Skip to main content Accessibility help
×
Hostname: page-component-848d4c4894-nmvwc Total loading time: 0 Render date: 2024-06-14T04:08:01.681Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  17 December 2018

Christopher H. Scholz
Affiliation:
Lamont-Doherty Earth Observatory, Columbia University
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

Abe, Y., and Kato, N. 2013. Complex earthquake cycle simulations using a two-degree-of-freedom spring-block model with a rate- and state-friction law. Pure Appl. Geophys 170(5): 745765, doi: 10.1007/s00024-011–0450-8.Google Scholar
Aben, F. M., Doan, M. L., Mitchell, T. M., et al. 2016. Dynamic fracturing by successive coseismic loadings leads to pulverization in active fault zones. J. Geophys. Res.-Solid Earth 121(4): 23382360, doi: 10.1002/2015jb012542.CrossRefGoogle Scholar
Abercrombie, R. E. 1995. Earthquake source scaling relationships from −1 to 5 M(L) using seismograms recorded at 2.5-km depth. J. Geophys. Res.-Solid Earth 100: 2401524036.Google Scholar
Abercrombie, R. E., Agnew, D. C., and Wyatt, F. K. 1995. Testing a model of earthquake nucleation. Bull. Seismol. Soc. Amer. 85: 18731878.Google Scholar
Abercrombie, R. E., and Ekström, G. 2001. Earthquake slip on oceanic transform faults. Nature 410: 7476.CrossRefGoogle ScholarPubMed
Abercrombie, R. E., and Ekström, G. 2003. A reassessment of the rupture characteristics of oceanic transform earthquakes. J. Geophys. Res.-Solid Earth 108(B5): doi: 10.1029/2001jb000814.Google Scholar
Abercrombie, R. E., and Mori, J. 1996. Occurrence patterns of foreshocks to large earthquakes in the western United States. Nature 381(6580): 303307.Google Scholar
Abercrombie, R. E., and Rice, J. R. 2005. Can observations of earthquake scaling constrain slip weakening? Geophys. J. Int. 162: 406424,Google Scholar
Abercrombie, R. E., Antolik, M., Felzer, K., and Ekström, G. 2001. The 1994 Java tsunami earthquake: Slip over a subduction seamount. J. Geophys. Res. 106: 65956608.CrossRefGoogle Scholar
Abers, G. A. 1991. Possible seismogenic shallow-dipping normal faults in the Woodlark-Dentrecasteaux Extensional Province, Papua-New-Guinea. Geology 19: 12051208.2.3.CO;2>CrossRefGoogle Scholar
Abers, G. A. 2005. Seismic low-velocity layer at the top of subducting slabs: Observations, predictions, and systematics. Phys. Earth Planet. Int. 149(1–2): 729, doi: 10.1016/j.pepi.2004.10.002.CrossRefGoogle Scholar
Acebron, J. A., Bonilla, L. L., Vincente, J. P., Ritort, F., and Spigler, R. 2005. The Kuramoto model: A simple paradigm for synchronization phenomena. Rev. Mod. Phys. 77: 137185.Google Scholar
Achenbach, J. D. 1973. Dynamic effects in brittle fracture. In Mechanics Today, New York: Pergamon, pp. 157.Google Scholar
Ackermann, R. V., and Schlische, R. W. 1997. Anticlustering of small normal faults around larger faults. Geology 25: 11271130.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(11): 18031819.Google Scholar
Adams, F. D. 1938. The Growth and Development of the Geological Sciences. Baltimore: Williams and Wilkins.Google Scholar
Adams, J., Wetmiller, R. J., Hasegawa, H. S., and Drysdale, J. 1991. The 1st surface faulting from a historical intraplate earthquake in North-America. Nature 352: 617619.Google Scholar
Aderhold, K., and Abercrombie, R. E. 2016a. The 2015 M-w 7.1 earthquake on the Charlie-Gibbs transform fault: Repeating earthquakes and multimodal slip on a slow oceanic transform. Geophys. Res. Lett. 43(12): 61196128, doi: 10.1002/2016gl068802.Google Scholar
Aderhold, K., and Abercrombie, R. E. 2016b. Seismotectonics of a diffuse plate boundary: Observations off the Sumatra-Andaman trench. J. Geophys. Res.-Solid Earth 121(5): 34623478, doi: 10.1002/2015jb012721.Google Scholar
Aggarwal, Y. P., Sykes, L. R., Simpson, D. W., and Richards, P. G. 1973. Spatial and temporal variations of tJtp and in P wave residuals at Blue Mountain Lake, New York: Application to earthquake prediction. J. Geophys. Res. 80: 718732.Google Scholar
Aguiar, A. C., Melbourne, T. I., and Scrivner, C. W. 2009. Moment release rate of Cascadia tremor constrained by GPS. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb005909.Google Scholar
Aharonov, E., and Scholz, C. H. 2018a. A physics-based friction constitutive law: Steady state frictionJ. Geophys. Res. 123: 15911614, doi: 10.1002/2016JB013829.Google Scholar
Aharonov, E., and Scholz, C. H. 2018b. The brittle-ductile transition predicted by a physics-based friction law, J. Geophys. Res. subm.Google Scholar
Aki, K. 1966. Generation and propagation of G waves from the Niigata earthquake of June 16, 1964, 2, Estimation of earthquake moment, released energy, and stress-strain drop from G wave spectrum. Bull. Earthquake Res. Inst., Univ. Tokyo 44: 7388.Google Scholar
Aki, K. 1967. Scaling law of seismic spectrum. J. Geophys. Res. 72: 12171231.CrossRefGoogle Scholar
Aki, K. 1979. Characterization of barriers on an earthquake fault. J. Geophys. Res. 84: 61406148.Google Scholar
Aki, K. 1981. A probabilistic synthesis of precursory phenomena. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 566574.Google Scholar
Aki, K. 1984. Asperities, barriers and characteristics of earthquakes. J. Geophys. Res. 89: 58675872.Google Scholar
Aki, K. 1985. Theory of earthquake prediction with special reference to monitoring of the quality factor of lithosphere by coda method. Earthquake Pred. Res. 3: 219230.Google Scholar
Aki, K. 1987. Magnitude frequency relation for small earthquakes: A clue for the origin of fmax of large earthquakes. J. Geophys. Res. 92: 13491355.Google Scholar
Aki, K., and Richards, P. 2002. Quantitative Seismology, 2nd edn., University Science Books.Google Scholar
Alessio, G., et al. 2010. Evidence for surface rupture associated with the Mw 6.3 L’Aquila earthquake sequence of April 2009 (central Italy). Terra Nova 22(1): 4351, doi: 10.1111/j.1365–3121.2009.00915.x.Google Scholar
Allegre, C. J., Le Mouel, J. L., and Provost, A. 1982. Scaling rules in rock fracture and possible implications for earthquake prediction. Nature 297: 4749.CrossRefGoogle Scholar
Allen, C. R. 1968. The tectonic environment of seismically active and inactive areas along the San Andreas fault system. In Proc. Conf. on Geological Problems of the San Andreas Fault System, ed. Kovach, R., Stanford Univ. Publ Geol. Sci., pp. 7082.Google Scholar
Allen, C. R. 1969. Active Faulting in Northern Turkey. Contr. No. 1577. Div. Geol. Sci. Calif. Inst. Tech. 32.Google Scholar
Allen, C. R., Wyss, M., Brune, J., Grantz, A., and Wallace, R. 1972. Displacements on the Imperial, Superstition Hills, and San Andreas faults triggered by the Borrego Mountain earthquake: The Borrego Mountain Earthquake of April 9, 1968. US Geol Surv. Prof. Paper 787: 87104.Google Scholar
Allen, R. M., and Kanamori, H. 2003. The potential for earthquake early warning in southern California. Science 300(5620): 786789, doi: 10.1126/science.1080912.Google Scholar
Allen, R. M., Gasparini, P. Kamigaichi, O., and Böse, M. 2009. The status of earthquake early warning around the world: An introductory overview. Seismol. Res. Lett. 80(5): 682693.Google Scholar
Allmann, B. P., and Shearer, P. M. 2009. Global variations of stress drop for moderate to large earthquakes. J. Geophys. Res.-Solid Earth 114(B1): doi: 10.1029/2008jb005821.Google Scholar
Alt, R. C., and Zoback, M. D. 2017. In Situ Stress and Active Faulting in Oklahoma. Bull. Seismol. Soc. Am. 107(1): 216228, doi: 10.1785/0120160156.Google Scholar
Amato, A., Azzara, R., Chiarabba, C., et al. 1998. The 1997 Unbria-Marche, Italy, earthquake sequence: A first look at the mainshocks and aftershocks. Geophys. Res. Lett. 25: 28612864.Google Scholar
Ambraseys, N., and Finkel, C. 1988. The Anatolian earthquake of 17 August 1668. In Historic Seismograms and Earthquakes of the World, eds. Lee, H. M. W. H. K., and Shimazaki, K., Academic, San Diego, Calif., pp. 173180.Google Scholar
Ambrayseys, N. H. 1970. Some characteristic features of the Anatolian fault zone. Tectonophysics 9: 143165.Google Scholar
Ambrayseys, N. H. 1975. Studies of historical seismicity and tectonics. In Geodynamics Today, A Review of the Earth’s Dynamic Processes, London: The Royal Society, pp. 716.Google Scholar
Amelung, F., and King, G. C. P. 1997. Earthquake scaling laws for creeping and non-creeping faults. Geophys. Res. Lett. 24: 507510.CrossRefGoogle Scholar
Amitrano, D. 2003. Brittle-ductile transition and associated seismicity: Experimental and numerical studies and relationship with the b value. J. Geophys. Res. 108(B1): 2044, doi: 10.1029/2001JB000680.Google Scholar
Amitrano, D., and Helmstetter, A. 2006. Brittle creep, damage, and time to failure in rocks. J. Geophys. Res.-Solid Earth 111 (B11): 117, doi: 10.1029/2005jb004252.Google Scholar
Ammon, C. J., Ji, C., Thio, H.-K., Robinson, D., Ni, S., Hjorleifsdottir, V., Kanamori, H., Lay, T., Das, S.,and Helmberger, D. 2005. Rupture process of the 2004 Sumatra-Andaman earthquake. Science 308(5725): 11331139.Google Scholar
Ammon, C. J., Kanamori, H., and Lay, T. 2008. A great earthquake doublet and seismic stress transfer cycle in the central Kuril islands. Nature 451(7178): 561564, doi: 10.1038/nature06521.Google Scholar
Ammon, C. J., Kanamori, H., Lay, T., and Velasco, A. A. 2006. The 17 July 2006 Java tsunami earthquake. Geophys. Res. Lett. 33(24): L24308.Google Scholar
Ampuero, J. P., and Ben-Zion, Y. 2008. Cracks, pulses and macroscopic asymmetry of dynamic rupture on a bimaterial interface with velocity-weakening friction. Geophys. J. Int. 173(2): 674692, doi: 10.1111/j.1365-246X.2008.03736.x.Google Scholar
Ampuero, J.-P., and Rubin, A. M. 2008. Earthquake nucleation on rate and state faults: Aging and slip laws. J. Geophys. Res. 113, doi: 10.1029/2007JB005082.Google Scholar
Anandakrishnan, S., Voigt, D. E., Alley, R. B. and King, M. A. 2003. Ice stream D flow speed is strongly modulated by the tide beneath the Ross Ice Shelf. Geophys. Res. Lett. 30(7): 1361, doi: 10.1029/2002gl016329.CrossRefGoogle Scholar
Anderlini, L., Serpelloni, E., and Belardinelli, M. E. 2016. Creep and locking of a low-angle normal fault: Insights from the Altotiberina fault in the Northern Apennines (Italy). Geophys. Res. Lett. 43(9): 43214329, doi: 10.1002/2016gl068604.Google Scholar
Anders, M. H., and Schlische, R. W. 1994. Overlapping faults, intrabasin highs, and the growth of normal faults. J. Geol. 102(2): 165180.CrossRefGoogle Scholar
Anders, M. H., and Wiltschko, D. V. 1994. Microfracturing, paleostress and the growth of faults. J. Struct. Geol. 16: 795815.Google Scholar
Anders, M. H., Laubach, S. E., and Scholz, C. H. 2014. Microfractures: A review. J. Struct. Geol. 69: 377394, doi: 10.1016/j.jsg.2014.05.011.Google Scholar
Anders, M. H., Spiegelman, M., Rodgers, D. W., and Hagstrum, J. T. 1993. The growth of fault-bounded tilt blocks. Tectonics 12(6): 14511459.Google Scholar
Anderson, E. M. 1905. The dynamics of faulting. Trans. Edinburgh Geol. Soc. 8: 387–340.Google Scholar
Anderson, E. M. 1936. The dynamics of the formation of cone-sheets, ring-dykes, and cauldron-subsidences. Proc. Roy. Soc. Edinburgh 56: 128–125.Google Scholar
Anderson, E. M. 1951. The Dynamics of Faulting. Edinburgh: Oliver and Boyd.Google Scholar
Anderson, J. G., and Bodin, P. 1987. Earthquake recurrence models and historical seismicity in the Mexicali-Imperial Valley. Bull. Seismol. Soc. Am. 77: 562578.Google Scholar
Anderson, J. G., Brune, J. N., Louie, J. N., Zeng, Y., Savage, M., Yu, G., Chen, Q., and dePolo, D. 1994. Seismicity in the western Great Basin apparently triggered by the Landers, California earthquake, 28 June, 1992. Bull. Seismol. Soc. Amer. 84: 863891.Google Scholar
Anderson, J. G., Wesnousky, S. G., and Stirling, M. W. 1996. Earthquake size as a function of fault slip rate. Bull. Seismol. Soc. Am. 86(3): 683690.CrossRefGoogle Scholar
Anderson, J. L., Osborne, R., and Palmer, D. 1983. Cataclastic rocks of the San Gabriel fault zone and expression of deformation at deeper crustal levels in the San Andreas fault zone. Tectonophysics 98: 209251.Google Scholar
Anderson, O. L., and Grew, P. C. 1977. Stress-corrosion theory of crack-propagation with applications to geophysics. Rev. Geophys. 15(1): 77104, doi: 10.1029/RG015i001p00077.CrossRefGoogle Scholar
Anderson, R. N., Langseth, M. G., and Sclater, J. G. 1977. The mechanisms of heat transfer through the floor of the Indian Ocean. J. Geophys. Res. 82: 33913409.Google Scholar
Ando, M. 1974. Seismotectonics of the 1923 Kanto earthquake. J. Phys. Earth 22: 263277.Google Scholar
Ando, M. 1975. Source mechanisms and tectonic significance of historical earthquakes along the Nankai trough. Tectonophysics 27: 119140.Google Scholar
Ando, R., and Imanishi, K. 2011. Possibility of M-w 9.0 mainshock triggered by diffusional propagation of after-slip from M-w 7.3 foreshock. Earth Planets and Space, 63(7): 767771, doi: 10.5047/eps.2011.05.016.Google Scholar
Ando, R., Nakata, R., and Hori, T. 2010. A slip pulse model with fault heterogeneity for low-frequency earthquakes and tremor along plate interfaces. Geophys. Res. Lett. 37, doi: 10.1029/2010gl043056.Google Scholar
Andrews, D. 1976a. Rupture propagation with finite stress in antiplane strain. J. Geophys. Res. 81: 35753582.CrossRefGoogle Scholar
Andrews, D. 1976b. Rupture velocity of plane strain shear cracks. J. Geophys. Res. 81: 56795687.Google Scholar
Andrews, D. 1980. A stochastic fault model, static case. J. Geophys. Res. 85: 38673887.CrossRefGoogle Scholar
Andrews, D. 1985. Dynamic plane strain shear rupture with a slip-weakening friction law calculated by a boundary integral method. Bull. Seismol. Soc. Am. 75: 121.Google Scholar
Andrews, D. J. 2005. Rupture dynamics with energy loss outside the slip zone, J. Geophys. Res. 110: 114, doi: 10.1029/2004JB003191Google Scholar
Andrews, D. J., and Ben-Zion, Y. 1997. Wrinkle-like slip pulse on a fault between different materials. J. Geophys. Res.-Solid Earth 102: 553571.Google Scholar
Andrews, D. J., and Harris, R. A. 2005. The wrinkle-like slip pulse is not important in earthquake dynamics, Geophys. Res. Lett. 32: L23303, doi: 10.1029/2005GL023996.Google Scholar
Angellier, J. 1984. Tectonic analysis of fault slip data sets. J. Geophys. Res. 89: 58355848.CrossRefGoogle Scholar
Antolik, M., Abercrombie, R. E., Pan, J., and Ekström, G. 2006. Rupture characteristics of the 2003 Mw 7.6 mid‐Indian Ocean earthquake: Implications for seismic properties of young oceanic lithosphere. J. Geophys. Res.-Solid Earth 111(B4).Google Scholar
Antolik, M., Dreger, D., and Romanowicz, B. 1999. Rupture processes of large deep-focus earthquakes from inversion of moment rate functions. J. Geophys. Res.-Solid Earth 104(B1): 863894, doi: 10.1029/1998jb900042.Google Scholar
Antonellini, M. A., Aydin, A., and Pollard, D. D. 1994. Microstructure of deformation bands in porous sandstone at Arches National Park, Utah. J. Struct. Geol. 16(7): 941959, doi: 10.1016/0191–8141(94)90077–9.Google Scholar
Aoki, Y., and Scholz, C. H. 2003. Interseismic deformation at the Nankai subduction zone and the Median Tectonic Line, southwest Japan. J. Geophys. Res.-Solid Earth 108(B10).Google Scholar
Archard, J. F. 1953. Contact and rubbing of flat surfaces. J. App. Phy. 24: 981988.Google Scholar
Archard, J. F. 1957. Elastic deformation and the laws of friction. Proc. R. Soc. London Ser. A 243: 190205.Google Scholar
Arnadottir, T., and Segal, P. 1994. The 1989 Loma-Prieta earthquake imaged from inversion of geodetic data. J. Geophys. Res.-Solid Earth 99: 2183521855.Google Scholar
Asano, Y., Saito, T., Ito, Y., Shiomi, K., Hirose, H., Matsumoto, T., Aoi, S., Hori, S., and Sekiguchi, S. 2011. Spatial distribution and focal mechanisms of aftershocks of the 2011 off the Pacific coast of Tohoku Earthquake. Earth Planets and Space, 63(7): 669673, doi: 10.5047/eps.2011.06.016.Google Scholar
Ashby, M. F., and Sammis, C. G. 1990. The damage mechanics of brittle solids in compression. Pure Appl. Geophys. 133(3): 489521, doi: 10.1007/bf00878002.Google Scholar
Astiz, L., and Kanamori, H. 1986. Interplate coupling and temporal variation of mechanisms of intermediate-depth earthquakes in Chile. Bull. Seismol. Soc. Am. 76(6): 16141622.Google Scholar
Atkinson, B. K. 1980. Stress corrosion and the rate-dependent tensile failure of a fine-grained quartz rock. TTectonophysics 65: 281290.Google Scholar
Atkinson, B. K. 1984. Subcritical crack growth in geological materials. J. Geophys. Res. 89: 40774114.Google Scholar
Atkinson, B. K. 1987. Introduction to fracture mechanics and its geophysical applications. In Fracture Mechanics of Rock, ed. Atkinson, B. K.. London: Academic Press, pp. 126.Google Scholar
Atkinson, B. K., and Meredith, P. G. 1987. Experimental fracture mechanics data for rocks and minerals. In Fracture Mechanics of Rock, ed. Atkinson, B. K., Academic Press, London, pp. 477525.Google Scholar
Atwater, B. F. 1990. Evidence for great Holocene earthquakes along the outer coast of Washington State. Science 236: 942944.Google Scholar
Atwater, T. 1970. Implications of plate tectonics for the Cenozoic tectonic evolution of western North America. Geol. Soc. Am. Bull. 81: 35133536.Google Scholar
Audet, P., Bistock, M. G., Boyarko, D. C., Brudzinski, M. R., and Allen, R. M. 2010. Slab morphology in the Cascadia forearc and its relation to episodic tremor and slip. J. Geophys. Res. 115, b00A16: doi: 10.1029/2008JB006053.Google Scholar
Audet, P., Bostock, M. G., Christensen, N. I., and Peacock, S. M. 2009. Seismic evidence for overpressured subducted oceanic crust and megathrust fault sealing. Nature 457(7225): 7678, doi: 10.1038/nature07650.Google Scholar
Auzende, J. M., Bideau, D., Bonatti, E., et al. 1989. Direct observation of a section through slow-spreading oceanic-crust. Nature 337: 726729.Google Scholar
Aviles, C. A., Scholz, C. H., and Boatwright, J. 1987. Fractal analyis applied to characteristic segments of the San Andreas fault. J. Geophys. Res.-Solid Earth and Planets 92(B1): 331344, doi: 10.1029/JB092iB01p00331.Google Scholar
Avouac, J. P. 2015. From geodetic imaging of seismic and aseismic fault slip to dynamic modeling of the seismic cycle. In Ann. Rev. Earth Planet. Sci, Vol 43, eds. Jeanloz, R. and Freeman, K. H., pp. 233271, doi: 10.1146/annurev-earth-060614–105302.Google Scholar
Axen, G. J. 1992. Pore pressure, stress increase, and fault weakening in low-angle normal faulting. J. Geophys. Res.-Solid Earth 97: 89798991.Google Scholar
Aydin, A. 1978. Small faults formed as deformation bands in sandstone. Pure Appl. Geophys. 116(4–5): 913930, doi: 10.1007/bf00876546.Google Scholar
Aydin, A., and Johnson, A. M. 1978. Development of faults as zones of deformation bands and as slip surfaces in sandstone. Adv. Appl. Mech. 116: 922929.Google Scholar
Aydin, A., and Johnson, A. M. 1983. Analysis of faulting in porous sandstones. J. Struct. Geol. 5(1): 1931 doi: 10.1016/0191–8141(83)90004–4.Google Scholar
Aydin, A., and Nur, A. 1982. Evolution of pull-apart basins and their scale independence. Tectonics 1: 91105.Google Scholar
Bai, T. X., and Pollard, D. D. 2000a. Fracture spacing in layered rocks: A new explanation based on the stress transition. J. Struct. Geol. 22(1): 4357, doi: 10.1016/s0191-8141(99)00137–6.Google Scholar
Bai, T. X., Pollard, D. D., and Gao, H. J. 2000b. Spacing of edge fractures in layered materials. Int. J. Fract. 103(4): 373395, doi: 10.1023/a:1007659406011.Google Scholar
Baisch, S., Vörös, R., Rothert, E., Stang, H., Jung, R., and Schellschmidt, R. 2010. A numerical model for fluid injection induced seismicity at Soultz-sous-Forêts. Int. J. Rock Mech. Min. Sci. 47(3): 405413.Google Scholar
Bak, J., Korstgard, J. and Sorensen, K. 1975. Major shear zone within Nagssugtoqidian of West Greenland. Tectonophysics 27(3): 191209.Google Scholar
Bak, P., and Tang, C. 1989. Earthquakes as a self-organized critical phenomenon. J. Geophys. Res. 94: 1563515637.Google Scholar
Bak, P., Tang, C., and Weisenfeld, K. 1987. Self-organized criticality: An explanation of l/f noise. Phys. Rev. Lett. 59: 381384.Google Scholar
Bak, P., Tang, C., and Wiesenfeld, K. 1988. Self-organized criticality. Phys. Rev. A 38: 364374.Google Scholar
Baker, B. H., and Wohlenberg, J. 1971. Structure and evolution of the Kenya rift valley. Nature 229: 538.Google Scholar
Bakun, W. H., and Lindh, A. 1985. The Parkfield, California, earthquake prediction experiment. Science 229: 619624.Google Scholar
Bakun, W. H., and McEvilly, T. V. 1984. Recurrence models and Parkfield, California, earthquakes. J. Geophys. Res. 89: 30513058.Google Scholar
Bakun, W. H., King, G. C. P., and Cockerham, R. S. 1986. Seismic slip, aseismic slip, and the mechanics of repeating earthquakes on the Calaveras fault, California. In Earthquake Source Mechanics. Geophys, AGU. Mono. 37, eds. Das, S., Scholz, C., and Boatwright, J.. Washington, DC: American Geophysical Union, pp. 195207.Google Scholar
Bakun, W. H., Stewart, R. M., Bufe, C. G., and Marks, S. M. 1980. Implication of seismicity for failure of a section of the San Andreas fault. Bull. Seismol. Soc. Amer. 70: 185201.Google Scholar
Bakun, W., Aagaard, B., Dost, B., et al. 2005. Implications for prediction and hazard assessment from the 2004 Parkfield earthquake. Nature 437(7061): 969974.Google Scholar
Baltay, A. S., Hanks, T. C., and Beroza, G. C. 2013. Stable Stress-Drop Measurements and their Variability: Implications for Ground-Motion Prediction. Bull. Seismol. Soc. Amer. 103: 211222, doi: 10.1785/0120120161.Google Scholar
Banerjee, P., Pollitz, F. F., and Bürgmann, R. 2005. The size and duration of the Sumatra-Andaman earthquake from far-field static offsets. Science 308(5729): 17691772.Google Scholar
Barbot, S., Lapusta, N., and Avouac, J.-P. 2012. Under the hood of the earthquake machine: Toward predictive modeling of the seismic cycle. Science 336(6082): 707710.CrossRefGoogle ScholarPubMed
Barenblatt, G. I. 1962. The mathematical theory of equilibrium cracks in brittle fracture. Adv. Appl. Mech. 7: 5580.Google Scholar
Barka, A., and Kadinsky-Cade, K. 1988. Strike-slip fault geometry in Turkey and its influence on earthquake activity. Tectonics 7: 663684.Google Scholar
Barnett, D. E., Bowman, J. R., Pavlis, T. L., Rubenstone, J. R., Snee, L. W., and Onstott, T. C. 1994. Metamorphism and near-trench plutonism during initial accretion of the cretaceous Alaskan fore-arc. J. Geophys. Res.-Solid Earth 99: 2400724024.Google Scholar
Barnett, J. A., Mortimer, J., Rippon, J. H., Walsh, J. J., and Watterson, J. 1987. Displacement geometry in the volume containing a single normal fault. Am. Assoc. Pet. Geol. Bull. 71: 925937.Google Scholar
Barnhart, W. D. 2017. Fault creep rates of the Chaman fault (Afghanistan and Pakistan) inferred from InSAR. J. Geophys. Res.-Solid Earth, 122: 372386, doi: 10.1002/2016JB013656.Google Scholar
Barr, T. D., and Dahlen, F. A. 1989. Brittle frictional mountain building 2. Thermal structure and heat budget. J. Geophys. Res. 94: 39233947.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
Bassett, D., Sutherland, R., and Henrys, S. 2014. Slow wavespeeds and fluid overpressure in a region of shallow geodetic locking and slow slip, Hikurangi subduction margin. Earth Planet. Sci. Lett. 389: 113.Google Scholar
Baud, P., and Meredith, P. G. 1997. Damage accumulation during triaxial creep of Darley Dale sandstone from pore volumometry and acoustic emission. Int. J. Rock Mech. Min. Sci. 34: 34.Google Scholar
Baumberger, T., and Caroli, C. 2006. Solid friction from stick-slip down to pinning and aging. Adv. Phys. 55(3–4): 279348, doi: 10.1080/00018730600732186.Google Scholar
Baumberger, T., Berthoud, P., and Caroli, C. 1999. Physical analysis of the state- and rate-dependent friction law. II. Dynamic friction. Phys. Rev. B-Condens Matter 60(6): 39283939.Google Scholar
Beach, A., Welbon, A. I., Brockbank, P. J., and McCallum, J. E. 1999. Reservoir damage around faults: Outcrop examples from the Suez rift. Petrol. Geosci. 5(2): 109116, doi: 10.1144/petgeo.5.2.109.Google Scholar
Beavan, J., Bilham, R., and Hurst, K. 1984. Coherent tilt signals observed in the Shumagin seismic gap: Detection of time-dependent subduction at depth? J. Geophys. Res. 89: 44784492.Google Scholar
Beavan, J., Wang, X., Holden, C., Wilson, K., Power, W., Prasetya, G., Bevis, M., and Kautoke, R. 2010. Near-simultaneous great earthquakes at Tongan megathrust and outer rise in September 2009. Nature 466(7309): 959978, doi: 10.1038/nature09292.Google Scholar
Bécel, A., Delescluse, M., NedimovićM. R., et al. 2017. Tsunamigenic structures in a creeping section of the Alaska subduction zone. Nat. Geosci. 10: 609613, doi: 10.1038/ngeo2990.Google 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.-Solid Earth 108(B8): doi: 10.1029/2001jb001518.Google Scholar
Beeler, N. M., Hickman, S. H., and Wong, T. F. 2001. Earthquake stress drop and laboratory-inferred interseismic strength recovery. J. Geophys. Res. 106: 3070130713Google Scholar
Beeler, N. M., Hirth, G., Thomas, A., and Buergmann, R. 2016. Effective stress, friction, and deep crustal faulting. J. Geophys. Res.-Solid Earth 121(2): 10401059, doi: 10.1002/2015jb012115.Google Scholar
Beeler, N. M., Lockner, D. L., and Hickman, S. H. 2001. A simple stick-slip and creep-slip model for repeating earthquakes and its implication for microearthquakes at Parkfield. Bull. Seismol. Soc. Am. 91(6): 17971804, doi: 10.1785/0120000096.Google Scholar
Beeler, N. M., Tullis, T. E., and Goldsby, D. L. 2008. Constitutive relationships and physical basis of fault strength due to flash heating. J. Geophys. Res.-Solid Earth 113(B1): doi: 10.1029/2007jb004988.Google Scholar
Beeler, N. M., Tullis, T. E., and Weeks, J. D. 1994. The roles of time and displacement in the evolution effect in rock friction. Geophys. Res. Lett. 21: 19871990.Google Scholar
Behr, W. M., and Platt, J. P. 2011. A naturally constrained stress profile through the middle crust in an extensional terrane. Earth Planet. Sci. Lett. 303(3–4): 181192, doi: 10.1016/j.epsl.2010.11.044.Google Scholar
Belardinelli, M. E., Bonafede, M., and Gudmundsson, A. 2000. Secondary earthquake fractures generated by a strike-slip fault in the South Iceland Seismic Zone. J. Geophys. Res. 105: 1361313630.Google Scholar
Belardinelli, M. E., Cocco, M., Coutant, O., and Cotton, F. 1999. Redistribution of dynamic stress during coseismic ruptures: Evidence for fault interaction and earthquake triggering. J. Geophys. Res.-Solid Earth 104: 1492514945.Google Scholar
Bell, J. W., Caskey, S. J., Ramelli, A. R., and Guerrieri, l. 2004. Pattern and rates of faulting in the central Nevada seismic belt, and paleoseismic evidence for prior beltlike behavior. Bull. Seismol. Soc. Amer. 94: 12291254.Google Scholar
Bell, M. L., and Nur, A. 1978. Strength changes due to reservoir induced pore pressure and stresses and application to Lake Oroville. J. Geophys. Res. 83: 44694483.Google Scholar
Bell, T. H., and Etheridge, M. A. 1973. Microstructures of mylonites and their descriptive terminology. Lithos 6: 337348.Google Scholar
Ben-David, O., and Fineberg, J. 2011. Static friction coefficient is not a material constant. Phys. Rev. Lett. 106(25): doi: 10.1103/PhysRevLett.106.254301.Google Scholar
Ben-David, O., Cohen, G., and Fineberg, J. 2010a. The dynamics of the onset of frictional slip. Science 330(6001): 211214, doi: 10.1126/science.1194777.Google Scholar
Ben-David, O., Rubinstein, S. M., and Fineberg, J. 2010b. Slip-stick and the evolution of frictional strength. Nature 463(7277): 7679, doi: 10.1038/nature08676.Google Scholar
Bennett, M., Schatz, M. F., Rockwood, H., and Wiesenfeld, K. 2002. Huygens’s clocks. P. Roy. Soc. A-Math. Phy, 458(2019), 563579, doi: 10.1098/rspa.2001.0888.Google Scholar
Bennett, R. A., Davis, J. L., and Wernicke, B. P. 1999. Present-day pattern of Cordilleran deformation in the western United States. Geology 27: 371374.Google Scholar
Bennett, R. A., Rodi, W., and Reilinger, R. 1996. Global Positioning System constraints on fault slip rates in southern California and northern Baja, Mexico. J. Geophys. Res. 101: 21, 943–21, 960.Google Scholar
Ben-Zion, Y., and Rice, J. R. 1993. Earthquake failure sequences along a cellular fault zone in a 3-dimensional elastic solid containing asperity and non-asperity regions. J. Geophys. Res.-Solid Earth 98(B8): 1410914131, doi: 10.1029/93jb01096.Google Scholar
Ben-Zion, Y., and Rice, J. R. 1995. Slip patterns and earthquake populations along different classes of faults in elastic solids. J. Geophys. Res.-Solid Earth 100: 1295912983.Google Scholar
Ben-Zion, Y., and Shi, Z. Q. 2005. Dynamic rupture on a material interface with spontaneous generation of plastic strain in the bulk. Earth Planet. Sci. Lett. 236(1–2): 486496, doi: 10.1016/j.epsl.2005.03.025.Google Scholar
Ben-Zion, Y., Rice, J. R., and Dmowska, R. 1993. Interaction of the San Andreas fault creeping segment with adjacent great rupture zones and earthquake recurrence at Parkfield. J. Geophys. Res. 98: 21342144.Google Scholar
Bergman, E. A. 1986. Intraplate earthquakes and the state of stress in oceanic lithosphere. Tectonophysics 132: 135.Google Scholar
Bergman, E. A., and Solomon, S. 1988. Transform fault earthquakes in the North Atlantic: Source mechanism and depth of faulting. J. Geophys. Res. 93: 90279057.Google Scholar
Bergman, E. A., and Solomon, S. C. 1980. Oceanic intraplate earthquakes – Implications for local and regional intraplate stress. J. Geophys. Res. 85(NB10): 53895410, doi: 10.1029/JB085iB10p05389.Google Scholar
Berkhemer, H., Zschau, J., and Ergunay, O. 1988. The German-Turkish project on earthquake prediction research, concept and first results. Paper presented at Proceedings of the ECE/UN Seminar on Prediction of Earthquakes, Lisbon.Google Scholar
Beroza, G. C. 1991. Near-source modeling of the Loma-Prieta earthquake – Evidence for heterogeneous slip and implications for earthquake hazard. Bull. Seismol. Soc. Amer. 81: 16031621.Google Scholar
Beroza, G. C., and Ide, S. 2011. Slow Earthquakes and Nonvolcanic Tremor. In Ann. Rev. Earth Planet. Sci, Vol 39, eds. R. Jeanloz and K. H. Freeman, pp. 271296, doi: 10.1146/annurev-earth-040809–152531.Google Scholar
Beroza, G. C., and Mikumo, T. 1996. Short slip duration in dynamic rupture in the presence of heterogeneous fault properties. J. Geophys. Res.-Solid Earth 101(B10): 2244922460, doi: 10.1029/96jb02291.Google Scholar
Berthaud, P., Baumberger, T., G’sell, C., and Hiver, J. M. 1999. Physical analysis of the state and rate-dependent friction law: Static friction. Phys. Rev. B. 59: 14313.Google Scholar
Berthe, D., Choukroune, P., and Jegouzo, P. 1979. Orthogneiss, mylonite and non-coaxial deformation of granites: The example of the South Armorican shear zone. J. Struct. Geol. 1: 3142.Google Scholar
Bhat, H. S., Dmowska, R., Rice, J. R., and Kame, N. 2004. Dynamic slip transfer from the Denali to Totschunda faults, Alaska: Testing theory for fault branching. Bull. Seismol. Soc. Am. 94(6): S202S213, doi: 10.1785/0120040601.Google Scholar
Bhattacharya, P., Rubin, A. M., Bayart, E., Savage, H. M., and Marone, C. 2015. Critical evaluation of state evolution laws in rate and state friction: Fitting large velocity steps in simulated fault gouge with time-, slip-, and stress-dependent constitutive laws. J. Geophys. Res.-Solid Earth 120(9): 63656385, doi: 10.1002/2015jb012437.Google Scholar
Biasi, G. P., and Wesnousky, S. 2017. Bends and ends of surface ruptures. Bull. Seismol. Soc. Amer. 107: 25432560, doi: 10.1785/0120160292.Google Scholar
Biasi, G. P., and Wesnousky, S. G. 2016. Steps and gaps in ground ruptures: Empirical bounds on rupture propagation. Bull. Seismol. Soc. Am. 106(3): 11101124, doi: 10.1785/0120150175.Google Scholar
Biegel, R. L., Wang, W., Scholz, C. H., Boitnott, G. N., and Yoshioka, N. 1992. Micromechanics of rock friction. 1. Effects of surface-roughness on initial friction and slip hardening in westerly granite. J. Geophys. Res.-Solid Earth 97: 89518964.Google Scholar
Bilek, S. L. 2010. The role of subduction erosion on seismicity. Geology 38(5): 479480, doi: 10.1130/focus052010.1.Google Scholar
Bilek, S. L., and Engdahl, E. R. 2007. Rupture characterization and aftershock relocations for the 1994 and 2006 tsunami earthquakes in the Java subduction zone. Geophys. Res. Lett. 34(20): doi: 10.1029/2007gl031357.Google Scholar
Bilek, S. L., and Lay, T. 2002. Tsunami earthquakes possibly widespread manifestations of frictional conditional stability. Geophys. Res. Lett. 29(14): doi: 10.1029/2002gl015215.Google Scholar
Bilek, S. L., Engdahl, E. R., DeShon, H. R., and El Hariri, M. 2011. The 25 October 2010 Sumatra tsunami earthquake: Slip in a slow patch. Geophys. Res. Lett. 38, doi: 10.1029/2011gl047864.Google Scholar
Bilham, R. G., and Beavan, J. 1978. Tilts and strains on crustal blocks. Tectonophysics 52: 121138.Google Scholar
Bilham, R. G., and Williams, P. 1985. Sawtooth segmentation and deformation processes on the southern San Andreas fault, California. Geophys. Res. Lett. 12: 557560.Google Scholar
Bilham, R., and Yu, T. T. 2000. The morphology of thrust faulting in the 21 September 1999, Chichi, Taiwan earthquake. J. Asian Earth Sci. 18(3): 351367, doi: 10.1016/s1367-9120(99)00071–1.Google Scholar
Bilham, R., Engdahl, R., Feldl, N., and Satyabala, S. P. 2005. Partial and complete rupture of the Indo-Andaman plate boundary 1847–2004. Seismol. Res. Lett. 76(3): 299311.Google Scholar
Bilham, R., Ozener, H., Mencin, D., et al. 2016. Surface creep on the North Anatolian Fault at Ismetpasa, Turkey, 1944–2016, J. Geophys. Res., 121, doi: 10.1002/2016JB013394.Google Scholar
Bird, P., Kagan, Y. Y., and Jackson, D. D. 2002. Plate tectonics and earthquake potential of spreading ridges and oceanic transform faults, Wiley Online Library: https://doi.org/10.1029/GD030p0203.Google Scholar
Bizzarri, A. 2010. Pulse-like dynamic earthquake rupture propagation under rate-, state- and temperature-dependent friction. Geophys. Res. Lett. 37, doi: 10.1029/2010gl044541.Google Scholar
Bjarnason, I. T., Cowie, P., Anders, M. H., Seeber, L., and Scholz, C. H. 1993. The 1912 Iceland earthquake rupture – Growth and development of a nascent transform system. Bull. Seismol. Soc. Amer. 83: 416435.Google Scholar
Blanpied, M. L., Lockner, D. A., and Byerlee, J. D. 1992. An earthquake mechanism based on rapid sealing of faults. Nature 358: 574–6.Google Scholar
Blanpied, M. L., Lockner, D. A., and Byerlee, J. D. 1995. Frictional slip of granite at hydrothermal conditions. J. Geophys. Res.-Solid Earth 100: 1304513064.Google Scholar
Blanpied, M. L., Marone, C. J., Lockner, D. A., Byerlee, J. D., and King, D. P. 1998. Quantitative measure of the variation in fault rheology due to fluid-rock interactions. J. Geophys. Res. 103: 96919712.Google Scholar
Blenkinsop, T. G. 1989. Thickness – Displacement relationships for deformation zones: Discussion, J. Struct. Geol., 8: 10511054.Google Scholar
Boatwright, J. 1980. A spectral theory for circular seismic sources: Simple estimates of source dimension, dynamic stress-drop and radiated seismic energy. Bull Seismol. Soc. Am. 70: 127.Google Scholar
Boatwright, J. 1985. Characteristics of the aftershock sequence of the Borah Peak, Idaho, earthquake determined from digital recordings of the events. Bull. Seismol. Soc. Am. 75: 12651284.Google Scholar
Boatwright, J., and Choy, G. L. 1986. Teleseismic estimates of the energy radiated by shallow earthquakes. J. Geophys. Res.-Solid Earth and Planets 91(B2): 20952112, doi: 10.1029/JB091iB02p02095.Google Scholar
Bodin, P., and Brune, J. N. 1996. On the scaling of slip with rupture length for shallow strike-slip earthquakes: Quasi-static models and dynamic rupture propagation. Bull. Seismol. Soc. Amer. 86: 12921299.Google Scholar
Bodin, P., Bilham, R., Behr, J., Gomberg, J., and Hudnut, K. W. 1994. Slip triggered on southern California faults by the 1992 Joshua-Tree, Landers, and Big-Bear earthquakes. Bull. Seismol. Soc. Amer. 84: 806816.Google Scholar
Boettcher, M. S., and Jordan, T. H. 2004. Earthquake scaling relations for mid-ocean ridge transform faults. J. Geophys. Res.-Solid Earth 109(B12), doi: 10.1029/2004jb003110.Google Scholar
Boettcher, M. S., and McGuire, J. J. 2009. Scaling relations for seismic cycles on mid-ocean ridge transform faults. Geophys. Res. Lett. 36, doi: 10.1029/2009gl040115.Google Scholar
Boettcher, M. S., Hirth, G., and Evans, B. 2007. Olivine friction at the base of oceanic seismogenic zones. J. Geophys. Res.-Solid Earth 112(B1): doi: 10.1029/2006jb004301.Google Scholar
Bohnenstiehl, D. R., Tolstoy, M., and Chapp, E. 2004. Breaking into the plate: A 7.6 Mw fracture-zone earthquake adjacent to the Central Indian Ridge. Geophys. Res. Lett. 31(2): doi: 10.1029/2003gl018981.Google Scholar
Boitnott, G. N., Biegel, R. L., Scholz, C. H., Yoshioka, N., and Wang, W. 1992. Micromechanics of rock friction 2. Quantitative modeling of initial friction with contact theory. J. Geophys. Res.-Solid Earth 97: 89658978.Google Scholar
Boland, J. N., and Tullis, T. E. 1986. Deformation behavior of wet and dry clinopyroxenite in the brittle to ductile transition region. In Mineral and Rock Deformation: Laboratory Studies AGU Geophys. Mono, eds. Hobbs, B. E. and Heard, H. C.. Washington, DC: American Geophysical Union, pp. 3550.Google Scholar
Bolt, B. A. 1978. Earthquakes: A Primer. San Francisco: Freeman.Google Scholar
Boncio, P., Pizzi, A., Brozzetti, F., Pomposo, G., Lavecchia, G., Di Naccio, D., and Ferrarini, F. 2010. Coseismic ground deformation of the 6 April 2009 L’Aquila earthquake. central Italy, M(w)6.3, Geophys. Res. Lett., 37, doi: 10.1029/2010gl042807.Google Scholar
Boneh, Y, and Reches, Z. 2017. Geotribology – Friction, wear, and lubrication of faults. Tectonophysics 733: 171181: doi: 10.1016/jtecto.2017.11.022.Google Scholar
Boneh, Y., Chang, J. C., Lockner, D. A., and Reches, Z. 2014. Evolution of wear and friction along experimental faults. Pure Appl. Geophys, 171(11): 31253141, doi: 10.1007/s00024-014–0801-3.Google Scholar
Boneh, Y., Sagy, A., and Reches, Z. 2013. Frictional strength and wear-rate of carbonate faults during high-velocity, steady-state sliding, Earth Planet. Sci. Lett. 381: 127137, doi: 10.1016/j.epsl.2013.08.050.Google Scholar
Borghi, A., Aoudia, A., Javed, F., and Barzaghi, R. 2016. Precursory slow-slip loaded the 2009 L’Aquila earthquake sequence, Geophys. J. Int., 205(2): 776784, doi: 10.1093/gji/ggw046.Google Scholar
Bos, B., and Spiers, C. J. 2000. Effect of phyllosilicates on fluid-assisted healing of gouge-bearing faults. Earth Planet. Sci. Lett. 184(1): 199210, doi: 10.1016/s0012-821x(00)00304–6.Google Scholar
Bos, B., and Spiers, C. J. 2002. Frictional-viscous flow of phyllosilicate-bearing fault rock: Microphysical model and implications for crustal strength profiles. J. Geophys. Res.-Solid Earth 107(B2): 2028, doi: 10.1029/2001jb000301.Google Scholar
Bose, M., Allen, R., Brown, H., et al. 2014. CISN ShakeAlert: An Earthquake Early Warning demonstration systme for California. In Early Warnings for Geological Disasters – Scientific methods and Current Practice, ed. Zschau, E. W. A. J., Springer, Berlin, Germany.Google Scholar
Böse, M., Felizardo, C., and Heaton, T. H. 2015. Finite-fault rupture detector (FinDer): Going real-time in Californian ShakeAlert warning system. Seismol. Res. Lett. 86(6): 16921704.Google Scholar
Bouchon, M. 1982. The rupture mechanism of the Coyote Lake earthquake of 6 Aug. 1979 inferred from near field data. Bull. Seismol. Soc. Amer. 72: 745757.Google Scholar
Bouchon, M. 1997. The state of stress on some faults of the San Andreas system as inferred from near-field strong motion data. J. Geophys. Res. 102: 1173111744.Google Scholar
Bouchon, M., Bouin, M. P., Karabulut, H., Toksoz, M. N., Dietrich, M., and Rosakis, A. J. 2001. How fast is rupture during an earthquake? New insights from the 1999 Turkey earthquakes. Geophys. Res. Lett. 28(14): 27232726, doi: 10.1029/2001gl013112.Google Scholar
Bouchon, M., Campillo, M., and Cotton, F. 1998. Stress field associated with the rupture of the 1992 Landers, California, earthquake and its implications concerning the fault strength at the onset of the earthquake. J. Geophys. Res. 103: 2109121097.Google Scholar
Bouchon, M., Durand, V., Marsan, D., Karabulut, H., and Schmittbuhl, J. 2013. The long precursory phase of most large interplate earthquakes. Nat Geosci. 6(4): 299302, doi: 10.1038/ngeo1770.Google Scholar
Bouchon, M., Karabulut, H., Aktar, M., Ozalaybey, S., Schmittbuhl, J., and Bouin, M. P. 2011. Extended Nucleation of the 1999 Mw Izmit Earthquake. Science 331: 878882.Google Scholar
Bouchon, M., Karabulut, H., Bouin, M. P et al. 2010. Faulting characteristics of supershear earthquakes. Tectonophysics 493(3–4): 244253, doi: 10.1016/j.tecto.2010.06.011.Google Scholar
Bouchon, M., Marsan, D., Durand, V., et al. 2016. Potential slab deformation and plunge prior to the Tohoku, Iquique and Maule earthquakes. Nat. Geosci. 9(5): 380+, doi: 10.1038/ngeo2701.Google Scholar
Bourne, S. J., Arnadottir, T., Beavan, J., et al. 1998. Crustal deformation of the Marlborough fault zone in the South Island of New Zealand: Geodetic constraints over the interval 1982–1994. J. Geophys. Res.-Solid Earth 103: 3014730165.Google Scholar
Bourne, S. J., England, P. C., and Parsons, B. 1998. The motion of crustal blocks driven by flow of the lower lithosphere and implications for slip rates of continental strike-slip faults. Nature 391: 655659.Google Scholar
Bowden, F. P., and Tabor, D. 1950. The Friction and Lubrication of Solids: Part 1. Oxford: Clarendon Press.Google Scholar
Bowden, F. P., and Tabor, D. 1964. The Friction and Lubrication of Solids. Part II. Oxford: Clarendon Press.Google Scholar
Bowman, D. D., Ouillon, G., Sammis, C. G., Sornette, A., and Sornette, D. 1998. An observational test of the critical earthquake concept. J. Geophys. Res.-Solid Earth 103: 2435924372.Google Scholar
Bowman, J., Jones, T., Gibson, G., Corke, A., Thompson, R., and Comacho, A. 1988. Tennant Creek earthquakes of 22 January 1988: Reactivation of a fault zone in the Proterozoic Australian shield. Eos 69: 400.Google Scholar
Brace, W. F. 1960. An extension of the Griffith theory of fracture to rocks. J. Geophys. Res. 65: 34773480.CrossRefGoogle 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. 1963. Behavior of quartz during indentation. J. Geol. 71(5): 581595.Google Scholar
Brace, W. F. 1972. Laboratory studies of stick-slip and their application to earthquakes. Tectonophysics 14: 189200.Google Scholar
Brace, W. F. 1980. Permeability of crystalline and argillaceous rocks. Int. J. Rock Mech. Min. Sci. 17: 241251.Google Scholar
Brace, W. F. 1984. Permeability of crystalline rocks – New in situ measurements. J. Geophys. Res. 89 (NB6): 43274330.Google Scholar
Brace, W. F., and Bombalakis, 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 Byerlee, J. D. 1970. California earthquakes – Why only shallow focus? Science 168: 15731575.Google Scholar
Brace, W. F., and Kohlstedt, D. 1980. Limits on lithospheric stress imposed by laboratory experiments. J. Geophys. Res. 85: 62486252.Google Scholar
Brace, W. F., and Martin, R. J. 1968. A test of the law of effective stress for crystalline rocks of low porosity. Int. J. Rock Mech. Min. Sci. 5: 415426.Google Scholar
Brace, W. F., and Walsh, J. B. 1962. Some direct measurements of the surface energy of quartz and orthoclase. Am. Mineral 47: 11111122.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. 1969. A statistical theory of brittle fracture of rock materials. Int. J. Rock Mech. Min. Sci. I. 6: 2142.Google Scholar
Brantut, N., Heap, M. J., Baud, P., and Meredith, P. G. 2014. Rate- and strain-dependent brittle deformation of rocks. J. Geophys. Res.-Solid Earth 119(3): 18181836, doi: 10.1002/2013jb010448.Google Scholar
Brantut, N., Heap, M. J., Meredith, P. G., and Baud, P. 2013. Time-dependent cracking and brittle creep in crustal rocks: A review. J. Struct. Geol. 52: 1743, doi: 10.1016/j.jsg.2013.03.007.Google Scholar
Brantut, N., Schubnel, A., Corvisier, J., and Sarout, J. 2010. Thermochemical pressurization of faults during coseismic slip. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb006533.Google Scholar
Brantut, N., Stefanou, I., and Sulem, J. 2017. Dehydration-induced instabilities at intermediate depths in subduction zones. J. Geophys. Res.-Solid Earth 122(8): 60876107, doi: 10.1002/2017jb014357.Google Scholar
Braunmiller, J., and Nabelek, J. 1996. Geometry of continental normal faults: Seismological constraints. J. Geophys. Res.-Solid Earth 101: 30453052.Google Scholar
Braunmiller, J., and Nabelek, J. 2008. Segmentation of the Blanco Transform Fault Zone from earthquake analysis: Complex tectonics of an oceanic transform fault. J. Geophys. Res.-Solid Earth 113(B7): doi: 10.1029/2007jb005213.Google Scholar
Brechet, Y., and Estrin, Y. 1994. The effect of strain rate sensitivity on dynamic friction of metals, Scripta Met. Mater. 30: 14491454.Google Scholar
Bridgman, P. W. 1945. Polymorphic transitions and geological phenomena. Am. J. Sci. 243A: 9097.Google Scholar
Briggs, R. W., Sieh, K., Meltzner, A. J., et al. Deformation and slip along the Sunda megathrust in the great 2005 Nias-Simeulue earthquake. Science 311(5769): 1897–1901.Google Scholar
Brock, N. G. and Engelder, T. 1977. Deformation associated with movement of Muddy Mountain overthrust in Buffington Window, Southeastern Nevada. Geol. Soc. Am. Bull. 88 (11): 16671677.Google Scholar
Brodsky, E. E. 2006. Long-range triggered earthquakes that continue after the wave train passes. Geophys. Res. Lett. 33(15): L15313, doi: 10.1029/2006gl026605.Google Scholar
Brodsky, E. E., and Kanamori, H. 2001. Elastohydrodynamic lubrication of faults. J. Geophys. Res.-Solid Earth 106(B8): 1635716374, doi: 10.1029/2001jb000430.Google Scholar
Brodsky, E. E., and Mori, J. 2007. Creep events slip less than ordinary earthquakes. Geophys. Res. Lett. 34: L16309, doi: 16310.11029/12007GL030917.Google Scholar
Brodsky, E. E., and van der Elst, N. J. 2014. The Uses of Dynamic Earthquake Triggering. In Ann. Rev. Earth Planet. Sci. Vol 42, ed. R. Jeanloz, pp. 317339, doi: 10.1146/annurev-earth-060313–054648.Google Scholar
Brodsky, E. E., Gilchrist, J. J., Sagy, A., and Collettini, C. 2011. Faults smooth gradually as function of slip. Earth Planet. Sci. Lett. 302(1–2): 185193, doi: 10.1016/j.epsl.2010.12.010.Google Scholar
Brodsky, E. E., Kirkpatrick, J. D., and Candela, T. 2016. Constraints from fault roughness on the scale-dependent strength of rocks. Geology 44(1): 1922.Google Scholar
Brown, K. M., and Fialko, Y. 2012. “Melt welt” mechanism of extreme weakening of gabbro at seismic slip rates. Nature 488: 638641, doi: 10.1038/nature11370.Google Scholar
Brown, S. R., and Scholz, C. H. 1985a. Closure of random elastic surfaces in contact. J. Geophys. Res. 90: 55315545.Google Scholar
Brown, S. R., and Scholz, C. H. 1985b. Broad bandwidth study of the topography of natural rock surfaces. J. Geophys. Res. 90: 1257512582.Google Scholar
Brown, S. R., and Scholz, C. H. 1986. Closure of rock joints. J. Geophys. Res. 91: 49394948.Google Scholar
Brown, S. R., Scholz, C. H., and Rundle, J. B. 1991. A simplified spring-block model of earthquakes. Geophys. Res. Lett. 18, 215–18,218:Google Scholar
Brudy, M., Zoback, M. D., Fuchs, K., Rummel, F., and Baumgartner, J. 1997. Estimation of the complete stress tensor to 8 km depth in the KTB scientific drill holes: Implications for crustal strength. J. Geophys. Res.-Solid Earth 102: 1845318475.Google Scholar
Brudzinski, M. R., Thurber, C. H., Hacker, B. R., and Engdahl, E. R. 2007. Global prevalence of double Benioff zones. Science 316(5830): 14721474, doi: 10.1126/science.1139204.Google Scholar
Bruhat, L., Barbot, S., and Avouac, J.-P. 2011. Evidence for postseismic deformation of the lower crust following the 2004 Mw6.0 Parkfield earthquake. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb008073.Google Scholar
Brun, J. P., and Cobbold, P. R. 1980. Strain heating and thermal softening in continental shear zones: A review. J. Struct. Geol. 2: 149158.Google Scholar
Brune, J. 1968. Seismic moment, seismicity, and rate of slip along major fault zones. J. Geophys. Res. 73: 777784.Google Scholar
Brune, J. 1970. Tectonic stress and the spectra of seismic shear waves from earthquakes. J. Geophys. Res. 75: 49975009.Google Scholar
Brune, J. N. 1996. Particle motions in a physical model of shallow angle thrust faulting. P. Indian As-Math. Sci. -- Earth and Planetary Sciences 105(2): L197L206.Google Scholar
Brune, J. N. 2001. Fault normal dynamic loading and unloading: An explanation for “non-gouge” rock powder and lack of fault-parallel shear bands along the San Andreas Fault; EOS Transactions, American Geophysical Union. AGU Fall Meeting Abstracts. 8. 0655.Google Scholar
