cataclastic slip band distribution in normal fault damage zones, Nubian sandstones, Suez rift. J. Geophys. Res. 107:2141; doi: 10.129/2001JB000493.Google Scholar
Du Bernard, X., Eichhubl, P. and Aydin, A. (2002b). Dilation bands: a new form of localized failure in granular media. Geophys. Res. Lett. 29: 2176; doi: 10.1029/2002GL015966.Google Scholar
Dugdale, D. S. (1960). Yielding in steel sheets containing slits. J. Mech. Phys. Solids 8:100104.Google Scholar
Dunand, D. C., Schuh, C. and Goldsby, D. L. (2001). Pressure-induced transformation plasticity of H2O ice. Phys. Rev. Lett. 86:668.Google Scholar
Dunn, D. E., LaFountain, L. J. and Jackson, R. E. (1973). Porosity dependence and mechanism of brittle fracture in sandstones. J. Geophys. Res. 78:24032417.Google Scholar
Dunne, W. M. and Ferrill, D. A. (1988). Blind thrust systems. Geology 16:3336.Google Scholar
Dunne, W. M. and Hancock, P. L. (1994). Paleostress analysis of small-scale brittle structures. In Continental Deformation (ed. Hancock, P. L.), pp. 101120, Pergamon, New York.Google Scholar
Durney, D. W. (1974). The influence of stress concentrations on the lateral propagation of pressure solution zones and surfaces. Geol. Soc. Austrail. Tecton. Struct. Newslett. 3:19.Google Scholar
Durney, D. W. and Kisch, H. (1994). A field classification and intensity scale for first generation cleavages. J. Aust. Geol. Geophys. 15:257295.Google Scholar
Dyer, R. (1988). Using joint interactions to estimate paleostress ratios. J. Struct. Geol. 10:685699.Google Scholar
Ebinger, C. J. (1989). Geometric and kinematic development of border faults and accommodation zones, Kivu-Rusizi rift, Africa. Tectonics 8:117133.Google Scholar
Edwards, M., Fornari, D., Malinverno, A. and Ryan, W. (1991). The regional tectonic fabric of the East Pacific Rise from 12°50’N to 15°10’N. J. Geophys. Res. 96:79958017.Google Scholar
Eftis, J. and Liebowitz, H. (1972). On the modified Westergaard equations for certain plane crack problems. Int. J. Fracture Mech. 8:383392.Google Scholar
Eichhubl, P. (2004). Growth of ductile opening-mode fractures in geomaterials. In The Initiation, Propagation, and Arrest of Joints and Other Fractures: Interpretations Based on Field Observations (eds. Cosgrove, J. W. and Engelder, T.), pp. 1124. Geol. Soc. London Spec. Publ. 231.Google Scholar
Eichhubl, P. and Boles, J. R. (2000). Focused fluid flow along faults in the Monterey Formation, coastal California. Geol. Soc. Am. Bull. 112:16671679.Google Scholar
Eichhubl, P., Davatzes, N. C. and Becker, S. P. (2009). Structural and diagenetic control of fluid migration along the Moab fault, Utah. Amer. Assoc. Petrol. Geol. Bull. 93:653681.Google Scholar
Eichhubl, P., Hooker, J. N. and Laubach, S. E. (2010). Pure and shear-enhanced compaction bands in Aztec sandstone. J. Struct. Geol. 32:18731886.Google Scholar
Eisenstadt, G. and DePaor, D. G. (1987). Alternative model of thrust-fault propagation. Geology 15:630633.Google Scholar
Elfgren, L. (editor) (1989). Fracture Mechanics of Concrete Structures. Chapman and Hall, London.Google Scholar
Elliott, D. (1976). The energy balance and deformation mechanism of thrust sheets. Phil. Trans. Royal Soc. London A, 283: 289312.Google Scholar
Elliott, S. J., Eichhubl, P. and Landry, C. J. (2014). Effects of coupled structural and diagenetic processes on deformation localization and fluid flow properties in sandstone reservoirs of the southwestern United States (abstract). Eos (Trans. AGU), MR23A–4336.Google Scholar
Ellis, S., Schreurs, G. and Panien, M. (2004). Comparisons between analogue and numerical models of thrust wedge development. J. Struct. Geol. 26:16591675.Google Scholar
Ellsworth, W. L. (2013). Injection-induced earthquakes. Science, 341, no. 6142; doi:10.1126/science.1225942.Google Scholar
Elmo, D., Rogers, S., Stead, D. and Eberhardt, E. (2014). Discrete Fracture Network approach to characterise rock mass fragmentation and implications for geomechanical upscaling. Mining Technol. 123:149161.Google Scholar
Elyasi, A. and Goshtasbi, K. (2016). Using different failure criteria in wellbore stability analysis. Geomech. Energy Environ. 2:1521.Google Scholar
Engelder, T. (1974). Cataclasis and the generation of fault gouge. Geol. Soc. Am. Bull. 85:15151522.Google Scholar
Engelder, T. (1985). Loading paths to joint propagation during a tectonic cycle: an example from the Appalachian Plateau, USA. J. Struct. Geol. 7:459476.Google Scholar
Engelder, T. (1987). Joints and shear fractures in rock. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 2769, Academic Press, New York.Google Scholar
Engelder, T. (1989). The analysis of pinnate joints in the Mount Desert Island Granite: Implications for post-intrusion kinematics in the coastal volcanic belt, Maine. Geology 17:564567.Google Scholar
Engelder, T. (1993). Stress Regimes in the Lithosphere. Princeton University Press, Princeton, New Jersey.Google Scholar
Engelder, T. (1994). Deviatoric stressitis: A virus infecting the Earth science community. Eos (Trans. Am. Geophys. Un.) 75:209212.Google Scholar
Engelder, T. (1999). Transitional-tensile fracture propagation: A status report. J. Struct. Geol. 21:10491055.Google Scholar
Engelder, T. (2007). Propagation velocity of joints: a debate over stable vs. unstable growth of cracks in the Earth. In Fractography of Glasses and Ceramics V (eds. Quinn, G. D., Varner, J. R. and Wightman, M.), pp. 500525, American Ceramic Society, Westerville, OH.Google Scholar
Engelder, T. and Fischer, M. P. (1996). Loading configurations and driving mechanisms for joints based on the Griffith energy-balance concept. Tectonophysics 256:253277.Google Scholar
Engelder, T. and Geiser, P. (1980). On the use of regional joint sets as trajectories of paleostress fields during the development of the Appalachian plateau, New York. J. Geophys. Res. 85:63196341.Google Scholar
Engelder, T. and Gross, M. R. (1993). Curving cross joints and the lithospheric stress field in eastern North America. Geology 21:817820.Google Scholar
Engelder, T. and Lacazette, A. (1990). Natural hydraulic fracturing. In Rock Joints (eds. Barton, N. and Stephansson, O.), pp. 3543, Balkema, Rotterdam.Google Scholar
Engelder, T. and Marshak, S. (1985). Disjunctive cleavage formed at shallow depths in sedimentary rocks. J. Struct. Geol. 7:327343.Google Scholar
Engelder, T., Fischer, M. P. and Gross, M. R. (1993). Geological aspects of fracture mechanics. Geol. Soc. Am. Short Course Notes.Google Scholar
Engelder, T., Lash, G. G. and Uzcátegui, R. S. (2009). Joint sets that enhance production from Middle and Upper Devonian gas shales of the Appalachian Basin. Am. Assoc. Petrol. Geol. Bull. 93:857889.Google Scholar
Erdogan, F. and Biricikoglu, V. (1973). Two bonded half planes with a crack going through the interface. Int. J. Eng. Sci. 11:745766.Google Scholar
Erdogan, F. and Sih, G. C. (1963). On the crack extension in plates under plane loading and transverse shear. J. Basic Eng. 85:519527.Google Scholar
Erickson, S. G. (1996). Influence of mechanical stratigraphy on folding vs. faulting. J. Struct. Geol. 18:443450.Google Scholar
Erslev, E. A. (1991). Trishear fault-propagation folding. Geology 19:617620.Google Scholar
Evans, A. G. and Wiederhorn, S. M. (1974). Crack propagation and failure prediction in silicon nitride at elevated temperatures. J. Mat. Sci. 9:270278.Google Scholar
Evans, B. and Kohlstedt, D. L. (1995). Rheology of rocks. In Rock Physics and Phase Relations—A Handbook of Physical Constants (ed. Ahrens, T. J.), American Geophysical Union Reference Shelf 3, pp. 148165.Google Scholar
Evans, B., Fredrich, J. T. and Wong, T.-f. (1990). The brittle–ductile transition in rocks: recent experimental and theoretical progress. In The Brittle–Ductile Transition in Rocks (eds. Duba, A. G., Durham, W. B., Handin, J. W. and Wang, H. F.), American Geophysical Union Geophys. Monog. 56, pp. 120.Google Scholar
Evans, J. P. and Bradbury, K. K. (2004). Faulting and fracturing of nonnwelded Bishop Tuff, eastern California: deformation mechanisms in very porous materials in the vadose zone. Vadose Zone J. 3:602623.Google Scholar
Exner, U. and Grassemann, B. (2010). Deformation bands in gravels: displacement gradients and heterogeneous strain. J. Geol. Soc. London 167:905913.Google Scholar
Exner, U. and Tschegg, C. (2012). Preferential cataclastic grain size reduction of feldspar in deformation bands in poorly consolidated arkosic sands. J. Struct. Geol. 43:6372.Google Scholar
Exner, U., Kaiser, J. and Gier, S. (2013). Deformation bands evolving from dilation to cementation bands in a hydrocarbon reservoir (Vienna Basin, Austria). Marine Petrol. Geol. 43:504515.Google Scholar
Fall, A., Eichhubl, P., Cumella, S. P., et al. (2012). Testing the basin-centered gas accumulation model using fluid inclusion observations: southern Piceance Basin, Colorado. Am. Assoc. Petrol. Geol. Bull. 96:22972318.Google Scholar
Farmer, I. W. (1983). Engineering Behaviour of Rocks (2nd edn), Chapman and Hall, London.Google Scholar
Fassett, C. I. (2016). Analysis of impact crater populations and the geochronology of planetary surfaces in the inner solar system. J. Geophys. Res. 121:19001926.Google Scholar
Faulds, J. E. and Varga, R. J. (1998). The role of accommodation zones and transfer zones in the regional segmentation of extended terranes. In Accommodation Zones and Transfer Zones: The Regional Segmentation of the Basin and Range Province (eds. Faulds, J. E. and Stewart, J. H.), pp. 145, Geol. Soc. Am. Spec. Pap. 323.Google Scholar
Faulkner, D. R., Jackson, C. A. I., Lunn, R. J.. et al. (2010). A review of recent developments concerning the structure, mechanics and fluid flow properties of fault zones. J. Struct. Geol. 32:15571575.Google Scholar
Felbeck, D. K. and Atkins, A. G. (1984). Strength and Fracture of Engineering Solids, Prentice-Hall, Englewood Cliffs, NJ.Google Scholar
Fender, M., Lechenault, F., and Daniels, K. E. (2010). Universal shapes formed by interacting cracks. Phys. Rev. Lett. 105:125505.Google Scholar
Ferrill, D. A. and Morris, A. P. (2003). Dilational normal faults. J. Struct. Geol. 25:183196.Google Scholar
Ferrill, D. A., Stamatakos, J. A. and Sims, D. (1999). Normal fault corrugation: implications for growth and seismicity of active normal faults. J. Struct. Geol. 21:10271038.Google Scholar
Ferrill, D. A., McGinnis, R. N., Morris, A. P., et al. (2014). Control of mechanical stratigraphy on bed-restricted jointing and normal faulting: Eagle Ford Formation, south-central Texas. Am. Assoc. Petrol. Geol. Bull. 98:24772506.Google Scholar
Ferrill, D. A., Morris, A. P., McGinnis, R. N., et al. (2016). Observations on normal-fault scarp morphology and fault system evolution of the Bishop Tuff in the Volcanic Tableland, Owens Valley, California, USA. Lithosphere 8:238253.Google Scholar
Ferrill, D. A., Morris, A. P., McGinnis, R. N., Smart, K. J. and Wigginton, S. S. (2017). Mechanical stratigraphy and normal faulting. J. Struct. Geol. 94:275302.Google Scholar
Fialko, Y. I. and Rubin, A. M. (1998). Thermodynamics of lateral dike propagation: implications for crustal accretion at slow spreading mid-ocean ridges. J. Geophys. Res. 103:25012514.Google Scholar
Fidan, M. Nielsen, H., Anderson, N., et al. (2012). Characterization of overburden anisotropy improves wellbore stability in North Sea field. World Oil, May 2012:13.Google Scholar
Field, J. E. (1971). Brittle fracture: its study and application. Contemp. Phys. 12:131.Google Scholar
Fields, R. J. and Ashby, M. F. (1976). Finger-like crack growth in solids and liquids. Philos. Mag. 33:3348.Google Scholar
Fink, J. H. (1985). Geometry of silicic dikes beneath the Inyo Domes, California. J. Geophys. Res. 90:11,127–11,133.Google Scholar
Finkbeiner, T., Zoback, M., Stump, B. B. and Flemings, P. B. (1998). In situ stress, pore pressure, and hydrocarbon migration in the South Eugene Island field, Gulf of Mexico. In Overpressures in Petroleum Exploration (eds. Mitchell, A. and Grauls, D.), pp. 103110, Proc. Workshop Pau, France, April 1998, Bull. Centre Rech. Elf Explor. Prod., Mém. 22.Google Scholar
Finkbeiner, T., Zoback, M., Flemings, P. and Stump, B. (2001). Stress, pore pressure, and dynamically constrained hydrocarbon columns in the South Eugene Island 330 field, northern Gulf of Mexico. Bull. Am. Assoc. Petrol. Geol. 85:10071031.Google Scholar
Fischer, G. J. and Paterson, M. S. (1989). Dilatancy during rock deformation at high temperatures and pressures. J. Geophys. Res. 94:17,60717,617.Google Scholar
Fischer, M. P., Gross, M. R., Engelder, T. and Greenfield, R. J. (1995). Finite-element analysis of the stress distribution around a pressurized crack in a layered elastic medium: implications for the spacing of fluid-driven joints in bedded sedimentary rock. Tectonophysics 247:4964.Google Scholar
Fletcher, R. C. (1982). Coupling of diffusional mass transport and deformation in a tight rock. Tectonophysics 83:275291.Google Scholar
Fletcher, R. C. and Pollard, D. D. (1981). Anticrack model for pressure solution surfaces. Geology 9:419424.Google Scholar
Fletcher, R. C. and Pollard, D. D. (1999). Can we understand structural and tectonic processes and their products without appeal to a complete mechanics? J. Struct. Geol. 21:10711088.Google Scholar
Flodin, E. and Aydin, A. (2004a). Faults with asymmetric damage zones in sandstone, Valley of Fire State Park, southern Nevada. J. Struct. Geol. 26:983988.Google Scholar
Flodin, E. and Aydin, A. (2004b). Evolution of a strike-slip fault network, Valley of Fire State Park, southern Nevada. Geol. Soc. Am. Bull. 116:4259.Google Scholar
Folger, P. and Tiemann, M. (2016). Human-induced earthquakes from deep-well injection: a brief overview. Congressional Research Service Report for Members of Congress 7-5700, R43836, 29 pp.Google Scholar
Fookes, P. G. and Parrish, D. G. (1969). Observations on small-scale structural discontinuities in the London Clay and their relationship to regional geology. Quart. J. Eng. Geol. 1:217240.Google Scholar
Fortin, J., Stanchits, S., Dresen, G. and Gueguen, Y. (2009). Acoustic emissions monitoring during inelastic deformation of porous sandstone: comparison of three modes of deformation. Pure Appl. Geophys. 166:823841.Google Scholar
Fossen, H. (2010a). Structural Geology, Cambridge University Press, 463 pp.Google Scholar
Fossen, H. (2010b). Deformation bands formed during soft-sediment deformation: observations from SE Utah. Marine Petrol. Geol. 27:215222.Google Scholar
Fossen, H. and Bale, A. (2007). Deformation bands and their influence on fluid flow. Amer. Assoc. Petrol. Geol. Bull. 91:16851700.Google Scholar
Fossen, H. and Cavalcante, G. C. G. (2017). Shear zones: a review. Earth-Sci. Rev. 171:434455.Google Scholar
Fossen, H. and Gabrielsen, R. H. (1996). Experimental modeling of extensional fault systems by use of plaster. J. Struct. Geol. 18:673687.Google Scholar
Fossen, H. and Gabrielsen, R. H. (2005). Strukturgeologi, Fagbokforlaget (in Norwegian), Bergen, Norway, 375 pp.Google Scholar
Fossen, H. and Hesthammer, J. (1997). Geometric analysis and scaling relations of deformation bands in porous sandstone. J. Struct. Geol. 19:14791493.Google Scholar
Fossen, H. and Hesthammer, J. (1998). Deformation bands and their significance in porous sandstone reservoirs. First Break 16:2125.Google Scholar
Fossen, H. and Hesthammer, J. (2000). Possible absence of small faults in the Gullfaks Field, northern North Sea: implications for downscaling of faults in some porous sandstones. J. Struct. Geol. 22:851863.Google Scholar
Fossen, H. and Rotevatn, A. (2016). Fault linkage and relay structures in extensional settings—a review. Earth-Sci. Rev. 154:1428.Google Scholar
Fossen, H. and Tikoff, B. (1998). Extended models of transpression and transtension, and application to tectonic settings. In Continental Transpressional and Transtensional Tectonics (eds. Holdsworth, R. E., Strachan, R. E. and Dewey, J. F.), pp. 1533, Geol. Soc. London Spec. Publ. 135.Google Scholar
Fossen, H., Johansen, T. E. S., Hesthammer, J. and Rotevatn, A. (2005). Fault interaction in porous sandstone and implications for reservoir management: examples from southern Utah. Amer. Assoc. Petrol. Geol. Bull. 89:15931606.Google Scholar
Fossen, H., Schultz, R. A., Shipton, Z. K. and Mair, K. (2007). Deformation bands in sandstone: a review. J. Geol. Soc. London 164:755769.Google Scholar
Fossen, H., Schultz, R. A., Rundhovde, E., Rotevatn, A. and Buckley, S. J. (2010). Fault linkage and graben stepovers in Canyonlands (Utah) and the North Sea Viking Graben, with implications for hydrocarbon migration and accumulation. Amer. Assoc. Petrol. Geol. Bull. 94:597613.Google Scholar
Fossen, H., Schultz, R. A. and Torabi, A. (2011). Conditions and implications for compaction band formation in Navajo Sandstone, Utah. J. Struct. Geol. 33:14771490.Google Scholar
Fossen, H., Zuluaga, L. F., Ballas, G., Soliva, R. and Rotevatn, A. (2015). Contractional deformation of porous sandstone: insights from the Aztec Sandstone, SE Nevada, USA. J. Struct. Geol. 74:172184.Google Scholar
Fossen, H., Soliva, R., Ballas, G., et al. (2017). A review of deformation bands in reservoir sandstones: geometries, mechanisms and distribution. In Subseismic-Scale Reservoir Deformation (eds. Ashton, M., Dee, S. J. and Wennberg, O. P.), Geol. Soc. London Spec. Publ. 459; doi:10.1144/SP459.4.Google Scholar
Fox, P. J. and Gallo, D. G. (1984). A tectonic model for ridge-transform-ridge plate boundaries: implications for the structure of oceanic lithosphere. Tectonophysics 104:205242.Google Scholar
Franklin, J. A. (1993). Empirical design and rock mass characterization. In Comprehensive Rock Engineering (ed. Hudson, J. A.), vol. 2 (ed. Fairhurst, C.), pp. 795806, Pergamon, New York.Google Scholar
Frankowicz, E. and McClay, K. R. (2010). Extensional fault segmentation and linkages, Bonaparte Basin, outer North West Shelf, Australia. Am. Assoc. Petrol. Geol. Bull. 94:9771010.Google Scholar
Fréchette, V. D. (1972). The fractography of glass. In Introduction to Glass Science (ed. Pye, D. L.), pp. 432450, Plenum Press, New York.Google Scholar
Fréchette, V. D. (1990). Failure Analysis of Brittle Materials. Advances in Ceramics, vol. 28, Am. Ceram. Soc., Westerville, Ohio.Google Scholar
Fredrich, J. T., Evans, B. and Wong, T.-f. (1989). Micromechanics of the brittle to plastic transition in Carrara marble. J. Geophys. Res. 94:41294145.Google Scholar
Freed, A. M. (2005). Earthquake triggering by static, dynamic, and postseismic stress transfer. Annu. Rev. Earth Planet. Sci. 33:335367.Google Scholar
Freed, A. M. and Lin, J. (1998). Time-dependent changes in failure stress following thrust earthquakes. J. Geophys. Res. 103:24,39324,409.Google Scholar
Freeze, A. R. and Cherry, J. A. (1979). Groundwater, Prentice-Hall, New Jersey.Google Scholar
Freund, L. B. (1990). Dynamic Fracture Mechanics, Cambridge University Press, New York, 563 pp.Google Scholar
Freund, R. (1970). Rotation of strike-slip faults in Sistan, southeast Iran. J. Geol. 78:188200.Google Scholar
Freund, R. (1974). Kinematics of transform and transcurrent faults. Tectonophysics 21:93134.Google Scholar
Friedman, M. and Logan, J. M. (1973). Lüders’ bands in experimentally deformed sandstone and limestone. Geol. Soc. Am. Bull. 84:14651476.Google Scholar
Frohlich, C. (2012). Two-year survey comparing earthquake activity and injection-well locations in the Barnett shale, Texas. Proc. US Nat. Acad. Sci. 109:13,93413,938.Google Scholar
Fueten, F. and Robin, P.-Y. F. (1992). Finite element modeling of the propagation of a pressure solution cleavage seam. J. Struct. Geol. 14:953962.Google Scholar
Gabrielsen, R. H. and Kløvjan, O. S. (1997). Late Jurassic–early Cretaceous caprocks of the southwestern Barents Sea: fracture systems and rock mechanical properties. In Hydrocarbon Seals: Importance for Exploration and Production (eds. Møller-Pedersen, P. and Koestler, A. G.), pp. 7389, Norwegian Petrol. Soc. Spec. Publ. 7.Google Scholar
Gabrielsen, R. H. and Koestler, A. G. (1987). Description and structural implications of fractures in the late Jurassic sandstones of the Troll Field, northern North Sea. Norsk Geologisk Tidsskrift 67:371381.Google Scholar
Gabrielsen, R. H., Aarland, R.-K. and Alsaker, E. (1998). Identification and spatial distribution of fractures in porous, silisciclastic sediments. In Structural Geology in Reservoir Characterization (eds. Coward, M. P., Daltaban, T. S. and Johnson, H.), pp. 4964, Geol. Soc. London Spec. Publ. 127.Google Scholar
Gale, J. F. W., Laubach, S. E., Olson, J. E., Eichhubl, P. and Fall, A. (2014). Natural fractures in shale—a review and new observations. Am. Assoc. Petrol. Geol. Bull. 98:21652216.Google Scholar
Gallagher, J. J., Friedman, M., Handin, J. and Sowers, G. M. (1974). Experimental studies relating to microfracture in sandstone. Tectonophysics 21:203247.Google Scholar
Gamond, J. F. (1983). Displacement features associated with fault zones: a comparison between observed examples and experimental models. J. Struct. Geol. 5:3345.Google Scholar
Gamond, J. F. (1987). Bridge structures as sense of displacement criteria on brittle faults. J. Struct. Geol. 9:609620.Google Scholar
Garfunkel, Z. (1981). Internal structure of the Dead Sea leaky transform (rift) in relation to plate tectonics. Tectonophysics 80:81108.Google Scholar
Gawthorpe, R. L. and Leeder, M. R. (2000). Tectono-sedimentary evolution of active extensional basins. Basin Res. 12:195218.Google Scholar
Geiser, P. A. (1988). The role of kinematics in the construction and analysis of geological cross sections in deformed terranes. In Geometries and Mechanisms of Thrusting, with Special Reference to the Appalachians (eds. Mitra, G. and Wojtal, S.), Geol. Soc. Am. Spec. Pap. 222, pp. 4776.Google Scholar
Geiser, P. A. and Sansone, S. (1981). Joints, microfractures, and the formation of solution cleavage in limestone. Geology 9:280285.Google Scholar
Gercek, H. (2007). Poisson’s ratio for rocks. Int. J. Rock Mech. Min. Sci. 44:113.Google Scholar
Germanovich, L. N. and Cherepanov, G. P. (1995). On some general properties of strength criteria. Int. J. Fracture 71:3756.Google Scholar
Germanovich, L. N., Salganik, R. L., Dyskin, A. V. and Lee, K. K. (1994). Mechanisms of brittle fracture of rock with pre-existing cracks in compression. Pure Appl. Geophys. 143:117149.Google Scholar
Gerya, T. (2012). Origin and models of oceanic transform faults. Tectonophysics 522523:3454.Google Scholar
Giba, M., Walsh, J. J. and Nicol, A. (2012). Segmentation and growth of an obliquely reactivated normal fault. J. Struct. Geol. 39:253267.Google Scholar
Gibbs, A. D. (1984). Structural evolution of extensional basin margins. J. Geol. Soc. London 141:609620.Google Scholar
Gibbs, A. D. (1990). Linked fault families in basin formation. J. Struct. Geol. 12:795803.Google Scholar
Gibson, R. G. (1998). Physical character and fluid-flow properties of sandstone-derived fault zones. In Structural Geology in Reservoir Characterization (eds. Coward, M. P., Daltaban, T. S. and Johnson, H.), pp. 8397, Geol. Soc. London Spec. Publ. 127.Google Scholar
Gillespie, P. A., Walsh, J. J. and Watterson, J. (1992). Limitations of dimension and displacement data from single faults and the consequences for data analysis and interpretation. J. Struct. Geol. 14:11571172.Google Scholar
Gillespie, P. A., Howard, C. B., Walsh, J. J. and Watterson, J. (1993). Measurement and characterisation of spatial distributions of fractures. Tectonophysics 226:113141.Google Scholar
Ginsberg, J. H. and Genin, J. (1977). Statics, Wiley, New York.Google Scholar
Goehring, L., Mahadevan, L. and Morris, S. W. (2009). Nonequilibrium scale selection mechanism for columnar jointing. Proc. US Nat. Acad. Sci. 106:387392.Google Scholar
Goetze, C. and Evans, B. (1979). Stress and temperature in the bending lithosphere as constrained by experimental rock mechanics. Geophys. J. Royal Astron. Soc. 59:463478.Google Scholar
Gokceoglu, C., Somnez, H. and Kayabasi, A. (2003). Predicting the deformation moduli of rock masses. Int. J. Rock Mech. Min. Sci. 40:701710.Google Scholar
Golombek, M. P. and Phillips, R. J. (2010). Mars tectonics. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 183232, Cambridge University Press.Google Scholar
Golombek, M. P., Anderson, F. S. and Zuber, M. T. (2001). Martian wrinkle ridge topography: evidence for subsurface faults from MOLA. J. Geophys. Res. 106:23,81123,821.Google Scholar
Gomberg, J., Blanpied, M. L. and Beeler, N. M. (1997). Transient triggering of near and distant earthquakes. Bull. Seismol. Soc. Am. 87:294309.Google Scholar
Goodier, J. N. (1968). Mathematical theory of equilibrium cracks. Fracture, An Advanced Treatise 2:166.Google Scholar
Goodman, R. E. (1976). Methods of Engineering in Discontinuous Rocks, West, St. Paul.Google Scholar
Goodman, R. E. (1989). Introduction to Rock Mechanics (2nd edn), Wiley, New York.Google Scholar
Goodman, R. E. and Shi, G. (1985). Block Theory and its Application to Rock Engineering, Prentice-Hall, New Jersey.Google Scholar
Gordon, F. R. and Lewis, J. D. (1980). The Meckering and Calingiri earthquakes October 1968 and March 1970. West. Aust. Geol. Surv. Bull. 126, 229 pp.Google Scholar
Goscombe, B. D., Passchier, C. W. and Hand, M. (2004). Boudinage classification: end-member boudin types and modified boudin structures. J. Struct. Geol. 26:739763.Google Scholar
Goudy, C. L., and Schultz, R. A. (2005). Dike intrusions beneath grabens south of Arsia Mons, Mars. Geophys. Res. Lett. 32:5, doi: 10.1029/2004GL021977.Google Scholar
Goudy, C. L., Schultz, R. A., and Gregg, T. K. P. (2005). Coulomb stress changes in Hesperia Planum, Mars, reveal regional thrust fault reactivation. J. Geophys. Res. 110:E10005; doi:10.1029/2004JE002293.Google Scholar
Gowd, T. N. and Rummel, F. (1980). Effect of confining pressure on fracture behavior of a porous rock. Int. J. Rock Mech. Min. Sci. 17:225229.Google Scholar
Grady, D. E. and Kipp, M. E. (1987). Dynamic rock fragmentation. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 429475, Academic Press, New York.Google Scholar
