Skip to main content Accessibility help
×
Hostname: page-component-8448b6f56d-xtgtn Total loading time: 0 Render date: 2024-04-19T22:15:03.424Z Has data issue: false hasContentIssue false

6 - Parallel finite element modeling of multi-timescale faulting and lithospheric deformation in western USA

from Part II - Modeling software and community codes

Published online by Cambridge University Press:  25 October 2011

Mian Liu
Affiliation:
University of Missouri
Youqing Yang
Affiliation:
University of Missouri
Qingsong Li
Affiliation:
Lunar and Planetary Institute
Gang Luo
Affiliation:
University of Missouri
Huai Zhang
Affiliation:
Graduate University of the Chinese Academy of Sciences
G. Randy Keller
Affiliation:
University of Oklahoma
Chaitanya Baru
Affiliation:
University of California, San Diego
Get access

Summary

Introduction

The lithosphere, which includes the crust and the uppermost part of the mantle, is the mechanically strong outer shell of the Earth. The lithosphere is broken into a dozen or so pieces, called the tectonic plates. Constant motions between these tectonic plates cause lithospheric deformation, which leads to mountain building, earthquakes, and many other geological processes, including geohazards. Hence understanding lithospheric deformation has been a major focus of modern geosciences research.

Most studies of lithospheric dynamics are based on continents, where the lithosphere is relatively soft and hence deforms strongly, and where direct observation is relatively easy. One major observational technique developed in the past decades is space-based geodesy (e.g., Gordon and Stein, 1992; Minster and Jordan, 1987). The Global Positioning System (GPS), the Interferometric Synthetic Aperture Radar (InSAR), and other space-based geodetic measurements are providing unprecedented details of deformation at the Earth's surface, thus revolutionizing the studies of lithospheric dynamics.

As space geodetic data grew rapidly in recent years, it has become increasingly clear that interpreting the tectonic implications of these data may not be straightforward. In many places, space-geodetic measurements of crustal deformation differ significantly from that reflected in geological records (Dixon et al., 2003; Friedrich et al., 2003; He et al., 2003; Liu et al., 2000; Pollitz, 2003; Shen et al., 1999). This could be due to many factors; an important one is timescale-dependent lithospheric rheology and deformation (Jeanloz and Morris, 1986; Liu et al., 2000; Pollitz, 1997, 2003; Ranalli, 1995).

Type
Chapter
Information
Geoinformatics
Cyberinfrastructure for the Solid Earth Sciences
, pp. 68 - 94
Publisher: Cambridge University Press
Print publication year: 2011

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

Ainsworth, M. (1996). A preconditioner based on domain decomposition for h-p finite-element approximation on quasi-uniform meshes. SIAM Journal on Numerical Analysis, 33: 1358–1376.CrossRefGoogle Scholar
Allmendinger, R. W., Jordan, T. E., Kay, S. M., and Isacks, B. L. (1997). The evolution of the Altiplano-Puna plateau of the Central Andes. Annual Reviews of Earth and Planetary Science, 25: 139–174.CrossRefGoogle Scholar
Anderson, G., Aagaard, B., and Hudnut, K. (2003). Fault interactions and large complex earthquakes in the Los Angeles area. Science, 302: 1946–1949.CrossRefGoogle ScholarPubMed
Argus, D. F. and Gordon, R. G. (1991). Current Sierra Nevada-North America motion from very long baseline interferometry: Implications for the kinematics of the Western United States. Geology (Boulder), 19: 1085–1088.2.3.CO;2>CrossRefGoogle Scholar
Armstrong, R. L. and Ward, P. (1991). Evolving geographic patterns of Cenozoic magmatism in the North American Cordillera: The temporal and spatial association of magmatism and metamorphic core complexes. Journal of Geophysical Research, 96: 13 201–13 224.CrossRefGoogle Scholar