Brune, J. N., Brown, S., and Johnson, P. A. 1993. Rupture mechanism and interface separation in foam rubber models of earthquakes – A possible solution to the heat flow paradox and the paradox of large overthrusts. Tectonophysics 218(1–3): 5967, doi: 10.1016/0040–1951(93)90259-m.Google Scholar
Brune, J., Henyey, T., and Roy, R. 1969. Heat flow, stress, and rate of slip along the San Andreas fault, California. J. Geophys. Res. 74: 38213827.Google Scholar
Bucholc, M., and Steacy, S. 2016. Tidal stress triggering of earthquakes in Southern California, Geophys. J. Int. 205(2) 681693, doi: 10.1093/gji/ggw045.Google Scholar
Buck, W. R. 1988. Flexural rotation of normal faults. Tectonics 7: 959973.Google Scholar
Buck, W. R., Lavier, L. L., and Poliakov, A. N. B. 2005. Modes of faulting at mid-ocean ridges, Nature, 434(7034): 719723, doi: 10.1038/nature03358.Google Scholar
Bufe, C. G., and Varnes, D. J. 1993. Predictive modeling of the seismic cycle of the greater San Francisco Bay region. J. Geophys. Res.-Solid Earth 98: 98719883.Google Scholar
Burford, R. 1972. Continued slip on the Coyote Creek fault after the Borrego Mountain earthquake. In The Borrego Mountain Earthquake of April 9, 1968, ed. US Geol. Surv. Prof. Paper, pp. 105111.Google Scholar
Burford, R. 1988. Retardations in fault creep rates before local moderate earthquakes along the San Andreas fault system, central California. Pageoph 126: 499529.Google Scholar
Burford, R., and Harsh, P. W. 1980. Slip on the San Andreas fault in central California from alinement array surveys. Bull. Seismol. Soc. Am. 70: 12331261.Google Scholar
Burgmann, R., and Dresen, G. 2008. Rheology of the lower crust and upper mantle: Evidence from rock mechanics, geodesy, and field observations. Ann. Rev. Earth Planet. Sci. 38: 531567, doi: 10.1146/annurev.earth.36.031207.124326.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
Bürgmann, R., Rosen, P. A., and Fielding, E. J. 2000. Synthetic aperture radar interferometry to measure Earth’s surface topography and its deformation. Ann. Rev. Earth Planet. Sci. 28: 169209, doi: 10.1146/annurev.earth.28.1.169.Google Scholar
Bürgmann, R., Schmidt, D., Nadeau, R. M., et al. 2000. Earthquake potential along the northern Hayward fault, California. Science 289: 11781182.Google Scholar
Burnley, P. C., Green, H. W., and Prior, D. J. 1991. Faulting associated with the olivine to spinel transformation in Mg2GeO4 and its implications for deep‐focus earthquakes. J. Geophys. Res.-Solid Earth 96(B1): 425443.Google Scholar
Burr, N., and Solomon, S. 1978. The relationship of source parameters of oceanic transform earthquakes to plate velocity and transform length. J. Geophys. Res. 83: 11931205.Google Scholar
Burridge, R. 1973. Admissible speeds for plane strain self-similar shear cracks with friction but lacking cohesion. Geophys. J. R.A.S. 35: 439455.Google Scholar
Burridge, R., and Knopoff, L. 1967. Model and theoretical seismicity. Bull. Seism. Soc. Amer. 57: 341362.Google Scholar
Burroughs, S. M., and Tebbens, S. F. 2001. Upper-truncated power laws in natural systems. Pure Appl. Geophys. 158(4): 741757.Google Scholar
Byerlee, J. D. 1967a. Frictional characteristics of granite under high confining pressure. J. Geophys. Res. 72: 36393648.Google Scholar
Byerlee, J. D. 1967b. Theory of friction based on brittle fracture. J. Appl. Phys. 38: 29282934.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. 1990. Friction, overpressure, and fault normal compression. Geophys. Res. Lett. 17: 2109–12.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
Byerlee, J. D., and Savage, J. C. 1992. Coulomb plasticity within the fault zone. Geophys. Res. Lett. 19: 23412344.Google Scholar
Bykov, V. G. 2014. Sine-Gordon equation and its application to tectonic stress transfer. J. Seismol. 18(3): 497510, doi: 10.1007/s10950-014–9422-7.Google Scholar
Byrne, D. E., Davis, D. M., and Sykes, L. R. 1988. Loci and maximum size of thrust earthquakes and the mechanics of the shallow region of subduction zones. Tectonics 7: 833857.Google Scholar
Cailleux, A. 1958. Etude quantitative de failles. Revue de Geomorphologie Dynamique IX(9–10): 129145.Google Scholar
Caine, J. S., Evans, J. P., and Forster, C. B. 1996. Fault zone architecture and permeability structure. Geology 24: 10251028.Google Scholar
Camacho, A., Vernon, R. H., and Fitz Gerald, J. D. 1995. Large volumes of anhydrous pseudotachylyte in the Woodroffe Thrust, Eastern Musgrave Ranges, Australia, J. Struct. Geol., 17(3): 371383.Google Scholar
Campos, J., Madariaga, R., and Scholz, C. H. 1996. Faulting process of the August 8, 1993 Guam earthquake: A thrust event in an otherwise weakly coupled subduction zone. J. Geophys. Res. 101: 17,58117,596.Google Scholar
Campus, P., and Das, S. 2000. Comparison of the rupture and radiation characteristics of intermediate and deep earthquakes. J. Geophys. Res.-Solid Earth 105: 61776189.Google Scholar
Candela, T., and Brodsky, E. E. 2016. The minimum scale of grooving on faults. Geology 44(8): 603606, doi: 10.1130/g37934.1.Google Scholar
Candela, T., Renard, F., Bouchon, M., Schmittbuhl, J., and Bodsky, E. E. 2011a. Stress Drop during Earthquakes: Effect of Fault Roughness Scaling. Bull. Seismol. Soc. Amer. 101: 23692387, doi: 10.1785/0120100298.Google Scholar
Candela, T., Renard, F., Klinger, Y., Mair, K., Schmittbuhl, J., and Brodsky, E. E. 2012. Roughness of fault surfaces over nine decades of length scales. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb009041.Google Scholar
Candela, T., Renard, F., Schmittbuhl, J., Bouchon, M., and Brodsky, E. E. 2011b. Fault slip distribution and fault roughness. Geophys. J. Int. 187: 959968, doi: 10.1111/j.1365-246X.2011.05189.x.Google Scholar
Cao, T., and Aki, K. 1985. Seismicity simulation with a mass-spring model and a displacement hardening-softening friction law. Pageoph 122: 1023.Google Scholar
Cao, T., and Aki, K. 1986. Effect of slip rate on stress drop. Pageoph 124: 515530.Google Scholar
Carder, D. S. 1945. Seismic investigations in the Boulder Dam area, 1940–1945, and the influence of reservoir loading on earthquake activity. Bull. Seismol. Soc. Am. 35: 175192.Google Scholar
Cardwell, R. K., Chinn, D. S., Moore, G. F., and Turcotte, D. L. 1978. Frictional heating on a fault zone with finite thickness. Geophys. J. Roy. Astron. Soc. 52: 525530.Google Scholar
Carlson, J. M., and Langer, J. S. 1989a. Properties of earthquakes generated by fault dynamics. Phys. Rev. Lett. 62: 26322635.Google Scholar
Carlson, J. M., and Langer, J. S. 1989b. Mechanical model of an earthquake fault. Phys. Rev. A 40: 64706484.Google Scholar
Carlson, R. L., Hilde, T. W. C., and Uyeda, S. 1983. The driving mechanism of plate tectonics – relation to age of the lithosphere at trenches. Geophys. Res. Lett. 10: 297300.Google Scholar
Carpenter, B. M., Collettini, C., Viti, C., and Cavallo, A. 2016. The influence of normal stress and sliding velocity on the frictional behaviour of calcite at room temperature: Insights from laboratory experiments and microstructural observations. Geophys. J. Int. 205(1): 548561, doi: 10.1093/gji/ggw038.Google Scholar
Carpenter, B. M., Saffer, D. M., and Marone, C.. 2015. Frictional properties of the active San Andreas Fault at SAFOD: Implications for fault strength and slip behavior. J. Geophys. Res.-Solid Earth 120(7): 52735289, doi: 10.1002/2015jb011963.Google Scholar
Carter, N. L., and Kirby, S. 1978. Transient creep and semi-brittle behavior of crystalline rocks. Pure Appl. Geophys. 116: 807839.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 tracelength data from the Canyonlands graben of SE Utah. J. Struct. Geol. 17(9): 13191326, doi: 10.1016/0191–8141(95)00033-a.Google Scholar
Caskey, S. J., and Wesnousky, S. G. 1997. Static stress changes and earthquake triggering during the 1954 Fairview peak and Dixie valley earthquakes, central Nevada. Bull. Seismol. Soc. Amer. 87: 521527.Google Scholar
Castillo, D., and Hickman, S. H. 2000. Systematic near-field stress rotations adjacent to the Carrizo Plain segment of the San Andreas fault, paper presented at Proceedings of the 3rd conference on tectonic problems of the San Andreas fault system, School of Earth Sciences, Stanford University.Google Scholar
Castle, R., Elliot, M., Church, J., and Wood, S. 1984. The Evolution of the Southern California Uplift, 1955 through 1976. In US Geol. Surv. Prof. Paper 1342.Google Scholar
Causse, M., Cotton, F., and Mai, P. M. 2010. Constraining the roughness degree of slip heterogeneity. J. Geophys. Res. 115, doi: 10.1029/2009JB006747.Google Scholar
Cembrano, J., Jensen, E., Griffith, W. A., Stanton-Yonge, A., Mitchell, T., and Anzaldo, Y. 2018. Self-similar displacement-lenth scaling of strike-slip faults in crystalline rocks: Mechanical implications. J. Struct. Geol. In print.Google Scholar
Cessaro, R. K., and Hussong, D. M. 1986. Transform seismicity at the intersection of the oceanographer fracture zone and the Mid-Atlantic ridge. J. Geophys. Res. 91: 48394853.Google Scholar
Cetin, E., Cakir, Z., Meghraoul, M., Ergintav, S., and Akoglu, A. M. 2014. Extent and distribution of aseismic slip on the Ismetpasa segment of the North Anatolian fault. turkey) from persistent scatterer in In SAR. Geochem. Geophys. Geosyst 15: 28832894 doi: 2810.1002/2014GC005307.Google Scholar
Challen, J. M., and Oxley, P. L. B. 1979. An explanation of the different regimes of friction and wear using asperity deformation models. Wear 53: 229243.Google Scholar
Chaussard, E., Burgmann, R., Fattahi, H., Johnson, C. W., Nadeau, R., Taira, T., and Johanson, I. 2015. Interseismic coupling and refined earthquake potential on the Hayward-Calaveras fault zone. J. Geophys. Res.-Solid Earth 120(12): 85708590, doi: 10.1002/2015jb012230.Google Scholar
Cheloni, D., D’Agostino, N., D’Anastasio, E., et al. 2010. Coseismic and initial post-seismic slip of the 2009 M-w 6.3 L’Aquila earthquake, Italy, from GPS measurements. Geophys. J. Int. 181(3): 15391546, doi: 10.1111/j.1365-246X.2010.04584.x.Google Scholar
Chen, A., Frohlich, C., and Latham, G. 1982. Seismicity of the forearc marginal wedge (accretionary prism). J. Geophys. Res. 87: 36793690.Google Scholar
Chen, C. H., Tang, C.-C., Cheng, K.-C., et al. 2015. Groundwater-strain coupling before the 1999 M-W 7.6 Taiwan Chi-Chi earthquake. J. Hydrology 524: 378384, doi: 10.1016/j.jhydrol.2015.03.006.Google Scholar
Chen, J., Verberne, B. A., and Spiers, C. J. 2015a. Interseismic re-strengthening and stabilization of carbonate faults by “non-Dieterich” healing under hydrothermal conditions. Earth Planet. Sci. Lett. 423: 112, doi: 10.1016/j.epsl.2015.03.044.Google Scholar
Chen, J. Y., Verberne, B. A., and Spiers, C. J. 2015b. Effects of healing on the seismogenic potential of carbonate fault rocks: Experiments on samples from the Longmenshan Fault, Sichuan, China. J. Geophys. Res.-Solid Earth 120(8): 54795506, doi: 10.1002/2015jb012051.Google Scholar
Chen, L., and Talwani, P. 2001. Mechanism of initial seismicity following impoundment of the Monticello Reservoir, South Carolina. Bull. Seismol. Soc. Amer. 91: 94101.Google Scholar
Chen, T., and Lapusta, N. 2009. Scaling of small repeating earthquakes explained by interaction of seismic and aseismic slip in a rate and state fault model. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb005749.Google Scholar
Chen, W. P., and Molnar, P. 1983. Focal depths of intracontinental and intraplate earthquakes and their implications for the thermal and mechanical properties of the lithosphere. J. Geophys. Res. 88: 41834215.Google Scholar
Chen, X., and Shearer, P. M. 2013. California foreshock sequences suggest aseismic triggering process. Geophys. Res. Lett. 40(11): 26022607.Google Scholar
Chen, Y. 1988. Thermal model of oceanic transform faults. J. Geophys. Res. 93: 88398851.Google Scholar
Chernak, L. J., and Hirth, G. 2010. Deformation of antigorite serpentinite at high temperature and pressure. Earth Planet. Sci. Lett. 296(1): 2333.Google Scholar
Chernyshev, S. N., and Dearman, W. R. 1991. Rock Fractures. London: Butterworth-Heinemann Ltd.Google Scholar
Chester, F. M. 1994. Effect of temperature on friction – Constitutive equations and experiments with quartz gouge. J. Geophys. Res. 99: 72477261.Google Scholar
Chester, F. M. 1995. A rheologic model for wet crust applied to strike-slip faults. J. Geophys. Res.-Solid Earth 100: 1303313044.Google Scholar
Chester, F. M., and Chester, J. S. 2000. Stress and deformation along wavy frictional faults. J. Geophys. Res. 105: 23,421423,430.Google Scholar
Chester, F. M., and Higgs, N. G. 1992. Multimechanism friction constitutive model for ultrafine quartz gouge at hypocentral conditions. J. Geophys. Res.-Solid Earth 97: 18591870.Google Scholar
Chester, F. M., and Logan, J. M. 1986. Implications for mechanical properties of brittle faults from observations of the Punchbowl fault zone, California. Pageoph 124: 79106.Google Scholar
Chester, F. M., Biegel, R. L., and Evans, J. P. 1993. Internal structure and weakening mechanisms of the San-Andreas fault. J. Geophys. Res.-Solid Earth 98: 771786.Google Scholar
Chester, F. M., Friedman, M., and Logan, J. M. 1985. Foliated cataclasites. Tectonophysics 111: 134146.Google Scholar
Chester, F. M., Rowe, C. D., Ujiie, K., Kirkpatrick, J. D., Regalla, C., Remitti, F., Moore, J. C., Toy, V. G., Wolfson-Schwehr, M., and Bose, S. 2013. Structure and composition of the plate-boundary slip zone for the 2011 Tohoku-oki earthquake. Science 342: 12081211, doi: 10.1126/science.1243719.Google Scholar
Chester, J. S., Chester, F. M., and Kronenberg, A. K. 2005. Fracture surface energy of the Punchbowl fault, San Andreas system. Nature 437(7055): 133136.Google Scholar
Chiaraluce, L. 2012. Unravelling the complexity of Apenninic extensional fault systems: A review of the 2009 L’Aquila earthquake (Central Apennines, Italy). J. Struct. Geol. 42: 218, doi: 10.1016/j.jsg.2012.06.007.Google Scholar
Chiaraluce, L., Chiarabba, C., Collettini, C., Piccinini, D., and Cocco, M. 2007. Architecture and mechanics of an active low-angle normal fault: Alto Tiberina Fault, northern Apennines, Italy. J. Geophys. Res.-Solid Earth 112(B10): doi.org/10.1029/2007JB005015.Google Scholar
Chinnery, M. A. 1961. Deformation of the ground around surface faults. Bull. Seismol. Soc. Am. 51: 355372.Google Scholar
Chinnery, M. A. 1964. The strength of the earth’s crust under horizontal shear stress. J. Geophys. Res. 69: 20852089.Google Scholar
Chlieh, M., Avouac, J.-P., Hjorleifsdottir, V.,et al. 2007. Coseismic slip and afterslip of the great M-w 9.15 Sumatra-Andaman earthquake of 2004. Bull. Seismol. Soc. Am. 97(1): S152S173, doi: 10.1785/0120050631.Google Scholar
Chlieh, M., Avouac, J.-P., Sieh, K., Natawidjaja, D. H., and Galetzka, J. 2008. Heterogeneous coupling of the Sumatran megathrust constrained by geodetic and paleogeodetic measurements. J. Geophys. Res.-Solid Earth 113(B5): doi.org/10.1029/2007JB004981.Google Scholar
Chlieh, M., Perfettini, H., Tavera, H., et al. 2011. Interseismic coupling and seismic potential along the Central Andes subduction zone. J. Geophys. Res.-Solid Earth 116(B12): doi.org/10.1029/2010JB008166.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-Science Reviews 152: 7087, doi: 10.1016/j.earscirev.2015.11.006.Google Scholar
Chopra, P. N. 1997. High-temperature transient creep in olivine rocks. Tectonophysics 279(1–4): 93111, doi: 10.1016/s0040-1951(97)00134–0.Google Scholar
Choy, G. L., and Boatwright, J. L. 1995. Global patterns of radiated energy and apparent stress. J. Geophys. Res.-Solid Earth 100(B9): 1820518228, doi: 10.1029/95jb01969.Google Scholar
Choy, G. L., and McGarr, A. 2002. Strike-slip earthquakes in the oceanic lithosphere: Observations of exceptionally high apparent stress. Geophys. J. Int. 150(2): 506523, doi: 10.1046/j.1365-246X.2002.01720.x.Google Scholar
Choy, G. L., McGarr, A., Kirby, S. H., and Boatwright, J. 2006. An overview of the global variability in radiated energy and apparent stress. In Earthquakes: Radiated Energy and the Physics of Faulting, eds. R. Abercrombie, A. McGarr, G. DiToro and H. Kanamori, pp. 4357, doi: 10.1029/170gm01.Google Scholar
Christensen, D. H., and Ruff, L. 1983. Outer rise earthquakes and seismic coupling. Geophys. Res. Lett. 10: 697700.Google Scholar
Christie, J. M. 1960. Mylonitic rocks of the Moine thrust zone in the Assynt district, northwest Scotland. Trans. Geol. Soc. Edinburgh 18: 7993.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. Soc. Econ. Pal. Miner. Spec. Publ. 37, eds. K. Biddle and N. Christie-Blick. pp. 134.Google Scholar
Cicerone, R. D., Ebel, J. E., and Britton, J. 2009. A systematic compilation of earthquake precursors. Tectonophysics 476(3–4): 371396, doi: 10.1016/j.tecto.2009.06.008.Google Scholar
Cifuentes, I. 1989. The 1960 Chilean Earthquakes. J. Geophys. Res. 94: 665680, doi: 10.1029/JB094iB01p00665.Google Scholar
Cifuentes, I. L., and Silver, P. G. 1989. Low-frequency source characteristics of the great 1960 Chilean earthquake. J. Geophys. Res. 94: 643663.Google Scholar
Cirella, A., Piatanesi, A., Cocco, M., Tinti, E., Scognamiglio, L., Michelini, A., Lomax, A., and Boschi, E. 2009. Rupture history of the 2009 L’Aquila (Italy) earthquake from non-linear joint inversion of strong motion and GPS data. Geophys. Res. Lett. 36, doi: 10.1029/2009gl039795.Google Scholar
Cladouhos, T. T., and Allmendinger, R. W. 1993. Finite strain and rotation from fault-slip data. J. Struct. Geol. 15(6): 771784.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
Cloetingh, S., and Wortel, R. 1986. Stress in the Indo-Australian plate. Tectonophysics 132(1–3): 4967.Google Scholar
Cocco, M., and Rice, J. R. 2002. Pore pressure and poroelasticity effects in Coulomb stress analysis of earthquake interactions. J. Geophys. Res. 107: ESE 2-1-ESE 2-17, doi: 10.1029/2000JB000138.Google Scholar
Cochard, A., and Madariaga, R. 1996. Complexity of seismicity due to highly rate-dependent friction. J. Geophys. Res. 101: 2532125336.Google Scholar
Cochran, E. S., Li, Y. G., Shearer, P. M., Barbot, S., Fialko, Y., and Vidale, J. E. 2009. Seismic and geodetic evidence for extensive, long-lived fault damage zones. Geology 37(4): 315318, doi: 10.1130/g25306a.1.Google Scholar
Cochran, E. S., Vidale, J. E., and Tanaka, S. 2004. Earth tides can trigger shallow thrust fault earthquakes. Science 306(5699): 11641166, doi: 10.1126/science.1103961.Google Scholar
Cockerham, R. S., and Eaton, J. P. 1984. The April 24, 1984 Morgan Hill earthquake and its aftershocks. In The 1984 Morgan Hill, California, Earthquake, eds. Bennett, J. and Sherburne, R.. Sacramento, California: Calif. Div. of Mines: Calif. Div. Mines and Geol. Spec. Publ. 68, pp. 209213.Google Scholar
Collettini, C. 2011. The mechanical paradox of low-angle normal faults: Current understanding and open questions. Tectonophysics 510(3–4): 253268, doi: 10.1016/j.tecto.2011.07.015.Google Scholar
Collettini, C., and Holdsworth, R. E. 2004. Fault zone weakening and character of slip along low-angle normal faults: Insights from the Zuccale fault, Elba, Italy. J. Geol. Soc. London 161: 10391051.Google Scholar
Collettini, C., and Sibson, R. H. 2001. Normal faults, normal friction. Geology 29: 927930.Google Scholar
Collettini, C., Niemeijer, A., Viti, C., and Marone, C. 2009a. Fault zone fabric and fault weakness. Nature 462(7275): 907998, doi: 10.1038/nature08585.Google Scholar
Collettini, C., Viti, C., Smith, S. A. F., and Holdsworth, R. E. 2009b. Development of interconnected talc networks and weakening of continental low-angle normal faults. Geology, 37(6): 567570, doi: 10.1130/g25645a.1.Google Scholar
Collings, R., Lange, D., Rietbrock, A., et al. 2012. Structure and seismogenic properties of the Mentawai segment of the Sumatra subduction zone revealed by local earthquake traveltime tomography. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb00846Google Scholar
Colombelli, S., Zollo, A., Festa, G., and Picozzi, M. 2014. Evidence for a difference in rupture initiation between small and large earthquakes. Nat. Commun. 5, doi: 10.1038/ncomms4958.Google Scholar
Conneally, J., Childs, C., and Walsh, J. J. 2014. Contrasting origins of breached relay zone geometries. J. Struct. Geol. 58: 5968.Google Scholar
Contreras, J., Anders, M. H., and Scholz, C. H. 2000. Growth of a normal fault system: Observations from the Lake Malawi basin of the east African rift. J. Struct. Geol. 22: 159168.Google Scholar
Cook, R. F. 1986. Crack propagation thresholds: A measure of surface energy. J. Mater. Res. 1: 852860.Google Scholar
Cornell, C. A. 1968. Engineering seismic risk analysis. Bull. Seismol. Soc. Amer. 58: 15831606.Google Scholar
Cotton, F., and Coutant, O. 1997. Dynamic stress variations due to shear faults in a plane-layered medium. Geophys. J. Int. 128: 676688.Google Scholar
Cottrell, A. H. 1953. Dislocations and Plastic Flow in Crystals. Oxford: Clarendon Press.Google Scholar
Coward, M. P., Dewey, J. F., and Hancock, P. L. (eds.). 1987. Continental Extensional Tectonics. London: Blackwell.Google Scholar
Cowie, P. A., and Roberts, G. P. 2001. Constraining slip rates and spacings for active normal faults. J. Struct. Geol. 23(12): 19011915, doi: 10.1016/s0191-8141(01)00036–0.Google Scholar
Cowie, P. A., and Scholz, C. H. 1992a. 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. 1992b. Growth of faults by accumulation of seismic slip. J. Geophys. Res. 97 (B7): 1108511095.Google Scholar
Cowie, P. A., and Scholz, C. H. 1992c. Displacement-length scaling relationship for faults: Data synthesis and discussion. J. Struct. Geol. 14: 11491156.Google Scholar
Cowie, P. A., and Shipton, Z. K. 1998. Fault tip displacement gradients and process zone dimensions. J. Struct. Geol. 20(8): 983997, doi: 10.1016/s0191-8141(98)00029–7.Google Scholar
Cowie, P. A., Knipe, R. J., and Main, I. G. 1996. Special issue: Scaling laws for fault and fracture populations – Analyses and applications – Introduction. J. Struct. Geol. 18: R5R11.Google Scholar
Cowie, P. A., Scholz, C. H., Edwards, M., and Malinverno, A. 1993. Fault strain and seismic coupling on midocean ridges. J. Geophys. Res.-Solid Earth 98: 1791117920.Google Scholar
Cowie, P. A., Vanneste, C., and Sornette, D. 1993. Statistical physics model for the spacitemporal evolution of faults. J. Geophys. Res.-Solid Earth 98(B12): 2180921821, doi: 10.1029/93jb02223.Google Scholar
Cowie, P., Phillips, R., Roberts, G., et al. 2017. Orogen-scale uplift drives episodic behaviour of earthquake faults. Sci. Rep. 7: 44858.Google Scholar
Cowie, P., Scholz, C., Roberts, G. P., Walker, J. F., and Steer, P. 2013. Viscous roots of active seismogenic faults revealed by geologic slip rate variations. Nat. Geosci. 6(12): 10361040.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: An experimental study. J. Struct. Geol. 10: 413430.Google Scholar
Crampin, S. 1987. Geological and industrial applications of extensive-dilatancy anisotropy. Nature 328: 491496.Google Scholar
Crampin, S., Evans, R., and Atkinson, B. K. 1984. Earthquake prediction: A new physical basis. Geophys. J. R.A.S. 76: 147156.Google Scholar
Crampin, S., Gao, Y., and Bukits, J. 2015. A review of retrospective stress-forecasts of earthquakes and eruptions. Phys. Earth Planet. Int. 245: 7687, doi: 10.1016/j.pepi.2015.05.008.Google Scholar
Crampin, S., Volti, T., and Stefansson, R. 1999. A successfully stress-forecast earthquake. Geophys. J. Int. 138(1): F1-F5, doi: 10.1046/j.1365-246x.1999.00891.x.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(4): 691707, doi: 10.1016/j.jsg.2003.07.007.Google Scholar
Crider, J. G., and Pollard, D. D. 1998. Fault linkage: Three-dimensional mechanical interaction between echelon normal faults. J. Geophys. Res.-Solid Earth 103: 2437324391.Google Scholar
Crone, A. J., and Luza, K. V. 1986. Holocene deformation associated with the Meers fault, southwestern Oklahoma. In The Slick Hills of Southwestern Oklahoma – Fragments of an Aulachogen? ed. Donovan, R. N.. Norman, Oklahoma: University of Oklahoma, pp. 6874.Google Scholar
Crone, A. J., Machette, M., Bonilla, M., Lienkaemper, J., Pierce, K., Scott, W., and Bucknam, R. 1987. Surface faulting accompanying the Borah Peak earthquake and segmentation of the Lost River fault central Idaho. Bull. Seismol. Soc. Am. 77: 739770.Google Scholar
Crone, A., and Machette, M. 1984. Surface faulting accompanying the Borah Peak earthquake, central Idaho. Geology 12: 664667.Google Scholar
Crook, C. N. 1984. Geodetic measurement of the horizontal crustal deformation associated with the Oct. 15, 1979 Imperial Valley (California) earthquake. PhD thesis, University of London,Google Scholar
Crowell, J. C. 1984. Origin of late Cenozoic basins in southern California. In Tectonics and Sedimentation. ed. W. Dickinson. Soc. Econ. Pal. Miner., pp. 190204.Google Scholar
Cruden, D. M. 1970. A theory of brittle creep in rock under uniaxial compression. J. Geophys. Res. 75: 34313442.Google Scholar
Cubas, N., Avouac, J. P., Leroy, Y. M., and Pons, A. 2013. Low friction along the high slip patch of the 2011 Mw 9.0 Tohoku-Oki earthquake required from the wedge structure and extensional splay faults. Geophys. Res. Lett. 40(16): 42314237, doi: 10.1002/grl.50682.Google Scholar
D’Agostino, N., Avallone, A., Cheloni, D., D’Anastasio, E., Mantenuto, S., and Selvaggi, G. 2008. Active tectonics of the Adriatic region from GPS and earthquake slip vectors. J. Geophys. Res.-Solid Earth 113(B12): doi: 10.1029/2008jb005860.Google Scholar
D’Agostino, N., Cheloni, D., Fornaro, G., Giuliani, R., and Reale, D. 2012. Space-time distribution of afterslip following the 2009 L’Aquila earthquake. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb008523.Google Scholar
Dahlen, F. A., and Barr, T. D. 1989. Frictional mountain building 1. Deformation and mechanical energy budget. J. Geophys. Res. 94: 39063922.Google Scholar
Dahmen, K. A., Ben-Zion, Y., and Uhl, J. T. 2011. A simple analytic theory for the statistics of avalanches in sheared granular materials. Nature Physics 7(7): 554557, doi: 10.1038/nphys1957.Google Scholar
Das, S. 1981. Three-dimensional rupture propagation and implications for the earthquake source mechanism. Geophys. J. R.A.S. 67: 375393.Google Scholar
Das, S. 1982. Appropriate boundary conditions for modeling very long earthquakes and physical consequences. Bull. Seismol. Soc. Am. 72: 19111926.Google Scholar
Das, S. 1993. The Macquarie Ridge earthquake of 1989. Geophys. J. Int. 115: 778798.Google Scholar
Das, S., and Aki, K. 1977. Fault planes with barriers: A versatile earthquake model. J. Geophys. Res. 82: 56585670.Google Scholar
Das, S., and Kostrov, B. 1983. Breaking of a single asperity: Rupture process and seismic radiation. J. Geophys. Res. 88: 42774288.Google Scholar
Das, S., and Scholz, C. 1981a. Off-fault aftershock clusters caused by shear stress increase? Bull Seismol. Soc. Am. 71: 16691675.Google Scholar
Das, S., and Scholz, C. 1981b. Theory of time-dependent rupture in the earth. J. Geophys. Res. 86: 60396051.Google Scholar
Das, S., and Scholz, C. H. 1983. Why large earthquakes do not nucleate at shallow depths. Nature 305: 621623.Google Scholar
Daub, E. G., Manning, M. L., and Carlson, J. M. 2010. Pulse-like, crack-like, and supershear earthquake ruptures with shear strain localization. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb006388.Google Scholar
Davidsen, J., Gu, C., and Baiesi, M. 2015. Generalized Omori-Utsu law for aftershock sequences in southern California. Geophys. J. Int. 201(2): 965978, doi: 10.1093/gji/ggv061.Google Scholar
Davies, G., and Brune, J. N. 1971. Global plate motion rates from seismicity data. Nature 229: 101107.Google Scholar
Davis, D., Dahlen, F. A., and Suppe, J. 1983. Mechanics of fold-and-thrust belts and accretionary wedges. J. Geophys. Res. 88: 11531172.Google Scholar
Davison, F., and Scholz, C. 1985. Frequency-moment distribution of earthquakes in the Aleutian Arc: A test of the characteristic earthquake model. Bull. Seismol. Soc. Am. 75: 13491362.Google Scholar
Davy, P., Sornette, A., and Sornette, D. 1990. Some consequences of a proposed fractal nature of continental faulting. Nature 348: 5658.Google Scholar
Dawers, N. H., and Anders, M. H. 1995. Displacement–length scaling and fault linkage. J. Struct. Geol. 17: 607611.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
Day, S. M. 1982. Three-dimensional simulation of spontaneous rupture: The effect of nonuniform prestress. Bull. Seismol. Soc. Am. 72: 18811902.Google Scholar
Day, S. M., Yu, G., and Wald, D. J. 1998. Dynamic stress changes during earthquake rupture. Bull. Seismol. Soc. Amer. 88: 512522.Google Scholar
De Paola, N., Hirose, T., Mitchell, T., Di Toro, G., Viti, C., and Shimamoto, T. 2011a. Fault lubrication and earthquake propagation in carbonate rocks. In Multiscale and Multiphysics Processes in Geomechanics: Results of the Workshop on Multiscale and Multiphysics Processes in Geomechanics, ed. R. I. Borja, pp. 153156.Google Scholar
De Paola, N., Hirose, T., Mitchell, T., Di Toro, G., Viti, C., and Shimamoto, T. 2011b. Fault lubrication and earthquake propagation in thermally unstable rocks, Geology, 39: 3538, doi: 10.1130/G31398.1.Google Scholar
Deacon, R. F., and Goodman, J. F. 1958. Lubrication by lamilar solids. Proc. R. Soc. Lond. A Math. Phys. Sci. 243 (1235): 464+, doi: 10.1098/rspa.1958.0013.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.-Solid Earth and Planets 91(B5): 49204938.Google Scholar
Delaplace, A., Schmittbuhl, J., and Maloy, K. J. 1999. High resolution description of a crack front in a heterogeneous Plexiglas block. Physical Review E 60(2): 13371343, doi: 10.1103/PhysRevE.60.1337.Google Scholar
DeMets, C., Gordon, R. G., Argus, D. F., and Stein, S. 1990. Current plate motions. Geophys. J. Int. 101: 425478.Google Scholar
DeMets, C., Gordon, R. G., Stein, S., and Argus, D. F. 1987. A revised estimate of Pacific-North America motion and implications for western North America Plate boundary tectonics. Geophys. Res. Lett. 14: 911914.Google Scholar
den Hartog, S. A. M., and Spiers, C. J. 2014. A microphysical model for fault gouge friction applied to subduction megathrusts. J. Geophys. Res.-Solid Earth 119(2): 15101529, doi: 10.1002/2013jb010580.Google Scholar
den Hartog, S. A. M., Niemeijer, A. R., and Spiers, C. J. 2012. New constraints on megathrust slip stability under subduction zone P-T conditions. Earth Planet. Sci. Lett. 353: 240252, doi: 10.1016/j.epsl.2012.08.022.Google Scholar
den Hartog, S. A. M., Niemeijer, A. R., and Spiers, C. J. 2013. Friction on subduction megathrust faults: Beyond the illite-muscovite transition. Earth Planet. Sci. Lett. 373: 819, doi: 10.1016/j.epsl.2013.04.036.Google Scholar
den Hartog, S. A., Saffer, D. M., and Spiers, C. J. 2014. The roles of quartz and water in controlling unstable slip in phyllosilicate-rich megathrust fault gouges. Earth Planets and Space 66(78): doi: 10.1186/1880–5981-66–78.Google Scholar
Deng, J. S., and Sykes, L. R. 1997. Evolution of the stress field in southern California and triggering of moderate-size earthquakes: A 200-year perspective. J. Geophys. Res.-Solid Earth 102: 98599886.Google Scholar
Deng, J. S., Gurnis, M., Kanamori, H. et al. 1998. Viscoelastic flow in the lower crust after the 1992 Landers California earthquake. Science 282: 16891692.Google Scholar
Deng, Q., Wu, D., Zhang, P., and Chen, S. 1986. Structure and deformational character of strike-slip fault zones. Pageoph 124: 203224.Google Scholar
Denolle, M. A., and Shearer, P. M. 2016. New perspectives on self-similarity for shallow thrust earthquakes. J. Geophys. Res.-Solid Earth 121(9): 65336565, doi: 10.1002/2016jb013105.Google Scholar
Deplus, C., Diament, M., Hébert, H., et al. 1998. Direct evidence of active deformation in the eastern Indian oceanic plate. Geology 26(2): 131134.Google Scholar
DePolo, C. M. 2008. Quaternary Faults in Nevada, Nevada Bureau of Mines and Geology Map 167.Google Scholar
Detrick, R., White, R., and Purdy, G. 1993. Crustal structure of North Atlantic fracture zones. Rev. Geophys. 31(4): 439458.Google Scholar
Dewey, J. F., and Bird, J. M. 1970. Mountain belts and the new global tectonics. J. Geophys. Res. 75: 26252647.Google Scholar
Di Luccio, F., Ventura, G., Di Giovambattista, R., Piscini, A., and Cinti, F. R. 2010. Normal faults and thrusts reactivated by deep fluids: The 6 April 2009 M-w 6.3 L’Aquila earthquake, central Italy. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb007190.Google Scholar
Di Toro, G., and Pennacchioni, G. 2005. Fault plane processes and mesoscopic structure of a strong-type seismogenic fault in tonalites (Adamello batholith, Southern Alps). Tectonophysics 402(1–4): 5580.Google Scholar
Di Toro, G., Goldsby, D. L., and Tullis, T. E. 2004. Friction falls towards zero in quartz rock as slip velocity approaches seismic rates. Nature 427(6973): 436439.Google Scholar
Di Toro, G., Han, R., Hirose, T., et al. 2011. Fault lubrication during earthquakes. Nature 471: 494498, doi: 10.1038/nature09838.Google Scholar
Di Toro, G., Hirose, T., Nielsen, S., Pennacchioni, G., and Shimamoto, T. 2006. Natural and experimental evidence of melt lubrication of faults during earthquakes. Science 311: 647649.Google Scholar
Di Toro, G., Pennacchioni, G., and Nielsen, S. 2009. Pseudotachylytes and Earthquake Source Mechanics. In Fault-Zone Properties and Earthquake Rupture Dynamics, ed. Fukuyama, E.. London: Academic Press, pp. 87133.Google Scholar
Di Toro, G., Pennacchioni, G., and Teza, G. 2005. Can pseudotachylytes be used to infer earthquake source parameters? An example of limitations in the study of exhumed faults. Tectonophysics 402(1–4): 320.Google Scholar
DiCapria, C. J., Simons, M., Kenner, S. J., and Williams, C. A. 2008. Post-seismic reloading and temporal clustering on a single fault. Geophys. J. Int. 172: 581592, doi: 10.1111/j.1365-246X.2007.03622.x.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. 1979a. Modelling of rock friction: 1. Experimental results and constitutive equations. J. Geophys. Res. 84: 21612168.Google Scholar
Dieterich, J. H. 1979b. Modelling of rock friction: 2. Simulation of preseismic slip. J. Geophys. Res. 84: 21692175.Google Scholar
Dieterich, J. H. 1981. Constitutive properties of faults with simulated gouge. In Mechanical Behavior of Crustal Rocks. AGU Geophys. Mono., eds. Carter, M. F. N., Logan, J., and Sterns, D.. Washington, DC: American Geophysical Union, pp. 103120.Google Scholar
Dieterich, J. H. 1986. A model for the nucleation of earthquake slip. In Earthquake Source Mechanics. AGU Geophys. Mono., eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 3749.Google Scholar
Dieterich, J. H. 1992. Earthquake nucleation on faults with rate-dependent and state-dependent strength. Tectonophysics 211: 115134.Google Scholar
Dieterich, J. H. 1994. A constitutive law for rate of earthquake production and its application to earthquake clustering, J. Geophys. Res., 99: 26012618.Google Scholar
Dieterich, J. H., and Conrad, G. 1984. Effect of humidity on time- and velocity-dependent friction in rocks. J. Geophys. Res. 89: 41964202.Google Scholar
Dieterich, J. H., and Kilgore, B. D. 1994. Direct observations of frictional contacts: New insights for state-dependent properties. In Faulting, Friction, and Earthquake Mechanics, Part II, eds. Marone, C. J. and Blanpied, M. L.. Basel: Birkhauser, pp. 283302.Google Scholar
Dieterich, J. H., and Kilgore, B. 1996. Implications of fault constitutive properties for earthquake prediction. Proc. Natl Acad. Sci. USA 93: 37873794.Google Scholar
Dieterich, J. H., and Smith, D. E. 2009. Nonplanar Faults: Mechanics of Slip and Off-fault Damage. Pure Appl. Geophys 166(10–11): 17991815, doi: 10.1007/s00024-009–0517-y.Google Scholar
Dmowska, R., Rice, J. R., Lovison, L. C., and Josell, D. 1988. Stress transfer and seismic phenomena in coupled subduction zones during the earthquake cycle. J. Geophys. Res. 93: 78697885.Google Scholar
Doan, M. L., and Gary, G. 2009. Rock pulverization at high strain rate near the San Andreas fault. Nat. Geosci. 2(10): 709712, doi: 10.1038/ngeo640.Google Scholar
Dobson, D. P., Meredith, P. G., and Boon, S. A. 2002. Simulation of subduction zone seismicity by dehydration of serpentine. Science 298(5597): 14071410.Google Scholar
Dodd, R. K., Eibeck, J. C., Gibbon, J. D., and Morris, H. C. 1982. Solitons and Nonlinear Wave Equations, London: Academic Press.Google Scholar
Dodge, D. A., Beroza, G. C., and Ellsworth, W. L. 1996. Detailed observations of California foreshock sequences: Implications for the earthquake initiation process. J. Geophys. Res.-Solid Earth 101: 2237122392.Google Scholar
Doglioni, C. 1990. The global tectonic pattern. J. Geodyn. 12: 2138.Google Scholar
Doglioni, C., Merlini, S., and Cantarella, G. 1999. Foredeep geometries at the front of the Apennines in the Ionian Sea (central Mediterranean). Earth Planet. Sci. Lett. 168: 243254.Google Scholar
Dokka, R. K., and Travis, C. J. 1990. Role of the eastern California shear zone in accommodating Pacific-North-American plate motion. Geophys. Res. Lett. 17: 13231326.Google Scholar
Donath, F. A. 1961. Experimental study of shear failure in anisotropic rocks. Bull. Geol. Soc. Am. 72: 985990.Google Scholar
Donnellan, A., and Lyzenka, G. A. 1998. GPS observations of fault afterslip and upper crustal deformation following the Northridge earthquake. J. Geophys. Res. 103: 21,28521,297.Google Scholar
Dor, O., Ben-Zion, Y., Rockwell, T. K., and Brune, J. 2006. Pulverized rocks in the Mojave section of the San Andreas Fault Zone. Earth Planet. Sci. Lett. 245(3–4): 642654, doi: 10.1016/j.epsl.2006.03.034.Google Scholar
Dorbath, C., Gerbault, M., Carlier, G., and Guiraud, M.. 2008. Double seismic zone of the Nazca plate in northern Chile: High-resolution velocity structure, petrological implications, and thermomechanical modeling. Geochem. Geophys. Geosystems 9, doi: 10.1029/2008gc002020.Google Scholar
Doser, D. I. 1988. Source parameters of earthquakes in the Nevada seismic zone, 1915–43. J. Geophys. Res. 93: 1500115015.Google Scholar
Doser, D., and Kanamori, H. 1986. Depth of seismicity in the Imperial Valley region. 1977–83) and its relationship to heat flow, crustal structure and the October 15, 1979 earthquake. J. Geophys. Res. 91: 675688.Google Scholar
Doser, I. 1986. Earthquake processes in the Rainbow Mountain–Fairview Peak–Dixie Valley, Nevada, region 1954–1959. J. Geophys. Res. 91: 1257212586.Google Scholar
Dreger, D. S., Oglesby, D. D., Harris, R., Ratchkovski, N., and Hansen, R. 2004. Kinematic and dynamic rupture models of the November 3, 2002 Mw7.9 Denali, Alaska, earthquake. Geophys. Res. Lett. 31(4): doi: 10.1029/2003gl018333.Google Scholar
Duan, B., and Oglesby, D. D. 2005. Multicycle dynamics of nonplanar strike‐slip faults. J. Geophys. Res.-Solid Earth 110(B3): doi.org/10.1029/2004JB003298.Google Scholar
Duan, B., and Oglesby, D. D. 2006. Heterogeneous fault stresses from previous earthquakes and the effect on dynamics of parallel strike‐slip faults. J. Geophys. Res.-Solid Earth 111(B5): doi.org/10.1029/2005JB004138.Google Scholar
Dugdale, D. S. J. 1960. Yielding of steel sheets containing slits. Mech. Phys. Solids 8: 100115.Google Scholar
Dunham, E. M. 2007. Conditions governing the occurrence of supershear ruptures under slip-weakening friction. J. Geophys. Res.-Solid Earth 112(B7): doi: 10.1029/2006jb004717.Google Scholar
Dunham, E. M., Kozdon, J. E., Belanger, D., and Cong, L. 2011. Earthquake ruptures on rough faults, in multiscale and multiphysics processes in geomechanics: Results of the workshop on multiscale and multiphysics processes. In Geomechanics, ed. Borja, R. I., pp. 145-+, Berlin: Springer-Verlag.Google Scholar
Dunlap, W. J., Hirth, G., and Teyssier, C. 1997. Thermomechanical evolution of a ductile duplex. Tectonics 16(6): 9831000, doi: 10.1029/97tc00614.Google Scholar
Dunning, J. D., Petrovski, D., Schuyler, J., and Owens, A. 1984. The effects of aqueous chemical environments on crack growth in quartz. J. Geophys Res. 89: 41154124.Google Scholar
Duquesnoy, T., Barrier, E., Kasser, M., et al. 1994. Detectionof creep along the Philippine fault – 1st results of geodetic measurements on Leyte Island, Central Philippine. Geophys. Res. Lett. 21(11): 975978, doi: 10.1029/94gl00640.Google Scholar
Durney, D. W., and Ramsay, J. G. 1973. Incremental strains measured by syntectonic crystal growths. In Gravity and Tectonics, eds. de Jong, K. A. and Scholten, R.. New York: John Wiley, pp. 6796.Google Scholar
DuRoss, C. B., Personius, S. F., Crone, A. J., et al. 2016. Fault segmentation: New concepts from the Wasatch Fault Zone, Utah, USA. J. Geophys. Res.-Solid Earth 121(2): 11311157, doi: 10.1002/2015jb012519.Google Scholar
Dziak, R. P., Fox, C. G., and Embley, R. W. 1991. Relationship between the seismicity and geologic structure of the Blanco Transform Fault Zone. Marine Geophysical Research 13(3): 203208.Google Scholar
Eaton, J. P., O’Neill, M. E. and Murdock, J. N. 1970. Aftershocks of the 1966 Parkfield–Cholame, California earthquake. A detailed study. Bull. Seismol. Soc. Am. 60: 11511197.Google Scholar
Eaton, J., Cockerham, R., and Lester, F. 1983. Study of the May 2, 1983 Coalinga earthquake and its aftershocks, based on the U.S.G.S. seismic network in northern California. In The 1983 Coalinga, California, Earthquakes. Spec. Pub., eds. Bennet, J. and Sherbume, R.. Sacramento: California Department of Conservation, Division of Mines, pp. 923.Google Scholar
Eberhart-Phillips, D., and Reyners, M. 1999. Plate interface properties in the northeast Hikurangi subduction zone, New Zealand, from converted seismic waves. Geophysical Research Letters, 26(16): 25652568, doi: 10.1029/1999gl900567.Google Scholar
Eberhart-Phillips, D., Haeussler, P. J., and Freymueller, J. T., et al. 2003. The 2002 Denali fault earthquake, Alaska: A large magnitude, slip-partitioned event. Science, 300(5622): 11131118, doi: 10.1126/science.1082703.Google Scholar
Ebinger, C. J., and Hayward, N. J. 1996. Soft plates and hot spots: Views from afar. J. Geophys. Res. 101: 2185921976.Google Scholar
Edmond, J. M., and Paterson, M. S. 1972. Volume changes during the deformation of rocks at high pressure. Int. J. Rock Mech. Min. Sci. 9: 161182.Google Scholar
Eggler, D. H., and Ehmann, A. N. 2010. Rate of antigorite dehydration at 2 GPa applied to subduction zones. Am. Mineral. 95(5–6): 761769.Google Scholar
Einarsson, P., and Eiriksson, J. 1982. Earthquake fractures in the districts Land and Rangarvellin in the South Iceland seismic zone. Jokull 32: 113120.Google Scholar
Einarsson, P., Bjornsson, S., Foulger, G., Stefansson, R., and Skaftadottir, T. 1981. Seismicity pattern in the south Iceland seismic zone. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 141152.Google Scholar
Einav, I., Rognon, P., Miller, T., and Sulem, J. 2018., Faults get colder through transient granular vortices. Geophys. Res. Lett, 45: 26252632, doi: 10.1002/2017GL076029.Google Scholar
Ekstrom, G., and Romanowicz, B. 1990. The 23 May 1989 Macquarie Ridge earthquake: A very broad band analysis. Geophys. Res. Lett. 17: 993996.Google Scholar
Ekstrom, G., Stein, R. S., Eaton, J. P., and Eberhart-Phillips, D. 1992. Seismicity and geometry of a 110-km-long blind thrust-fault 1. The 1985 Kettleman Hills, California, Earthquake. J. Geophys. Res.-Solid Earth 97: 48434864.Google Scholar
Elkhoury, J. E., Brodsky, E. E., and Agnew, D. C. 2006. Seismic waves increase permeability. Nature 441(7097): 11351138, doi: 10.1038/nature04798.Google Scholar
Elliott, A. J., Dolan, J. F., and Oglesby, D. D. 2009. Evidence from coseismic slip gradients for dynamic control on rupture propagation and arrest through stepovers. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb005969.Google Scholar
Elliott, D. 1976. The energy balance and deformation mechanisms of thrust sheets. Phil. Trans. Roy. Soc. London Ser. A 283: 289312.Google Scholar
Ellsworth, W. L. 2013. Injection-Induced Earthquakes. Science 341(6142): 142+, doi: 10.1126/science.1225942.Google Scholar
Ellsworth, W. L., and Beroza, G. C. 1995. Seismic evidence for an earthquake nucleation phase. Science 268: 851855.Google Scholar
Ellsworth, W. L., and Bulut, F. 2018. Nucleation of the 1999 Izmit earthquake by a triggered cascade of foreshocks. Nature Geoscience 11(7): 531535, doi: 10.1038/s41561-018-0145-1.Google Scholar
Ellsworth, W. L., Lindh, A. G., Prescott, W. H., and Herd, D. G. 1981. The 1906 San Francisco earthquake and the seismic cycle. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 126140.Google Scholar
Emmanuel, S., and Berkowitz, B. 2006. Suppression and stimulation of seafloor hydrothermal convection by exothermic mineral hydration. Earth Planet. Sci. Lett. 243(3): 657668.Google Scholar
Emter, C. 1997. Tidal triggering of earthquakes and volcanic events. In Tidal Phenomena, Lecture Notes in Earth Sciences, vol. 66, ed. Wilhelm, H., Zurn, W. and Wenzel., H. G. New York: Springer, pp. 293310.Google Scholar
Engdahl, E. R. 1977. Seismicity and plate subduction in the central Aleutians. In Island Arcs and Deep Sea Trenches and Back-arc Basins. Ewing, M.; Ser. 1, eds. Talwani, I. M. and Pittman, W.. Washington, DC: American Geophysical Union, pp. 259272.Google Scholar
Engdahl, E. R., and Scholz, C. H. 1977. A double Benioff zone beneath the central Aleutians: An unbending of the lithosphere. Geophys. Res. Lett. 4(10): 473476.Google Scholar
Engdahl, E. R., Villaseñor, A., DeShon, H. R., and Thurber, C. H. 2007. Teleseismic relocation and assessment of seismicity (1918–2005) in the region of the 2004 Mw 9.0 Sumatra–Andaman and 2005 Mw 8.6 Nias Island great earthquakes. Bull. Seismol. Soc. Am. 97(1A): S43S61.Google Scholar
Engelder, J. T. 1974a. Cataclasis and the generation of fault gouge. Geol. Soc. Am. Bull. 85: 15151522.Google Scholar
Engelder, T. 1974b. Microscopic wear grooves on slickensides: Indicators of paleoseismicity. J. Geophys. Res. 79: 4387.Google Scholar
Engelder, T. 2014. Stress regimes in the Lithosphere. New Jersey: Princeton University Press.Google Scholar
Engelder, T., and Scholz, C. H. 1976. The role of asperity indentation and ploughing in rock friction – II. Influence of relative hardness and normal load. Int. J. Rock Mech. Min. Sci. and Geomech. 13: 155163.Google Scholar
Engelder, T., Logan, J., and Handin, J. 1975. The sliding characteristics of sandstone on quartz fault gouge. J. Pure Appl. Geophys. 113: 6986.Google Scholar
Engeln, J. F., Wiens, D. A., and Stein, S. 1986. Mechanisms and depths of Atlantic transform earthquakes. J. Geophys. Res.-Solid Earth 91(B1): 548577.Google Scholar
England, P. C., and McKenzie, D. P. 1982. A thin viscous sheet model for continental deformation. Geophys. J. R.A.S. 70: 295321.Google Scholar
Engvik, A. K., Bertram, A., Kalthoff, J. F., Stockhert, B., Austrheim, H., and Elvevold, S. 2005. Magma-driven hydraulic fracturing and infiltration of fluids into the damaged host rock, an example from Dronning Maud Land, Antarctica. J. Struct. Geol. 27(5): 839854.Google Scholar
Engvik, L., Stockhert, B., and Engvik, A. K. 2009. Fluid infiltration, heat transport, and healing of microcracks in the damage zone of magmatic veins: Numerical modeling. J. Geophys. Res.-Solid Earth 114: doi.org/10.1029/2008JB005880.Google Scholar
Erismann, T., Heuberger, H., and Preuss, E. 1977. Der Bimstein von Kofels (Tirol), ein Bergsturz-“Frictionit.” Tschermaks Mineral. Petrogr. Mitt. 24: 67119.Google Scholar
Escartin, J., Andreani, M., Hirth, G., and Evans, B. 2008. Relationships between the microstructural evolution and the rheology of talc at elevated pressures and temperatures. Earth Planet. Sci. Lett. 268: 468475.Google Scholar
Escartin, J., Hirth, G., and Evans, B. 1997. Nondilatant brittle deformation of serpentinites: Implications for Mohr-Coulomb theory and the strength of faults. J. Geophys. Res. 102: 28972913.Google Scholar
Escartin, J., Smith, D. K., Cann, J., Schouten, H., Langmuir, C. H., and Escrig, S. 2008. Central role of detachment faults in accretion of slow-spreading oceanic lithosphere. Nature 455(7214): 790.Google Scholar
Eshelby, J. 1957. The determination of the elastic field of an ellipsoidal inclusion and related problems. Proc. Roy. Soc. London Series A 241: 376396.Google Scholar
Espinosa-Aranda, J. M., Cuellar, A., Garcia, G., et al. 2009. Evolution of the Mexican Seismic Alert System (SASMEX). Seismol. Res. Lett. 80: 694706.Google Scholar
Etchecopar, A., Granier, T., and Larroque, J.-M. 1986. Origine des fentes en echelon: Propogation des failles. R. Acad. Sci. Paris 302 (II): 479484.Google Scholar