Graham, B., Antonellini, M. and Aydin, A. (2003). Formation and growth of normal faults in carbonates within a compressive environment. Geology 31:1114.Google Scholar
Grasemann, B., Martel, S. J. and Passchier, C. (2005). Reverse and normal drag along a single dip-slip fault. J. Struct. Geol. 27:9991010.Google Scholar
Grasso, J. R., and Wittlinger, G. (1990). 10 years of seismic monitoring over a gas field area. Seismol. Soc. Am. Bull. 80: 450473.Google Scholar
Graveleau, F., Malavieille, J. and Dominguez, S. (2012). Experimental modelling of orogenic wedges: a review. Tectonophysics 538540:166.Google Scholar
Gray, D., Anderson, P., Logel, J., et al. (2012). Estimation of stress and geomechanical properties using 3D seismic data. First Break 30:5968.Google Scholar
Green, A. E. and Sneddon, I. N. (1950). The distribution of stresses in the neighborhood of a flat elliptical crack in an elastic solid. Proc. Cambridge Phil. Soc. 46:159163.Google Scholar
Green, H. W., II (1984). How and why does olivine transform to spinel? Geophys. Res. Lett. 11:817820.Google Scholar
Green, H. W., II and Burnley, P. C. (1989). A new, self-organizing, mechanism for deep-focus earthquakes. Nature 341:733737.Google Scholar
Green, H. W., II, Young, T. E., Walker, D. and Scholz, C. H. (1990). Anticrack-associated faulting at very high pressure in natural olivine. Nature 348:720722.Google Scholar
Green, H. W., II and Houston, H. (1995). The mechanics of deep earthquakes. Ann. Rev. Earth Planet. Sci. 23:169213.Google Scholar
Green, H. W., II and Marone, C. (2002). Instability of deformation. In Plastic Deformation of Minerals and Rocks (eds. Karato, S.-i. and Wenk, H.-R.), pp. 181199, Mineral. Soc. Am. Reviews in Mineral. and Geochem. v. 51.Google Scholar
Grieve, R. A. F. (1987). Terrestrial impact structures. Ann. Rev. Earth Planet. Sci. 15:245270.Google Scholar
Griffith, A. A. (1921). The phenomena of rupture and flow in solids. Phil. Trans. Royal Soc. London A 221:163198.Google Scholar
Griffith, A. A. (1924). The theory of rupture. In Proc. 1st Int. Congress Appl. Mech. (eds. Biezeno, C. B. and Burgers, J. M.), pp. 5563.Google Scholar
Griggs, D. T. (1936). Deformation of rocks under high confining pressures. J. Geol. 44:541577.Google Scholar
Griggs, D. T. and Handin, J. (1960). Observations on fracture and a hypothesis of earthquakes. In Rock Deformation (eds. Griggs, D. and Handin, J.), pp. 347373, Geol. Soc. Am. Memoir 79.Google Scholar
Griggs, D. T., Turner, F. J. and Heard, H. C. (1960). Deformation of rocks at 500° to 800° C. In Rock Deformation (eds. Griggs, D. and Handin, J.), pp. 39104, Geol. Soc. Am. Memoir 79.Google Scholar
Grimm, R. E. and Phillips, R. J. (1991). Gravity anomalies, compensation mechanisms, and the geodynamics of western Ishtar Terra, Venus. J. Geophys. Res. 96:83058324.Google Scholar
Grosfils, E. B., Schultz, R. A. and Kroeger, G. (2003). Geophysical exploration within northern Devils Lane graben, Canyonlands National Park, Utah: implications for sediment thickness and tectonic evolution. J. Struct. Geol. 25:455467.Google Scholar
Groshong, R. H. Jr. (1988). Low-temperature deformation mechanisms and their interpretation. Geol. Soc. Am. Bull. 100:13291360.Google Scholar
Groshong, R. H. Jr. (1989). Half-graben structures: balanced models of extensional fault-bend folds. Geol. Soc. Am. Bull. 101:96105.Google Scholar
Groshong, R. H. Jr. and Usdansky, S. I. (1988). Kinematic models of plane-roofed duplex styles. In Geometries and Mechanisms of Thrusting, with Special Reference to the Appalachians (eds. Mitra, G. and Wojtal, S.), pp. 197206, Geol. Soc. Am. Spec. Pap. 222.Google Scholar
Gross, M. R. (1993). The origin and spacing of cross joints: examples from the Monterey Formation, Santa Barbara coastline, California. J. Struct. Geol. 15:737751.Google Scholar
Gross, M. R. (1995). Fracture partitioning: failure mode as a function of lithology in the Monterey Formation of coastal California. Geol. Soc. Am. Bull. 107:779792.Google Scholar
Gross, M. R., Gutierrez-Alonzo, G., Bai, T., et al. (1997). Influence of mechanical stratigraphy and kinematics on fault scaling relationships. J. Struct. Geol. 19:171183.Google Scholar
Grueschow, E. and Rudnicki, J. W. (2005). Elliptic yield cap constitutive modeling for high porosity sandstone. Int. J. Solids Structures 42:45744587.Google Scholar
Gu, Y. and Wong, T.-f. (1994). Development of shear localization in simulated quartz gouge: effects of cumulative slip and gouge particle size. Pure Appl. Geophys. 143:387423.Google Scholar
Gudmundsson, A. (1992). Formation and growth of normal faults at the divergent plate boundary in Iceland. Terra Nova 4:464471.Google Scholar
Gudmundsson, A. (1995). Stress fields associated with oceanic transform faults. Earth Planet. Sci. Lett. 136:603614.Google Scholar
Gudmundsson, A. (2000). Fracture dimensions, displacements and fluid transport. J. Struct. Geol. 22:12211231.Google Scholar
Gudmundsson, A. (2004). Effects of Young’s modulus on fault displacement. Comptes Rendus Geoscience 336:8592.Google Scholar
Gudmundsson, A. (2011). Rock Fractures in Geological Processes, Cambridge University Press, Cambridge, 560 pp.Google Scholar
Gudmundsson, A. and Bäckström, K. (1991). Structure and development of the Sveeinagja graben, Northeast Iceland. Tectonophysics 200:111125.Google Scholar
Gupta, A. and Scholz, C. H. (2000a). A model of normal fault interaction based on observations and theory. J. Struct. Geol. 22:865879.Google Scholar
Gupta, A. and Scholz, C. H. (2000b). Brittle strain regime transition in the Afar depression: implications for fault growth and seafloor spreading. Geology 28:10781090.Google Scholar
Gupta, S. and Cowie, P. (2000). Processes and controls in the stratigraphic development of extensional basins. Basin Res. 12:185194.Google Scholar
Gupta, S., Cowie, P. A., Dawers, N. H. and Underhill, J. R. (1998). A mechanism to explain rift-basin subsidence and stratigraphic patterns through fault-array evolution. Geology 26:595598.Google Scholar
Gürbüz, A. (2014). Geometric characteristics of pull-apart basins. Lithosphere 2:199206.Google Scholar
Gutierrez, M. and Homand, S. (1998). Formulation of a basic chalk constitutive model. In Chalk V – Chalk Geomechanics. Report prepared for Joint Chalk Research Phase V, Norwegian Geotechnical Institute, Oslo, 50 pp.Google Scholar
Haddad, M. and Sepehrnoori, K. (2014). Cohesive fracture analysis to model multiple-stage fracturing in quasibrittle shale formations. 2014 SIMULIA Conference, www.3ds.com/simulia, 15 pp.Google Scholar
Haddad, M. and Sepehrnoori, K. (2015). Simulation of hydraulic fracturing in quasi-brittle shale formations using characterized cohesive layer: stimulation controlling factors. J. Unconv. Oil Gas Res. 9:6583.Google Scholar
Haimson, B. C. (2001). Fracture-like borehole breakouts in high-porosity sandstone: are they caused by compaction bands? Physics Chem. Earth 26:1520.Google Scholar
Haimson, B. C. and Fairhurst, C. (1967). Initiation and extension of hydraulic fractures in rocks. Soc. Petrol. Eng. J. 7:310318.Google Scholar
Haimson, B. C. and Kim, R. Y. (1971). Mechanical behavior of rock under cyclic failure. Proc. US Rock Mech. Symp. 13:845862.Google Scholar
Haimson, B. C. and Rummel, F. (1982). Hydrofracturing stress measurements in the Iceland research drilling project drill hole at Reydarfjördur, Iceland. J. Geophys. Res. 87:66316649.Google Scholar
Hancock, P. L. (1972). The analysis of en echelon veins. Geol. Mag. 109:269276.Google Scholar
Hancock, P. L. (1985). Brittle microtectonics: principles and practice. J. Struct. Geol. 7:437457.Google Scholar
Handin, J. (1969). On the Coulomb–Mohr failure criterion. J. Geophys. Res. 74:53435348.Google Scholar
Handin, J., Hager, R. V. Jr., Friedman, M. and Feather, J. N. (1963). Experimental deformation of sedimentary rocks under confining pressure: pore pressure tests. Am. Assoc. Petrol. Geol. Bull. 47:717755.Google Scholar
Hansen, B. (1958). Line ruptures regarded as narrow rupture zones: basic equations based on kinematic considerations. Proc. Brussels Conf. 58 on Earth Pressure Problems 1:3948.Google Scholar
Hansen, F. D., Hardin, E. L., Rechard, R. P., et al. (2010). Shale disposal of U.S. high-level radioactive waste. Report SAND2010–2843, Sandia National Laboratories, Albuquerque, New Mexico, 148 pp.Google Scholar
Harding, T. P. (1974). Petroleum traps associated with wrench faults. Am. Assoc. Petrol. Geol. Bull. 58:12901304.Google Scholar
Harding, T. P. and Lowell, J. D. (1979). Structural styles, their plate-tectonic habitats, and hydrocarbon traps in petroleum provinces. Am. Assoc. Petrol. Geol. Bull. 63:10161058.Google Scholar
Harding, T. P., Vierbuchen, R. C. and Christie-Blick, N. (1985). Structural styles, plate-tectonic settings, and hydrocarbon traps of divergent (transtensional) wrench faults. In Strike-Slip Deformation, Basin Formation, and Sedimentation (eds. Biddle, K. T. and Christie-Blick, N.), pp. 5177, Soc. Econ. Paleon. Miner., Spec. Publ. 37.Google Scholar
Hardy, S. and Ford, M. (1997). Numerical modeling of trishear fault-propagation folding. Tectonics 16:841854.Google Scholar
Hardy, S. and Allmendinger, R.W. (2011). Trishear: a review of kinematics, mechanics, and applications. In Thrust Fault-Related Folding (eds. McClay, K. R., Shaw, J. H. and Suppe, J.), pp. 95119, Amer. Assoc. Petrol. Geol. Mem. 94.Google Scholar
Harland, W. B. (1957). Exfoliation joints and ice action. J. Glaciol. 3:810.Google Scholar
Harland, W. B. (1971). Tectonic transpression in Caledonian Spitzbergen. Geol. Mag. 108:2742.Google Scholar
Harris, R. A. (1998). Introduction to special section: Stress triggers, stress shadows, and implications for seismic hazard. J. Geophys. Res. 103:24,34724,358.Google Scholar
Harris, R. A. and Day, S. M. (1993). Dynamics of fault interaction: parallel strike-slip faults. J. Geophys. Res. 98:44614472.Google Scholar
Harris, R. A. and Simpson, R. W. (1998). Suppression of large earthquakes by stress shadows: a comparison of Coulomb and rate-and-state failure. J. Geophys. Res. 103:24,43924,451.Google Scholar
Hashida, T., Oghikubo, H., Takahashi, H. and Shoji, T. (1993). Numerical simulation with experimental verification of the fracture behavior in granite under confining pressures based on the tension-softening model. Int. J. Fracture 59:227244.Google Scholar
Hatheway, A. W. (1996). Fractures; discontinuities that control your project (Perspective No. 28). AEG News 39/4:1922.Google Scholar
Hatton, C. G., Main, I. G. and Meredith, P. G. (1993). A comparison of seismic and structural measurements of scaling exponents during tensile subcritical crack growth. J. Struct. Geol. 15:14851495.Google Scholar
Hawkes, C. D., Bachu, S. and Mclellan, P. J. (2005). Geomechanical factors affecting geologic storage of CO2 in depleted oil and gas reservoirs. J. Can. Petrol. Technol. 44:5261.Google Scholar
Hawkes, I. and Mellor, M. (1970). Uniaxial testing in rock mechanics laboratories. Eng. Geol. 4:177285.Google Scholar
Hayward, N. and Ebinger, C. J. (1996). Variations in the along-axis segmentation of the Afar rift system. Tectonics 15:244257.Google Scholar
Heald, M. T. (1956). Cementation of Simpson and St. Peter sandstones in parts of Oklahoma, Arkansas, and Missouri. J. Geol. 64:1630.Google Scholar
Heald, M. T. (1959). Significance of stylolites in permeable sandstones. J. Sedimentary Res. 29:251253.Google Scholar
Heald, P. T., Spink, G. M. and Worthington, P. J. (1972). Post yield fracture mechanics. Mat. Sci. Engng. 10:129138.Google Scholar
Healy, D., Blenkinsop, T. G., Timms, N. E., et al. (2015). Polymodal faulting: time for a new angle on shear failure. J. Struct. Geol. 80:5771.Google Scholar
Heidbach, O., Tingay, M., Barth, A., et al. (2010). Global crustal stress pattern based on the World Stress Map database release 2008. Tectonophysics 482:215.Google Scholar
Helgeson, D. E. and Aydin, A. (1991). Characteristics of joint propagation across layer interfaces in sedimentary rocks. J. Struct. Geol. 13:897911.Google Scholar
Hennings, P. H., Olson, J. E. and Thompson, L. B. (2000). Combining outcrop data and three-dimensional structural models to characterize fractured reservoirs: an example from Wyoming. Amer. Assoc. Petrol. Geol. Bull. 84:830849.Google Scholar
Hennings, P. H., Allwardt, P., Paul, P., et al. (2012). Relationship between fractures, fault zones, stress, and reservoir productivity in the Suban gas field, Sumatra, Indonesia. Amer. Assoc. Petrol. Geol. Bull. 96:753772.Google Scholar
Hergarten, S. and Kenkmann, T. (2015). The number of impact craters on Earth: any room for further discoveries? Earth Planet. Sci. Lett. 425:187192.Google Scholar
Hesthammer, J., Johansen, T. E. S. and Watts, L. (2000). Spatial relationships within fault damage zones in sandstone. Marine Petrol. Geol. 17:873893.Google Scholar
Heuzé, F. E. (1980). Scale effects in the determination of rock mass strength and deformability. Rock Mech. 12:167192.Google Scholar
Hickman, R. J. (2004). Formulation and implementation of a constitutive model for soft rock, PhD dissertation, Virginia Polytechnic Institute and State University, Blacksburg, Virginia.Google Scholar
Hickman, R. J., and Gutierrez, M. S. (2007). Formulation of a three-dimensional rate-dependent constitutive model for chalk and porous rocks. Int. J. Numer. Anal. Meth. Geomech. 31: 583605.Google Scholar
Higgins, R. I. and Harris, L. B. (1997). The effect of cover composition on extensional faulting above re-activated basement faults; results from analog modeling. J. Struct. Geol. 19:8998.Google Scholar
Hill, D. P. and Prejean, S. (2007). Dynamic triggering. In  Earthquake Seismology, V. 4 (ed. Kanamori, H.), pp. 258288, Treatise on Geophysics (Schubert, G., ed. in chief), Elsevier, Amsterdam.Google Scholar
Hill, M. L. (1984). Earthquakes and folding, Coalinga, California. Geology 12:711712.Google Scholar
Hill, R. (1963). Elastic properties of reinforced solids: some theoretical principles. J. Mech. Phys. Solids 11:357372.Google Scholar
Hill, R. E. (1989). Analysis of deformation bands in the Aztec Sandstone, Valley of Fire, Nevada, MS Thesis, Geosciences Department, University of Nevada, Las Vegas.Google Scholar
Hillerborg, A. (1991). Application of the ficticious crack model to different types of materials. Int. J. Fracture 51:95102.Google Scholar
Hillerborg, A., Modéer, M. and Petersson, P.-E. (1976). Analysis of crack formation and crack growth in concrete by means of fracture mechanics and finite elements. Cement Concrete Res. 6:773782.Google Scholar
Hobbs, B. E., Means, W. D. and Williams, P. F. (1976). An Outline of Structural Geology, Wiley, New York.Google Scholar
Hodgkinson, K. M., Stein, R. S. and King, G. C. P. (1996). The 1954 Rainbow Mountain–Fairview Peak–Dixie Valley earthquakes: a triggered normal faulting sequence. J. Geophys. Res. 101:25,45925,471.Google Scholar
Hodgson, R. A. (1961). Classification of structures on joint surfaces. Am. J. Sci. 259:493502.Google Scholar
Hoek, E. (1983). Strength of jointed rock masses. Géotechnique 33:187223.Google Scholar
Hoek, E. (1990). Estimating Mohr–Coulomb friction and cohesion from the Hoek-Brown failure criterion. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 27:227229.Google Scholar
Hoek, E. (2002). A brief history of the development of the Hoek–Brown failure criterion. Online document, www.rocscience.com, 7 pages.Google Scholar
Hoek, E. (2007). Practical Rock Engineering. RocScience, available online at www.rocscience.com.Google Scholar
Hoek, E. and Bieniawski, Z. T. (1965). Brittle fracture propagation in rock under compression. Int. J. Fracture Mech. 1:137155.Google Scholar
Hoek, E. and Brown, E. T. (1980). Empirical strength criterion for rock masses. J. Geotech. Engng. Div. Am. Soc. Civ. Engrs. 106:10131035.Google Scholar
Hoek, E. and Brown, E. T. (1997). Practical estimates of rock mass strength. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 34:11651186.Google Scholar
Hoek, E. and Diederichs, M. (2006). Empirical estimates of rock mass modulus. Int. J. Rock Mech. Min. Sci. 43:203215.Google Scholar
Hoek, E., Kaiser, P. K. and Bawden, W. F. (1995). Support of Underground Excavations in Hard Rock, Balkema, Rotterdam.Google Scholar
Hoek, E., Carranza-Torres, C. and Corkum, B. (2002). Hoek–Brown failure criterion—2002 edition, 5th N. Am. Rock Mech. Symp. and 17th Tunnel. Assoc. Canada Conf., NARMS-TAC, 267–273.Google Scholar
Hoek, E., Carter, T. G. and Diederichs, M. S. (2013). Quantification of the Geological Strength Index chart. Paper presented at the 47th US Rock Mech./Geomech. Symp., ARMA, 13672.Google Scholar
Holcomb, D., Rudnicki, J. W., Issen, K. A. and Sternlof, K. (2007). Compaction localization in the Earth and the laboratory: state of the research and research directions. Acta Geotech. 2:115.Google Scholar
Holder, J., Olson, J. E. and Philip, Z. (2001). Experimental determination of subcritical crack growth parameters in sedimentary rock. Geophys. Res. Lett. 28:599602.Google Scholar
Holt, R. M., Fjaer, E., Nes, O.-M. and Alassi, H. T. (2011). A shaley look at brittleness. Paper ARMA 11–366 presented at the 45th US Rock Mechanics/Geomechanics Symposium, San Francisco, California, 26–29 June 2011.Google Scholar
Holt, R. M., Fjaer, E., Stenebråten, J. F. and Nes, O.-M. (2015). Brittleness of shales: relevance to borehole collapse and hydraulic fracturing. J. Petrol. Sci. Eng. 131:200209.Google Scholar
Holzhausen, G. R. and Johnson, A. M. (1979). Analyses of longitudinal splitting of uniaxially compressed rock cylinders. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 16:163177.Google Scholar
Hook, J. R. (2003). An introduction to porosity. Petrophysics 44:205212.Google Scholar
Hooker, J. N., Laubach, S. E. and Marrett, R. (2014). A universal power-law scaling exponent for fracture apertures in sandstones. Geol. Soc. Am. Bull. 126:13401362.Google Scholar
Hooper, D. M., Bursik, M. I. and Webb, F. H. (2003). Application of high-resolution, interferometric DEMs to geomorphic studies of fault scarps, Fish Lake Valley, Nevada-California, USA. Rem. Sens. Environ. 84:255267.Google Scholar
Hoppa, G. V., Tufts, B. R., Greenberg, R. and Geissler, P. E. (1999). Formation of cycloid features on Europa. Science 285:18991902.Google Scholar
Hoppin, R. A. (1961). Precambrian rocks and their relationship to Laramide structure along the east flank of the Bighorn Mountains near Buffalo, Wyoming. Geol. Soc. Am. Bull. 72:351368.Google Scholar
Horii, H. and Nemat-Nasser, S. (1985). Compression-induced microcrack growth in brittle solids: axial splitting and shear failure. J. Geophys. Res. 90:31053125.Google Scholar
Hornbach, M. J., DeShon, H. R., Ellsworth, W. L., et al. (2015). Causal factors for seismicity near Azle, Texas. Nature Comm. 6: 6728; doi:10.1038/ncomms7728.Google Scholar
Hubbert, M. K. and Rubey, W. W. (1959). Role of fluid pressure in mechanics of overthrust faulting. Pts. I & II. Geol. Soc. Am. Bull. 70:115205.Google Scholar
Hubbert, M. K. and Willis, D. G. (1957). Mechanics of hydraulic fracturing. AIME Petrol. Trans. 210:153163.Google Scholar
Hucka, V. and Das, B. (1974). Brittleness determination of rocks by different methods. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 11:389392.Google Scholar
Hudleston, P. (1999). Strain compatibility and shear zones: is there a problem? J. Struct. Geol. 21:923932.Google Scholar
Hudson, J. A., Brown, E. T. and Fairhurst, C. (1971). Shape of the complete stress-strain curve for rock. Proc. US Rock Mech. Symp. 13:773795.Google Scholar
Hudson, J. A., Crouch, S. L. and Fairhurst, C. (1972). Soft, stiff, and servo-controlled testing machines: a review with reference to rock failure. Eng. Geol. 6:155189.Google Scholar
Hudson, J. A. and Harrison, J. P. (1997). Engineering Rock Mechanics: An Introduction to the Principles, Pergamon, New York.Google Scholar
Hughes, A. N. and Shaw, J. H. (2014). Fault displacement-distance relationships as indicators of contractional fault-related folding style. Am. Assoc. Petrol. Geol. Bull. 98:227251.Google Scholar
Ikari, M. J., Marone, C. and Saffer, D. M. (2011). On the relation between fault strength and frictional stability. Geology 39:8386.Google Scholar
Inglis, C. E. (1913). Stresses in a plate due to the presence of cracks and sharp corners. Royal Inst. Naval Architec. Trans. 55:219241.Google Scholar
Ingraffea, A. R. (1987). Theory of crack initiation and propagation in rock. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 71110, Academic Press, New York.Google Scholar
Ingram, G. M. and Urai, J. L. (1999). Top-seal leakage through faults and fractures: the role of mudrock properties. In Muds and Mudstones—Physical and Fluid Flow Properties (eds. Aplin, A. C., Fleet, A. J. and Macquaker, J. H. S.), pp. 125135, Geol. Soc. London Spec. Publ. 158.Google Scholar
Ingram, G. M., Urai, J. L. and Naylor, M. A. (1997). Sealing processes and top seal assessment. In Hydrocarbon Seals: Importance for Exploration and Production (eds. Møller-Pedersen, P. and Koestler, A. G.), pp. 165174, Elsevier, Amsterdam.Google Scholar
International Society for Rock Mechanics, Commission on Standardization of Laboratory and Field Tests (1978). Suggested methods for the quantitative description of discontinuities in rock masses. Int. J. Rock Mech. Min. Sci. Geomech. Abs. 15:319368.Google Scholar
Irwin, G. R. (1957). Analysis of stresses and strains near the end of a crack traversing a plate. J. Appl. Mech. 24:361364.Google Scholar
Irwin, G. R. (1960). Fracture mode transition for a crack traversing a plate. J. Basic Engng. 82:417425.Google Scholar
Irwin, G. R. (1962). The crack extension force for a part-through crack in a plate. J. Appl. Mech. 29:651654.Google Scholar
Irwin, G. R., Kies, J. A. and Smith, H. L. (1958). Fracture strengths relative to onset and arrest of crack propagation. Proc. Amer. Soc. Test. Mater. 58:640657.Google Scholar
Ishii, E. (2012). Microstructure and origin of faults in siliceous mudstone at the Horonobe Underground Research laboratory site, Japan. J. Struct. Geol. 34:2029.Google Scholar
Ishii, E., Sanada, H., Funaki, J., Sugita, Y. and Kurikami, H. (2011). The relationships among brittleness, deformation behavior, and transport properties in mudstones: an example from the Horonobe Underground Research Laboratory, Japan. J. Geophys. Res. 116: B09206; doi 10.1029/2011JB008279.Google Scholar
Issen, K. A. (2002). The influence of constitutive models on localization conditions for porous rock. Eng. Fracture Mech. 69:18911906.Google Scholar
Issen, K. A. and Challa, V. (2003). Conditions for dilation band formation in granular materials. In Proc. 16th ASCE Eng. Mech. Conf. 16th, pp. 14.Google Scholar
Issen, K. A. and Rudnicki, J. W. (2000). Conditions for compaction bands in porous rock. J. Geophys. Res. 105:21,52921,536.Google Scholar
Issen, K. A. and Rudnicki, J. W. (2001). Theory of compaction bands in porous rock. Phys. Chem. Earth A26:95100.Google Scholar
Jackson, C. A.-L., Bell, R. E., Rotevatn, A. and Tvedt, A. B. M. (2016). Techniques to determine the kinematics of synsedimentary normal faults and implications for fault growth models. In The Geometry and Growth of Normal Faults (eds. Childs, C., Holdsworth, R. E., Jackson, C. A.-L. et al.), Geol. Soc. London Spec. Publ. 439; doi:10.1144/SP429.22.Google Scholar
Jackson, J. A. and White, N. J. (1989). Normal faulting in the upper continental crust: observations from regions of active extension. J. Struct. Geol. 11:1536.Google Scholar
Jackson, M. D. and Pollard, D. D. (1988). The laccolith-stock controversy: new results from the southern Henry Mountains, Utah. Geol. Soc. Am. Bull. 100:117139.Google Scholar
Jackson, M. D. and Pollard, D. D. (1990). Flexure and faulting of sedimentary host rocks during growth of igneous domes, Henry Mountains, Utah. J. Struct. Geol. 12:185206.Google Scholar
Jackson, M. P. A. (1995). Retrospective salt tectonics. In Salt Tectonics (eds. Jackson, M. P. A., Roberts, D. G. and Snelson, S.), pp. 128, AAPG Memoir 65, Tulsa, Oklahoma.Google Scholar
Jackson, M. P. A. and Vendeville, B. C. (1994). Regional extension as a geologic trigger for diapirism. Geol. Soc. Am. Bull. 106:5773.Google Scholar
Jackson, P. and Sanderson, D. J. (1992). Scaling of fault displacement from the Badajoz-Córdoba shear zone, SW Spain. Tectonophysics 210:179190.Google Scholar
Jaeger, J. C. (1969). Elasticity, Fracture and Flow, with Engineering and Geological Applications, Chapman and Hall, London, 268 pp.Google Scholar
Jaeger, J. C. (1971). Friction of rocks and stability of rock slopes. Rankine lecture. Géotechnique 21:97134.Google Scholar
Jaeger, J. C. and Cook, N. G. W. (1979). Fundamentals of Rock Mechanics (3rd edn), Chapman and Hall, New York, 593 pp.Google Scholar
Jaeger, J. C., Cook, N. G. W. and Zimmerman, R. W. (2007). Fundamentals of Rock Mechanics (4th edn), Blackwell, Oxford, 475 pp.Google Scholar
Jamison, W. R. (1989). Fault-fracture strain in Wingate Sandstone. J. Struct. Geol. 11:959974.Google Scholar
Jamison, W. R. and Stearns, D. W. (1982). Tectonic deformation of Wingate Sandstone, Colorado National Monument. Amer. Assoc. Petrol. Geol. Bull. 66:25842608.Google Scholar
Jaroszewski, W. (1984). Fault and Fold Tectonics, Wiley, New York, 565 pp.Google Scholar
Jarrard, R. D. (1986). Terrane motion by strike-slip faulting of forearc slivers. Geology 14:780783.Google Scholar
Jing, L. and Stephansson, O. (2007). 10 – Discrete fracture network (DFN) method. Dev. Geotech. Eng. 85:365398.Google Scholar
Johansen, T. E. S. and Fossen, H. (2008). Internal geometry of fault damage zones in interbedded siliciclastic sediments. In The Internal Structure of Fault Zones: Implications for Mechanical and Fluid-Flow Properties (eds. Wibberley, C. A. J., Kurz, W., Imber, J., Holdsworth, R. E. and Collettini, C.), pp. 3556, Geol. Soc. London Spec. Publ. 299.Google Scholar
Johnson, A. M. (1970). Physical Processes in Geology: a Method for Interpretation of Natural Phenomena—Intrusions in Igneous Rocks, Fractures and Folds, Flow of Debris and Ice, Freeman, Cooper, and Co., San Francisco, California, 577 pp.Google Scholar