Atwater, T. (1970). Implications of plate tectonics for the Cenozoic tectonic evolution of western North America. Geolical Society of America Bulletin, 81: 3513–3536.CrossRefGoogle Scholar
Atwater, T. and Stock, J. (1998). Pacific North America plate tectonics of the Neogene southwestern United States: An update. International Geology Review, 40: 375–402.CrossRefGoogle Scholar
Axelsson, O. and Larin, M. (1998). An algebraic multilevel iteration method for finite element matrices. Journal of Computational and Applied Mathematics, 89: 135–153.CrossRefGoogle Scholar
Axen, G. J., Taylor, W., and Bartley, J. M. (1993). Space-time patterns and tectonic controls of Tertiary extension and magmatism in the Great Basin of the western United States. Geological Society of America Bulletin, 105: 56–72.2.3.CO;2>CrossRefGoogle Scholar
Becker, T. W., Hardebeck, J. L., and Anderson, G. (2005). Constraints on fault slip rates of the southern California plate boundary from GPS velocity and stress inversions. Geophysical Journal International, 160: 634–650.CrossRefGoogle Scholar
Bennett, R. A., Davis, J. L., Normandeau, J. E., and Wernicke, B. P. (2002). Space Geodetic Measurements of Plate Boundary Deformation in the Western U.S. Cordillera: Plate Boundary Zones. Vol. 30. Washington, D.C.: American Geophysical Union, pp. 27–55.Google Scholar
Bennett, R. A., Davis, J. L., and Wernicke, B. P. (1999). The present-day pattern of Werstern U.S. Cordillera deformation. Geology, 27: 371–374.2.3.CO;2>CrossRefGoogle Scholar
Bennett, R. A., Wernicke, B. P., Niemi, N. A., Friedrich, A. M., and Davis, J. L. (2003). Contemporary strain rates in the northern Basin and Range Province from GPS data. Tectonics, 22: doi:10.1029/2001TC001355.CrossRefGoogle Scholar
Benz, H. M., Smith, R. B., and Mooney, W. D. (1990). Crustal structure of the northwestern Basin and Range province from the 1986 PASSCAL seismic experiment. Journal of Geophysical Research, 95: 21 823–21 842.CrossRefGoogle Scholar
Bevis, M., Kendrick, E., Smalley, R.et al. (2001). On the strength of interplate coupling and the rate of back arc convergence in the central Andes: An analysis of the interseismic velocity field. Geochemistry Geophysics Geosystems, 2: doi:10.129/2001GC000198.CrossRefGoogle Scholar
Bhat, H. S., Dmowska, R., King, G. C. P., Klinger, Y., and Rice, J. R. (2007). Off-fault damage patterns due to supershear ruptures with application to the 2001 Mw 8.1 Kokoxili (Kunlun) Tibet earthquake. Journal of Geophysical Research, 112: doi:10.1029/2006JB004425.CrossRefGoogle Scholar
Bird, P. and Kong, X. (1994). Computer simulations of California tectonics confirm very low strength of major faults. GSA Bulletin, 106: 159–174.2.3.CO;2>CrossRefGoogle Scholar
Bird, P. and Piper, K. (1980). Plane-stress finite. Physics of the Earth Planetary Interiors, 21: 158–175.CrossRefGoogle Scholar
Bokelmann, G. H. R. (2002). Convection-driven motion of the North American Craton: Evidence from P-wave anisotropy. Geophysical Journal International, 148: 278–287.Google Scholar
Brace, W. F. and Kohlstedt, D. L. (1980). Limits on lithospheric stress imposed by laboratory experiments. Journal of Geophysical Research, 85: 6248–6252.CrossRefGoogle Scholar
Burchfiel, B. C., Cowan, D. S., and Davis, G. A. (1992). Tectonic overview of the Cordilleran orogen in western United States. In The Cordilleran Orogen: Conterminous U.S., ed. Burchfiel, B. C., Zoback, M. L. and Lipman, P. W.. Volume G-3. Boulder, CO: Geological Society of America, pp. 407–480.Google Scholar
Burchfiel, C. B. and Davis, G. A. (1975). Nature and controls of cordilleran orogenesis, western United States: Extensions of an earlier synthesis. American Journal of Science, 275-A: 363–396.Google Scholar