Etheridge, M. A., Wall, V. J., Cox, S. F., and Vernon, R. H. 1984. High fluid pressures during regional metamorphism: Implications for mass transport and deformation mechanisms. J. Geophys. Res. 89: 43444358.Google Scholar
Evans, A. G. 1990. Perspective on the development of high-toughness ceramics. J. Amer. Ceramics Soc. 73: 187206.Google Scholar
Evans, A. G., Heuer, A. H., and Porter, D. L 1977. The Fracture Toughness of Ceramics. Canada: Waterloo, pp. 529556.Google Scholar
Evans, B., and Goetze, C. 1979. Temperature variation of hardness of olivine and its implication for polycrystalline yield stress, J. Geophys. Res., 84(NB10): 55055524, doi: 10.1029/JB084iB10p05505.Google Scholar
Evans, J. P. 1990. Thickness displacement relationships for fault zones. J. Struct. Geol. 12(8): 10611065.Google Scholar
Evison, F. 1977. Fluctuations of seismicity before major earthquakes. Nature 266: 710712.Google Scholar
Ewing, M., and Heezen, B. 1956. Some problems of Antarctic submarine geology in Antarctica. In The International Geophysical Year. AGU Geophys. Mono., ed. Crary, A.. Washington, DC: American Geophysical Union, p. 75.Google Scholar
Fan, W. Y., and Shearer, P. M. 2016. Local near instantaneously dynamically triggered aftershocks of large earthquakes. Science 353(6304): 11331136, doi: 10.1126/science.aag0013.Google Scholar
Fattahi, H., and Amelung, F. 2016. InSAR observations of strain accumulation and fault creep along the Chaman Fault system, Pakistan and Afghanistan. Geophys. Res. Lett. 43(16): 83998406, doi: 10.1002/2016gl070121.Google Scholar
Faulkner, D. R., and Rutter, E. H. 2001. Can the maintenance of overpressured fluids in large strike-slip fault zones explain their apparent weakness? Oeology 29: 503506.Google Scholar
Faulkner, D. R., Lewis, A. C., and Rutter, E. H. 2003. On the internal structure and mechanics of large strike-slip fault zones: Field observations of the Carboneras fault in southeastern Spain. Tectonophysics 367: 235251.Google Scholar
Faulkner, D. R., Mitchell, T. M., Behnsen, J., Hirose, T., and Shimamoto, T. 2011. Stuck in the mud? Earthquake nucleation and propagation through accretionary forearcs. Geophys. Res. Lett. 38, doi: 10.1029/2011gl048552.Google Scholar
Faulkner, D. R., Mitchell, T. M., Jensen, E., and Cembrano, J. 2011. Scaling of fault damage zones with displacement and the implications for fault growth processes. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb007788.Google Scholar
Fedotov, S. A. 1965. Regularities in the distribution of strong earthquakes in Kamchatka, the Kuriles, and northeastern Japan. Akad. Nauk USSR Inst. Fiz. Zeml.: Trudy 36: 6695.Google Scholar
Felzer, K. R., Abercrombie, R. E., and Ekstrom, G. 2004. A common origin for aftershocks, foreshocks, and multiplets. Bull. Seismol. Soc. Am. 94(1): 8898, doi: 10.1785/0120030069.Google Scholar
Felzer, K. R., and Brodsky, E. E. 2006. Decay of aftershock density with distance indicates triggering by dynamic stress. Nature 441(7094): 735738, doi: 10.1038/nature04799.Google Scholar
Feng, R., and McEvilly, T. V. 1983. Interpretation of seismic reflection profiling data for the structure of the San Andreas fault zone. Bull. Seismol. Soc. Am. 73: 17011720.Google Scholar
Ferrand, T. P., Hilairet, N., Incel, S., et al. 2017. Dehydration-driven stress transfer triggers intermediate-depth earthquakes. Nat. Commun. 8: 15247, doi: 10.1038/ncomms15247.Google Scholar
Ferri, F., Di Toro, G., Hirose, T., and Shimamoto, T. 2010. Evidence of thermal pressurization in high-velocity friction experiments on smectite-rich gouges. Terra Nova 22(5): 347353, doi: 10.1111/j.1365–3121.2010.00955.x.Google Scholar
Ferrill, D. A., Stamatakos, J. R., and Sims, D. 1999. Normal fault corrugation: Implications for growth and seismicity of active normal faults. J. Struct. Geol. 21(8–9): 10271038, doi: 10.1016/s0191-8141(99)00017–6.Google Scholar
Festa, G., Picozzi, M., Caruso, A., et al. 2018. Performance of Earthquake Early Warning systems during the 2016–2017 Mw 5–6.5 Central Italy Sequence. Seismol. Res. Lett. 89: 112, doi: 10.1785/0220170150.Google Scholar
Fialko, Y. 2004. Evidence of fluid-filled upper crust from observations of postseismic deformation due to the 1992 M(w)7.3 Landers earthquake. J. Geophys. Res.-Solid Earth 109(B8): doi: 10.1029/2004jb002985.Google Scholar
Fialko, Y., and Khazan, Y. 2005. Fusion by earthquake fault friction: Stick or slip? J. Geophys. Res.-Solid Earth 110(B12): doi: 10.1029/2005jb003869.Google Scholar
Field, E. H., Arrowsmith, R. J., Biasi, G. P., et al. 2014. Uniform California earthquake rupture forecast, version 3 (UCERF3) – The time‐independent model. Bull. Seismol. Soc. Am. 104(3): 11221180.Google Scholar
Field, E. H., Biasi, G. P., Bird, P., et al. 2015. Long‐term time‐dependent probabilities for the third Uniform California Earthquake Rupture Forecast (UCERF3). Bull. Seismol. Soc. Am. 105(2A): 511543.Google Scholar
Field, E. H., Jordan, T. H., Jones, L. M., Michael, A. J., Blanpied, M. L., and Workshop, P. 2016. The potential uses of operational earthquake forecasting. Seismol. Res. Lett. 87(2): 313322, doi: 10.1785/0220150174.Google Scholar
Field, E. H., Milner, K. R., Hardebeck, J. L., et al. 2017. A Spatiotemporal Clustering Model for the Third Uniform California Earthquake Rupture Forecast. UCERF3-ETAS): Toward an Operational Earthquake Forecast. Bull. Seismol. Soc. Am. 107(3): 10491081, doi: 10.1785/0120160173.Google Scholar
Fielding, E. J., Lundgren, P. R., Burgmann, R., and Funning, G. J. 2009. Shallow fault-zone dilatancy recovery after the 2003 Bam earthquake in Iran. Nature 458(7234): 6468, doi: 10.1038/nature07817.Google Scholar
Fisher, D. S. 1998. Collective transport in random media: From superconductors to earthquakes. Phys. Rep.-- Review Section of Physics Letters 301(1–3): 113150, doi: 10.1016/s0370-1573(98)00008–8.Google Scholar
Fitch, T. J., and Scholz, C. H. 1971. Mechanism of underthrusting in southwest Japan: A model of convergent plate interactions. J. Geophys. Res. 76: 72607292.Google Scholar
Fleitout, L., and Froidevaux, J. C. J. 1980. Thermal and mechanical evolution of shear zones. J. Struct. Geol. 2: 159164.Google Scholar
Fletcher, J. B. 1982. A comparison between the tectonic stress measured in situ and stress parameters from induced seismicity at Monticello Reservoir, South Carolina. J. Geophys. Res 87: 69316944.Google Scholar
Fletcher, R., and Pollard, D. D. 1981. An anticrack mechanism for stylolites. Geology 9: 419424.Google Scholar
Fleuty, M. J. 1975. Slickensides and slickenlines. Geol. Mag. 112: 319322.Google Scholar
Floyd, J. S., Mutter, J. C., Goodliffe, A. M. et al. 2001. Evidence for fault weakness and fluid flow within an active low-angle normal fault. Nature 411 (6839): 779783.Google Scholar
Forsyth, D. W. 1992. Finite extension and low-angle normal faulting. Geology 20(1): 2730.Google Scholar
Forsyth, D., and Uyeda, S. 1975. On the relative importance of driving forces of plate motion. Geophys. J. R.A.S. 43: 163200.Google Scholar
Fossen, H., and Hesthammer, J. 1997. Geometric analysis and scaling relations of deformation bands in porous sandstone. J. Struct. Geol. 19(12): 14791493, doi: 10.1016/s0191-8141(97)00075–8.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, doi: 10.1144/0016–76492006-036.Google Scholar
Fox, C. G., Matsumoto, H., and Lau, T.-K. A. 2001. Monitoring Pacific Ocean seismicity from autonomous hydrophone array. J. Geophys. Res., 106: 41834206.Google Scholar
Francis, T. J. G. 1981. Serpentinization faults and their role in the tectonics of slow spreading ridges. J. Geophys.l Res. 86(NB12): 16161622, doi: 10.1029/JB086iB12p11616.Google Scholar
Frank, F. C. 1965. On dilatancy in relation to seismic sources. Rev. Geophys. Space Phys. 3: 485503.Google Scholar
Frank, W. B., Radiguet, M., Rousset, B., et al. 2015. Uncovering the geodetic signature of silent slip through repeating earthquakes. Geophys. Res. Lett. 42(8): 27742779, doi: 10.1002/2015gl063685.Google Scholar
Frankel, A. 1991. High-frequency spectral falloff of earthquakes, fractal dimension of complex rupture, b value, and the scaling of strength of faults. J. Geophys. Res. 96: 62916302.Google Scholar
Frankel, A. D. 2004. Rupture process of the M 7.9 Denali fault, Alaska, earthquake: Subevents, directivity, and scaling of high-frequency ground motions. Bull. Seismol. Soc. Amer. 94: S234-S255.Google Scholar
Fraser-Smith, A. C., Bernardi, A., McGill, P. R., Ladd, M. E., Helliwell, R. A., and Villard, O. G. 1990. Low-frequency magnetic-field measurements near the epicenter of the Ms 7.1 Loma-Prieta earthquake. Geophys. Res. Lett. 17: 14651468.Google Scholar
Fraser-Smith, A. C., McGill, P. R., Helliwell, R. A., and Villard, O. G. 1994. Ultra-low frequency magnetic-field measurements in southern California during the Northridge earthquake of 17 January 1994. Geophys. Res. Lett. 21: 21952198.Google Scholar
Freed, A. M. 2005. Earthquake triggering by static, dynamic, and postseismic stress transfer. Ann. Rev. Earth Planet. Sci. 33: 335367, doi: 10.1146/annurev.earth.33.092203.122505.Google Scholar
Freed, A. M., and Burgmann, R. 2004. Evidence of power-law flow in the Mojave desert mantle. Nature 430(6999): 548551, doi: 10.1038/nature02784.Google Scholar
Freed, A. M., and Lin, J. 2001. Delayed triggering of the 1999 Hector Mine earthquake by Viscoelastic stress transfer. Nature 411:180183.Google Scholar
Freed, A. M., and Lin, J. 2002. Accelerated stress buildup on the southern San Andreas fault and surrounding regions caused by Mojave Desert earthquakes. Geology 30: 571574.Google Scholar
Freed, A. M., Burgmann, R., and Herring, T. 2007. Far-reaching transient motions after Mojave earthquakes require broad mantle flow beneath a strong crust. Geophys. Res. Lett. 34(19): doi: 10.1029/2007gl030959.Google Scholar
Freed, A. M., Burgmann, R., Calais, E., and Freymueller, J. 2006b. Stress-dependent power-law flow in the upper mantle following the 2002 Denali, Alaska, earthquake. Earth Planet. Sci. Lett. 252(3–4): 481489, doi: 10.1016/j.epsl.2006.10.011.Google Scholar
Freed, A. M., Burgmann, R., Calais, E., Freymueller, J., and Hreinsdottir, S. 2006a. Implications of deformation following the 2002 Denali, Alaska, earthquake for postseismic relaxation processes and lithospheric rheology. J. Geophys. Res.-Solid Earth 111(B1): doi: 10.1029/2005jb003894.Google Scholar
Freed, A. M., Hirth, G., and Behn, M. D. 2012. Using short-term postseismic displacements to infer the ambient deformation conditions of the upper mantle. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb008562.Google Scholar
Freiman, S. W. 1984. Effects of chemical environments on slow crack growth in glasses and ceramics. J. Geophys. Res. 89: 40724076.Google Scholar
Freund, L. B. 1990. Dynamic Fracture Mechanics. New York: Cambridge University Press.Google Scholar
Friedman, M., Handin, J., and Alani, G. 1972. Fracture-surface energy of rocks. Int. J. Rock Mech. Min. Sci. 9: 757766.Google Scholar
Friedman, M., Logan, J., and Rigert, J. 1974. Glass-indurated quartz gouge in sliding-friction experiments on sandstone. Bull. Geol. Soc. Am. 85: 937942.Google Scholar
Friedrich, A. M., Wernicke, B. P., Niemi, N. A., Bennett, R. A., and Davis, J. L. 2003. Comparison of geodetic and geologic data from the Wasatch region, Utah, and implications for the spectral character of Earth deformation at periods of 10 to 10 million years. J. Geophys. Res.-Solid Earth 108(B4): doi: 10.1029/2001jb000682.Google Scholar
Frohlich, C. 1989. The nature of deep-focus earthquakes. Ann. Rev. Earth Planet. Phys. 17: 227254.Google Scholar
Frohlich, C. 2006. Deep Earthquakes. Cambridge: Cambridge University Press, doi: 10.2277/ 0521828694.Google Scholar
Froment, B., McGuire, J. J., van der Hilst, R. D., et al. 2014. Imaging along-strike variations inmechanical properties of the Gofar transform fault, East Pacific Rise. J. Geophys. Res.-Solid Earth 119(9): 71757194, doi: 10.1002/2014jb011270.Google Scholar
Frye, K. M., and Marone, C. 2002. Effect of humidity on granular friction at room temperature. J. Geophys. Res.-Solid Earth 107(B11): ETG 11-1–ETG 11-13, doi: 10.1029/2001jb000654.Google Scholar
Fukao, Y. 1979. Tsunami earthquakes and subduction processes near deep-sea trenches. J. Geophys. Res. 84: 23032314.Google Scholar
Fulton, P. M., Brodsky, E. E., Kano, Y., et al. 2013. Low Coseismic Friction on the Tohoku-Oki Fault Determined from Temperature Measurements. Science 342(6163): 12141217, doi: 10.1126/science.1243641.Google Scholar
Fulton, P. M., Saffer, D. M., Harris, R. N., and Bekins, B. A. 2004. Re-evaluation of heat flow data near Parkfield, CA: Evidence for a weak San Andreas Fault. Geophys. Res. Lett. 31(15): doi: 10.1029/2003GLØ19378.Google Scholar
Fyfe, W. S., Price, N. J., and Thompson, A. B. 1978. Fluids in the Earth’s Crust. Amsterdam: Elsevier.Google Scholar
Gabuchian, V., Rosakis, A. J., Bhat, H. S., Madariaga, R., and Kanamori, H. 2017. Experimental evidence that thrust earthquake ruptures might open faults. Nature 545: 336338 doi: 10.1038/nature22045.Google Scholar
Galli, P., Galadini, F., and Calzoni, F. 2005. Surface faulting in Norcia (central Italy): A “paleoseismological perspective.” Tectonophysics 403(1–4): 117130, doi: 10.1016/j.tecto.2005.04.003.Google Scholar
Gamond, J. F., and Giraud, A. 1982. Identification des zones de faille a l’aide des associations de fractures de second ordre. Bull. Soc. Geol. France 24: 755762.Google Scholar
Gan, W., Svarc, L., Savage, J. C., and Prescott, W. H. 2000. Strain accumulation across the eastern California shear zone at latitude 36° 30′. N. J. Geophys. Res. 105: 16,22916,236.Google Scholar
Gao, S. S., Silver, P. G., Linde, A. T., and Sacks, I. S. 2000. Annual modulation of triggered seismicity following the 1992 Landers earthquake in California. Nature 406: 500504.Google Scholar
Gao, X., and Wang, K. 2014. Strength of stick-slip and creeping subduction megathrusts from heat flow observations. Science 345(6200): 10381041.Google Scholar
Gao, X., and Wang, K. L. 2017. Rheological separation of the megathrust seismogenic zone and episodic tremor and slip. Nature 543(7645): 416+, doi: 10.1038/nature21389.Google Scholar
Garfunkel, Z., and Ron, H. 1985. Block Rotation and Deformation by Strike-Slip Faults. 2. the Properties of a Type of Macroscopic Discontinuous Deformation. J. Geophys. Res.-Solid Earth and Planets 90. NB10): 85898602.Google Scholar
Gawthorpe, R. L., Jackson, C. A. L., Young, M. J., Sharp, I. R., Moustafa, A. R., and Leppard, C. W. 2003. Normal fault growth, displacement localisation and the evolution of normal fault populations: The Hammam Faraun fault block, Suez rift, Egypt. J. Struct. Geol. 25: 883895.Google Scholar
Geller, R. J., 1997. Earthquake prediction: A critical review. Geophys. J. Int. 131: 425450.Google Scholar
Geller, R. J. Jackson, D. D., Kagan, Y. Y., and Mulargia, F. 1997. Geoscience – Earthquakes cannot be predicted. Science 275: 16161617.Google Scholar
Gephardt, J. W., and Forsyth, D. W. 1984. An improved method for determining the regional stress tensor using earthquake focal mechanism data. J. Geophys. Res. 89: 93059320.Google Scholar
Gershenzon, N. I., Bambakidis, G., and Skinner, T. E. 2016. Sine-Gordon modulation solutions: Application to macroscopic non-lubricant friction. Physica D-Nonlinear Phenomena 333: 285292, doi: 10.1016/j.physd.2016.01.004.Google Scholar
Gerstenberger, M. C., Wiemer, S., Jones, L. M., and Reasenberg, P. A. 2005. Real-time forecasts of tomorrow’s earthquakes in California. Nature 435(7040): 328331, doi: 10.1038/nature03622.Google Scholar
Gerstenberger, M., McVerry, G., Rhoades, D., and Stirling, M. 2014. Seismic hazard modeling for the recovery of Christchurch. Earthquake Spectra 30(1): 1729, doi: 10.1193/021913eqs037m.Google Scholar
Ghosh, A., Vidale, J. E., Peng, Z. G., Creager, K. C., and Houston, H. 2009a. Complex nonvolcanic tremor near Parkfield, California, triggered by the great 2004 Sumatra earthquake. J. Geophys. Res. 114, doi: 10.1029/2008JB006062.Google Scholar
Ghosh, A., Vidale, J. E., Sweet, J. R., Creager, K. C., Wech, A. G., and Houston, H. 2010. Tremor bands sweep Cascadia. Geophys. Res. Lett. 37, doi: 10.1029/2009gl042301.Google Scholar
Gilbert, G. K. 1884. A theory of the earthquakes of the Great Basin, with a practical application. Am. J. Sci. xxvii: 4954.Google Scholar
Gilbert, G. K. 1909. Earthquake forecasts. Science XXIX: 121138.Google Scholar
Gilbert, L. E., and Malinverno, A. 1988. A characterization of the spectral density of residual ocean floor topography. Geophys. Res. Lett. 15(12): 14011404.Google Scholar
Gilbert, L., Scholz, C. H., and Beavan, J. 1994. Strain localization along the San Andreas fault: consequences for loading mechanisms. J. Geophys. Res. 99: 975984.Google Scholar
Giorgetti, C., Carpenter, B. M., and Collettini, C. 2015. Frictional behavior of talc-calcite mixtures. J. Geophys. Res.-Solid Earth 120(9): 66146633, doi: 10.1002/2015jb011970.Google Scholar
Gleason, G. C., and Green, H. W. 2009. A general test of the hypothesis that transformation-induced faulting cannot occur in the lower mantle. Phys. Earth Planet. Int. 172(1–2): 91103, doi: 10.1016/j.pepi.2008.06.019.Google Scholar
Goebel, T. H. W., Kwiatek, G., Becker, T. W., Brodsky, E. E., and Dresen, G. 2017. What allows seismic events to grow big?: Insights from b-value and fault roughness analysis in laboratory stick-slip experiments. Geology 45(9): 815818, doi: 10.1130/g39147.1.Google Scholar
Goebel, T. H. W., Schorlemmer, D., Becker, T. W., Dresen, G., and Sammis, C. G. 2013. Acoustic emissions document stress changes over many seismic cycles in stick-slip experiments. Geophys. Res. Lett. 40: 20492054, doi: 10.1002/grl.50507.Google Scholar
Goebel, T. H. W., Weingarten, M., Chen, X., Haffener, J., and Brodsky, E. E. 2017. The 2016 Mw5.1 Fairview, Oklahoma earthquakes: Evidence for long-range poroelastic triggering at > 40 km from fluid disposal wells. Earth Planet. Sci. Lett. 472: 5061, doi: 10.1016/j.epsl.2017.05.011.Google Scholar
Goetze, C. 1971. High temperature rheology of Westerly granite. J. Geophys. Res. 76: 12231230.Google Scholar
Goetze, C., and Evans, B. 1979. Stress and temperature in the bending lithosphere as constrained by experimental rock mechanics. Geophysical Journal R. Astron. Soc. 59: 463478.Google Scholar
Gold, R. D., and Cowgill, E. 2011. Deriving fault-slip histories to test for secular variation in slip, with examples from the Kunlun and Awatere faults. Earth Planet. Sci. Lett. 301(1–2): 5264, doi: 10.1016/j.epsl.2010.10.011.Google Scholar
Goldfinger, C., Ikeda, Y., Yeats, R. S., and Ren, J. J. 2013. Superquakes and Supercycles. Seismol. Res. Lett. 84(1): 2432, doi: 10.1785/0220110135.Google Scholar
Goldfinger, C., Nelson, C. H., and Johnson, J. E. 2003. Holocene earthquake records from the Cascadia subduction zone and northern San Andreas fault based on precise dating of offshore turbidites. Ann. Rev. Earth. Planet Sci. 31: 555578.Google Scholar
Goldfinger, C., Nelson, C. H., Morey, A. E., et al. 2012. Turbidite Event History – Methods and Implications for Holocene Paleoseismicity of the Cascadia Subduction Zone, U. S. Geol. Surv. Prof. Pap. 1661-F: 170.Google Scholar
Goldsby, D. L., and Tullis, T. E. 2002. Low frictional strength of quartz rocks at subseismic slip rates. Geophys. Res. Lett. 29(17): 25-1–25-4, doi: 10.1029/2002gl015240.Google Scholar
Goldsby, D. L., and Tullis, T. E. 2011. Flash Heating Leads to Low Frictional Strength of Crustal Rocks at Earthquake Slip Rates. Science 334(6053): 216218, doi: 10.1126/science.1207902.Google Scholar
Gomberg, J. 2001. The failure of earthquake failure models, J. Geophys. Res.-Solid Earth, 106(B8): 1625316263, doi: 10.1029/2000jb000003.Google Scholar
Gomberg, J. 2010. Slow-slip phenomena in Cascadia from 2007 and beyond: A review. Geol. Soc. Amer. Bull. 122(7–8): 963978.Google Scholar
Gomberg, J. 2018. Unsettled earthquake nucleation. Nature Geoscience 11(7): 463464, doi: 10.1038/s41561-018-0149-x.Google Scholar
Gomberg, J., and Bodin, P. 1994. Triggering of the Ms = 5.4 Little Skull Mountain, Nevada, earthquake with dynamic strains. Bull. Seismol. Soc. Amer. 84: 844853.Google Scholar
Gomberg, J., Beeler, N. M., Blanpied, M. L., and Bodin, P. 1998. Earthquake triggering by transient and static deformations. J. Geophys. Res.-Solid Earth 103: 2441124426.Google Scholar
Gomberg, J., Beeler, N., and Blanpied, M. 2000. On rate-state and Coulomb failure models. J. Geophys. Res.-Solid Earth 105: 78577871.Google Scholar
Gomberg, J., Blanpied, M. L., and Beeler, N. M. 1997. Transient triggering of near and distant earthquakes. Bull. Seismol. Soc. Amer. 87: 294309.Google Scholar
Gomberg, J., Bodin, P., Larson, K., and Dragert, H. 2004. Earthquake nucleation by transient deformations caused by the M=7.9 Denali, Alaska, earthquake. Nature 427(6975): 621624, doi: 10.1038/nature02335.Google Scholar
Gomberg, J., Reasenberg, P. A., Bodin, P., and Harris, R. A. 2001. Earthquake triggering by seismic waves following the Landers and Hector Mine earthquakes. Nature 411: 462466.Google Scholar
Gomberg, J., Rubenstein, J. L., Peng, Z., Creager, K. C., Vidale, J. E., and Bodin, P. 2008. Widespread Triggering of Nonvolcanic Tremor in California. Science 319: 173.Google Scholar
González, G., Salazar, P., Loveless, J. P., Allmendinger, R. W., Aron, F., and Shrivastava, M. 2015. Upper plate reverse fault reactivation and the unclamping of the megathrust during the 2014 northern Chile earthquake sequence. Geology 43(8): 671674.Google Scholar
Goodier, J. N. 1968. Mathematical theory of equilibrium cracks. In Fracture V.II, ed. Liebowitz, H., New York: Academic, pp. 166.Google Scholar
Goodier, J. N. and Field, F. A. 1963. Plastic energy dissipation in crack propagation. In Fracture of Solids, eds. Drucker, D. C. and Gilman, J. J.. New York: Wiley, pp. 103118.Google Scholar
Gordon, F., and Lewis, J. 1980. The Meckering and Caligari earthquakes of October 1968 and March, 1970. Geol Surv. Western Australia Bull 126: https://d28rz98at9flks.cloudfront.net/12254/Rec1968_142.pdf.Google Scholar
Gordon, R. G. 1998. The plate tectonic approximation: Plate nonrigidity, diffuse plate boundaries, and global plate reconstructions. Ann. Rev. Earth Planet. Sci. 26(1): 615642.Google Scholar
Gorgun, E. 2013. Analysis of the b-values before and after the 23 October 2011 M-w 7.2 Van-Ercis, Turkey earthquake. Tectonophysics 603: 213221, doi: 10.1016/j.tecto.2013.05.030.Google Scholar
Gough, D. I., Fordjor, C. K., and Bell, J. S. 1983. A stress province boundary and tractions on the North American plate. Nature 305: 619621.Google Scholar
Govers, R., Furlong, K. P., van de Wiel, L., Herman, M. W., and Broerse, T. 2017. The geodetic signature of the earthquake cycle at subduction zones: Model constraints on the deep processes. Rev. Geophys. 55, doi: 10.1002/2017RG000589.Google Scholar
Granier, T. 1985. Origin, damping and pattern of development of faults in granite. Tectonics 4: 721737.Google Scholar
Grant, L. B., and Sieh, K. 1994. Paleoseismic evidence of clustered earthquakes on the San Andreas fault in the Carrizo Plain, California. J. Geophys. Res.-Solid Earth 99(B4): 68196841.Google Scholar
Grapenthin, R., Johanson, I. A., and Allen, R. M. 2014. Operational real‐time GPS‐enhanced earthquake early warning. J. Geophys. Res.-Solid Earth 119(10): 79447965.Google Scholar
Grapes, R. H. 1995. Uplift and exhumation of Alpine Schist, Southern Alps, New Zealand: Thermobarometric constraints. N. Z. J. Geol. Geophys. 38: 525533.Google Scholar
Grapes, R. H., and Downes, G. L. 2010. Charles Lyell and the great 1855 earthquake in New Zealand: First recognition of active fault tectonics. J. Geol. Soc. London 167(1): 3547, doi: 10.1144/0016–76492009-104.Google Scholar
Grapes, R. H., Sissons, B. A., and Wellman, H. W. 1987. Widening of the Taupo volcanic zone, New Zealand and the Edgecumbe earthquake of March, 1987. Geology 15: 11231125.Google Scholar
Grasso, J. R., and Sornette, D. 1998. Testing self‐organized criticality by induced seismicity. J. Geophys. Res.-Solid Earth 103(B12): 2996529987.Google Scholar
Green, H., Scholz, C., Tingle, T., Young, T., and Koczynski, T. 1992. Acoustic emissions produced by anticrack faulting during the olivine→ spinel transformation. Geophys. Res. Lett. 19(8): 789792.Google Scholar
Green, H. W. 2007. Shearing instabilities accompanying high-pressure phase transformations and the mechanics of deep earthquakes. Proc. Nat. Acad. Sci. USA 104(22): 91339138, doi: 10.1073/pnas.0608045104.Google Scholar
Green, H. W., and Burnley, P. C. 1989. A new self-organizing mechanism for deep-focus earthquakes. Nature 341: 733737.Google Scholar
Green, H. W., and Houston, H. 1995. The mechanics of deep earthquakes. Ann. Rev. Earth Planet. Sci. 23: 169213.Google Scholar
Green, H. W., Chen, W. P., and Brudzinski, M. R. 2010. Seismic evidence of negligible water carried below 400-km depth in subducting lithosphere. Nature 467(7317): 828831, doi: 10.1038/nature09401.Google Scholar
Green, H. W., Shi, F., Bozhilov, K., Xia, G., and Reches, Z. 2015. Phase transformation and nanometric flow cause extreme weakening during fault slip. Nature Geoscience 8: 484489.Google Scholar
Green, H. W., 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
Greenwood, J. A., and Williamson, J. B. P. 1966. Contact of nominally flat surfaces. J. Proc. Roy. Soc. London 295: 300319.Google Scholar
Gretener, P. E. 1977. On the character of thrust sheets with particular reference to the basal tongues. Bull. Can. Pet. Geol. 25: 110122.Google Scholar
Griffith, A. A. 1920. The phenomena of rupture and flow in solids. Trans. Roy. Soc. Phil. Ser. A 221: 163198.Google Scholar
Griffith, A. A. 1924. The theory of rupture. In Proc. Ist. Int. Congr. Appl. Mech., eds. Biezeno, C. B. and Burgers, J. M.. Delft: Tech. Boekhandel en Drukkerij J. Walter Jr., pp. 5463.Google Scholar
Griggs, D. T., and Blacic, J. D. 1965. Quartz-anomalous weakness of synthetic crystals. Science 147: 292295.Google Scholar
Grocott, J. 1981. Fracture geometry of pseudotachylyte generation zones: A study of shear fractures formed during seismic events. J. Struct. Geol. 3: 169178.Google Scholar
Gu, J. C. 1984. Frictional resistance to accelerating slip. Pageoph 122: 662679.Google Scholar
Gu, J. C., Rice, J. R., Ruina, A. L., and Tse, S. T. 1984. Slip motion and stability of a single degree of freedom elastic system with rate and state dependent friction. J. Mech. Phys. Sol. 32: 167196.Google Scholar
Gudmundsson, G. H. 2006. Fortnightly variations in the flow velocity of Rutford Ice Stream, West Antarctica. Nature 444(7122): 10631064, doi: 10.1038/nature05430.Google Scholar
Guidoboni, E., and Valensise, G. 2015. On the complexity of earthquake sequences: A historical seismology perspective based on the L’Aquila seismicity (Abruzzo, Central Italy), 1315–1915. Earthquakes and Structures 8(1): 153184.Google Scholar
Gulia, L., and Wiemer, S. 2010. The influence of tectonic regimes on the earthquake size distribution: A case study for Italy. Geophys. Res. Lett. 37: 10305, doi: 10.1029/2010GL043066.Google Scholar
Gulia, L., Tormann, T., Wiemer, S., Herrmann, M., and Seif, S. 2016. Short-term probabilistic earthquake risk assessment considering time-dependent b values. Geophys. Res. Lett. 43(3): 11001108, doi: 10.1002/2015gl066686.Google Scholar
Gupta, A., and Scholz, C. H. 1998. Utility of elastic models in predicting fault displacement fields. J. Geophys. Res. 103: 823834.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: 10871090.Google Scholar
Gupta, H. K. 2002. A review of recent studies of triggered earthquakes by artificial water reservoirs with special emphasis on earthquakes in Koyna, India. Earth-Science Reviews 58(3–4): 279310, doi: 10.1016/s0012-8252(02)00063–6.Google Scholar
Gusev, A. A. 2013. High-frequency radiation from an earthquake fault: A review and a hypothesis of fractal rupture front geometry. Pageoph 170: 6593, doi: 10.1007/s00024-012–0455-y.Google Scholar
Habermann, R. E. 1981. Precursory seismicity patterns: Stalking the mature seismic gap. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 2942.Google Scholar
Habermann, R. E. 1988. Precursory seismic quiescence: Past, present, and future. Pageoph. 126: 277318.Google Scholar
Hacker, B. R., Peacock, S. M., Abers, G. A., and Holloway, S. D. 2003. Subduction factory – 2. Are intermediate-depth earthquakes in subducting slabs linked to metamorphic dehydration reactions? J. Geophys. Res.-Solid Earth. 108(B1): doi: 10.1029/2001jb001129.Google Scholar
Hadley, K. 1973. Laboratory investigation of dilatancy and motion of fault surfaces at low confining pressures. In Proc. Conf. Tect. Problems San Andreas Fault System. Publ. Geol. Sci. vol. XIII, eds. Kovach, R. and Nur, A.. Stanford: Stanford University, pp. 427435.Google Scholar
Hadley, K. J. 1975. Azimuthal variation of dilatancy. J. Geophys. Res. 80: 48454850.Google Scholar
Haeussler, P. J., Schwartz, D. P., Dawson, T. E., et al. 2004. Surface rupture and slip distribution of the Denali and Totschunda faults in the 3 November 2002 M 7.9 earthquake, Alaska. Bull. Seismol. Soc. Am. 94(6): S23S52.Google Scholar
Hafner, W. 1951. Stress distributions and faulting. Bull. Geol. Soc. Am. 62: 373398.Google Scholar
Hagiwara, Y. 1974. Probability of earthquake occurrence as obtained from a Weibull distribution analysis of crustal strain. Tectonophysics 23: 313318.Google Scholar
Hagstrum, J. T., Atwater, B. F., and Sherrod, B. L. 2004, Paleomagnetic correlation of late Holocene earthquakes among estuaries in Washington and Oregon. Geochem. Geophys. Geosystems 5, doi: 10.1029/2004GC000736.Google Scholar
Haines, A. J., and Holt, W. E. 1993. A procedure for obtaining the complete horizontal motions within zones of distributed deformation from the inversion of strain-rate data. J. Geophys. Res.-Solid Earth 98: 1205712082.Google Scholar
Hall, S. S. 2011. Scientists on trial: At fault? Nature 477: 264269, doi: 10.1038/477264a.Google Scholar
Han, J. G., Vidale, J. E., Houston, H., Chao, K., and Obara, K. 2014. Triggering of tremor and inferred slow slip by small earthquakes at the Nankai subduction zone in southwest Japan. Geophys. Res. Lett. 41(22): 80538060, doi: 10.1002/2014gl061898.Google Scholar
Han, R., Hirose, T., Shimamoto, T., Lee, Y., and Ando, J. 2011. Granular nanoparticles lubricate faults during seismic slip. Geology 39(6): 599602, doi: 10.1130/g31842.1.Google Scholar
Han, R., Shimamoto, T., Hirose, T., Ree, J.-H., and Ando, J. 2007. Ultralow friction of carbonate faults caused by thermal decomposition. Science 316: 878881.Google Scholar
Hanks, T. C. 1971. Kuril trench-Hokkaido rise system – Large shallow earthquakes and simple models of deformation. Geophys. J. Roy. Astron. Soc. 23(2): 173-&, doi: 10.1111/j.1365-246X.1971.tb01811.x.Google Scholar
Hanks, T. C. 1977. Earthquake stress drops, ambient tectonic stresses and stresses that drive plate motions. Pure Appl. Geophys. 115: 441458.Google Scholar
Hanks, T. C. 1979. b values and ω−γ seismic source models: Implications for tectonic stress variations along active crustal fault zones and the estimation of high frequency strong ground motion. J. Geophys. Res. 84: 22352242.Google Scholar
Hanks, T. C. 1982. fmax. Bull. Seismol. Soc. Am. 72: 18671880.Google Scholar
Hanks, T. C., and Bakun, W. H. 2002. A bilinear source-scaling model for M-log A observations of continental earthquakes. Bull. Seismol. Soc. Am. 92(5): 18411846, doi: 10.1785/0120010148.Google Scholar
Hanks, T. C., and Bakun, W. H. 2008. M-log A observations for recent large earthquakes. Bull. Seismol. Soc. Am. 98(1): 490494, doi: 10.1785/0120070174.Google Scholar
Hanks, T. C., and Bakun, W. H. 2014. M-log A Models and Other Curiosities. Bull. Seismol. Soc. Am. 104(5): 26042610, doi: 10.1785/0120130163.Google Scholar
Hanks, T. C., and Johnson, D. A. 1976. Geophysical assessment of peak accelerations. Bull. Seismol. Soc. Am. 66: 959968.Google Scholar
Hanks, T. C., and Raleigh, C. B. 1980. Stress in the lithosphere. J. Geophys. Res. 85: 60836435.Google Scholar
Hanks, T., and Kanamori, H. 1979. A moment-magnitude scale. J. Geophys. Res. 84: 23482352.Google Scholar
Hanks, T., and McGuire, R. 1981. The character of high-frequency strong ground motion. Bull. Seismol. Soc. Am. 71: 20712095.Google Scholar
Hanks, T., and Schwartz, D. 1987. Morphological dating of the pre-1983 fault scarp on the Lost River fault at Doublesprings Pass road, Custer County, Idaho. Bull. Seismol. Soc. Am. 77: 837846.Google Scholar
Hanmer, S. 1988. Great Slave Lake shear zone, Canadian shield – Reconstructed vertical profile of a crustal-scale fault zone. Tectonophysics 149(3–4): 245264.Google Scholar
Hanmer, S., Williams, M., and Kopf, C. 1995. Modest movements, spectacular fabrics in an intracontinental deep-crustal strike-slip-fault: Striding–Athabasca mylonite Zone, NW Canadian Shield. J. Struct. Geol. 17: 493507.Google Scholar
Hardebeck, J. L., and Hauksson, E. 1999. Role of fluids in faulting inferred from stress field signatures. Science 285: 236239.Google Scholar
Hardebeck, J. L., and Michael, A. J. 2004. Stress orientations at intermediate angles to the San Andreas Fault, California. J. Geophys. Res.-Solid Earth 109(B11): doi: 10.1029/2004jb003239.Google Scholar
Harris, R. A. 1998a. Forecasts of the 1989 Loma Prieta, California, earthquake. Bull. Seismol. Soc. Amer. 88: 898916.Google Scholar
Harris, R. A. 1998b. Introduction to special section: Stress triggers, stress shadows, and implications for seismic hazard. J. Geophys. Res.-Solid Earth 103: 2434724358.Google Scholar
Harris, R. A. 2017. Large earthquakes and creeping faults. Rev. Geophys. 55: 169198, doi: 10.1002/201/2016RG000539.Google Scholar
Harris, R. A., and Day, S. M. 1993. Dynamics of fault interaction: Parallel strike-slip faults. J. Geophys. Res. 98: 4461–72.Google Scholar
Harris, R. A., and Day, S. M. 1999. Dynamic 3D simulations of earthquakes on en echelon faults. Geophys. Res. Lett. 26(14): 20892092, doi: 10.1029/1999gl900377.Google Scholar
Harris, R. A., and Segall, P. 1987. Detection of a locked zone at depth on the Parkfield, California, segment of the San Andreas fault. J. Geophys. Res. 92: 79457962.Google Scholar
Harris, R. A., and Simpson, R. W. 1996. In the shadow of 1857 – The effect of the great Ft Tejon earthquake on subsequent earthquakes in southern California. Geophys. Res. Lett. 23: 229232.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: 2443924451.Google Scholar
Harris, R. A., and Simpson, R. W. 2002. The 1999 M-w 7.1 Hector Mine, California, earthquake: A test of the stress shadow hypothesis? Bull. Seismol. Soc. Am. 92(4): 14971512, doi: 10.1785/0120000913.Google Scholar
Harrison, T. M., Grove, M., Lovera, O. M., and Catlos, E. J. 1998. A model for the origin of Himalayan anatexis and inverted metamorphism. J. Geophys. Res.-Solid Earth 103: 2701727032.Google Scholar
Hartleb, R. D., Dolan, J. F., Kozaci, O., Akyuz, H. S., and Seitz, G. G. 2006. A 2500-yr-long paleoseismologic record of large, infrequent earthquakes on the North Anatolian fault at Cukurcimen, Turkey. Geol. Soc. Amer. Bull. 118: 823840.Google Scholar
Hartzell, S., and Heaton, T. 1989. The fortnightly tide and the tidal triggering of earthquakes. Bull. Seismol. Soc. Am. 79(4): 12821286.Google Scholar
Hasegawa, A., and Nakajima, J. 2017. Seismic imaging of slab metamorphism and genesis of intermediate-depth intraslab earthquakes. Progress in Earth and Planetary Science 4(1): 12.Google Scholar
Hasegawa, A., Umino, N., and Takagi, A. 1978. Double-planed structure of the deep seismic zone in the northeastern Japan Arc. Tectonophysics 47: 4358.Google Scholar
Hashimoto, C., Noda, A., Sagiya, T., and Matsu’ura, M. 2009. Interplate seismogenic zones along the Kuril-Japan trench inferred from GPS data inversion. Nat. Geosci. 2(2): 141144.Google Scholar
Haskell, N. 1964. Total energy and energy spectral density of elastic wave radiation from propagating faults. Bull. Seismol. Soc. Am. 54: 18111842.Google Scholar
Hauksson, E., Jones, L. M., and Hutton, K. 1995. The 1994 Northridge earthquake sequence in California – Seismological and tectonic aspects. J. Geophys. Res.-Solid Earth 100: 1233512355.Google Scholar
Hauksson, E., Jones, L. M., Hutton, K., and Eberhartphillips, D. 1993. The 1992 Landers earthquake sequence – Seismological observations. J. Geophys. Res.-Solid Earth 98: 1983519858.Google Scholar
Hauksson, E. Jones, L. M., Davis, T. L., et al. 1988. The 1987 Whittier Narrows earthquake in the Los Angeles metropolitan area, California. Science 239: 14091412.Google Scholar
Hayward, N. J., and Ebinger, C. J. 1996. Variation in the along-axis segmentation of the Afar rift system. Tectonics 15: 244257.Google Scholar
Hazzard, J. F., Young, R. P., and Maxwell, S. C. 2000. Micromechanical modeling of cracking and failure in brittle rocks. J. Geophys. Res. 105: 16,68316,698.Google Scholar
He, C. R., Wang, Z. L., and Yao, W. M. 2007. Frictional sliding of gabbro gouge under hydrotherynal conditions. Tectonophysics 445(3–4): 353362, doi: 10.1016/j.tecto.2007.09.008.Google Scholar
Heap, M. J., Baud, P., Meredith, P. G., Bell, A. F., and Main, I. G. 2009. Time-dependent brittle creep in Darley Dale sandstone. J. Geophys. Res. 114, doi: 10.1029/2008JB006212.Google Scholar
Hearn, E. H., and Thatcher, W. R. 2015. Reconciling viscoelastic models of postseismic and interseismic deformation: Effects of viscous shear zones and finite length ruptures. J. Geophys. Res.-Solid Earth 120(4): 27942819, doi: 10.1002/2014jb011361.Google Scholar
Hearn, E. H., McClusky, S., Ergintav, S., and Reilinger, R. E. 2009. Izmit earthquake postseismic deformation and dynamics of the North Anatolian Fault Zone. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb006026.Google Scholar
Heaton, T. H. 1990. Evidence for and implications of self-healing pulses of slip in earthquake rupture. Phys. Earth Planet. Int. 64: 120.Google Scholar
Heinrich, P., Piatanesi, A., Okal, E., and Hebert, H. 2000. Near-field modeling of the July 17, 1998 tsunami in Papua New Guinea. Geophys. Res. Lett. 27: 30373040.Google Scholar
Heki, K. 2004. Space geodetic observation of deep basal subduction erosion in northeastern Japan. Earth Planet. Sci. Lett. 219(1–2): 1320, doi: 10.1016/s0012-821x(03)00693–9.Google Scholar
Heki, K., and Enomoto, Y. 2015. Mw dependence of the preseismic ionospheric electron enhancements. J. Geophys. Res.: Space Physics 120(8): 70067020.Google Scholar
Heki, K., and Miyazaki, S. 2001. Plate convergence and long-term crustal deformation in Central Japan. Geophys. Res. Lett. 28(12): 23132316, doi: 10.1029/2000gl012537.Google Scholar
Helmstetter, A., and Sornette, D. 2002. Subcritical and supercritical regimes in epidemic models of earthquake aftershocks. J. Geophys. Res.-Solid Earth 107(B10): ESE 10-1–ESE 10-21, doi: 10.1029/2001jb0001580.Google Scholar
Helmstetter, A., and Sornette, D. 2003a. Bath’s law derived from the Gutenberg-Richter law and from aftershock properties. Geophys. Res. Lett. 30(20): doi: 10.1029/2003gl018186.Google Scholar
Helmstetter, A., and Sornette, D. 2003b. Foreshocks explained by cascades of triggered seismicity. J. Geophys. Res.-Solid Earth 108(B10): doi: 10.1029/2003jb002409.Google Scholar
Helmstetter, A., Sornette, D., and Grasso, J. R. 2003. Mainshocks are aftershocks of conditional foreshocks: How do foreshock statistical properties emerge from aftershock laws. J. Geophys. Res.-Solid Earth 108(B1): 2046, doi: 10.1029/2002jb001991.Google Scholar
Hemelsdael, R., and Ford, M. 2016. Relay zone evolution: A history of repeated fault propagation and linkage, central Corinth rift, Greece. Basin Research 28(1): 3456, doi: 10.1111/bre.12101.Google Scholar
Henry, C., Das, S., and Woodhouse, J. H. 2000. The great March 25, 1998, Antarctic Plate earthquake: Moment tensor and rupture history. J. Geophys. Res.-Solid Earth 105: 1609716118.Google Scholar
Henstock, T. J., and Levander, A. 2000. Lithospheric evolution in the wake of the Mendocino triple junction: Structure of the San Andreas Fault system at 2 Ma. Geophys. J. Int. 140: 233247.Google Scholar
Henstock, T. J., Levander, A., and Hole, J. A. 1997. Deformation in the lower crust of the San Andreas fault system in northern California. Science 278: 650653.Google Scholar
Herrendorfer, R., van Dinther, Y., Gerya, T., and Dalguer, L. A. 2015. Earthquake supercycle in subduction zones controlled by the width of the seismogenic zone. Nature Geoscience 8(6): 471473, doi: 10.1038/ngeo2427.Google Scholar
Herz, A. V. M., and Hopfield, J. J. 1995. Earthquake cycles and neural reverberations – Collective oscillations in systems with pulse-coupled threshold elements. Phys. Rev. Lett. 75: 12221225.Google Scholar
Heslot, F., Baumberger, T., Perrin, B., Caroli, B., and Caroli, C. 1994. Creep, stick-slip, and dry-friction dynamics – Experiments and a heuristic model. Phys. Rev. E 49: 49734988.Google Scholar
Heuret, A., Lallemand, S., Funiciello, F., Piromallo, C., and Faccenna, C. 2011. Physical characteristics of subduction interface type seismogenic zones revisited. Geochem. Geophys. Geosystems 12, doi: 10.1029/2010gc003230.Google Scholar
Hickman, S., and Zoback, M. 2004. Stress orientations and magnitudes in the SAFOD pilot hole. Geophys. Res. Lett. 31(15): doi.org/10.1029/2004GL020043.Google Scholar
Higgs, N. G. 1981. Mechanical properties of ultrafine quartz, chlorite, and bentonite in environments appropriate to upper-crustal earthquakes. PhD, Texas AandM.Google Scholar
Hill, D. P., Reasenberg, P. A., Michael, A., et al. 1993. Seismicity remotely triggered by the magnitude 7.3 Landers, California, earthquake. Science 260: 16171623.Google Scholar
Hill, E. M., Barbot, S., Lay, T., et al. 2015. The 2012 M(w)8.6 Wharton Basin sequence: A cascade of great earthquakes generated by near-orthogonal, young, oceanic mantle faults. J. Geophys. Res.-Solid Earth 120(5): 37233747, doi: 10.1002/2014jb011703.Google Scholar
Hino, R., Inazu, D., Ohta, Y., et al. 2014. Was the 2011 Tohoku-Oki earthquake preceded by aseismic preslip? Examination of seafloor vertical deformation data near the epicenter. Marine Geophysical Research 35(3): 181190, doi: 10.1007/s11001-013–9208-2.Google Scholar
Hirata, T. 1989. Fractal dimension of fault systems in Japan: Fractal structure in rock fracture geometry at various scales. Pageoph 131: 157170.Google Scholar
Hirose, T., and Shimamoto, T. 2005. Growth of molten zone as a mechanism of slip weakening of simulated faults in gabbro during frictional melting. J. Geophys. Res. 110: B05202.Google Scholar
Hirose, T., Mizoguchi, K., and Shimamoto, T. 2012. Wear processes in rocks at slow to high slip rates. J. Struct. Geol. 38: 102116, doi: 10.1016/j.jsg.2011.12.007.Google Scholar
Hirth, G., and Tullis, J. 1992. Dislocation creep regimes in quartz aggregates. J. Struct. Geol. 14: 145159.Google Scholar
Hirth, G., and Tullis, J. 1994. The brittle-plastic transition in experimentally deformed quartz aggregates. J. Geophys. Res.-Solid Earth 99(B6): 1173111747, doi: 10.1029/93jb02873.Google Scholar
Hirth, G., Teyssier, C., and Dunlap, W. J. 2001. An evaluation of quartzite flow laws based on comparisons between experimentally and naturally deformed rocks. Int. J. Earth Sciences 90(1): 7787, doi: 10.1007/s005310000152.Google Scholar
Hjörleifsdóttir, V., Kanamori, H., and Tromp, J. 2009. Modeling 3‐D wave propagation and finite slip for the 1998 Balleny Islands earthquake. J. Geophys. Res.-Solid Earth 114(B3): B03301, doi.org/10.1029/2008JB005975.Google Scholar
Hobbs, B. E., Ord, A., and Teyssier, C. 1986. Earthquakes in the ductile regime? Pageoph 124: 309336.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.-Solid Earth 101: 2545925471.Google Scholar
Holdsworth, R. E. 2004. Weak faults–rotten cores. Science 303(5655): 181182.Google Scholar
Hollingsworth, J., Ye, L., and Avouac, J. P. 2017. Dynamically triggered slip on a splay fault in the Mw 7.8, 2016 Kaikoura (New Zealand) earthquake. Geophys. Res. Lett. 44: 35173525, doi: 10.1002/2016GL072228.Google Scholar
Holt, W. E., Chamot-Rooke, N., Le Pichon, X., Haines, A. J., Shen-Tu, B., and Ren, J. 2000. Velocity field in Asia inferred from Quaternary fault slip rates and Global Positioning System observations. J. Geophys. Res.-Solid Earth 105: 1918519209.Google Scholar
Horowitz, F., and Ruina, A. 1985. Frictional slip patterns generated in a spatially homogeneous elastic fault model. Eos 66: 1027910298.Google Scholar
Houston, H. 2015. Low friction and fault weakening revealed by rising sensitivity of tremor to tidal stress. Nature Geoscience 8(5): 409+, doi: 10.1038/ngeo2419.Google Scholar
Houston, H., Benz, H. M., and Vidale, J. E. 1998. Time functions of deep earthquakes from broadband and short-period stacks. J. Geophys. Res.-Solid Earth 103: 2989529913.Google Scholar
Hreinsdottir, S., and Bennett, R. A. 2009. Active aseismic creep on the Alto Tiberina low-angle normal fault, Italy. Geology 37(8): 683686.Google Scholar
Hreinsdottir, S., Freymueller, J. T., Fletcher, H. J., Larsen, C. F., and Burgmann, R. 2003. Coseismic slip distribution of the 2002 M(w)7.9 Denali fault earthquake, Alaska, determined from GPS measurements. Geophys. Res. Lett. 30(13): doi: 10.1029/2003gl017447.Google Scholar
Hsieh, P. A., and Bredehoeft, J. D. 1981. A reservoir analysis of the Denver earthquakes – A case of induced seismicity. J. Geophys. Res 86(NB2): 903920, doi: 10.1029/JB086iB02p00903.Google Scholar
Hsu, Y. J., Bechor, N., Segall, P., Yu, S. B., Kuo, L. C., and Ma, K. F. 2002. Rapid afterslip following the 1999 Chi-Chi, Taiwan earthquake. Geophys. Res. Lett. 29(16): doi: 10.1029/2002gl014967.Google Scholar
Hsu, Y. J., Segall, P., Yu, S. B., Kuo, L. C., and Williams, C. A. 2007. Temporal and spatial variations of post-seismic deformation following the 1999 Chi-Chi, Taiwan earthquake. Geophys. J. Int., 169(2): 367379, doi: 10.1111/j.1365-246X.2006.03310.x.Google Scholar
Hsu, Y. J., Simons, M., Yu, S. B., Kuo, L. C., and Chen, H. Y. 2003. A two-dimensional dislocation model for interseismic deformation of the Taiwan mountain belt. Earth Planet. Sci. Lett. 211(3–4): 287294, doi: 10.1016/s0012-821x(03)00203–6.Google Scholar
Hu, M. S., and Evans, A. G. 1989. The cracking and decohesion of thin-films on ductile substrates. Acta Metall. 37: 917925.Google Scholar
Huang, J., and Turcotte, D. L. 1990. Evidence for chaotic fault interactions in the seismicity of the San-Andreas fault and Nankai trough. Nature 348: 234236.Google Scholar
Hubbard, J., Barbot, S., Hill, E. M., and Tapponnier, P. 2015. Coseismic slip on shallow decollement megathrusts: Implications for seismic and tsunami hazard. Earth-Science Reviews 141: 4555, doi: 10.1016/j.earscirev.2014.11.003.Google Scholar
Hubbert, M. K., and Rubey, W. W. 1959. Role of fluid pressure in the mechanics of overthrust faulting. Bull. Geol. Soc. Am. 70: 115166.Google Scholar
Hudnut, K., Seeber, L., and Pacheco, J. F. 1989. Cross-fault triggering in the November 1987 Superstition Hills earthquake sequence, southern California. J. Geophys. Res. Lett. 16: 199202.Google Scholar
Hull, J. 1988. Thickness-displacement relationships for deformation zones. J. Struct. Geol. 10: 431435.Google Scholar
Hundley-Goff, E., and Moody, J. 1980. Microscopic characteristics of orthoquartzite from sliding friction experiments, I. Sliding surfaces. Tectonophysics 62: 279299.Google Scholar
Hurwitz, S., Sohn, R. A., Luttrell, K., and Manga, M. 2014. Triggering and modulation of geyser eruptions in Yellowstone National Park by earthquakes, earth tides, and weather. J. Geophys. Res.-Solid Earth 119(3): 17181737, doi: 10.1002/2013jb010803.Google Scholar
Husseini, M. 1977. Energy balance for motion along a fault. Geophys. J. R.A.S. 49: 699714.Google Scholar
Hyndman, R. D. 2013. Downdip landward limit of Cascadia great earthquake rupture. J. Geophys. Res.-Solid Earth 118(10): 55305549, doi: 10.1002/jgrb.50390.Google Scholar
Hyndman, R. D., and Wang, K. 1993. Thermal constraints on the zone of major thrust earthquake failure: The Cascadia subduction zone. J. Geophys. Res. 98: 20392060.Google Scholar
Hyndman, R. D., McCrory, P. A., Wech, A., Kao, H., and Ague, J. 2015. Cascadia subducting plate fluids channelled to fore-arc mantle corner: ETS and silica deposition. J. Geophys. Res.-Solid Earth 120(6): 43444358, doi: 10.1002/2015jb011920.Google Scholar
Ida, Y. 1972. Cohesive force across tip of a longitudinal shear crack and Griffith’s specific energy balance. J. Geophys. Res. 77: 37963805.Google Scholar
Ida, Y. 1973. Stress concentration and unsteady propagation of longitudinal shear cracks. J. Geophys. Res. 78: 34183429.Google Scholar