Johnson, A. M. (1980). Folding and faulting of strain-hardening sedimentary rocks. Tectonophysics 62:251278.Google Scholar
Johnson, A. M. (1995). Orientations of faults determined by premonitory shear zones. Tectonophysics 247:161238.Google Scholar
Johnson, A. M. (2001). Propagation of deformation bands in porous sandstones. Unpubl. manuscript, Purdue University, West Lafayette, Indiana, 73 pp.Google Scholar
Johnson, A. M. and Fletcher, R. C. (1994). Folding of Viscous Layers—Mechanical Analysis and Interpretation of Structures in Deformed Rock, Columbia University Press, New York, 461 pp.Google Scholar
Johnson, A. M. and Pollard, D. D. (1973). Mechanics of growth of some laccolithic intrusions in the Henry Mountains, Utah, I. Tectonophysics 18:261309.Google Scholar
Johnson, E. and Cleary, M. P. (1991). Implications of recent laboratory experimental results for hydraulic fractures. Paper presented at SPE Joint Rocky Mountain Regional Meeting and Low Permeability Reservoir Symposium, Soc. of Pet. Eng., Denver, Colorado, paper SPE 21846.Google Scholar
Johri, M., Zoback, M. D. and Hennings, P. (2014). A scaling law to characterize fault-damage zones at reservoir depths. Amer. Assoc. Petrol. Geol. Bull. 98:20572079.Google Scholar
Justo, J. L., Justo, E., Azañón, J. M., Durand, P. and Morales, A. (2010). The use of rock mass classification systems to estimate the modulus and strength of jointed rock. Rock Mech. Rock Eng. 43:287304.Google Scholar
Kachanov, M. (1992). Effective elastic properties of cracked solids: critical review of some basic concepts. Appl. Mech. Rev. 45:304335.Google Scholar
Kakimi, T. (1980). Magnitude–frequency relation for displacement of minor faults and its significance in crustal deformation. Bull. Geol. Survey Japan 31:467487.Google Scholar
Kamb, W. B. (1959). Petrofabric observations from Blue Glacier, Washington, in relation to theory and experiment. J. Geophys. Res. 64:18911909.Google Scholar
Kalthoff, J. F. (1971). On the characteristic angle or crack branching in ductile materials. Int. J. Fracture Mech. 7:478480.Google Scholar
Kanninen, M. F. and Popelar, C. H. (1985). Advanced Fracture Mechanics, Oxford University Press, New York, 563 pp.Google Scholar
Kaplan, M. F. (1961). Crack propagation and the fracture of concrete. Am. Concrete Inst. J. 58:591610.Google Scholar
Kaproth, B. M., Cashman, S. M. and Marone, C. (2010). Deformation band formation and strength evolution in unlithified sand: the role of grain breakage. J. Geophys. Res. 115:B12103; doi:10.1029/2010JB007406.Google Scholar
Karato, S.-I. (2008). Deformation of Earth Materials: an Introduction to the Rheology of Solid Earth, Cambridge University Press, New York, 463 pp.Google Scholar
Karcz, Z. and Scholz, C. H. (2003). The fractal geometry of some stylolites from the Calcare Massiccio Formation, Italy. J. Struct. Geol. 25:13011316.Google Scholar
Karner, S. L. (2006). An extension of rate and state theory to poromechanics. Geophys. Res. Lett. 33:L03308, 10.1029/2005GL024934.Google Scholar
Karner, S. L., Marone, C. and Evans, B. (1997). Laboratory study of fault healing and lithification in simulated fault gouge under hydrothermal conditions. Tectonophysics 277:4155.Google Scholar
Karner, S. L., Chester, F. M., Kronenberg, A. K. and Chester, J. S. (2003). Subcritical compaction and yielding of granular quartz sand. Tectonophysics 377:357381.Google Scholar
Karner, S. L., Chester, J. S., Chester, F. M., Kronenberg, A. K. and Hajash, A. Jr. (2005). Laboratory deformation of granular quartz sand: implications for the burial of clastic rocks. Am. Assoc. Petrol. Geol. Bull. 89:603625.Google Scholar
Karson, J. A. and Dick, H. (1983). Tectonics of ridge-transform intersections at the Kane fracture zone. Mar. Geophys. Res. 6:5198.Google Scholar
Karson, K. A. (2002). Geologic structure of the uppermost oceanic crust created at fast- to intermediate-rate spreading centers. Ann. Rev. Earth Planet. Sci. 30:347384.Google Scholar
Kassir, M. K. and Sih, G. C. (1966). Three-dimensional stress distribution around an elliptical crack under arbitrary loadings. J. Appl. Mech. 33:601611.Google Scholar
Kastens, K. A. (1987). A compendium of causes and effects of processes at transform faults and fracture zones. Rev. Geophys. 25:15541562.Google Scholar
Katsman, R. and Aharonov, E. (2006). A study of compaction bands originating from cracks, notches, and compacted defects. J. Struct. Geol. 28:508518.Google Scholar
Katsman, R., Aharanov, E. and Scher, H. (2004). Numerical simulation of compaction bands in high-porosity sedimentary rock. Mech. Mater. 37:371390.Google Scholar
Katsman, R., Aharanov, E. and Scher, H. (2006a). A numerical study on localized volume reduction in elastic media: some insights on the mechanics of anticracks. J. Geophys. Res. 111:B03204; doi:10.1029/2004JB003607.Google Scholar
Katsman, R., Aharanov, E. and Scher, H. (2006b). Localized compaction in rocks: Eshelby’s inclusion and the spring network model. Geophys. Res. Lett. 33:L10311; doi:10.1029/2005GL025628.Google Scholar
Katsube, T. J. and Williamson, M. A. (1998). Shale petrophysical characteristics: permeability history of subsiding shales. In Shales and Mudstones, vol. II (eds. Schieber, J., Zimmerle, W. and Sethi, P.), pp. 6991, E. Schweizerbart Science Publishers, Stuttgart, Germany.Google Scholar
Kattenhorn, S. A. and Marshall, S. T. (2006). Fault-induced stress fields and associated tensile and compressive deformation at fault tips in the ice shell of Europa: implications for fault mechanics. J. Struct. Geol. 28:22042221.Google Scholar
Kattenhorn, S. A. and Pollard, D. D. (1999). Is lithostatic loading important for the slip behavior and evolution of normal faults in the Earth’s crust? J. Geophys. Res. 104:28,879–28,898.Google Scholar
Kattenhorn, S. A., Aydin, A. and Pollard, D. D. (2000). Joints at high angles to normal fault strike: an explanation using 3-D numerical models of fault-perturbed stress fields. J. Struct. Geol. 22:123.Google Scholar
Kattenhorn, S. A. and Watkeys, M. K. (1995). Blunt-ended dyke segments. J. Struct. Geol. 17:15351542.Google Scholar
Katz, Y., Weinberger, R. and Aydin, A. (2004). Geometry and kinematic evolution of Riedel shear structures, Capitol Reef National Park, Utah. J. Struct. Geol. 26:491501.Google Scholar
Katzman, R., ten Brink, U. S. and Lin, J. (1995). Three-dimensional modeling of pull-apart basins: implications for the tectonics of the Dead Sea basin. J. Geophys. Res. 100:62956312.Google Scholar
Kemeny, J. W. (1991). A model for non-linear rock deformation under compression due to sub-critical crack growth. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 28:459467.Google Scholar
Kemeny, J. (2003). The time-dependent reduction of sliding cohesion due to rock bridges along discontinuities: a fracture mechanics approach. Rock Mech. Rock Eng. 36:2738.Google Scholar
Kemeny, J. and Cook, N. G. W. (1986). Effective moduli, non-linear deformation and strength of a cracked elastic solid. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 23:107118.Google Scholar
Kemeny, J. M. and Cook, N. G. W. (1987). Crack models for the failure of rock under compression. Proc. 2nd. Int. Conf. Constitutive Laws for Engng. Materials 2:879887.Google Scholar
Kenkmann, T. (2002). Folding within seconds. Geology 30:231234.Google Scholar
Kenkmann, T. (2003). Dike formation, cataclastic flow, and rock fluidization during impact cratering: an example from the Upheaval Dome structure, Utah. Earth Planet. Sci. Lett. 214:4358.Google Scholar
Keranen, K. M., Savage, H. M., Abers, G. A. and Cochran, E. S. (2013). Potentially induced earthquakes in Oklahoma, USA—links between wastewater injection and the 2011 Mw 5.7 earthquake sequence. Geology 41; doi:10.1130/G34045.1.Google Scholar
Key, W. R. O. and Schultz, R. A. (2011). Fault formation in porous rocks at high strain rates: first results from the Upheaval Dome impact crater, Utah, USA. Geol. Soc. Am. Bull. 123:11611170.Google Scholar
Khabbazi, A., Ghafoori, M., Lashkaripour, G. R. and Cheshomi, A. (2012). Estimation of the rock mass deformation modulus using a rock classification system. Geomech. Geoeng.; doi:10.1080/17486025.2012.695089.Google Scholar
Khan, A. S., Xiang, Y. and Huang, S. (1991). Behavior of Berea Sandstone under confining pressure part 1: Yield and failure surfaces, and nonlinear elastic response. Int. J. Plasticity 7:607624.Google Scholar
Khazan, Y. M. and Fialko, Y. A. (1995). Fracture criteria at the tip of fluid-driven cracks in the earth. Geophys. Res. Lett. 22:25412544.Google Scholar
Kidambi, T. and Kumar, G. S. (2016). Mechanical Earth Modeling for a vertical well drilled in a naturally fractured tight carbonate gas reservoir in the Persian Gulf. J. Petrol. Sci. Eng. 141:3851.Google Scholar
Kies, J. A., Krafft, J. M., Sanford, R. J., Smith, H. L. and Sullivan, A. M. (1975). Historical note on the development of fracture mechanics by G. R. Irwin. In Linear Fracture Mechanics: Historical Developments and Applications of Linear Fracture Mechanics Theory (eds. Sih, G. C., Wei, R. P. and Erdogan, F.), pp. 127, Envo Publishing Company, Lehigh, Pennsylvania.Google Scholar
Kilburn, C. R. J. and Voight, B. (1998). Slow rock fracture as eruption precursor at Soufriere Hills volcano, Montserrat. Geophys. Res. Lett. 25:36653668.Google Scholar
Kim, W.-Y. (2013). Induced seismicity associated with fluid injection into a deep well in Youngstown, Ohio. J. Geophys. Res. 118:35063518.Google Scholar
Kim, Y.-S. and Sanderson, D. J. (2005). The relationship between displacement and length of faults: a review. Earth-Sci. Rev. 68:317334.Google Scholar
Kim, Y.-S., Andrews, J. R. and Sanderson, D. J. (2003). Reactivated strike-slip faults: examples from north Cornwall, UK. Tectonophysics 340:173194.Google Scholar
Kim, Y.-S., Peacock, D. C. P. and Sanderson, D. J. (2004). Fault damage zones. J. Struct. Geol. 26:503517.Google Scholar
King, G. C. P. (1978). Geological faulting: fracture, creep and strain. Phil. Trans. Royal Soc. London A, 288:197212.Google Scholar
King, G. C. P. (1983). The accommodation of large strains in the upper lithosphere of the Earth and other solids by self-similar fault systems—the geometrical origin of b-value. Pure Appl. Geophys. 121:762815.Google Scholar
King, G. and Ellis, M. (1990). The origin of large local uplift in extensional regions. Nature 348:689693.Google Scholar
King, G. C. P. and Nábêlek, J. L. (1985). The role of fault bends in the initiation and termination of earthquake rupture. Science 228:984987.Google Scholar
King, G. and Yielding, G. (1984). The evolution of a thrust fault system: processes of rupture initiation, propagation and termination in the 1980 El Asnam (Algeria) earthquake. Geophys. J. Royal Astron. Soc. 77:915933.Google Scholar
King, G. C. P., Stein, R. S. and Lin, J. (1994). Static stress changes and the triggering of earthquakes. Bull. Seismol. Soc. Am. 84:935953.Google Scholar
Kingma, J. T. (1958). Possible origin of piercement structures, local unconformities, and secondary basins in the Eastern Geosyncline, New Zealand. N. Z. J. Geol. Geophys. 1:269274.Google Scholar
Kirby, S. H. and Kronenberg, A. K. (1987). Rheology of the lithosphere: selected topics. Rev. Geophys. 25:12191244.Google Scholar
Kirby, S. H., Durham, W. B. and Stern, L. (1991). Mantle phase changes and deep earthquake faulting in subducting lithosphere. Science 252:216225.Google Scholar
Kirby, S. H., Durham, W. B. and Stern, L. (1992). The ice I–II transformation: mechanisms and kinetics under hydrostatic and nonhydrostatic conditions. In Physics and Chemistry of Ice (eds. Maeno, N. and Hondoh, T.), pp. 456463, Hokkaido University Press, Sapporo, Japan.Google Scholar
Klimczak, C. (2014). Geomorphology of lunar grabens requires igneous dikes at depth. Geology 42:963966.Google Scholar
Klimczak, C. and Schultz, R. A. (2013a). Shear-enhanced compaction in dilating granular materials. Int. J. Rock Mech. Min. Sci. 64:139147.Google Scholar
Klimczak, C. and Schultz, R. A. (2013b). Fault damage zone origin of the Teufelsmauer, Subhercynian Cretaceous Basin, Germany. Int. J. Earth Sci. (Geologische Rundschau) 102:121138; doi:10.1007/s00531-012-0794-z.Google Scholar
Klimczak, C., Schultz, R. A. Parashar, R. and Reeves, D. M. (2010). Cubic law with aperture-length correlation: implications for network scale fluid flow. Hydrogeol. J. 18:851862; doi:10.1007/s10040–009–0572–6.Google Scholar
Klimczak, C., Soliva, R., Schultz, R. A. and Chéry, J. (2011). Growth of deformation bands in a multilayer sequence. J. Geophys. Res. 116, B09209; doi:10.1029/2011JB008365.Google Scholar
Klimczak, C., Ernst, C. M., Byrne, P. K., et al. (2013). Insights into the subsurface structure of the Caloris basin, Mercury, from assessments of mechanical layering and changes in long-wavelength topography. J. Geophys. Res. 118:20302044; doi:10.1002/JGRE.20157.Google Scholar
Klimczak, C., Byrne, P. K. and Solomon, S. C. (2015). A rock-mechanical assessment of Mercury’s global tectonic fabric. Earth Planet. Sci. Lett. 416:8290.Google Scholar
Knapmeyer, M., Oberst, J., Hauber, E., et al. (2006). Working models for spatial distribution and level of Mars’ seismicity. J. Geophys. Res. 111:E11006, doi:10.1029/2006JE002708.Google Scholar
Knipe, R. J. (1989). Deformation mechanisms—recognition from natural tectonites. J. Struct. Geol. 11:127146.Google Scholar
Knott, S. D. (1993). Fault seal analysis in the North Sea. Am. Assoc. Petrol. Geol. Bull. 77:778792.Google Scholar
Knott, S. D., Beach, A., Brockbank, P. J., et al. (1996). Spatial and mechanical controls on normal fault populations. J. Struct. Geol. 18:359372.Google Scholar
Ko, T. Y. and Kemeny, J. (2011). Subcritical crack growth in rocks under shear loading. J. Geophys. Res. 116: B01407; doi:10.1029/2010JB000846.Google Scholar
Koehn, D., Renard, F., Toussaint, R. and Passchier, C. W. (2007). Growth of stylolite teeth patterns depending on normal stress and finite compaction. Earth Planet. Sci. Lett. 257:582595.Google Scholar
Koenig, E. and Aydin, A. (1998). Evidence for large-scale strike-slip faulting on Venus. Geology 26:551554.Google Scholar
Kohlstedt, D. L. and Mackwell, S. J. (2010). Strength and deformation of planetary lithospheres. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 397456, Cambridge University Press.Google Scholar
Kohlstedt, D. L., Evans, B. and Mackwell, S. J. (1995). Strength of the lithosphere: constraints imposed by laboratory experiments. J. Geophys. Res. 100:17,58717,602.Google Scholar
Kolosov, G. V. (1935). Application of a complex variable to the theory of elasticity. Objed. Nauchno-tekhn. Izd., Moscow-Leningrad (in Russian).Google Scholar
Kolyukhin, D. and Torabi, A. (2012). Statistical analysis of the relationships between faults attributes. J. Geophys. Res. 117, B05406; doi:10.1029/2011JB008880.Google Scholar
Kostrov, B. (1974). Seismic moment and energy of earthquakes, and seismic flow of rock. Izvestiya, Phys. Solid Earth 13:1321.Google Scholar
Kramer, E. J. (1974). The stress–strain curve of shear-banding polystyrene. J. Macromol. Sci. B10:191202.Google Scholar
Krantz, R. L. (1983). Microcracks in rocks: a review. Tectonophysics 100:449480.Google Scholar
Krantz, R. W. (1988). Multiple fault sets and three-dimensional strain: theory and application. J. Struct. Geol. 10:225237.Google Scholar
Krantz, R. W. (1989). Orthorhombic fault patterns: the odd axis model and slip vector orientations. Tectonics 8:483495.Google Scholar
Krantz, R. W. (1995). The transpressional strain model applied to strike-slip, oblique-convergent and oblique-divergent deformation. J. Struct. Geol. 17:11251137.Google Scholar
Kreemer, C., Blewett, G. and Klein, E. C. (2014). A geodetic plate motion and global strain rate model. Geochem. Geophys. Geosyst. 15:38493889.Google Scholar
Kulander, B. R. and Dean, S. L. (1985). Hackle plume geometry and joint propagation dynamics. In Proceedings of the International Symposium on Fundamentals of Rock Joints (ed. Stephansson, O.), pp. 8594, Centek.Google Scholar
Kulander, B. R. and Dean, S. L. (1995). Observations on fractography with laboratory experiments for geologists. In Fractography: Fracture Topography as a Tool in Fracture Mechanics and Stress Analysis (ed. Ameen, M. S.), pp. 97–147, Geol. Soc. Spec. Publ. 92.Google Scholar
Kulander, B. R., Barton, C. C. and Dean, S. L. (1979). The application of fractography to core and outcrop fracture investigations. US Department of Energy, METC/SP–79/3, National Technical Information Service, US Department of Commerce, Springfield, VA.Google Scholar
Kulhawy, F. H. (1975). Stress deformation properties of rock and rock discontinuities. Eng. Geol. 9:327350.Google Scholar
Labuz, J. F., Zeng, F., Makhnenko, R. and Li, Y. (2018). Brittle failure of rock: a review and general linear criterion. J. Struct. Geol. 112:728.Google Scholar
Lacazette, A. and Engelder, T. (1992). Fluid-driven cyclic propagation of a joint in the Ithaca Siltstone, Appalachian Basin, New York. In Fault Mechanics and Transport Properties of Rocks (eds. Evans, B. and Wong, T.-f.), pp. 297324, Academic Press, New York.Google Scholar
Lachenbruch, A. H. (1961). Depth and spacing of tension cracks. J. Geophys. Res. 66:42734292.Google Scholar
Lahee, F. H. (1961). Field Geology (6th edn), McGraw-Hill, New York.Google Scholar
Lajtai, E. Z. (1969). Mechanics of second order faults and tension gashes. Geol. Soc. Am. Bull. 80:22532272.Google Scholar
Lajtai, E. Z. (1974). Brittle fracture in compression. Int. J. Fracture 10:525536.Google Scholar
Lajtai, E. Z. (1991). Time dependent behaviour of the rock mass. Geotech. Geol. Eng. 9:109124.Google Scholar
Lamarche, J., Chabani, A. and Gauthier, B. D. M. (2018). Dimensional threshold for fracture linkage and hooking. J. Struct. Geol. 108:171179.Google Scholar
Lander, R. H. and Laubach, S. E. (2015). Insight into rates of fracture growth and sealing from a model for quartz cementation in fractured sandstones. Geol. Soc. Am. Bull. 127:516538.Google Scholar
Landes, J. D. (2000). The contributions of George Irwin to elastic-plastic fracture mechanics development. In Fatigue and Fracture Mechanics: 31st Volume (eds. Halford, G. R. and Gallagher, J. P.), pp. 5463, Am. Soc. Test. Mat. Spec. Publ. 1389, West Conshohocken, Penn.Google Scholar
Laney, R. T. and Van Schmus, W. R. (1978). A structural study of the Kentland, Indiana, impact site. In Proc. 9th Lunar Planet. Sci. Conf. pp. 26092632.Google Scholar
Langford, J. C. and Diederichs, M. S. (2015). Quantifying uncertainty in Hoek–Brown intact strength envelopes. Int. J. Rock Mech. Min. Sci. 74:91102.Google Scholar
Larsen, P. H. (1988). Relay structures in a Lower Permian basement-involved extension system, East Greenland. J. Struct. Geol. 10:38.Google Scholar
Latzko, D. G. H. (editor) (1979). Post-Yield Fracture Mechanics, Elsevier Science Ltd, 364 pp.Google Scholar
Laubach, S. E., Reed, R. M., Olson, J. E., Lander, R. H. and Bonnell, L. M. (2004). Coevolution of crack-seal texture and fracture porosity in sedimentary rocks: cathodoluminescence observations of regional fractures. J. Struct. Geol. 26:967982.Google Scholar
Laubach, S. E., Olson, J. E. and Gross, M. R. (2009). Mechanical and fracture stratigraphy. Am. Assoc. Petrol. Geol. Bull. 93:14131426.Google Scholar
Laubach, S. E., Eichhubl, P., Hilgers, C. and Lander, R. H. (2010). Structural diagenesis. J. Struct. Geol. 32:18661872.Google Scholar
Lawn, B. (1993). Fracture of Brittle Solids (2nd edn), Cambridge University Press, 378 pp.Google Scholar
Leblond, J.-B., Karma, A. and Lazarus, V. (2011). Theoretical analysis of crack front instability in mode I+III. J. Mech. Phys. Solids 59:18721887.Google Scholar
Lecampion, B., Bunger, A. and Zhang, X. (2018). Numerical methods for hydraulic fracture propagation: a review of recent trends. Journal of Natural Gas Science and Engineering 49:6683; doi:10.1016/j.jngse.2017.10.012.Google Scholar
Lee, H. P. and Olson, J. E. (2017). The effect of remote and internal crack stresses on mixed-mode brittle fracture propagation of open cracks under compressive loading. Int. J. Fracture 207; doi:10.1007/s10704-017-0231-1.Google Scholar
Lemaitre, J. and Desmorat, R. (2005). Engineering Damage Mechanics: Ductile, Creep, Fatigue and Brittle Failures, Springer, Berlin.Google Scholar
Lensen, G. J. (1958). A method of horst and graben formation. J. Geol. 66:579587.Google Scholar
Li, A. J., Merifield, R. S. and Lyamin, A. V. (2008). Stability charts for rock slopes based on the Hoek–Brown failure criterion. Int. J. Rock Mech. Min. Sci. 45:689700.Google Scholar
Li, V. C. (1987). Mechanics of shear rupture applied to earthquake zones. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 351428, Academic Press, New York.Google Scholar
Li, V. C. and Liang, E. (1986). Fracture processes in concrete and fiber reinforced cementitious composites. J. Eng. Mech. 112:566586.Google Scholar
Lin, B., Mear, M. E. and Ravi-Chandar, K. (2010). Criterion for initiation of cracks under mixed-mode I + III loading. Int. J. Fracture 165:175188.Google Scholar
Lin, J. and Parmentier, E. M. (1988). Quasistatic propagation of a normal fault: a fracture mechanics model. J. Struct. Geol. 10:249262.Google Scholar
Lin, J. and Stein, R. S. (1989). Coseismic folding, earthquake recurrence, and the 1987 source mechanism at Whittier Narrows, Los Angeles basin, California. J. Geophys. Res. 94:96149632.Google Scholar
Lin, M., Hardy, M. P., Agapito, J. F. T. et al. (1993). Rock mass mechanical property estimations for the Yucca Mountain site characterization project. Sandia Nat. Lab. Rept. SAND92–0450, Albquerque, New Mexico.Google Scholar
Linker, M. F. and Dieterich, J. H. (1992). Effects of variable normal stress on rock friction: observations and constitutive equations. J. Geophys. Res. 97:49234940.Google Scholar
Lisle, R. J. and Leyshon, P. R. (2004). Stereographic Projection Techniques for Geologists and Civil Engineers (2nd edn), Cambridge University Press, Cambridge.Google Scholar
Liu, C., Pollard, D. D., Deng, S. and Aydin, A. (2016). Mechanism of formation of wiggly compaction bands in porous sandstone: 1. observations and conceptual model. J. Geophys. Res. 120:81388152; doi:10.1002/2015JB012372.Google Scholar
Lockner, D. (1993). The role of acoustic emission in the study of rock fracture. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 30:883899.Google Scholar
Lockner, D. A. (1995). Rock failure. In Rock Physics and Phase Relations—A Handbook of Physical Constants (ed. Ahrens, T. J.), pp. 127147, American Geophysical Union Reference Shelf 3.Google Scholar
Lockner, D. A. (1998). A generalized law for brittle deformation of Westerly granite. J. Geophys. Res. 103:51075123.Google Scholar
Lockner, D. A. and Beeler, N. M. (1999). Premonitory slip and tidal triggering of earthquakes. J. Geophys. Res. 104:20,13320,151.Google Scholar
Lockner, D. A., Byerlee, J. D., Kuksenko, V., Ponomarev, A. and Sidorin, A. (1991). Quasi-static fault growth and shear fracture energy in granite. Nature 350:3942.Google Scholar
Loizzo, M., Lecampion, B. and Mogilevskaya, S. (2017). The role of geological barriers in achieving robust well integrity. Energy Procedia 114:51935205.Google Scholar
Lonergan, L. and Cartwright, J. A. (1999). Polygonal faults and their influence on deep-water sandstone reservoir geometries, Alba field, United Kingdom central North Sea. Am. Assoc. Petrol. Geol. Bull. 83:410432.Google Scholar
Long, J. J. and Imber, J. (2011). Geological controls on fault relay zone scaling. J. Struct. Geol. 33:17901800.Google Scholar
Lorenz, J. C. and Cooper, S. P. (2017) Deformation bands. In Atlas of Natural and Induced Fractures in Core, John Wiley & Sons, Ltd, Chichester, UK; doi:10.1002/9781119160014.ch11.Google Scholar
Lorenz, J. C., Sterling, J. L., Schechter, D. S., Whigham, C. L. and Jensen, J. J. (2002). Natural fractures in the Sprayberry Formation, Midland Basin, Texas: the effects of mechanical stratigraphy on fracture variability and reservoir behavior. Am. Assoc. Petrol. Geol. Bull., 86:505524.Google Scholar
Lucchitta, B. K. (1976). Mare ridges and related highland scarps—results of vertical tectonism? Proc. Lunar Sci. Conf. 7:27612782.Google Scholar
Lucchitta, B. K. (1977). Topography, structure, and mare ridges in southern Mare Imbrium and northern Oceanus Procellarum. Proc. Lunar Sci. Conf. 8:26912703.Google Scholar
Luyendyk, B. P., Kamerling, M. J. and Terres, R. (1980). Geometric model for Neogene crustal rotations in southern California. Geol. Soc. Am. Bull. 91:211217.Google Scholar
Lyakhovsky, V., Zhu, W. and Shalev, E. (2015). Visco-poroelastic damage model for brittle-ductile failure of porous rocks. J. Geophys. Res. 120; doi:10.1002/2014JB011805.Google Scholar
Lyzenga, G. A., Wallace, K. S., Faneslow, J. L., Raefsky, A. and Groth, P. M. (1986). Tectonic motions in California inferred from VLBI observations, 1980–1984. J. Geophys. Res. 91:94739487.Google Scholar
Ma, X. Q. and Kusznir, N. J. (1993). Modelling of near-field subsurface displacements for generalized faults and fault arrays. J. Struct. Geol. 15:14711484.Google Scholar
Macdonald, G. A. (1957). Faults and monoclines on Kilauea Volcano, Hawaii. Geol. Soc. Am. Bull. 68:269271.Google Scholar
Macdonald, K. (1982). Mid-ocean ridges: fine scale tectonic, volcanic, and hydrothermal processes within the plate boundary zone. Annu. Rev. Earth Planet. Sci. 10:155190.Google Scholar
Macdonald, K. C., Fox, P. J., Perram, L. J., et al. (1988). A new view of the mid-ocean ridge from the behavior of ridge-axis discontinuities. Nature 335:217225.Google Scholar
Mack, G. H. and Seager, W. R. (1995). Transfer zones in the southern Rio Grande rift. J. Geol. Soc. London 152:551560.Google Scholar
Maerten, L., Willemse, E. J. M., Pollard, D. D. and Rawnsley, K. (1999). Slip distributions on intersecting normal faults. J. Struct. Geol. 21:259272.Google Scholar
Main, I. G. (1991). A modified Griffith criterion for the evolution of damage with a fractal distribution of crack lengths: application to seismic event rates and b-values. Geophys. J. Int. 107:353362.Google Scholar