Chlieh, M., Chabalier, J. B., Ruegg, J. C. et al. (2004). Crustal deformation and fault slip during the seismic cycle in the North Chile subduction zone, from GPS and InSAR observations. Geophysical Journal International, 158: 695–711.CrossRefGoogle Scholar
Choi, E. S. and Gurnis, M. (2003). Deformation in transcurrent and extensional environments with widely spaced weak zones. Geophysical Research Letters, 30: doi:10.1029/2002GL016129.CrossRefGoogle Scholar
Coney, P. J. (1978). Mesozoic-Cenozoic Cordilleran plate tectonics:Memoirs of the Geological Society of America, 152: 33–50.CrossRefGoogle Scholar
Coney, P. J. and Harms, T. A. (1984). Cordilleran metamorphic core complexes: Cenozoic extensional relics of Mesozoic compression. Geology, 12: 550–554.2.0.CO;2>CrossRefGoogle Scholar
Cooke, M. L. and Kameda, A. (2002). Mechanical fault interaction within the Los Angeles Basin: A two-dimensional analysis using mechanical efficiency. Journal of Geophysical Research – Solid Earth, 107.CrossRefGoogle Scholar
Davis, J. L., Wernicke, B. P., Bisnath, S., Niemi, N. A., and Elosegui, P. (2006). Subcontinental-scale crustal velocity changes along the Pacific-North America plate boundary. Nature, 441: 1131–1134.CrossRefGoogle ScholarPubMed
Dickinson, W. R. (2002). The basin and range province as a composite extensional domain. International Geology Review, 44: 1–38.CrossRefGoogle Scholar
Dieterich, J. (1979). Modeling of rock friction: 1. Experimental results and constitutive equations. Journal of Geophysical Research, 84.CrossRefGoogle Scholar
Dieterich, J. (1994). A constitutive law for rate of earthquake production and its application to earthquake clustering. Journal of Geophysical Research, 99: 2601–2618.CrossRefGoogle Scholar
Dixon, T. H., Miller, M., Farina, F., Wang, H., and Johnson, D. (2000). Present-day motion of the Sierra Nevada block and some tectonic implications for the Basin and Range Province, North American Cordillera. Tectonics, 19: 1–24.CrossRefGoogle Scholar
Dixon, T. H., Norabuena, E., and Hotaling, L. (2003) Paleoseismology and Global Positioning System: Earthquake-cycle effects and geodetic versus geologic fault slip rates in the Eastern California shear zone. Geology (Boulder), 31: 55–58.2.0.CO;2>CrossRefGoogle Scholar
Dixon, T. H., Robaudo, S., Lee, J., and Reheis, M. C. (1995). Constraints on present-day Basin and Range deformation from space geodesy. Tectonics, 14: 755–772.CrossRefGoogle Scholar
Dokka, R. K. and Travis, C. J. (1990). Role of the eastern California shear zone in accommodating Pacific-North American plate motion. Geophysical Research Letters, 17: 1323–1326.CrossRefGoogle Scholar
Dragert, H., Wang, K., and James, T. S. (2001). A silent slip event on the deeper Cascadia subduction interface. Science, 292: 1525–1528.CrossRefGoogle ScholarPubMed
Du, Y. and Aydin, A. (1996). Is the San Andreas big bend responsible for the Landers earthquake and the Eastern California shear zone?Geology, 24: 219–222.2.3.CO;2>CrossRefGoogle Scholar
Duan, B. and Oglesby, D. D. (2006). Heterogeneous fault stresses from previous earthquakes and the effect on dynamics of parallel strike-slip faults. Journal of Geophysical Research, 111: doi:10.1029/2005JB004138.CrossRefGoogle Scholar
England, P. and Molnar, P. (1997). Active deformation of Asia: From kinematics to dynamics. Science, 278: 647–650.CrossRefGoogle Scholar
England, P. C. and McKenzie, D. P. (1982). A thin viscous sheet model for continental deformation. Geophysical Journal, Royal Astronomical Society, 70: 295–321.CrossRefGoogle Scholar
Flesch, L. M., Holt, W. E., Haines, A. J., and Shen, T. B. (2000). Dynamics of the Pacific-North American plate boundary in the Western United States. Science, 287: 834–836.CrossRefGoogle ScholarPubMed