Ide, S. 2010. Striations, duration, migration and tidal response in deep tremor. Nature 466(7304): 356-359.Google Scholar
Ide, S., and Aochi, H. 2005. Earthquakes as multiscale dynamic ruptures with heterogeneous fracture surface energy. J. Geophys. Res. 110, doi: 10.1029/2004JB003591.Google Scholar
Ide, S., and Beroza, G. C. 2001. Does apparent stress vary with earthquake size? Geophys. Res. Lett. 28(17): 33493352, doi: 10.1029/2001gl013106.Google Scholar
Ide, S., Baltay, A., and Beroza, G. C. 2011. Shallow dynamic overshoot and energetic deep rupture in the 2011 M-w 9.0 Tohoku-Oki earthquake. Science 332(6036): 14261429, doi: 10.1126/science.1207020.Google Scholar
Ide, S., Beroza, G. C., Shelly, D. R., and Uchide, T. 2007. A scaling law for slow earthquakes, Nature, 447(7140): 7679.Google Scholar
Ide, S., Imamura, F., Yoshida, Y., and Abe, K. 1993. Source characteristics of the Nicaraguan tsunami earthquake of September 2, 1992. Geophys. Res. Lett. 20(9): 863866.Google Scholar
Ide, S., Yabe, S., and Tanaka, Y. 2016. Earthquake potential revealed by tidal influence on earthquake size–frequency statistics, Nature Geoscience, 9: 834838.Google Scholar
Igarashi, T., Matsuzawa, T., and Hasegawa, A. 2003. Repeating earthquakes and interplate aseismic slip in the northeastern Japan subduction zone. J. Geophys. Res.-Solid Earth 108(B5): 2249, doi: 10.1029/2002jb001920.Google Scholar
Ihmle, P. F., and Jordan, T. H. 1994. Teleseismic search for slow precursors to large earthquakes. Science 266: 15471551.Google Scholar
Ihmle, P. F., Harabaglia, P., and Jordan, T. H. 1993. Teleseismic detection of a slow precursor to the great 1989 Macquarie Ridge earthquake. Science 261(5118): 177183, doi: 10.1126/science.261.5118.177.Google Scholar
Iio, Y. 1995.Observations of a slow emergent phase generated by microearthquakes – Implications for earthquake nucleation and propagation. J. Geophys. Res.-Solid Earth 100(B8): 1533315349, Doi: 10.1029/95jb01150.Google Scholar
Ikari, M. J., Ito, Y., Ujiie, K., and Kopf, A. J. 2015. Spectrum of slip behaviour in Tohoku fault zone samples at plate tectonic slip rates. Nature Geoscience 8(11): 870+, doi: 10.1038/ngeo2547.Google Scholar
Ikari, M. J., Kameda, J., Saffer, D. M., and Kopf, A. J. 2015. Strength characteristics of Japan Trench borehole samples in the high-slip region of the 2011 Tohoku-Oki earthquake. Earth Planet. Sci. Lett. 412: 3541Google Scholar
Ikari, M. J., Marone, C., and Saffer, D. M. 2011.On the relation between fault strength and frictional stability. Geology 39(1): 8386, doi: 10.1130/g31416.1.Google Scholar
Ikari, M. J., Marone, C., Saffer, D. M., and Kopf, A. J. 2013. Slip weakening as a mechanism for slow earthquakes. Nat. Geosci. 6(6): 468472, doi: 10.1038/ngeo1818.Google Scholar
Imamura, A. 1937. Theoretical and Applied Seismology. Tokyo: Maruzen.Google Scholar
Imber, J., Holdsworth, R. E., Smith, S. A. F., Jefferies, S. P., and Collettini, C. 2008. Frictional-visous flow, seismicity and the geology of weak faults: A review and future directions. In The Internal Stucture 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., The Geological Society of London, doi: 10.1144/SP299.10.Google Scholar
Incel, S., Hilairet, N., Labrousse, L., John, T., Deldicque, D., Ferrand, T., Wang, Y. B., Renner, J., Morales, L., and Schubnel, A. 2017. Laboratory earthquakes triggered during eclogitization of lawsonite-bearing blueschist. Earth Planet. Sci. Lett. 459: 320331, doi: 10.1016/j.epsl.2016.11.047.Google Scholar
Irwin, G. R. 1958. Fracture. In Handbuch der Physik, ed. Flugge, S.. Berlin: Springer-Verlag, pp. 551590.Google Scholar
Irwin, W. P., and Barnes, I. 1975. Effects of geological structure and metamorphic fluids on seismic behavior of the San Andreas fault system in central and northern California. Geology 3: 713716.Google Scholar
Isacks, B. L., and Barazangi, M. 1977. Geometry of Benioff zones. Island arcs, deep sea Trenches and Back-arc Basins, 99114: doi: 10.1029/ME001.Google Scholar
Isacks, B. L., Oliver, J., and Sykes, L. R. 1968. Seismology and the new global tectonics. J. Geophys. Res. 73: 58555899.Google Scholar
Isacks, B., and Molnar, P. 1971. Distribution of stresses in the descending lithosphere from a global survey of focal‐mechanism solutions of mantle earthquakes. Rev. Geophys., 9(1): 103174.Google Scholar
Ishibashi, K. 1981. Specification of soon-to-occur seismic faulting in the Tokai District, central Japan, based on seismotectonics. In Earthquake Prediction: An International Review, eds. Simpson, D. W. and Richards, P. G.. Washington, DC: Amer. Geophys. Union, pp. 297332.Google Scholar
Ishibashi, K. 1988. Two categories of earthquake precursors, physical and tectonic, and their role in intermediate-term earthquake prediction. Pageoph 126: 687700.Google Scholar
Ishibashi, K. 2004. Status of historical seismology in Japan. Annals of Geophysics 47: 339368.Google Scholar
Ishlinski, A. Y., and Kraghelsky, I. V. 1944. On stick-slip in friction. ZhurTekhn. Fiz. 14: 276282.Google Scholar
Israelachvili, J. N., McGuiggan, P. M., and Homola, A. M. 1988. Dynamic properties of molecularly thin liquid films. Science 240(4849): 189191, doi: 10.1126/science.240.4849.189.Google Scholar
Ito, A., Fujie, G., Miura, S., Kodaira, S., Kaneda, Y., and Hino, R. 2005. Bending of the subducting oceanic plate and its implication for rupture propagation of large interplate earthquakes off Miyagi, Japan, in the Japan Trench subduction zone. Geophys. Res. Lett. 32(5): doi: 10.1029/2004gl022307.Google Scholar
Ito, T., and Zoback, M. D. 2000. Fracture permeability and in situ stress to 7 km depth in the KTB Scientific Drillhole. Geophys. Res. Lett. 27: 10451048.Google Scholar
Ito, Y., and Obara, K. 2006. Very low frequency earthquakes within accretionary prisms are very low stress-drop earthquakes. Geophys. Res. Lett. 33(9): L09302.Google Scholar
Ito, Y., Hino, R., Suzuki, S., et al. 2013. Episodic slow slip events in the Japan subduction zone before the 2011 Tohoku-Oki earthquake. Tectonophysics 600: 1426, doi: 10.1016/j.tecto.2012.08.022.Google Scholar
Ito, Y., Obara, K., Shiomi, K., Sekine, S., and Hirose, H. 2007. Slow earthquakes coincident with episodic tremors and slow slip events Science 315: 503506.Google Scholar
Iyer, K., Rupke, L. H., Morgan, J. P., and Grevemeyer, I. 2012. Controls of faulting and reaction kinetics on serpentinization and double Benioff zones. Geochem, Geophys., Geosystems 13, doi: 10.1029/2012GC004304.Google Scholar
Jackson, D., and Kagan, Y. 2006. The 2004 Parkfield earthquake, the 1985 prediction, and characteristic earthquakes: Lessons for the future. Bull. Seismol. Soc. Am. 96(4B): S397S409.Google Scholar
Jackson, J. A. 1987. Active normal faulting and crustal extension. In Continental Extensional Tectonics, eds. Coward, M., Dewey, J., and Hancock, P.. London: Blackwell, pp. 318.Google Scholar
Jackson, J., and McKenzie, D. 1988. The relationship between plate motions and seismic moment tensors, and the rates of active deformation in the Mediterranean and Middle East. Geophys. J. R.A.S. 93: 4573.Google Scholar
Jackson, J., and McKenzie, D. 1999. A hectare of fresh striations on the Arkitsa Fault, central Greece. J. Struct. Geol. 21: 16.Google Scholar
Jackson, J., and Molnar, P. 1990. Active Faulting and Block Rotations in the Western Transverse Ranges, California. J. Geophys. Res.-Solid Earth and Planets 95(B13): 2207322087.Google Scholar
Jacob, K. H. 1984. Estimates of long-term probabilities for future great earthquakes in the Aleutians. Geophys. Res. Lett. 11: 295298.Google Scholar
Jacob, K. H., Armbruster, J., Seeber, L., Pennington, W., and Farhatulla, S. 1979. Tarbella reservoir, Pakistan: A region of compressive tectonics and reduced seismicity upon initial reservoir filling. Bull. Seismol. Soc. Am. 69: 11751192.Google Scholar
Jaeger, J. C., and Cook, N. G. W. 1976. Fundamentals of Rock Mechanics. London: Chapman and Hall.Google Scholar
Jaeger, J. C., Cook, N. G. W., Zimmerman, R. G. 2007. Fundamentals of Rock Mechanics, 4th edn. Malden; Blackwell.Google Scholar
Jankaew, K., Atwater, B. F., Sawai, Y., Choowong, M., Charoentitirat, T., Martin, M. E., and Prendergast, A. 2008. Medieval forewarning of the 2004 Indian Ocean tsunami in Thailand. Nature 455(7217): 12281231.Google Scholar
Jaoul, O., Tullis, J. A., and Kronenberg, A. K. 1984. The effect of varying water content on the creep behavior of Heavitree quartzite. J. Geophys. Res. 89: 42894312.Google Scholar
Jara, J., Socquet, A., Marsan, D., and Bouchon, M. 2017. Long-term interactions between intermediate depth and shallow seismicity in North Chile subduction zone. Geophys. Res. Lett 44: 92839292, doi: 10.1002/2017GL075029.Google Scholar
Jarrard, R. D. 1986a. Relations among subduction parameters. Rev. Geophys. 24: 217284.Google Scholar
Jarrard, R. D. 1986b. Causes of compression and extension behind trenches. Tectonophysics 132: 89102.Google Scholar
Jaumé, S. C., and Lillie, R. J. 1988. Mechanics of the Salt Range–Potwar Plateau, Pakistan: A fold and thrust belt underlain by evaporites. Tectonics 7: 5771.Google Scholar
Jaumé, S. C., and Sykes, L. R. 1992. Changes in state of stress on the southern San Andreas fault resulting from the California earthquake sequence of April to June 1992. Science 258: 13251328.Google Scholar
Jaumé, S. C., and Sykes, L. R. 1996. Evolution of moderate seismicity in the San Francisco Bay region, 1850 to 1993: Seismicity changes related to the occurrence of large and great earthquakes. J. Geophys. Res.-Solid Earth 101: 765789.Google Scholar
Jaumé, S. C., and Sykes, L. R. 1999. Evolving towards a critical point: A review of accelerating seismic moment/energy release prior to large and great earthquakes. Pure Appl. Geophys. 155: 279305.Google Scholar
Ji, C., Helmberger, D. V., Wald, D. J., and Ma, K. F. 2003. Slip history and dynamic implications of the 1999 Chi-Chi, Taiwan, earthquake. J. Geophys. Res.-Solid Earth, 108(B9): doi: 10.1029/2002jb001764.Google Scholar
Jiang, J., and Lapusta, N. 2017. Deeper penetration of large earthquakes on seismically quiescent faults. Science 352: 12931297, doi: 10.1126/science.aaf1496.Google Scholar
Jiao, W. J., Silver, P. G., Fei, Y. W., and Prewitt, C. T. 2000. Do intermediate- and deep-focus earthquakes occur on preexisting weak zones? An examination of the Tonga subduction zone. J. Geophys. Res.-Solid Earth 105(B12): 2812528138, doi: 10.1029/2000jb900314.Google Scholar
Johnson, A. M., Fleming, R. W., and Cruikshank, K. M. 1994. Shear zones formed along long, straight traces of fault zones during the 28 June Landers, California, earthquake. Bull. Seismol. Soc. Amer. 84: 499510.Google Scholar
Johnson, J. M., and Satake, K. 1997. Estimation of seismic moment and slip distribution of the April 1, 1946, Aleutian tsunami earthquake. J. Geophys. Res.-Solid Earth 102: 1176511774.Google Scholar
Johnson, K. M., Hsu, Y. J., Segall, P., and Yu, S. B. 2001. Fault geometry and slip distribution of the 1999 Chi-Chi, Taiwan earthquake imaged from inversion of GPS data. Geophys. Res. Lett. 28(11): 22852288, doi: 10.1029/2000gl012761.Google Scholar
Johnson, T. L. 1981. Time dependent friction of granite: Implications for precursory slip on faults. J. Geophys. Res. 86: 60176028.Google Scholar
Johnson, T. L., and Scholz, C. H. 1976. Dynamic properties of stick-slip friction in rock. J. Geophys. Res. 81: 881888.Google Scholar
Johnson, T. L., Wu, F. T., and Scholz, C. H. 1973. Source parameters for stick-slip and for earthquakes. Science 179: 278280.Google Scholar
Johnston, A. C. 1996a. Seismic moment assessment of earthquakes in stable continental regions 1. Instrumental seismicity. Geophys. J. Int. 124: 381414.Google Scholar
Johnston, A. C. 1996b. Seismic moment assessment of earthquakes in stable continental regions 2. Historical seismicity. Geophys. J. Int. 125: 639678.Google Scholar
Jolivet, R., Lasserre, C., Doin, M. P., Peltzer, G., Avouac, J. P., Sun, J., and Dailu, R. 2013. Spatio-temporal evolution of aseismic slip along the Haiyuan fault, China: Implications for fault frictional properties. Earth Planet. Sci. Lett. 377: 2333, doi: 10.1016/j.eps1.2013.07.020.Google Scholar
Jones, L. M. 1988. Focal mechanisms and the state of stress on the San Andreas fault in Southern California. J. Geophys. Res. 93: 88698892.Google Scholar
Jones, L. M., and Molnar, P. 1979. Some characteristics of foreshocks and their possible relationship to earthquake prediction and premonitory slip on faults. J. Geophys. Res. 84: 35963608.Google Scholar
Jonsson, S., Segall, P., Pedersen, R., and Bjornsson, G. 2003. Post-earthquake ground movements correlated to pore-pressure transients. Nature 424(6945): 179183, doi: 10.1038/nature01776.Google Scholar
Jordan, P. 1988. The rheology of polymineralic rocks – An approach. Geol. Runds. 77: 285294.Google Scholar
Jordan, T., Chen, Y.-T., Gasparini, P., et al. 2011. Operational Earthquake Forecasting: State of Knowledge and Guidelines for Implementation. Ann. Geophys. 54(4): 315391.Google Scholar
Jordan, T., Marzocchi, W., Michael, A., and Gerstenberger, M. 2014. Operational earthquake forecasting can enhance earthquake preparedness, Seism. Res. Lett. 85(5): 955959.Google Scholar
Jordan, T. H. 2013. Lessons of L’Aquila for operational earthquake forecasting, Seism. Res. Lett. 84(1): 47.Google Scholar
Jordan, T. H., and Jones, L. M. 2010. Operational earthquake forecasting: Some thoughts on why and how. Seismol. Res. Lett. 81(4): 571574.Google Scholar
Jung, H., Green, H. W., and Dobrzhinetskaya, L. F. 2004. Intermediate-depth earthquake faulting by dehydration embrittlement with negative volume change. Nature 428(6982): 545549, doi: 10.1038/nature02412.Google Scholar
Kaduri, M., Gratier, J. P., Renard, F., and Lasserre, C. 2017. The implications of fault zone transformation on aseismic creep: Example of the North Anatolian Fault, Turkey: Rock transformation and aseismic creep. J. Geophys. Res. doi: 10.1002/2016JB013803.Google Scholar
Kagan, Y. Y. 1991. Likelihood analysis of earthquake catalogues. Geophys. J. Int., 106(1): 135148, doi: 10.1111/j.1365-246X.1991.tb04607.x.Google Scholar
Kagan, Y. Y., and Jackson, D. D. 1991. Seismic gap hypothesis – 10 years after. J. Geophys. Res.-Solid Earth 96: 2141921431.Google Scholar
Kagan, Y. Y., and Jackson, D. D. 2000. Probabilistic forecasting of earthquakes. Geophys. J. Int. 143(2): 438453.Google Scholar
Kagan, Y., and Knopoff, L. 1978. Statistical study of the occurrence of shallow earthquakes. Geophys. J. R.A.S. 55: 6786.Google Scholar
Kaiser, A., Balfour, N., Fry, B., et al., 2017. The 2016 Kaikōura, New Zealand, Earthquake: Preliminary Seismological Report. Seismol. Res. Lett. 88(3): 727739.Google Scholar
Kame, N., and Yamashita, T. 2003. Dynamic branching, arresting of rupture and the seismic wave radiation in self-chosen crack path modelling. Geophys. J. Int. 155(3): 10421050.Google Scholar
Kame, N., Fujita, S., Nakatani, M., and Kusakabe, T. 2015 Earthquake Nucleation on Faults with a Revised Rate- and State-Dependent Friction Law. Pure Appl. Geophys 172(8): 22372246, doi: 10.1007/s00024-013–0744-0.Google Scholar
Kanamori, H. 1971. Great earthquakes at island arcs and the lithosphere. Tectonophysics 12: 187198.Google Scholar
Kanamori, H. 1972. Mechanism of tsunami earthquakes. Phys. Earth Planet. Int. 6: 346359.Google Scholar
Kanamori, H. 1973. Mode of strain release associated with major earthquakes in Japan. Ann. Rev. Earth Planet. Sci. 5: 129139.Google Scholar
Kanamori, H. 1977. The energy release in great earthquakes. J. Geophys. Res. 82: 29812987.Google Scholar
Kanamori, H. 1981. The nature of seismicity patterns before large earthquakes. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 119.Google Scholar
Kanamori, H. 1986. Rupture process of subduction zone earthquakes. Ann. Rev. Earth Planet. Sci. 14: 293322.Google Scholar
Kanamori, H. 2005. Real-time seismology and earthquake damage mitigation. Ann. Rev. Earth Planet. Sci. 33: 195214, doi: 10.1146/annurev.earth.33.092203.122626.Google Scholar
Kanamori, H., and Allen, C. 1986. Earthquake repeat time and average stress drop. In Earthquake Source Mechanics. Geophys, AGU. Mono. 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 227236.Google Scholar
Kanamori, H., and Anderson, D. 1975. Theoretical basis of some empirical relations in seismology. Bull. Seismol. Soc. Am. 65: 10731095.Google Scholar
Kanamori, H., and Brodsky, E. E. 2004. The physics of earthquakes. Rep. Prog. Phys. 67(8): 14291496, doi: 10.1088/0034–4885/67/8/r03.Google Scholar
Kanamori, H., and Cipar, J. 1974. Focal process of the great Chilean earthquake May 22,1960. Phys. Earth Planet. Int. 9: 128136.Google Scholar
Kanamori, H., and Fuis, G. 1976. Variations of P-wave velocity before and after the Galway Lake earthquake (ML = 5.2) and the Goat Mountain earthquakes (ML = 4.7), 1975, in the Mojave Desert, California. Bull. Seismol. Soc. Am. 66: 20172038.Google Scholar
Kanamori, H., and Heaton, T. H. 2000. Microscopic and macroscopic physics of earthquakes. In GeoComplexity and the Physics of Earthquakes, Washington, DC: Amer. Geophys. Un., pp. 147163.Google Scholar
Kanamori, H., and Kikuchi, M. 1993. The 1992 Nicaragua earthquake: A slow tsunami earthquake associated with subducted sediments. Nature 361: 714716.Google Scholar
Kanamori, H., and McNally, K. C. 1982. Variable rupture mode of the subduction zone along the Equador-Colombia coast. Bull. Seismol. Soc. Am. 72: 12411253.Google Scholar
Kanamori, H., and Rivera, L. 2006. Energy partitioning during an earthquake. In Earthquakes: Radiated Energy and the Physics of Faulting, eds. Abercrombie, R. et al., Amer. Geophys. Union. Monograph 170, Washington D.C., doi: 10.1029/170GM03.Google Scholar
Kanamori, H., and Stewart, G. S. 1976. Mode of strain release along the Gibbs fracture zone, Mid-Atlantic ridge. Phys. Earth Planet. Int. 11: 312332.Google Scholar
Kanamori, H., Anderson, D. L., and Heaton, T. H. 1998. Frictional melting during the rupture of the 1994 Bolivian earthquake. Science 279(5352): 839842, doi: 10.1126/science.279.5352.839.Google Scholar
Kanamori, H., Hauksson, E., and Heaton, T. 1997. Real-time seismology and earthquake hazard mitigation. Nature 390: 461464.Google Scholar
Kanaori, Y. 1991. Grain boundary microcracking of granitic rocks from the northeastern region of the Atotsugawa fault, central Japan: SEM backscattered electron images. Engin. Geol. 30: 221235.Google Scholar
Kaneko, Y., Avouac, J. P., and Lapusta, N. 2010. Towards inferring earthquake patterns from geodetic observations of interseismic coupling, Nature Geoscience, 3(5): 363–324, doi: 10.1038/ngeo843.Google Scholar
Kanninen, M. F., and Popelar, C. H. 1985. Advanced Fracture Mechanics. Oxford: Oxford University Press.Google Scholar
Kao, H., and Chen, W. P. 2000. The Chi-Chi earthquake sequence: Active, out-of-sequence thrust faulting in Taiwan. Science 288(5475): 23462349, doi: 10.1126/science.288.5475.2346.Google Scholar
Kao, H., Shan, S. J., Dragert, H., and Rogers, G. 2009. Northern Cascadia episodic tremor and slip: A decade of tremor observations from 1997 to 2007. J. Geophys. Res.-Solid Earth 114.Google Scholar
Karato, S.-i., Riedel, M. R., and Yuen, D. A. 2001. Rheological structure and deformation of subducted slabs in the mantle transition zone: Implications for mantle circulation and deep earthquakes. Phys. Earth Planet. Int. 127(1): 83108.Google Scholar
Karig, D. E. 1970. Kermadec arc-New Zealand tectonic confluence. N.Z. J. Geol. Geophys. 13: 2129.Google Scholar
Karig, D. E., and Hou, G. 1992. High-stress consolidation experiments and their geological implications. J. Geophys. Res. 97: 289300.Google Scholar
Kasahara, K. 1981. Earthquake Mechanics. Cambridge: Cambridge University Press.Google Scholar
Kato, A., and Igarashi, T. 2012. Regional extent of the large coseismic slip zone of the 2011 Mw 9.0 Tohoku-Oki earthquake delineated by on-fault aftershocks. Geophys. Res. Lett. 39, doi: 10.1029/2012gl052220.Google Scholar
Kato, A., Fukuda, J. I., Kumazawa, T., and Nakagawa, S. 2016. Accelerated nucleation of the 2014 Iquique, Chile Mw 8.2 earthquake. Scientific Reports 6: 24792, doi: 10.1038/srep24792.Google Scholar
Kato, A., Obara, K., Igarashi, T., Tsuruoka, H., Nakagawa, S., and Hirata, N. 2012. Propagation of slow slip leading up to the 2011 m-w 9.0 Tohoku-Oki earthquake. Science 335(6069): 705708, doi: 10.1126/science.1215141.Google Scholar
Kato, N. 2007. Expansion of aftershock areas caused by propagating post-seismic sliding. Geophys. J. Int. 168(2): 797808, doi: 10.1111/j.1365-246X.2006.03255.x.Google Scholar
Kawamoto, E., and Shimamoto, T. 1998. The strength profile for bimineralic shear zones: An insight from high-temperature shearing experiments on calcite-halite mixtures. Tectonophysics 295: 114.Google Scholar
Kawamura, H., Hatano, T., Kato, N., Biswas, S., and Chakrabarti, B. K. 2012. Statistical physics of fracture, friction, and earthquakes. Rev. Modern Phys. 84(2): 839884, doi: 10.1103/RevModPhys.84.839.Google Scholar
Kawasaki, I., Kawahara, Y., Takata, I., and Kosugi, N. 1985. Mode of seismic moment release at transform faults. Tectonophysics 118: 313327.Google Scholar
Keilis-Borok, V. I., and Kossobokov, V. G. 1990. Premonitory observations of earthquake flow: Algorithm M8. Phys. Earth Planet. Int. 61: 7383.Google Scholar
Kelemen, P. B., and Hirth, G. 2007. A periodic shear-heating mechanism for intermediate-depth earthquakes in the mantle. Nature 446(7137): 787.Google Scholar
Kelleher, J., and McCann, W. 1976. Bouyant zones, great earthquakes, and unstable boundaries of subduction. J. Geophys. Res. 81: 48854896.Google Scholar
Kelleher, J., Savino, J., Rowlett, H., and McCann, W. 1974. Why and where great thrust earthquakes occur along island arcs. J. Geophys. Res. 79: 48894899.Google Scholar
Keller, E. A., Gurrola, L., and Tierney, T. E. 1999. Geomorphic criteria to determine direction of lateral propagation of reverse faulting and folding. Geology 27: 515518.Google Scholar
Kelsey, H. M., Nelson, A. R., Hemphill-Haley, E., and Witter, R. C. 2005. Tsunami history of an Oregon coastal lake reveals a 4600 yr record of great earthquakes on the Cascadia subduction zone. Geol. Soc. Amer. Bull. 117: 10091032.Google Scholar
Kemeny, J. M. 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
Kenner, S. J., and Segall, P. 2000. Postseismic deformation following the 1906 San Francisco earthquake. J. Geophys. Res. 105: 13,19513,209.Google Scholar
Kenner, S. J., and Segall, P. 2003. Lower crustal structure in northern California: Implications from strain rate variations following the 1906 San Francisco earthquake. J. Geophys. Res.-Solid Earth 108(B1): ETG 5-1–ETG 5-17.Google Scholar
Kenner, S. J., and Simons, M. 2005. Temporal clustering of major earthquakes along individual faults due to post-seismic reloading. Geophys. J. Int. 160: 179194, doi: 10.1111/j.1365-246X.2005.02460.xGoogle Scholar
Keranen, K. M., and Weingarten, M. 2017. Induced seismicity, Ann. Rev. Earth Planet. Sci. 46:149174.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(6): 699702.Google Scholar
Keranen, K. M., Weingarten, M., Abers, G. A., Bekins, B. A., and Ge, S. 2014. Sharp increase in central Oklahoma seismicity since 2008 induced by massive wastewater injection. Science 345(6195): 448451, doi: 10.1126/science.1255802.Google Scholar
Kerrich, R. 1986. Fluid infiltration into fault zones: Chemical, isotopic, and mechanical effects. Pageoph 124: 225268.Google Scholar
Kerrich, R., Beckinsdale, R. D., and Durham, J. J. 1977. The transition between deformation regimes dominated by intercrystalline diffusion and intracrystalline creep evaluated by oxygen isotope thermometry. Tectonophysics 38: 241257.Google Scholar
Kerrick, D., and Connolly, J. 2001. Metamorphic devolatilization of subducted oceanic metabasalts: implications for seismicity, arc magmatism and volatile recycling. Earth Planet. Sci. Lett. 189(1): 1929.Google Scholar
Kessler, D. W. 1933. Wear resistance of natural stone flooring. US Bureau of Standards, Res. Pap. RP612 635648.Google Scholar
Kikuchi, M., and Fukao, Y. 1987. Inversion of long period P waves from great earthquakes along subduction zones. Tectonophysics 144: 231248.Google Scholar
Kikuchi, M., and Kanamori, H. 1995. Source characteristics of the 1992 Nicaragua tsunami earthquake inferred from teleseismic body waves. Pure Appl. Geophys 144(3): 441453.Google Scholar
Kilb, D., Gomberg, J., and Bodin, P. 2000. Triggering of earthquake aftershocks by dynamic stresses. Nature 408: 570574.Google Scholar
Kilgore, B. D., Blanpied, M. L. and Dieterich, J. H. 1993. Velocity dependent friction of granite over a wide range of conditions. Geophys. Res. Lett. 20 (10): 903906.Google Scholar
Kim, W. Y. 2013. Induced seismicity associated with fluid injection into a deep well in Youngstown, Ohio. J. Geophys. Res.-Solid Earth 118(7): 35063518, doi: 10.1002/jgrb.50247.Google Scholar
Kim, Y. S., and Sanderson, D. J. 2006. Structural similarity and variety at the tips in a wide range of strike-slip faults: A review. Terra Nova 18(5): 330344.Google Scholar
King, D. S. H., and Marone, C. 2012. Frictional properties of olivine at high temperature with applications to the strength and dynamics of the oceanic lithosphere. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2012jb009511.Google Scholar
King, G. C. P. 1986. Speculations on the geometry of the initiation and termination processes of earthquake rupture and its relation to morphology and geological structure. Pageoph 124: 567586.Google Scholar
King, G. C. P., and Cocco, M. 2000. Fault interaction by elastic stress changes: New clues from earthquake sequences. Advances in Geophys. 44: 138.Google Scholar
King, G. C. P., and Nabelek, J. 1985. Role of fault bends in the initiation and termination of earthquake rupture. Science 228: 984987.Google Scholar
King, G. C. P., Stein, R. S., and Lin, J. 1994. Static stress changes and the triggering of earthquakes. Bull. Seismol. Soc. Amer. 84: 935953.Google Scholar
King, N. E., and Savage, J. C. 1984. Regional deformation near Palmdale, California. J. Geophys. Res. 89: 24712477.Google Scholar
King, N. E., and Thatcher, W. 1998. The coseismic slip distributions of the 1940 and 1979 Imperial Valley, California, earthquakes and their implications. J. Geophys. Res.-Solid Earth 103: 1806918086.Google Scholar
Kirby, S. 1980. Tectonic stress in the lithosphere: Constraints provided by the experimental deformation of rock. J. Geophys. Res. 85: 63536363.Google Scholar
Kirby, S. 1983. Rheology of the lithosphere. Rev. Geophys. Space Phys. 21: 14581487.Google Scholar
Kirby, S. 1987. Localized polymorphic phase transformations in high pressure faults and applications to the physical mechanism of deep earthquakes. J. Geophys. Res. 92: 1378913800.Google Scholar
Kirby, S. H., Stein, S., Okal, E. A., and Rubie, D. C. 1996. Metastable mantle phase transformations and deep earthquakes in subduction oceanic lithosphere. Rev. Geophys., 34: 261306.Google Scholar
Kirkpatrick, J. D., and Brodsky, E. E. 2014. Slickenline orientations as a record of fault rock rheology. Earth Planet. Sci. Lett. 408: 2434, doi: 10.1016/j.eps1.2014.09.040.Google Scholar
Kirkpatrick, J. D., Shipton, Z. K., and Persano, C. 2009. Pseudotachylytes: Rarely Generated, Rarely Preserved, or Rarely Reported? Bull. Seismol. Soc. Am. 99(1): 382388, doi: 10.1785/0120080114.Google Scholar
Kirschner, D. L., and Chester, F. M. 1999. Are fluids important in seismogenic faulting? Geochemical evidence for limited fluid-rock interaction in two strike-slip faults of the San Andreas system. Eos 80, suppl. F727.Google Scholar
Kisslinger, C. 1975. Processes during the Matsushiro swarm as revealed by leveling, gravity, and spring-flow observations. Geology 3: 5762.Google Scholar
Kita, S., Okada, T., Nakajima, J., Matsuzawa, T., and Hasegawa, A. 2006. Existence of a seismic belt in the upper plane of the double seismic zone extending in the along-arc direction at depths of 70–100 km beneath NE Japan. Geophys. Res. Lett. 33(24): doi: 10.1029/2006gl028239.Google Scholar
Klinger, Y. 2010. Relation between continental strike‐slip earthquake segmentation and thickness of the crust. J. Geophys. Res.-Solid Earth 115(B7).Google Scholar
Knipe, R. J., and White, S. H. 1979. Deformation in low grade shear zones in the Old Red Sandstone, SW Wales. J. Struct. Geol. 1: 5366.Google Scholar
Knopoff, L. 1958. Energy release in earthquakes. Geophys. J. R.A.S. 1: 4452.Google Scholar
Kodaira, S., Iidaka, T., Kato, A., Park, J. O., Iwasaki, T., and Kaneda, Y. 2004. High pore fluid pressure may cause silent slip in the Nankai Trough. Science 304(5675): 12951298.Google Scholar
Kodaira, S., Takahashi, N., Nakanishi, A., Miura, S., and Kaneda, Y. 2000. Subducted seamount imaged in the rupture zone of the 1946 Nankaido earthquake. Science 289: 104106.Google Scholar
Kogan, M. G., Vasilenko, N. F., Frolov, D. I., et al. 2013. Rapid postseismic relaxation after the great 2006–2007 Kuril earthquakes from GPS observations in 2007–2011 J. Geophys. Res.-Solid Earth 118(7): 36913706, doi: 10.1002/jgrb.50245.Google Scholar
Kohlstedt, D. L., and Weathers, M. S. 1980. Deformation-induced microstructures, paleopiezometers, and differential stresses in deeply eroded fault zones. J. Geophys. Res. 85: 62696285.Google Scholar
Koketsu, K., YokotaY., NishimuraN., et al. 2011. A unified source model for the 2011 Tohoku earthquake. Earth Planet. Sci. Lett. 310(3–4): 480487, doi: 10.1016/j.epsl.2011.09.009.Google Scholar
Komori, J., Shishikura, M., Ando, R., Yokoyama, Y., and Miyairi, Y. 2017. History of the great Kanto earthquakes inferred from the ages of Holocene marine terraces revealed by a comprehensive drilling survey. Earth Planet. Sci. Lett. 471: 7484.Google Scholar
Konca, A. O., Hjorleifsdottir, V., Song, T.-R. A., et al. 2007. Rupture kinematics of the 2005 Mw 8.6 Nias–Simeulue earthquake from the joint inversion of seismic and geodetic data. Bull. Seismol. Soc. Am. 97(1A): S307S322.Google Scholar
Konca, A. O., Kaneko, Y., Lapusta, N., and Avouac, J. P. 2013. Kinematic Inversion of Physically Plausible Earthquake Source Models Obtained from Dynamic Rupture Simulations. Bull. Seismol. Soc. Am. 103(5): 26212644, doi: 10.1785/0120120358.Google Scholar
Kondo, H., Awata, Y., Emre, O., et al. 2005. Slip distribution, fault geometry, and fault segmentation of the 1944 Bolu-Gerede earthquake rupture, North Anatolian fault, Turkey. Bull. Seismol. Soc. Am. 95(4): 12341249, doi: 10.1785/0120040194.Google Scholar
Kondo, H., Ozaksoy, V., and Yildirim, C. 2010. Slip history of the 1944 Bolu-Gerede earthquake rupture along the North Anatolian fault system: Implications for recurrence behavior of multisegment earthquakes. J. Geophys. Res.-Solid Earth 115: B04316, doi: 10.1029/2009jb006413.Google Scholar
Koper, K. D., Hutko, A. R., Lay, T., Ammon, C. J., and Kanamori, H. 2011. Frequency-dependent rupture process of the 2011 M-w 9.0 Tohoku Earthquake: Comparison of short-period P wave backprojection images and broadband seismic rupture models. Earth Planets and Space 63(7): 599602, doi: 10.5047/eps.2011.05.026.Google Scholar
Kostoglodov, V., Singh, S. K., Santiago, J. A., Larsen, K. M., and Bilham, R. 2003. A large silent earthquake in the Guerrero seismic gap, Mexico. Geophys. Res. Lett. 30: 1807, doi: 10.1029/2003GL017219.Google Scholar
Kostrov, B. 1964. Self-similar problems of propagation of shear cracks. J. Appl. Math. Mech. 28: 10771087.Google Scholar
Kostrov, B. 1966. Unsteady propagation of longitudinal shear cracks. J. Appl. Math. Mech. 30: 12411248.Google Scholar
Kostrov, B. 1974. Seismic moment and energy of earthquakes, and seismic flow of rock. Izv. Acad. Sci. USSR Phys. Solid Earth 1: 2340.Google Scholar
Kostrov, B., and Das, S. 1988. Principles of Earthquake Source Mechanics. Cambridge: Cambridge University Press.Google Scholar
Koto, B. 1893. On the cause of the Great Earthquake in central Japan. J. Coll. Sci. Imp. Univ. 5. pt 4): 294353.Google Scholar
Koulali, A., McClusky, S., Susilo, S., et al. 2017. The kinematics of crustal deformation in Java from GPS observations: Implications for fault slip partitioning. Earth Planet. Sci. Lett. 458: 6979, doi: 10.1016/j.epsl.2016.10.039.Google Scholar
Kozaci, O., Dolan, J. F., Yonlu, O., and Hartleb, R. D. 2011. Paleoseismologic evidence for the relatively regular recurrence of infrequent, large-magnitude earthquakes on the eastern North Anatolian fault at Yaylabeli, Turkey. Lithosphere 3(1): 3754, doi: 10.1130/l118.1.Google Scholar
Kozaci, O., Dolan, J., Finkel, R., and Hartleb, R. 2007. Late Holocene slip rate for the North Anatolian fault, Turkey, from cosmogenic Cl-36 geochronology: Implications for the constancy of fault loading and strain release rates. Geology 35(10): 867870, doi: 10.1130/g23187a.1.Google Scholar
Kozdon, J. E., and Dunham, E. M. 2013. Rupture to the trench: Dynamic rupture simulatons of the 11 March 2011 Tohoku earthquake. Bull. Seismol. Soc. Am. 103: 12751289.Google Scholar
Krantz, R. W. 1988. Multiple fault sers and 3-dimensional strain – Theory and application. J. Struct. Geol. 10 (3): 225.Google Scholar
Kranz, R. L. 1979. Crack growth and development during creep of Barre granite. Int. J. Rock Mech. Min. Sci. 16: 2335.Google Scholar
Kranz, R. L. 1980. The effects of confining pressure and stress difference on static fatigue of granite. J. Geophys. Res. 85: 18541866.Google Scholar
Kranz, R. L., and Scholz, C. H. 1977. Critical dilatant volume of rocks at the onset of tertiary creep. J. Geophys. Res. 82: 48934898.Google Scholar
Kranz, R. L., Frankel, A., Engelder, T., and Scholz, C. 1979. The permeability of whole and jointed Barre granite. Int. J. Rock Mech. Min. Soc. 16: 225234.Google Scholar
Kranz, R. L., Harris, W. J., and Carter, N. L. 1982. Static fatigue of granite at 200°C. Geophys. Res. Lett. 9(1): 14, doi: 10.1029/GL009i001p00001.Google Scholar
Kristy, M., Burdick, L., and Simpson, D. 1980. The focal mechanisms of the Gazli, USSR, earthquakes. Bull. Seismol. Soc. Am. 70: 17371750.Google Scholar
Kubo, T., and Katayama, I. 2015. Effect of temperature on the frictional behavior of smectite and illite. J.Min. Pet. Sci. 110(6): 293299, doi: 10.2465/jmps.150421.Google Scholar
Kumar, S., and Singh, A. K. 2017. Ionospheric precursors observed in TEC due to earthquake of Tamenglong on 3 January 2016. Current Science 113(4): 795801, doi: 10.18520/cs/v113/i04/795–801.Google Scholar
Kwiatek, G., Plenkers, K., Nakatani, M., Yabe, Y., and Dresen, G. 2010. Frequency-magnitude characteristics down to magnitude-4.4 for induced seismicity recorded at Mponeng Gold Mine, South Africa. Bull. Seismol. Soc. Am. 100(3): 11651173.Google Scholar
Labuz, J. F., Shah, S. P., and Dowding, C. H. 1985. Experimental-analysis of crack-propagation in granite. Int. J. Rock Mech. Min. Sci. 22(2): 8598.Google Scholar
Labuz, J. F., Shah, S. P., and Dowding, C. H. 1987. The fracture process zone in granite – Evidence and effect. Int. J. Rock Mech. Min. Sci. Geomech. Abstr. 24(4): 235246.Google Scholar
Lachenbruch, A. 1980. Frictional heating, fluid pressure, and the resistance to fault motion. J. Geophys. Res. 85: 60976112.Google Scholar
Lachenbruch, A., and Sass, J. 1973. Thermo-mechanical aspects of the San Andreas. In Proc. Conf. on the Tectonic Problems of the San Andreas Fault System., eds. Kovach, R. and Nur, A.. Palo Alto, California: Stanford University Press, pp. 192205.Google Scholar
Lachenbruch, A. H., and McGarr, A. 1990. Stress and heat flow, in The San Andreas fault system, California. US Geol. Surv. Prof. Pop. 1515, pp. 261–77.Google Scholar
Lachenbruch, A. H., and Sass, J. 1980. Heat flow and energetics of the San Andreas fault zone. J. Geophys. Res. 85: 61856222.Google Scholar
Lachenbruch, A. H., and Sass, J. 1988. The stress heat-flow paradox and thermal results from Cajon Pass. Geophys. Res. Lett. 15: 981984.Google Scholar
Lachenbruch, A. H., and Sass, J. 1992. Heat flow from Cajon Pass, Fault strength, and tectonic implications. J. Geophys. Res. 97: 49955015.Google Scholar
Lachenbruch, A. H., and Sass, J. H. 1977. Heat flow in the United States and the thermal regime of the crust. In The Earth’s Crust, Washington, DC: American Geophysical Union, pp. 626675.Google Scholar
LaFemina, P., Dixon, T. H., Govers, R., et al. 2009. Fore‐arc motion and Cocos Ridge collision in Central America. Geochem., Geophys., Geosystems 10(5).Google Scholar
Lambert, A., Kao, H., Rogers, G., and Courtier, N. 2009. Correlation of tremor activity with tidal stress in the northern Cascadia subduction zone. J. Geophys. Res.-Solid Earth 114: B00A08, doi: 10.1029/2008jb006038.Google Scholar
Landis, C. A., and Coombs, D. S. 1967. Metamorphic belts and orogenesis in southern New Zealand. Tectonophysics 4: 501518.Google Scholar
Langseth, M., and Moore, J. C. 1990. Introduction to special section on the role of fluids in sediment accretion, deformation, diagenesis, and metamorphism in subduction zones. J. Geophys. Res. 95: 8737–41.Google Scholar
Lapworth, C. 1885. The highland controversy in British geology; its causes, course and consequence. Nature 32: 558559.Google Scholar
Larsen, P.-H. 1988. Relay structures in a Lower Permian basement-involved extensional system, East Greenland. J. Struct. Geol. 10: 38.Google Scholar
Lawn, B. R. 1967. Partial cone crack formation in a brittle material loaded with a sliding spherical indenter. Proc. Roy. Soc. London Ser. A 299: 307316.Google Scholar
Lawn, B. R. 2010. Fracture of Brittle Solids, 2nd Edn., Cambridge: Cambridge University Press.Google Scholar
Lay, T. 2015. The surge of great earthquakes from 2004 to 2014. Earth Planet. Sci. Lett. 409: 133146, doi: 10.1016/j.epsl.2014.10.047.Google Scholar
Lay, T., and Kanamori, H. 1980. Earthquake doublets in the Solomon Islands. Phys. Earth Planet. Int. 21: 283304.Google Scholar
Lay, T., and Kanamori, H. 1981. An asperity model of great earthquake sequences. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 579592.Google Scholar
Lay, T., and Wallace, T. C. 1995. Modern Global Seismology. San Diego, Calif.: Academic Press.Google Scholar
Lay, T., Ammon, C. J., Hutko, A. R., and Kanamori, H. 2010. Effects of kinematic constraints on teleseismic finite-source rupture inversions: Great Peruvian earthquakes of 23 June 2001 and 15 August 2007. Bull. Seismol. Soc. Am. 100(3): 969994, doi: 10.1785/0120090274.Google Scholar
Lay, T., Ammon, C. J., Kanamori, H., Rivera, L., Koper, K. D., and Hutko, A. R. 2010, The 2009 Samoa-Tonga great earthquake triggered doublet. Nature 466: 964968, doi: 10.1038/nature09214.Google Scholar
Lay, T., Astiz, L., Kanamori, H., and Christensen, D. H. 1989. Temporal variation of large intraplate earthquakes in coupled subduction zones. Phys. Earth Planet. Int. 54(3–4): 258312.Google Scholar
Lay, T., Duputel, Z., Ye, L. L., and Kanamori, H. 2013. The December 7, 2012 Japan Trench intraplate doublet (M-w 7.2, 7.1) and interactions between near-trench intraplate thrust and normal faulting. Phys. Earth Planet. Int. 220: 7378, doi: 10.1016/j.pepi.2013.04.009.Google Scholar
Lay, T., Kanamori, H., Ammon, C. J., et al. 2005. The great Sumatra-Andaman earthquake of 26 December 2004, Science, 308(5725): 11271133.Google Scholar
Lay, T., Kanamori, H., Ammon, C. J., et al. 2012. Depth-varying rupture properties of subduction zone megathrust faults. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb009133.Google Scholar
Lay, T., Kanamori, H., Ammon, C. J., Hutko, A. R., Furlong, K., and Rivera, L. 2009. The 2006–2007 Kuril Islands great earthquake sequence. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb006280.Google Scholar
Lay, T., Kanamori, H., and Ruff, L. 1982. The asperity model and the nature of large subduction zone earthquakes. Earthquake Pred. Res. 1: 171.Google Scholar
Lee, A. G. G., and Rutter, E. H. 2004. Experimental rock-on-rock frictional wear: Application to subglacial abrasion. J. Geophys. Res.-Solid Earth 109(B9): doi: 10.1029/2004jb003059.Google Scholar
Lee, J. C., Chu, H. T., Angelier, J., et al. 2002. Geometry and structure of northern surface ruptures of the 1999 Mw=7.6 Chi-Chi Taiwan earthquake: Influence from inherited fold belt structures. J. Struct. Geol. 24(1): 173192, doi: 10.1016/s0191-8141(01)00056–6.Google Scholar
Leeman, J. R., Saffer, D. M., Scuderi, M. M., and Marone, C. 2016. Laboratory observations of slow earthquakes and the spectrum of tectonic fault slip modes. Nat. Comm. 7: 11104, doi: 10.1038/ncomms11104.Google Scholar
Lei, X., Kusunose, K., Satoh, T., and Nishizawa, O. 2003. The hierarchical rupture process of a fault: An experimental study. Phys. Earth Planet. Int. 137: 213228.Google Scholar
Leloup, P. H., Ricard, Y., Battaglia, J., and Lacassin, R. 1999. Shear heating in continental strike-slip shear zones: Model and field examples. Geophys. J. Int. 136: 1940.Google Scholar
Lensen, G. J. 1975. Earth deformation studies in New Zealand. Tectonophysics 29: 541551.Google Scholar
Lensen, G. J. 1981. Tectonic strain and drift. Tectonophysics 71: 173188.Google Scholar
Leonard, M. 2010. Earthquake fault scaling: Self-consistent relating of rupture length, width, average displacement, and moment release. Bull. Seismol. Soc. Am. 100(5A): 19711988.Google Scholar
Levander, A., Henstock, T. J., Meltzer, A. S., Beaudoin, B. C., Trehu, A. M., and Klemperer, S. L. 1998. Fluids in the lower crust following Mendocino triple junction migration: Active basaltic intrusion? Geology 26: 171174.Google Scholar
Li, G. Y., Vidale, J. E., Aki, K., and Xu, F. 2000. Depth-dependent structure of the Landers fault zone from trapped waves generated by aftershocks. J. Geophys. Res. 105: 62376254.Google Scholar
Li, V. C. 1987. Mechanics of shear rupture applied to earthquake zones. In Fracture Mechanics of Rock, ed. Atkinson, B.. London: Academic Press, pp. 351428.Google Scholar
Li, V. C., and Rice, J. R. 1983a. Preseismic rupture progression and great earthquake instabilities at plate boundaries. J. Geophys. Res. 88: 42314246.Google Scholar
Li, V. C., and Rice, J. R. 1983b. Precursory surface deformation in great plate boundary earthquake sequences. Bull. Seismol. Soc. Amer. 73: 14151434.Google Scholar
Li, V. C., and Rice, J. R. 1987. Crustal deformation in great California earthquake cycles. J. Geophys. Res. 92: 1153311551.Google Scholar
Li, Y. G., Chen, P., Cochran, E. S., Vidale, J. E., and Burdette, T. 2006. Seismic evidence for rock damage and healing on the San Andreas fault associated with the 2004 M 6.0 Parkfield earthquake. Bull. Seismol. Soc. Am. 96(4): S349S363, doi: 10.1785/0120050803.Google Scholar
Li, Y. G., De Pascale, G. P., Quigley, M. C., and Gravley, D. M. 2014. Fault damage zones of the M7.1 Darfield and M6.3 Christchurch earthquakes characterized by fault-zone trapped waves. Tectonophysics 618: 79101, doi: 10.1016/j.tecto.2014.01.029.Google Scholar
Li, Y. G., Vidale, J. E., Aki, K., and Xu, F. 2000. Depth-dependent structure of the Landers fault zone from trapped waves generated by aftershocks. J. Geophys. Res.-Solid Earth 105(B3): 62376254, doi: 10.1029/1999jb900449.Google Scholar
Li, Y. G., Vidale, J. E., Aki, K., Xu, F., and Burdette, T. 1998. Evidence for shallow fault zone strengthening after the 1992 M 7.5 Landers earthquake. Science 279: 217219.Google Scholar
Li, Y.-G., Vidale, J. E., Day, S. M., Oglesby, D. D., and Cochran, E. 2003. Postseismic fault healing on the rupture zone of the 1999 M 7.1 Hector Mine, California, earthquake. Bull. Seismol. Soc. Am. 93(2): 854869.Google Scholar
Lienkaemper, J. J., Galehouse, J. S., and Simpson, R. W. 1997. Creep response of the Hayward fault to stress changes caused by the Loma Prieta earthquake. Science 276: 20142016.Google Scholar
Lienkaemper, J. J., McFarland, F. S., Simpson, R. W., and Caskey, S. J. 2014. Using surface creep rate to infer fraction locked for sections of the san andreas fault system in Northern California from alignment array and GPS data. Bull. Seismol. Soc. Amer. 104, doi: 10.1785/0120140117.Google Scholar
Lin, A. 2008. Fossil Earthquakes: The Formation and Preservation of Pseudotachylytes. Heidelberg: Springer.Google Scholar
Lin, A. M., Maruyama, T., Aaron, S., Michibayashi, K., Camacho, A., and Kano, K. I. 2005. Propagation of seismic slip from brittle to ductile crust: Evidence from pseudotachylyte of the Woodroffe thrust, central Australia. Tectonophysics 402(1–4): 2135.Google Scholar
Linde, A. T., and Sacks, I. S. 2002. Slow earthquakes and great earthquakes along the Nankai trough. Earth Planet. Sci. Lett. 203(1): 265275, doi: 10.1016/s0012-821x(02)00849-x.Google Scholar
Linde, A. T., and Silver, P. G. 1989. Elevation changes and the great 1960 Chilean earthquake – support for aseismic slip. Geophys. Res. Lett. 16(11): 13051308, doi: 10.1029/GL016i011p01305.Google Scholar
Linde, A. T., Gladwin, M. T., Johnston, M. J. S., Gwyther, R. L., and Bilham, R. G. 1996. A slow earthquake sequence on the San Andreas fault. Nature 383: 6568.Google Scholar
Linde, A. T., Sacks, I. S., Johnston, M. J. S., Hill, D. P., and Bilham, R. G. 1994. Increased pressure from rising bubbles as a mechanism for remotely triggered seismicity. Nature 371: 408410.Google Scholar
Linde, A. T., Suyehiro, K., Miura, S., Sacks, I. S., and Takagi, A. 1988. Episodic aseismic earthquake precursors. Nature 334: 513515.Google Scholar
Lindh, A. G. 1983. Preliminary assessment of long-term probabilities for large earthquakes along the San Andreas fault system in California. US Geol Surv. Open File Report 83–63.Google Scholar
Lindh, A. G., and Boore, D. M. 1981. Control of rupture by fault geometry during the1966 Parkfield earthquake. Bull. Seismol. Soc. Am. 71(1): 95116.Google Scholar
Lindsey, E. O., Fialko, Y., Bock, Y., Sandwell, D. T., and Bilham, R. 2014. Localized and distributed creep along the southern San Andreas Fault. J. Geophys. Res.-Solid Earth 119(10): 79097922, doi: 10.1002/2014jb011275.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.-Solid Earth 97: 49234940.Google Scholar
Lipovsky, B. P., and Dunham, E. M. 2016. Tremor during ice-stream stick slip. The Cryosphere 10: 385399, doi: 10.5194/tc-10–385-2016.Google Scholar
Lippiello, E., Giacco, F., Marzocchi, W., Godano, C., and de Arcangelis, L. 2015. Mechanical origin of aftershocks. Sci. Rep. 5: 15560, doi: 10.1038/srep15560.Google Scholar
Lisowski, M., and Prescott, W. H. 1981. Short-range distance measurements along the San Andreas fault in central California, 1975 to 1979. Bull. Seismol. Soc. Am. 71: 16071624.Google Scholar
Lisowski, M., Savage, J. C., Prescott, W. H., and Gross, W. K. 1988. Absence of strain accumulation in the Shumagin seismic gap, Alaska, 1980–1987. J. Geophys. Res. 93: 79097922.Google Scholar
Lister, G. S., and Snoke, A. 1984. S-C mylonites. J. Struct. Geol. 6: 617638.Google Scholar
Lister, G. S., and Williams, P. F. 1979. Fabric development in shear zones: Theoretical controls and observed phenomena. J. Struct. Geol. 1: 283297.Google Scholar
Little, T. A. 1995. Brittle deformation adjacent to the Awatere strike-slip-fault in New-Zealand – Faulting patterns, scaling relationships, and displacement partitioning. Geol. Soc. Am. Bull. 107: 12551271.Google Scholar
Liu, Y. J., McGuire, J. J., and Behn, M. D. 2012. Frictional behavior of oceanic transform faults and its influence on earthquake characteristics. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb009025.Google Scholar
Liu, Y., and Rice, J. R. 2007. Spontaneous and triggered aseismic deformation transients in a subduction fault model. J. Geophys. Res.-Solid Earth 112: B0904Google Scholar
Liu-Zeng, J., Klinger, Y., Sieh, K., Rubin, C., and Seitz, G. 2006. Serial ruptures of the San Andreas fault, Carrizo Plain, California, revealed by three-dimensional excavations. Journal of Geophysical Research-Solid Earth 111(B2): doi: 10.1029/2004jb003601.Google Scholar
Livio, F., Michetti, A., Vittori, E., et al. 2016. Surface faulting during the August 24, 2016, central Italy earthquake. Mw 6.0): Preliminary results. Annals of Geophysics 59(Fast Track 5), 18.Google Scholar
Lockner, D. A. 1993. Room temperature creep in saturated granite. J. Geophys. Res. 98: 475–87.Google Scholar
Lockner, D. A. 1998. A generalized law for brittle deformation of Westerly granite. J. Geophys. Res.-Solid Earth 103: 51075123.Google Scholar
Lockner, D. A., and Beeler, N. M. 1999. Premonitory slip and tidal triggering of earthquakes. J. Geophys. Res.-Solid Earth 104: 2013320151.Google Scholar
Lockner, D. A., and Byerlee, J. D. 1977. Acoustic emission and fault formation in rocks. In Proc. 1st Conf. Microseismic Activity in Geological Structures and Materials, eds. Hardy, H. R. and Leighton, F. W.. Claustal: Trans-tech Publ., pp. 99107.Google Scholar