Main, I. G. (1995). Earthquakes as critical phenomena: implications for probabilistic seismic hazard analysis. Bull. Seismol. Soc. Am. 85:12991308.Google Scholar
Main, I. G. (2000). A damage mechanics model for power-law creep and earthquake aftershock and foreshock sequences. Geophys. J. Int. 142:151161.Google Scholar
Main, I. G. and Meredith, P. G. (1991). Stress corrosion constitutive laws as a possible mechanism of intermediate-term and short-term seismic quiescence. Geophys. J. Int. 107:363372.Google Scholar
Main, I. G., Sammonds, P. R. and Meredith, P. G. (1993). Application of a modified Griffith criterion to the evolution of fractal damage during compressional rock failure. Geophys. J. Int. 115:367380.Google Scholar
Main, I. G., Leonard, T., Papasouliotis, O., Hatton, C. G. and Meredith, P. G. (1999). One slope or two? Detecting statistically significant breaks of slope in geophysical data, with applications to fracture scaling relationships. Geophys. Res. Lett. 26:28012804.Google Scholar
Mair, K., Main, I. and Elphick, S. (2000). Sequential growth of deformation bands in the laboratory. J. Struct. Geol. 22:2542.Google Scholar
Mair, K., Frye, K. M. and Marone, C. (2002). Influence of grain characteristics on the friction of granular shear zones. J. Geophys. Res. 107; doi:10.1029/2001JB000516, 2219.Google Scholar
Maiti, S. K. and Smith, T. A. (1983). Comparison of the criteria for mixed mode brittle failure based on the preinstability stress-strain field—part I: slit and elliptical cracks under uniaxial tensile loading. Int. J. Fracture 23:281295.Google Scholar
Maiti, S. K. and Smith, T. A. (1984). Comparison of the criteria for mixed mode brittle failure based on the preinstability stress-strain field—part II: pure shear and uniaxial compressive loading. Int. J. Fracture 24:522.Google Scholar
Majer, E. L., Baria, R., Stark, M., et al. (2007). Induced seismicity associated with Enhanced Geothermal Systems. Geothermics 36:185222.Google Scholar
Malik, J. N., Shah, A. A., Sahoo, A. K., et al. (2010). Active fault, fault growth and segment linkage along the Janauri anticline (frontal foreland fold), NW Himalaya, India. Tectonophysics 483:327343.Google Scholar
Maltman, A. J. (1984). On the term ‘soft-sediment deformation.’ J. Struct. Geol. 6:589592.Google Scholar
Maltman, A. J. (1988). The importance of shear zones in naturally deformed wet sediments. Tectonophysics 145:163175.Google Scholar
Maltman, A. J. (1994). Prelithification deformation. In Continental Deformation (ed. Hancock, P. L.), pp. 143158, Pergamon, New York.Google Scholar
Mandl, G. (1987a). Discontinuous fault zones. J. Struct. Geol. 9:105110.Google Scholar
Mandl, G. (1987b). Tectonic deformation by rotating parallel faults: the “bookshelf” mechanism. Tectonophysics 141:277316.Google Scholar
Mandl, G. (1988). Mechanics of Tectonic Faulting—Models and Basic Concepts, Elsevier, Amsterdam.Google Scholar
Mandl, G. (2000). Faulting in Brittle Rocks: an Introduction to the Mechanics of Tectonic Faults, Springer-Verlag, Heidelberg.Google Scholar
Mandl, G. (2005). Rock Joints: the Mechanical Genesis, Springer-Verlag, Heidelberg.Google Scholar
Manighetti, I., King, G. C. P., Gaudemer, Y., Scholz, C. H. and Doubre, C. (2001). Slip accumulation and lateral propagation of active normal faults in Afar. J. Geophys. Res. 106:13,66713,696.Google Scholar
Manighetti, I., King, G. and Sammis, C. G. (2004). The role of off-fault damage in the evolution of normal faults. Earth Planet. Sci. Lett. 217:399408.Google Scholar
Mangold, N., Allemand, P. and Thomas, P. G. (1998). Wrinkle ridges of Mars: structural analysis and evidence for shallow deformation controlled by ice-rich décollements. Planet. Space Sci. 46:345356.Google Scholar
Mann, P., Hempton, M. R., Bradley, D. C. and Burke, K. C. (1983). Development of pull-apart basins. J. Geol. 91:529554.Google Scholar
Mann, P., Draper, G. and Burke, K. (1985). Neotectonics of a strike-slip restraining bend system, Jamaica. In Strike-Slip Deformation, Basin Formation, and Sedimentation (eds. Biddle, K. T. and Christie-Blick, N.), pp. 211226, Soc. Econ. Paleon. Miner., Spec. Publ. 37.Google Scholar
Mansinha, L. and Smylie, D. E. (1971). The displacement fields of inclined faults. Bull. Seismol. Soc. Am. 61:14331440.Google Scholar
Marder, M. and Fineberg, J. (1996). How things break. Phys. Today 49:2429.Google Scholar
Marinos, P. and Hoek, E. (2001). Estimating the geotechnical properties of heterogeneous rock masses such as flysch. Bull. Eng. Geol. Env. 60:8592.Google Scholar
Marinos, V., Marinos, P. and Hoek, E. (2005). The geologic strength index: applications and limitations. Bull. Eng. Geol. Env. 64:5565.Google Scholar
Marketos, G. and Bolton, M. D. (2009). Compaction bands simulated in discrete element models. J. Struct. Geol. 31:479490.Google Scholar
Marone, C. (1995). Fault zone strength and failure criteria. Geophys. Res. Lett. 22:723726.Google Scholar
Marone, C. (1998a). The effect of loading rate on static friction and the rate of fault healing during the earthquake cycle. Nature 391:6972.Google Scholar
Marone, C. (1998b). Laboratory-derived friction laws and their application to seismic faulting. Annu. Rev. Earth Planet. Sci. 26:643696.Google Scholar
Marone, C. and Scholz, C. H. (1988). The depth of seismic faulting and the transition from stable to unstable slip regimes. Geophys. Res. Lett. 15:621624.Google Scholar
Marone, C., Scholz, C. H. and Bilham, R. (1991). On the mechanics of earthquake afterslip. J. Geophys. Res. 96:84418452.Google Scholar
Marrett, R. (1996). Aggregate properties of fracture populations. J. Struct. Geol. 18:169178.Google Scholar
Marrett, R. and Allmendinger, R. W. (1990). Kinematic analysis of fault slip data. J. Struct. Geol. 12:973986.Google Scholar
Marrett, R. and Peacock, D. C. P. (1999). Strain and stress. J. Struct. Geol. 21:10571063.Google Scholar
Marshak, S. and Mitra, G. (eds.) (1988). Basic Methods of Structural Geology, Prentice-Hall, New York.Google Scholar
Martel, S. J. (1990). Formation of compound strike-slip fault zones, Mount Abbot quadrangle, California. J. Struct. Geol. 12:869882.Google Scholar
Martel, S. J. (1997). Effects of cohesive zones on small faults and implications for secondary fracturing and fault trace geometry. J. Struct. Geol. 19:835847.Google Scholar
Martel, S. J. (2004). Mechanics of landslide initiation as a shear fracture phenomenon. Marine Geol. 203:319339.Google Scholar
Martel, S. J. (2017). Progress in understanding sheeting joints over the past two centuries. J. Struct. Geol. 94:6886.Google Scholar
Martel, S. J. and Boger, W. A. (1998). Geometry and mechanics of secondary fracturing around small three-dimensional faults in granitic rock. J. Geophys. Res. 103:21,299–21,314.Google Scholar
Martel, S. J. and Pollard, D. D. (1989). Mechanics of slip and fracture along small faults and simple strike-slip fault zones in granitic rock. J. Geophys. Res. 94:94179428.Google Scholar
Martel, S. J., Pollard, D. D. and Segall, P. (1988). Development of simple strike-slip fault zones, Mount Abbot quadrangle, Sierra Nevada, California. Geol. Soc. Am. Bull. 100:14511465.Google Scholar
Mastin, L. G. and Pollard, D. D. (1988). Surface deformation and shallow dike intrusion processes at Inyo Craters, Long Valley, California. J. Geophys. Res. 93:13,22113,235.Google Scholar
Masuda, K. (2001). Effects of water on rock strength in a brittle regime. J. Struct. Geol. 23:16531657.Google Scholar
Matthäi, S., Aydin, A., Pollard, D. D. and Roberts, S. (1998). Numerical simulation of departures from radial drawdown in a faulted sandstone reservoir with joints and deformation bands, pp. 157191, Geol. Soc. London Spec. Publ. 147.Google Scholar
Mazars, J. and Pijaudier-Cabot, G. (1996). From damage to fracture mechanics and conversely: a combined approach. Int. J. Solids Structures 33:33273342.Google Scholar
Mauldon, M. and Dershowitz, W. (2000). A multi-dimensional system of fracture abundance measures (abstract). Paper presented at the annual meeting, Geol. Soc. Amer., Reno, Nevada, November 2000.Google Scholar
McCaffrey, R. (1992). Oblique plate convergence, slip vectors, and forearc deformation. J. Geophys. Res. 97:89058915.Google Scholar
McClay, K. R. (1992). Glossary of thrust tectonics terms. In Thrust Tectonics (ed. McClay, K. R.), pp. 419433, Chapman and Hall, London.Google Scholar
McClay, K. R. (2011). Introduction to thrust fault-related folding. In Thrust Fault-Related Folding (eds. McClay, K. R., Shaw, J. H. and Suppe, J.), pp. 119, Amer. Assoc. Petrol. Geol. Mem. 94.Google Scholar
McClay, K. R., Dooley, T., Whitehouse, P. S. and Anadon-Ruiz, S. (2005). 4D analogue models of extensional fault systems in asymmetric rifts: 3D visualizations and comparisons with natural examples. In Petroleum Geology: North-West Europe and Global PerspectivesProc. 6th Petrol. Geol. Conf. (eds. Doré, A. G. and Vining, B. A.), pp. 15431556, Geological Society, London.Google Scholar
McClintock, F. A. and Walsh, J. B. (1962). Friction on Griffith cracks in rocks under pressure. Proc., 4th US National Congress of Appl. Mech., pp. 10151021.Google Scholar
McConaughy, D. T. and Engelder, T. (2001). Joint initiation in bedded clastic rocks. J. Struct. Geol. 23:203221.Google Scholar
McCoss, A. M. (1986). Simple constructions for deformation in transpression/transtension zones. J. Struct. Geol. 8:715718.Google Scholar
McGarr, A. (1988). On the state of lithospheric stress in the absence of applied tectonic forces. J. Geophys. Res. 93:13,60913,617.Google Scholar
McGarr, A. (2014). Maximum magnitude earthquakes induced by fluid injection. J. Geophys. Res. 119:10081019.Google Scholar
McGarr, A. and Gay, N. C. (1978). State of stress in the Earth’s crust. Ann. Rev. Earth Planet. Sci. 6:405436.Google Scholar
McGarr, A., Pollard, D. D., Gay, N. C. and Ortlepp, W. D. (1979). Observations and analysis of structures in exhumed mine-induced faults. US Geological Survey Open-File Report, 79(1239), pp.101120.Google Scholar
McGarr, A., Simpson, D. and Seeber, L. (2002). Case histories of induced and triggered seismicity. In International Handbook of Earthquake and Engineering Seismology, vol. 81A, pp. 647661, Academic Press, San Francisco.Google Scholar
McGill, G. E. (1993). Wrinkle ridges, stress domains and kinematics of Venusian plains. Geophys. Res. Lett. 20:24072410.Google Scholar
McGill, G. E. and Stromquist, A. W. (1979). The grabens of Canyonlands National Park, Utah: geometry, mechanics, and kinematics. J. Geophys. Res. 84:45474563.Google Scholar
McGill, G. E., Stofan, E. R. and Smrekar, S. E. (2010). Venus tectonics. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 81120, Cambridge University Press.Google Scholar
McGovern, P. J. and Solomon, S. C. (1993). State of stress, faulting, and eruption characteristics of large volcanoes on Mars. J. Geophys. Res. 98:23,55323,579.Google Scholar
McGrath, A. G. and Davison, I. (1995). Damage zone geometry around fault tips. J. Struct. Geol. 17:10111024.Google Scholar
McKinstry, H. E. (1948). Mining Geology, Prentice-Hall, New York, 680 pp.Google Scholar
McLeod, A. E., Dawers, N. H. and Underhill, J. R. (2000). The propagation and linkage of normal faults: insights from the Strathsprey-Brent-Statfjord fault array, northern North Sea. Basin Res. 12:263284.Google Scholar
Means, W. D. (1976). Stress and Strain—Basic Concepts of Continuum Mechanics for Geologists, Springer-Verlag, New York.Google Scholar
Medwedeff, D. A. (1992). Geometry and kinematics of an active, laterally propagating wedge thrust, Wheeler Ridge, California. In Structural Geology of Fold and Thrust Belts (ed. Mitra, S. and Fisher, G. W.), pp. 328, Johns Hopkins University Press, Baltimore, Maryland.Google Scholar
Melin, S. (1982). Why do cracks avoid each other? Int. J. Fracture 23:3745.Google Scholar
Melin, S. (1986). When does a crack grow under mode II conditions? Int. J. Fracture 30:103114.Google Scholar
Melosh, H. J. (1989). Impact Cratering: A Geologic Process, Oxford University Press, New York.Google Scholar
Melosh, H. J. and Williams, C. A. Jr. (1989). Mechanics of graben formation in crustal rocks: a finite element analysis. J. Geophys. Res. 94:13,96113,973.Google Scholar
Menéndez, B., Zhu, W. and Wong, T.-f. (1996). Micromechanics of brittle faulting and cataclastic flow in Berea sandstone. J. Struct. Geol. 18:116.Google Scholar
Meng, C. and Pollard, D. D. (2014). Eshelby’s solution for ellipsoidal inhomogeneous inclusions with applications to compaction bands. J. Struct. Geol. 67:119.Google Scholar
Meng, L., Fu, X., Lv, Y., et al. (2016). Risking fault reactivation induced by gas injection into depleted reservoirs based on the heterogeneity of geomechanical properties of fault zones. Petrol. Geosci. 23:2938.Google Scholar
Mercier, E., Outtani, F. and Frizon de Lamotte, D. (1997). Late-stage evolution of fault-propagation folds: principles and example. J. Struct. Geol. 19:185193.Google Scholar
Meredith, P. G. and Atkinson, B. K. (1983). Stress corrosion and acoustic emission during tensile crack propagation in Whin Sill dolerite and other basic rocks. Geophys. J. R. Astron. Soc. 75:121.Google Scholar
Meredith, P. G., Main, I. G. and Jones, C. (1990). Temporal variations in seismicity during quasi-static and dynamic rock failure. Tectonophysics 175:249268.Google Scholar
Mériaux, C., Lister, J. R., Lyakhovsky, V. and Agnon, A. (1999). Dyke propagation with distributed damage of the host rock. Earth Planet. Sci. Lett. 165:177185.Google Scholar
Michalske, T. A., Singh, M. and Fréchette, V. D. (1981). Experimental observation of crack velocity and crack front shape effects in double-torsion fracture mechanics tests. In Fracture Mechanics for Ceramics, Rocks, and Concrete (eds. Freiman, S. W. and Fuller, E. R. Jr.), pp. 312, Am. Soc. Test. Mat. Spec. Publ. 745, Philadelphia, Penn.Google Scholar
Micklethwaite, S., Ford, A., Witt, W. and Sheldon, H. A. (2014). The where and how of faults, fluids and permeability: insights from fault stepovers, scaling properties and gold mineralisation. Geofluids; doi 10.1111/gfl.12102.Google Scholar
Mitra, G. (1979). Ductile deformation zones in Blue Ridge basement rocks and estimation of finite strain. Geol. Soc. Am. Bull. 90:935951.Google Scholar
Mitra, G. and Sussman, A. J. (1997). Structural evolution of connecting splay duplexes and their implications for critical taper: an example based on geometry and kinematics of the Canyon Range culmination, Sevier Belt, central Utah. J. Struct. Geol. 19:503521.Google Scholar
Mitra, S. (1986). Duplex structures and imbricate thrust systems: geometry, structural position, and hydrocarbon potential. Am. Assoc. Petrol. Geol. Bull. 70:10871112.Google Scholar
Mitra, S. (1990). Fault-propagation folds: geometry, kinematic evolution, and hydrocarbon traps. Am. Assoc. Petrol. Geol. Bull. 74:921945.Google Scholar
Mitra, S. and Mount, V. S. (1998). Foreland basement-involved structures. Am. Assoc. Petrol. Geol. Bull. 82:70109. (Errata, Am. Assoc. Petrol. Geol. Bull. 83:2024–2027, 1999.)Google Scholar
Mitra, S. and Mount, V. S. (1999). Foreland basement-involved structures: Reply. Am. Assoc. Petrol. Geol. Bull. 83:20172023.Google Scholar
Moeck, I. and Backers, T. (2011). Fault reactivation potential as a critical factor during reservoir stimulation. First Break 29:7380.Google Scholar
Moeck, I., Kwiatek, G. and Zimmermann, G. (2009). Slip tendency analysis, fault reactivation potential and induced seismicity in a deep geothermal reservoir. J. Struct. Geol. 31:11741182.Google Scholar
Mollema, P. N. and Antonellini, M. A. (1996). Compaction bands: a structural analog for anti-mode I cracks in aeolian sandstone. Tectonophysics 267:209228.Google Scholar
Mollema, P. N. and Antonellini, M. A. (1999). Development of strike-slip faults in the dolomites of the Sella Group, northern Italy. J. Struct. Geol. 21:273292.Google Scholar
Molnar, P. (1983). Average regional strain due to slip on numerous faults of different orientations. J. Geophys. Res. 88:64306432.Google Scholar
Molnar, P. (1988). Continental tectonics in the aftermath of plate tectonics. Nature 335:131137.Google Scholar
Molnar, P. and Tapponnier, P. (1975). Cenozoic tectonics of Asia: effects of a continental collision. Science 189:419426.Google Scholar
Moody, J. D. and Hill, M. L. (1956). Wrench-fault tectonics. Geol. Soc. Am. Bull. 67:12071246.Google Scholar
Moore, D. E. and Lockner, D. A. (1995). The role of microcracking in shear-fracture propagation in granite. J. Struct. Geol. 17:95114.Google Scholar
Moore, J. M. and Schultz, R. A. (1999). Processes of faulting in jointed rocks of Canyonlands National Park, Utah. Geol. Soc. Am. Bull. 111:808822.Google Scholar
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. 95:93059325.Google Scholar
Morewood, N. C. and Roberts, G. P. (2002). Surface observations of active normal fault propagation: implications for growth. J. Geol. Soc. London 159:263272.Google Scholar
Morgan, J. K. and Boettcher, M. S. (1999). Numerical simulations of granular shear zones using the distinct element method, 1. Shear zone kinematics and the micromechanics of localization. J. Geophys. Res. 104:27032719.Google Scholar
Morgan, J. K. and McGovern, P. J. (2005). Discrete element simulations of gravitational volcanic deformation: 2. Mechanical analysis. J. Geophys. Res. 110:B05403; doi:10.1029/2004JB003253.Google Scholar
Morgan, W. J. (1968). Rises, trenches, great faults, and crustal blocks. J. Geophys. Res. 73:19591982.Google Scholar
Morley, C. K., Nelson, R. A., Patton, T. L. and Munn, S. G. (1990). Transfer zones in the East African rift system and their relevance to hydrocarbon exploration in rifts. Am. Assoc. Petrol. Geol. Bull. 74:12341253.Google Scholar
Morley, C. K. (1999). How successful are analogue models in addressing the influence of pre-existing fabrics on rift structure? J. Struct. Geol. 21:12671274.Google Scholar
Moros, J. G. (1999). Relationship between fracture aperture and length in sedimentary rocks. Unpublished MS thesis, The University of Texas, Austin.Google Scholar
Moustafa, A. R. (1997). Controls on the development and evolution of transfer zones: the influence of basement structure and sedimentary thickness in the Suez rift and Red Sea. J. Struct. Geol. 19:755768.Google Scholar
Mount, V. S. and Suppe, J. (1987). State of stress near the San Andreas fault: implications for wrench tectonics. Geology 15:11431146.Google Scholar
Mount, V. S., Suppe, J. and Hook, S. C. (1990). A forward modeling strategy for balancing cross sections. Am. Assoc. Petrol. Geol. Bull. 74:521531.Google Scholar
Muehlberger, W. R. (1961). Conjugate joint sets of small dihedral angle. J. Geol. 69:211218.Google Scholar
Mueller, K. and Golombek, M. P. (2004). Compressional structures on Mars. Ann. Rev. Earth Planet. Sci. 32:435464.Google Scholar
Mueller, K. and Talling, P. (1997). Geomorphic evidence for tear faults accommodating lateral propagation of an active fault-bend fold, Wheeler Ridge, California. J. Struct. Geol. 19:397411.Google Scholar
Mühlhaus, H. B. and Vardoulakis, I. (1988). The thickness of shear bands in granular materials. Géotechnique 38:271284.Google Scholar
Mukhopadhyay, B. (2002). Water-rock interactions in the basement beneath Long Valley caldera: an oxygen isotope study of the Long Valley Exploratory Well drill cores. J. Volcanol. Geotherm. Res. 116:325359.Google Scholar
Muir Wood, D. (1990). Soil Behaviour and Critical State Soil Mechanics, Cambridge University Press, New York, 462 pp.Google Scholar
Mullen, M., Roundtree, R., Barree, B. and Turk, G. (2007). A composite determination of mechanical rock properties for stimulation design (what to do when you don’t have a sonic log). Paper SPE 108139 presented at the 2007 SPE Rocky Mountain Oil and Gas Technology Symposium, Denver, Colorado, 16–18 April 2007.Google Scholar
Muller, J. R. and Martel, S. J. (2000). Numerical models of translational landslide rupture surface growth. Pure Appl. Geophys. 157:10091038.Google Scholar
Murgatroyd, J. B. (1942). The significance of surface marks on fractured glass. J. Soc. Glass Technol. 26:155171.Google Scholar
Muraoka, H. and Kamata, H. (1983). Displacement distribution along minor fault traces. J. Struct. Geol. 5:483495.Google Scholar
Murrell, S. A. F. (1958). The strength of coal under triaxial compression. In Mechanical Properties of Non-Metallic Materials (ed. Walton, W.), pp. 123153, Butterworths.Google Scholar
Murrell, S. A. F. (1964). The theory of the propagation of elliptical Griffith cracks under various conditions of plane strain or plane stress, Pts. II, III. Br. J. Appl. Phys. 15:12111223.Google Scholar
Muskhelishvili, N. I. (1963). Some Basic Problems of the Mathematical Theory of Elasticity (4th edn), Noordhoff, Groningen, Holland.Google Scholar
Myers, R. and Aydin, A. (2004). The evolution of faults formed by shearing across joint zones in sandstone. J. Struct. Geol. 26:947966.Google Scholar
Nadai, A. (1950). Theory of Flow and Fracture of Solids (Vol. 1), McGraw-Hill, New York.Google Scholar
Nahm, A. K. and Schultz, R. A. (2007). Outcrop-scale properties of Burns Formation at Meridiana Planum, Mars. Geophys. Res. Lett. 34, L20203; doi:10.1029/2007GL031005.Google Scholar
Nahm, A. L., and Schultz, R. A. (2010). Evaluation of the orogenic belt hypothesis for the origin of the Thaumasia Highlands, Mars. J. Geophys. Res. 115:E04008; doi:10.1029/2009JE003327.Google Scholar
Nahm, A. L. and Schultz, R. A. (2011). Magnitude of global contraction on Mars from analysis of surface faults: implications for Martian thermal history. Icarus 211:389400; doi:10.1016/j.icarus.2010.11.003.Google Scholar
Nakamura, K. (1977). Volcanoes as possible indicators of tectonic stress orientation: principle and proposal. J. Volcanol. Geotherm. Res. 2:116.Google Scholar
Nara, Y., Yamanaka, H., Oe, Y. and Kaneko, K. (2013). Influence of temperature and water on subcritical crack growth parameters and long-term strength for igneous rocks. Geophys. J. Int. 193:4760.Google Scholar
Narr, W. and Suppe, J. (1991). Joint spacing in sedimentary rocks. J. Struct. Geol. 13:10371048.Google Scholar
National Academy of Sciences (1996). Rock Fractures and Fluid Flow: Contemporary Understanding and Applications, National Academy Press, Washington, DC, 551 pp.Google Scholar
Naylor, M. A., Mandl, G. and Sijpesteijn, C. H. K. (1986). Fault geometries in basement-induced wrench faulting under different initial stress states. J. Struct. Geol. 8:737752.Google Scholar
Naylor, M., Sinclair, H. D., Willett, S. and Cowie, P. A. (2005). A discrete element model for orogenesis and accretionary wedge growth. J. Geophys. Res. 110:B12403; doi:10.1029/2003JB002940.Google Scholar
Nehari, Z. (1952). Conformal Mapping, Dover Publications, New York.Google Scholar
Nemat-Nasser, S. and Horii, H. (1982). Compression-induced nonplanar crack extension with application to splitting, exfoliation, and rockburst. J. Geophys. Res. 87:68056821.Google Scholar
Nemčok, M., Schamel, S. and Gayer, R. (2005). Thrustbelts: Structural Architecture, Thermal Regimes, and Petroleum Systems, Cambridge University Press, 541 pp.Google Scholar
Nenna, F. and Aydin, A. (2011). The formation and growth of pressure solution seams in clastic rocks: a field and analytical study. J. Struct. Geol. 33:633643.Google Scholar
Neuffer, D. P., Schultz, R. A. and Watters, R. J. (2006). Mechanisms of slope failure on Pyramid Mountain, a subglacial volcano in Wells Gray Provincial Park, British Columbia. Can. J. Earth Sci. 43:147155.Google Scholar
Neuffer, D. P. and Schultz, R. A. (2006). Mechanisms of slope failure in Valles Marineris, Mars. Quart. J. Eng. Geol. Hydrogeol. 39:227240.Google Scholar
Neuendorf, K. K. E., Mehl, J. P., Jr. and Jackson, J. A. (2005). Glossary of Geology (5th edn), Am. Geol. Inst., Alexandria, VA.Google Scholar
Newman, J. C. Jr. (2000). Irwin’s stress intensity factor—A historical perspective. In Fatigue and Fracture Mechanics: 31st Volume (eds. Halford, G. R. and Gallagher, J. P.), pp. 3953, Am. Soc. Test. Mat. Spec. Publ. 1389, West Conshohocken, Penn.Google Scholar
Nguyen, G. D., Nguyen, C. T., Bui, H. H. and Nguyen, V. P. (2016). Constitutive modelling of compaction localization in porous sandstones. Int. J. Rock Mech. Min. Sci. 83:5772.Google Scholar
Nguyen, O., Repetto, E. A., Ortiz, M. and Radovitzky, R. A. (2001). A cohesive model of fatigue crack growth. Int. J. Fracture 110:351369.Google Scholar
Nicol, A., Walsh, J. J., Watterson, J. and Bretan, P. G. (1995). Three-dimensional geometry and growth of conjugate normal faults. J. Struct. Geol. 17:847862.Google Scholar
Nicol, A., Watterson, J., Walsh, J. J. and Childs, C. (1996). The shapes, major axis orientations and displacement patterns of fault surfaces. J. Struct. Geol. 18:235248.Google Scholar
Nicol, A., Gillespie, P. A., Childs, C. and Walsh, J. J. (2002). Relay zones between mesoscopic thrust fault in layered sedimentary sequences. J. Struct. Geol. 24:709727.Google Scholar
Nicol, A., Walsh, J., Berryman, K. and Nodder, S. (2005). Growth of a normal fault by the accumulation of slip over millions of years. J. Struct. Geol. 27:327342.Google Scholar
Nicol, A., Childs, C., Walsh, J. J. and Schafer, K. W. (2013). A geometric model for the formation of deformation band clusters. J. Struct. Geol. 55:2133.Google Scholar
Nicol, A., Childs, C., Walsh, J. J., Manzocchi, T. and Schöpfer, M. P. J. (2016). Interactions and growth of faults in an outcrop-scale system. In The Geometry and Growth of Normal Faults (eds. Childs, C., Holdsworth, R. E., Jackson, C. A.-L. et al.), Geol. Soc. London Spec. Publ. 439. (First published online March 10, 2016; https://doi.org/10.1144/SP439.9.)Google Scholar
Nicolas, A. (1995). The Mid-Oceanic Ridges: Mountains Below Sea Level, Springer-Verlag, Berlin.Google Scholar
Nicholson, C., Seeber, L., Williams, P. and Sykes, L. R. (1986). Seismic evidence for conjugate slip and block rotation within the San Andreas fault system, southern California. Tectonics 5:629648.Google Scholar
Nicholson, R. and Pollard, D. D. (1985). Dilation and linkage of echelon cracks. J. Struct. Geol. 7:583590.Google Scholar