Flueh, E. R., Fisher, M. A., Bialas, J.et al. (1998). New seismic images of the cascadia subduction zone from cruise SO 108-ORWELL. Tectonophysics, 293: 69–84.CrossRefGoogle Scholar
Freed, A. M., Ali, S. T., and Burgmann, R. (2007). Evolution of stress in southern California for the past 200 years from coseismic, postseismic, and interseismic processes. Geophysical Journal International, 169: 1164–1179.CrossRefGoogle Scholar
Friedrich, A. M., Wernicke, B. P., Niemi, N. A., Bennett, R. A., and Davis, J. L. (2003). Comparison of geodetic and geological data from the Wasatch region, Utah, and implications for the spectral characteristics of Earth deformation at periods of 10 to 10 million years. Journal of Geophysical Research, 108: doi:10.1029/2001JB000682.CrossRefGoogle Scholar
Gabrielse, H. and Yorath, C. J. (1992). Tectonic Synthesis: Geology of the Cordilleran Orogen in Canada, ed. Gabrielse, H. and Yorath, C. J.. Vol. G-2. Ottawa: Geological Survey of Canada, pp. 679–705;.Google Scholar
Gan, W. J., Svarc, J. L., Savage, J. C., and Prescott, W. H. (2000). Strain accumulation across the Eastern California Shear Zone at latitude 36 degrees 30′ N. Journal of Geophysical Research, 105: 16 229–16 236.CrossRefGoogle Scholar
Goodman, R. E., Taylor, R. L., and Brekke, T. L. (1968). A Model for the Mechanics of Jointed Rock. Vol. 94. New York: American Society of Civil Engineers, pp. 637–659.Google Scholar
Gordon, R. G. and Stein, S. (1992). Global tectonics and space geodesy. Science, 256: 333–342.CrossRefGoogle ScholarPubMed
Hamilton, W. and Myers, W. B. (1966). Cenozoic tectonics of the western United States. Reviews of Geophysics, 5: 509–549.CrossRefGoogle Scholar
He, J., Liu, M., and Li, Y. (2003). Is the Shanxi rift of northern China extending?Geophysical Research Letters, 30: doi:10.1029/2003GL018764.CrossRefGoogle Scholar
Hu, Y., Wang, K., He, J., Klotz, J., and Khazaradze, G. (2004). Three-dimensional viscoelastic finite element model for postseismic deformation of the great 1960 Chile earthquake. Journal of Geophysical Research – Solid Earth, 109: doi:10.1029/2004JB003163.CrossRefGoogle Scholar
Isacks, B. L. (1988). Uplift of the central Andean plateau and bending of the Bolivian orocline. Journal of Geophysical Research, 93: 3211–3231.CrossRefGoogle Scholar
Jaeger, J. C., Cook, N. G. W., and Zimmerman, R. W. (2007). Fundamentals of Rock Mechanics. Oxford: Blackwell Publishing, pp. 261–262.Google Scholar
Jeanloz, R. and Morris, S. (1986). Temperature distribution in the crust and mantle. Annual Reviews of Earth and Planetary Science, 14: 377–415.CrossRefGoogle Scholar
Jones, C. H., Unruh, J. R., and Sonder, L. J. (1996). The role of gravitational potential energy in active deformation in the southwestern United States. Nature, 381: 37–41.CrossRefGoogle Scholar
Kirby, S. H. and Kronenberg, A. K. (1987). Rheology of the lithosphere: Selected topics. Reviews of Geophysics, 25: 1219–1244.CrossRefGoogle Scholar
Kong, X. and Bird, P. (1996). Neotectonics of Asia: Thin-shell finite-element models with faults. In The Tectonic Evolution of Asia: World and Regional Geology Series, ed. Yin, A. and Harrison, T. M.. Cambridge: Cambridge University Press, pp. 18–36.Google Scholar
Lamb, S. (2006). Shear stresses on megathrusts: Implications for mountain building behind subduction zones. Journal of Geophysical Research, 111: doi:10.1029/2005JB003916.CrossRefGoogle Scholar
Lavier, L. L. and Buck, W. R. (2002) Half graben versus large-offset low-angle normal fault: Importance of keeping cool during normal faulting. Journal of Geophysical Research, 107: doi:10.1029/2001JB000513.CrossRefGoogle Scholar
Lewis, J. C., Unruh, J. R., and Twiss, R. J. (2003). Seismogenic strain and motion of the Oregon coast block. Geology, 31: 183–186.2.0.CO;2>CrossRefGoogle Scholar