Lockner, D. A., and Byerlee, J. D. 1985. Complex resistivity of fault gouge and its significance for earthquake lights and induced polarization. Geophys. Res. Lett. 12: 211214.Google Scholar
Lockner, D. A., and Byerlee, J. D. 1993. How geometrical constraints contribute to the weakness of mature faults. Nature 363: 250252.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
Lockner, D. A., Byerlee, J. D., Kuksenko, V., Ponomarev, A., and Sidorin, A. 1992. Observations of quasistatic fault growth from acoustic emissions. In Fault Mechanics and Transport Properties of Rock, eds. Evans, B. and Wong, T.-F., San Diego: Academic Press, pp. 332.Google Scholar
Lockner, D. A., Morrow, C. A., Moore, D. E., and Hickman, S. 2011. Low strength of deep San Andreas fault gouge from SAFOD core. Nature 472: 8285, doi: 10.1038/nature09927.Google Scholar
Logan, J. M., and Teufel, L. W. 1986. The effect of normal stress on the real area of contact during frictional sliding in rocks. Pure Appl. Geophys. 124: 471486.Google Scholar
Logan, J. M., Dengo, C. A., Higgs, N. G., and Wang, Z. Z. 1992. Fabrics of experimental fault zones: their development and relationship to mechanical behavior. In Fault Mechanics and Transport Properties of Rock, eds. Evans, B. and Wong, T.-F., New York: Academic Press, pp. 3368.Google Scholar
Logan, J. M., Higgs, N., and Friedman, M. 1981. Laboratory studies on natural gouge from the U.S.G.S. Dry Lake Valley No. 1 Well, San Andreas fault zone. In Mechanical Behavior of Crustal Rocks. AGU Geophys. Mono., eds. Carter, M. F. N., Logan, J., and Sterns, D.. Washington, DC: American Geophysical Union, pp. 121134.Google Scholar
Lohman, R. B., and McGuire, J. J. 2007. Earthquake swarms driven by aseismic creep in the Salton Trough, California. J. Geophys. Res.-Solid Earth 112(B4): doi: 10.1029/2006jb004596.Google Scholar
Lomnitz, C. 1996. Search of a worldwide catalog for earthquakes triggered at intermediate distances. Bull. Seismol. Soc. Amer. 86: 293298.Google Scholar
Long, J. J., and Imber, J. 2011. Geological controls on fault relay zone scaling. J. Struct. Geol. 33(12): 17901800, doi: 10.1016/j.jsg.2011.09.011.Google Scholar
Long, J. J., and Imber, J. 2012. Strain compatibility and fault linkage in relay zones on normal faults. J. Struct. Geol. 36: 1626, doi: 10.1016/j.jsg.2011.12.013.Google Scholar
Lorenzo-Martin, F., Roth, F., and Wang, R. J. 2006. Inversion for rheological parameters from post-seismic surface deformation associated with the 1960 Valdivia earthquake, Chile. Geophys. J. Int. 164(1): 7587, doi: 10.1111/j.1365-246X.2005.02803.x.Google Scholar
Louderback, G. D. 1942. Faults and earthquakes. Bull. Seismol. Soc. Amer. 32: 305330.Google Scholar
Louie, J. N., Allen, C. R., Johnson, D., Haase, P., and Cohn, S. 1985. Fault slip in southern California. Bull. Seismol. Soc. Am. 75: 811834.Google Scholar
Loveless, J. P., and Meade, B. J. 2011. Spatial correlation of interseismic coupling and coseismic rupture extent of the 2011 M-W=9.0 Tohoku-oki earthquake. Geophys. Res. Lett. 38, doi: 10.1029/2011gl048561.Google Scholar
Loveless, J. P., and Meade, B. J. 2016. Two decades of spaciotemporal variations in subduction zone coupling offshore Japan. Earth Planet. Sci. Lett. 436: 1930, doi: 10.1016/j.epsl.2015.12.033.Google Scholar
Lu, X., Lapusta, N., and Rosakis, A. J. 2010. Pulse-like and crack-like dynamic shear ruptures on frictional interfaces: Experimental evidence, numerical modeling, and implications. Int. J. Fracture 163(1–2): 2739, doi: 10.1007/s10704-010–9479-4.Google Scholar
Lu, Z., and He, C. R. 2014. Frictional behavior of simulated biotite fault gouge under hydrothermal condition. Tectonophysics 622: 6280, doi: 10.1016/j.tecto.2014.03.002.Google Scholar
Lucente, F. P., De Gori, P., Margheriti, L., Piccinini, D., Di Bona, M., Chiarabba, C., and Agostinetti, N. P. 2010. Temporal variation of seismic velocity and anisotropy before the 2009 M-W 6.3 L’Aquila earthquake, Italy. Geology 38(11): 10151018, doi: 10.1130/g31463.1.Google Scholar
Lund, B., and Zoback, M. D. 1999. Orientation and magnitude of in situ stress to 6.5 km depth in the Baltic Shield. Int. J. Rock Mech. Min. Sci. 36: 169190.Google Scholar
Luthra, T., Anandakrishnan, S., Winberry, J. P., Alley, R. B., and Holschuh, N. 2016. Basal characteristics of the main sticky spot on the ice plain of Whillans Ice Stream, Antarctica. Earth Planet. Sci. Lett. 440: 1219, doi: 10.1016/j.epsl.2016.01.035.Google Scholar
Lyell, S. C. 1868. Principles of Geology. London: John Murray.Google Scholar
Lykotrafitis, G. and Rosakis, A. J. 2006. Dynamic sliding of frictionally held bimaterial interfaces subjected to impact shear loading. Proceedings of the Royal Society a-Mathematical Physical and Engineering Sciences 462(2074): 29973026, doi: 10.1098/rspa.2006.1703.Google Scholar
Ewing, M.; Ser. 4, eds. D. Simpson and P. Richards. Washington, DC: American Geophysical Union, pp. 2028.Google Scholar
Ma, K. F., Brodsky, E. E., Mori, J., Ji, C., Song, T. R. A., and Kanamori, H. 2003. Evidence for fault lubrication during the 1999 Chi-Chi, Taiwan, earthquake (Mw7.6). Geophysical Research Letters 30(5): doi: 10.1029/2002gl015380.Google Scholar
Ma, K. F., Mori, J., Lee, S. J., and Yu, S. B. 2001. Spatial and temporal distribution of slip for the 1999 Chi-Chi, Taiwan, earthquake. Bull. Seismol. Soc. Am. 91(5): 10691087.Google Scholar
Ma, K. F., Song, A., Lee, S. J., and Wu, H. I. 2000. Spatial slip distribution of the September 20, 1999, Chi-Chi, Taiwan, earthquake (Mw7.6): Inverted from teleseismic data. Geophys. Res. Lett. 27: 34173420.Google Scholar
Ma, S., and Beroza, G. C. 2008. Rupture dynamics on a bimaterial interface for dipping faults. Bull. Seismol. Soc. Am. 98(4): 16421658, doi: 10.1785/0120070201.Google Scholar
Macelwane, J. 1936. Problems and progress on the geologico-seismological frontier. Science 83: 193198.Google Scholar
Machette, M. N., Personius, S. F., Nelson, A. R., Schwartz, D. P., and Lund, W. R. 1991. The Wasatch fault zone, Utah segmentation and history of Holocene earthquakes. J. Struct. Geol. 13(2): 137149, doi: 10.1016/0191–8141(91)90062-n.Google Scholar
Madariaga, R. 1976. Dynamics of an expanding circular fault. Bull. Seismol. Soc. Am. 66: 636666.Google Scholar
Madariaga, R., Metois, M., Vigny, C., and Campos, J. 2010. Central Chile finally breaks. Science 328(5975): 181182, doi: 10.1126/science.1189197.Google Scholar
Maddock, R. H. 1983. Melt origin of fault-generated pseudotachylytes demonstrated by textures. Geology 11: 105108.Google Scholar
Maemoku, H. 2006. Quaternary crustal movement and coseismic uplift along Muroto Peninsula. J. Geol. Soc. Japan 112 , Supplement (in Japanese), 1726.Google Scholar
Mai, P. M., and Beroza, G. C. 2000. Source scaling properties from finite-fault-rupture models. Bull. Seismol. Soc. Amer. 90: 604615.Google Scholar
Mai, P. M., and Thingbaijam, K. 2014. SRCMOD: An online database of finite‐fault rupture models. Seismol. Res. Lett. 85(6): 13481357.Google Scholar
Main, I. G. 1997. Earthquakes – Long odds on prediction. Nature 385: 1920.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
Mainprice, D. H., and Paterson, M. S. 1984. Experimental studies of the role of water in the plasticity of quartzites. J. Geophys. Res. 89: 42574270.Google Scholar
Mair, K., Elphick, S., and Main, I. 2002. Influence of confining pressure on the mechanical and structural evolution of laboratory deformation bands. Geophys. Res. Lett. 29(10): 49-1–49-4, doi: 10.1029/2001gl013964.Google Scholar
Mair, K., Main, I., and Elphick, S. 2000. Sequential growth of deformation bands in the laboratory. J. Struct. Geol. 22(1): 2542.Google Scholar
Mandelbrot, B. B. 1977. Fractals: Form, Chance, and Dimension. San Francisco: W. H. Freeman.Google Scholar
Mandelbrot, B. B. 1983. The Fractal Geometry of Nature. San Francisco: W. H. Freeman.Google Scholar
Manga, M. 1998. Advective heat transport by low-temperature discharge in the Oregon Cascades. Geology 26: 799802.Google Scholar
Manga, M., Beresnev, I., Brodsky, E. E., Elkhoury, J. E., Elsworth, D., Ingebritsen, S. E., Mays, D. C., and Wang, C. Y. 2012. Changes in permeability caused by transient stresses: Field observations, experiments, and mechanisms. Rev. Geophys. 50, doi: 10.1029/2011rg000382.Google Scholar
Manga, M., Wang, C. Y., and Shirzaei, M. 2016. Increased stream discharge after the 3 September 2016 M-w 5.8 Pawnee, Oklahoma earthquake. Geophys. Res. Lett. 43(22): 1158811594, doi: 10.1002/2016gl071268.Google Scholar
Manighetti, I., Campillo, M., Bouley, S., and Cotton, F. 2007. Earthquake scaling, fault segmentation, and structural maturity. Earth Planet. Sci. Lett. 253(3–4): 429438, doi: 10.1016/j.epsl.2006.11.004.Google Scholar
Manighetti, I., Campillo, M., Sammis, C., Mai, P. M., and King, G. 2005. Evidence for self-similar, triangular slip distributions on earthquakes: Implications for earthquake and fault mechanics. J. Geophys. Res.-Solid Earth 110(B5): doi: 10.1029/2004jb003174.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.-Solid Earth 106(B7): 1366713696.Google Scholar
Manning, C. E., and Ingebritsen, S. E. 1999. Permeability of the continental crust: Implications of geothermal data and metamorphic systems. Rev. Geophys. 37: 127150.Google Scholar
Marcellini, A. 1997. Physical model of aftershock temporal behaviour. Tectonophysics 277: 137146.Google Scholar
Marchal, D., Guiraud, M., and Rives, T. 2003. Geometric and morphologic evolution of normal fault planes and traces from 2D to 4D data. J. Struct. Geol. 25(1): 135158, doi: 10.1016/s0191-8141(02)00011–1.Google Scholar
Marco, S., Stein, M., Agnon, A., and Ron, H. 1996. Long-term earthquake clustering: A 50,000-year paleoseismic record in the Dead Sea Graben. J. Geophys. Res.-Solid Earth 101(B3): 61796191, doi: 10.1029/95jb01587.Google Scholar
Marone, C. 1998a. Laboratory-derived friction laws and their application to seismic faulting. Ann. Rev. Earth Planet. Sci. 26: 643696.Google Scholar
Marone, C. 1998b. The effect of loading rate on static friction and the rate of fault healing during the earthquake cycle. Nature 391(6662): 6972, doi: 10.1038/34157.Google Scholar
Marone, C., and Kilgore, B. 1993. Scaling of the critical slip distance for seismic faulting with shear strain in fault zones. Nature 362: 618621.Google Scholar
Marone, C., and Scholz, C. 1988. The depth of seismic faulting and the upper transition from stable to unstable slip regimes. Geophys. Res. Lett. 15: 621624.Google Scholar
Marone, C., and Scholz, C. H. 1989. Particle-size distribution and microstructures within simulated fault gouge. J. Struct. Geol. 11: 799814.Google Scholar
Marone, C., Ellsworth, W., and Vidale, J. 1995. Fault healing inferred from time-dependent variations of source properties of repeating earthquakes. Geophys. Res. Lett. 22: 30953099.Google Scholar
Marone, C., Raleigh, C. B., and Scholz, C. H. 1990. Frictional behavior and constitutive modeling of simulated fault gouge. J. Geophys. Res. 95: 70077025.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., and Allmendinger, R. W. 1990. Kinematic analysis of fault-slip data. J. Struct. Geol. 12 (8): 973986.Google Scholar
Marrett, R., and Allmendinger, R. W. 1991. Estimates of strain due to brittle faulting – Sampling of fault populations. J. Struct. Geol. 13: 735738.Google Scholar
Marrett, R., and Allmendinger, R. W. 1992. Amount of extension on “small” faults: An example from the Viking graben. Geology 20: 4750.Google Scholar
Martel, S. J. 1999. Mechanical controls on fault geometry. J. Struct. Geol. 21(6): 585596.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
Martin, R. J., Price, R. H., Boyd, P. J., and Noel, J. S. 1995. Bulk and Mechanical Properties of the Paintbrush Tuff Recovered From Borehole USW NRG-7/7A: Data Report Rep., Sandia Report SAND94-1996 UC-814 Sandia National Laboratories.Google Scholar
Maruyama, T. 1963. On the force equivalents of dynamic elastic dislocations with reference to the earthquake mechanism. Bull. Earthquake Res. Inst., Univ. Tokyo 41: 467486.Google Scholar
Massonnet, D., Rossi, M., Carmona, C., Adragna, F., Peltzer, G., Feigl, K., and Rabaute, T. 1993. The displacement field of the Landers earthquake mapped by radar interferometry. Nature 364: 138142.Google Scholar
Massonnet, D., Thatcher, W., and Vadon, H. 1996. Detection of postseismic fault-zone collapse following the Landers earthquake. Nature 382: 612616.Google Scholar
Masterlark, T., and Wang, H. 2000. Poroelastic coupling between the 1992 Landers and Big Bear earthquakes. Geophys. Res. Lett. 27(22): 36473650, doi: 10.1029/2000gl011472.Google Scholar
Masterlark, T., and Wang, H. F. 2002. Transient stress-coupling between the 1992 Landers and 1999 Hector Mine, California, earthquakes. Bull. Seismol. Soc. Am. 92(4): 14701486, doi: 10.1785/0120000905.Google Scholar
Matsu’ura, T. 2015. Late Quaternary uplift rate inferred from marine terraces,Muroto Peninsula, southwest Japan: Forearc deformation in an oblique subduction zone. Geomorphology 234: 133150.Google Scholar
Matsu’ura, T., Furusawa, A., and Saomoto, H. 2008. Late Quaternary uplift rate of the northeastern Japan arc inferred from fluvial terraces. Geomorphology 95(3): 384397.Google Scholar
Matsubara, M., Obara, K., and Kasahara, K. 2009. High-V-P/V-S zone accompanying non-volcanic tremors and slow-slip events beneath southwestern Japan. Tectonophysics 472(1–4): 617.Google Scholar
Matsuda, T. 1977. Estimation of future destructive earthquakes from active faults in Japan. J. Phys. Earth, Suppl. 25: 795855.Google Scholar
Matsuda, T., Ota, Y., Ando, M., and Yonekura, N. 1978. Fault mechanism and recurrence time of major earthquakes in the southern Kanto district. Geol. Soc. Am. Bull. 89: 161016118.Google Scholar
Matsuura, M., and Iwasaki, T. 1983. Study on coseismic and postseismic crustal movements associated with the 1923 Kanto earthquake. Tectonophysics 97(1–4): 201215, doi: 10.1016/0040–1951(83)90148–8.Google Scholar
Matthews, M. V., Ellsworth, W. L., and Reasenberg, P. A. 2002. A Brownian model for recurrent earthquakes. Bull. Seismol. Soc. Am. 92(6): 22332250.Google Scholar
Maurer, J., and Johnson, K. 2014. Fault coupling and potential for earthquakes on the creeping section of the central San Andreas Fault. J. Geophys. Res.-Solid Earth 119(5): 44144428, doi: 10.1002/2013jb010741.Google Scholar
Mavko, G. M. 1981. Mechanics of motion on major faults. Ann. Rev. Earth Planet. Sci. 9: 81111.Google Scholar
Mavrommatis, A. P., Segall, P., and Johnson, K. M. 2014b. A decadal‐scale deformation transient prior to the 2011 Mw 9.0 Tohoku‐oki earthquake. Geophys. Res. Lett. 41(13): 44864494.Google Scholar
Mavrommatis, A. P., Segall, P., Uchida, N., and Johnson, K. M. 2015. Long‐term acceleration of aseismic slip preceding the Mw 9 Tohoku‐oki earthquake: Constraints from repeating earthquakes. Geophys. Res. Lett. 42(22): 97179725.Google Scholar
McAdoo, D. C., and Sandwell, D. T. 1985. Folding of oceanic lithosphere. J. Geophys. Res.-Solid Earth and Planets 90(NB10): 85638569, doi: 10.1029/JB090iB10p08563.Google Scholar
McCaffrey, R. 1992. Oblique plate convergence, slip vectors, and fore-arc deformation. J. Geophys. Res. 97: 89058915.Google Scholar
McCalpin, J., and Nishenko, S. 1996. Holocene paleoseismicity, temporal clustering, and probabilities of future large. M> 7) earthquakes on the Wasatch fault zone, Utah. J. Geophys. Res.-Solid Earth 101(B3): 62336253.Google Scholar
McCann, W. R., Nishenko, S. P., Sykes, L. R., and Krause, J. 1979. Seismic gaps and plate tectonics: Seismic potential for major boundaries. Pageoph 117: 10821147.Google Scholar
McClintock, F. A., and Walsh, J. B. 1962. Friction of Griffith cracks in rock under pressure. In Proc. 4th US Natl Congr. Appl. Mech., vol II. New York, New York: North American Society of Mechanical Engineering, pp. 10151021.Google Scholar
McCrory, P. A., Blair, J. L., Oppenheimer, D. H., and Walter, S. R. 2004. Depth to the Juan de Fuca slab beneath the Cascadia subduction margin: A 3-D model for sorting earthquakes [CD-Rom]. U.S. Geological Survey, Data Ser. 91.Google Scholar
McEvilly, T. V., and Johnson, L. R. 1974. Stability of P and S velocities from central California quarry blasts. Bull. Seismol. Soc. Amer. 64: 343353.Google Scholar
McGarr, A. 1976. Seismic moments and volume changes. J. Geophys. Res. 81: 14871494.Google Scholar
McGarr, A. 1977. Seismic moments of earthquakes beneath island arcs, phase changes, and subduction velocities. J. Geophys. Res. 82: 256264.Google Scholar
McGarr, A. 1984. Some applications of seismic source mechanism studies to assessing underground hazard. In Proc. Ist Int. Cong. Rockbursts and Seismicity in Mines, eds. Gay, N. C. and Wainwright, E. H.. Johannesburg: pp. 199208.Google Scholar
McGarr, A. 1999. On relating apparent stress to the stress causing earthquake fault slip. J. Geophys. Res.-Solid Earth 104: 30033011.Google Scholar
McGarr, A. 2014. Maximum magnitude earthquakes induced by fluid injection. J. Geophys. Res.-Solid Earth 119(2): 10081019, doi: 10.1002/2013jb010597.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., and Green, R. W. E. 1975. Measurement of tile in a deep-level gold mine and its relationship to mining and seismicity. Geophys. J. R.A. S. 43: 327345.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. In Proc. Conf. VIII – Analysis of Actual Fault-zones in Bedrock. US Geol. Surv. Open-file Rept. 791239, pp. 101120.Google Scholar
McGarr, A., Spottiswoode, S. M., Gay, N. C., and Ortlepp, W. D. 1979. Observations relevant to seismic driving stress, stress drop, and efficiency. J. Geophys. Res. 84: 22512261.Google Scholar
McGarr, A., Zoback, M. D., and Hanks, T. C. 1982. Implications of an elastic analysis of in situ stress measurements near the San Andreas fault. J. Geophys. Res. 87: 77977806.Google Scholar
McGill, S. F., Owen, L. A., Weldon, R. J., and Kendrick, K. J. 2013. Latest Pleistocene and Holocene slip rate for the San Bernardino strand of the San Andreas fault, Plunge Creek, Southern California: Implications for strain partitioning within thesouthern San Andreas fault system for the last ~35 k. Geol. Soc. Amer. Bull. 125: 4872, doi: 10.1130/B30647.1.Google Scholar
McGrath, A., 1992. Fault propagation and growth: A study of the Triassic and Jurassic from Wachet and Klive, North Somerset. Master’s thesis: London, Royal Holloway, University of London, 435.Google Scholar
McGrath, A. G., and Davison, I. 1995. Damage zone geometry around fault tips. J. Struct. Geol. 17(7): 10111024, doi: 10.1016/0191–8141(94)00116-h.Google Scholar
McGuire, J. J. 2008. Seismic cycles and earthquake predictability on east pacific rise transform faults. Bull. Seismol. Soc. Am. 98(3): 10671084, doi: 10.1785/0120070154.Google Scholar
McGuire, J. J., and Beroza, G. C. 2012. A rogue earthquake off Sumatra. Science 336(6085): 11181119.Google Scholar
McGuire, J. J., and Collins, J. A. 2013. Millimeter-level precision in a seafloor geodesy experiment at the Discovery transform fault, East Pacific Rise. Geochem. Geophys. Geosystems 14(10): 43924402, doi: 10.1002/ggge.20225.Google Scholar
McGuire, J. J., and Jordan, T. H. 2000. Further evidence for the compound nature of slow earthquakes: The Prince Edward Island earthquake of April 28, 1997. J. Geophys. Res.-Solid Earth 105: 78197827.Google Scholar
McGuire, J. J., Boettcher, M. S., and Jordan, T. H. 2005. Foreshock sequences and short-term earthquake predictability on East Pacific Rise transform faults (vol 434, p. 457, 2005). Nature 435(7041): 528528, doi: 10.1038/nature03621.Google Scholar
McGuire, J. J., Collins, J. A., Gouédard, P., Roland, E., Lizarralde, D., Boettcher, M. S., Behn, M. D., and Van Der Hilst, R. D. 2012. Variations in earthquake rupture properties along the Gofar transform fault, East Pacific Rise. Nat. Geosci. 5(5): 336.Google Scholar
McGuire, J. J., Ihmle, P. F., and Jordan, T. H. 1996. Time-domain observations of a slow precursor to the 1994 Romanche Transform earthquake. Science 274(5284): 8285.Google Scholar
McGuire, J. J., Wiens, D. A., Shore, P. J., and Bevis, M. G. 1997. The March 9, 1994 (M-w 7.6), deep Tonga earthquake: Rupture outside the seismically active slab. J. Geophys. Res.-Solid Earth 102(B7): 1516315182, doi: 10.1029/96jb03185.Google Scholar
McKay, A. 1890. On earthquakes of September, 1888, in the Amuri and Marlborough Districts of the South Island, New Zealand. New Zealand Geol Surv. Geol. Exploration 1888–1889 Rept. 20: 116.Google Scholar
McKenzie, D. P. 1969. The relation between fault plane solutions for earthquakes and the directions of the principal stresses. Bull. Seismol. Soc. Am. 59: 591601.Google Scholar
McKenzie, D. P., and Brune, J. N. 1972. Melting on fault planes during large earthquakes. Geophys. J. Roy. Astron. Soc. 29: 6578.Google Scholar
McLeod, A., Dawers, N. H., and Underhill, J. R. 2000. The propagation and linkage of normal faults: Insights from the Strathspey-Brent-Statfjord fault array, North Sea. Basin Res. 12: 263284.Google Scholar
McNally, K. 1981. Plate subduction and prediction of earthquakes along the Middle America trench. In Earthquake Prediction, an International Review, eds. Simpson, D. and Richards, P. G.. Washington, DC: Amer. Geophys. Union, pp. 6371.Google Scholar
McNally, K. C., and Gonzalez-Ruis, J. R. 1986. Predictability of the whole earthquake cycle and source mechanics for large (7.0 <Mw< 8.1) earthquakes along the Middle America trench offshore Mexico. Earthquake Notes 57: 22.Google Scholar
McSaveney, M. J., Graham, I. J., Begg, J. G., Beu, A. G., Hull, A. G., Kim, K., and Zondervan, A. 2006. Late Holocene uplift of beach ridges at Turakirae Head, south Wellington Coast, New Zealand. New Zeal. J. Geol. Geophys. 49(3): 337354.Google Scholar
Meade, C., and Jeanloz, R. 1989. Acoustic emissions and shear instabilities during phase-transformations in Si and Ge at ultrahigh pressures. Nature 339: 616618.Google Scholar
Meade, C., and Jeanloz, R. 1991. Deep-focus earthquakes and recycling of water into the Earth’s mantle. Science 252: 6872.Google Scholar
Means, W. D. 1984. Shear zones of type I and II and their significance for reconstructing rock history. Geol. Soc. Am. Prog. Abstr. 16: 50.Google Scholar
Mehta, A. P., Dahmen, K. A., and Ben-Zion, Y. 2006. Universal mean moment rate profiles of earthquake ruptures. Physical Review E 73(5): 056104, doi: 10.1103/PhysRevE.73.056104.Google Scholar
Mei, S., Suzuki, A. M., Kohlstedt, D. L., Dixon, N. A., and Durham, W. B. 2010. Experimental constraints on the strength of the lithospheric mantle. J. Geophys. Res. 115. B8).Google Scholar
Meier, M. A., Ampuero, J. P., and Heaton, T. 2017. The hidden simplicity of subduction megathrust earthquakes. Science 357: 12771281, doi: 10.1126/Science.aan5643.Google Scholar
Meier, M. A., Heaton, T., and Clinton, J. 2016. Evidence for universal earthquake rupture initiation behavior. Geophys. Res. Lett. 43(15): 79917996, doi: 10.1002/2016gl070081.Google Scholar
Meissner, R., and Strelau, J. 1982. Limits of stress in continental crust and their relation to the depth–frequency relation of shallow earthquakes. Tectonics 1: 7389.Google Scholar
Melbourne, T. I., Szeliga, W. M., Miller, M. M., and Santillan, V. M. 2005. Extent and duration of the 2003 Cascadia slow earthquake. Geophys. Res. Lett. 32(4): L04301.Google Scholar
Melgar, D., Fan, W., Riquelme, S., et al. 2016. Slip segmentation and slow rupture to the trench during the 2015, Mw8. 3 Illapel, Chile earthquake. Geophys. Res. Lett. 43(3): 961966.Google Scholar
Melnick, D. 2016. Rise of the central Andean coast by earthquakes straddling the Moho, Nature Geoscience. 9. 401406.Google Scholar
Melosh, H. J. 1996. Dynamical weakening of faults by acoustic fluidization. Nature 379: 601606.Google Scholar
Meltzner, A. J., Sieh, K., Chiany, H.-W., et al. 2015. Time-varying interseismic strain rates and similar seismic ruptures on the Nias-Simeulue patch of the Sunda megathrust. Quaternary Science Reviews 122: 258281, doi: 10.1016/j.quascirev.2015.06.003.Google Scholar
Mendoza, C., and Hartzell, S. H. 1988. Aftershock patterns and main shock faulting. Bull. Seismol. Soc. Am. 78: 14381449.Google Scholar
Meng, L. S., Ampuero, J. P., and Burgmann, R. 2014. The 2013 Okhotsk deep-focus earthquake: Rupture beyond themetastable olivinewedge and thermally controlled rise time near the edge of a slab. Geophys. Res. Lett. 41(11): 37793785, doi: 10.1002/2014gl059968.Google Scholar
Meng, L. S., Inbal, A., and Ampuero, J. P. 2011. A window into the complexity of the dynamic rupture of the 2011 Mw 9 Tohoku-Oki earthquake. Geophys. Res. Lett. 38, doi: 10.1029/2011gl048118.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. J. Geophys. Res. 75: 121.Google Scholar
Meredith, P. G., and Atkinson, B. K. 1985. Fracture toughness and subcritical crack grown during high temperature tensile deformaiton of westerly granite and black gabbro. Phys. Earth Planet. Int. 39(1): 3351, doi: 10.1016/0031–9201(85)90113-x.Google Scholar
Metois, M., Vigny, C., and Socquet, A. 2016. Interseismic Coupling, Megathrust Earthquakes and Seismic Swarms Along the Chilean Subduction Zone (38A degrees-18A degrees S). Pure Appl. Geophys. 173(5): 14311449, doi: 10.1007/s00024-016–1280-5.Google Scholar
Michael, A. J. 1987. Use of focal mechanisms to determine stress – A control study. J. Geophys. Res. 92: 357368.Google Scholar
Michael, A. J. 1990. Energy constraints on kinematic models of oblique faulting: Loma Prieta versus Parkfield-Coalinga. Geophys. Res. Lett. 17: 14531456.Google Scholar
Michael, A. J. 2012. Fundamental questions of earthquake statistics, source behavior, and the estimation of earthquake probabilities from possible foreshocks. Bull. Seismol. Soc. Am. 102(6): 25472562.Google Scholar
Michael, A. J., and Eberhart-Phillips, D. 1991. Relations among fault behavior, subsurface geology, and three-dimensional velocity models, Science, 253(5020): 651654.Google Scholar
Michas, G., Vallianatos, F., and Sammonds, P. 2015. Statistical mechanics and scaling of fault populations with increasing strain in the Corinth Rift. Earth Planet. Sci. Lett. 431: 150163, doi: 10.1016/j.epsl.2015.09.014.Google Scholar
Michel, S., Avouac, J. P., Jolivet, R., and Wang, L. 2018. Seismic and aseismic moment budget and implications for the seismic potential of the Parkfield segment on The San Andreas fault. Bull. Seismol. Soc. Amer. 108(1): 1938.Google Scholar
Mildon, Z. K., Roberts, G. P., Walker, J. P. F., and Iezzi, F. 2017. Coulomb stress transfer and fault interaction over millennia on non-planar active normal faults: the M-w 6.5–5.0 seismic sequence of 2016–2017, central Italy. Geophys. J. Int. 210(2): 12061218, doi: 10.1093/gji/ggx213.Google Scholar
Miller, D. D. 1998. Distributed shear, rotation, and partitioned strain along the San Andreas fault, central California. Geology 26: 867870.Google Scholar
Miller, S. A. 1996. Fluid-mediated influence of adjacent thrusting on the seismic cycle at Parkfield. Nature 382: 799802.Google Scholar
Miller, S. A., Ben-Zion, Y., and Burg, J. P. 1999. A three-dimensional fluid-controlled earthquake model: Behavior and implications. J. Geophys. Res.-Solid Earth 104: 1062110638.Google Scholar
Miller, S. A., Collettini, C., Chiaraluce, L., Cocco, M., Barchi, M., and Kaus, B. J. P. 2004. Aftershocks driven by a high-pressure CO2 source at depth. Nature 427(6976): 724727, doi: 10.1038/nature02251.Google Scholar
Mindlin, R. D. 1949. Compliance of elastic bodies in contact. Trans. ASME, Series E, J. Appl. Mech. 16: 1334.Google Scholar
Mindlin, R. D., and Deresiewicz, H. 1953. Elastic spheres in contact under varying oblique forces. Trans. ASME, Series E, J. Appl. Mech. 20: 327353.Google Scholar
Minoura, K., Imamura, F., Sugawara, D., Kono, Y., and Isashita, T. 2001. The 869 Jogan Tsunami deposit and recurrence interval of large-scale tsunami on the Pacific coast of Japan. J. Nat. Disaster Sci. 23: 8388.Google Scholar
Minson, S. E., Brooks, B. A., Glennie, C. L., et al. 2015. Crowdsourced earthquake early warning. Science Advances 1(3): e1500036.Google Scholar
Minson, S. E., Wu, S., Beck, J. L., and Heaton, T. 2017. Combining Multiple Earthquake Models in Real Time for Earthquake Early Warning. Bull. Seismol. Soc. Amer. 107(4): 18681882, doi: 10.1785/0120160331.Google Scholar
Minster, J. B., and Jordan, T. H. 1978. Present day plate motions. J. Geophys. Res. 83: 55315554.Google Scholar
Minster, J. B., and Jordan, T. H. 1987. Vector constraints on western US deformation from space geodesy, neotectonics, and plate motions. J. Geophys. Res. 92: 47984809.Google Scholar
Misra, S., Boutareaud, S., and Burg, J. P. 2014. Rheology of talc sheared at high pressure and temperature: A case study for hot subduction zones. Tectonophysics 610: 5162, doi: 10.1016/j.tecto.2013.10.009.Google Scholar
Mitchell, E. K., Fialko, Y., and Brown, K. M. 2015. Frictional properties of gabbro at conditions corresponding to slow slip events in subduction zones. Geochem. Geophys. Geosystems 16(11): 40064020, doi: 10.1002/2015gc006093.Google Scholar
Mitchell, T. M., and Faulkner, D. R. 2009. The nature and origin of off-fault damage surrounding strike-slip fault zones with a wide range of displacements: A field study from the Atacama fault system, northern Chile. J. Struct. Geol. 31(8): 802816, doi: 10.1016/j.jsg.2009.05.002.Google Scholar
Mitchell, T. M., Ben-Zion, Y., and Shimamoto, T. 2011. Pulverized fault rocks and damage asymmetry along the Arima-Takatsuki Tectonic Line, Japan. Earth Planet. Sci. Lett. 308(3–4): 284297, doi: 10.1016/j.epsl.2011.04.023.Google Scholar
Mitchell, T. M., Billi, A., Miller, S. A., Goldsby, D. L., Scholz, C. H., Gran, J. K., and Simons, J. 2013. Dynamic pulverization by rapid decompression, edited, American Geophysical Union, Fall Meeting 2013, abstract #MR41B-04.Google Scholar
Mitra, G. 1984. Brittle to ductile transition due to large strains along the White Rock thrust, Wind River Mountains, Wyoming. J. Struct. Geol. 6: 5161.Google Scholar
Miyatake, T. 1992. Reconstruction of dynamic rupture of an earthquake with constraints of kinematic parameters. Geophys. Res. Lett. 19: 349352.Google Scholar
Mjachkin, V. I., Brace, W. F., Sobolev, G. A., and Dieterich, J. H. 1975. 2 Models for earthquake forerunners. Pure Appl. Geophys. 113: 169181.Google Scholar
Mogi, K. 1962. Study of elastic shocks caused by the fracture of heterogenous materials and their relation to earthquake phenomena. Bull. Earthquake Res. Inst. Univ. Tokyo 40: 125173.Google Scholar
Mogi, K. 1963. Some discussions on aftershocks, foreshocks and earthquake swarms – The fracture of a semi-infinite body caused by inner stress origin and its relation to the earthquake phenomena (3). Bull. Earthquake Res. Inst. Univ. Tokyo 41: 615658.Google Scholar
Mogi, K. 1966. Some features of recent seismicity in and near Japan (2). Activity before and after great earthquakes. Bull. Earthquake Res. Inst. Univ. Tokyo 47: 395417.Google Scholar
Mogi, K. 1969. Relationship between the occurrence of great earthquakes and tectonic structures. Bull. Earthquake Res. Inst. Univ. Tokyo 47: 429441.Google Scholar
Mogi, K. 1977. Seismic activity and earthquake prediction. Proceedings of the National Symposium on Earthquake Prediction Researches: pp. 203214.Google Scholar
Mogi, K. 1979. Two kinds of seismic gaps. Pageoph 117: 11721186.Google Scholar
Mogi, K. 1981. Seismicity in western Japan and long term earthquake forecasting. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 4351.Google Scholar
Mogi, K. 1982. Temporal variation of the precursory crustal deformation just prior to the 1944 Tonankai earthquake. J. Seismol. Soc. Japan 35: 145148.Google Scholar
Mogi, K. 1985. Earthquake Prediction. Tokyo: Academic Press.Google Scholar
Mogi, K. 1995. Earthquake prediction research in Japan. J. Phys. Earth 43: 533561.Google Scholar
Mohr, C. H., Manga, M., Wang, C. Y., and Korup, O. 2017. Regional changes in streamflow after a megathrust earthquake. Earth Planet. Sci. Lett. 458: 418428, doi: 10.1016/j.epsl.2016.11.013.Google Scholar
Molchan, G., Kronrod, T., and Nekrasova, A. 1999. Immediate foreshocks: Time variation of the b-value. Phys. Earth Planet. Int. 111(3): 229240.Google Scholar
Molchanov, O. A., Kopytenko, Y. A., Voronov, P. M., et al. 1992. Results of Ulf magnetic-field measurements near the epicenters of the Spitak (M(S) = 6.9) and Loma-Prieta. M(S) = 7.1) earthquakes – Comparative analysis. Geophys. Res. Lett. 19: 14951498.Google Scholar
Molnar, P. 1979. Earthquake recurrence intervals and plate tectonics. Bull. Seismol. Soc. Am. 69(1): 115133.Google Scholar
Molnar, P. 1992. Brace-Goetze strength profiles, the partitioning of strike-slip and thrust faulting at zones of oblique convergence, and the stress-heat flow paradox of the San Andreas fault. In Fault Mechanics and Transport Properties of Rocks, eds. Evans, B. and Wong, T.-F.. London: Academic Press, pp. 435460.Google Scholar
Molnar, P., and Deng, Q. 1984. Faulting associated with large earthquakes and the average rate of deformation in central and east Asia. J. Geophys. Res. 89: 62036214.Google Scholar
Molnar, P. and Wyss, M. 1972. Moments, source dimensions, and stress drops of shallow-focus earthquakes in the Tonga-Kermadec arc. Phys. Earth Planet. Int. 6: 263278.Google Scholar
Molnar, P., Atwater, T., Mammerickx, J., and Smith, S. M. 1975. Magnetic anomalies, bathymetry and tectonic evolution of the South Pacific since the Late Cretaceous. Geophys. J. R.A.S. 40: 383420.Google Scholar
Molnar, P., Chen, W.-P., and Padovani, E. 1983. Calculated temperatures in overthrust terrains and possible combinations of heat sources responsible for the tertiary granites in the Greater Himalaya. J. Geophys. Res. 88: 64156429.Google Scholar
Monecke, K., Finger, W., Klarer, D., Kongko, W., McAdoo, B. G., Moore, A. L., and Sudrajat, S. U. 2008. A 1,000-year sediment record of tsunami recurrence in northern Sumatra. Nature 455(7217): 12321234.Google Scholar
Mooney, W. D., and Ginsberg, A. 1986. Seismic measurements of the internal properties of fault zones. Pageoph 124: 141158.Google Scholar
Moore, D. E., Lockner, D. A., Tanaka, H., and Iwata, K.. 2004. The coefficient of friction of Chrysotile gouge at seismogenic depths. International Geology Review 46(5): 385398, doi: 10.2747/0020–6814.46.5.385.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, D. E., and Lockner, D. A. 2004. Crystallographic controls on the frictional behavior of dry and water-saturated sheet structure minerals. J. Geophys. Res.-Solid Earth 109(B3): B03401, doi: 10.1029/2003jb002582.Google Scholar
Moore, D. E., and Lockner, D. A. 2007a. Comparative deformation behavior of minerals in serpentinized ultramafic rock: Application to the slab-mantle interface in subduction zones. Int. Geol. Rev. 49(5): 401415, doi: 10.2747/0020–6814.49.5.401.Google Scholar
Moore, D. E., and Lockner, D. A. 2007b. Friction of the smectite clay montmorillonite. In The Seismogenic Zone of Subduction Thrust Faults, eds. Dixon, T. and Moore, C., New York: Columbia Univ. Press, pp. 317345.Google Scholar
Moore, D. E., and Lockner, D. A. 2008. Talc friction in the temperature range 25 degrees-400 degrees C: Relevance for fault-zone weakening. Tectonophysics 449(1–4): 120132, doi: 10.1016/j.tecto.2007.11.039.Google Scholar
Moore, D. E., and Lockner, D. A. 2011. Frictional strengths of talc-serpentine and talc-quartz mixtures. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb007881.Google Scholar
Moore, D. E., and Lockner, D. A. 2013. Chemical controls on fault behavior: Weakening of serpentinite sheared against quartz-bearing rocks and its significance for fault creep in the San Andreas system. J. Geophys. Res.-Solid Earth 118(5): 25582570, doi: 10.1002/jgrb.50140.Google Scholar
Moore, D. E., and Rymer, M. J. 2007. Talc-bearing serpentinite and the creeping section of the San Andreas fault. Nature 448(7155): 795797.Google Scholar
Moore, D. E., Lockner, D. A., and Hickman, S.. 2016. Hydrothermal frictional strengths of rock and mineral samples relevant to the creeping section of the San Andreas Fault. J. Struct. Geol. 89: 153167, doi: 10.1016/j.jsg.2016.06.005.Google Scholar
Moore, D. E., Lockner, D. A., Ma, S. L., Summers, R., and Byerlee, J. D. 1997. Strengths of serpentinite gouges at elevated temperatures. J. Geophys. Res.-Solid Earth 102: 1478714801.Google Scholar
Moore, D. E., Lockner, D. A., Summers, R., Shengli, M., and Byerlee, J. D. 1996. Strength of chrysotile-serpentinite gouge under hydrothermal conditions: Can it explain a weak San Andreas fault? Geology 24: 10411044.Google Scholar
Mori, J., and Abercrombie, R. E. 1997. Depth dependence of earthquake frequency‐magnitude distributions in California: Implications for rupture initiation. J. Geophys. Res.-Solid Earth 102(B7): 1508115090.Google Scholar
Mori, J., Hudnut, K., Jones, L., Hauksson, E., and Hutton, K. 1992. Rapid scientific response to Landers quake. Eos 73: 417418.Google Scholar
Moro, M., Saroli, M., Stramondo, S., et al. 2017. New insights into earthquake precursors from InSAR. Scientific Reports 7: 12035, doi: 10.1038/s41598-017–12058-3.Google Scholar
Morrow, C. A., Moore, D. E., and Lockner, D. A. 2000. The effect of mineral bond strength and adsorbed water on fault gouge frictional strength. Geophys. Res. Lett. 27: 815818.Google Scholar
Morrow, C. A., Radney, B., and Byerlee, J. D. 1992. Frictional strength and the effective pressure law of montmorillonite and illite clays. In Fault Mechanics and Transport Properties of Rocks, eds. Evans, B. and Wong, T.-F.. Academic Press, pp. 6988.Google Scholar
Mott, N. F. 1948. Brittle fracture in mild steel plates. Engineering 165: 1624.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
Mouslopoulou, V., Walsh, J. J., and Nicol, A. 2009. Fault displacement rates on a range of timescales. Earth and Planetary Science Letters 278(3–4): 186197, doi: 10.1016/j.epsl.2008.11.031.Google Scholar
Muir-Wood, R., and King, G. C. P. 1993. Hydrological signatures of earthquake strain. J. Geophys. Res. 98: 2203522068.Google Scholar
Muller, W., Prosser, G., Mancktelow, N. S., et al. 2001. Geochronological constraints on the evolution of the periadriatic faults system (Alps). Int. J. Earth Sciences (Geol. Rundsch.) 90(3): 623653.Google Scholar
Muraoka, H., and Kamata, H. 1983. Displacement distribution along minor fault traces. J. Struct. Geol. 4: 483495.Google Scholar
Murray, J., and Langbein, J. 2006. Slip on the San Andreas fault at Parkfield, California, over two earthquake cycles, and the implications for seismic hazard. Bull. Seismol. Soc. Am. 96(4): S283-S303, doi: 10.1785/0120050820.Google Scholar
Murray, J., and Segall, P. 2002. Testing time-predictable earthquake recurrence by direct measurement of strain accumulation and release. Nature 419(6904): 287291, doi: 10.1038/nature00984.Google Scholar
Musha, K. 1943. Zotei Dainihon Jisin Shiryo. Tokyo: Shinsai Yobo Hyogi Kai.Google Scholar
Myers, C., Shaw, B., and Langer, J. 1996. Slip complexity in a crustal plane model of an earthquake fault. Phys. Rev. Lett. 77: 972975.Google Scholar
Nabelek, J., Chen, W.-P., and Ye, H. 1987. The Tangshan earthquake sequence and its implications for the evolution of the North China basin. J. Geophys. Res. 92: 1261512628.Google Scholar
Nabelek, P. I., and Liu, M. A. 1999. Leucogranites in the Black Hills of South Dakota: The consequence of shear heating during continental collision. Geology 27: 523526.Google Scholar
Nadeau, R. M., and Dolenc, D. 2005. Nonvolcanic tremors deep beneath the San Andreas Fault. Science 307(5708): 389389.Google Scholar
Nadeau, R. M., and Johnson, L. R. 1998. Seismological studies at Parkfield VI: Moment release rates and estimates of source parameters for small repeating earthquakes. Bull. Seismol. Soc. Amer. 88: 790814.Google Scholar
Nadeau, R. M., and McEvilly, T. V. 1999. Fault slip rates at depth from recurrence intervals of repeating microearthquakes. Science 285: 718721.Google Scholar
Nadeau, R. M., Foxall, W., and McEvilly, T. V. 1995. Clustering and periodic recurrence of microearthquakes on the San-Andreas fault at Parkfield, California. Science 267: 503507.Google Scholar
Nagata, K., Nakatani, M., and Yoshida, S. 2012. A revised rate- and state-dependent friction law obtained by constraining constitutive and evolution laws separately with laboratory data. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb008818.Google Scholar
Nakajima, J., Tsuji, Y., Hasegawa, A., Kita, S., Okada, T., and Matsuzawa, T. 2009. Tomographic imaging of hydrated crust and mantle in the subducting Pacific slab beneath Hokkaido, Japan: Evidence for dehydration embrittlement as a cause of intraslab earthquakes. Gondwana Research 16(3–4): 470481, doi: 10.1016/j.gr.2008.12.010.Google Scholar
Nakamura, K. 1969. Arrangement of parasitic cones as a possible key to a regional stress field. Bull. Volcanol. Soc. Japan 14: 820.Google Scholar
Nakamura, K., Jacob, K., and Davies, J. 1977. Volcanoes as possible indicators of tectonic stress indicators – Aleutians and Alaska. Pageoph 115: 86112.Google Scholar
Nakamura, K., Shimazaki, K., and Yonekura, N. 1984. Subduction, bending, and eduction. Present and Quaternary tectonics of the northem border of the Philippine Sea plate. Bull. Soc. Geol. France 26: 221243.Google Scholar
Nakamura, Y. 1988. On the urgent earthquake detection and alarm system (UrEDAS), Proceedings of the 13th World Conference on Earthquake Engineering, VII: 673678.Google Scholar
Nakamura, Y., and Saita, J. 2007. UrEDAS, the earthquake warning system, Today and tomorrow. In Earthqauke Early Earning Sytems, eds. Gasparini, P., Manfredi, G. and Zschau, J., Berlin and Heidelberg: Springer, pp. 249282,Google Scholar
Nakamura, Y., Muto, J., Nagahama, H., Shimizu, I., Miura, T., and Arakawa, I. 2012. Amorphization of quartz by friction: Implication to silica-gel lubrication of fault surfaces. Geophys. Res. Lett. 39, doi: 10.1029/2012gl053228.Google Scholar
Nakano, H. 1923. Notes on the nature of the forces which give rise to the earthquake motions. Seismol. Bull. Central Meteorol. Obs., Japan 1: 92120.Google Scholar
Nakata, T., Takahashi, T., and Koba, M. 1978. Holocene emerged coral reefs and sea level changes in the Ryukyu Islands. Geograph. Rev. Japan 51: 87108.Google Scholar
Nakatani, M. 2001. Conceptual and physical clarification of rate and state variable friction. J. Geophys. Res. 106: 1334713380.Google Scholar
Nakatani, M., and Scholz, C. H. 2004a. Frictional healing of quartz gouge under hydrothermal conditions:1. Experimental evidence for solution transfer healing mechanism. J. Geophys. Res. 109, doi: 10.1029/2001JB001522.Google Scholar
Nakatani, M., and Scholz, C. H. 2004b. Frictional healing of quartz gouge under hydrothermal conditions: 2. Quantitative interpretation with a physical model. J. Geophys. Res.-Solid Earth 109(B7): doi: 10.1029/2003jb002938.Google Scholar
Nakatani, M., and Scholz, C. H. 2006. Intrinsic and apparent short-time limits for fault healing: Theory, observations, and implications for velocity-dependent friction. J. Geophys. Res.-Solid Earth 111(B12).Google Scholar
Nakatani, M., Kaneshima, S., and Fukao, Y. 2000. Size-dependent microearthquake initiation inferred from high-gain and low-noise observations at Nikko district, Japan. J. Geophys. Res.-Solid Earth 105(B12): 2809528109, doi: 10.1029/2000jb900255.Google Scholar
Nalbant, S. S., Hubert, A., and King, G. C. P. 1998. Stress coupling between earthquakes in northwest Turkey and the north Aegean Sea. J. Geophys. Res. 103: 24,46924,486.Google Scholar
Nanayama, F., Satake, K., Furukawa, R., et al. 2003. Unusually large earthquakes inferred from tsunami deposits along Kurile trench. Nature 424: 660663.Google Scholar
Nanjo, K. Z., Hirata, N., Obara, K., and Kasahara, K. 2012. Decade-scale decrease in b value prior to the M9-class 2011 Tohoku and 2004 Sumatra quakes. Geophys. Res. Lett. 39: 660663, doi: 10.1029/2012gl052997.Google Scholar
Naoi, M., Nakatani, M., Yabe, Y., Kwiatek, G., Igarashi, T., and Plenkers, K. 2011. Twenty thousand aftershocks of a very small (M 2) earthquake and their relation to the mainshock rupture and geological structures. Bull. Seismol. Soc. Am. 101(5): 23992407.Google Scholar
Narr, W., and Suppe, J. 1991. Joint spacing in sedimentary rocks. J. Struct. Geol. 13(9): 1037+, doi: 10.1016/0191–8141(91)90055-n.Google Scholar
Narteau, C., Byrdina, S., Shebalin, P., and Schorlemmer, D. 2009. Common dependence on stress for the two fundamental laws of statistical seismology. Nature 462(7273): 642645, doi: 10.1038/nature08553.Google Scholar
Naylor, M. A., Mandl, G., and Sijpesteijn, C. H. 1986. Fault geometries in basement-induced wrench faulting under different initial stress states. J. Struct. Geol. 8: 737752.Google Scholar
Nelson, A. R., Kelsey, H. M., and Witter, R. C. 2006. Great earthquakes of variable magnitude at the Cascadia subduction zone. Quaternary Research 65(3): 354365, doi: 10.1016/j.yqres.2006.02.009.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
Newcomb, K. R., and McCann, W. R. 1987. Seismic history and seismotectonics of the Sunda arc. J. Geophys. Res.-Solid Earth and Planets 92(B1): 421439, doi: 10.1029/JB092iB01p00421.Google Scholar
Newman, A. V., and Okal, E. A. 1998. Teleseismic estimates of radiated seismic energy: The E/M-0 discriminant for tsunami earthquakes. J. Geophys. Res.-Solid Earth 103(B11): 2688526898, doi: 10.1029/98jb02236.Google Scholar
Newman, M. E. J. 2006. Power laws, Pareto distributions and Zipf’s law. Contemp. Phys. doi.org/10.1080/00107510500052444.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(2): 327342, doi: 10.1016/j.jsg.2004.09.002.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
Nicolas, A., and Poirier, J.-P. 1976. Crystalline Plasticity and Solid State Flow in Metamorphic Rocks. New York: John Wiley.Google Scholar
Nielsen, S., Di Toro, G., Hirose, T., and Shimamoto, T. 2008. Frictional melt and seismic slip. J. Geophys. Res. 113: B01308.Google Scholar
Nielsen, S., Spagnolo, E., Smith, S. A. F., Violay, M., DiToro, G., and Bistacchi, A. 2016. Scaling in natural and laboratory earthquakes. Geophys. Res. Lett. 43: 15041510: doi.org/10.1002/2015GL067490.Google Scholar
Nielsen, S., Mosca, P., Giberti, G., Di Toro, G., Hirose, T., and Shimamoto, T. 2010. On the transient behavior of frictional melt during seismic slip. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb007020.Google Scholar
Niemeijer, A., Di Toro, G., Griffith, W. A., Bistacchi, A., Smith, S. A. F., and Nielsen, S. 2012. Inferring earthquake physics and chemistry using an integrated field and laboratory approach. J. Struct. Geol. 39: 236, doi: 10.1016/j.jsg.2012.02.018.Google Scholar
Niemeijer, A. R., and Spiers, C. J. 2006. Velocity dependence of strength and healing behaviour in simulated phyllosilicate-bearing fault gouge. Tectonophysics 427(1–4): 231253, doi: 10.1016/j.tecto.2006.03.048.Google Scholar
Nishenko, S. P. 1985. Seismic potential for large and great interplate earthquakes along the Chilean and southern Peruvian margins of South America – A quantitative reappraisal. J. Geophys. Res. 90: 35893615.Google Scholar
Nishenko, S. P., and Buland, R. 1987. A generic recurrence interval distribution for earthquake forecasting. Bull. Seismol. Soc. Am. 77: 13821399.Google Scholar
Nishenko, S. P., and McCann, W. R. 1981. Seismic potential of the world’s major plate boundaries. In Earthquake Prediction, an International Review,Google Scholar
Noda, H., and Lapusta, N. 2013. Stable creeping fault segments can become destructive as a result of dynamic weakening. Nature 493(7433): 518+, doi: 10.1038/nature11703.Google Scholar
Noda, H., Dunham, E. M., and Rice, J. R. 2009. Earthquake ruptures with thermal weakening and the operation of major faults at low overall stress levels. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb006143.Google Scholar
Noda, H., Nakatani, M., and Hori, T. 2013. Large nucleation before large earthquakes is sometimes skipped due to cascade-up-Implications from a rate and state simulation of faults with hierarchical asperities. J. Geophys. Res.-Solid Earth 118(6): 29242952, doi: 10.1002/jgrb.50211.Google Scholar
Nur, A. 1972. Dilatancy, pore fluids, and premonitory variations in ts/tp travel times. Bull. Seism. Soc. Amer. 62: 12171222.Google Scholar
Nur, A. 1974. Matsushiro, Japan earthquake swarm: Confirmation of the dilatancy-fluid diflusion model. Geology 2: 217221.Google Scholar
Nur, A. 1978. Nonuniform friction as a basis for earthquake mechanics. Pageoph 116: 964989.Google Scholar
Nur, A. 1981. Rupture mechanics of plate boundaries. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 629634.Google Scholar