Nickelsen, R. P. and Hough, V. N. D. (1967). Jointing in the Appalachian plateau of Pennsylvania. Geol. Soc. Am. Bull. 78:609630.Google Scholar
Niño, F., Philip, H. and Chéry, J. (1998). The role of bed-parallel slip in the formation of blind thrust faults. J. Struct. Geol. 20:503516.Google Scholar
Nova, R. (2005). A simple elastoplastic model for soils and soft rocks. In Soil Constitutive Models: Evaluation, Selection, and Calibration (eds. Yamamuro, J. A. and Kaliakin, V. N.), pp. 380399, Am. Soc. Civ. Eng. Geotech. Spec. Publ. 128.Google Scholar
Nova, R. and Lagioia, R. (2000). Soft rocks: behaviour and modeling. In Eurock ‘96, Prediction and Performance in Rock Mechanics and Rock Engineering (ed. Barla, G.), pp. 15211540, Proc. ISRM Symp., vol. 3, Balkema, Rotterdam.Google Scholar
Nur, A., Ron, H. and Scotti, O. (1986). Fault mechanics and the kinematics of block rotations. Geology 14:746749.Google Scholar
Nygård, R., Gutierrez, M., Bratli, R. K. and Høeg, K. (2006). Brittle-ductile transition, shear failure and leakage in shales and mudrocks. Marine Petrol. Geol. 23:201212.Google Scholar
Odé, H. (1957). Mechanical analysis of the dike pattern of the Spanish Peaks area, Colorado. Geol. Soc. Am. Bull. 68:567576.Google Scholar
Odling, N. E. (1997). Scaling and connectivity of joint systems in sandstones from western Norway. J. Struct. Geol. 19:12571271.Google Scholar
Odonne, F. and Massonnat, G. (1992). Volume loss and deformation around conjugate fractures: comparison between a natural example and analogue experiments. J. Struct. Geol. 14:963972.Google Scholar
Oertel, G. (1965). The mechanism of faulting in clay experiments. Tectonophysics 2:343393.Google Scholar
Oertel, G. (1996). Stress and Deformation: A Handbook on Tensors in Geology, Oxford University Press.Google Scholar
Ogilvie, S. R. and Glover, P. W. J. (2001). The petrophysical properties of deformation bands in relation to their microstructure. Earth Planet. Sci. Lett. 193:129142.Google Scholar
Ohlmacher, G. C. and Aydin, A. (1995). Progressive deformation and fracture patterns during foreland thrusting in the southern Appalachians. Am. J. Sci. 295:943987.Google Scholar
Ohlmacher, G. C. and Aydin, A. (1997). Mechanics of vein, fault and solution surface formation in the Appalachian Valley and Ridge, northeastern Tennessee, USA: implications for fault friction, state of stress and fluid pressure. J. Struct. Geol. 19:927944.Google Scholar
Okal, E. A. and Langenhorst, A. R. (2000). Seismic properties of the Etanin Transform System, South Pacific. Phys. Earth Planet. Inter. 119:185208.Google Scholar
Okubo, C. H. (2007). Strength and deformability of light-toned layered deposits observed by MER Opportunity: Eagle to Erebus craters, Mars. Geophys. Res. Lett. 34:L20205; doi:10.1029/2007GL031327.Google Scholar
Okubo, C. H. and Martel, S. J. (1998). Pit crater formation on Kilauea volcano, Hawaii. J. Volcanol. Geotherm. Res. 86:118.Google Scholar
Okubo, C. H. and Schultz, R. A. (2003). Thrust fault vergence directions on Mars: a foundation for investigating global-scale Tharsis-driven tectonics. Geophys. Res. Lett. 30:2154, 10.1029/2003GL018664.Google Scholar
Okubo, C. H. and Schultz, R. A. (2004). Mechanical stratigraphy in the western equatorial region of Mars based on thrust fault-related fold topography and implications for near-surface volatile reservoirs. Geol. Soc. Am. Bull. 116:594605.Google Scholar
Okubo, C. H. and Schultz, R. A. (2005). Evolution of damage zone geometry and intensity in porous sandstone: insight from strain energy density. J. Geol. Soc. London 162:939949.Google Scholar
Okubo, C. H. and Schultz, R. A. (2006). Near-tip stress rotation and the development of deformation band stepover geometries in mode II. Geol. Soc. Am. Bull. 118:343348.Google Scholar
Okubo, C. H. and Schultz, R. A. (2007). Compactional deformation bands in Wingate Sandstone; additional evidence of an impact origin for Upheaval Dome, Utah. Earth Planet. Sci. Lett. 256:169181.Google Scholar
Okubo, C. H., Schultz, R. A., Chan, M. A., Komatsu, G., and the High-Resolution Imaging Science Experiment (HiRISE) Team (2009). Deformation band clusters on Mars and implications for subsurface fluid flow. Geol. Soc. Am. Bull. 121:474482.Google Scholar
Okubo, P. G. and Dieterich, J. H. (1984). Effects of fault properties on frictional instabilities produced on simulated faults. J. Geophys. Res. 89:58175827.Google Scholar
Okubo, P. G. and Dieterich, J. H. (1986). State variable fault constitutive relations for dynamic slip. In Earthquake Source Mechanics (eds. Das, S., Boatwright, J. and Scholz, C. H.), pp. 2535, Am. Geophys. Un. Geophys. Monog. 37.Google Scholar
Oldenburg, D. W. and Brune, J. N. (1975). An explanation for the orthogonality of ocean ridges and transform faults. J. Geophys. Res. 80:25752585.Google Scholar
Olson, E. L. and Cooke, M. L. (2005). Application of three fault growth criteria to the Puente Hills thrust system, Los Angeles, California, USA. J. Struct. Geol. 27:17651777.Google Scholar
Olson, J. E. (1993). Joint pattern development: effects of subcritical crack growth and mechanical crack interaction. J. Geophys. Res. 98:12,25112,265.Google Scholar
Olson, J. E. (2003). Sublinear scaling of fracture aperture versus length: an exception or the rule? J. Geophys. Res. 108:2413; doi:10.1029/2001JB000419.Google Scholar
Olson, J. E. (2004). Predicting fracture swarms—the influence of subcritical crack growth and the crack-tip process zone on joint spacing in rock. In The Initiation, Propagation, and Arrest of Joints and Other Fractures: Interpretations Based on Field Observations (eds. Cosgrove, J. W. and Engelder, T.), pp. 7387, Geol. Soc. London Spec. Publ. 231.Google Scholar
Olson, J. E. (2007). Fracture aperture, length and pattern geometry development under biaxial loading: a numerical study with applications to natural, cross-jointed systems. In Fracture-Like Damage and Localisation (eds. Couples, G. and Lewis, H.), pp. 123142, Geol. Soc. London Spec. Publ. 289.Google Scholar
Olson, J. and Pollard, D. D. (1989). Inferring paleostresses from natural fracture patterns: a new method. Geology 17:345348.Google Scholar
Olson, J. and Pollard, D. D. (1991). The initiation and growth of en échelon veins. J. Struct. Geol. 13:595608.Google Scholar
Olson, J. E. and Schultz, R. A. (2011). Comment on “A note on the scaling relations for opening mode fractures in rock” by C.H. Scholz. J. Struct. Geol. 33:15231524.Google Scholar
Olson, J. E., Hennings, P. H. and Laubach, S. E. (1998). Integrating wellbore data and geomechanical modeling for effective characterization of naturally fractured reservoirs. SPE/ISRM Eurock Conference, Trondheim, Norway, July 8–10.Google Scholar
Olson, J. E., Laubach, S. E., and Lander, R. H. (2007). Combining diagenesis and mechanics to quantify fracture aperture distributions and fracture pattern permeability. In Fractured Reservoirs (eds. Lonergan, L., Jolly, R. J. H., Rawnsley, K. and Sanderson, D. J.), pp. 101116, Geol. Soc. London Spec. Publ. 270.Google Scholar
Olsson, W. A. (1999). Theoretical and experimental investigation of compaction bands in porous rock. J. Geophys. Res. 104:72197228.Google Scholar
Olsson, W. A. (2000). Origin of Lüders’ bands in deformed rock. J. Geophys. Res. 105:59315938.Google Scholar
Olsson, W. A. (2001). Quasistatic propagation of compaction fronts in porous rock. Mech. Mat. 33:659668.Google Scholar
Olsson, W. A. and Holcomb, D. J. (2000). Compaction localization in porous rock. Geophys. Res. Lett. 27:35373540.Google Scholar
Omdal, E., Madland, M. V., Kristiansen, T. G., et al. (2010). Deformation behavior of chalk studied close to in situ reservoir conditions. Rock Mech. Rock Eng. 43:557580.Google Scholar
Onasch, C. M., Farver, J. R. and Dunne, W. M. (2010). The role of dilation and cementation in the formation of cataclasite in low temperature deformation of well cemented quartz-rich rocks. J. Struct. Geol. 32:19121922.Google Scholar
Opheim, J. A. and Gudmundsson, A. (1989). Formation and geometry of fractures, and related volcanism, of the Krafla fissure swarm, Northeast Iceland. Geol. Soc. Am. Bull. 101:16081622.Google Scholar
Pachell, M. A. and Evans, J. P. (2002). Growth, linkage, and termination processes of a 10-km-long strike-slip fault in jointed granite: the Gemini fault zone, Sierra Nevada, California. J. Struct. Geol. 24:19031924.Google Scholar
Palmer, A. C. and Rice, J. R. (1973). The growth of slip surfaces in the progressive failure of over-consolidated clay. Proc. Royal Soc. London A, 332:527548.Google Scholar
Papamichos, E. and Vardoulakis, I. (1995). Shear band formation in sand according to non-coaxial plasticity model. Géotechnique 45:649661.Google Scholar
Papamichos, E., Brignoli, M. and Santerelli, F. J. (1997). An experimental and theoretical study of partially saturated collapsible rocks. Mech. Cohesive-Frictional Mat. 2:251278.Google Scholar
Pappalardo, R. T. and Collins, G. C. (2005). Strained craters on Ganymede. J. Struct. Geol. 27:827838.Google Scholar
Pappalardo, R. T., Head, J. W., Collins, G. C., et al. (1998). Grooved terrain on Ganymede: first results from Galileo high-resolution imaging. Icarus 135:276302.Google Scholar
Paris, P., Gomez, M. and Anderson, W. (1961). A rational analytic theory of fatigue. The Trend in Eng. 13:914.Google Scholar
Paris, P. and Erdogan, F. (1963). A critical analysis of crack propagation laws. J. Basic Eng. 85:528534.Google Scholar
Paris, P. C. and Sih, G. C. (1965). Stress analysis of cracks. In Fracture Toughness Testing and its Applications. Amer. Soc. Testing Mater., Spec. Tech. Publ. 381:3081.Google Scholar
Park, K. and Paulino, G. H. (2011). Cohesive zone models: a critical review of traction-separation relationships across fracture surfaces. Appl. Mech. Rev. 64: 061002; doi:10.1115/1.4023110.Google Scholar
Park, R. G. (1989). Foundations of Structural Geology (2nd edn), Chapman and Hall, New York, 148 pp.Google Scholar
Parker, J. M. (1942). Regional systematic jointing in slightly deformed sedimentary rocks. Geol. Soc. Am. Bull. 53:381408.Google Scholar
Parnell, J. (2010). Potential of palaeofluid analysis for understanding oil charge history. Geofluids 10:7382.Google Scholar
Parnell, J., Watt, G. R., Middleton, D., Kelley, J. and Baron, M. (2004). Deformation band control on hydrocarbon migration. J. Sed. Res. 74:552560.Google Scholar
Parry, W. T., Chan, M. A. and Beitler, B. (2004). Chemical weathering indicates episodes of fluid flow in deformation bands in sandstone. Am. Assoc. Petrol. Geol. Bull. 88:175191.Google Scholar
Passchier, C. W. and Trouw, R. A. J. (1996). Microtectonics, Springer-Verlag, Berlin.Google Scholar
Paterson, M. S. (1978). Experimental Rock Deformation—the Brittle Field, Springer-Verlag, Heidelberg.Google Scholar
Paterson, M. S. and Wong, T.-f. (2005). Experimental Rock Deformation—the Brittle Field (2nd edn), Springer-Verlag, Heidelberg.Google Scholar
Paul, B. (1961). A modification of the Coulomb–Mohr theory of fracture. J. Appl. Mech. 28:259268.Google Scholar
Pawar, R. J., Bromhal, G. S., Carey, J. W., et al. (2015). Recent advances in risk assessment and risk management of geologic CO2 storage. Int. J. Greenhouse Gas Control 40:292311.Google Scholar
Peacock, D. C. P. (2002). Propagation, interaction and linkage in normal fault systems. Earth Sci. Rev. 58:121142.Google Scholar
Peacock, D. C. P. and Azzam, I. N. (2006). Development and scaling relationships of a stylolite population. J. Struct. Geol. 28:18831889.Google Scholar
Peacock, D. C. P. and Sanderson, D. J. (1991). Displacements, segment linkage and relay ramps in normal fault systems. J. Struct. Geol. 13:721733.Google Scholar
Peacock, D. C. P. and Sanderson, D. J. (1992). Effects of layering and anisotropy on fault geometry. J. Geol. Soc. London 149:793802.Google Scholar
Peacock, D. C. P. and Sanderson, D. J. (1993). Estimating strain from fault slip using a line sample. J. Struct. Geol. 15:15131516.Google Scholar
Peacock, D. C. P. and Sanderson, D. J. (1994). Geometry and development of relay ramps in normal fault systems. Am. Assoc. Petrol. Geol. Bull. 78:147165.Google Scholar
Peacock, D. C. P. and Sanderson, D. J. (1995). Strike-slip relay ramps. J. Struct. Geol. 17:13511360.Google Scholar
Peacock, D. C. P., Jones, G., Knipe, F. J. and McAllister, E. (1998). Large lateral ramps in the Eocene Valkyr shear zone: extensional ductile faulting controlled by plutonism in southern British Columbia: Discussion. J. Struct. Geol. 20:487488.Google Scholar
Peacock, D. C. P., Knipe, R. J. and Sanderson, D. J. (2000). Glossary of normal faults. J. Struct. Geol. 22:291305.Google Scholar
Peacock, D. C. P., Nixon, C. W., Rotevatn, A., Sanderson, D. J. and Zuluaga, L. F. (2016). Glossary of fault and other fracture networks. J. Struct. Geol. 92:1229.Google Scholar
Peacock, D. C. P., Nixon, C. W., Rotevatn, A., Sanderson, D. J. and Zuluaga, L. F. (2017). Interacting faults. J. Struct. Geol. 97:122.Google Scholar
Peacock, D. C. P., Sanderson, D. J. and Rotevatn, A. (2018). Relationships between fractures. J. Struct. Geol. 106:4153.Google Scholar
Peng, S. D. (1971). Stresses within elastic circular cylinders loaded uniaxially and triaxially. Int. J. Rock Mech. Min. Sci. & Geomech. Abstr. 8:399432.Google Scholar
Peng, S. and Johnson, A. M. (1972). Crack growth and faulting in cylindrical specimens of Chelmsford granite. Int. J. Rock Mech. Min. Sci. 9:3786.Google Scholar
Pennacchioni, G. (2005). Control of the geometry of precursor brittle structures on the type of ductile shear zone in the Adamello tonalites, Southern Alps (Italy). J. Struct. Geol. 27:627644.Google Scholar
Perfettini, H., Schmittbul, J. and Cochard, A. (2003). Shear and normal load perturbations on a two-dimensional continuous fault: 2. Dynamic triggering. J. Geophys. Res. 108:2409; doi:10.1029/2002JB001805.Google Scholar
Perkins, T. K. and Krech, W. W. (1966). Effect of cleavage rate and stress level on apparent surface energies of rocks. Soc. Petrol. Eng. J. 6:308314.Google Scholar
Peters, J. F., Muthuswamy, M., Wibowo, J. and Tordesillas, A. (2005). Characterization of force chains in granular material. Phys. Rev. E 72:041307.Google Scholar
Petersen, M. D., Mueller, C. S., Moschetti, M. P., et al. (2015). Incorporating induced seismicity in the 2014 United States National Seismic Hazard Model—Results of 2014 workshop and sensitivity studies: US Geological Survey Open-File Report 2015–1070, 69 pp.; http://dx.doi.org/10.3133/ofr20151070.Google Scholar
Petit, J.-P. (1987). Criteria for the sense of movement on fault surfaces in brittle rocks. J. Struct. Geol. 9:597608.Google Scholar
Petit, J.-P. and Barquins, M. (1988). Can natural faults propagate under mode-II conditions? Tectonics 7:12431256.Google Scholar
Petit, J.-P. and Mattauer, M. (1995). Palaeostress superimposition deduced from mesoscale structures in limestone: the Matelles exposure, Languedoc, France. J. Struct. Geol. 17:245256.Google Scholar
Petrie, E. S., Evans, J. P. and Bauer, S. J. (2014). Failure of cap-rock seals as determined from mechanical stratigraphy, stress history, and tensile-failure analysis of exhumed analogs. Am. Assoc. Petrol. Geol. Bull. 98:23652389.Google Scholar
Philip, H. and Meghraoui, M. (1983). Structural analysis and interpretation of the surface deformations of the El Asnam earthquake of October 10, 1980. Tectonics 2:1749.Google Scholar
Phipps Morgan, J. and Parmentier, E. M. (1985). Causes and rate limiting mechanisms of ridge propagation: a fracture mechanics model. J. Geophys. Res. 90:86038612.Google Scholar
Philit, S. (2017). Elaboration of a structural, petrophysical and mechanical model of faults in porous sandstones; implication for migration and fluid entrapment, PhD thesis, University of Montpellier, France, 298 pp.Google Scholar
Philit, S., Soliva, R., Ballas, G. and Fossen, H. (2017). Grain deformation processes in porous quartz sandstones: insight from the clusters of cataclastic deformation bands. EPJ Web of Conferences 140:07002, Powders and Grains 2017; doi:10.1051/epjconf/201714007002.Google Scholar
Philit, S., Soliva, R., Castilla, R., Ballas, G. and Taillefer, A. (2018). Clusters of deformation bands in porous sandstones. J. Struct. Geol. 114:235250.Google Scholar
Pinter, N. (1995). Faulting on the volcanic tableland, Owens Valley, California. J. Geol. 103:7383.Google Scholar
Pittman, E. D. (1981). Effect of fault related granulation on porosity and permeability of quartz sandstones, Simpson Group (Ordovician), Oklahoma. Am. Assoc. Petrol. Geol. Bull. 65:23812387.Google Scholar
Plescia, J. B. and Golombek, M. P. (1986). Origin of planetary wrinkle ridges based on the study of terrestrial analogs. Geol. Soc. Am. Bull. 97:12891299.Google Scholar
Plumb, R. A. (1994). Variations of the least horizontal stress magnitude in sedimentary basins. In Rock Mechanics: Models and Measurements, Challenges from Industry (eds. Nelson, P. and Laubach, S. E.), pp. 7177, Balkema, Rotterdam.Google Scholar
Polit, A. T., Schultz, R. A. and Soliva, R. (2009). Geometry, displacement-length scaling, and extensional strain of normal faults on Mars with inferences on mechanical stratigraphy of the Martian crust. J. Struct. Geol. 31:662673.Google Scholar
Pollard, D. D. (1987). Elementary fracture mechanics applied to the structural interpretation of dykes. In Mafic Dyke Swarms (eds. Halls, H. C. and Fahrig, W. F.), pp. 524, Geol. Assoc. Canada Spec. Paper 34.Google Scholar
Pollard, D. D. and Aydin, A. (1984). Propagation and linkage of oceanic ridge segments. J. Geophys. Res. 89:10,01710,128.Google Scholar
Pollard, D. D. and Aydin, A. (1988). Progress in understanding jointing over the past century. Geol. Soc. Am. Bull. 100:11811204.Google Scholar
Pollard, D. D. and Fletcher, R. C. (2005). Fundamentals of Structural Geology, Cambridge University Press, Cambridge.Google Scholar
Pollard, D. D. and Holzhausen, G. (1979). On the mechanical interaction between a fluid-filled fracture and the Earth’s surface. Tectonophysics 53:2757.Google Scholar
Pollard, D. D. and Johnson, A. M. (1973). Mechanics of growth of some laccolithic intrusions in the Henry Mountains, Utah II: Bending and failure of overburden layers and sill formation. Tectonophysics 18:311354.Google Scholar
Pollard, D. D. and Muller, O. H. (1976). The effect of regional gradients in stress and magma pressure on the form of sheet intrusions in cross section. J. Geophys. Res. 81:975984.Google Scholar
Pollard, D. D. and Segall, P. (1987). Theoretical displacements and stresses near fractures in rock: with applications to faults, joints, dikes, and solution surfaces. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 277349, Academic Press, New York.Google Scholar
Pollard, D. D., Muller, O. H. and Dockstader, D. R. (1975). The form and growth of fingered sheet intrusions. Geol. Soc. Am. Bull. 86:351363.Google Scholar
Pollard, D. D., Segall, P. and Delaney, P. T. (1982). Formation and interpretation of dilatant echelon cracks. Geol. Soc. Am. Bull. 93:12911303.Google Scholar
Pollard, D. D., Delaney, P. T., Duffield, W. A., Endo, E. T. and Okamura, A. T. (1983). Surface deformation in volcanic rift zones. Tectonophysics 94:541584.Google Scholar
Pollard, D. D., Saltzer, S. D. and Rubin, A. M. (1993). Stress inversion methods: are they based on faulty assumptions? J. Struct. Geol. 15:10451054.Google Scholar
Polun, S. G., Gomez, F. and Tesfaye, S. (2018). Scaling properties of normal faults in the central Afar, Ethiopia and Djibouti: implications for strain partitioning during the final stages of continental breakup. J. Struct. Geol., 155:178–189.Google Scholar
Poncelet, E. F. (1958). The markings on fracture surfaces. J. Soc. Glass Tech. 42:279 T288 T.Google Scholar
Pook, L. P. (1971). The effect of crack angle on fracture toughness. Eng. Frac. Mech. 3:205218.Google Scholar
Preston, F. W. (1926). A study of the rupture of glass. J. Soc. Glass Technol. 10:234269.Google Scholar
Price, N. J. (1966). Fault and Joint Development in Brittle and Semi-Brittle Rock, Pergamon Press, Oxford.Google Scholar
Price, N. J. and Cosgrove, J. W. (1990). Analysis of Geological Structures, Cambridge University Press.Google Scholar
Priest, S. D. (1993). Discontinuity Analysis for Rock Engineering, Chapman and Hall, London.Google Scholar
Priest, S. D. and Hudson, J. A. (1976). Discontinuity spacings in rock. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 13:135148.Google Scholar
Priest, S. D. and Hudson, J. A. (1981). Estimation of discontinuity spacing and trace length using scanline surveys. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 18:183197.Google Scholar
Proffett, J. M. Jr. (1977). Cenozoic geology of the Yerington district, Nevada, and implications for the nature and origin of Basin and Range faulting. Geol. Soc. Am. Bull. 88:247266.Google Scholar
Prost, G. L. and Newsome, J. (2015). Caprock integrity determination at the Christina Lake thermal recovery project, Alberta.: Canadian Soc. Petrol. Geol. Bull. 64:309323.Google Scholar
Pugno, N., Ciavarella, M., Cornetti, P. and Carpinteri, A. (2006). A generalized Paris’ law for fatigue crack growth. J. Mech. Phys. Solids 54:13331349.Google Scholar
Qu, D. and Tveranger, J. (2016). Incorporation of deformation band fault damage zones in reservoir models. Am. Assoc. Petrol. Geol. Bull. 100:423443.Google Scholar
Quennell, A. M. (1958). The structural and geomorphic evolution of the Dead Sea rift. J. Geol. Soc. Lond. 114:124.Google Scholar
Quigley, M., Van Dissen, R., Litchfield, N., et al. (2012). Surface rupture during the 2010 Mw 7.1 Darfield (Canterbury) earthquake: implications for fault rupture dynamics and seismic-hazard analysis. Geology 40:5558.Google Scholar
Rabinovitch, A. and Bahat, D. (1999). Model of joint spacing distribution based on shadow compliance. J. Geophys. Res. 104:48774886.Google Scholar
Rabinowicz, E. (1958). The intrinsic variables affecting the stick-slip process. Proc. Phys. Soc. 71:668675.Google Scholar
Raduha, S., Butler, D., Mozley, P. S., et al. (2016). Potential seal bypass and caprock storage produced by deformation-band-to-opening-mode-fracture transition at the reservoir/caprock interface. Geofluids; doi:10.1111/gfl.12177.Google Scholar
Ragan, D. M. (2009). Structural Geology: an Introduction to Geometrical Techniques (4th edn), Cambridge University Press, 602 pp.Google Scholar
Ramulu, M. and Kobayashi, A. S. (1983). Dynamic crack curving: a photoelastic evaluation. Exp. Mech. 23:19.Google Scholar
Ramsay, J. G. and Allison, I. (1979). Structural analysis of shear zones in an alpinised Hercynian granite (Maggia Lappen, Pennine zone, central Alps). Schweitz. mineral. petrogr. Mitt. 59:251279.Google Scholar
Ramsay, J. G. and Huber, M. I. (1987). The Techniques of Modern Structural Geology—Volume 2: Folds and Fractures, Academic, San Diego, 700 pp.Google Scholar
Ramsay, J. G. and Lisle, R. J. (2000). The Techniques of Modern Structural Geology—Volume 3: Applications of Continuum Mechanics in Structural Geology, Academic, San Diego, 1061 pp.Google Scholar
Ramsey, J. M. and Chester, F. M. (2004). Hybrid fracture and the transition from extension fracture to shear fracture. Nature 428:6366.Google Scholar
Rawnsley, K. D., Rives, T., Petit, J.-P., Henscher, S. R. and Lumsden, A. C. (1992). Joint development in perturbed stress fields near faults. J. Struct. Geol. 14:939951.Google Scholar
Raynaud, S. and Carrio-Schaffhauser, E. (1992). Rock matrix structures in a zone influenced by a stylolite. J. Struct. Geol. 14:973980.Google Scholar
Reading, H. G. (1980). Characteristics and recognition of strike-slip systems. In Sedimentation in Oblique-Slip Mobile Zones (eds. Ballance, P. F. and Reading, H. G.) pp. 726, Int. Assoc. Sedimentol. Spec. Publ. 4.Google Scholar
Reasenberg, P. A. and Simpson, R. W. (1992). Response of regional seismicity to the static stress change produced by the Loma Prieta earthquake. Science 255:16871690.Google Scholar
Reches, Z. (1978). Analysis of faulting in three-dimensional strain field. Tectonophysics 47:109129.Google Scholar
Reches, Z. (1983). Faulting of rocks in three-dimensional strain fields II. Theoretical analysis. Tectonophysics 95:133156.Google Scholar
Reches, Z. (1987). Mechanical aspects of pull-apart basins and push-up swells with applications to the Dead Sea transform. Tectonophysics 141:7588.Google Scholar
Reches, Z. and Dieterich, J. H. (1983). Faulting of rocks in three-dimensional strain fields I. Failure of rocks in polyaxial, servo-control experiments. Tectonophysics 95:111132.Google Scholar
Reches, Z. and Lockner, D. A. (1994). Nucleation and growth of faults in brittle rocks. J. Geophys. Res. 99:18,15918,173.Google Scholar
Reid, H. F., Davis, W. M., Lawson, A. C. and Ransome, F. L. (1913). Report of the committee on the nomenclature of faults. Geol. Soc. Am. Bull. 24:163186.Google Scholar
Rekach, V. G. (1979). Manual of the Theory of Elasticity. Mir Publishers, Moscow (trans. M. Konyaeva).Google Scholar
Renard, F., Gratier, J.-P. and Jamtveit, B. (2000). Kinetics of crack-sealing, intergranular pressure solution, and compaction around active faults. J. Struct. Geol. 22:13951407.Google Scholar
Renard, F., Schmittbuhl, J., Gratier, J.-P., Meakin, P. and Merino, E. (2004). Three-dimensional roughness of stylolites in limestones. J. Geophys. Res. 109:B03209.Google Scholar
Renshaw, C. E. (1996). Influence of subcritical fracture growth on the connectivity of fracture networks. Water Resour. Res. 32:15191530.Google Scholar
Renshaw, C. E. (1997). Mechanical controls on the spatial density of opening-mode fracture networks. Geology 25:923926.Google Scholar
Renshaw, C. E. and Pollard, D. D. (1994a). Are large differential stresses required for straight fracture propagation? J. Struct. Geol. 16:817822.Google Scholar
Renshaw, C. E. and Pollard, D. D. (1994b). Numerical simulation of fracture set formation: a fracture mechanics model consistent with experimental observations. J. Geophys. Res. 99:93599372.Google Scholar