Li, Q. and Liu, M. (2006). Geometrical impact of the San Andreas Fault on stress and seismicity in California. Geophysical Research Letters, 33, L08302, doi:10.1029/2005GL025661.CrossRefGoogle Scholar
Li, Q. and Liu, M. (2007). Initiation of the San Jacinto Fault and its interaction with the San Andreas Fault. Pure and Applied Geophysics, 164: 1937–1945.CrossRefGoogle Scholar
Li, Q., Liu, M., Qie, Z., and Sandvol Eric, A. (2006). Stress evolution and seismicity in the central-eastern United States: Insights from geodynamic modeling. In Continental Intraplate Earthquakes: Science, Hazard, and Policy Issues, ed. Stein, S. and Mazzotti, S.. Special Paper 425. Boulder, CO: Geological Society of America, pp. 149–166, doi:110.1130/2007.2425(11).Google Scholar
Li, Q., Liu, M., and Zhang, H. (2009). A 3D visco-elasto-plastic model for simulating long-term slip on non-planar faults. Geophysical Journal International, 176: 293–306.CrossRefGoogle Scholar
Lipman, P. W., Prostka, W. J., and Christiansen, R. L. (1972). Cenozoic volcanism and plate tectonic evolution of the western United States: Part I. Early and Middle Cenozoic. Royal Society of London Philosophical Transactions, 271-A: 217–248.CrossRefGoogle Scholar
Liu, M. (2001). Cenozoic extension and magmatism in the North America Cordillera: The role of gravitational collapse. Tectonophysics, 342: 407–433.CrossRefGoogle Scholar
Liu, M. and Shen, Y. (1998). Crustal collapse, mantle upwelling, and Cenozoic extension in the North American Cordillera. Tectonics, 17: 311–321.CrossRefGoogle Scholar
Liu, M. and Yang, Y. (2003). Extensional collapse of the Tibetan Plateau: Results of three-dimensional finite element modeling. Journal of Geophysical Research – Solid Earth, 108: 2361, doi:10.1029/2002JB002248.CrossRefGoogle Scholar
Liu, M., Wang, H., and Li, Q. (2010). Inception of the Eastern California Shear Zone and evolution of the San Andreas Fault plate boundary zone: From kinematics to geodynamics. Journal of Geophysical Research, 115, B07401, doi:10.1029/2009JB007055.Google Scholar
Liu, M., Yang, Y., Li, Q., and Zhang, H. (2007). Parallel computing of multi-scale continental deformation in the western United States: Preliminary results. Physics of the Earth and Planetary Interiors, 163: 35–51.CrossRefGoogle Scholar
Liu, M., Yang, Y., Stein, S., Zhu, Y., and Engeln, J. (2000). Crustal shortening in the Andes: Why do GPS rates differ from geological rates?Geophysical Research Letters, 27: 3005–3008.CrossRefGoogle Scholar
Liu, Z. and Bird, P. (2002). North America Plate is driven westward by lower mantle flow. Geophysical Research Letters, 29: doi:10.1029/2002GL016002.CrossRefGoogle Scholar
Luo, G. and Liu, M. (2009). Why short-term crustal shortening leads to mountain building in the Andes, but not in the Cascadia?Geophysical Research Letters, 36: L08301, doi:10.1029/2009GL037347.CrossRefGoogle Scholar
Luo, G. and Liu, M. (2010). Stress evolution and fault interactions before and after the 2008 Great Wenchuan Earthquake. Tectonophysics, 491, 127–140, doi:110.1016/j.tecto.2009.1012.1019.CrossRefGoogle Scholar
Lv, X., Zhao, Y., Huang, X. Y., Xia, G. H., and Wang, Z. J. (2006). An efficient parallel/unstructured-multigrid preconditioned implicit method for simulating 3D unsteady compressible flows with moving objects. Journal of Computational Physics, 215: 661–690.CrossRefGoogle Scholar
Mazzotti, S., Dragert, H., Henton, J.et al. (2003). Current tectonics of northern Cascadia from a decade of GPS measurements. Journal of Geophysical Research, 108: doi:10.1029/2003JB002653.CrossRefGoogle Scholar
McCaffrey, R., Qamar, A. I., King, R. W.et al. (2007). Fault locking, block rotation and crustal deformation in the Pacific Northwest. Geophysical Journal International, 169: 1315–1340.CrossRefGoogle Scholar