Nur, A., and Byerlee, J. D. 1971. An exact effective stress law for elastic deformation of rock with fluids. J. Geophys. Res. 76: 64146419.Google Scholar
Nur, A., and Mavko, G. 1974. Postseismic viscoelastic rebound. Science 183: 204206.Google Scholar
Nur, A., and Ron, H. 2003. Material and stress rotations: The key to reconciling crustal faulting complexity with rock mechanics. Int. Geol. Rev. 45(8): 671690, doi: 10.2747/0020–6814.45.8.671Google Scholar
Nuttli, O. W. 1973. The Mississippi Valley earthquakes of 1811–12: Intensities, ground motion and magnitudes. Bull. Seismol. Soc. Am. 63: 227248.Google Scholar
O’Connell, R., and Budiansky, B. 1974. Seismic velocities in dry and saturated cracked solids. J. Geophys. Res. 79: 54125426.Google Scholar
O’Neil, J. R., and Hanks, T. C. 1980. Geochemical evidence for water-rock interaction along the San Andreas and Garlock faults of California. J. Geophys. Res. 85: 62866292.Google Scholar
O’Reilly, W., and Rastogi, B. K. 1986. Induced seismicity. Phys. Earth Planet int. 44: 73199.Google Scholar
Obana, K., Fujie, G., Takahashi, T., et al. 2012. Normal‐faulting earthquakes beneath the outer slope of the Japan Trench after the 2011 Tohoku earthquake: Implications for the stress regime in the incoming Pacific plate. Geophys. Res. Lett. 39(7).Google Scholar
Obana, K., Kodaira, S., Shinohara, M., et al. 2013. Aftershocks near the updip end of the 2011 Tohoku-Oki earthquake. Earth Planet. Sci. Lett. 382: 111116.Google Scholar
Obara, K. 2002. Nonvolcanic deep tremor associated with subduction in southwest Japan. Science 296(5573): 16791681.Google Scholar
Obara, K. 2010. Phenomenology of deep slow earthquake family in southwest Japan: Spatiotemporal characteristics and segmentation. J. Geophys. Res.-Solid Earth, 115.Google Scholar
Obara, K., and Kato, A. 2016. Connecting slow earthquakes to huge earthquakes. Science 353(6296): 253257, doi: 10.1126/science.aaf1512.Google Scholar
Obriemoff, J. W. 1930. The splitting strength of mica. Proc. Roy. Soc. London, ser. A 127: 290302.Google Scholar
Ogata, Y. 1998. Space-time point-process models for earthquake occurrences. Annals of the Institute of Statistical Mathematics 50(2): 379402, doi: 10.1023/a:1003403601725Google Scholar
Ogata, Y. 1999. Seismicity analysis through point-process modeling: A review. Pure Appl. Geophys 155(2–4): 471507, doi: 10.1007/s000240050275.Google Scholar
Oglesby, D. 2008. Rupture termination and jump on parallel offset faults. Bull. Seismol. Soc. Am. 98(1): 440447, doi: 10.1785/0120070163.Google Scholar
Oglesby, D. D., and Day, S. M. 2001. Fault geometry and the dynamics of the 1999 Chi-Chi (Taiwan) earthquake. Bull. Seismol. Soc. Am. 91(5): 10991111.Google Scholar
Ohnaka, M. 1975. Frictional characteristics of typical rocks. J. Phys Earth 23: 87102.Google Scholar
Ohnaka, M. 1992. Earthquake source nucleation – A physical model for short-term precursors. Tectonophysics 211: 149178.Google Scholar
Ohnaka, M. 1993. Critical size of the nucleation zone of earthquake rupture inferred from immediate foreshock activity. J. Phys. Earth 41: 4556.Google Scholar
Ohnaka, M. 2000. A physical scaling relation between the size of an earthquake and its nucleation zone size. Pure Appl. Geophys 157(11–12): 22592282, doi: 10.1007/pl00001084.Google Scholar
Ohnaka, M., Kuwahana, Y., Yamamoto, K., and Hirasawa, T. 1986. Dynamic breakdown processes and the generating mechanism for high frequency elastic radiation during stick-slip instabilities. In Earthquake Source Mechanics. AGU Geophys. Mono. 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 1324.Google Scholar
Ohta, Y., HinoR., Inazu, D., et al. 2012. Geodetic constraints on afterslip characteristics following the March 9, 2011, Sanriku-oki earthquake, Japan. Geophys. Res. Lett. 39, doi: 10.1029/2012gl052430.Google Scholar
Ohtake, M., Matumoto, T., and Latham, G. V. 1977. Seismicity gap near Oaxaca, southern Mexico as a probable precursor to a large earthquake. Pure Appl. Geophys. 115: 375385.Google Scholar
Ohtake, M., Matumoto, T., and Latham, G. 1981. Evaluation of the forecast of the 1978 Oaxaca, southern Mexico earthquake based on a precursory seismic quiescence. In Earthquake Prediction: An International Review, Ewing, M., Ser. 4 eds. Simpson, D. P. and Richards, P. G.. Washington, DC: American Geophysical Union, pp. 5362.Google Scholar
Ohuchi, T., Lei, X., Ohfuji, H., et al. 2017. Intermediate-depth earthquakes linked to localized heating in dunite and harzburgite. Nat. Geosci. 10: 771776, doi: 10.1038/NGEO3011.Google Scholar
Okal, E. A., and Langenhorst, A. R. 2000. Seismic properties of the Eltanin transform system, South Pacific. Phys. Earth Planet. Int. 119(3): 185208.Google Scholar
Okal, E. A., and Stein, S. 1987. The 1942 southwest Indean Ocean ridge earthquake – Largest ever recorded on oneanic transform. Geophys. Res. Lett. 14(2): 147150, doi: 10.1029/GL014i002p00147.Google Scholar
Okal, E. A., and Stewart, L. M. 1982. Slow earthquakes along oceanic fracture-zones – Evidence for asthenospheric flow away from hotspots. Earth Planet. Sci. Lett. 57: 7587.Google Scholar
Okazaki, K., and Hirth, G. 2016. Dehydration of lawsonite could directly trigger earthquakes in subducting oceanic crust. Nature 530(7588): 81+, doi: 10.1038/nature16501.Google Scholar
Okubo, P. 1988. Rupture propagation in a nonuniform stress field following a state variable friction model. Eos 69: 400401.Google Scholar
Okubo, P., and Dieterich, J. H. 1984. Effects of physical fault properties on frictional instabilities produced on simulated faults. J. Geophys. Res. 89: 58155827.Google Scholar
Okubo, P., and Dieterich, J. H. 1986. State variable fault constitutive relations for dynamic slip. In Earthquake Source Mechanics. AGU Geophys. Mono., eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 2536.Google Scholar
Okui, Y. and Horii, H. 1997. Stress and time-dependent failure of brittle rocks under compression: A theoretical prediction. J. Geophys. Res-Sd EA 102 (B7): 1486914881.Google Scholar
Oleskevich, D. A., Hyndman, R. D., and Wang, K. 1999. The updip and downdip limits to great subduction earthquakes: Thermal and structural models of Cascadia, south Alaska, SW Japan, and Chile. J. Geophys. Res.-Solid Earth 104: 1496514991.Google Scholar
Olgaard, D. L., and Brace, W. F. 1983. The microstructure of gouge from a mining-induced seismic shear zone. Int. J. Rock Mech. Min. Sci. 20: 1119.Google Scholar
Olive, J. A., and Escartin, J. 2016. Dependence of seismic coupling on normal fault style along the Northern Mid-Atlantic Ridge. Geochem. Geophys. Geosystems 17(10): 41284152, doi: 10.1002/2016gc006460.Google Scholar
Olsen, K. B., Madariaga, R., and Archuleta, R. J. 1997. Three-dimensional dynamic simulation of the 1992 Landers earthquake. Science 278: 834838.Google Scholar
Olson, E. L., and Allen, R. M. 2005. The deterministic nature of earthquake rupture. Nature 438(7065): 212215, doi: 10.1038/nature04214.Google Scholar
Olson, J. E. 2003. Sublinear scaling of fracture aperture versus length: An exception or the rule? J. Geophys. Res.-Solid Earth 108(B9).Google Scholar
Olson, J. E., and Schultz, R. A. 2011. Comments on “A note on the scaling relations for opening mode fractures in rock” by C. H. Scholz. J. Struct. Geol. 33(10): 15231524.Google Scholar
Oohashi, K., Hirose, T., Kobayashi, K., and Shimamoto, T. 2012. The occurrence of graphite-bearing fault rocks in the Atotsugawa fault system, Japan: Origins and implications for fault creep. J. Struct. Geol. 38: 3950, doi: 10.1016/j.jsg.2011.10.011.Google Scholar
Oohashi, K., Hirose, T., Takahashi, M., and Tanikawa, W. 2015. Dynamic weakening of smectite-bearing faults at intermediate velocities: Implications for subduction zone earthquakes. J. Geophys. Res.-Solid Earth 120(3): 15721586, doi: 10.1002/2015jb011881.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
Oppenheimer, D. H. 1990. Aftershock slip behavior of the 1989 Loma-Prieta, California earthquake. Geophys. Res. Lett. 17: 11991202.Google Scholar
Oppenheimer, D. H., Reasenberg, P. A., and Simpson, R. W. 1988. Fault plane solutions for the 1984 Morgan Hill, California, earthquake sequence: Evidence for the state of stress on the Calaveras fault. J. Geophys. Res. 93: 90079026.Google Scholar
Ord, A., and Christie, J. M. 1984. Flow structures from microstructures of mylonitic quartzites of the Moine thrust zone, Assynt area, Scotland. J. Struct. Geol. 6: 639654.Google Scholar
Orowan, E. 1944. The fatigue of glass under stress, Nature 154: 341343.Google Scholar
Orowan, E. 1949. Fracture and strength of solids. Rep. Prog. Phys. 12: 4874.Google Scholar
Ota, Y., and Yamaguchi, M. 2004. Holocene coastal uplift in the western Pacific Rim in the context of late Quaternary uplift. Quaternary International 120: 105117.Google Scholar
Ota, Y., Machida, H., Hori, N., Kunishi, K., and Omur, A. 1978. Holocene raised coral reefs of Kinkaijima (Ryukyu Is.) – An approach to Holocene sea level study. Geograph. Rev. Japan. 51: 109130.Google Scholar
Ota, Y., Matsuda, T., and Naganuma, K. 1976. Tilted marine terraces of the Ogi Peninsula, Sado Island, related to the Ogi earthquake of 1802. Zisin II 29: 5570.Google Scholar
Owen, S., Anderson, G., Agnew, D. C., et al. 2002. Early postseismic deformation from the 16 October 1999 M-w 7.1 Hector Mine, California, earthquake as measured by survey-mode GPS. Bull. Seismol. Soc. Am. 92(4): 14231432, doi: 10.1785/0120000930.Google Scholar
Özalaybey, S., and Savage, M. K. 1995. Shear‐wave splitting beneath western United States in relation to plate tectonics. J. Geophys. Res.-Solid Earth 100(B9): 1813518149.Google Scholar
Ozawa, S., Miyazaki, S., Hatanaka, Y., Imakiire, T., Kaidzu, M., and Murakami, M. 2003. Characteristic silent earthquakes in the eastern part of the Boso peninsula, Central Japan. Geophys. Res. Lett. 30, doi: 10.1029/2002GL016665.Google Scholar
Ozawa, S., Nishimura, T., Munekane, H., Suito, H., Kobayashi, T., Tobita, M., and Imakiire, T. 2012. Preceding, coseismic, and postseismic slips of the 2011 Tohoku earthquake, Japan. J. Geophys. Res.-Solid Earth 117(B7): doi.org/10.1029/2011JB009120.Google Scholar
Pacheco, J. F., and Sykes, L. R. 1992. Seismic moment catalog of large shallow earthquakes, 1900 to 1989. Bull. Seismol. Soc. Amer. 82: 13061349.Google Scholar
Pacheco, J. F., Scholz, C. H., and Sykes, L. R. 1992. Changes in frequency-size relationship from small to large earthquakes. Nature 355: 7173.Google Scholar
Pacheco, J. F., Sykes, L. R., and Scholz, C. H. 1993. Nature of seismic coupling along simple plate boundaries of the subduction type. J. Geophys. Res. 98: 1413314159.Google Scholar
Page, B. M., Thompson, G. A., and Coleman, R. G. 1998. Late Cenozoic tectonics of the central and southern coast ranges of California. Geol. Soc. Am. Bull. 110: 846876.Google Scholar
Papazachos, B. 1975. Foreshocks and earthquake prediction. Tectonophysics 28: 213226.Google Scholar
Pareto, V. 1896. Cours d ‘Economic Politique. Reprinted as a volume of Oeuures Completes. 1965. Geneva: Droz.Google Scholar
Parks, G. 1984. Surface and interfacial free energies of quartz. J. Geophys. Res. 89: 39974008.Google Scholar
Parsons, T. 1998. Seismic-reflection evidence that the Hayward fault extends into the lower crust of the San Francisco Bay Area, California. Bull. Seismol. Soc. Amer. 88: 1212–23.Google Scholar
Parsons, T., and Dreger, D. S. 2000. Static-stress impact of the 1992 Landers earthquake sequence on nucleation and slip at the site of the 1999 M = 7.1 Hector Mine earthquake, southern California. Geophys. Res. Lett. 27: 19491952.Google Scholar
Parsons, T., Malagnini, L., and Akinci, A. 2017. Nucleation speed limit on remote fluid-induced earthquakes. Science Advances 3(8): e1700660, doi: 10.1126/sciadv.1700660.Google Scholar
Parsons, T., Segou, M., and Marzocchi, W. 2014. The global aftershock zone. Tectonophysics 618: 134, doi: 10.1016/j.tecto.2014.01.038.Google Scholar
Parsons, T., Toda, S., Stein, R. S., Barka, A., and Dieterich, J. H. 2000. Heightened odds of large earthquakes near Istanbul: An interaction-based probability calculation. Science 288: 661665.Google Scholar
Passchier, C. 1984. The generation of ductile and brittle deformation bands in a low-angle mylonite zone. J. Struct. Geol. 6: 273281.Google Scholar
Passelègue, F., Spagnuolo, E., Violay, M., Nielsen, S., Di Toro, G., and Schubnel, A. 2016. Frictional evolution, acoustic emissions activity, and off‐fault damage in simulated faults sheared at seismic slip rates. J. Geophys. Res.-Solid Earth 121 (10): 74907513.Google Scholar
Paterson, M. S. 1987. Problems in the extrapolation of laboratory rheological data. Tectonophys. 133: 3343.Google Scholar
Paterson, M. S. and Weaver, C. W. 1970. Deformation of polycrystalline MgO under pressure. J. Am. Ceram. Soc. 53: 463471.Google Scholar
Paterson, M. S., and Wong, T.-F. 2005. Experimental Rock Deformation: The Brittle Field, ed. Paterson, M. S.. New York: Springer.Google Scholar
Peacock, D. C. P. 1991. Displacements and segment linkage in strike-slip-fault zones. J. Struct. Geol. 13: 10251035.Google Scholar
Peacock, D. C. P., and Sanderson, D. J. 1991. Displacements, segment linkage and relay ramps in normal-fault zones. J. Struct. Geol. 13: 721734.Google Scholar
Peacock, D. C. P., and Sanderson, D. J. 1994. Geometry and development of relay ramps in normal-fault systems. AAPG Bull. Am. Assoc. Petr. Geol. 78: 147165.Google Scholar
Peacock, S. M. 2001. Are the lower planes of double seismic zones caused by serpentine dehydration in subducting oceanic mantle? Geology 29(4): 299302, doi: 10.1130/0091-7613(2001)029<0299:ATLPOD>2.0.CO;2.Google Scholar
Peacock, S. M. 2009. Thermal and metamorphic environments of subduction zone episodic tremor and slip. J. Geophys. Res. 114: B00A07, doi: 10.1029/2008JB005978.Google Scholar
Peck, L., Barton, C. C., and Gordon, R. B. 1985a. Microstructure and the Resistance of Rock to Tensile Fracture. J. Geophys. Res.-Solid Earth and Planets 90 (NB13):15331546.Google Scholar
Peck, L., Nolenhoeksema, R. C., Barton, C. C., and Gordon, R. B. 1985b. Measurement of the Resistance of Imperfectly Elastic Rock to the Propagation of Tensile Cracks. J. Geophys. Res.-Solid Earth and Planets 90(NB9):78277836.Google Scholar
Pegler, G., and Das, S. 1996. Analysis of the relationship between seismic moment and fault length for large crustal strike-slip earthquakes between 1977–92. Geophys. Res. Lett. 23: 905908.Google Scholar
Pelayo, A. M., and Wiens, D. A. 1992. Tsunami earthquakes – Slow thrust-faulting events in the accretionary wedge. J. Geophys. Res.-Solid Earth 97: 1532115337.Google Scholar
Peltzer, G., Rosen, P., Rogez, F., and Hudnut, K. 1996. Postseismic rebound in fault step-overs caused by pore fluid flow. Science 273: 12021204.Google Scholar
Peltzer, G., Rosen, P., Rogez, F., and Hudnut, K. 1998. Poroelastic rebound along the Landers 1992 earthquake surface rupture. J. Geophys. Res.-Solid Earth 103: 3013130145.Google Scholar
Peng, S. S. 1975. Note on Fracture Propagation and Time-Dependent Behavior of Rocks in Uniaxial Tension. Int. J. Rock Mech. Min. Sci. 12(4): 125127.Google Scholar
Peng, Z. G., and Zhao, P. 2009. Migration of early aftershocks following the 2004 Parkfield earthquake. Nature Geoscience 2(12): 877881, doi: 10.1038/ngeo697.Google Scholar
Peng, Z., and Gomberg, J. 2010. An integrated perspective of the continuum between earthquakes and slow-slip phenomena. Nature Geoscience 3(9): 599607, doi: 10.1038/ngeo940.Google Scholar
Peresan, A., Kossobokov, V. G., and Panza, G. F. 2012. Operational earthquake forecast/prediction. Rendiconti Lincei 23(2): 131138.Google Scholar
Perez, O. J., and Scholz, C. H. 1984. Heterogeneities of the instrumental seismicity catalog (1904–1980) for strong shallow earthquakes. Bull. Seismol. Soc. Amer. 74: 669686.Google Scholar
Perez, O. J., and Scholz, C. H. 1997. Long-term seismic behavior of the focal and adjacent regions of great earthquakes during the time between two successive shocks. J. Geophys. Res.-Solid Earth 102: 82038216.Google Scholar
Perez-Campos, X., McGuire, J. J., and Beroza, G. C. 2003. Resolution of the slow earthquake/high apparent stress paradox for oceanic transform fault earthquakes. J. Geophys. Res.-Solid Earth 108(B9): 18, doi: 10.1029/2002jb002312.Google Scholar
Perfettini, H., and Avouac, J. P. 2004. Postseismic relaxation driven by brittle creep: A possible mechanism to reconcile geodetic measurements and the decay rate of aftershocks, application to the Chi-Chi earthquake, Taiwan. J. Geophys. Res.-Solid Earth 109(B2): doi: 10.1029/2003jb002488.Google Scholar
Perfettini, H., and Avouac, J. P.. 2007. Modeling afterslip and aftershocks following the 1992 Landers earthquake, J. Geophys. Res.-Solid Earth, 112(B7), doi.org/10.1029/2006JB004399.Google Scholar
Perfettini, H., Avouac, J.-P., Tavera, H., et al. 2010. Seismic and aseismic slip on the Central Peru megathrust. Nature 465(7294): 7881, doi: 10.1038/nature09062.Google Scholar
Perouse, E., and Wernicke, B. P. 2017. Spatiotemporal evolution of fault slip rates in deforming continents: The case of the Great Basin region, northern Basin and Range province. Geosphere 13(1): 112135, doi: 10.1130/ges01295.1.Google Scholar
Perrin, C., Waldhauser, F., and Scholz, C. H. 2018. Characterizing the structural maturity of fault zones using high resolution earthquake locations, Geophys. Res. Lett. submitt.Google Scholar
Peterson, E. T., and Seno, T. 1984. Factors affecting seismic moment release rates in subduction zones. J. Geophys. Res. 89: 1023310248.Google Scholar
Philibosian, B., Sieh, K., Avouac, J.‐P., et al. 2017. Earthquake supercycles on the Mentawai segment of the Sunda megathrust in the seventeenth century and earlier. J. Geophys. Res.-Solid Earth 122(1): 642676, doi: 10.1002/2016jb013560.Google Scholar
Phillips, W. S., House, L. S., and Fehler, M. C. 1997. Detailed joint structure in a geothermal reservoir from studies of induced microearthquake clusters. J. Geophys. Res.-Solid Earth 102(B6): 1174511763.Google Scholar
Plafker, G., and Rubin, M. 1978. Uplift history and earthquake recurrence as deduced from marine terraces on Middleton Island, Alaska. In Proc. Conf. Vl: Methodology for Identifying Seismic Gaps and Soon-to-break Gaps, U.S. Geol Surv. Open-file Rept. 78–943, pp. 857868.Google Scholar
Platt, J. P., and Behr, W. M. 2011. Deep structure of lithospheric fault zones. Geophys. Res. Lett. 38, doi: 10.1029/2011gl049719.Google Scholar
Podolskiy, E. A., and Walter, F. 2016. Cryoseismology. Rev. Geophys. 54, doi: 10.1002/2016RG000526.Google Scholar
Poirier, J. P. 1980. Shear localization and shear instability in materials in the ductile field. J. Struct. Geol. 2(1–2): 135+, doi: 10.1016/0191–8141(80)90043–7.Google Scholar
Poirier, J.-P. 1985. Creep of Crystals. Cambridge: Cambridge University Press.Google Scholar
Polet, J., and Kanamori, H. 2000. Shallow subduction zone earthquakes and their tsunamigenic potential. Geophys. J. Int. 142: 684702.Google Scholar
Poliakov, A. N., Dmowska, R., and Rice, J. R. 2002. Dynamic shear rupture interactions with fault bends and off‐axis secondary faulting. J. Geophys. Res.-Solid Earth 107(B11): doi.org/10.1029/2001JB000572.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.. London: Academic Press, pp. 277348.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
Pollitz, F. F. 2003. Transient rheology of the uppermost mantle beneath the Mojave Desert, California. Earth Planet. Sci. Lett. 215(1–2): 89104, doi: 10.1016/s0012-821x(03)00432–1.Google Scholar
Pollitz, F. F. 2015. Postearthquake relaxation evidence for laterally variable viscoelastic structure and water content in the Southern California mantle. J. Geophys. Res.-Solid Earth 120(4): 26722696, doi: 10.1002/2014jb011603.Google Scholar
Pollitz, F. F., and Thatcher, W. 2010. On the resolution of shallow mantle viscosity structure using postearthquake relaxation data: Application to the 1999 Hector Mine, California, earthquake. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2010jb007405.Google Scholar
Pollitz, F. F., Bürgmann, R., and Banerjee, P. 2006. Post-seismic relaxation following the great 2004 Sumatra-Andaman earthquake on a compressible self-gravitating Earth. Geophys. J. Int. 167(1): 397420.Google Scholar
Pollitz, F. F., Burgmann, R., and Segall, P. 1998. Joint estimation of afterslip rate and postseismic relaxation following the 1989 Loma Prieta earthquake. J. Geophys. Res. 103: 2697526992.Google Scholar
Pollitz, F. F., Peltzer, G., and Burgmann, R. 2000. Mobility of continental mantle: Evidence from postseismic geodetic observations following the 1992 Landers earthquake. J. Geophys. Res. 105: 80358054.Google Scholar
Pollitz, F. F., Stein, R. S., Sevilgen, V., and Burgmann, R. 2012. The 11 April 2012 east Indian Ocean earthquake triggered large aftershocks worldwide. Nature 490(7419): 250+, doi: 10.1038/nature11504.Google Scholar
Pomeroy, P. W., Simpson, D. W., and Sbar, M. L. 1976. Earthquakes triggered by surface quarrying – Wappinger Falls, New York sequence of June, 1974. Bull. Seismol. Soc. Am. 66: 685700.Google Scholar
Popek, M. A., and Saffer, D. M. 2011. Heat advection by groundwater flow through a heterogeneous permeability crust: A potential cause of scatter in surface heat flow near Parkfield, California. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb008081.Google Scholar
Poulimenos, G. 2000. Scaling properties of normal fault populations in the western Corinth Graben, Greece: Implications for fault growth in large strain settings. J. Struct. Geol. 22(3): 307322, doi: 10.1016/s0191-8141(99)00152–2.Google Scholar
Power, W. L., and Tullis, T. E. 1989. The relationship between slickenside surfaces in fine grained quartz and the seismic cycle. J. Struct. Geol. 11(7): 879893, doi: 10.1016/0191–8141(89)90105–3.Google Scholar
Power, W. L., Tullis, T. E., and Weeks, J. D. 1988. Roughness and wear during brittle faulting. J. Geophys. Res. 93: 1526815278.Google Scholar
Power, W. L., Tullis, T. E., Brown, S., Boitnott, G. N., and Scholz, C. H. 1987. Roughness of natural fault surfaces. Geophys. Res. Lett. 14: 2932.Google Scholar
Powers, P. M., and Jordan, T. H. 2010. Distribution of seismicity across strike-slip faults in California. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2008jb006234.Google Scholar
Pratt, H. R., Black, A. D., Brown, W. S., and Brace, W. F. 1972. The effect of specimen size on the strength of unjointed diorite. Int. J. Rock Mech. Min. Sci. 9: 513529.Google Scholar
Pratt, M. J., Winberry, J. P., Wiens, D. A., Anandakrishnan, S., and Alley, R. B. 2014. Seismic and geodetic evidence for grounding-line control of Whillans Ice Stream stick-slip events. J. Geophys. Res. Earth Surf. 119: 333348, doi: 10.1002/2013JF002842.Google Scholar
Prejean, S. G., Hill, D. P., Brodsky, E. E., et al. 2004. Remotely triggered seismicity on the United States west coast following the M-W 7.9 Denali fault earthquake. Bull. Seismol. Soc. Am. 94(6): S348S359, doi: 10.1785/0120040610.Google Scholar
Price, N. J. 1966. Fault and Joint Development in Brittle and Semi-brittle Rock. Oxford: Pergamon.Google Scholar
Price, R. A. 1988. The mechanical paradox of large overthrusts. Geol. Soc. Amer. Bull. 100(12): 18981908.Google Scholar
Procter, B. A., Whitney, I., and Johnson, J. W. 1967. The strength of fused silica. Proc. Roy. Soc. London Ser. A 297: 534547.Google Scholar
Proctor, B., and Lockner, D. A. 2016. Pseudotachylyte increases the post-slip strength of faults. Geology 44 (12): 10031006, doi: 10.1130/g38349.1.Google Scholar
Proffett, J. M. 1977. Cenozoic geology of the Yerington District, Nevada, and implications for the nature and origin of Basin and Range faulting. Geol. Soc. Amer. Bull. 88: 247266.Google Scholar
Prothero, W. A., Jr., and Reid, I. D. 1982. Microearthquakes on the East Pacific Rise at 21° N and the Rivera fracture zone. J. Geophys. Res. 87: 85098518.Google Scholar
Provost, A. S., and Houston, H. 2001. Orientation of the stress field surrounding the creeping section of the San Andreas Fault: Evidence for a narrow mechanically weak fault zone. J. Geophys. Res.-Solid Earth 106(B6): 1137311386, doi: 10.1029/2001jb900007.Google Scholar
Provost, A. S., and Houston, H. 2003. Stress orientations in northern and central California: Evidence for the evolution of frictional strength along the San Andreas plate boundary system. J. Geophys. Res.-Solid Earth 108(B3), doi.org/10.1029/2001JB001123.Google Scholar
Purcaru, G., and Berckhemer, H. 1978. A magnitude scale for very large earthquakes. Tectonophysics 49: 189198.Google Scholar
Putelat, T., and Dawes, J. H. P. 2015. Steady and transient sliding under rate-and-state friction. J. Mech. Phys. Sol. 78: 7093, doi: 10.1016/j.jmps.2015.01.016.Google Scholar
Putelat, T., Dawes, J. H. P., and Willis, J. R. 2011. On the microphysical foundations of rate-and-state friction. J. Mech. Phys. Sol. 59(5): 10621075, doi: 10.1016/j.jmps.2011.02.002.Google Scholar
Rabinowicz, E. 1951. The nature of the static and kinetic coefficients of friction. J. Appl. Phys. 22: 13731379.Google Scholar
Rabinowicz, E. 1956. Autocorrelation analysis of the sliding process. J. Appl. Phys. 27: 131135.Google Scholar
Rabinowicz, E. 1958. The intrinsic variables affecting the stick-slip process. Proc. Phys. Soc. London) 71: 668675.Google Scholar
Rabinowicz, E. 1965. Friction and Wear of Materials. New York: John Wiley.Google Scholar
Rabinowitz, H. S. 2017. The seismogenic potential of subducting sediments, 185 pp, PhD thesis, Columbia University.Google Scholar
Rabotnov, Y. N. 1969. Creep Problems in Structural Members. Amsterdam: North Holland.Google Scholar
Radiguet, M., Cotton, F., Vergnolle, M., et al. 2012. Slow slip events and strain accumulation in the Guerrero gap, Mexico. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb008801.Google Scholar
Radiguet, M., Perfettini, H., Cotte, N., et al. 2016. Triggering of the 2014 Mw7.3 Papanoa earthquake by a slow slip event in Guerrero, Mexico. Nature Geoscience 9: 829833.Google Scholar
Raeesi, M., and Atakan, K. 2009. On the deformation cycle of a strongly coupled plate interface: The triple earthquakes of 16 March 1963, 15 November 2006, and 13 January 2007 along the Kurile subduction zone. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb006184.Google Scholar
Rajendran, K., Rajendran, C., Earnest, A., Prasad, G. R., Dutta, K., Ray, D., and Anu, R. 2008. Age estimates of coastal terraces in the Andaman and Nicobar Islands and their tectonic implications. Tectonophysics 455(1): 5360.Google Scholar
Raleigh, C. B., and Patterson, M. S. 1965. Experimental deformation of serpentinite and its tectonic implications. J. Geophys. Res. 70: 39653985.Google Scholar
Raleigh, C. B., et al. 1977. Prediction of the Haicheng earthquake. Eos 58: 236272.Google Scholar
Raleigh, C. B., Healy, J. H., and Bredehoeft, J. 1972. Faulting and crustal stress at Rangely, Colorado. In Flow and Fracture of Rocks, AGU Geophysical Mono. 16, eds. Heard, H., Borg, I., Carter, N., and Raleigh, C.. Washington, DC: American Geophysical Union, pp. 277284.Google Scholar
Raleigh, C. B., Healy, J., and Bredehoeft, J. 1976. An experiment in earthquake control at Rangely, Colorado. Science 191: 12301237.Google Scholar
Ramsay, J. G. 1983. The crack–seal mechanism of rock deformation. Nature 284: 135139.Google Scholar
Ramsay, J. G., and Graham, R. H. 1970. Strain variation in shear belts. Can. J. Earth Sci. 7: 786813.Google Scholar
Ranero, C. R., Morgan, J. P., McIntosh, K., and Reichert, C. 2003. Bending-related faulting and mantle serpentinization at the Middle America trench. Nature 425(6956): 367.Google Scholar
Rayleigh, L. 1877. The Theory of Sound. New York: Dover.Google Scholar
Reasenberg, P. A. 1999. Foreshock occurrence before large earthquakes. J. Geophys. Res.-Solid Earth 104(B3): 47554768.Google Scholar
Reasenberg, P. A., and Jones, L. M. 1989. Earthquake hazzard after a mainshock in California. Science 243(4895): 11731176, doi: 10.1126/science.243.4895.1173.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., and Dewers, T. A. 2005. Gouge formation by dynamic pulverization during earthquake rupture. Earth Planet. Sci. Lett. 235(1–2): 361374, doi: 10.1016/j.epsl.2005.04.009.Google Scholar
Reches, Z., and Lockner, D. A. 2010. Fault weakening and earthquake instability by powder lubrication. Nature 467(7314): 452-455, doi: 10.1038/nature09348.Google Scholar
Reid, H. F. 1910. The mechanism of the earthquake. In The California Earthquake of April 18, 1906, Report of the State Earthquake Investigation Commission, Vol. 2. Washington, DC: Carnegie Institution, pp. 1192.Google Scholar
Reilinger, R., McClusky, S., Vernant, P., et al. 2006. GPS constraints on continental deformation in the Africa‐Arabia‐Eurasia continental collision zone and implications for the dynamics of plate interactions. J. Geophys. Res.-Solid Earth 111(B5): doi.org/10.1029/2005JB004051.Google Scholar
Reinecke, T. 1998. Prograde high- to ultrahigh-pressure metamorphism and exhumation of oceanic sediments at Lago di Cignana, Zermatt-Saas Zone, western Alps. Lithos 42: 147189.Google Scholar
Reinen, L. A. 2000. Slip styles in a spring-slider model with a laboratory-derived constitutive law for serpentinite. Geophys. Res. Lett. 27: 20372040.Google Scholar
Reinen, L. A., Tullis, T. E., and Weeks, J. D. 1992. 2-mechanism model for frictional sliding of serpentinite. Geophys. Res. Lett. 19(15): 15351538, doi: 10.1029/92gl01388.Google Scholar
Reinen, L. A., Weeks, J. D., and Tullis, T. E. 1991. The frictional behavior of serpentinite – implications for aseismic creep on shallow crustal faults. Geophys. Res. Lett. 18: 19211924.Google Scholar
Reinen, L. A., Weeks, J. D., and Tullis, T. E. 1994. The frictional behavior of lizardite and antigorite serpentinites – Experiments, constitutive models, and implications for natural faults. Pure Appl. Geophys. 143: 317358.Google Scholar
Remy, D., Perfettini, H., Cotte, N., et al. 2016. Postseismic relocking of the subduction megathrust following the 2007 Pisco, Peru, earthquake. J. Geophys. Res.-Solid Earth 121(5): 39783995, doi: 10.1002/2015jb012417.Google Scholar
Renard, F., and Ortoleva, P. 1997. Water films at grain-grain contacts: Debye-Hückel, osmotic model of stress, salinity, and mineralogy dependence. Geochimica et Cosmochimica Acta 61 (10): 19631970.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
Research Group for Active Faults in Japan. 1980. Active Faults in Japan, Sheet Maps and Inventories. Tokyo: Tokyo University Press.Google Scholar
Reynolds, S. J., and Lister, G. S. 1987. Structural aspects of fluid-rock interactions in detachment zones. Geology 15: 362366.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. 1980. The mechanics of earthquake rupture. In Physics of the Earth’s Interior, eds. Dziewonski, A. and Boschi, E.. Amsterdam: Elsevier Science, pp. 555649.Google Scholar
Rice, J. R. 1983. Constitutive relations for fault slip and earthquake instabilities. Pageoph 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.. London: Academic Press, pp. 475504.Google Scholar
Rice, J. R. 2006. Heating and weakening of faults during earthquake slip. J. Geophys. Res. 111(B5), doi.org/10.1029/2005JB004006.Google Scholar
Rice, J. R., and Cleary, M. P. 1976. Some basic stress diffusion solutions for fluid-saturated elastic porous media with compressible constituents. Rev. Geophys. Space Phys. 14: 227241.Google Scholar
Rice, J. R., and Rudnicki, J. W. 1979. Earthquake precursory effects due to pore fluid stabilization of a weakened fault zone. J. Geophys. Res. 84: 21772184.Google Scholar
Rice, J. R., and Tse, S. T. 1986. Dynamic motion of a single degree of freedom system following a rate and state dependent friction law. J. Geophys. Res. 91: 521530.Google Scholar
Rice, J. R., Lapusta, N., and Ranjith, K. 2001. Rate and state dependent friction and the stability of sliding between elastically deformable solids. J. Mech. Phys. Sol. 49: 18651898.Google Scholar
Rice, J. R., Sammis, C. G., and Parsons, R. 2005. Off-fault secondary failure induced by a dynamic slip pulse. Bull. Seismol. Soc. Am. 95(1): 109134, doi: 10.1785/0120030166.Google Scholar
Richards, P. 1976. Dynamic motions near an earthquake fault: A three-dimensional solution. Bull. Seismol. Soc. Am. 66: 132.Google Scholar
Richards-Dinger, K., Stein, R. S., and Toda, S. 2010. Decay of aftershock density with distance does not indicate triggering by dynamic stress. Nature 467(7315): 583-586, doi: 10.1038/nature09402.Google Scholar
Richins, W., Pechmann, J., Smith, R., Langer, C., Goter, S., Zollweg, J., and King, J. 1987. The 1983 Borah Peak, Idaho, earthquake and its aftershocks. Bull. Seismol. Soc. Am. 77: 694723.Google Scholar
Richter, C. F. 1958. Elementary Seismology. San Francisco: W. H. Freeman.Google Scholar
Richter, F. M., and McKenzie, D. P. 1978. Simple plate models of mantle convection. J. Geophys. 44: 441471.Google Scholar
Riedel, W. 1929. Zur Mechanik geologischer Brucherscheinungen. Centralbl. f. Mineral. Geol. u. Pal. 1929B: 354368.Google Scholar
Rikitake, T. 1976. Earthquake Prediction. Amsterdam: Elsevier.Google Scholar
Rikitake, T. 1982. Earthquake Forecasting and Warning. Tokyo: D. Reidel.Google Scholar
Rippon, J. H. 1985. Contoured pattems of the throw and hade of normal faults in the Coal Measures (Westphalian) of northwest Derbyshire. Proc. Yorkshire Geol. Soc. 45: 147161.Google Scholar
Ritz, J.-F., and Taboada, A. 1993. Revolution stress ellipsoids in brittle tectonics resulting from uncritical use of inverse methods. Bull. Soc. Geol. France 164: 519531.Google Scholar
Rives, T., Razack, M., Petit, J. P., and Rawnsley, K. D. 1992. Joint spacing – Analog and numerical simulations. J. Struct. Geol. 14: 925937.Google Scholar
Rivet, D., Campillo, M., Shapiro, N. M., et al. 2011. Seismic evidence of nonlinear crustal deformation during a large slow slip event in Mexico. Geophys. Res. Lett. 38(8).Google Scholar
Roberts, G. P., and Michetti, A. M. 2004. Spatial and temporal variations in growth rates along active normal fault systems: An example from The Lazio-Abruzzo Apennines, central Italy. J. Struct. Geol. 26(2): 339376, doi: 10.1016/s0191-8141(03)00103–2.Google Scholar
Roberts, G. P., Cowie, P., Papanikolaou, 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(2): 377398, doi: 10.1016/s0191-8141(03)00104–4.Google Scholar
Robertson, E. C. 1982. Continuous formation of gouge and breccia during fault displacement. In Issues in Rock Mechanics, Proc. Symp. Rock Mech. 23rd, eds. Goodman, R. E. and Hulse, F.. New York: Am. Inst. Min. Eng., pp. 397404.Google Scholar
Robertson, E. C. 1987. Fault breccia, displacement, and rock type, paper presented at the 28th US Symposium on Rock Mechanics (USRMS), American Rock Mechanics Association.Google Scholar
Robin, P. Y. 1973. Note on effective stress. J. Geophys. Res. 78: 24342437.Google Scholar
Robin, P. Y. 1978. Pressure solution-to-grain contacts. Geochim. Cosmochim. Acta 42: 13831389.Google Scholar
Robinson, D. P. 2011. A rare great earthquake on an oceanic fossil fracture zone. Geophys. J. Int. 186(3): 11211134, doi: 10.1111/j.1365-246X.2011.05092.x.Google Scholar
Rodgers, D. W., and Little, T. A. 2006. World’s largest coseismic strike-slip offset: The 1855 rupture of the Wairarapa Fault, New Zealand, and implications for displacement/length scaling of continental earthquakes. J. Geophys. Res.-Solid Earth 111 (B12), doi: 10.1029/2005jb004065.Google Scholar
Roeloffs, E. 1988a. Hydrological precursors to earthquakes. Pageoph 126: 177209.Google Scholar
Roeloffs, E. A. 1988b. Fault stability changes induced beneath a reservoir with cyclic variations in water level. J. Geophys. Res. 93: 21072124.Google Scholar
Roeloffs, E. A. 2006. Evidence for aseismic deformation rate changes prior to earthquakes. Ann. Rev. Earth Planet. Sci. 34: 591627, doi: 10.1146/annurev.earth.34.031405.124947.Google Scholar
Roeloffs, E. and Langbein, J. 1994. The earthquake prediction experiment at Parkfield, California. Rev. Geophys. 32: 315336.Google Scholar
Roeloffs, E., and Quilty, E. 1997. Case 21: Water level and strain changes preceding and following the August 4, 1985 Kettleman Hills, California, earthquake. Pure Appl. Geophys. 149: 2160.Google Scholar
Rogers, G., and Dragert, H. 2003. Episodic tremor and slip on the Cascadia subduction zone: The chatter of silent slip. Science 300(5627): 19421943.Google Scholar
Rojstaczer, S. A., and Wolf, S. 1992. Permeability changes associated with large earthquakes: An example from Loma Prieta, California. Geology 20: 211214.Google Scholar
Rojstaczer, S. A., Wolf, S., and Michel, R. 1995. Permeability enhancement in the shallow crust as a cause of earthquake-induced hydrological changes. Nature 373: 237239.Google Scholar
Roland, E., and McGuire, J. J. 2009. Earthquake swarms on transform faults. Geophys. J. Int. 178(3): 16771690, doi: 10.1111/j.1365-246X.2009.04214.x.Google Scholar
Roland, E., Lizarralde, D., McGuire, J. J., and Collins, J. A. 2012. Seismic velocity constraints on the material properties that control earthquake behavior at the Quebrada-Discovery-Gofar transform faults, East Pacific Rise. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2012jb009422.Google Scholar
Romanowicz, B. 1992. Strike-slip earthquakes on quasi-vertical transcurrent faults: Inferences for general scaling relations. Geophys. Res. Lett. 19: 481484.Google Scholar
Rosakis, A. J., Samudrala, O., and Coker, D. 1999. Cracks faster than the shear wave speed. Science 284: 13371340.Google Scholar
Ross, Z. E., Rollins, C., Cochran, E. S., Hauksson, E., Avouac, J. P., and Ben-Zion, Y. 2017. Aftershocks driven by afterslip and fluid pressure sweeping through a fault-fracture mesh. Geophys. Res. Lett. 44(16): 82608267, doi: 10.1002/2017gl074634.Google Scholar
Rowe, C. D., and Griffith, W. A. 2015. Do faults preserve a record of seismic slip: A second opinion. J. Struct. Geol. 78: 126, doi: 10.1016/j.jsg.2015.06.006.Google Scholar
Royer, J.-Y., and Gordon, R. G. 1997. The motion and boundary between the Capricorn and Australian plates. Science 277(5330): 12681274.Google Scholar
Rubin, A. M. 2008. Episodic slow slip events and rate-and-state friction. J. Geophys. Res.-Solid Earth 113(B11): doi.org/10.1029/2008JB005642.Google Scholar
Rubin, A., and Ampuero, J. P. 2005. Earthquake nucleation on (aging) rate and state faults. J. Geophys. Res.-Solid Earth 110(B11), doi.org/10.1029/2005JB003686.Google Scholar
Rubin, C. M., Horton, B. P., Sieh, K., Pilarczyk, J. E., Daly, P., Ismail, N., and Parnell, A. C. 2017. Highly variable recurrence of tsunamis in the 7,400 years before the 2004 Indian Ocean tsunami. Nature Communications 8, doi: 10.1038/ncomms16019.Google Scholar
Rubinstein, J. L., Ellsworth, W. L., Chen, K. H., and Uchida, N. 2012. Fixed recurrence and slip models better predict earthquake behavior than the time- and slip-predictable models: 1. Repeating earthquakes. J. Geophys. Res., 117, doi: 10.1029/2011JB008724.Google Scholar
Rubinstein, J. L., LaRocca, M., Vidale, J. E., Creager, K. C., and Wech, A. G. 2008. Tidal modulation of non-volcanic tremor. Science 319: 186189.Google Scholar
Rubinstein, J. L., Vidale, J. E., Gomberg, J., Bodin, P., Creager, K. C., and Malone, S. D. 2007. Non-volcanic tremor driven by large transient shear stresses. Nature 448: 579582.Google Scholar
Rubinstein, S. M., Cohen, G., and Fineberg, J. 2004. Detachment fronts and the onset of dynamic friction. Nature, 430(7003): 10051009, doi: 10.1038/nature02830.Google Scholar
Rudnicki, J. W. 1984. Effects of dilatancy hardening on the develoment of localized shear in fissured rock masses. J.Geophys. Res. 89(NB11): 92599270, doi: 10.1029/JB089iB11p09259.Google Scholar
Rudnicki, J. W. 1988. Physical models of earthquake instability and precursory processes. Pure Appl. Geophys. 126: 531554.Google Scholar
Rudnicki, J. W., and Rice, J. R. 1975. Conditions for localization of deformation in pressure sensitive dilatant materials. J. Mech. Phys. Sol. 23(6): 371394, doi: 10.1016/0022–5096(75)90001–0.Google Scholar
Ruff, L. D. 1989. Do trench sediments affect great earthquake occurrence in subduction zones? Pageoph 129: 263282.Google Scholar
Ruff, L., and Kanamori, H. 1980. Seismicity and the subduction process. Phys. Earth Planet. Int. 23: 240252.Google Scholar
Ruff, L., and Kanamori, H. 1983. Seismic coupling and uncoupling at subduction zones. Tectonophysics 99: 99117.Google Scholar
Ruina, A. L. 1983. Slip instability and state variable friction laws. J. Geophys. Res. 88: 10,35910,370.Google Scholar
Ruiz, J. A., Baumont, D., Bernard, P., and Berge-Thierry, C. 2013. Combining a kinematic fractal source model with hybrid green’s functions to model broadband strong ground motion. Bull. Seismol. Soc. Amer. 103: 31153130, doi: 10.1785/0120110135.Google Scholar
Ruiz, S., Metois, M., Fuenzalida, A., et al. 2014. Intense foreshocks and a slow slip event preceded the 2014 Iquique M-w 8.1 earthquake. Science 345(6201): 11651169, doi: 10.1126/science.1256074.Google Scholar
Rundle, J. B. 1989. Derivation of the complete Gutenberg–Richter magnitude – Frequency relation using the principle of scale invariance. J. Geophys. Res. 94: 12,33712,342.Google Scholar
Rundquist, D., and Sobolev, P. 2002. Seismicity of mid-oceanic ridges and its geodynamic implications: A review. Earth-Science Reviews 58(1): 143161.Google Scholar
Rutter, E. H. 1986. On the nomenclature of mode of failure transitions in rocks. Tectonophysics 122: 381387.Google Scholar
Rutter, E. H. 1999. On the relationship between the formation of shear zones and the form of the flow law for rocks undergoing dynamic recrystallization. Tectonophysics 303(1–4): 147158, doi: 10.1016/s0040-1951(98)00261–3.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. Pageoph 124: 330.Google Scholar
Ryall, A., and Malone, S. D. 1971. Earthquake distribution and mechanism of faulting in the Rainbow Mountain–Dixie Valley–Fairview Peak area, central Nevada. J. Geophys. Res. 76: 72417248.Google Scholar
Rydelek, P. A., and Sacks, I. S. 1989. Testing the completeness of earthquake catalogs and the hypothesis of self-similarity. Nature 337: 251253.Google Scholar
Rymer, M. J., TreimanJ. A.Kendrick, K. J., et al. 2011. Triggered Surface Slips in Southern California Associated with the 2010 El Mayor-Cucapah, Baja California, Mexico, Earthquake. U.S. Geol. Surv. Open-file report 2010–1333, available at http://pubs.usgs.gov./of/2010/1333/.Google Scholar
Sacks, I. S., Suyehiro, S., Linde, A. T., and Snoke, J. A. 1978. Slow earthquakes and stress redistribution. Nature 275: 599602.Google Scholar
Saffer, D. M., and Marone, C. 2003. Comparison of smectite- and illite-rich gouge frictional properties: Application to the updip limit of the seismogenic zone along subduction megathrusts. Earth Planet. Sci. Lett. 215(1–2): 219235, doi: 10.1016/s0012-821x(03)00424–2.Google Scholar
Saffer, D. M., and Wallace, L. M. 2015. The frictional, hydrologic, metamorphic and thermal habitat of shallow slow earthquakes. Nature Geoscience, 8(8): 594600, doi: 10.1038/ngeo2490.Google Scholar
Sagiya, T. 1999. Interplate coupling in the Tokai District, Central Japan, deduced from continuous GPS measurements. Geophys. Res. Lett. 26: 23152318.Google Scholar
Sagiya, T. 2004. Interplate coupling in the Kanto District, Central Japan, and the Boso Peninsula Silent earthquake in May 1996. Pageoph 161: 23272342, doi: 10.1007/s00024-004–2566-6.Google Scholar
Sagiya, T., and Thatcher, W. 1999. Coseismic slip resolution along a plate boundary megathrust: The Nankai Trough, southwest Japan. J. Geophys. Res.-Solid Earth 104(B1): 11111129.Google Scholar
Sagy, A., and Brodsky, E. E. 2009. Geometric and rheological asperities in an exposed fault zone. J. Geophys. Res.-Solid Earth 114, doi: 10.1029/2008jb005701.Google Scholar
Sagy, A., Brodsky, E. E., and Axen, G. J. 2007. Evolution of fault-surface roughness with slip. Geology 35(3): 283286, doi: 10.1130/g23235a.1.Google Scholar
Saillard, M., Audin, L., Rousset, B., Avouac, J. P., Chlieh, M., Hall, S. R., Husson, L., and Farber, D. L. 2017. From the seismic cycle to long-term deformation: Linking seismic coupling and Quaternary coastal geomorphology along the Andean megathrust. Tectonics 36(2): 241256, doi: 10.1002/2016tc004156.Google Scholar
Saleur, H., Sammis, C. G., and Sornette, D. 1996. Discrete scale invariance, complex fractal dimensions, and log-periodic fluctuations in seismicity. J. Geophys. Res.-Solid Earth 101: 1766117677.Google Scholar
Sammis, C. G., and King, G. C. P. 2007. Mechanical origin of power law scaling in fault zone rock. Geophys. Res. Lett. 34(4): doi: 10.1029/2006gl028548.Google Scholar
Sammis, C. G. Osborne, R., Anderson, J., Banerdt, M., and White, P. 1986. Self-similar cataclasis in the formation of fault gauge. Pure Appl. Geophys. 124: 5378.Google Scholar
Samudrala, O., Huang, Y., and Rosakis, A. J. 2002. Subsonic and intersonic shear rupture of weak planes with a velocity weakening cohesive zone. J. Geophys. Res.-Solid Earth 107(B8): ESE 7-1–ESE 7-32, doi: 10.1029/2001jb000460.Google Scholar
Sanford, A. R. 1959. Analytical and experimental study of simple geologic structures. Bull. Geol. Soc. Am. 70: 1951.Google Scholar
Satake, K. 1994. Mechanism of the 1992 Nicaragua tsunami earthquake. Geophys. Res. Lett. 21(23): 25192522.Google Scholar
Satake, K. 2015. Geological and historical evidence of irregular recurrent earthquakes in Japan. Philos. Trans. A Math. Phys. Eng. Sci. 373(2053): doi: 10.1098/rsta.2014.0375.Google Scholar
Satake, K., and Tanioka, Y. 1999. Sources of tsunami and tsunamigenic earthquakes in subduction zones. Pure Appl. Geophys. 154: 467483.Google Scholar
Satake, K., Wang, K., and Atwater, B. F. 2003. Fault slip and seismic moment of the 1700 Cascadia earthquake inferred from the Japanese tsunami descriptions. J. Geophys. Res. 108 (B11), doi: 190.1029/2003JB002521.Google Scholar
Sato, H. 1988. Temporal change in scattering and attenuation associated with the earthquake occurrence – A review of recent studies on coda waves. Pageoph 126: 465498.Google Scholar
Sato, T., Higuchi, H., Miyauchi, T., et al. 2016. The source model and recurrence interval of Genroku-type Kanto earthquakes estimated from paleo-shoreline data. Earth Planets and Space 68: 17, doi: 10.1186/s40623-016–0395-3.Google Scholar
Saucier, F., Humphreys, E., and Weldon, R. 1992. Stress near geometrically complex strike-slip faults: Application to the San Andreas fault at Cajon Pass, southern California. J. Geophys. Res. 97: 50815094.Google Scholar
Savage, H. M., and Brodsky, E. E. 2011. Collateral damage: Evolution with displacement of fracture distribution and secondary fault strands in fault damage zones. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb007665.Google Scholar
Savage, H. M., and Marone, C. 2008. Potential for earthquake triggering from transient deformations. J. Geophys. Res.-Solid Earth 113(B5): doi: 10.1029/2007jb005277.Google Scholar
Savage, H. M., Polissar, P. J., Sheppard, R., Rowe, C. D., and Brodsky, E. E. 2014. Biomarkers heat up during earthquakes: New evidence of seismic slip in the rock record. Geology 42(2): 99102, doi: 10.1130/g34901.1.Google Scholar
Savage, J. C. 1969. Mechanics of deep-focus faulting. Tectonophysics 8(2): 115+, doi: 10.1016/0040–1951(69)90085–7.Google Scholar
Savage, J. C. 1983. A dislocation model of strain accumulation and release at a subduction zone. J. Geophys. Res. 88: 49844996.Google Scholar
Savage, J. C. 1993. The Parkfield prediction fallacy. Bull. Seismol. Soc. Amer. 83: 16.Google Scholar
Savage, J. C. 1994. Empirical earthquake probabilities from observed recurrence intervals. Bull. Seismol. Soc. Amer. 84: 219221.Google Scholar
Savage, J. C. 1995. Interseismic uplift at the Nankai subduction zone, southwest Japan, 1951–1990. J. Geophys. Res.-Solid Earth 100: 63396350.Google Scholar
Savage, J. C. 2007. Postseismic relaxation associated with transient creep rheology. J. Geophys. Res.-Solid Earth 112(B5): doi.org/10.1029/2006JB004688.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., and Hastie, L. M. 1966. Surface deformation associated with dip-slip faulting. J. Geophys. Res. 71: 48974904.Google Scholar
Savage, J. C., and Lisowski, M. 1984. Deformation in the White Mountain seismic gap, California–Nevada, 1972–1982. J. Geophys. Res. 89: 76717688.Google Scholar
Savage, J. C., and Lisowski, M. 1993. Inferred depth of creep on the Hayward fault, central California. J. Geophys. Res.-Solid Earth 98: 787793.Google Scholar
Savage, J. C., and Lisowski, M. 1995. Interseismic deformation along the San-Andreas fault in southern California. J. Geophys. Res.-Solid Earth 100: 1270312717.Google Scholar