Renshaw, C. E. and Pollard, D. D. (1995). An experimentally verified criterion for propagation across unbonded frictional interfaces. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 32:237249.Google Scholar
Reynolds, O. (1885). On the dilatancy of media composed of rigid particles in contact. Philosophical Mag. 5:469.Google Scholar
Rice, J. R. (1968). A path-independent integral and the approximate analysis of strain concentration by notches and cracks. J. Appl. Mech. 35:379386.Google Scholar
Rice, J. R. (1978). Thermodynamics of the quasi-static growth of Griffith cracks. J. Mech. Phys. Solids 26:6178.Google Scholar
Rice, J. R. (1979). Theory of precursory processes in the inception of earthquake rupture. Beitr. Geophys. 88:91127.Google Scholar
Rice, J. R. (1980). The mechanics of earthquake rupture. In Physics of the Earth’s Interior (eds. Dziewonski, A. M. and Boschi, E.), pp. 555649, Proc. Int. School of Phys. “Enrico Fermi” Course 78, Italian Physical Society, North Holland Publ. Co.Google Scholar
Rice, J. R. (1983). Constitutive relations for fault slip and earthquake instabilities. Pure Appl. Geophys. 121:443475.Google Scholar
Rice, J. R. (1992). Fault stress states, pore-pressure distributions, and the weakness of the San Andreas fault. In Fault Mechanics and Transport Properties of Rocks (eds. Evans, B. and Wong, T.-f.), pp. 435459, Academic, New York.Google Scholar
Richards, L. R. and Read, S. A. L. (2013). Estimation of the Hoek–Brown parameter mi using Brazilian tensile test. Paper presented at the 47th US Rock Mech./Geomech. Symp., ARMA, 13465.Google Scholar
Rickman, R., Mullen, M., Petre, E., Grieser, B. and Kundert, D. (2008). A practical use of shale petrophysics for stimulation design optimization: all shale plays are not clones of the Barnett Shale. Paper SPE 115258 presented at the 2008 SPE Annual Technical Conference and Exhibition, Denver, Colorado, 21–24 September 2008.Google Scholar
Rickman, R., Mullen, M., Petre, E., Grieser, B. and Kundert, D. (2009). Petrophysics key in stimulating shales. American Oil and Gas Reporter, March 2009.Google Scholar
Riedel, J. J. and Labuz, J. F. (2006). Propagation of a shear band in sandstone. Int. J. Numer. Analyt. Methods Geomech. 31:12811299.Google Scholar
Riedel, W. (1929). Zur Mechanik geologischer Brucherscheinungen ein Beitrag zum Problem der “Fiederspalten.” Centralbl. f. Mineral. Geol. u. Paläont. Abt. 1929B:354368.Google Scholar
Riggs, E. M. and Green, H. W., II (2001). Shear localization in transformation-induced faulting: first order similarities to brittle shear failure. Tectonophysics 340:95107.Google Scholar
Rijken, M. C. M. (2005). Modeling naturally fractured reservoirs—From experimental rock mechanics to flow simulation, PhD dissertation, The University of Texas at Austin, 239 pp.Google Scholar
Rijken, P. and Cooke, M. L. (2001). Role of shale thickness on vertical connectivity of fractures: application of crack-bridging theory to the Austin Chalk, Texas. Tectonophysics 337:117133.Google Scholar
Risnes, R. (2001). Deformation and yield in high porosity outcrop chalk. Phys. Chem. Earth 26:5357.Google Scholar
Rispoli, R. (1981). Stress fields about strike-slip faults inferred from stylolites and tension gashes. Tectonophysics 75:T29T36.Google Scholar
Ritchie, R. O. (2011). The conflicts between strength and toughness. Nature Mater. 10:817822.Google Scholar
Rivero, C. and Shaw, J. H. (2011). Active folding and blind thrust faulting induced by basin inversion processes, inner California borderlands. In Thrust Fault-Related Folding (eds. McClay, K. R., Shaw, J. H. and Suppe, J.), pp. 187214, Amer. Assoc. Petrol. Geol. Mem. 94.Google Scholar
Robert, R., Robion, P., Souloumiac, P., David, C. and Saillet, E. (2018). Deformation bands, early markers of tectonic activity in front of a fold-and-thrust belt: example from the Tremp-Graus basin, southern Pyrenees, Spain. J. Struct. Geol. 110:6585.Google Scholar
Roberts, A. and Yielding, G. (1994). Continental extensional tectonics. In Continental Deformation (ed. Hancock, P. L.), pp. 223250, Pergamon, New York.Google Scholar
Roberts, G. P. (2007). Fault orientation variations along the strike of active normal fault systems in Italy and Greece: implications for predicting the orientations of subseismic-resolution faults in hydrocarbon reservoirs. Am. Assoc. Petrol. Geol. Bull. 91:120.Google Scholar
Roberts, G. P., Cowie, P., Papanikoulaou, I. and Michetti, A. M. (2004). Fault scaling relationships, deformation rates and seismic hazards: An example from the Lazio-Abruzzo Apennines, central Italy. J. Struct. Geol. 26:377398.Google Scholar
Roberts, J. C. (1961). Feather-fracture, and the mechanics of rock-jointing. Amer. J. Sci. 259:481492.Google Scholar
Roche, V., Homberg, C. and Rocher, M. (2012). Architecture and growth of normal fault zones in multilayer systems: a 3D field analysis in the South-Eastern Basin, France. J. Struct. Geol. 37:1935.Google Scholar
Roche, V., Homberg, C. and Rocher, M. (2013). Fault nucleation, restriction, and aspect ratio in layered sections: quantification of the strength and stiffness roles using numerical modeling. J. Geophys. Res. 118:44464460; doi:10.1002/JGRB.50279.Google Scholar
Rodgers, D. A. (1980). Analysis of pull-apart basin development produced by en echelon strike-slip faults. In Sedimentation in Oblique-Slip Mobile Zones (eds. Ballance, P. F. and Reading, H. G.), pp. 2741, Int. Assoc. Sedimentol. Spec. Publ. 4.Google Scholar
Rögnvaldsson, S. T., Gudmundsson, A. and Slunga, R. (1998). Seismotectonic analysis of the Tjörnes Fracture Zone, an active transform fault in north Iceland. J. Geophys. Res. 103:30,11730,129.Google Scholar
Roering, C. (1968). The geometrical significance of natural en echelon crack arrays. Tectonophysics 5:107123.Google Scholar
Roering, J. J., Cooke, M. L. and Pollard, D. D. (1997). Why blind thrust faults do not propagate to the Earth’s surface: numerical modeling of coseismic deformation associated with thrust-related anticlines. J. Geophys. Res. 102:11,90111,912.Google Scholar
Rogers, R. D. and Bird, D. K. (1987). Fracture propagation associated with dike emplacement at the Skaergaard intrusion, East Greenland. J. Struct. Geol. 9:7186.Google Scholar
Rogers, S. F. (2003). Critical stress-related permeability in fractured rocks. In Fracture and in-situ Stress Characterization of Hydrocarbon Reservoirs (ed. Ameen, M.), pp. 716, Geol. Soc. London Spec. Publ. 209.Google Scholar
Rogers, S. F., Elmo, D., Webb, G. and Moreno, C. G. (2016). DFN modelling of major structural instabilities in a large open pit for end of life planning purposes. Paper ARMA 16–882 presented at the 50th US Rock Mechanics/Geomechanics Symposium, Houston, Texas, June 26–29, 2016.Google Scholar
Romijn, R. and Groenenboom, J. (1997). Acoustic monitoring of hydraulic fracture growth. First Break 15:295303.Google Scholar
Ron, H., Aydin, A. and Nur, A. (1986). Strike-slip faulting and block rotation in the Lake Mead fault system. Geology 14:10201023.Google Scholar
Ron, H., Freund, R., Garfunkel, Z. and Nur, A. (1984). Block rotation by strike-slip faulting: structural and paleomagnetic evidence. J. Geophys. Res. 89:62566270.Google Scholar
Ron, H., Nur, A. and Aydin, A. (1993). Rotation of stress and blocks in the Lake Mead, Nevada, fault system. Geophys. Res. Lett. 20:17031706.Google Scholar
Roscoe, K. H. (1970). Influence of strains in soil mechanics. Géotechnique 20:129170.Google Scholar
Roscoe, K. H. and Burland, J. B. (1968). On the generalized stress-strain behaviour of ‘wet’ clay. In Engineering Plasticity (eds. Heyman, J. and Leckie, F. A.), pp. 535609, Cambridge University Press.Google Scholar
Roscoe, K. H. and Poorooshasb, H. B. (1963). A theoretical and experimental study of strains in triaxial tests on normally consolidated clays. Géotechnique 13:1238.Google Scholar
Roscoe, K. H., Schofield, A. N. and Wroth, C. P. (1958). On the yielding of soils. Géotechnique 8:2253.Google Scholar
Roscoe, K. H., Schofield, A. N. and Thurairajah, A. (1963). Yielding of clays in states wetter than critical. Géotechnique 13:211240.Google Scholar
Rosendahl, B. R. (1987). Architecture of continental rifts with special reference to East Africa. Annu. Rev. Earth Planet. Sci. 15:445503.Google Scholar
Rotevatn, A., Tveranger, J., Howell, J. and Fossen, H. (2009). Dynamic investigation of the effect of a relay ramp on simulated fluid flow: geocellular modeling of the Delicate Arch Ramp, Utah. Petrol. Geoscience 15:4558.Google Scholar
Rowe, P. W. (1962). The stress-dilatancy relation for static equilibrium of an assembly of particles in contact. Proc. Royal Soc. London A, 269:500527.Google Scholar
Rowland, S. M. and Duebendorfer, E. M. (1994). Structural Analysis and Synthesis: A Laboratory Course in Structural Geology (2nd edn), Blackwell Science, Cambridge, Mass.Google Scholar
Royden, L. H. (1985). The Vienna Basin: a thin-skinned pull-apart basin. In Strike-Slip Deformation, Basin Formation, and Sedimentation (eds. Biddle, K. T. and Christie-Blick, N.), pp. 319338, Soc. Econ. Paleon. Miner. Spec. Publ. 37.Google Scholar
Roznovsky, T. A. and Aydin, A. (2001). Concentration of shearing deformation related to changes in strike of monoclinal fold axes: the Waterpocket monocline, Utah. J. Struct. Geol. 23:15671579.Google Scholar
Rubin, A. M. (1990). A comparison of rift-zone tectonics in Iceland and Hawaii. Bull. Volcanol. 52:302319.Google Scholar
Rubin, A. M. (1992). Dike-induced faulting and graben subsidence in volcanic rift zones. J. Geophys. Res. 97:18391858.Google Scholar
Rubin, A. M. (1993a). Tensile fracture of rock at high confining pressure: implications for dike propagation. J. Geophys. Res. 98:15,91915,935.Google Scholar
Rubin, A.M. (1993b). Dikes vs. diapirs in viscoelastic rock. Earth Planet. Sci. Lett. 119:641–569.Google Scholar
Rubin, A. M. (1995a). Getting granitic dikes out of the source region. J. Geophys. Res. 100:59115929.Google Scholar
Rubin, A. M. (1995b). Propagation of magma-filled cracks. Ann. Rev. Earth Planet. Sci. 23:287336.Google Scholar
Rubin, A. M. (1998). Dike ascent in partially molten rock. J. Geophys. Res. 103:20,90120,919.Google Scholar
Rubin, A. M. and Pollard, D. D. (1987). Origins of blade-like dikes in volcanic rift zones. In Volcanism in Hawaii (eds. Decker, R. W., Wright, T. L. and Stauffer, P. H.), pp. 14491470, US Geol. Surv. Prof. Pap. 1350.Google Scholar
Rubin, A. M. and Pollard, D. D. (1988). Dike-induced faulting in rift zones of Iceland and Afar. Geology 16:413417.Google Scholar
Rudnicki, J. W. (1977). The inception of faulting in a rock mass with a weakened zone. J. Geophys. Res. 82:844854 (correction, J. Geophys. Res. 82:3437).Google Scholar
Rudnicki, J. W. (1980). Fracture mechanics applied to the Earth’s crust. Ann. Rev. Earth Planet. Sci. 8:489525.Google Scholar
Rudnicki, J. W. (2004). Shear and compaction band formation on an elliptic yield cap. J. Geophys. Res. 109:B03402, doi:10.1029/2003JB002633.Google Scholar
Rudnicki, J. W. (2007). Models for compaction band propagation. In Rock Physics and Geomechanics in the Study of Reservoirs and Repositories (eds. David, C. and Le Ravalec-Dupin, M.), pp. 107125, Geol. Soc. London Spec. Publ. 284.Google Scholar
Rudnicki, J. W. and Rice, J. R. (1975). Conditions for the localization of deformation in pressure-sensitive dilatant materials. J. Mech. Phys. Solids 23:371394.Google Scholar
Rudnicki, J. W. and Sternlof, K. R. (2005). Energy release model of compaction band propagation. Geophys. Res. Lett. 32:L16303; doi 10.1029/2005GL023602.Google Scholar
Ruegg, J. C., Kasser, M., Tarantola, A., Lepine, L. C. and Chouikrat, B. (1982). Deformations associated with the El Asnam earthquake of October 1980: geodetic determination of vertical and horizontal movements. Bull. Seismol. Soc. Am. 72:22272244.Google Scholar
Ruina, A. L. (1983). Slip instability and state variable friction laws. J. Geophys. Res. 88:10,35910,370.Google Scholar
Ruf, J. C., Rust, K. A. and Engelder, T. (1998). Investigating the effect of mechanical discontinuities on joint spacing. Tectonophysics 295:245257.Google Scholar
Rummel, F. (1987). Fracture mechanics approach to hydraulic fracturing stress measurements. In Fracture Mechanics of Rock (ed. Atkinson, B. K.), pp. 217240, Academic Press, New York.Google Scholar
Rutqvist, J. (2012). The geomechanics of CO2 storage in deep sedimentary formations. Geotech. Geol. Eng. 30:525551.Google Scholar
Rutqvist, J., Tsang, C. F. and Stephansson, O. (2000). Uncertainty in the maximum principal stress estimated from hydraulic fracturing measurements due to the presence of the induced fracture. Int. J. Rock Mech. Min. Sci. 37:107120.Google Scholar
Rutter, E. H. (1983). Pressure solution in nature, theory and experiment. J. Geol. Soc. London 140:725740.Google Scholar
Rutter, E. H. (1986). On the nomenclature of mode of failure transitions in rocks. Tectonophysics 122:381387.Google Scholar
Rutter, E. H. and Brodie, K. H. (1991). Lithosphere rheology—a note of caution. J. Struct. Geol. 13:363367.Google Scholar
Rutter, E. H. and Glover, C. T. (2012). The deformation of porous sandstones; are Byerlee friction and the critical state line equivalent? J. Struct. Geol. 44:129140.Google Scholar
Rutter, E. H., Maddock, R. H., Hall, S. H. and White, S. H. (1986). Comparative microstructures of natural and experimentally produced clay-bearing fault gouges. Pure Appl. Geophys. 124:330.Google Scholar
Ryan, M. P. and Sammis, C. G. (1978). Cyclic fracture mechanisms in cooling basalt. Geol. Soc. Am. Bull. 89:12951308.Google Scholar
Rybacki, E., Meier, T. and Dresen, G. (2016). What controls mechanical properties of rocks? —part II: brittleness. J. Petrol. Sci. Eng. 144:3958.Google Scholar
Saada, A. S., Liang, L., Figueroa, J. L. and Cope, C. T. (1999). Bifurcation and shear band propagation in sands. Géotechnique 49:367385.Google Scholar
Saade, A., Abou-Jaoude, G. and Wartman, J. (2016). Regional-scale co-seismic landslide assessment using limit equilibrium analysis. Eng. Geol. 204:5364.Google Scholar
Safaricz, M. and Davison, I. (2005). Pressure solution in chalk. Amer. Assoc. Petrol. Geol. Bull. 89:383401.Google Scholar
Saffer, D. M., Frye, K. M., Marone, C. and Mair, K. (2001). Laboratory results indicating complex and potentially unstable frictional behavior of smectite clay. Geophys. Res. Lett. 28:22972300.Google Scholar
Sagy, A., Reches, Z. and Roman, I. (2001). Dynamic fracturing: field and experimental observations. J. Struct. Geol. 23:12231239.Google Scholar
Saillet, E. and Wibberley, C. A. J. (2013). Permeability and flow impacts of faults and deformation bands in high-porosity sand reservoirs: Southeast Basin, France, analog. Am. Assoc. Petrol. Geol. Bull. 97:437464.Google Scholar
Sammis, C., King, G. and Biegel, R. (1987). The kinematics of gouge deformation. Pure Appl. Geophys. 125:777812.Google Scholar
Sammonds, P. R., Meredith, P. G. and Main, I. G. (1992). Role of pore fluids in the generation of seismic precursors to shear fracture. Nature 359:228230.Google Scholar
Sample, J. C., Woods, S., Bender, E. and Loveall, M. (2006). Relationship between deformation bands and petroleum migration in an exhumed reservoir rock, Los Angeles Basin, California, USA. Geofluids 6:105112.Google Scholar
Sanderson, D. J. and Marchini, W. R. D. (1984). Transpression. J. Struct. Geol. 6:449458.Google Scholar
Sandwell, D. (1986). Thermal stress and the spacings of transform faults. J. Geophys. Res. 91:64056417.Google Scholar
Sauber, J., Thatcher, W. and Solomon, S. (1986). Geodetic measurement of deformation in the central Mojave Desert, California. J. Geophys. Res. 91:12,68312,693.Google Scholar
Savage, J. C., and Hastie, L. M. (1966). Surface deformation associated with dip-slip faulting. J. Geophys. Res. 71:48974904.Google Scholar
Savage, J. C., and Burford, R. O. (1973). Geodetic determination of relative plate motion in central California. J. Geophys. Res. 78:832845.Google Scholar
Savage, J. C., Byerlee, J. D. and Lockner, D. A. (1996). Is internal friction friction? Geophys. Res. Lett. 23:487490.Google Scholar
Savalli, L. and Engelder, T. (2005). Mechanisms controlling rupture shape during subcritical growth of joints in layered rock. Geol. Soc. Am. Bull. 117:436449.Google Scholar
Schardin, E. B. (1959). Velocity effects in fracture. In Proceedings, International Conference on Atomic Mechanisms of Fracture (eds. Averback, B. L., Felbeck, D. K., Hahn, G. T. and Thomas, D. A.), Wiley, New York.Google Scholar
Schenk, P. M. and McKinnon, W. B. (1989). Fault offsets and lateral crustal movement on Europa: evidence for a mobile ice shell. Icarus 79:75100.Google Scholar
Schlische, R. W. (1995). Geometry and origin of fault-related folds in extensional settings. Am. Assoc. Petrol. Geol. Bull. 79:16611678.Google Scholar
Schlische, R. W., Young, S. S., Ackermann, R. V. and Gupta, A. (1996). Geometry and scaling relations of a population of very small rift-related normal faults. Geology 24:683686.Google Scholar
Schmidt, R. A. (1980). A microcrack model and its significance to hydraulic fracturing and fracture toughness testing. Paper ARMA 80–581 presented at the 21st US Symposium on Rock Mechanics, 27–30 May 1980, Rolla, Missouri.Google Scholar
Schmidt, R. A. and Huddle, C. W. (1977). Effect of confining pressure on fracture toughness of Indiana limestone. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 14:289293.Google Scholar
Schmitt, D. R., Currie, C. A. and Zhang, L. (2012). Crustal stress determination from boreholes and rock cores: fundamental principles. Tectonophysics 580:126 (corrigendum, Tectonophysics 586:206–207, 2013).Google Scholar
Schofield, A. (2005). Disturbed Soil Properties and Geotechnical Design, Thomas Telford, London.Google Scholar
Schofield, A. and Wroth, P. (1968). Critical State Soil Mechanics, McGraw-Hill, New York, 310 pp.Google Scholar
Scholz, C. H. (1968a). Microcracking and the inelastic deformation of rock in compression. J. Geophys. Res. 73:14171432.Google Scholar
Scholz, C. H. (1968b). Experimental study of the fracturing process in brittle rock. J. Geophys. Res. 73:14471454.Google Scholar
Scholz, C. H. (1982). Scaling laws for large earthquakes: consequences for physical models. Bull. Seismol. Soc. Amer. 72:114.Google Scholar
Scholz, C. H. (1990). The Mechanics of Earthquakes and Faulting, Cambridge University Press, 439 pp.Google Scholar
Scholz, C. H. (1997). Earthquake and fault populations and the calculation of brittle strain. Geowissenschaften 15:124130.Google Scholar
Scholz, C. H. (1998). Earthquakes and friction laws. Nature 391:3742.Google Scholar
Scholz, C. H. (2002). The Mechanics of Earthquakes and Faulting (2nd edn), Cambridge University Press, 471 pp.Google Scholar
Scholz, C. H. (2003). Good tidings. Nature 425:670671.Google Scholar
Scholz, C. H. and Cowie, P. A. (1990). Determination of total strain from faulting using slip measurements. Nature 346:837838.Google Scholar
Scholz, C. H. and Lawler, T. M. (2004). Slip tapers at the tips of faults and earthquake ruptures. Geophys. Res. Lett. 31:L21609; doi:10.1029/2004GL021030.Google Scholar
Scholz, C. H., Dawers, N. H., Yu, J.-Z. and Anders, M. H. (1993). Fault growth and fault scaling laws: preliminary results. J. Geophys. Res. 98:21,95121,961.Google Scholar
Schöpfer, M. P. J. and Childs, C. (2013). The orientation and dilatancy of shear bands in a bonded particle model for rock. Int. J. Rock Mech. Min. Sci. 57:7588.Google Scholar
Schöpfer, M. P. J., Childs, C. and Walsh, J. J. (2006). Localisation of normal faults in multilayer sequences. J. Struct. Geol. 28:816823.Google Scholar
Schöpfer, M. P. J., Arslan, A., Walsh, J. J. and Childs, C. (2011). Reconciliation of contrasting theories for fracture spacing in layered rocks. J. Struct. Geol. 33:551565.Google Scholar
Schreurs, G. and Colletta, B. (1998). Analogue modelling of faulting in zones of continental transpression and transtension. In Continental Transpressional and Transtensional Tectonics (eds. Holdsworth, R. E., Strachan, R. E. and Dewey, J. F.), pp. 5979, Geol. Soc. London Spec. Publ. 135.Google Scholar
Schreurs, G., Buiter, S .J. H., Boutelier, D., et al. (2006). Analogue benchmarks of shortening and extension experiments. In Analogue and Numerical Modeling of Crustal-Scale Processes (eds. Buiter, S. J. H. and Schreurs, G.), pp. 127, Geol. Soc. London Spec. Publ. 253.Google Scholar
Schueller, S., Braathen, A., Fossen, H. and Tveranger, J. (2013). Spatial distribution of deformation bands in damage zones of extensional faults in porous sandstones: statistical analysis of field data. J. Struct. Geol. 52:148162.Google Scholar
Schultz, R. A. (1989). Strike-slip faulting of ridged plains near Valles Marineris, Mars. Nature 341:424426.Google Scholar
Schultz, R. A. (1992). Mechanics of curved slip surfaces in rock. Eng. Anal. Bound. Elements 10:147154.Google Scholar
Schultz, R. A. (1993). Brittle strength of basaltic rock masses with applications to Venus. J. Geophys. Res. 98:10,88310,895.Google Scholar
Schultz, R. A. (1995a). Limits on strength and deformation properties of jointed basaltic rock masses. Rock Mech. Rock Eng. 28:115.Google Scholar
Schultz, R. A. (1995b). Gradients in extension and strain at Valles Marineris, Mars. Planet. Space Sci. 43:15611566.Google Scholar
Schultz, R. A. (1996). Relative scale and the strength and deformability of rock masses. J. Struct. Geol. 18:11391149.Google Scholar
Schultz, R. A. (1997). Displacement-length scaling for terrestrial and Martian faults: implications for Valles Marineris and shallow planetary grabens. J. Geophys. Res. 102:12,00912,015.Google Scholar
Schultz, R. A. (1998). Integrating rock mechanics into traditional geoscience curricula. J. Geosci. Ed. 46:141145.Google Scholar
Schultz, R. A. (1999). Understanding the process of faulting: selected challenges and opportunities at the edge of the 21st century. J. Struct. Geol. 21:985993.Google Scholar
Schultz, R. A. (2000a). Fault-population statistics at the Valles Marineris Extensional Province, Mars: implications for segment linkage, crustal strains, and its geodynamical development. Tectonophysics 316:169193.Google Scholar
Schultz, R. A. (2000b). Growth of geologic fractures into large-strain populations: nomenclature, subcritical crack growth, and implications for rock engineering. Int. J. Rock Mech. Min. Sci. 37:403411.Google Scholar
Schultz, R. A. (2000c). Localization of bedding plane slip and backthrust faults above blind thrust faults: keys to wrinkle ridge structure. J. Geophys. Res. 105:12,03512,052.Google Scholar
Schultz, R. A. (2002). Stability of rock slopes in Valles Marineris, Mars. Geophys. Res. Lett. 30:1932; doi:10.1029/2002GL015728.Google Scholar
Schultz, R. A. (2003a). Seismotectonics of the Amenthes Rupes thrust fault population, Mars. Geophys. Res. Lett. 30:1303, 10.1029/2002GL016475.Google Scholar
Schultz, R. A. (2003b). A method to relate initial elastic stress to fault population strains. Geophys. Res. Lett. 30:1593, 10.1029/2002GL016681.Google Scholar
Schultz, R. A. (2009). Scaling and paleodepth of compaction bands, Nevada and Utah. J. Geophys. Res. 114:B03407; doi:10.1029/2008JB005876.Google Scholar
Schultz, R. A. (2011a). Relationship of compaction bands in Utah to Laramide fault-related folding. Earth Planet. Sci. Lett. 304, 2935.Google Scholar
Schultz, R. A. (2011b). The influence of near-surface stratigraphy on the growth and scaling of normal faults in Bishop Tuff, California, with planetary implications (abstr.), in Lunar and Planetary Science XLII, #1471.Google Scholar
Schultz, R. A. and Aydin, A. (1990). Formation of interior basins associated with curved faults in Alaska. Tectonics 9:13871407.Google Scholar
Schultz, R. A. and Balasko, C. M. (2003). Growth of deformation bands into echelon and ladder geometries. Geophys. Res. Lett. 30:2033; doi:10.1029/2003GL018449.Google Scholar
Schultz, R. A. and Fori, A. N. (1996). Fault-length statistics and implications of graben sets at Candor Mensa, Mars. J. Struct. Geol. 18:373383.Google Scholar
Schultz, R. A. and Fossen, H. (2002). Displacement-length scaling in three dimensions: the importance of aspect ratio and application to deformation bands. J. Struct. Geol. 24:13891411.Google Scholar
Schultz, R. A. and Fossen, H. (2008). Terminology for structural discontinuities. Am. Assoc. Petrol. Geol. Bull. 92:853867.Google Scholar
Schultz, R. A. and Li, Q. (1995). Uniaxial strength testing of non-welded Calico Hills tuff, Yucca Mountain, Nevada. Eng. Geol. 40:287299.Google Scholar
Schultz, R. A. and Moore, J. M. (1996). New observations of grabens from the Needles District, Canyonlands National Park, Utah. In Geology and Resources of the Paradox Basin (eds. Huffman, A. C., Jr., Lund, W. R. and Godwin, L. H.), pp. 295302, Utah Geological Association and Four Corners Geological Society, Salt Lake City.Google Scholar
Schultz, R. A. and Siddharthan, R. (2005). A general framework for the occurrence and faulting of deformation bands in porous granular rocks. Tectonophysics 411:118.Google Scholar
Schultz, R. A. and Soliva, R. (2012). Propagation energies inferred from deformation bands in sandstone. Int. J. Fracture 176:135149.Google Scholar
Schultz, R. A. and Watters, T. R. (1995). Elastic buckling of fractured basalt on the Columbia Plateau, Washington State. In Rock Mechanics: Proceedings of the 35th US Symposium (eds. Daemen, J. J. K. and Schultz, R. A.), pp. 855860, Balkema, Rotterdam.Google Scholar
Schultz, R. A. and Watters, T. R. (2001). Forward mechanical modeling of the Amenthes Rupes thrust fault on Mars. Geophys. Res. Lett. 28:46594662.Google Scholar
Schultz, R. A. and Zuber, M. T. (1994). Observations, models, and mechanisms of failure of surface rocks surrounding planetary surface loads. J. Geophys. Res. 99:14,69114,702.Google Scholar
Schultz, R. A., Ward, K. A. and Hibbard, M. J. (1993). Dike emplacement in the Donner Summit pluton, northern Sierra Nevada, California. In Crustal Evolution of the Great Basin and Sierra Nevada (eds. Lahren, M. M., Trexler, J. H., Jr. and Spinosa, C.), pp. 277284, Geol. Soc. Am. Guidebook, Cordilleran/Rocky Mountains section, Reno, Nevada.Google Scholar