McClusky, S. C., Bjornstad, S. C., Hager, B. H.et al. (2001). Present day kinematics of the Eastern California Shear Zone from a geodetically constrained block model. Geophysical Research Letters, 28: 3369–3372.CrossRefGoogle Scholar
McQuarrie, N. and Wernicke, B. P. (2005). An animated tectonic reconstruction of southwestern North America since 36 Ma. Geosphere, 1: 147–172.CrossRefGoogle Scholar
Meade, B. J. (2005). Block models of crustal motion in southern California constrained by GPS measurements. Journal of Geophysical Research, 110, B03403, doi:10.1029/2004JB003209.CrossRefGoogle Scholar
Meade, B. J. and Hager, B. H. (2004). Viscoelastic deformation for a clustered earthquake cycle. Geophysical Research Letters, 31, L10610, doi:10.1029/2004GL019643.CrossRefGoogle Scholar
Miller, M. M., Johnson, D. J., Dixon, T. H., and Dokka, R. K. (2001a). Refined kinematics of the Eastern California shear zone from GPS observations, 1993–1998. Journal of Geophysical Research, 106: 2245–2263.CrossRefGoogle Scholar
Miller, M. M., Johnson, D. J., Rubin, C. M.et al. (2001b). GPS-determination of along-strike variation in Cascadia margin kinematics; implications for relative plate motion, subduction zone coupling, and permanent deformation. Tectonics, 20: 161–176.CrossRefGoogle Scholar
Minster, J. B. and Jordan, T. H. (1987). Vector constraints on Western U.S. deformation from space geodesy, neotectonics, and plate motions. Journal of Geophysical Research – Solid Earth and Planets, 92: 4798–4804.CrossRefGoogle Scholar
Murray, M. H. and Lisowski, M. (2000). Strain accumulation along the Cascadia subduction zone. Geophysical Research Letters, 27: 3631–3634.CrossRefGoogle Scholar
Okada, Y. (1985). Surface deformation due to shear and tensile faults in a half-space. Bulletin of the Seismological Society of America, 75: 1135–1154.Google Scholar
Parsons, T. (2002). Post-1906 stress recovery of the San Andreas fault system calculated from three-dimensional finite element analysis. Journal of Geophysical Research – Solid Earth, 107.CrossRefGoogle Scholar
Pollitz, F. F. (1997). Gravitational viscoelastic postseismic relaxation on a layered spherical Earth. Journal of Geophysical Research – Solid Earth and Planets, 102: 17,921–17 941.CrossRefGoogle Scholar
Pollitz, F. F. (2003). The relationship between the instantaneous velocity field and the rate of moment release in the lithosphere. Geophysical Journal International, 153: 596–608.CrossRefGoogle Scholar
Pollitz, F. F., Banerjee, P., Burgmann, R., Hashimoto, M., and Choosakul, N. (2006). Stress changes along the Sunda trench following the 26 December 2005 Sumatra-Andaman and 28 March 2005 Nias earthquake. Geophysical Research Letters, 33: doi:10.1029/2005gl024558.CrossRefGoogle Scholar
Pope, D. C. and Willett, S. D. (1998). Thermal-mechanical model for crustal thickening in the central Andes driven by ablative subduction. Geology, 26: 511–514.2.3.CO;2>CrossRefGoogle Scholar
Quarteroni, A. and Valli, A. (1999). Domain Decomposition Methods for Partial Differential Equations. Oxford, New York: The Clarendon Press (Oxford University Press), 376pp.Google Scholar
Ranalli, G. (1995). Rheology of the Earth. London: Chapman & Hall, 413pp.Google Scholar
Savage, J. C. (1983). A dislocation model of strain accumulation and release at a subduction zone. Journal of Geophysical Research, 88: 4984–4996.CrossRefGoogle Scholar
Savage, J. C. and Burford, R. (1973). Geodetic determination of relative plate motion in central California. Journal of Geophysical Research, 78: 832–845.CrossRefGoogle Scholar
Savage, J. C., Gan, W. J., and Svarc, J. L. (2001). Strain accumulation and rotation in the Eastern California Shear Zone. Journal of Geophysical Research, 106: 21 995–22 007.CrossRefGoogle Scholar