Savage, J. C., and Prescott, W. H. 1978. Asthenosphere readjustment and the earthquake cycle. J. Geophys. Res. 83: 33693376.Google Scholar
Savage, J. C., and Simpson, R. W. 1997. Surface strain accumulation and the seismic moment tensor. Bull. Seismol. Soc. Amer. 87: 13451353.Google Scholar
Savage, J. C., and Svarc, J. L. 1997. Postseismic deformation associated with the 1992 M-w = 7.3 Landers earthquake, southern California. J. Geophys. Res.-Solid Earth 102: 75657577.Google Scholar
Savage, J. C., and Thatcher, W. 1992. Interseismic deformation at the Nankai trough, Japan, subduction zone. J. Geophys. Res. 97: 1111711135.Google Scholar
Savage, J. C., and Wood, M. 1971. The relation between apparent stress and stress drop. Bull. Seismol. Soc. Am. 61: 13811388.Google Scholar
Savage, J. C., Lisowski, M., and Prescott, W. H. 1990. An apparent shear zone trending north-northwest across the Mojave-Desert into Owens-Valley, eastern California. Geophys. Res. Lett. 17: 21132116.Google Scholar
Savage, J. C., Lisowski, M., Svarc, L., and Gross, W. K. 1995. Strain accumulation across the central Nevada seismic zone. J. Geophys. Res. 100: 20,25720,269.Google Scholar
Savage, J. C., Simpson, R. W., and Murray, M. H. 1998. Strain accumulation rates in the San Francisco Bay area, 1972–1989. J. Geophys. Res.-Solid Earth 103: 1803918051.Google Scholar
Savage, J. C., Svarc, J. L., and Prescott, W. H. 1999. Geodetic estimates of fault slip rates in the San Francisco Bay area. J. Geophys. Res.-Solid Earth 104: 49955002.Google Scholar
Savage, J. C., Svarc, J. L., and Yu, S. B. 2005. Postseismic relaxation and transient creep. J. Geophys. Res.-Solid Earth 110(B11): doi: 10.1029/2005jb003687.Google Scholar
Savage, J., and Walsh, J. 1978. Gravitational energy and faulting. Bull. Seismol. Soc. Am. 68(6): 16131622.Google Scholar
Sawai, Y., Kamataki, T., Shishikura, M., Nasu, H., Okamura, Y., Satake, K., Thomson, K. H., Matsumoto, D., Fujii, Y., and Komatsubara, J. 2009. Aperiodic recurrence of geologically recorded tsunamis during the past 5500 years in eastern Hokkaido, Japan. J. Geophys. Res.-Solid Earth 114(B1).Google Scholar
Sbar, M. L., and Sykes, L. R. 1977. Seismicity and lithospheric stress in New York and adjacent areas. J. Geophys. Res. 82: 57715786.Google Scholar
Scharer, K. M., Biasi, G. P., Weldon, R. J., and Fumal, T. E. 2010. Quasi-periodic recurrence of large earthquakes on the southern San Andreas fault. Geology 38(6): 555558, doi: 10.1130/g30746.1.Google Scholar
Schlindwein, V., and Schmid, F. 2016. Mid-ocean-ridge seismicity reveals extreme types of ocean lithosphere. Nature 535(7611): 276279, doi: 10.1038/nature18277.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
Schmid, S. M., Boland, J. N., and Paterson, M. S. 1977. Superplastic flow in finegrained limestone. Tectonophysics 43: 257291.Google Scholar
Schmitt, S. V., Segall, P., and Matsuzawa, T. 2011. Shear heating-induced thermal pressurization during earthquake nucleation. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb008035.Google Scholar
Schmittbuhl, J., and Maloy, K. J. 1997. Direct observation of a self-affine crack propagation. Phys. Rev. Lett. 78(20): 38883891, doi: 10.1103/PhysRevLett.78.3888.Google Scholar
Scholz, C. H. 1968a. Microfracturing 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. 1968c. The frequency-magnitude relation of microfracturing in rock and its relation to earthquakes. Bull. Seismol. Soc. Am. 58: 399415.Google Scholar
Scholz, C. H. 1968d. Mechanism of creep in brittle rock. J. Geophys. Res. 73: 32953302.Google Scholar
Scholz, C. H. 1968e. Microfractures, aftershocks, and seismicity. Bull. Seismol. Soc. Am. 58: 11171130.Google Scholar
Scholz, C. H. 1972a. Static fatigue of quartz. J. Geophys. Res. 77: 21042114.Google Scholar
Scholz, C. H. 1972b. Crustal movements in tectonic areas. Tectonophysics 14: 201217.Google Scholar
Scholz, C. H. 1974. Post-earthquake dilatancy recovery. Geology 2: 551554.Google Scholar
Scholz, C. H. 1977. A physical interpretation of the Haicheng earthquake prediction. Nature 267: 121124.Google Scholar
Scholz, C. H. 1978. Velocity anomalies in dilatant rock. Science 201: 441442.Google Scholar
Scholz, C. H. 1980. Shear heating and the state of stress on faults. J. Geophys. Res. 85: 61746184.Google Scholar
Scholz, C. H. 1982. Scaling laws for large earthquakes: Consequences for physical models. Bull. Seismol. Soc. Am. 72: 114.Google Scholar
Scholz, C. H. 1985. The Black Mountain asperity: Seismic hazard on the San Francisco peninsula, California. Geophys. Res. Lett. 12: 717719.Google Scholar
Scholz, C. H. 1986. The Black Mountain asperity: Seismic hazard on the San Francisco peninsula, California. Geophys. Res. Lett., 12: 717719.Google Scholar
Scholz, C. H. 1987. Wear and gouge formation in brittle faulting. Geology 15: 493495.Google Scholar
Scholz, C. H. 1988a. The critical slip distance for seismic faulting. Nature 336: 761763.Google Scholar
Scholz, C. H. 1988b. The brittle-plastic transition and the depth of seismic faulting. Geol. Runds. 77: 319328.Google Scholar
Scholz, C. H. 1988c. Mechanisms of seismic quiescences. Pure Appl. Geophys. 126: 701718.Google Scholar
Scholz, C. H. 1989. Mechanics of faulting. Ann. Rev. Earth Planet. Sci. 17: 309334.Google Scholar
Scholz, C. H. 1991. Earthquakes and Faulting: Self-organized critical phenomena with a characteristic dimension. In Spontaneous formation of Space-time Structures and Criticality, eds. Riste, T. and Sherrington., D. Dortrect, Netherlands: Springer, pp. 4156.Google Scholar
Scholz, C. H. 1994a. Fractal transitions on geological surfaces. In Fractal Geometry and Its Use in the Earth Sciences, eds. Barton, C. and LaPointe, P.. New York: Plenum Press, pp. 131140.Google Scholar
Scholz, C. H. 1994b. A reappraisal of large earthquake scaling. Bull. Seismol. Soc. Amer. 84: 215218.Google Scholar
Scholz, C. H. 1994c. Reply to comments on “a reappraisal of large earthquake scaling.” Bull. Seismol. Soc. Amer. 84: 16771678.Google Scholar
Scholz, C. H. 1997a. Earthquake and fault populations and the calculation of brittle strain. Geowissenshaften 3–4: 124130.Google Scholar
Scholz, C. H. 1997b. Size distributions for large and small earthquakes. Bull. Seismol. Soc. Amer. 87: 10741077.Google Scholar
Scholz, C. H. 1998a. Earthquakes and friction laws. Nature 391: 3742.Google Scholar
Scholz, C. H. 1998b. A further note on earthquake size distributions. Bull. Seismol. Soc. Amer. 88: 13251326.Google Scholar
Scholz, C. H. 2000. Evidence for a strong San Andreas fault. Geology 28: 163166.Google Scholar
Scholz, C. H. 2006. The strength of the San Andreas Fault: A critical analysis, paper presented at Earthquakes: Radiated energy and the Physics of Faulting, AGU Monograph 170.Google Scholar
Scholz, C. H. 2010a. Large Earthquake Triggering, Clustering, and the Synchronization of Faults. Bull. Seismol. Soc. Am. 100(3): 901909, doi: 10.1785/0120090309.Google Scholar
Scholz, C. H. 2010b. A note on the scaling relations for opening mode fractures in rock. J. Struct. Geol. 32: 14851487.Google Scholar
Scholz, C. H. 2011a. Reply to Jon Olson and Richard Schultz. J. Struct. Geol. 33: 15251526.Google Scholar
Scholz, C. H. 2011b. First-order splay faults: Dip-slip examples. In Geology of the Earthquake Source: A Volume in Honour of Rick Sibson, eds. A. Fagereng, V. G. Toy, and J. V. Rowland, pp. 313318, doi: 10.1144/sp359.17.Google Scholar
Scholz, C. H. 2014a. Holocene earthquake history of Cascadia: A quantitative test. Bull. Seismol. Soc. Amer. 104: 21202124, doi: 10.1785/0120140002.Google Scholar
Scholz, C. H. 2014b. The rupture mode of the shallow large-slip surge of the Tohoku-oki Earthquake. Bull. Seismol. Soc. Amer. 104, doi: 10.1785/0120140130.Google Scholar
Scholz, C. H. 2015. On the stress dependence of the earthquake b value. Geophys. Res. Lett. 42: 1399-1402, doi: 10.1002/2014GL062863.Google Scholar
Scholz, C. H., and Aviles, C. 1986. The fractal geometry of faults and faulting. In Earthquake Source Mechanics AGU Geophys. Mono. 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 147155.Google Scholar
Scholz, C. H., and Campos, J. 1995. On the mechanism of seismic decoupling and back-arc spreading in subduction zones. J. Geophys. Res. 100: 2210322115.Google Scholar
Scholz, C. H., and Campos, J. 2012. The seismic coupling of subduction zones revisited. J. Geophys. Res. 117, doi: 10.1029/2011JB009003.Google Scholar
Scholz, C. H., and Contreras, J. C. 1998. Mechanics of continental rift architecture. Geology 26: 967970.Google Scholar
Scholz, C. H., and Cowie, P. A. 1990. Determination of total strain from faulting using slip measurements. Nature 346: 837839.Google Scholar
Scholz, C. H., and Engelder, T. 1976. Role of asperity indentation and ploughing in rock friction. Int. J. Rock Mech. Min. Sci. 13: 149–54.Google Scholar
Scholz, C. H., and Hanks, T. C. 2004. The strength of the San Andreas Fault: A discussion. In Rheology and Deformation of the Lithosphere at Continental Margins, eds. Karner, G. D., Taylor, B., Driscoll, N. W., and Kohlstedt, D. L., New York: Columbia University Press, pp. 261283.Google Scholar
Scholz, C. H., and Kato, T. 1978. The behavior of a convergent plate boundary: Crustal deformation in the south Kanto District, Japan. J. Geophys. Res. 83: 783791.Google Scholar
Scholz, C. H., and Koczynski, T. A. 1979. Dilatancy anisotropy and the response of rock to large cyclic loads. J. Geophys. Res. 84: 55255534.Google Scholar
Scholz, C. H., and Lawler, T. M. 2004. Slip tapers at the tips of faults and earthquake ruptures. Geophys. Res. Lett. 31(21): doi.org/10.1029/2004GL021030.Google Scholar
Scholz, C. H., and Martin, R. J. 1971. Crack growth and static fatigue in quartz. J. Am. Ceram. Soc. 54: 474.Google Scholar
Scholz, C. H., and Saucier, F. J. 1993. What do the Cajon Pass stress measurements say about stress on the San-Andreas fault – In-situ stress measurements to 3.5 km depth in the Cajon Pass scientific-research borehole – Implications for the mechanics of crustal faulting – comment. J. Geophys. Res.-Solid Earth 98: 1786717869.Google Scholar
Scholz, C. H., and Small, C. 1997. The effect of seamount subduction on seismic coupling. Geology 25: 487490.Google Scholar
Scholz, C. H., Ando, R., and Shaw, B. E. 2010. The mechanics of first order splay faulting: The strike-slip case. J. Struct. Geol. 32(1): 118126, doi: 10.1016/j.jsg.2009.10.007.Google Scholar
Scholz, C. H., Aviles, C., and Wesnousky, S. 1986. Scaling differences between large intraplate and interplate earthquakes. Bull. Seismol. Soc. Am. 76: 6570.Google Scholar
Scholz, C. H., Barazangi, M., and Sbar, M. L. 1971. Late Cenozoic evolution of the Great Basin, western United States, as an ensialic interarc basin. Geol. Soc. Am. Bull. 82: 29792990.Google Scholar
Scholz, C. H., Beavan, J., and Hanks, T. C. 1979. Frictional metamorphism, argon depletion, and tectonic stress on the Alpine fault, New Zealand. J. Geophys. Res. 84: 67706782.Google Scholar
Scholz, C. H., Boitnott, G. A., and Nemart-Nasser, S. 1986. The Bridgman ring paradox revisited. Pageoph 124: 587600.Google Scholar
Scholz, C. H., Dawers, N. H., Yu, J. Z., Anders, M. H., and Cowie, P. A. 1993. Fault growth and fault scaling laws – Preliminary results. J. Geophys. Res.-Solid Earth 98: 2195121961.Google Scholar
Scholz, C. H., Koczynski, T., and Hutchins, J. 1976. Evidence for incipient rifting in southern Africa. Geophys. J. R.A.S. 44: 135144.Google Scholar
Scholz, C. H., Molnar, P., and Johnson, T. 1972. Detailed studies of frictional sliding of granite and implications for the earthquake mechanism. J. Geophys. Res. 77: 63926406.Google Scholar
Scholz, C. H., Sykes, L. R., and Aggarwal, Y. P. 1973. Earthquake prediction: A physical basis. Science 181: 803810.Google Scholar
Scholz, C. H., Wyss, M., and Smith, S. W. 1969. Seismic and aseismic slip on the San Andreas fault. J. Geophys. Res. 74: 20492069.Google Scholar
Schorlemmer, D., Wiemer, S., and Wyss, M. 2005. Variations in earthquake-size distribution across different stress regimes. Nature 437(7058): 539542.Google Scholar
Schubnel, A., Brunet, F., Hilairet, N., Gasc, J., Wang, Y. B., and Green, H. W. 2013. Deep-focus earthquake analogs recorded at high pressure and temperature in the laboratory. Science 341(6152): 13771380, doi: 10.1126/science.1240206.Google Scholar
Schulson, E. M., and Fortt, A. L. 2012. Friction of ice on ice. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2012jb009219.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(11): 14051411.Google Scholar
Schulz, S., Burford, R. O., and Mavko, B. 1983. Influence of seismicity and rainfall on episodic creep on the San Andreas fault system in central California. J. Geophys. Res. 88(NB9): 74757484, doi: 10.1029/JB088iB09p07475.Google Scholar
Schulz, S. S., Mavko, G., Burford, R. O., and Stuart, W. D. 1982. Long-term fault creep observations in central California. J. Geophys. Res. 87: 69776982.Google Scholar
Schurr, B. Asch, G. Hainzl, S., et al. 2014. Gradual unlocking of plate boundary controlled initiation of the 2014 Iquique earthquake. Nature 512(7514): 299302, doi: 10.1038/nature13681.Google Scholar
Schwartz, D. P., and Coppersmith, K. J. 1984. Fault behavior and characteristic earthquakes: Examples from the Wasatch and San Andreas fault zones. J. Geophys. Res. 89: 56815698.Google Scholar
Schwartz, D. P., Hanson, K., and Swan, F. H. 1983. Paleoseismic investigations along the Wasatch Fault Zone: An update. In Paleoseismicity along the Wasatch Fault Zone and Adjacent Areas, Central Utah, ed. Crone, A. J.. Utah Geological and Mineral Survey Special Studies 62, pp. 4549.Google Scholar
Schwartz, S. Y., and Rokosky, J. M. 2007. Slow slip events and seismic tremor at circum-pacific subduction zones. Rev. Geophys., 45(3), doi.org/10.1029/2006RG000208.Google Scholar
Sclater, J. G., Jaupart, C., and Galson, D. 1980. The heat-flow through oceanic and continental-crust and the heat-loss of the Earth. Rev. Geophys. 18: 269311.Google Scholar
Secor, D. T. 1965. The role of fluid pressure in jointing. Am. J. Sci. 263: 633646.Google Scholar
Seeber, L., and Armbruster, J. 1981. The 1886 Charleston, South Carolina earthquake and the Appalachian detachment. J. Geophys. Res. 86: 78747894.Google Scholar
Seeber, L., and Armbruster, J. 1986. A study of earthquake hazard in New York State and adjacent areas. (NUREG/CR-4750). US Nuclear Regulatory Commission.Google Scholar
Seeber, L., and Armbruster, J. G. 1990. Fault kinematics in the 1989 Loma-Prieta rupture area during 20 years before that event. Geophys. Res. Lett. 17: 14251428.Google Scholar
Seeber, L., and Armbruster, J. G. 1998. Earthquakes, faults, and stress in Southern California. Southern California Earthquake Center.Google Scholar
Seeber, L., and Armbruster, J. G. 2000. Earthquakes as beacons of stress change. Nature 407: 6972.Google Scholar
Seeber, L., Armbruster, J. G., and Quittmeyer, R. C. 1981. Seismicity and continental subduction along the Himalayan arc. In Geodynamics Series, V., eds. Gupta, H. K. and Delany, F. M.. Washington, DC: American Geophysical Union, pp. 215242.Google Scholar
Segall, P. 1984. Formation and Growth of Extensional Fracture Sets. Geol. Soc. Amer. Bull. 95(4): 454462.Google Scholar
Segall, P. 1989. Earthquakes triggered by fluid extraction. Geology 17(10): 942946.Google Scholar
Segall, P., and Du, Y. 1993. How similar were the 1934 and 1966 Parkfield earthquakes? J. Geophys. Res. 98: 45274538.Google Scholar
Segall, P., and Harris, R. 1987. Earthquake deformation cycle on the San-Andreas fault near Parkfield, California. J. Geophys. Res. 92: 1051110525.Google Scholar
Segall, P., and Lisowski, M. 1990. Surface displacements in the 1906 San Francisco and 1989 Loma Prieta earthquakes. Science 250: 12411244.Google Scholar
Segall, P., and Lu, S. 2015. Injection-induced seismicity: Poroelastic and earthquake nucleation effects. J. Geophys. Res.-Solid Earth 120(7): 50825103, doi: 10.1002/2015jb012060.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 Rice, J. R. 1995. Dilatancy, compaction, and slip instability of a fluid-infiltrated fault. J. Geophys. Res. 100: 22,155122,171.Google Scholar
Segall, P., and Simpson, C. 1986. Nucleation of ductile shear zones on dilatant fractures. Geology 14: 56–9.Google Scholar
Segall, P., Burgmann, R., and Matthews, M. 2000. Time-dependent triggered afterslip following the 1989 Loma Prieta earthquake. J. Geophys. Res. 105: 56155634.Google Scholar
Segall, P., Desmarais, E. K., Shelly, D., Miklius, A., and Cervelli, P. 2006. Earthquakes triggered by silent slip events on Kilauea volcano, Hawaii. Nature 442(7098): 7174, doi: 10.1038/nature04938.Google Scholar
Segall, P., Grasso, J. R., and Mossop, A. 1994. Poroelastic stressing and induced seismicity near the Lacq gas field, southwestern France. J. Geophys. Res.-Solid Earth, 99(B8): 1542315438.Google Scholar
Semenov, A. N. 1969. Variations of the travel time of transverse and longitudinal waves before violent earthquakes. Izv. Acad. Sci. USSR, Phys. Solid Earth (Eng. Transl.) 3: 245258.Google Scholar
Seno, T. 1979. Pattern of intraplate seismicity in southwest Japan before and after great interplate earthquakes. Tectonophysics 57: 267283.Google Scholar
Seno, T., and Yamanaka, Y. 1996. Double seismic zones, compressional deep trench‐outer rise events, and superplumes, in Subduction Top to Bottom, Geophys. Monograph 96. Washington, D.C.: AGU, pp. 347355.Google Scholar
Shamir, G., and Zoback, M. D. 1992. Stress orientation profile to 3.5 km depth near the San Andreas fault at Cajon Pass, California. J. Geophys. Res. 97: 50595080.Google Scholar
Shapiro, S. A. 2015. Fluid-induced Seismicity. Cambridge: Cambridge University Press.Google Scholar
Shapiro, S. A., Patzig, R., Rothert, E., and Rindschwentner, J. 2003. Triggering of seismicity by pore-pressure perturbations: Permeability-related signatures of the phenomenon. Pure Appl. Geophys 160(5–6): 10511066, doi: 10.1007/pl00012560.Google Scholar
Sharp, R. V. 1979. Implications of surficial strike-slip fracture patterns for simplification and widening with depth. In Proc. Conf. VII – Analysis of Actual Fault-zones in Bedrock, US Geol. Surv. Open-file Rept 79–1239, pp. 6678.Google Scholar
Sharp, R. V. 1982. Comparison of 1979 surface faulting with earlier displacements on the Imperial fault, U.S. Geol. Surv. Prof. Paper, 1254, pp. 213221.Google Scholar
Sharp, R. V., and Clark, M. M. 1972. Geologic evidence of previous faulting near the 1968 rupture on the Coyote Creek fault. The Borrego Mountain, California Earthquake of April 9, 1968. US Geol Surv. Prof. Pap. 787: 131140.Google Scholar
Sharp, R. V., Lienkaemper, J., Bonilla, M., et al. 1982. Surface faulting in the central Imperial Valley. In The Imperial Valley, California, Earthquake of October 15, 1979. US Geol. Surv. Prof. Paper 1254, pp. 119144.Google Scholar
Shaw, B. E. 1995. Frictional weakening and slip complexity on earthquake faults. J. Geophys. Res. 100: 18,23918,248.Google Scholar
Shaw, B. E. 1997. Model quakes in the two-dimensional wave equation. J. Geophys. Res.-Solid Earth 102: 2736727377.Google Scholar
Shaw, B. E., and Scholz, C. H. 2001. Slip-length scaling in large earthquakes: Observations and theory and implications for earthquake physics. Geophys. Res. Lett. 28: 29952998.Google Scholar
Shaw, B. E., and Wesnousky, S. G. 2008. Slip-length scaling in large earthquakes: The role of deep-penetrating slip below the seismogenic layer. Bull. Seismol. Soc. Am. 98(4): 16331641, doi: 10.1785/0120070191.Google Scholar
Shearer, P. M. 2012. Self-similar earthquake triggering, Bath’s law, and foreshock/aftershock magnitudes: Simulations, theory, and results for southern California. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2011jb008957.Google Scholar
Shearer, P., and Burgmann, R. 2011. Lessons learned from the 2004 Sumatra-Andaman Megathrust Rupture. Ann. Rev. Earth Planet. Sci. 38: 103131, doi: 10.1146/annurev-earth-040809–152537.Google Scholar
Shebalin, P., and Narteau, C. 2017. Depth dependent stress revealed by aftershocks Nat. Comm. 8(1): 1317.Google Scholar
Shedlock, K. M., Giardini, D., Grünthal, G., and Zhang, P. 2000. The GSHA global seismic hazard map. Seismol. Res. Lett. 71: 679686.Google Scholar
Shelly, D. R. 2010. Migrating tremors illuminate complex deformation beneath the seismogenic San Andreas fault. Nature 463(7281): 648675.Google Scholar
Shelly, D. R., and Johnson, K. M. 2011. Tremor reveals stress shadowing, deep postseismic creep, and depth-dependent slip recurrence on the lower-crustal San Andreas fault near Parkfield. Geophys. Res. Lett. 38, doi: 10.1029/2011gl047863.Google Scholar
Shelly, D. R., Beroza, G. C., and Ide, S. 2007. Complex evolution of transient slip derived from precise tremor locations in western Shikoku, Japan. Geochem. Geophys. Geosystems 8, doi: 10.1029/2007gc001640.Google Scholar
Shen, Z.-K., Dong, D., Herring, T., et al. 1997. Crustal deformation measured in southern California. Eos 78: 477.Google Scholar
Shennan, I., Barlow, N., Carver, G., Davies, F., Garrett, E., and Hocking, E. 2014. Great tsunamigenic earthquakes during the past 1000 yr on the Alaska megathrust. Geology 42(8): 687690.Google Scholar
Shi, Z. Q., and Ben-Zion, Y. 2006. Dynamic rupture on a bimaterial interface governed by slip-weakening friction. Geophys. J. Int. 165(2): 469484, doi: 10.1111/j.1365-246X.2006.02853.x.Google Scholar
Shibazaki, B., and Iio, Y. 2003. On the physical mechanism of silent slip events along the deeper part of the seismogenic zone. Geophysical Research Letters 30(9).Google Scholar
Shiina, T., Nakajima, J., Matsuzawa, T., Toyokuni, G., and Kita, S. 2017. Depth variations in seismic velocity in the subducting crust: Evidence for fluid-related embrittlement for intermediate-depth earthquakes. Geophys. Res. Lett. 44(2): 810817, doi: 10.1002/2016gl071798.Google Scholar
Shimamoto, T. 1986. A transition between frictional slip and ductile flow undergoing large shearing deformation at room temperature. Science 231: 711714.Google Scholar
Shimamoto, T., and Logan, J. 1986. Velocity-dependent behavior of simulated halite shear zones: An analog for silicates. In Earthquake Source Mechanics. ACU Ceophys. Mono. 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 4964.Google Scholar
Shimamoto, T., and Noda, H. 2014. A friction to flow constitutive law and its application to a 2-D modeling of earthquakes. J. Geophys. Res.-Solid Earth 119(11): 80898106, doi: 10.1002/2014jb011170.Google Scholar
Shimazaki, K. 1976. Intraplate seismicity and interplate earthquakes – Historical activity in southwest Japan. Tectonophysics 33: 3342.Google Scholar
Shimazaki, K. 1986. Small and large earthquakes: The effects of the thickness of the seismogenic layer and the free surface. In Earthquake Source Mechanics. AGU Geophys. Mono. 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 209216.Google Scholar
Shimazaki, K., and Nakata, T. 1980. Time-predictable recurrence model for large earthquakes. Geophys. Res. Lett. 7: 279282.Google Scholar
Shipton, Z. K., and Cowie, P. A. 2001. Damage zone and slip-surface evolution over mu m to km scales in high-porosity Navajo sandstone, Utah. J. Struct. Geol. 23(12): 18251844, doi: 10.1016/s0191-8141(01)00035–9.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(3): 333344, doi: 10.1016/s0191-8141(02)00037–8.Google Scholar
Shipton, Z. K., Soden, A. M., Kirkpatrick, J. D., Bright, A. M., and Lunn, R. J. 2006. How thick is a fault? Fault displacement-thickness scaling revisited. In Earthquakes: Radiated Energy and the Physics of Faulting, eds. R. Abercrombie, A. McGarr, G. DiToro and H. Kanamori, pp. 193198, doi: 10.1029/170gm19.Google Scholar
Shlomai, H., and Fineberg, J. 2016. The structure of slip-pulses and supershear ruptures driving slip in bimaterial friction. Nature Communications 7: 11787, doi: 10.1038/ncomms11787.Google Scholar
Sibson, R. H. 1973. Interactions between temperature and pore fluid pressure during an earthquake faulting and a mechanism for partial or total stress relief. Nature 243: 6668.Google Scholar
Sibson, R. H. 1975. Generation of pseudotachylyte by ancient seismic faulting. Geophys. J. Roy. Astron. Soc. 43: 775794.Google Scholar
Sibson, R. H. 1977. Fault rocks and fault mechanisms. J. Geol. Soc. London 133: 191213.Google Scholar
Sibson, R. H. 1980a. Transient discontinuities in ductile shear zones. J. Struct Geol. 2: 165171.Google Scholar
Sibson, R. H. 1980b. Power dissipation and stress levels on faults in the upper crust. J. Geophys. Res. 85: 62396247.Google Scholar
Sibson, R. H. 1981. Fluid flow accompanying faulting: Field evidence and models. In Earthquake Prediction, an International Review. M. Ewing, Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 593604.Google Scholar
Sibson, R. H. 1982. Fault zone models, heat flow, and the depth distribution of earthquakes in the continental crust of the United States. Bull. Seismol. Soc. Am. 72: 151163.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. Stopping of earthquake ruptures at dilatational jogs. Nature 316: 248251.Google Scholar
Sibson, R. H. 1986a. Brecciation processes in fault zones: Inferences from earthquake rupturing. Pageoph 124: 159176.Google Scholar
Sibson, R. H. 1986b. Earthquakes and rock deformation in crustal fault zones. Ann. Rev. Earth Planet. Sci. 14: 149175.Google Scholar
Sibson, R. H. 1986c. Rupture interaction with fault jogs. In Earthquake Source Mechanics, AGU Geophys. Mono. 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 157168.Google Scholar
Sibson, R. H. 1987. Earthquake rupturing as a mineralizing agent in hydrothermal systems. Geology 15: 701704.Google Scholar
Sibson, R. H. 2003. Thickness of the seismic slip zone. Bull. Seismol. Soc. Amer. 93: 11691178.Google Scholar
Sibson, R. H., and Toy, V. G. 2006. The habitat of fault-generated pseudotachylyte: Presence vs. absence of friction melt. In Earthquakes: Radiated Energy and the Physics of Faulting, eds. Abercrombie, R., McGarr, A., DiToro, G. and Kanamori, H., Amer. Geophys. Union, Geophys. Monograph 170, pp. 153166.Google Scholar
Sibson, R. H., and Xie, G. Y. 1998. Dip range for intracontinental reverse fault ruptures: Truth not stranger than friction? Bull. Seismol. Soc. Amer. 88: 10141022.Google Scholar
Sibson, R. H., Roberts, F., and Poulsen, K. H. 1988. High-angle reverse faults, fluid pressure cycling, and mesothermal gold-quartz deposits. Geology 16: 551555.Google Scholar
Sieh, K. 1978. Slip along the San Andreas fault associated with the great 1857 earthquake. Bull. Seismol. Soc. Amer. 68: 14211428.Google Scholar
Sieh, K. 1981. A review of geological evidence for recurrence times of large earthquakes. In Earthquake Prediction: An International Review, eds. Simpson, D. and Richards, P. G.. Washington, DC: American Geophysical Union, pp. 209216.Google Scholar
Sieh, K. 1984. Lateral offsets and revised dates of large prehistoric earthquakes at Pallett Creek, southern California. J. Geophys. Res. 89: 76417670.Google Scholar
Sieh, K., and Jahns, R. 1984. Holocene activity of the San Andreas fault at Wallace Creek, California. Geol. Soc. Amer. Bull. 95: 883896.Google Scholar
Sieh, K., Jones, L., Hauksson, E., et al. 1993. Near-field investigations of the Landers earthquake sequence, April to July 1992. Science 260: 171176.Google Scholar
Sieh, K., Natawidjaja, D. H., Meltzner, A. J., et al. 2008. Earthquake supercycles inferred from sea-level changes recorded in the corals of West Sumatra. Science 322(5908): 16741678, doi: 10.1126/science.1163589.Google Scholar
Sieh, K., Stuiver, M., and Brillinger, D. 1989. A more precise chronology of earthquakes produced by the San Andreas fault in southern California. J. Geophys. Res. 94: 603623.Google Scholar
Silver, P. G., Beck, S. L., Wallace, T. C., et al. 1995. Rupture characteristics of the deep Bolivian earthquake of 9 June 1994 and the mechanism of deep focus earthquakes. Science 268(5207): 6973, doi: 10.1126/science.268.5207.69.Google Scholar
Siman-Tov, S., Aharonov, E., Boneh, Y., and Reches, Z. 2015. Fault mirrors along carbonate faults: Formation and destruction during shear experiments. Earth Planet. Sci. Lett. 430: 367376, doi: 10.1016/j.epsl.2015.08.031.Google Scholar
Siman-Tov, S., Aharonov, E., Sagy, A., and Emmanuel, S. 2013. Nanograins form carbonate fault mirrors. Geology 41(6): 703706, doi: 10.1130/g34087.1.Google Scholar
Simons, M., Minson, S. E., Sladen, A., et al. 2011. The 2011 magnitude 9.0 Tohoku-Oki Earthquake: Mosaicking the megathrust from seconds to centuries. Science 332(6036): 14211425.Google Scholar
Simpson, C. 1984. Borrego Springs-Santa Rosa mylonite zone: A late Cretaceous west-directed thrust in southern California. Geology 12: 811.Google Scholar
Simpson, C. 1985. Deformation of granitic rocks across the brittle-ductile transition. J. Struct. Geol. 5: 503512.Google Scholar
Simpson, C., and Schmid, S. M. 1983. An evaluation of criteria to deduce the sense of movement in sheared rock. Geol. Soc. Am. Bull. 94: 12811288.Google Scholar
Simpson, D. W. 1986. Triggered earthquakes. Ann. Rev. Earth Planet. Sci. 14: 2142.Google Scholar
Simpson, D. W., and Negmatullaev, S. K. 1981. Induced seismicity at Nurek reservoir, Tadjikistan, USSR. Bull. Seismol. Soc. Am. 71: 15611586.Google Scholar
Simpson, D. W., Leith, W. S., and Scholz, C. H. 1988. Two types of reservoir induced seismicity. Bull. Seismol. Soc. Am. 78: 20252040.Google Scholar
Singh, S. C., Hananto, N., Mukti, M., et al. 2011. Aseismic zone and earthquake segmentation associated with a deep subducted seamount in Sumatra. Nature Geoscience 4(5): 308.Google Scholar
Singh, S., and Suárez, G. 1988. Regional variation in the number of aftershocks (mb≧ 5) of large, subduction-zone earthquakes (Mw≧ 7.0). Bull. Seismol. Soc. Am. 78(1): 230242Google Scholar
Singh, S., Rodriguez, M., and Esteva, L. 1983. Statistics of small earthquakes and frequency of occurrence of large earthquakes along the Mexican subduction zone. Bull. Seismol. Soc. Am. 73: 17791796.Google Scholar
Sladen, A., Tavera, H., Simons, M., et al. 2010. Source model of the 2007 M-w 8.0 Pisco, Peru earthquake: Implications for seismogenic behavior of subduction megathrusts. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb006429.Google Scholar
Slemmons, D. B. 1957. Geological effects of the Dixie Valley–Fairview Peak, Nevada, earthquakes of December 16, 1954. Bull. Seismol. Soc. Am. 47: 353375.Google Scholar
Smith, D. E., and Dieterich, J. H. 2010. Aftershock Sequences Modeled with 3-D Stress Heterogeneity and Rate-State Seismicity Equations: Implications for Crustal Stress Estimation. Pure Appl. Geophys 167(8–9): 10671085, doi: 10.1007/s00024-010–0093-1.Google Scholar
Smith, D. K., Escartin, J., Cannat, M., et al. 2003. Spatial and temporal distribution of seismicity along the northern Mid-Atlantic Ridge (15 degrees-35 degrees N). J. Geophys. Res.-Solid Earth 108(B3): doi: 10.1029/2002jb001964.Google Scholar
Smith, R. B. 1975. Unified theory of onset of folding, boudinage, and mullion structure. Geol. Soc. Amer. Bull. 86(11): 16011609, doi: 10.1130/0016–7606(1975)86<1601:utotoo>2.0.co;2.Google Scholar
Smith, R. B. 1977. Formation of folds, boudinage, and mullions in non-Newtonian materials. Geol. Soc. Amer. Bull. 88(2): 312320, doi: 10.1130/0016–7606(1977)88<312:fofbam>2.0.co;2.Google Scholar
Smith, R. B., and Bruhn, R. L. 1984. Intraplate extensional tectonics of the eastern Basin-Range: Inferences on structural style from seismic reflection data, regional tectonics, and thermo-mechanical models of brittle-ductile transition. J. Geophys. Res. 89: 57335762.Google Scholar
Smith, S., and Wyss, M. 1968. Displacement on the San Andreas fault subsequent to the 1966 Parkfield earthquake. Bull. Seismol. Soc. Am. 58: 19551973.Google Scholar
Smith, W. D. 1981. The b-value as an earthquake precursor. Nature 289: 136139.Google Scholar
Snoke, A. W., Tullis, J., and Todd, V. R. 1998. Fault-related Rocks. New Jersey: Princeton University Press.Google Scholar
Snow, D. T. 1972. Geodynamics of seismic reservoirs. In Proc. Symp. Percolation through Fissured Rocks. Stuttgart: Ges. Erd- und Grundbau T2–J: 119.Google Scholar
Socquet, A., Valdes, J. P., Jara, J., Cotton, F., Walpersdorf, A., Cotte, N., Specht, S., Ortega-Culaciati, F., Carrizo, D., and Norabuena, E. 2017. An 8 month slow slip event triggers progressive nucleation of the 2014 Chile megathrust. Geophys. Res. Lett. 44(9): 40464053, doi: 10.1002/2017gl073023.Google Scholar
Soga, N., Mizutani, H., Spetzler, H., and Martin, R. J. 1978. The effect of dilatancy on velocity anisotropy in Westerly granite. J. Geophys. Res. 83: 44514458.Google Scholar
Soliva, R., Benedicto, A., and Maerten, L. 2006. Spacing and linkage of confined normal faults: Importance of mechanical thickness. J. Geophys. Res.-Solid Earth 111(B1): doi: 10.1029/2004jb003507.Google Scholar
Solomon, S. C., Huang, P. Y., and Meinke, L. 1988. The seismic moment budget of slowly spreading ridges. Nature 334: 5860.Google Scholar
Somerville, P. 1978. The accommodation of plate collision by deformation in the Izu block, Japan. Bull. Earthquake Res. Inst., Univ. Tokyo 53: 629648.Google Scholar
Somerville, P. G., McLaren, J. P., LeFevre, L. V., Burger, R. W., and Helmberger, D. V. 1987. Comparison of source scaling relations of eastern and western North American earthquakes. Bull. Seismol. Soc. Am. 77: 322346.Google Scholar
Sondergeld, C. H., and Esty, L. H. 1982. Source mechanisms and microfracturing during axial cycling of rock. Pure Appl. Geophys. 120: 151166.Google Scholar
Sone, H., Shimamoto, T., and Moore, D. E. 2012. Frictional properties of saponite-rich gouge from a serpentinite-bearing fault zone along the Gokasho-Arashima Tectonic Line, central Japan. J. Struct. Geol. 38: 172182, doi: 10.1016/j.jsg.2011.09.007.Google Scholar
Sornette, A., and Sornette, S. 1989. Self-organized criticality and earthquakes. Europhys. Lett. 9: 197202.Google Scholar
Sornette, D. 2003 Critical Phenomena in Natural Science, 2nd edn, Heidelberg: Springer.Google Scholar
Sornette, D., and Sammis, C. G. 1995. Complex critical exponents from renormalization-group theory of earthquakes – Implications for earthquake predictions. J. Phys. I 5: 607619.Google Scholar
Sornette, D., and Virieux, J. 1992. Linking short-timescale deformation to long-timescale tectonics. Nature 357: 401403.Google Scholar
Sowers, J. M., Unruh, J. R., Lettis, W. R., and Rubin, T. D. 1994. Relationship of the Kickapoo Fault to the Johnson Valley and Homestead Valley Faults, San-Bernardino County, California. Bull. Seismol. Soc. Amer. 84: 528542.Google Scholar
Spada, M., Tormann, T., Wiemer, S., and Enescu, B. 2013. Generic dependence of the frequency-size distribution of earthquakes on depth and its relation to the strength profile of the crust. Geophys. Res. Lett. 40: 709714, doi: 10.1029/2012GL054198.Google Scholar
Spagnuolo, E., Plumper, O., Violay, M., Cavallo, A., and Di Toro, G. 2015. Fast-moving dislocations trigger flash weakening in carbonate-bearing faults during earthquakes. Sci. Reports 5: 16112, doi: 10.1038/srep16112.Google Scholar
Spotilla, J. A., House, M. A., Niemi, N. A., Brady, R. C., Oskin, M., and Buscher, J. T. 2007. Patterns of bedrock uplift along the San Andreas fault and implications for mechanisms of transpression, in Exhumation Associated with Continental Strike-Slip Fault Systems: Geological Society of America Special Paper 434, eds. A. B. Till, S. M. Roeske, J. C. Sample, and D. A. Foster, pp. 1533, doi: 10.1130/2007.2434(02Google Scholar
Spottiswoode, S. M. 1984. Seismic deformation around Blyvooruitzicht Gold Mine. In Proc. Ist Int. Cong. Rockbursts and Seismicity in Mines South African Inst. Min. Met, eds. Gay, N. C. and Wainwright, E. H.. Johannesburg: pp. 2937.Google Scholar
Spottiswoode, S. M., and McGarr, A. 1975. Source parameters of tremors in a deep-level gold mine. Bull. Seismol. Soc. Am. 65: 93112.Google Scholar
Spray, J. G. 1987. Artificial generation of pseudotachylyte using friction welding apparatus: Simulation of melting on a fault plane. J. Struct. Geol. 9: 4960.Google Scholar
Spudich, P., Steck, L. K., Hellweg, M., Fletcher, J. B., and Baker, L. M. 1995. Transient stresses at Parkfield, California, produced by the M-7.4 Landers earthquake of June 28, 1992 – Observations from the Upsar Dense Seismograph Array. J. Geophys. Res.-Solid Earth 100: 675690.Google Scholar
Spyropoulos, C., Griffith, W. J., Scholz, C. H., and Shaw, B. E. 1999. Experimental evidence for different strain regimes of crack populations in a clay model. Geophys. Res. Lett. 26: 10811084.Google Scholar
Spyropoulos, C., Scholz, C. H. and Shaw, B. E. 2002. Transient regimes in growing crack populations. Phys. Rev. E., 65: 103105.Google Scholar
Staff, U. S. G. S. 1990. The Loma Prieta, California, earthquake: An anticipated event. Science 247: 286293.Google Scholar
Starr, A. 1928. Slip in a crystal and rupture in a solid due to shear. Proc. Cambridge Philos. Soc. 24: 489500.Google Scholar
Stauder, W. 1968. Mechanism of the Rat Island earthquake sequence of February 4, 1965, with relationships to island arcs and sea-floor spreading. J. Geophys. Res. 73: 38473854.Google Scholar
Steacy, S., Gerstenberger, M., Williams, C., Rhoades, D., and Christophersen, A. 2014. A new hybrid Coulomb/statistical model for forecasting aftershock rates. Geophys. J. Int. 196(2): 918923, doi: 10.1093/gji/ggt404.Google Scholar
Steacy, S., Gomberg, J., and Cocco, M. 2005. Introduction to special section: Stress transfer, earthquake triggering, and time-dependent seismic hazard. J. Geophys. Res.-Solid Earth 110(B5): doi: 10.1029/2005jb003692.Google Scholar
Steblov, G. M., Kogan, M. G., Levin, B. V., Vasilenko, N. F., Prytkov, A. S., and Frolov, D. I. 2008. Spatially linked asperities of the 2006–2007 great Kuril earthquakes revealed by GPS. Geophys. Res. Lett., 35(22), doi.org/10.1029/2008GL035572.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. 1984. Seismic potential revealed by surface folding: The 1983 Coalinga, California earthquake. Science 224: 869871.Google Scholar
Stein, R. S., and Lisowski, M. 1983. The 1979 Homestead Valley earthquake sequence, California: Control of aftershocks and postseismic deformations. J. Geophys. Res. 88: 64776490.Google Scholar
Stein, R. S., Barka, A. A., and Dieterich, J. H. 1997. Progressive failure on the North Anatolian fault since 1939 by earthquake stress triggering. Geophys. J. Int. 128: 594604.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 265: 14321435.Google Scholar
Stein, S., and Okal, E. A. 1978. Seismicity and tectonics of the ninetyeast ridge area – Evidence for internal deformation of the Indian plate. J. Geophys. Res. 83(NB5), 22332245, doi: 10.1029/JB083iB05p02233.Google Scholar
Stein, S., and Okal, E. A. 2007. Ultralong period seismic study of the december 2004 Indian Ocean Earthquake and the implications for regional tectonics and the subduction process. Bull. Seismol. Soc. Am. 97(1A): S279-S295.Google Scholar
Stein, S., and Pelayo, A. 1991. Seismological constraints on stress in the oceanic lithosphere. Phil. Trans. R. Soc. Lond. A 337: 5372.Google Scholar
Steinbrugge, K. V., Zacher, E. G., Tocher, D., Whitten, C. A., and Claire, C. N. 1960. Creep on the San Andreas fault. Bull. Seismol. Soc. Am. 50: 389415.Google Scholar
Stel, H. 1981. Crystal growth in cataclasites: Diagnostic microstructures and implications. Tectonophysics 78: 585600.Google Scholar
Stel, H. 1986. The effect of cyclic operation of brittle and ductile deformation on the metamorphic assemblage in cataclasites and mylonites. Pageoph 124: 289307.Google Scholar
Stesky, R. 1975. The mechanical behavior of faulted rock at high temperature and pressure. PhD, Cambridge, Massachusetts: Massachusetts Institute of Technology.Google Scholar
Stesky, R. 1978. Mechanisms of high temperature frictional sliding in Westerly granite. Can. J. Earth Sci. 15: 361375.Google Scholar
Stesky, R., and Hannan, G. 1987. Growth of contact areas between rough surfaces under normal stress. Geophys. Res. Lett. 14: 550553.Google Scholar
Stesky, R., Brace, W., Riley, D., and Robin, P-Y. 1974. Friction in faulted rock at high temperature and pressure. Tectonophysics 23: 177203.Google Scholar
Stipp, M., Stunitz, H., Heilbronner, R., and Schmid, S. M. 2002a. Dynamic recrystallization of quartz: Correlation between natural and experimental conditions. In Deformation Mechanisms, Rheology and Tectonics: Current Status and Future Perspectives, eds. S. DeMeer, M. R. Drury, J. H. P. DeBresser and G. M. Pennock, pp. 171190, doi: 10.1144/gsl.sp.2001.200.01.11.Google Scholar
Stipp, M., Stunitz, H., Heilbronner, R., and Schmid, S. M. 2002b. The eastern Tonale fault zone: A “natural laboratory” for crystal plastic deformation of quartz over a temperature range from 250 to 700 degrees C. J. Struct. Geol. 24(12): 18611884, doi: 10.1016/s0191-8141(02)00035–4.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
St-Laurent, F. 2000. The Sanguenay, Quebec, Earthquake lights of November 1988–January 1989. Seismol. Res. Lett. 71: 160183.Google Scholar
Stoker, J. J. 1950. Nonlinear Vibrations. New York: Interscience.Google Scholar
Strelau, J. 1986. A discussion of the depth extent of rupture in large continental earthquakes. In Earthquake Source Mechanics. AGU Geophys. Mono. 37, eds. Das, J. B. S. and Scholz, C.. Washington, DC: American Geophysical Union, pp. 131146.Google Scholar
Strogatz, S. 2003. Sync. New York: Theia.Google Scholar
Strogatz, S. H. 2001. Exploring complex networks. Nature 410: 268276.Google Scholar
Stroup, D. F., Tolstoy, M., Crone, T. J., Malinverno, A., Bohnenstiehl, D. R., and Waldhauser, F. 2009. Systematic along-axis tidal triggering of microearthquakes observed at 9 degrees 50 N East Pacific Rise. Geophys. Res. Lett. 36, doi: 10.1029/2009gl039493.Google Scholar
Stuart, W. D. 1979. Strain softening prior to 2-dimensional strike slip earthquakes. J. Geophys. Res. 84: 10631070.Google Scholar
Stuart, W. D. 1988. Forecast model for great earthquakes at the Nankai trough, southwest Japan. Pageoph 126: 619642.Google Scholar
Stuart, W. D., and Aki, K. 1988. Intermediate-term earthquake prediction – Introduction. Pure Appl. Geophys. 126: 175176.Google Scholar
Stuart, W. D., and Mavko, G. M. 1979. Earthquake instability on a strike-slip fault. J. Geophys. Res. 84: 21532160.Google Scholar
Stuwe, K. 1998. Heat sources of Cretaceous metamorphism in the Eastern Alps – A discussion. Tectonophysics 287: 251269.Google Scholar
Subarya, C., Chlieh, M., Prawirodirdjo, L., et al. 2006. Plate-boundary deformation associated with the great Sumatra–Andaman earthquake. Nature 440(7080): 4651.Google Scholar
Sugiura, N., Hori, T., and Kawamura, Y. 2014. Synchronization of coupled stick-slip oscillators. Nonlin. Processes Geophys. 21: 251267, doi: 10.5194/npg-21–251-2014.Google Scholar
Suh, N. P., and Sin, H. C. 1981. The genesis of friction. Wear 69: 91114.Google Scholar
Summers, R., and Byerlee, J. 1977. Summary of results of frictional sliding studies, at confining pressures up to 6.98 kb, in selected rock materials. In US Geol. Surv. Open-file Rept., pp. 77142.Google Scholar
Sundaram, P., Goodman, R., and Wang, C.-Y. 1976. Precursory and coseismic water pressure variations in stick-slip experiments. Geology 4: 108110.Google Scholar
Suppe, J. 1985. Principles of Structural Geology. Englewood Cliffs, New Jersey: Prentice-Hall.Google Scholar
Susong, D. D., Janecke, S. U., and Bruhn, R. L. 1990. Structure of a fault segment boundary in the Lost River fault zone, Idaho, a possible effect on the 1983 Borah Peak earthquake rupture. Bull. Seismol. Soc. Amer. 80: 5768.Google Scholar
Suyehiro, S., Asada, T., and Ohtake, M. 1964. Foreshocks and aftershocks accompanying a perceptible earthquake in central Japan – On a peculiar nature of foreshocks. Papers Meteorol. Geophys. 15: 7188.Google Scholar
Svetlizky, I., and Fineberg, J. 2014. Classical shear cracks drive the onset of dry frictional motion. Nature 509(7499): 205208, doi: 10.1038/nature13202.Google Scholar
Swan, F. H. 1988. Temporal clustering of paleoseismic events on the Oued Fodda fault, Algeria. Geology 16(12): 10921095, doi: 10.1130/0091–7613(1988)016<1092:tcopeo>2.3.co;2.Google Scholar
Swan, F. H., Schwartz, D. P., and Cluff, L. S. 1980. Recurrence of moderate to large magnitude earthquakes produced by surface faulting on the Wasatch fault zone, Utah. Bull. Seismol. Soc. Amer. 70: 14311462.Google Scholar
Swan, G., and Zongqi, S. 1985. Prediction of shear behavior of joints using profiles. Rock Mech. Rock Eng. 18: 183212.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 ultrasonic and microscopic investigation. J. Geophys. Res. 92: 80158036.Google Scholar
Sykes, L. R. 1967. Mechanism of earthquakes and nature of faulting on the mid-oceanic ridges. J. Geophys. Res. 72: 2131.Google Scholar
Sykes, L. R. 1970a. Earthquake swarms and sea-floor spreading. J. Geophys. Res. 75: 65986611.Google Scholar
Sykes, L. R. 1970b. Seismicity of the Indian Ocean and a possible nascent island arc between Ceylon and Australia. J. Geophys. Res. 75: 50415055.Google Scholar
Sykes, L. R. 1971. Aftershock zones of great earthquakes, seismicity gaps, and earthquake prediction for Alaska and the Aleutians. J. Geophys. Res. 76: 80218041.Google Scholar
Sykes, L. R. 1978. Intra-plate seismicity, reactivation of pre-existing zones of weakness, alkaline magmatism, and other tectonics post-dating continental separation. Rev. Geophys. Space Phys. 16: 621688.Google Scholar
Sykes, L. R., and Ekstrom, G. 2012. Earthquakes along Eltanin transform system, SE Pacific Ocean: Fault segments characterized by strong and poor seismic coupling and implications for long-term earthquake prediction. Geophys. J. Int. 188(2): 421434, doi: 10.1111/j.1365-246X.2011.05284.x.Google Scholar
Sykes, L. R., and Jaumé, S. C. 1990. Seismic activity on neighboring faults as a long-term precursor to large earthquakes in the San Francisco Bay area. Nature 348: 595599.Google Scholar
Sykes, L. R., and Menke, W. 2006. Repeat times of large earthquakes: Implications for earthquake mechanics and long-term prediction. Bull. Seismol. Soc. Am. 96(5): 15691596, doi: 10.1785/0120050083.Google Scholar
Sykes, L. R., and Nishenko, S. P. 1984. Probabilities of occurrence of large plate rupturing earthquakes for the San Andreas, San Jacinto, and Imperial faults, California. J. Geophys. Res. 89: 5905–5027.Google Scholar
Sykes, L. R., and Quittmeyer, R. C. 1981. Repeat times of great earthquakes along simple plate boundaries. In Earthquake Prediction, an International Review. Ewing, M., Ser. 4., eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 217247.Google Scholar
Sykes, L. R., and Sbar, M. L. 1973. Intraplate earthquakes, lithospheric stresses and the driving mechanism of plate tectonics. Nature 245: 298302.Google Scholar
Sykes, L. R., Kisslinger, J., House, L., Davies, J., and Jacob, K. 1981. Rupture zones and repeat times of great earthquakes along the Alaska-Aleutian arc. In Earthquake Prediction: An International Review, Ewing, M., Ser. 4., eds. Simpson, D. and Richards, P. G.. Washington, DC: American Geophysical Union, pp. 7380.Google Scholar
Sykes, L. R., Shaw, B. E., and Scholz, C. H. 1999. Rethinking earthquake prediction. Pageoph 155: 207232.Google Scholar
Tada, H., Paris, P., and Irwin, G. 1973. The Stress Analysis of Cracks Handbook. Hellertown Pennsylvania: Del Research Corp.Google Scholar
Tajima, F., and Kanamori, H. 1985. Global survey of aftershock area expansion patterns. Phys. Earth Planet. Int. 40: 77134.Google Scholar
Talwani, P. 1997. On the nature of reservoir-induced seismicity. Pageoph 150: 473492.Google Scholar
Talwani, P., and Acree, S. 1985. Pore-pressure diffusion and the mechanism of reservoir-induced seismicity. Pageoph 122: 947965.Google Scholar
Talwani, P., and Rajendran, K. 1991. Some seismological and geometric features of intraplate earthquakes. Tectonophysics 186: 1941.Google Scholar
Tanaka, S. 2010. Tidal triggering of earthquakes precursory to the recent Sumatra megathrust earthquakes of 26 December 2004 (M-w 9.0), 28 March 2005 (M-w 8.6), and 12 September 2007 (M-w 8.5). Geophys. Res. Lett. 37, doi: 10.1029/2009gl041581.Google Scholar
Tanaka, S. 2012. Tidal triggering of earthquakes prior to the 2011 Tohoku-Oki earthquake (M-w 9.1). Geophys. Res. Lett. 39, doi: 10.1029/2012gl051179.Google Scholar
Tanioka, Y., and Seno, T. 2001. Sediment effect on tsunami generation of the 1896 Sanriku tsunami earthquake. Geophys. Res. Lett. 28(17): 33893392, doi: 10.1029/2001gl013149.Google Scholar