Schultz, R. A., Jensen, M. and Bradt, R. C. (1994). Single-crystal cleavage of brittle materials. Int. J. Fracture 65:291312.Google Scholar
Schultz, R. A., Okubo, C. H., Goudy, C. L. and Wilkins, S. J. (2004). Igneous dikes on Mars revealed by MOLA topography. Geology 32:889892.Google Scholar
Schultz, R. A., Okubo, C. H. and Wilkins, S. J. (2006). Displacement-length scaling relations for faults on the terrestrial planets. J. Struct. Geol. 28:21822193.Google Scholar
Schultz, R. A., Soliva, R., Fossen, H., Okubo, C. H. and Reeves, D. M. (2008a). Dependence of displacement–length scaling relations for fractures and deformation bands on the volumetric changes across them. J. Struct. Geol. 30:14051411.Google Scholar
Schultz, R. A., Mège, D. and Diot, H. (2008b). Emplacement conditions of igneous dikes in Ethiopian Traps. J. Volcanol. Geotherm. Res. 178:673692.Google Scholar
Schultz, R. A., Hauber, E., Kattenhorn, S., Okubo, C. H. and Watters, T. R. (2010a). Interpretation and analysis of planetary structures. J. Struct. Geol. 32:855875.Google Scholar
Schultz, R. A., Okubo, C. H. and Fossen, H. (2010b). Porosity and grain size controls on compaction band formation in Jurassic Navajo Sandstone. Geophys. Res. Lett. 37:L22306; doi:2010GL044909.Google Scholar
Schultz, R. A., Soliva, R., Okubo, C. H. and Mège, D. (2010c). Fault populations. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 457510, Cambridge University Press.Google Scholar
Schultz, R. A., Klimczak, C., Fossen, H., et al. (2013). Statistical tests of scaling relationships for geologic structures. J. Struct. Geol. 48:8594.Google Scholar
Schultz, R. A., Tong, X., Soofi, K. A., Sandwell, D. A. and Hennings, P. H. (2014). Using InSAR to detect active deformation associated with faults in Suban field, South Sumatra Basin, Indonesia. The Leading Edge 33:882888.Google Scholar
Schultz-Ela, D. D. and Walsh, P. (2002). Modeling of grabens extending above evaporites in Canyonlands National Park, Utah. J. Struct. Geol. 24:247275.Google Scholar
Searle, R. and Laughton, A. (1981). Fine-scale sonar study of tectonics and volcanism on the Reykjanes Ridge. Oceanol. Acta 4(suppl.):513.Google Scholar
Secor, D. T. (1965). Role of fluid pressure in jointing. Am. J. Sci. 263:633646.Google Scholar
Secor, D. T. and Pollard, D. D. (1975). On the stability of open hydraulic fractures in the Earth’s crust. Geophys. Res. Lett. 2:510513.Google Scholar
Segall, P. (1984a). Formation and growth of extensional fracture sets. Geol. Soc. Am. Bull. 95:454462.Google Scholar
Segall, P. (1984b). Rate-dependent extensional deformation resulting from crack growth in rock. J. Geophys. Res. 89:41854195.Google Scholar
Segall, P. (1989). Earthquakes triggered by fluid extraction. Geology 17: 942946.Google Scholar
Segall, P. (2010). Earthquake and Volcano Deformation, Princeton University Press, 432 pp.Google Scholar
Segall, P. and Fitzgerald, S. D. (1998). A note on induced stress changes in hydrocarbon and geothermal reservoirs. Tectonophysics 289:117128.Google Scholar
Segall, P. and Pollard, D. D. (1980). Mechanics of discontinuous faults. J. Geophys. Res. 85:43374350.Google Scholar
Segall, P. and Pollard, D. D. (1983a). Joint formation in granitic rock of the Sierra Nevada. Geol. Soc. Am. Bull. 94:563575.Google Scholar
Segall, P., and Pollard, D. D. (1983b). Nucleation and growth of strike-slip faults in granite. J. Geophys. Res. 88:555568.Google Scholar
Segall, P. and Pollard, D. D. (1983c). From joints and faults to photolineaments. Proc. Intern. Conf. Basement Tectonics 4: 1120.Google Scholar
Segall, P. and Simpson, C. (1986). Nucleation of ductile shear zones on dilatant fractures. Geology 14:5659.Google Scholar
Segall, P., Grasso, J. R., and Mossop, A. (1994). Poroelastic stressing and induced seismicity near the Laq gas field, southwestern France. J. Geophys. Res. 99:15,423–15,438.Google Scholar
Sempere, J.-C. and Macdonald, K. C. (1986). Overlapping spreading centers: implications from crack growth simulation by the displacement discontinuity method. Tectonics 5:151163.Google Scholar
Serafim, J. L. and Pereira, J. P. (1983). Considerations on the geomechanical classification of Bieniawski. Proc. Int. Symp. on Eng. Geol. and Underground Construction, vol. I, Lisbon, Portugal, pp. 3344.Google Scholar
Shah, R. C. and Kobayashi, A. S. (1973). Stress intensity factors for an elliptical crack approaching the surface of a semi-infinite solid. Int. J. Fracture 9:133146.Google Scholar
Shand, E. B. (1959). Breaking strength of glass determined from dimensions of fracture mirrors. J. Am. Ceramic Soc. 42:474477.Google Scholar
Sharon, E., Gross, S. P. and Fineberg, J. (1995). Local crack branching as a mechanism for instability in dynamic fracture. Phys. Rev. Lett. 74, 50965099.Google Scholar
Sharp, R. P. (1968). Sherwin till-Bishop tuff geological relationships, Sierra Nevada, California. Geol. Soc. Amer. Bull. 79, 351364.Google Scholar
Shaw, J. H., Bilotti, F. and Brennan, P. A. (1999). Patterns of imbricate thrusting. Geol. Soc. Am. Bull. 111:11401154.Google Scholar
Shaw, J. H., Plesch, A., Dolan, J. F., Pratt, T. L. and Fiore, P. (2002). Puente Hills blind-thrust system, Los Angeles, California. Seismol. Soc. Am. Bull. 92:29462960.Google Scholar
Shaw, P. R. and Lin, J. (1993). Causes and consequences of variations in faulting style at the Mid-Atlantic Ridge. J. Geophys. Res. 98:21,83921,851.Google Scholar
Sheldon, H. A., Barnicoat, A. C. and Ord, A. (2006). Numerical modelling of faulting and fluid flow in porous rocks: an approach based on critical state soil mechanics. J. Struct. Geol. 28:14681482.Google Scholar
Sheriff, R. E. and Geldart, L. P. (1995). Exploration Seismology, Cambridge University Press, New York.Google Scholar
Shimazaki, K. (1986). Small and large earthquakes: the effects of the thickness of seismogenic layer and the free surface. Earthquake Source Mech37:209216.Google Scholar
Shin, H., Santamarina, J. C. and Cartwright, J. A. (2010). Displacement field in contraction-driven faults. J. Geophys. Res. 115, B07408; doi:10.1029/2009JB006572.Google Scholar
Shipton, Z. K. and Cowie, P. A. (2001). Damage zone and slip-surface evolution over μm to km scales in high-porosity Navajo Sandstone, Utah. J. Struct. Geol. 23:18251844.Google Scholar
Shipton, Z. K. and Cowie, P. A. (2003). A conceptual model for the origin of fault damage zone structures in high-porosity sandstone. J. Struct. Geol. 25: 333344 (erratum, J. Struct. Geol. 25:1343–1345).Google Scholar
Shipton, Z. K., Evans, J. P., Robeson, K., Forster, C. B. and Snelgrove, S. (2002). Structural heterogeneity and permeability in faulted eolian sandstone: implications for subsurface modelling of faults. Am. Assoc. Petrol. Geol. Bull. 86:863883.Google Scholar
Shipton, Z. K., Evans, J. P. and Thompson, L. B. (2005). The geometry and thickness of deformation-band fault core and its influence on sealing characteristics of deformation-band fault zones. In Faults, Fluid Flow, and Petroleum Traps (eds. Sorkhabi, R. and Tsuji, Y.), pp. 181195, Am. Assoc. Petrol. Geol. Mem. 85.Google Scholar
Sibson, R. H. (1974). Frictional constraints on thrust, wrench and normal faults. Nature 249:542544.Google Scholar
Sibson, R. H. (1984). Roughness at the base of the seismogenic zone: contributing factors. J. Geophys. Res. 89:57915799.Google Scholar
Sibson, R. H. (1985). A note on fault reactivation. J. Struct. Geol. 7:751754.Google Scholar
Sibson, R. H. (1986a). Earthquakes and rock deformation in crustal fault zones. Annu. Rev. Earth Planet. Sci. 14:149175.Google Scholar
Sibson, R. H. (1986b). Brecciation processes in fault zones: inferences from earthquake rupturing. Pure Appl. Geophys. 124:159175.Google Scholar
Sibson, R. H. (1986c). Rupture interaction with fault jogs. In Earthquake Source Mechanics (eds. Das, S., Boatwright, J. and Scholz, C.H.), pp. 157167, Am. Geophys. Union. Geophys. Monog. 37.Google Scholar
Sibson, R. H. (1989). Earthquake faulting as a structural process. J. Struct. Geol. 11:114.Google Scholar
Sibson, R. H. (1994). An assessment of field evidence for ‘Byerlee’ friction. Pure Appl. Geophys. 142:645662.Google Scholar
Sibson, R. H. (1998). Brittle failure mode plots for compressional and extensional tectonic regimes. J. Struct. Geol. 20:655660.Google Scholar
Siddans, A. (1972). Slaty cleavage—a review of research since 1815. Earth Sci. Rev. 8:205232.Google Scholar
Sih, G. C. (1973). Some basic problems in fracture mechanics and new concepts. J. Eng. Frac. Mech. 5:365377.Google Scholar
Sih, G. C. (1974). Strain energy density factor applied to mixed mode crack problems. Int. J. Fracture 10:305322.Google Scholar
Sih, G. C. (1991). Mechanics of Fracture Initiation and Propagation—Surface and Volume Energy Density Applied as Failure Criterion, Kluwer Academic, Boston.Google Scholar
Sih, G. C., Paris, P. C. and Erdogan, F. (1962). Crack-tip, stress-intensity factors for plane extension and plate bending problems. J. Appl. Mech. 29:306312.Google Scholar
Sih, G. C. and Liebowitz, H. (1968). Mathematical theories of brittle fracture. In Fracture: An Advanced Treatise (ed. Liebowitz, H.), vol. II, pp. 67190, Academic, New York.Google Scholar
Sih, G. C. and Macdonald, B. (1974). Fracture mechanics applied to engineering problems: strain energy density fracture criterion. Eng. Fract. Mech. 6:361386.Google Scholar
Simony, P. S. and Carr, S. D. (1998). Large lateral ramps in the Eocene Valkyr shear zone: extensional ductile faulting controlled by plutonism in southern British Columbia: Reply. J. Struct. Geol. 20:489490.Google Scholar
Simpson, R. W. (1997). Quantifying Anderson’s fault types. J. Geophys. Res. 102:17,90917,919.Google Scholar
Singh, R. N. and Sun, G. (1990). Applications of fracture mechanics to some mining engineering problems. Mining Sci. Technol. 10:5360.Google Scholar
Skempton, A. W. (1954). The pore-pressure coefficients A and B. Géotechnique 4:143147.Google Scholar
Skempton, A. W. (1966). Some observations on tectonic shear zones. Paper presented at the 1st ISRM Congress, 25 September–1 October 1966, Lisbon, Portugal.Google Scholar
Skurtveit, E., Torabi, A., Gabrielsen, R. H. and Zoback, M. D. (2013). Experimental investigation of deformation mechanisms during shear-enhanced compaction in poorly lithified sandstone and sand. J. Geophys. Res. 118:40834100; doi:10.1002/jgrb.50342.Google Scholar
Smith, D. J., Ayatollahi, M. R. and Pavier, M. J. (2001). The role of T-stress in brittle fracture for linear elastic materials under mixed-mode loading. Fatigue Fracture Eng. Mater. Struct. 24:137150.Google Scholar
Smith, G. A. (1983). Porosity dependence of deformation bands in the Entrada Sandstone, La Plata County, Colorado. Mountain Geologist 20:8285.Google Scholar
Sneddon, I. N. (1946). The distribution of stress in the neighborhood of a crack in an elastic solid. Proc. Royal Soc. London A 187:229260.Google Scholar
Soliva, R. and Benedicto, A. (2004). A linkage criterion for segmented normal faults. J. Struct. Geol. 26:22512267.Google Scholar
Soliva, R. and Benedicto, A. (2005). Geometry, scaling relations and spacing of vertically restricted normal faults. J. Struct. Geol. 27:317325.Google Scholar
Soliva, R. and Schultz, R. A. (2008). Distributed and localized faulting in extensional settings: insight from the North Ethiopian Rift – Afar transition area. Tectonics 27:TC2003; doi:10.1029/2007TC002148.Google Scholar
Soliva, R., Schultz, R. A. and Benedicto, A. (2005). Three-dimensional displacement-length scaling and maximum dimension of normal faults in layered rocks. Geophys. Res. Lett. 32:L16302, 10.1029/2005GL023007.Google Scholar
Soliva, R., Benedicto, A. and Maerten, L. (2006). Spacing and linkage of confined normal faults: importance of mechanical thickness. J. Geophys. Res. 111:B01402, 10.1029/2004JB003507.Google Scholar
Soliva, R., Benedicto, A., Schultz, R. A., Maerten, L. and Micarelli, L. (2008). Displacement and interaction of normal fault segments branched at depth: implications for fault growth and potential earthquake rupture size. J. Struct. Geol. 30:12881299.Google Scholar
Soliva, R., Maerten, F., Petit, J.-P. and Auzias, V. (2010). Field evidences for the role of static friction on fracture orientation in extensional relays along strike-slip faults: comparison with photoelasticity and 3-D numerical modeling. J. Struct. Geol. 32:17211731.Google Scholar
Soliva, R., Schultz, R. A., Ballas, G., et al. (2013). A model of strain localization in porous sandstone as a function of tectonic setting, burial and material properties; new insight from Provence (southern France). J. Struct. Geol. 49:5063.Google Scholar
Soliva, R., Ballas, G., Fossen, H. and Philit, S. (2016). Tectonic regime controls clustering of deformation bands in porous sandstone. Geology 44:423426.Google Scholar
Solomon, S., Huang, P. and Meinke, L. (1988). The seismic moment budget of slowly spreading ridges. Nature 334:5860.Google Scholar
Solomon, S. C., McNutt, R. L. Jr., Watters, T. R., et al. (2008). Return to Mercury: a global perspective on MESSENGER’s first Mercury flyby. Science 321:5962.Google Scholar
Solum, J. G., Brandenburg, J. P., Naruk, S. J., et al. (2010). Characterization of deformation bands associated with normal and reverse stress states in Navajo Sandstone, Utah. Amer. Assoc. Petrol. Geol. Bull. 94:14531474.Google Scholar
Sommer, E. (1969). Formation of fracture lances in glass. Eng. Fracture Mech. 1:539546.Google Scholar
Sone, H. and Zoback, M. D. (2014). Time-dependent deformation of shale gas reservoir rocks and its long-term effect on the in situ state of stress. Int. J. Rock Mech. Min. Sci. 69:120132.Google Scholar
Sonmez, H., Ulusay, R. and Gokceoglu, C. (1998). A practical procedure for the back analysis of slope failures in closely jointed rock masses. Int. J. Rock Mech. Min. Sci. 35:219233.Google Scholar
Sonmez, H., Gokceoglu, C., Nefeslioglu, H. A., and Kayabasi, A. (2006). Estimation of rock modulus: for intact rocks with an artificial neural network and for rock masses with a new empirical equation. Int. J. Rock Mech. Min. Sci. 43:224235.Google Scholar
Stanchits, S., Fortin, J., Gueguen, Y. and Dresen, G. (2009). Initiation and propagation of compaction bands in dry and wet Bentheim Sandstone. Pure Appl. Geophys. 166:843868.Google Scholar
Steefel, C. I., DePaolo, D. J. and Lichtner, P. C. (2005). Reactive transport modeling: an essential tool and a new research approach for the Earth sciences. Earth Planet. Sci. Lett. 240:539558.Google Scholar
Stefanov, Y. P. and Bakeev, R. A. (2014). Deformation and fracture structures in strike-slip faulting. Eng. Fract. Mech. 129:102111.Google Scholar
Stefanov, Y. P., Bakeev, R. A., Rebetsky, Y. L. and Kontorovich, V. A. (2014). Structure and formation stages of a fault zone in a geomedium layer in strike-slip displacement of the basement. Physical Mesomechanics 17:112; translated from original Russian text in Fizicheskaya Mezomekhanika 16:41–52, 2013.Google Scholar
Stein, R. S. (1999). The role of stress transfer in earthquake occurrence. Nature 402:605609.Google Scholar
Stein, R. S. and King, G. C. P. (1984). Seismic potential revealed by folding: 1983 Coalinga, California, earthquake. Science 224:869872.Google Scholar
Stein, R. S., King, G. C. P. and Lin, J. (1992). Change in failure stress on the southern San Andreas fault system caused by the magnitude = 7.4 Landers earthquake. Science 258:13281332.Google Scholar
Stein, R. S., King, G. C. P. and Lin, J. (1994). Stress triggering of the 1994 M = 6.7 Northridge, California, earthquake by its predecessors. Science 165:14321435.Google Scholar
Sternlof, K. R., Rudnicki, J. W. and Pollard, D. D. (2005). Anticrack inclusion model for compaction bands in sandstone. J. Geophys. Res. 110:B11403; doi:10.1029/2005JB003764.Google 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 Resour. Res. 42:W07425; doi:10.1029/2005WR004664.Google Scholar
Stesky, R. M., Brace, W. F., Riley, D. K. and Robin, P.-Y. F. (1975). Friction in faulted rock at high temperature and pressure. Tectonophysics 23:177203.Google Scholar
Stevens, B. (1911). The laws of intrusion. Bull. Am. Inst. Min. Eng. 49:123.Google Scholar
Stewart, S. A., Harvey, M. J., Otto, S. C. and Weston, P. J. (1996). Influence of salt on fault geometry: examples from the UK salt basins. In Salt Tectonics (eds. Alsop, G. I., Blundell, D. J. and Davison, I.), pp. 175202, Geol. Soc. Spec. Publ. 100.Google Scholar
Stirling, M. W., Wesnousky, S. G. and Shimazaki, K. (1996). Fault trace complexity, cumulative slip, and the shape of the magnitude-frequency distribution for strike-slip faults: a global survey. Geophys. J. Int. 124:833868.Google Scholar
Stockdale, P. B. (1922). Stylolites: their nature and origin. Indiana Univ. Stud. 9:197.Google Scholar
Stone, D. S. (1999). Foreland basement-involved structures: discussion. Am. Assoc. Petrol. Geol. Bull. 83:20062016.Google Scholar
Storti, F., Holdsworth, R. E. and Salvini, F. (2003). Intraplate strike-slip deformation belts. In Intraplate Strike-Slip Deformation Belts (eds. Storti, F., Holdsworth, R. E. and Salvini, F.), pp. 114, Geol. Soc. London Spec. Publ. 210.Google Scholar
Strokova, L. (2013). Effect of the overconsolidation ratio of soils in surface settlements due to tunneling. Sciences in Cold and Actic Regions 5:637643.Google Scholar
Strom, R. G., Trask, N. J. and Guest, J. E. (1975). Tectonism and volcanism on Mercury. J. Geophys. Res. 80:24782507.Google Scholar
Strom, R. G., Malhotra, R., Xiao, Z., et al. (2015). The inner solar system cratering record and the evolution of impactor populations. Res. Astron. Astrophys. 15:407.Google Scholar
Sumy, D. F., Cochran, E. S., Keranen, K. M., Wei, M. and Abers, G. A. (2014). Observations of static Coulomb stress triggering of the November 2011 M5.7 Oklahoma earthquake sequence. J. Geophys. Res. 119; doi:10.1002/2013JB010612.Google Scholar
Suppe, J. (1983). Geometry and kinematics of fault-bend folding. Am. J. Sci. 283:648721.Google Scholar
Suppe, J. (1985). Principles of Structural Geology, Prentice-Hall, Englewood Cliffs, New Jersey.Google Scholar
Suppe, J. (2007). Absolute fault and crustal strength from wedge tapers. Geology 35:11271130.Google Scholar
Suppe, J. and Connors, C. (1992). Critical taper wedge mechanics of fold-and-thrust belts on Venus: initial results from Magellan. J. Geophys. Res. 97:13,545–13,561.Google Scholar
Suppe, J. and Medwedeff, D. (1990). Geometry and kinematics of fault propagation folding. Eclogae Geol. Helv. 83:409454.Google Scholar
Swain, M. V. and Hagan, J. T. (1978). Some observations of overlapping interacting cracks. Eng. Fracture Mech. 10:299304.Google Scholar
Swain, M. V., Lawn, B. R. and Burns, S. J. (1974). Cleavage step deformation. J. Mat. Sci. 9:175183.Google Scholar
Swanson, P. L. (1984). Subcritical crack growth and other time- and environment-dependent behavior in crustal rocks. J. Geophys. Res. 89:41374152.Google Scholar
Swanson, P. L. (1987). Tensile fracture resistance mechanisms in brittle polycrystals: an ultrasonics and in situ microscopy investigation. J. Geophys. Res. 92:80158036.Google Scholar
Sykes, L. R. (1967). Mechanism of earthquakes and nature of faulting on the Mid-Atlantic Ridge. J. Geophys. Res. 72:21312153.Google Scholar
Sylvester, A. G. (1988). Strike-slip faults. Geol. Soc. Am. Bull. 100:16661703.Google Scholar
Sylvester, A. G. and Smith, R. R. (1976). Tectonic transpression and basement-controlled deformation in San Andreas fault zone, Salton trough, California. Am. Assoc. Petrol. Geol. Bull. 60:20812102.Google Scholar
Taboada, A., Bousquet, J. C. and Philip, H. (1993). Coseismic elastic models of folds above blind thrusts in the Betic Cordilleras (Spain) and evaluation of seismic hazard. Tectonophysics 220:223241.Google Scholar
Tada, H., Paris, P. C. and Irwin, G. R. (2000). The Stress Analysis of Cracks Handbook (3rd edn), Am Soc. Mat. Eng., New York.Google Scholar
Tanaka, K. L., Anderson, R., Dohm, J. M., et al. (2010). Planetary structural mapping. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 351396, Cambridge University Press.Google Scholar
Tapp, B. and Cook, J. (1988). Pressure solution zone propagation in naturally deformed carbonate rocks. Geology 16:182185.Google Scholar
Tapponier, P. and Brace, W. F. (1976). Development of stress-induced microcracks in Westerly granite. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 13:103112.Google Scholar
Tavani, S. and Storti, F. (2011). Layer-parallel shortening templates associated with double-edge fault-propagation folding. In Thrust Fault-Related Folding (eds. McClay, K. R., Shaw, J. H. and Suppe, J.), pp. 121135, Amer. Assoc. Petrol. Geol. Mem. 94.Google Scholar
Taylor, B., Goodliffe, A. and Martinez, F. (2009). Initiation of transform faults at rifted continental margins. C. R. Geosci. 341:428438.Google Scholar
Taylor, W. L., Pollard, D. D. and Aydin, A. (1999). Fluid flow in discrete joint sets: field observations and numerical simulations. J. Geophys. Res. 104:28,983–29,006.Google Scholar
Tchalenko, J. S. (1970). Similarities between shear zones of different magnitudes. Geol. Soc. Am. Bull. 81:16251640.Google Scholar
Tchalenko, J. S. and Ambrasays, N. N. (1970). Structural analyses of the Dasht-e Bayaz (Iran) earthquake fractures. Geol. Soc. Am. Bull. 81:4160.Google Scholar
Tembe, S., Vajdova, V., Wong, T.-f. and Zhu, W. (2006). Initiation and propagation of strain localization in circumferentially notched samples of two porous sandstones. J. Geophys. Res. 111:B02409; doi 10.1029/2005JB003611.Google Scholar
Tembe, S., Baud, P. and Wong, T.-f. (2008). Stress conditions for the propagation of discrete compaction bands in porous sandstone. J. Geophys. Res. 113:B09409; doi:10.1029/2007JB005439.Google Scholar
ten Brink, U. S., Katzman, R. and Lin, J. (1996). Three-dimensional models of deformation near strike-slip faults. J. Geophys. Res. 101:16,20516,220.Google Scholar
Terzaghi, K. (1946). Rock defects and loads on tunnel support. In Rock Tunneling with Steel Supports (eds. Proctor, R. V. and White, T.), pp. 1599, Commercial Shearing Co., Youngstown, Ohio.Google Scholar
Terzaghi, K. (1943). Theoretical Soil Mechanics, Wiley, New York, 265 pp.Google Scholar
Terzaghi, K., Peck, R. B. and Mesri, G. (1996). Soil Mechanics in Engineering Practice (3rd edn), Wiley, New York, 565 pp.Google Scholar
Teufel, L. W., Rhett, D. W. and Farrell, H. E. (1991). Effect of reservoir depletion and pore pressure drawdown on in situ stress and deformation in the Ekofisk Field, North Sea. In Rock Mechanics as a Multidisciplinary Science (ed. Roegiers, J. C.), pp. 6372, Balkema, Rotterdam.Google Scholar
Thallak, S., Holder, J. and Gray, K. E. (1993). The pressure dependence of apparent hydrofracture toughness. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 30:831835.Google Scholar
Tharp, T. M. and Coffin, D. T. (1985). Field application of fracture mechanics analysis to small rock slopes. In Proc. 26th US Symp. Rock Mech., pp. 667–674.Google Scholar
Thomas, A. L. and Pollard, D. D. (1993). The geometry of echelon fractures in rock: implications from laboratory and numerical experiments. J. Struct. Geol. 15:323334.Google Scholar
Timoshenko, S. P. and Goodier, J. N. (1970). Theory of Elasticity, McGraw-Hill, New York.Google Scholar
Tirosh, J. and Catz, E. (1981). Mixed-mode fracture angle and fracture locus of materials subjected to compressive loading. Eng. Frac. Mech. 14:2738.Google Scholar
Tondi, E., Antonellini, M., Aydin, A., Marchegiani, L. and Cello, G. (2006). The role of deformation bands, stylolites and sheared stylolites in fault development in carbonate grainstones of Majella Mountain, Italy. J. Struct. Geol. 28:376391.Google Scholar
Torabi, A. and Alikarami, R. (2012). Heterogeneity within deformation bands in sandstone reservoirs. Paper ARMA-2012-347 presented at the 46th US Rock Mechanics/Geomechanics Symposium, 24–27 June, Chicago, Illinois.Google Scholar
Torabi, A. and Berg, S. S. (2011). Scaling of fault attributes: a review. Mar. Petrol. Geol. 28:14441460.Google Scholar
Torabi, A. and Fossen, H. (2009). Spatial variation of microstructure and petrophysical properties along deformation bands in reservoir sandstones. Am. Assoc. Petrol. Geol. Bull. 93:919938.Google Scholar
Toussaint, R., Aharonov, E., Koehn, D., et al. (2018). Stylolites: a review. J. Struct. Geol. 114:163195; doi:10.1016/j.jsg.2018.05.003.Google Scholar
Townend, E., Thompson, B. D., Benson, P. M., et al. (2008). Imaging compaction band propagation in Diemelstadt sandstone using acoustic emission locations. Geophys. Res. Lett. 35, L15301; doi:10.1029/2008GL034723.Google Scholar
Townend, J. and Zoback, M. D. (2000). How faulting keeps the crust strong. Geology 28:399402.Google Scholar
Treagus, S. H. and Lisle, R. J. (1997). Do principal surfaces of stress and strain always exist? J. Struct. Geol. 19:9971010.Google Scholar
Trudgill, B. and Cartwright, J. (1994). Relay-ramp forms and normal-fault linkages, Canyonlands National Park, Utah. Geol. Soc. Am. Bull. 106:11431157.Google Scholar
Tse, S. T. and Rice, J. R. (1986). Crustal earthquake instability in relation to the variation of frictional slip properties. J. Geophys. Res., 91:94529472.Google Scholar
Tucholke, B. E. and Lin, J. (1994). A geological model for the structure of ridge segments in slow-spreading ocean crust. J. Geophys. Res. 99:11,937–11,958.Google Scholar
Tucholke, B. E., Lin, J., Kleinrock, M. C., et al. (1997). Segmentation and crustal structure of the western Mid-Atlantic Ridge flank, 25°25’–27°10’N and 0–29 m.y. J. Geophys. Res. 102:10,203–10,223.Google Scholar
Tucholke, B. E., Lin, J. and Kleinrock, M. C. (1998). Megamullions and mullion structure defining oceanic metamorphic core complexes on the Mid-Atlantic Ridge. J. Geophys. Res. 103:98579866.Google Scholar
Tufts, B. R., Greenberg, R., Hoppa, G. and Geissler, P. (1999). Astypalaea Linea: a large-scale strike-slip fault on Europa. Icarus 141:5364.Google Scholar
Turcotte, D. L. and Schubert, G. (1982). Geodynamics: Applications of Continuum Physics to Geological Problems, Wiley, New York, 450 pp.Google Scholar
Turner, C. E. (1979). Methods for post-yield fracture safety assessment. In Post-Yield Fracture Mechanics (ed. Latzko, D. G. H.), pp. 23210, Applied Science Publishers, London.Google Scholar
Twidale, C. R. (1973). On the origin of sheet jointing. Rock Mech. 5:163187.Google Scholar
Twiss, R. J. and Marrett, R. (2010a). Determining brittle extension and shear strain using fault length and displacement systematics: part I: theory. J. Struct. Geol. 32:19601977.Google Scholar
Twiss, R. J. and Marrett, R. (2010b). Determining brittle extension and shear strain using fault length and displacement systematics: part II: data evaluation and test of the theory. J. Struct. Geol. 32:19781995.Google Scholar
Twiss, R. J. and Moores, E. M. (1992). Structural Geology, Freeman and Co., New York.Google Scholar
Twiss, R. J. and Moores, E. M. (2007). Structural Geology (2nd edn), Freeman and Co., New York.Google Scholar
Tworzydlo, W. W. and Hamzeh, O. N. (1997). On the importance of normal vibrations in modeling of stick slip in rock sliding. J. Geophys. Res. 102:15,091–15,103.Google Scholar
Ucar, R. (1986). Determination of shear failure envelope in rock masses. J. Geotech. Eng. Div. Am. Soc. Civ. Engrs. 112:303315.Google Scholar
Unger, D. J. (1995). Analytical Fracture Mechanics, Academic, New York.Google Scholar
Vajdova, V. and Wong, T.-f. (2003). Incremental propagation of discrete compaction bands: acoustic emission and microstructural observations on circumferentially notched samples of Bentheim sandstone. Geophys. Res. Lett. 30:1775; doi:10.1029/2003GL017750.Google Scholar
Vajdova, V., Baud, P. and Wong, T.-f. (2004a). Compaction, dilatancy, and failure in porous carbonate rocks. J. Geophys. Res. 109:B05204; doi:10.1029/2003JB002508.Google Scholar
Vajdova, V., Baud, P. and Wong, T.-f. (2004b). Permeability evolution during localized deformation in Bentheim sandstone. J. Geophys. Res. 109:B10406; doi:10.1029/2003JB002942.Google Scholar
Valkó, P. and Economides, M. J. (1995). Hydraulic Fracture Mechanics, Wiley, New York, 298 pp.Google Scholar
Vallianatos, F. and Sammonds, P. (2011). A non-extensive statistics of the fault-population at the Valles Marineris extensional province, Mars. Tectonophysics 509:5054.Google Scholar
van der Pluijm, B. A. and Marshak, S. (1997). Earth Structure: an Introduction to Structural Geology and Tectonics, McGraw-Hill.Google Scholar
Vermeer, P. A. (1990). The orientation of shear bands in biaxial tests. Géotechnique 40:223236.Google Scholar
Vermeer, P. A. and de Borst, R. (1984). Non-associated plasticity for soils, concrete, and rock. Heron 29:364.Google Scholar
Vermilye, J. M. and Scholz, C. H. (1995). Relation between vein length and aperture. J. Struct. Geol. 17:423434.Google Scholar
Vermilye, J. M. and Scholz, C. H. (1998). The process zone: a microstructural view of fault growth. J. Geophys. Res. 103:12,22312,237.Google Scholar
Vidale, J. E., Agnew, D. C., Johnson, M. J. S. and Oppenheimer, D. H. (1998). Absence of earthquake correlation with Earth tides: an indication of high preseismic fault stress rate. J. Geophys. Res. 103:24,56724,572.Google Scholar
Villemin, T., Angelier, J. and Sunwoo, C. (1995). Fractal distribution of fault length and offsets: implications of brittle deformation evaluation – the Lorrain coal basin. In: Fractals in the Earth Sciences (eds. Barton, C. C. and LaPointe, P. R.), pp. 205225, Plenum Press, New York.Google Scholar
Wachtman, J. B. (1996). Mechanical Properties of Ceramics, Wiley, New York.Google Scholar
Wallace, M. H. and Kemeny, J. (1992). Nucleation and growth of dip-slip faults in a stable craton. J. Geophys. Res. 97:71457157.Google Scholar
Wallner, H. (1939). Linienstrukturen an Bruchflachen. Zeitschrift für Physik 114:368378.Google Scholar
Walsh, F. R., III, and Zoback, M. D. (2015). Oklahoma’s recent earthquakes and saltwater disposal. Sci. Adv. 1:e1500195; doi:10.1126/sciadv.1500195.Google Scholar
Walsh, J. B. (1965). The effect of cracks on the uniaxial compression of rocks. J. Geophys. Res. 70:399411.Google Scholar
Walsh, J. J. and Watterson, J. (1987). Distributions of cumulative displacement and seismic slip on a single normal fault surface. J. Struct. Geol. 9:10391046.Google Scholar
Walsh, J. J. and Watterson, J. (1988). Analysis of the relationship between displacements and dimensions of faults. J. Struct. Geol. 10:239247.Google Scholar
Walsh, J. J. and Watterson, J. (1989). Displacement gradients on fault surfaces. J. Struct. Geol. 11:307316.Google Scholar
Walsh, J. J. and Watterson, J. (1991). Geometric and kinematic coherence and scale effects in normal fault systems. In The Geometry of Normal Faults (eds. Roberts, A. M., Yielding, G. and Freeman, B.), pp. 193203, Geol. Soc. London Spec. Publ. 56.Google Scholar
Walsh, J. J., Watterson, J. and Yielding, G. (1991). The importance of small-scale faulting in regional extension. Nature 351:391393.Google Scholar
Walsh, J. J., Watterson, J., Bailey, W. R. and Childs, C. (1999). Fault relays, bends and branch lines. J. Struct. Geol. 21:10191026.Google Scholar
Walsh, J. J., Nicol, A. and Childs, C. (2002). An alternative model for the growth of faults. J. Struct. Geol. 24:16691675.Google Scholar
Walsh, J. J., Bailey, W. R., Childs, C., Nicol, A. and Bonson, C. G. (2003). Formation of segmented normal faults: a 3-D perspective. J. Struct. Geol. 25:12511262.Google Scholar
Walsh, P. and Schultz-Ela, D. D. (2003). Mechanics of graben evolution in Canyonlands National Park, Utah. Geol. Soc. Am. Bull. 115:259270.Google Scholar
Wang, H., Marongiu-Porcu, M. and Economides, M. J. (2016). Poroelastic and poroplastic modeling of hydraulic fracturing in brittle and ductile formations. SPE Production and Operations 31:4759.Google Scholar
Wang, R. and Kemeny, J. M. (1995). A new empirical failure criterion for rock under polyaxial compressive stresses. In Rock Mechanics: Proceedings of the 35th US Symposium (eds. Daemen, J. J. K. and Schultz, R. A.), pp. 453458, Balkema, Rotterdam.Google Scholar
Warpinski, N. R. (1985). Measurement of width and pressure in a propagating hydraulic fracture. Soc. Petrol. Eng. J. 25:4654.Google Scholar
Watters, T. R. (1988). Wrinkle ridge assemblages on the terrestrial planets. J. Geophys. Res. 93:10,23610,254.Google Scholar
Watters, T. R. (1991). Origin of periodically spaced wrinkle ridges on the Tharsis Plateau of Mars. J. Geophys. Res. 96:15,59915,616.Google Scholar
Watters, T. R. (1993). Compressional tectonism on Mars. J. Geophys. Res. 98:17049–17060.Google Scholar
Watters, T. R. and Johnson, C. L. (2010). Lunar tectonics. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 121182, Cambridge University Press.Google Scholar
Watters, T. R. and Maxwell, T. A. (1983). Cross-cutting relations and relative age of ridges and faults in the Tharsis region of Mars. Icarus 56:278298.Google Scholar
Watters, T. R. and Maxwell, T. A. (1986). Orientation, relative age, and extent of the Tharsis Plateau ridge system. J. Geophys. Res. 91:81138125.Google Scholar
Watters, T. R. and Nimmo, F. (2010). The tectonics of Mercury. In Planetary Tectonics (eds. Watters, T. R. and Schultz, R. A.), pp. 1580, Cambridge University Press.Google Scholar
Watters, T. R., Robinson, M. S. and Cook, A. C. (1998). Topography of lobate scarps on Mercury: new constraints on the planet’s contraction. Geology 26:991994.Google Scholar
Watters, T. R., Schultz, R. A. and Robinson, M. S. (2000). Displacement-length relations of thrust faults associated with lobate scarps on Mercury and Mars: Comparison with terrestrial faults. Geophys. Res. Lett. 27:36593662.Google Scholar
Watterson, J. (1986). Fault dimensions, displacements and growth. Pure Appl. Geophys. 124:365373.Google Scholar
Watterson, J., Nicol, A., Walsh, J. J. and Meier, D. (1998). Strains at the intersections of synchronous conjugate normal faults. J. Struct. Geol. 20:363370.Google Scholar
Wawersik, W. R. and Brace, W. F. (1971). Post-failure behavior of a granite and diabase. Rock Mech. 3:6185.Google Scholar
Wawersik, W. R. and Fairhurst, C. (1970). A study of brittle rock fracture in laboratory compression experiments. Int. J. Rock Mech. Min. Sci. 7:561575.Google Scholar
Weijermars, R. (1997). Principles of Rock Mechanics, Alboran Science Publishing, Amsterdam.Google Scholar
Weinberger, R., Baer, G., Shamir, G. and Agnon, A. (1995). Deformation bands associated with dyke propagation in porous sandstone, Makhtesh Ramon, Israel. In Physics and Chemistry of Dikes (eds. Baer, G. and Heimann, A.), pp. 95112, Balkema, Rotterdam.Google Scholar
Weissel, J. K. and Karner, G. D. (1989). Flexural uplift of rift flanks due to mechanical unloading of the lithosphere during extension. J. Geophys. Res. 94:13,91913,950.Google Scholar
Welch, M. J., Knipe, R. J., Souque, C. and Davis, R. K. (2009). A quadshear kinematic model for folding and clay smear development in fault zones. Tectonophysics 471:186202.Google Scholar
Wells, A. A. (1961). Unstable crack propagation in metals: cleavage and fast fracture. Proc. Crack Prop. Symp. 1, paper 84, Cranfield, UK.Google Scholar
Wells, A. A. (1963). Application of fracture mechanics at and beyond general yielding. British Welding J. 10:563570.Google Scholar
Wells, D. L. and Coppersmith, K. J. (1994). New empirical relationships among magnitude, rupture length, rupture width, rupture area, and surface displacement. Bull. Seismol. Soc. Am. 84:9741002.Google Scholar
Wennberg, O. P., Casani, G., Jahanpanah, A., et al. (2013). Deformation bands in chalk, examples from the Shetland Group of the Oseberg Field, North Sea, Norway. J. Struct. Geol. 56:103117.Google Scholar
Wernicke, B. (1981). Low-angle normal faults in the Basin and Range province: nappe tectonics in an extending orogen. Nature 291:645648.Google Scholar
Wernicke, B. (1992). Cenozoic extensional tectonics of the U.S. Cordillera. In The Cordilleran Orogen: Conterminous U.S. (eds. Burchfiel, B. C., Lipman, P. W. and Zoback, M. L.) pp. 553581, Geol. Soc. Am., The Geology of North America, v. G–3.Google Scholar
Wesnousky, S. G. (1988). Seismological and structural evolution of strike-slip faults. Nature 335:340343.Google Scholar
Wesnousky, S. G. (1999). Crustal deformation and the stability of the Gutenberg-Richter relationship. Bull. Seismol. Soc. Am. 89:11311137.Google Scholar
Wesnousky, S. G. (2005). The San Andreas and Walker Lane fault systems, western North America: transpression, transtension, cumulative slip and the structural evolution of a major transform plate boundary. J. Struct. Geol. 27:15051512.Google Scholar
Wesnousky, S. G. (2008). Displacement and geometrical characteristics of earthquake surface ruptures: issues and implications for seismic hazard analysis and the earthquake rupture process. Bull. Seismol. Soc. Am. 98:16091632.Google Scholar
Westaway, R. (1992). Seismic moment summation for historical earthquakes in Italy: tectonic implications. J. Geophys. Res. 97:15,43715,464.Google Scholar
Westaway, R. (1994). Quantitative analysis of populations of small faults. J. Struct. Geol. 16:12591273.Google Scholar
Westergaard, H. M. (1939). Bearing pressures and cracks. J. Appl. Mech. 61:A49A53.Google Scholar
Whittaker, B. N., Singh, R. N. and Sun, G. (1992). Rock Fracture Mechanics: Principles, Design and Applications, Elsevier, New York, 570 pp.Google Scholar
Wibberley, C. A. J., Petit, J.-P. and Rives, T. (1999). Mechanics of high displacement gradient faulting prior to lithification. J. Struct. Geol. 21:251257.Google Scholar
Wibberley, C. A. J., Petit, J.-P. and Rives, T. (2000). Mechanics of cataclastic ‘deformation band’ faulting in high-porosity sandstone, Provence. Comptes Rendus Acad. Sci., Paris 331:419425.Google Scholar
Wibberley, C. A. J., Petit, J.-P. and Rives, T. (2007). The mechanics of fault distribution and localization in high-porosity sands, Provence, France. In The Relationship Between Damage and Localization (eds. Lewis, H. and Couples, G. D.), pp. 1946, J. Geol. Soc. London Spec. Publ. 289.Google Scholar
Wilcox, R. E., Harding, T. P. and Seely, D. R. (1973). Basic wrench tectonics. Am. Assoc. Petrol. Geol. Bull. 57:7496.Google Scholar
Wilkins, S. J. (2002). Mechanical and statistical aspects of faulting: from coseismic rupture to cumulative deformation. Unpublished PhD dissertation, University of Nevada, Reno, 181 pp.Google Scholar
Wilkins, S .J. and Gross, M. R. (2002). Normal fault growth in layered rocks at Split Mountain, Utah: influence of mechanical stratigraphy on dip linkage, fault restriction and fault scaling. J. Struct. Geol. 24:14131429 (erratum, J. Struct. Geol. 24, 2007).Google Scholar
Wilkins, S. J. and Naruk, S. J. (2007). Quantitative analysis of slip-induced dilation with application to fault seal. Am. Assoc. Petrol. Geol. Bull. 91:97113.Google Scholar
Wilkins, S. J. and Schultz, R. A. (2003). Cross faults in extensional settings: stress triggering, displacement localization, and implications for the origin of blunt troughs at Valles Marineris, Mars. J. Geophys. Res. 108:5056, 10.1029/2002JE001968.Google Scholar
Wilkins, S. J. and Schultz, R. A. (2005). 3-D cohesive end-zone model for source scaling of strike-slip interplate earthquakes. Bull. Seismol. Soc. Am. 95:22322258.Google Scholar
Wilkins, S. J., Gross, M. R., Wacker, M., Eyal, Y. and Engelder, T. (2001). Faulted joints: kinematics, displacement–length scaling relations and criteria for their identification. J. Struct. Geol. 23:315327.Google Scholar
Wilkins, S. J., Schultz, R. A., Anderson, R. C., Dohm, J. M., and Dawers, N. C. (2002). Deformation rates from faulting at the Tempe Terra extensional province, Mars. Geophys. Res. Lett. 29:1884, 10.1029/2002GL015391.Google Scholar
Willemse, E. J. M. (1997). Segmented normal faults: correspondence between three-dimensional mechanical models and field data. J. Geophys. Res. 102:675692.Google Scholar
Willemse, E. J. M. and Pollard, D. D. (1998). On the orientation and pattern of wing cracks and solution surfaces at the tips of a sliding flaw or fault. J. Geophys. Res. 103:24272438.Google Scholar
Willemse, E. J. M. and Pollard, D. D. (2000). Normal fault growth: evolution of tipline shapes and slip distributions. In Aspects of Tectonic Faulting (eds. Lehner, F. K. and Urai, J. L.), pp. 193–226, Springer, New York.Google Scholar
Willemse, E. J. M., Pollard, D. D. and Aydin, A. (1996). Three-dimensional analyses of slip distributions on normal fault arrays with consequences for fault scaling. J. Struct. Geol. 18:295309.Google Scholar
Willemse, E. J. M., Peacock, D. C. P. and Aydin, A. (1997). Nucleation and growth of strike-slip faults in limestones from Somerset, UK. J. Struct. Geol. 19:14611477.Google Scholar
Williams, A. (1958). Oblique-slip faults and rotated stress systems. Geol. Mag. 95:207218.Google Scholar
Williams, C. A., Connors, C., Dahlen, F. A., Price, E. J. and Suppe, J. (1994). Effect of the brittle-ductile transition on the topography of compressive mountain belts on Earth and Venus. J. Geophys. Res. 99:19,94719,974.Google Scholar
Williams, J. G. and Ewing, P. D. (1972). Fracture under complex stress: the angled crack problem. Int. J. Fracture Mech. 8:441446.Google Scholar
Williams, M. L. (1957). On the stress distribution at the base of a stationary crack. J. Appl. Mech. 24:109114.Google Scholar
Willis, B. and Willis, R. (1934). Geologic Structures, McGraw-Hill, New York.Google Scholar
Wilson, C. J. N. and Hildreth, W. (1997). The Bishop Tuff: new insights from eruptive stratigraphy. J. Geol. 105:407439; doi:10.1086/515937.Google Scholar
Wilson, J. E., Goodwin, L. B. and Lewis, C. J. (2003). Deformation bands in nonwelded ignimbrites: petrophysical controls on fault-zone deformation and evidence of preferential fluid flow. Geology 31:837840.Google Scholar
Wilson, J. T. (1965). A new class of faults and their bearing on continental drift. Nature 207:343347.Google Scholar
Wiprut, D. and Zoback, M. D. (2002). Fault reactivation, leakage potential, and hydrocarbon column heights in the North Sea. In Hydrocarbon Seal Quantification (eds. Koestler, A. G. and Hunsdale, R.), pp. 203219, Norwegian Petroleum Society Spec. Publ. 11.Google Scholar
Withjack, M. O., Schlische, R. W. and Henza, A. A. (2007). Scaled experimental models of extension: dry sand vs. wet clay. Bull. Houston Geol. Soc. 49:3149.Google Scholar
Wojtal, S. (1986). Deformation within foreland thrust sheets by populations of minor faults. J. Struct. Geol. 8:341360.Google Scholar
Wojtal, S. (1989). Measuring displacement gradients and strains in faulted rocks. J. Struct. Geol. 11:669678.Google Scholar
Wojtal, S. F. (1996). Changes in fault displacement populations correlated to linkage between faults. J. Struct. Geol. 18:265279.Google Scholar
Wojtal, S. and Mitra, G. (1988). Nature of deformation in some fault rocks from Appalachian thrusts. In Geometries and Mechanisms of Thrusting, with Special Reference to the Appalachians (eds. Mitra, G. and Wojtal, S.), pp. 1733, Geol. Soc. Am. Spec. Pap. 222.Google Scholar
Wolf, H., König, D. and Triantafyllidis, T. (2003). Experimental investigation of shear band patterns in granular material. J. Struct. Geol. 25:12291240.Google Scholar
Wood, D. S. (1974). Current views on the development of slaty cleavage. Rev. Earth Planet. Sci. 2:137.Google Scholar
Woodcock, N. H. and Fischer, M. (1986). Strike-slip duplexes. J. Struct. Geol. 8:725735.Google Scholar
Woodcock, N. H. and Schubert, C. (1994). Continental strike-slip tectonics. In Continental Deformation (ed. Hancock, P. L.), pp. 251263, Pergamon, New York.Google Scholar
Woodworth, J. B. (1896). On the fracture system of joints, with remarks on certain great fractures. Boston Soc. Natural History Proc. 27:163183.Google Scholar
Wong, T.-f. (1982). Shear fracture energy of Westerly granite from post-failure behavior. J. Geophys. Res. 87:9901000.Google Scholar
Wong, T.-f. and Baud, P. (1999). Mechanical compaction of porous sandstone. Oil & Gas Sci. Technol. – Rev. IFP 54:715727.Google Scholar
Wong, T.-f. and Baud, P. (2012). The brittle–ductile transition in rock: a review. J. Struct. Geol. 44:2553.Google Scholar
Wong, T.-f., Szeto, H. and Zhang, J. (1992). Effect of loading path and porosity on the failure mode of porous rocks. In Micromechanical Modelling of Quasi-Brittle Materials Behavior (ed. Li, V. C.), Appl. Mech. Rev. 45:281293.Google Scholar
Wong, T.-f., David, C. and Zhu, W. (1997). The transition from brittle faulting to cataclastic flow in porous sandstones: mechanical deformation. J. Geophys. Res. 102:30093025.Google Scholar
Wong, T.-f., Baud, P. and Klein, E. (2001). Localized failure modes in a compactant porous rock. Geophys. Res. Lett. 28:25212524.Google Scholar
Wong, T.-f., David, C. and Menéndez, B. (2004). Mechanical compaction. In Mechanics of Fluid-Saturated Rocks (eds. Guéguen, Y. and Boutéca, M.), pp. 55114, Elsevier, Amsterdam.Google Scholar
Worrall, D. M. and Snelson, S. (1989). Evolution of the northern Gulf of Mexico, with emphasis on Cenozoic growth faulting and the role of salt. In The Geology of North America, vol. A, The Geology of North America: an Overview (ed. Bally, A.W.), pp. 97137, Geol. Soc. Am., Boulder, CO.Google Scholar
Wu, J. E. and McClay, K. R. (2011). Two-dimensional analog modeling of fold and thrust belts: dynamic interactions with syncontractional sedimentation and erosion. In Thrust Fault-Related Folding (eds. McClay, K. R., Shaw, J. H. and Suppe, J.), pp. 301333, Amer. Assoc. Petrol. Geol. Mem. 94.Google Scholar
Wu, K. and Olson, J. E. (2015). Simultaneous multifracture treatments: fully coupled fluid flow and fracture mechanics for horizontal wells. Soc. Petrol. Eng. J. 20:337346.Google Scholar
Wyllie, D. C. and Norrish, N. I. (1996). Rock strength properties and their measurement. In Landslides: Investigation and Mitigation (eds. Turner, A. K. and Schuster, R. L.), pp. 372390, National Research Council, Transportation Research Board Special Report 247, National Academy Press, Washington, DC.Google Scholar
Xie, S. Y. and Shao, J. F. (2006). Elastoplastic deformation of a porous rock and water interaction. Int. J. Plasticity 22:21952225.Google Scholar
Xu, S.-S., Nieto-Samaniego, A. F., Alaniz-Álvarez, S. A. and Velasquillo-Martínez, L. G. (2005). Effect of sampling and linkage on fault length and length–displacement relationship. Int. J. Earth Sci. 95:841853.Google Scholar
Xue, Y. and Qu, J. (1999). Mixed-mode fracture mechanics parameters of elliptical interface cracks in anisotropic bimaterials. In Mixed-Mode Crack Behavior (eds. Miller, K. J., and McDowell, D. L.), pp. 143159, ASTM STP 1359, Am. Soc. Test. Mat., West Conshohocken, Penn.Google Scholar
Yang, Y., Sone, H., Hows, A. and Zoback, M. D. (2013). Comparison of brittleness indices in organic-rich shale formations. Paper ARMA 13–403 presented at the 47th US Symposium on Rock Mechanics/Geomechanics, San Francisco, California, 23–26 June 2013.Google Scholar
Yao, Y. (2012). Linear elastic and cohesive fracture analysis to model hydraulic fracture in brittle and ductile rocks. Rock Mech. Rock Eng. 45:375387; doi:10.1007/s00603-011-0211-0.Google Scholar
Yasuhara, H., Elsworth, D. and Polak, A. (2004). Evolution of permeability in a natural fracture: significant role of pressure solution. J. Geophys. Res. 109:3204; doi:10.1029/2003JB002663.Google Scholar
Yeats, R. S., Sieh, K. and Allen, C. R. (1997). The Geology of Earthquakes, Cambridge University Press, 568 pp.Google Scholar
Yerkes, R. F. and Castle, R. O. (1976). Seismicity and faulting attributable to fluid extraction. Eng. Geol. 10: 151167.Google Scholar
Yoffe, E. H. (1951). The moving Griffith crack. Philosoph. Mag. 42:739750.Google Scholar
Younes, A. I. and Engelder, T. (1999). Fringe cracks: key structures for the interpretation of the progressive Alleghanian deformation of the Appalachian Plateau. Geol. Soc. Am. Bull. 111:219239.Google Scholar
Zakharova, N. V. and Goldberg, D. S. (2014). In situ stress analysis in the northern Newark Basin: implications for induced seismicity from CO2 injection. J. Geophys. Res. 119; doi:10.1002/2013JB010492.Google Scholar
Zhang, D., Ranjith, P. G. and Perera, M. S. A. (2016). The brittleness indices used in rock mechanics and their application to shale hydraulic fracturing: a review. J. Petrol. Sci. Eng. 143:158170.Google Scholar
Zhang, J., Wong, T.-f. and Davis, D. M. (1990). Micromechanics of pressure-induced grain crushing in porous rocks. J. Geophys. Res. 95:341352.Google Scholar
Zhang, P., Burchfiel, B. C., Chen, S. and Deng, Q. (1989). Extinction of pull-apart basins. Geology 17:814817.Google Scholar
Zhao, G. and Johnson, A. M. (1991). Sequential and incremental formation of conjugate sets of faults. J. Struct. Geol. 13:887895.Google Scholar
Zhao, G. and Johnson, A. M. (1992). Sequence of deformation recorded in joints and faults, Arches National Park, Utah. J. Struct. Geol. 14:225236.Google Scholar
Zhao, X. G. and Cai, M. (2010). A mobilized dilation angle model for rocks. Int. J. Rock Mech. Min. Sci. 47:368384.Google Scholar
Zhou, X. and Aydin, A. (2010). Mechanics of pressure solution seam growth and evolution. J. Geophys. Res. 115:B12207; doi:12.1029/2010JB007614.Google Scholar
Zhou, X. and Aydin, A. (2012). Mechanics of the formation of orthogonal sets of solution seams, and solution seams and veins and parallel solution seams and veins. Tectonophysics 532535:242257.Google Scholar
Zhu, W. and Wong, T.-f. (1997). The transition from brittle faulting to cataclastic flow: permeability evolution. J. Geophys. Res. 102:30273041.Google Scholar
Zhu, X. K., Liu, G. T. and Chao, Y. J. (2001). Three-dimensional stress and displacement fields near an elliptical crack front. Int. J. Fracture 109:383401.Google Scholar
Ziv, A., Rubin, A. M. and Agnon, A. (2000). Stability of dike intrusion along preexisting fractures. J. Geophys. Res. 105:59475961.Google Scholar
Zoback, M. D. (2007). Reservoir Geomechanics, Cambridge University Press, New York.Google Scholar
Zoback, M. D. and Gorelick, S. M. (2012). Earthquake triggering and large-scale geologic storage of carbon dioxide. Proc. US Nat. Acad. Sci. 109:10,16410,168.Google Scholar
Zoback, M. D. and Healy, J. H. (1984). Friction, faulting, and “in situ” stress. Ann. Geophys. 2:689698.Google Scholar
Zoback, M. D., and Zinke, J. C. (2002). Production-induced normal faulting in the Valhall and Ekofisk oil fields. Pure Appl. Geophys. 159:403420.Google Scholar
Zoback, M. D., Zoback, M. L., Mount, V., et al. (1987). New evidence on the state of stress of the San Andreas fault system. Science 238:11051111.Google Scholar
Zoback, M. D., Apel, R., Baumgärtner, J., et al. (1993). Upper-crustal strength inferred from stress measurements to 6 km depth in the KTB borehole. Nature 365:633635.Google Scholar
Zoback, M. D., Townend, J. and Grollimund, B. (2002). Steady-state failure equilibrium and deformation of intraplate lithosphere. Inter. Geol. Rev. 44:383401.Google Scholar
Zoback, M. D., Barton, C. A., Brudy, M., et al. (2003). Determination of stress orientation and magnitude in deep wells. Int. J. Rock Mech. Min Sci. 40:10491076.Google Scholar
Zoback, M. D., Kohli, A., Das, I. and McClure, M. (2012). The importance of slow slip on faults during hydraulic fracturing stimulation of shale gas reservoirs. Paper SPE–155476 presented at the Americas Unconventional Resources Conference, Pittsburg, Pennsylvania, 5–7 June 2012.Google Scholar
Zoback, M. L. (1992). First and second order patterns of tectonic stress: the World Stress Map Project. J. Geophys. Res. 97:11,703–11,728.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

  • References
  • Richard A. Schultz
  • Book: Geologic Fracture Mechanics
  • Online publication: 03 August 2019
  • Chapter DOI: https://doi.org/10.1017/9781316996737.011
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
  • Richard A. Schultz
  • Book: Geologic Fracture Mechanics
  • Online publication: 03 August 2019
  • Chapter DOI: https://doi.org/10.1017/9781316996737.011
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
  • Richard A. Schultz
  • Book: Geologic Fracture Mechanics
  • Online publication: 03 August 2019
  • Chapter DOI: https://doi.org/10.1017/9781316996737.011
Available formats
×