Tape, C., Holtkamp, S., Silwal, V., Hawthorne, J., Kaneko, Y., Ampuero, J. P., Ji, C., Ruppert, N., Smith, K., and West, M. E. 2018. Earthquake nucleation and fault slip complexity in the lower crust of central Alaska. Nature Geoscience 11(7):536541, doi: 10.1038/s41561-018-0144-2.Google Scholar
Tapponnier, P., and Brace, W. F. 1976. Development of stress-induced microcracks in Westerly granite. Int. J. Rock Mech. Min. Sci. 13: 103112.Google Scholar
Taylor, F. W., Frohlich, C., Lecolle, J., and Strekler, M. 1987. Analysis of partially emerged corals and reef terraces in the central Vanuatu arc: Comparison of contemporary coseismic and nonseismic with Quarternary vertical movements. J. Geophys. Res. 92: 49054933.Google Scholar
Tchalenko, J. S. 1970. Similarities between shear zones of different magnitudes. Bull. Geol. Soc. Am. 81: 16251640.Google Scholar
Tchalenko, J. S., and Berberian, M. 1975. Dasht-e-Bayez fault, Iran: Earthquake and earlier related structures in bed rock. Geol. Soc. Am. Bull. 86: 703709.Google Scholar
Tembe, S., Lockner, D. A., and Wong, T. F. 2010. Effect of clay content and mineralogy on frictional sliding behavior of simulated gouges: Binary and ternary mixtures of quartz, illite, and montmorillonite. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb006383.Google Scholar
Terada, M., Yanigadani, T., and Ehara, S. 1984. AE rate controlled compression tests of rocks. In Proc. of the Third Conf on Acoustic Emission/Microseismic Activity in Geologic Structures and Materials, eds. Hardy Jr, M. R. and Leighton, F. W., Clausal Germany: Trans. Tech. Publ., pp. 159171.Google Scholar
Terakawa, T., Zoporowski, A., Galvan, B., and Miller, S. A. 2010. High-pressure fluid at hypocentral depths in the L’Aquila region inferred from earthquake focal mechanisms. Geology 38(11): 995998, doi: 10.1130/g31457.1.Google Scholar
Tesei, T., Collettini, C., Carpenter, B. M., Viti, C., and Marone, C. 2012. Frictional strength and healing behavior of phyllosilicate-rich faults. J. Geophys. Res.-Solid Earth 117: B09402, doi: 10.1029/2012jb009204.Google Scholar
Teufel, L. W., and Logan, J. M. 1978. Effect of displacement rate on the real area of contact and temperatures generated during frictional sliding of Tennessee sandstone. Pure Appl. Geophys. 116: 840872.Google Scholar
Teyssier, C., and Tikoff, B. 1998. Strike-slip partitioned transpression of the San Andreas fault system: A lithospheric-scale approach. In Continental Transpression and Transtensional Tectonics, eds. Holdsworth, R. E., Stachan, R. E., Dewey, R. A., Geological Society of London Special Publication 135, pp. 143158.Google Scholar
Teyssier, C., Tikoff, B., and Markley, M. 1995. Oblique plate motion and continental tectonics, Geology, 23: 447450.Google Scholar
Thatcher, W. 1972. Regional variations of seismic source parameters in northern Baja California area. J. Geophys Res 77: 1549.Google Scholar
Thatcher, W. 1975a. Nonlinear strain buildup and the earthquake cycle on the San Andreas fault. J. Geophys. Res. 88: 589358902.Google Scholar
Thatcher, W. 1975b. Strain accumulation and release mechanism of the 1906 San Francisco earthquake. J. Geophys. Res 80: 48624872.Google Scholar
Thatcher, W. 1983. Nonlinear strain buildup and the earthquake cycle on the San Andreas fault. J. Geophys. Res. 88: 58935902.Google Scholar
Thatcher, W. 1984. The earthquake deformation cycle, recurrence, and the time-predictable model. J. Geophys. Res. 89: 56745680.Google Scholar
Thatcher, W. 1990. Order and diversity in the modes of circum-Pacific earthquake recurrence. J. Geophys. Res 95: 26092623.Google Scholar
Thatcher, W., and Hanks, T. C. 1973. Source parameters of southern-Californian earthquakes. J. Geophys. Res. 78: 85478576.Google Scholar
Thatcher, W., and Lisowski, M. 1987. Long-term seismic potential of the San Andreas fault southeast of San Francisco, California. J. Geophys. Res. 92: 47714784.Google Scholar
Thatcher, W., Matsuda, T., Kato, T., and Rundle, J. B. 1980. Lithospheric loading by the 1896 Riku-u earthquake, northern Japan: Implications for plate flexure and asthenospheric rheology. J. Geophys. Res. 85: 64296435.Google Scholar
Thio, H. K., and Kanamori, H. 1996. Source complexity of the 1994 Northridge earthquake and its relation to aftershock mechanisms. Bull. Seismol. Soc. Amer. 86: S84S92.Google Scholar
Thomas, D. 1988. Geochemical precursors to seismic activity. Pageoph 126: 241267.Google Scholar
Thomas, M. Y., Avouac, J. P., Champenois, J., Lee, J. C., and Kuo, L. C. 2014a. Spatiotemporal evolution of seismic and aseismic slip on the Longitudinal Valley Fault, Taiwan. J. Geophys. Res.-Solid Earth 119(6): 51145139, doi: 10.1002/2013jb010603.Google Scholar
Thomas, M. Y., Avouac, J. P., Gratier, J. P., and Lee, J. C. 2014b. Lithological control on the deformation mechanism and the mode of fault slip on the Longitudinal Valley Fault, Taiwan. Tectonophysics 632: 4863, doi: 10.1016/j.tecto.2014.05.038.Google Scholar
Tichelaar, B. W., and Ruff, L. J. 1993. Depth of seismic coupling along subduction zones. J. Geophys. Res.-Solid Earth 98: 20172037.Google Scholar
Tingle, T. N., Green II, H. W., Scholz, C. H., and Koczynski, T. A. 1993. The rheology of faults triggered by the olivine-spinel transformation in Mg2GeO4 and its implications for the mechanism of deep-focus earthquakes. J. Struct. Geol. 15:12491256.Google Scholar
Tinti, E., Scognamiglio, L., Michelini, A., and Cocco, M. 2016. Slip heterogeneity and directivity of the ML 6.0, 2016, Amatrice earthquake estimated with rapid finite‐fault inversion. Geophys. Res. Lett. 43(20): 10,74510,752.Google Scholar
Tocher, D. 1959. Seismic history of the San Francisco region. In Calif. Div. Mines Spec. Rept 57. Sacramento, California: Calif. Div. of Mines, pp. 3949.Google Scholar
Tocher, D. 1960. Creep rate and related measurements at Vineyard, California. Bull. Seismol. Soc. Am. 50: 396404.Google Scholar
Toda, S., Stein, R. S., Reasonberg, P. A., Dieterich, J. H., and Yoshida, A. 1998. Stress transferred by the M = 6.9 Kobe, Japan, shock: Effect of aftershocks on future earthquake probabilities. J. Geophys. Res. 103: 2454324565.Google Scholar
Toda, S., Stein, R. S., Richards-Dinger, K., and Bozkurt, S. B. 2005. Forecasting the evolution of seismicity in southern California: Animations built on earthquake stress transfer. J. Geophys. Res.-Solid Earth 110(B5): doi: 10.1029/2004jb003415.Google Scholar
Tong, X., Sandwell, D. T., and Smith-Konter, B. 2013. High-resolution interseismic velocity data along the San Andreas Fault from GPS and InSAR. J. Geophys. Res., 118: 369389, doi: 10.1029/2012JB009442.Google Scholar
Toomey, D. R., Solomon, S. C., Purdy, G. M., and Murray, M. H. 1985. Microearthquakes beneath the median valley of the Mid-Atlantic ridge near 23° N: Hypocenters and focal mechanisms. J. Geophys. Res. 90: 54435458.Google Scholar
Toriumi, M. 1982. Strain, stress, and uplift. Tectonics 1: 5772.Google Scholar
Townend, J. 2006. What do faults feel? Observational constraints on the stresses acting on seismogenic faults. In Earthquakes: Radiated Energy and the Physics of Faulting, eds. Abercrombie, A. M. R., Di Toro, G., Kanamori, H., Amer. Geophys. Union Geophys. Mono 170, Washington, D.C, pp. 313327.Google Scholar
Townend, J., and Zoback, M. D. 2000. How faulting keeps the crust strong. Geology 28: 399402.Google Scholar
Townend, J., and Zoback, M. D. 2004. Regional tectonic stress near the San Andreas fault in central and southern California. Geophys. Res. Lett. 31(15): doi: 10.1029/2003gl018918.Google Scholar
Toyoda, H., and Noma, Y. 1952. Study for underground condition of the Dogo hot spring area, Ehime Prefecture. Mem. Ehime Univ., Sec. Il (Science) I: 139146.Google Scholar
Trehu, A. M., and Solomon, S. C. 1983. Earthquakes in the Orozco Transform zone: Seismicity, source mechanisms, and earthquakes. J. Geophys. Res. 88: 82038225.Google Scholar
Tributsch, H. 1983. When the Snakes Awake. Cambridge, Massachusetts: MIT Press.Google Scholar
Triep, E. G., and Sykes, L. R. 1997. Frequency of occurrence of moderate to great earthquakes in intracontinental regions: Implications for changes in stress, earthquake prediction, and hazards assessments. J. Geophys. Res.-Solid Earth 102: 99239948.Google Scholar
Tsang, L. L. H., Meltzner, A. J., Hill, E. M., Freymueller, J. T., and Sieh, K. 2015. A paleogeodetic record of variable interseismic rates and megathrust coupling at Simeulue Island, Sumatra. Geophys. Res. Lett. 42(24): 1058510594, doi: 10.1002/2015gl066366.Google Scholar
Tse, S., and Rice, J. 1986. Crustal earthquake instability in relation to the depth variation of frictional slip properties. J. Geophys. Res. 91: 94529472.Google Scholar
Tse, S., Dmowska, R., and Rice, J. R. 1985. Stressing of locked patches along a creeping fault. Bull. Seismol. Soc. Am. 75: 709736.Google Scholar
Tsuboi, C. 1933. Investigation of deformation of the crust found by precise geodetic means. Japan J. Astron. Geophys. 10: 93248.Google Scholar
Tsuboi, C. 1956. Earthquake energy, earthquake volume, aftershock area, and strength of the earth’s crust. J. Phys. Earthq. 4: 6366.Google Scholar
Tsumura, K., Karakama, I., Ogino, I., and Takahashi, M. 1978. Seismic activities before and after the Izu-Oshima-kinkai earthquake of 1978. Bull. Earthquake Res. Inst., Univ. Tokyo 53: 309315.Google Scholar
Tsuneishi, Y., Ito, T., and Kano, K. 1978. Surface faulting associated with the 1978 Izu-Oshima-kinkai earthquake. Bull. Earthquake Res. Inst., Univ. Tokyo 53: 649674.Google Scholar
Tsunogai, U., and Wakita, H. 1995. Precursory chemical changes in groundwater: Kobe earthquake, Japan. Science 269: 6163.Google Scholar
Tsuru, T., Park, J. O., Miura, S., Kodaira, S., Kido, Y., and Hayashi, T. 2002. Along‐arc structural variation of the plate boundary at the Japan Trench margin: Implication of interplate coupling. J. Geophys. Res.-Solid Earth 107(B12).Google Scholar
Tsutsumi, A., and Shimamoto, T. 1997. High-velocity frictional properties of gabbro. Geophys. Res. Lett. 24: 699702.Google Scholar
Tucker, B. C., and Brune, J. N. 1973. Seismograms, S-wave spectra and source parameters for aftershocks of the San Fernando earthquake, in San Fernando Earthquake of February 9, 1971, v.III, US Dept Commerce, pp: 69122.Google Scholar
Tullis, J. 2002. Deformation of granitic rocks: Experimental studies and natural examples. In Plastic Deformation of Minerals and Rocks, eds. S. Karato and H. R. Wenk, pp. 5195, doi: 10.2138/gsrmg.51.1.51.Google Scholar
Tullis, J., and Yund, R. A. 1977. Experimental deformation of dry Westerly granite. J. Geophys. Res. 82: 57055718.Google Scholar
Tullis, J., and Yund, R. A. 1980. Hydrolytic weakening of experimentally deformed Westerly granite and Hale albite rock. J. Struct. Geol. 2: 439451.Google Scholar
Tullis, J., and Yund, R. A. 1987. Transition from cataclastic flow to dislocation creep of feldspar: Mechanisms and microstructure. Geology 15: 606609.Google Scholar
Tullis, J., and Yund, R. A. 1991. Diffusion creep in feldspar aggregates-experimental evidence. J. Struct. Geol. 13(9): 9871000, doi: 10.1016/0191–8141(91)90051-j.Google Scholar
Tullis, T. E. 2015. Friction of rock at earthquake slip rates. In Treatise on Geophysics, 2nd edn., ed. Shubert, G., pp. 131152, Elsevier.Google Scholar
Tullis, T. E., and Weeks, J. D. 1986. Constitutive behavior and stability of frictional sliding of granite. Pageoph 124: 383414.Google Scholar
Turcotte, D. L., and Spence, D. A. 1974. An analysis of strain accumulation on a strike-slip fault. J. Geophys. Res. 79: 44074412.Google Scholar
Uchida, N., Hasegawa, A., Matsuzawa, T., and Igarashi, T. 2004. Pre- and post-seismic slow slip on the plate boundary off Sanriku, NE Japan associated with three interplate earthquakes as estimated from small repeating earthquake data. Tectonophysics 385(1–4): 115, doi: 10.1016/j.tecto.2004.04.015.Google Scholar
Ulomov, V. I., and Mavashev, B. Z. 1971. The Tashkent Earthquake of 26 April. Tashkent: Acad. Nauk. Uzbeck. SSR, FAN.Google Scholar
Usami, T. 1987. Descriptive Catalogue of Damaging Earthquakes in Japan (rev. edn.). Tokyo: University of Tokyo Press.Google Scholar
Utsu, T. 1970. Aftershocks and earthquake statistics II: Further investigations of aftershocks and other earthquake sequences based on a new classification of earthquake sequences. J. Fac. Sci. Hokkaido Univ. Ser. VII 3: 379441.Google Scholar
Utsu, T. 1971. Aftershocks and earthquake statistics (III). J. Fac. Science, Hokkaido Univ. Ser. VII. Geophysics) 3: 379441.Google Scholar
Uyeda, S. 1982. Subduction zones: An introduction to comparative subductology. Tectonophysics 81: 133159.Google Scholar
Uyeda, S., and Kanamori, H. 1979. Back-arc opening and the mode of subduction. J. Geophys. Res. 84: 10491061.Google Scholar
Vail, J. R. 1967. The southern extension of the East Africa rift system and related igneous activity. Geol. Runds. 57: 601614.Google Scholar
Vallianatos, F., Papadakis, G., and Michas, G. 2016. Generalized statistical mechanics approaches to earthquakes and tectonics. P. Roy. Soc. A-math. Phy. 472(2196): doi: 10.1098/rspa.2016.0497.Google Scholar
Valoroso, L., Chiaraluce, L., Piccinini, D., Di Stefano, R., Schaff, D., and Waldhauser, F. 2013. Radiography of a normal fault system by 64,000 high-precision earthquake locations: The 2009 L’Aquila (central Italy) case study. J. Geophys. Res.-Solid Earth 118(3): 11561176, doi: 10.1002/jgrb.50130.Google Scholar
van der Elst, N. J., and Brodsky, E. E. 2010. Connecting near-field and far-field earthquake triggering to dynamic strain. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2009jb006681.Google Scholar
van der Elst, N. J., and Savage, H. M. 2015. Frequency dependence of delayed and instantaneous triggering on laboratory and simulated faults governed by rate-state friction. J. Geophys. Res.-Solid Earth 120(5): 34063429, doi: 10.1002/2014jb011611.Google Scholar
van der Elst, N. J., and Shaw, B. E. 2015. Larger aftershocks happen farther away: Nonseparability of magnitude and spatial distributions of aftershocks. Geophys. Res. Lett. 42(14): 57715778.Google Scholar
van der Elst, N. J., Savage, H. M., Keranen, K. M., and Abers, G. A. 2013. Enhanced Remote Earthquake Triggering at Fluid-Injection Sites in the Midwestern United States. Science 341(6142): 164167, doi: 10.1126/science.1238948.Google Scholar
Van Diggelen, E. W. E., De Bresser, J. H. P., Peach, C. J., and Spiers, C. J. 2010. High shear strain behaviour of synthetic muscovite fault gouges under hydrothermal conditions. J. Struct. Geol. 32(11): 16851700, doi: 10.1016/j.jsg.2009.08.020.Google Scholar
van Stiphout, T., Wiemer, S., and Marzocchi, W. 2010. Are short-term evacuations warranted? Case of the 2009 L’Aquila earthquake. Geophys. Res. Lett. 37, doi: 10.1029/2009gl042352.Google Scholar
Vanorman, J., Cochran, J. R., Weissel, J. K., and Jestin, F. 1995. Distribution of shortening between the Indian and Australian plates in the central Indian Ocean. Earth Planet. Sci. Lett. 133(1–2): 3546, doi: 10.1016/0012-821x(95)00061-g.Google Scholar
Vening Meinesz, F. A. 1950. Les graben africains resultant de compression ou de tension dans le crout terrestre? Inst. Roy. Colon. Belge, Bull. 21: 539552.Google Scholar
Verberne, B. A., Niemeijer, A. R., De Bresser, J. H. P., and Spiers, C. J. 2015. Mechanical behavior and microstructure of simulated calcite fault gouge sheared at 20–600 degrees C: Implications for natural faults in limestones. J. Geophys. Res.-Solid Earth 120(12): 81698196, doi: 10.1002/2015jb012292.Google Scholar
Verberne, B. A., Plumper, O., de Winter, D. A. M., and Spiers, C. J. 2014. Superplastic nanofibrous slip zones control seismogenic fault friction. Science 346(6215): 13421344, doi: 10.1126/science.1259003.Google Scholar
Verberne, B. A., Spiers, C. J., Niemeijer, A. R., De Bresser, J. H. P., De Winter, D. A. M., and Plumper, O. 2014b. Frictional properties and microstructure of calcite-rich fault gouges sheared at sub-seismic sliding velocities. Pure Appl. Geophys 171 (10): 26172640, doi: 10.1007/s00024-013–0760-0.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.-Solid Earth 103: 1222312237.Google Scholar
Vermilye, J. M., and Scholz, C. H. 1999. Fault propagation and segmentation: Insight from the microstructural examination of a small fault. J. Struct. Geol. 21: 16231636.Google Scholar
Versfelt, J., and Rosendahl, B. R. 1989. Relationship between pre-rift structure and rift architecture in Lakes Tanganyika and Malawi, East Africa. Nature 337: 354357.Google Scholar
Vidale, J. E., and Houston, H. 1993. The depth dependence of earthquake duration and implications for rupture mechanisms. Nature 365: 4547.Google Scholar
Vidale, J. E., and Shearer, P. M. 2006. A survey of 71 earthquake bursts across southern California: Exploring the role of pore fluid pressure fluctuations and aseismic slip as drivers. J. Geophys. Res.-Solid Earth 111(B5): doi: 10.1029/2005jb004034.Google Scholar
Vidale, J. E., Agnew, D., Johnston, M., and Oppenheimer, D. 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
Vidale, J. E., Ellsworth, W. L., Cole, A., and Marone, C. 1994. Variations in rupture process with recurrence interval in a repeated small earthquake. Nature 368: 624626.Google Scholar
Viesca, R. C., and Garagash, D. I. 2015. Ubiquitous weakening of faults due to thermal pressurization. Nature Geoscience 8 (11): 875-+, doi: 10.1038/ngeo2554.Google Scholar
Villaggio, P. 1979. An elastic theory of Coulomb friction. Arch. Rational Mech. Anal. 70: 135143.Google Scholar
Villemin, T., Angelier, J., and Sunwoo, C. 1995. Fractal distribution of fault length and offsets: implications of brittle deformation evaluation – The Lorraine coal basin. In Fractals in the Earth Sciences, eds. Barton, C. C. and LaPointe, P. R.. New York: Plenum, pp. 205225.Google Scholar
Violay, M., Di Toro, G., Nielsen, S., Spagnuolo, E., and Burg, J. P. 2015. Thermo-mechanical pressurization of experimental faults in cohesive rocks during seismic slip. Earth Planet. Sci. Lett. 429: 110, doi: 10.1016/j.epsl.2015.07.054.Google Scholar
Violay, M., Nielsen, S., Gibert, B., et al. 2014. Effect of water on the frictional behavior of cohesive rocks during earthquakes. Geology 42 (1): 2730.Google Scholar
Voight, B. 1976. Mechanics of Thrust Faults and Decollements. Stroudsburg, Pennsylvania: Dowden, Huchinson, and Ross.Google Scholar
Voight, B. 1989. A relation to describe rate-dependent material failure. Science 243: 200203.Google Scholar
Voll, G. 1976. Recrystallization of quartz, biotite, and feldspars from Erstfeld to the Levantina nappe, Swiss Alps, and its geological implications. Schweiz. Miner. Petrogr. Mitt. 56: 641647.Google Scholar
Von Herzen, R., Ruppel, C., Molnar, P., Nettles, M., Nagihara, S., and Ekstrom, G. 2001. A constraint on the shear stress at the Pacific–Australia plate boundary from heat flow and seismicity at the Kermadec forearc. J. Geophys. Res. 106: 68176833.Google Scholar
Wadati, K. 1928. Shallow and deep earthquakes. Geophys. Mag. 1: 161202.Google Scholar
Wakita, H. 1988. Short term and intermediate term geochemical precursors. Pageoph 126: 267278.Google Scholar
Wakita, H. 1996. Geochemical challenge to earthquake prediction. Proc. NAT. Acad. Sci. USA 93: 37813786.Google Scholar
Wald, D. J., and Heaton, T. H. 1994. Spatial and temporal distribution of slip for the 1992 Landers, California, earthquake. Bull. Seismol. Soc. Amer. 84: 668691.Google Scholar
Wald, D. J., Heaton, T. H., and Hudnut, K. W. 1996. The slip history of the 1994 Northridge, California, earthquake determined from strong-motion, teleseismic, GPS, and leveling data. Bull. Seismol. Soc. Amer. 86: S49S70.Google Scholar
Wald, D. J., Helmberger, D. V. and Heaton, T. H. 1991. Rupture model of the 1989 Loma-Prieta earthquake from the inversion of strong-motion and broad-band teleseismic data. B Seismol. Soc. Am 81(5): 15401572.Google Scholar
Walker, A. N., Rutter, E. H., and Brodie, K. H. 1990. Experimental study of grain size sensitive flow of synthetic, hot-pressed calcite rocks. In Deformation mechanisms, Rheology and Tectonics, eds. Knipe, R. J. and Rutter, E. H.. Geol. Soc. London Spec. Publ. 54, pp. 259284.Google Scholar
Wallace, L. M., and Beavan, J. 2010. Diverse slow slip behavior at the Hikurangi subduction margin, New Zealand. J. Geophys. Res.-Solid Earth 115, doi: 10.1029/2010jb007717.Google Scholar
Wallace, L. M., and Eberhart-Phillips, D. 2013. Newly observed, deep slow slip events at the central Hikurangi margin, New Zealand: Implications for downdip variability of slow slip and tremor, and relationship to seismic structure. Geophys. Res. Lett. 40: 53935398, doi: 10.1002/2013GL057682.Google Scholar
Wallace, L. M., Beavan, J., Bannister, S., and Williams, C. 2012. Simultaneous long-term and short-term slow slip events at the Hikurangi subduction margin, New Zealand: Implications for processes that control slow slip event occurrence, duration, and migration. J. Geophys. Res.-Solid Earth 117, doi: 10.1029/2012jb009489.Google Scholar
Wallace, R. E. 1973. Surface fracture patterns along the San Andreas fault. In Proc. Conf. Tectonic Problems of the San Andreas Fault System. Spec. Publ. Geol. Sci. 13, eds. Kovach, R. and Nur. Palo Alto, A., California: Stanford University Press, pp. 248250.Google Scholar
Wallace, R. E. 1981. Active faults paleoseismology, and earthquake hazards in the Western United States. In Earthquake Prediction, an International Review. Ewing, M.; Ser. 4, eds. Simpson, D. and Richards, P.. Washington, DC: American Geophysical Union, pp. 209216.Google Scholar
Wallace, R. E. 1987. Grouping and migration of surface faulting and variations of slip rate on faults in the Great Basin province. Bull. Seismol. Soc. Am. 77: 868876.Google Scholar
Wallace, R. E. 1989. Segmentation in fault zones. In Proc. USGS Conference of Segmentation of Faults, US Geol Surv. Open-file Rept., pp. 89315.Google Scholar
Wallace, R. E., and Morris, H. T. 1986. Characteristics of faults and shear zones in deep mines. Pageoph 124: 107125.Google Scholar
Wallace, R. E., Davis, J. F., and McNally, K. C. 1984. Terms for expressing earthquake potential, prediction, and probability. Bull. Seismol. Soc. Amer. 74: 18191825.Google Scholar
Walsh, F. R., and Zoback, M. D. 2015. Oklahoma’s recent earthquakes and saltwater disposal. Sci. Adv., 1(5): e1500195.Google Scholar
Walsh, J. B. 1965. The effect of cracks on the compressibility of rocks. J. Geophys. Res. 70: 381389.Google Scholar
Walsh, J. B., and Rice, J. R. 1979. Local changes in gravity resulting from deformation. J. Geophys. Res.-Solid Earth 84(B1): 165170.Google Scholar
Walsh, J. J., and Watterson, J. 1987. Distribution of cumulative displacement and of 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: 238347.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(8): 12511262, doi: 10.1016/s0191-8141(02)00161-x.Google Scholar
Walsh, J. J., Nicol, A., and Childs, C. 2002. An alternative model for the growth of faults. J. Struct. Geol. 24(11): 16691675, doi: 10.1016/s0191-8141(01)00165–1.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
Walter, J. I., Brodsky, E. E., Tulaczyk, S., Schwartz, S. Y., and Pettersson, R. 2011. Transient slip events from near‐field seismic and geodetic data on a glacier fault, Whillans Ice Plain, West Antarctica. J.Geophys. Res. 116, doi: 10.1029/2010JF001754.Google Scholar
Walter, J. I., Svetlizky, I., Fineberg, J., et al. 2015. Rupture speed dependence on initial stress profiles: Insights from glacier and laboratory stick-slip. Earth Planet. Sci. Lett. 411: 112120.Google Scholar
Walters, R. J., Elliott, J. R., D’Agostino, N., et al. 2009. The 2009 L’Aquila earthquake. central Italy): A source mechanism and implications for seismic hazard. Geophys. Res. Lett. 36, doi: 10.1029/2009gl039337.Google Scholar
Wang, C.-Y. 1984. On the constitution of the San Andreas fault zone in central California. J. Geophys. Res. 89: 58585866.Google Scholar
Wang, C.-Y., Rui, F., Zhengshen, Y., and Xingjue, S. 1986. Gravity anomaly and density structure of the San Andreas fault zone. Pageoph 124: 127140.Google Scholar
Wang, D., Mori, J., and Uchide, T. 2012. Supershear rupture on multiple faults for the Mw 8.6 Off Northern Sumatra, Indonesia earthquake of April 11, 2012. Geophys. Res. Lett. 39(21): doi.org/10.1029/2012GL053622.Google Scholar
Wang, K. L., and Bilek, S. L. 2014. Invited review paper: Fault creep caused by subduction of rough seafloor relief. Tectonophysics 610: 124, doi: 10.1016/j.tecto.2013.11.024.Google Scholar
Wang, K., and Bilek, S. L. 2011. Do subducting seamounts generate or stop large earthquakes? Geology 39(9): 819822.Google Scholar
Wang, K., and Rogers, G. C. 2014. Earthquake preparedness should not fluctuate on a daily or weekly basis. Seismol. Res. Lett. 85: 569571Google Scholar
Wang, K., He, J., and Davis, E. E. 1997. Transform push, oblique subduction resistance, and intraplate stress of the Juan de Fuca plate. J. Geophys. Res.-Solid Earth 102(B1): 661674.Google Scholar
Wang, K., Mulder, T., Rogers, G. C., and Hyndman, R. D. 1995. Case for very low coupling stress on the Cascadia subduction fault. J. Geophys. Res. 100: 1290712918.Google Scholar
Wang, W. B., and Scholz, C. H. 1994a. Micromechanics of the velocity and normal stress dependence of rock friction. Pure Appl. Geophys. 143: 303315.Google Scholar
Wang, W. B., and Scholz, C. H. 1994b. Wear processes during frictional sliding of rock – A theoretical and experimental-study. J. Geophys. Res.-Solid Earth 99: 67896799.Google Scholar
Wang, W. B., and Scholz, C. H. 1995. Micromechanics of rock friction 3. Quantitative modeling of base friction. J. Geophys. Res.-Solid Earth 100: 42434247.Google Scholar
Wang, W., and Scholz, C. H. 1993. Scaling of constitutive parameters of friction for fractal surfaces. Int. J. Rock Mech. Min. Sci. 30: 13591365.Google Scholar
Wang, W., and Shearer, P. M. 2015. No clear evidence for localized tidal periodicities in earthquakes in the central Japan region. J. Geophys. Res.-Solid Earth 120(9): 63176328, doi: 10.1002/2015jb011937.Google Scholar
Ward, S., and Barrientos, S. 1986. An inversion for slip distribution and fault shape from geodetic observations of the 1983, Borah Peak, Idaho, earthquake. J. Geophys. Res. 91: 49094919.Google Scholar
Washington, DC: American Geophysical Union, pp. 629–634.Google Scholar
Watt, J. T., Ponce, D. A., Graymer, R. W., Jachens, R. C., and Simpson, R. W. 2014. Subsurface geometry of the San Andreas–Calaveras fault junction: Influence of serpentinite and the Coast Range Ophiolite. Tectonics 33: 20252044, doi: 10.1002/2014TC003561.Google Scholar
Watterson, J. 1986. Fault dimensions, displacements and growth. Pageoph 124: 365373.Google Scholar
Wawersik, W., and Brace, W. F. 1971. Post-failure behavior of a granite and a diabase. Rock Mech. 3: 6185.Google Scholar
Wech, A. G., and Creager, K. C. 2011. A continuum of stress, strength and slip in the Cascadia Subduction zone. Nature Geoscience 4: 624628.Google Scholar
Wech, A. G., Creager, K. C., and Melbourne, T. I. 2009. Seismic and geodetic constraints on cascadia slow slip. J. Geophys. Res., 114: B10316, doi.10310.11029/12008JB006090.Google Scholar
Weertman, J. 1980. Unstable slippage accross a fault that separates elastic media of different elastic constants. J. Geophys. Res. 85: 14551461.Google Scholar
Weertman, J., and Weertman, J. R. 1964. Elementary Dislocation Theory. New York: MacMillan.Google Scholar
Wei, S., Helmberger, D., and Avouac, J. P. 2013. Modeling the 2012 Wharton basin earthquakes off‐Sumatra: Complete lithospheric failure. J. Geophys. Res.-Solid Earth 118(7): 35923609.Google Scholar
Weissel, J. K., Anderson, R. N., and Geller, C. A. 1980. Deformation of the Indo-Australian plates. Nature 287(5780): 284291, doi: 10.1038/287284a0.Google Scholar
Weldon, R. J., Scharer, K. M., Fumal, T. E., and Biasi, G. P. 2004. Wrightwook and the earthquake cycle: What a long recurrence record tells us about how faults work. GSA Today 14, doi: 10.1130/1052–5173(2004)014<4:WATECW>2.0.CO;2.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. Amer. 84: 9741002.Google Scholar
Wenk, H. 1978. Are pseudotachylites products of fracture or fusion? Geology 6: 507511.Google Scholar
Wenk, H., and Weiss, L. 1982. Al-rich calcic pyroxene in pseudotachylyte: An indicator of high pressure and temperature? Tectonophysics 84: 329341.Google Scholar
Wernicke, B. 1995. Low-angle normal faults and seismicity – A review. J. Geophys. Res.-Solid Earth 100: 2015920174.Google Scholar
Wernicke, B., and Burchfiel, B. C. 1982. Modes of extensional tectonics. J. Struct. Geol. 4: 105115.Google Scholar
Wesnousky, S. G. 1986. Earthquakes, Quaternary faults, and seismic hazard in California. J. Geophys. Res. 91: 12,58712,631.Google Scholar
Wesnousky, S. G. 1988. Seismological and structural evolution of strike-slip faults. Nature 335: 340342.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(8): 15051512.Google Scholar
Wesnousky, S. G. 2006. Predicting the endpoints of earthquake ruptures. Nature 444(7117): 358360, doi: 10.1038/nature05275.Google Scholar
Wesnousky, S. G., and Scholz, C. H. 1980. The craton: Its effect on the distribution of seismicity and stress in North America. Earth Planet. Sci. Lett. 48: 348355.Google Scholar
Wesnousky, S. G., Scholz, C. H., and Shimazaki, K. 1982. Deformation of an Island Arc: Rates of moment release and crustal shortening in intraplate Japan determined from seismicity and quaternary fault data. J. Geophys. Res. 87: 68296852.Google Scholar
Wesnousky, S. G., Scholz, C. H., and Shimazaki, K. 1983. Earthquake frequency distribution and the mechanics of faulting. J. Geophys. Res. 88: 93319340.Google Scholar
Wesnousky, S. G., Scholz, C. H., Shimazaki, K., and Matsuda, T. 1984. Integration of geological and seismological data for the analysis of seismic hazard – A case-study of Japan. Bull. Seismol. Soc. Amer. 74: 687708.Google Scholar
Wesson, R. L., and Ellsworth, W. L. 1973. Seismicity preceding moderate earthquakes in California J. Geophys. Res. 78: 85278546.Google Scholar
Westbrook, J., and Jorgensen, P. 1968. Effects of water desorption on indentation hardness. Am. Mineral 53: 18991904.Google Scholar
Wetzler, N., Brodsky, E. E., and Lay, T. 2016. Regional and stress drop effects on aftershock productivity of large megathrust earthquakes. Geophys. Res. Lett. 43(23): 1201212020, doi: 10.1002/2016gl071104.Google Scholar
Whitcomb, J. H., Garmany, J. D., and Anderson, D. L. 1973b. Earthquake prediction – Variation of seismic velocities before San Francisco earthquake. Science 180: 632635.Google Scholar
Whitcomb, J., Allen, C., Garmany, J., and Hileman, J. 1973a. The 1971 San Fernando earthquake series: Focal mechanisms and tectonics. Rev. Geophys. Space Phys. 11: 693730.Google Scholar
White, S. 1975. Tectonic deformation and recrystallization of oligoclase. Contr. Mineral. Petrol. 50: 287304.Google Scholar
White, S., Burrows, S., Carreras, J., Shaw, N., and Humphreys, F. 1980. On mylonites in ductile shear zones. J. Struct. Geol. 2: 175187.Google Scholar
Whitten, C. A. 1948. Horizontal Earth movement, vicinity of San Francisco, California. Eos (Trans. American Geophysical Union), 29: 318323, doi: 10.1029/TR029i003p00318.Google Scholar
Wibberley, C. A. J., Petit, J. P., and Rives, T. 2000a. Mechanics of cataclastic “deformation band” faulting in high-porosity sandstone, Provence. Comptes Rendus De L Academie Des Sciences Serie Ii Fascicule a-Sciences De La Terre Et Des Planetes, 331(6): 419425, doi: 10.1016/s1251-8050(00)01423–3.Google Scholar
Wibberley, C. A. J., Petit, J. P., and Rives, T. 2000b. Micromechanics of shear rupture and the control of normal stress. J. Struct. Geol. 22(4): 411427, doi: 10.1016/s0191-8141(99)00158–3.Google Scholar
Wibberley, C. A. J., Yielding, G., and Di Toro, G. 2008. Recent advances in the understanding of fault zone internal structure: A review. In Internal Structure of Fault Zones: Implications for Mechanical and Fluid-Flow Properties, eds. C. A. J. Wibberley, W. Kurz, J. Imber, R. E. Holdsworth, and C. Collettini, pp. 533, doi: 10.1144/sp299.2.Google Scholar
Wiederhorn, S. M., and Bolz, L. H. 1970. Stress corrosion and static fatigue of glass. J. Am. Ceram. Soc. 53: 543551.Google Scholar
Wiens, D. A. 2001. Seismological constraints on the mechanism of deep earthquakes: Temperature dependence of deep earthquake source properties. Phys. Earth Planet. Int. 127(1–4): 145163, doi: 10.1016/s0031-9201(01)00225–4.Google Scholar
Wiens, D. A., Anandakrishnan, S., Winberry, J. P., and King, M. A. 2008. Simultaneous teleseismic and geodetic observations of the stick-slip motion of an Antarctic ice stream. Nature 453(7196): 770773, doi: 10.1038/nature06990.Google Scholar
Wiens, D. A., and McGuire, J. J. 1995. The 1994 Bolivia and Tonga events: Fundamentally different types of deep earthquakes? Geophys. Res. Lett. 22(16): 22452248.Google Scholar
Wiens, D. A., and McGuire, J. J. 2000. Aftershocks of the March 9, 1994, Tonga earthquake: The strongest known deep aftershock sequence. J. Geophys. Res.-Solid Earth 105: 1906719083.Google Scholar
Wiens, D. A., and Stein, S. 1983. Age dependence of oceanic intraplate seismicity and implications for lithospheric evolution. J. Geophys. Res.-Solid Earth 88(B8): 64556468.Google Scholar
Wiens, D. A., and Stein, S. 1984. Intraplate seismicity and stresses in young oceanic lithosphere. J. Geophys. Res. 89(NB13): 14421464, doi: 10.1029/JB089iB13p11442.Google Scholar
Wiens, D. A., and Stein, S. 1985. Implications for oceanic intraplate seismicity for plate stresses, driving forces and rheology. Tectonophysics 116(1–2): 143162, doi: 10.1016/0040–1951(85)90227–6.Google Scholar
Wiens, D., Stein, S., DeMets, C., Gordon, R., and Stein, C. 1986. Plate tectonic models for Indian Ocean “intraplate” deformation. Tectonophysics 132: 3748.Google Scholar
Wilcock, W. S. D. 2009. Tidal triggering of earthquakes in the Northeast Pacific Ocean. Geophys. J. Int. 179(2): 10551070, doi: 10.1111/j.1365-246X.2009.04319.x.Google Scholar
Wilcock, W. S. D., Archer, S. D., and Purdy, G. M. 2002. Microearthquakes on the Endeavour segment of the Juan de Fuca Ridge. J. Geophys. Res.-Solid Earth 107(B12): doi: 10.1029/2001jb000505.Google Scholar
Wilkinson, M. W., McCaffrey, K. J., Jones, R. R., et al. 2017. Near-field fault slip of the 2016 Vettore M w 6.6 earthquake (Central Italy) measured using low-cost GNSS. Scientific Reports 7(1): 4612.Google Scholar
Wilkinson, M., Roberts, G., McCaffrey, K., et al. 2015. Slip distributions on active normal faults measured from LiDAR and field mapping of geomorphic offsets: An example from L’Aquila, Italy, and implications for modelling seismic moment release. Geomorphology 237: 130141, doi: 10.1016/j.geomorph.2014.04.026.Google Scholar
Willemse, E. J. M. 1997. Segmented normal faults: Correspondence between three dimensional mechanical models and field data. J. Geophys. Res.-Solid Earth 102: 675692.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
Williams, C. F., and Narasimhan, T. N. 1989. Hydrogeologic constraints on heat-flow along the San Andreas fault – A testing of hypothesis. Earth Planet. Sci. Lett. 92: 131143.Google Scholar
Wilson, B., Dewers, T., Reches, Z., and Brune, J. 2005. Particle size and energetics of gouge from earthquake rupture zones. Nature 434(7034): 749752, doi: 10.1038/nature03433.Google Scholar
Wilson, J. E., Chester, J. S., and Chester, F. M. 2003. Microfracture analysis of fault growth and wear processes, Punchbowl Fault, San Andreas System, California. J. Struct. Geol. 25(11): 18551873, doi: 10.1016/s0191-8141(03)00036–1.Google Scholar
Wilson, J. T. 1965. A new class of faults and their bearing on continental drift. Nature, 207(4995): 343.Google Scholar
Winberry, J. P., Anandakrishnan, S., Alley, R. B., Bindschadler, R. A., and King, M. A. 2009. Basal mechanics of ice streams: Insights from the stick-slip motion of Whillans Ice Stream, West Antarctica. J. Geophys. Res. 114, doi: 10.1029/2008JF001035.Google Scholar
Winberry, J. P., Anandakrishnan, S., Alley, R. B., Wiens, D. A., and Pratt, M. J. 2014. Tidal pacing, skipped slips and the slowdown of Whillans Ice Stream, Antarctica. J. Glaciology 60, doi: 10.3189/2014JoG14J038.Google Scholar
Winberry, J. P., Anandakrishnan, S., Wiens, D. A., Alley, R. B., and Christianson, K. 2011. Dynamics of stick-slip motion, Whillans Ice Stream, Antarctica. Earth Planet. Sci. Lett. 305(3–4): 283289, doi: 10.1016/j.epsl.2011.02.052.Google Scholar
Winberry, J. P., Anandakrishnan, S., Wiens, D. A., and Alley, R. B. 2013. Nucleation and seismic tremor associated with the glacial earthquakes of Whillans Ice Stream, Antarctica. Geophys. Res. Lett, 40, doi: 10.1002/grl.50130.Google Scholar
Wise, D. U., Dunn, E., Engelder, T., et al. 1984. Fault-related rocks: Suggestions for terminology. Geology 12: 391394.Google Scholar
Wolfson-Schwehr, M., Boettcher, M. S., McGuire, J. J., and Collins, J. A. 2014. The relationship between seismicity and fault structure on the Discovery transform fault, East Pacific Rise. Geochem. Geophys. Geosystems 15(9): 36983712, doi: 10.1002/2014gc005445.Google Scholar
Wong, T.-F. 1982. Shear fracture energy of Westerly granite from postfailure behavior. J. Geophys. Res. 87: 9901000.Google Scholar
Wong, T.-F. 1986. On the normal stress dependence of the shear fracture energy. In Earthquake Source Mechanics. AGU Geophys. Mono 37, eds. Das, S., Boatwright, J., and Scholz, C.. Washington, DC: American Geophysical Union, pp. 112.Google Scholar
Woodcock, N. H., and Fischer, M. 1986. Strike-slip duplexes. J. Struct. Geol. 8: 725735.Google Scholar
Working Group on California Earthquake Probabilities. 1988. Probabilities of large earthquake occurring in California on the San Andreas fault. U.S. Geol Surv. Open-file Rept., pp. 88398.Google Scholar
Working Group on California Earthquake Probabilities. 1990. Probabilities of Large Earthquakes in the San Francisco Bay Region, California. U.S. Geol Survey Circular 1053.Google Scholar
Wu, C. J., Takeo, M., and Ide, S. 2001. Source process of the Chi-Chi earthquake: A joint inversion of strong motion data and global positioning system data with a multifault model. Bull. Seismol. Soc. Am. 91(5): 11281143.Google Scholar
Wyss, M. 1997. Second round of evaluations of proposed earthquake precursors. Pure Appl. Geophys. 149: 316.Google Scholar
Wyss, M., and Brune, J. 1967. The Alaska earthquake of 28 March 1964: A complex multiple rupture. Bull. Seismol. Soc. Am. 57: 10171023.Google Scholar
Wyss, M., and Brune, J. N. 1968. Seismic moment stress and source dimentions for earthquakes in the California-Nevada region. J. Geophys. Res. 73. 14): 46814694, doi: 10.1029/JB073i014p04681.Google Scholar
Wyss, M., and Habermann, R. E. 1988. Precursory seismic quiescence. Pure Appl. Geophys. 126: 319332.Google Scholar
Wyss, M., and Molnar, P. 1972. Efficiency, stress drop, apparent stress, effective stress, and frictional stress of Denver, Colorado, earthquakes. J. Geophys. Res. 77: 1433.Google Scholar
Wyss, M., Klein, F., and Johnston, A. 1981. Precursors of the Kalapana M = 7.2 earthquake. J. Geophys. Res. 86: 38813900.Google Scholar
Wyss, M., Shimazaki, K., and Ito, A. 1999. Seismicity patterns, their statistical significance and physical meaning. Pure Appl. Geophys. 155: 203205.Google Scholar
Wyss, M., Shimazaki, K., and Urabe, T. 1996. Quantitative mapping of a precursory seismic quiescence to the Izu-Oshima 1990 (M6.5) earthquake, Japan. Geophys. J. Int. 127: 735743.Google Scholar
Wyss, M., Westerhaus, M., Berckhemer, H., and Ates, R. 1995. Precursory seismic quiescence in the Mudurnu Valley, North Anatolian fault zone, Turkey. Geophys. J. Int. 123: 117124.Google Scholar
Yagi, Y., Okuwaki, R., Enescu, B., Hirano, S., Yamagami, Y., Endo, S., and Komoro, T. 2014. Rupture process of the 2014 Iquique Chile earthquake in relation with the foreshock activity. Geophys. Res. Lett. 41(12): 42014206.Google Scholar
Yamada, T., Mori, J. J., Ide, S., et al. 2007. Stress drops and radiated seismic energies of microearthquakes in a South African gold mine. J. Geophys. Res.-Solid Earth 112(B3).Google Scholar
Yamada, T., Takida, N., Kagami, J., and Naoi, T. 1978. Mechanisms of elastic contact and friction between rough surfaces. Wear 48: 1534.Google Scholar
Yamamoto, Y., Hino, R., and Shinohara, M. 2011. Mantle wedge structure in the Miyage Prefecture forearc region, central northeastern Japan arc, and its relation to corner-flow pattern and interplate coupling. J. Geophys. Res. 116(B10310): doi: 10.1029/2011JB008470.Google Scholar
Yamanaka, Y., and Shimazaki, K. 1990. Scaling relationship between the number of aftershocks and the size of the main shock. J. Phys. Earth 38: 305324.Google Scholar
Yamashita, T., and Knopoff, L. 1987. Models of aftershock occurrence. Geophys. J. R. A. S. 91(1): 1326, doi: 10.1111/j.1365-246X.1987.tb05210.x.Google Scholar
Yanigadani, T., Ehara, S., Nushizawa, O., Kusenose, K., and Terada, M. 1985. Localization of dilatancy in Oshima granite under constant uniaxial stress. J. Geophys. Res. 90: 6840–58.Google Scholar
Yao, L., Ma, S., Niemeijer, A., Shimamoto, T., and Platt, J. D. 2016. Is frictional heating needed to cause dramatic weakening of nanoparticle gouge during seismic slip? Insights from friction experiments with variable thermal evolutions. Geophys. Res. Lett. 43: 68526860, doi: 10.1002/2016GL069053.Google Scholar
Ye, L. L., Kanamori, H., Avouac, J. P., Li, L. Y., Cheung, K. F., and Lay, T. 2016a. The 16 April 2016, M-W 7.8 (M-S 7.5) Ecuador earthquake: A quasi-repeat of the 1942 M-S 7.5 earthquake and partial re-rupture of the 1906 M-S 8.6 Colombia-Ecuador earthquake. Earth Planet. Sci. Lett. 454: 248258, doi: 10.1016/j.epsl.2016.09.006.Google Scholar
Ye, L. L., Lay, T., Kanamori, H., and Rivera, L. 2016b. Rupture characteristics of major and great (M-w >= 7.0) megathrust earthquakes from 1990 to 2015: 1. Source parameter scaling relationships. J. Geophys. Res.-Solid Earth 121(2): 826844, doi: 10.1002/2015jb012426.Google Scholar
Ye, L., Lay, T., Kanamori, H., and Koper, K. D. 2013. Energy release of the 2013 Mw 8.3 Sea of Okhotsk earthquake and deep slab stress heterogeneity. Science 341(6152): 13801384.Google Scholar
Yeats, R. S. 1986. Faults related to folding with examples from New Zealand. R. Soc. N.2. Bull. 24: 273292.Google Scholar
Yokota, Y., and Koketsu, K. 2015. A very long-term transient event preceding the 2011 Tohoku earthquake. Nat. Comm. 6: 5934.Google Scholar
Yokota, Y., Ishikawa, T., Watanabe, S., Tashiro, T., and Asada, A. 2016. Seafloor geodetic constraints on interplate coupling of the Nankai Trough megathrust zone. Nature 534(7607): 374377, doi: 10.1038/nature17632.Google Scholar
Yonekura, N. 1975. Quaternary tectonic movements in the outer arc of southwest Japan with special reference to seismic crustal deformation. Bull. Dept. Geogr. Univ. Tokyo 7: 1971.Google Scholar
Yoshida, Y., and Abe, K. 1992. Source mechanism of the Luzon, Phillipines earthquake of July 16, 1990. Geophys. Res. Lett. 19: 545548.Google Scholar
Yoshii, T. 1979. A detailed cross-section of the deep seismic zone beneath northeastern Honshu, Japan. Tectonophysics 55: 349360.Google Scholar
Yoshioka, N. 1986. Fracture energy and the variation of gouge and surface roughness during frictional sliding of rocks. J. Phys. Earth 34: 335355.Google Scholar
Yoshioka, N., and Scholz, C. H. 1989a. Elastic properties of contacting surfaces under normal and shear loads. 1. Theory. J. Geophys. Res. 94: 1768117690.Google Scholar
Yoshioka, N., and Scholz, C. H. 1989b. Elastic properties of contacting surfaces under normal and shear loads. 2. Comparison of theory with experiment. J. Geophys. Res. 94: 1769117700.Google Scholar
Yuan, F., Prakash, V., and Tullis, T. 2011. Origin of pulverized rocks during earthquake fault rupture. J. Geophys. Res.-Solid Earth 116, doi: 10.1029/2010jb007721.Google Scholar
Yue, H., Lay, T., and Koper, K. D. 2012. En echelon and orthogonal fault ruptures of the 11 April 2012 great intraplate earthquakes. Nature 490(7419): 245-+, doi: 10.1038/nature11492.Google Scholar
Yue, H., Lay, T., Rivera, L., et al. 2014. Rupture process of the 2010 M-w 7.8 Mentawai tsunami earthquake from joint inversion of near-field hr-GPS and teleseismic body wave recordings constrained by tsunami observations. J. Geophys. Res.-Solid Earth 119(7): 55745593, doi: 10.1002/2014jb011082.Google Scholar
Zagrodzki, P. 2009. Thermoelastic instability in friction clutches and brakes – Transient modal analysis revealing mechanisms of excitation of unstable modes. Int. J. Solids and Structures 46: 24632476.Google Scholar
Zang, A., Wagner, F. C., Stanchits, S., Janssen, C., and Dresen, G. 2000. Fracture process zone in granite. J. Geophys. Res. 105: 23,65123,661.Google Scholar
Zen, E. 1974. Prehnite- and pumpellyite-bearing mineral assemblages, west side of the Appalachian metamorphic belt, Pennsylvania to Newfoundland. J. Petrology 15: 197242.Google Scholar
Zeng, Y. H., and Chen, C. H. 2001. Fault rupture process of the 20 September 1999 Chi-Chi, Taiwan, earthquake. Bull. Seismol. Soc. Am. 91(5): 10881098.Google Scholar
Zeng, Y., and Shen, Z.-K. 2014. Fault network modeling of crustal deformation in California constrained using GPS and geologic observations. Tectonophysics 612: 117.Google Scholar
Zhan, Z. W., Kanamori, H., Tsai, V. C., Helmberger, D. V., and Wei, S. J. 2014. Rupture complexity of the 1994 Bolivia and 2013 Sea of Okhotsk deep earthquakes. Earth Planet. Sci. Lett. 385: 8996, doi: 10.1016/j.epsl.2013.10.028.Google Scholar
Zhang, J., Hu, L., Pant, R., Yu, Y., Wei, Z., and Zhang, G. 2013. Effects of interlayer interactions on the nanoindentation behavior and hardness of 2:1 phyllosilicates. Applied Clay Science 80–81: 267280.Google Scholar
Zhao, L. Y., Zhu, Q. Z., Xu, W. Y., Dai, F., and Shao, J. F. 2016. A unified micromechanics-based damage model for instantaneous and time-dependent behaviors of brittle rocks. Int. J. Rock Mech. Min. Sci. 84: 187196, doi: 10.1016/j.ijrmms.2016.01.015.Google Scholar
Zhao, S., Wu, X., Hori, T., Kaneda, Y., and Takemoto, S. 2004. Crustal deformation and stress localization in Kanto-Tokai, central Japan revealed by GPS. Geophys. J. Int, 157(2): 737752, doi: 10.1111/j.1365-246X.2004.02227.x.Google Scholar
Ziv, A., and Rubin, A. M. 2000. Static stress transfer and earthquake triggering: No lower threshold in sight. J. Geophys. Res. 105: 1363113642.Google Scholar
Zoback, M. D., and Harjes, H.-P. 1997. Injection induced earthquakes and crustal stress at 9 km depth at the KTB deep drilling site, Germany. J. Geophys. Res. 102: 1847718491.Google Scholar
Zoback, M. D., and Healy, J. H. 1984. Friction, faulting, and insitu stress. Annales Geophysicae 2: 689698.Google Scholar
Zoback, M. D., and Healy, J. H. 1992. In situ stress measurements to 3.5 km depth in the Cajon Pass scientific research borehole: Implications for the mechanics of crustal faulting. J. Geophys. Res. 97: 50395057.Google Scholar
Zoback, M. D., and Townend, J. 2001. Implications of hydrostatic pore pressures and high crustal strength for the deformation of intraplate lithosphere. Tectonophysics 336(1–4): 1930, doi: 10.1016/s0040-1951(01)00091–9.Google Scholar
Zoback, M. D., Tsukahara, H., and Hickman, S. 1980. Stress measurements at depth in the vicinity of the San Andreas fault: Implications for the magnitude of shear stress at depth. J. Geophys. Res. 85: 61576173.Google Scholar
Zoback, M. D., Zoback, M. L., Mount, V. S., et al. 1987. New evidence on the state of stress of the San Andreas fault system. Science 238: 11051111.Google Scholar
Zoback, M. L. 1992. 1st-order and 2nd-order patterns of stress in the lithosphere – The world stress map project. J. Geophys. Res.-Solid Earth 97: 1170311728.Google Scholar
Zoback, M. L., and Zoback, M. D. 1980. State of stress in the conterminous United States. J. Geophys. Res. 85: 61136156.Google Scholar
Zoback, M. L., and Zoback, M. D. 1997. Crustal stress and intraplate deformation. Geowiss. 15: 116123.Google Scholar
Zoback, M. L., Jachens, R. C., and Olson, J. A. 1999. Abrupt along-strike change in tectonic style: San Andreas fault zone, San Francisco Peninsula. J. Geophys. Res.-Solid Earth 104: 1071910742.Google Scholar
Zöller, G., Hainzl, S., and Kurths, J. 2001. Observation of growing correlation length as an indicator for critical point behavior prior to large earthquakes. J. Geophys. Res. 106: 21672175.Google Scholar

Save book to Kindle

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

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

Find out more about the Kindle Personal Document Service.

Available formats
×

Save book to Dropbox

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

Available formats
×

Save book to Google Drive

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

Available formats
×