Shen, T. B., Holt, W. E., and Haines, A. J. (1999). Deformation kinematics in the Western United States determined from Quaternary fault slip rates and recent geodetic data. Journal of Geophysical Research – Solid Earth and Planets, 104: 28 927–28 956.CrossRefGoogle Scholar
Silver, P. G. and Holt, W. E. (2002). The mantle flow field beneath western North America. Science, 295: 1054–1058.CrossRefGoogle ScholarPubMed
Sobolev, S. V. and Babeyko, A. Y. (2005). What drives orogeny in the Andes?Geology, 33: 617–620.CrossRefGoogle Scholar
Sonder, L. J. and Jones, C. H. (1999). Western United States extension: How the west was widened. Annual Reviews of Earth and Planetary Science, 27: 417–462.CrossRefGoogle 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. Geophysical Journal International, 128: 594–604.CrossRefGoogle Scholar
Stein, S., Tomasello, J., and Newman, A. (2003). Should Memphis build for California's earthquakes?Eos Transactions AGU, 84: 177, 184–185.CrossRefGoogle Scholar
Stewart, J. H. (1978). Basin and Range structure in western North America: A review. In Cenozoic Tectonics and Regional Geophysics of the Western Cordillera, ed. Smith, R. B and Eaton, G. L.. Vol. 152. Boulder, CO: Memoirs of the Geological Society of America, pp. 1–31.Google Scholar
Thatcher, W., Foulger, G. R., Julian, B. R.et al. (1999). Present-day deformation across the Basin and Range Province, Western United States. Science, 283: 1714–1718.CrossRefGoogle ScholarPubMed
Thompson, G. A., Catchings, R., Goodwin, E.et al. (1989). Geophysics of the western Basin and Range province. In Geophysical Framework of the Continental United States, ed. Pakiser, L. C. and Mooney, W. D.. Vol. 172. Boulder, CO: Memoirs of the Geological Society of America, pp. 177–203.CrossRefGoogle Scholar
Wang, K. and He, J. (1999). Mechanics of low-stress forearcs: Nankai and Cascadia. Journal of Geophysical Research, 104: 15 191–15 205.CrossRefGoogle Scholar
Wang, K., Mulder, T., Rogers, G. C., and Hyndman, R. D. (1995). Case for very-low coupling stress on the Cascadia Subduction Fault. Journal of Geophysical Research, 100: 12,907–12 918.CrossRefGoogle Scholar
Wang, K., Wells, R., Mazzotti, S., Hyndman, R. D., and Sagiya, T. (2003). A revised dislocation model of interseismic deformation of the Cascadia subduction zone. Journal of Geophysical Research – Solid Earth, 108: doi:10.1029/2001JB001227.Google Scholar
Ward, S. N. (1990). Pacific-North America plate motions: New results from very long baseline interferometry. Journal of Geophysical Research – Solid Earth and Planets, 95: 21 965–21 981.CrossRefGoogle Scholar
Wdowinski, S. and Bock, Y. (1994). The evolution of deformation and topography of high elevated plateaus: 2. Application to the Central Andes. Journal of Geophysical Research, 99: 7121–7130.CrossRefGoogle Scholar
Weldon, R., Scharer, K., Fumal, T., and Biasi, G. (2004). Wrightwood and the earthquake cycle: What a long recurrence record tells us about how faults work. GSA Today, 14: 4–10.2.0.CO;2>CrossRefGoogle Scholar
Wells, R. E. and Simpson, R. W. (2001). Northward migration of the Cascadia forearc in the northwestern US and implications for subduction deformation. Earth Planets and Space, 53: 275–283.CrossRefGoogle Scholar
Wells, R. E., Weaver, C. S., and Blakely, R. J. (1998). Fore-arc migration in Cascadia and its neotectonic significance. Geology (Boulder), 26: 759–762.2.3.CO;2>CrossRefGoogle Scholar
Wernicke, B., Christiansen, R. L., England, P. C., and Sonder, L. J. (1987). Tectonomagmatic evolution of Cenozoic extension in the North American Cordillera. In Continental Extensional Tectonics, ed. Coward, M. P., Dewey, J. F. and Hancock, P. L.. Vol. 28, Geological Society of London, Special Publication, pp. 203–221.Google Scholar
Wernicke, B., Clayton, R., Ducea, M.et al. (1996). Origin of high mountains in the continents: The southern Sierra Nevada. Science, 271: 190–193.CrossRefGoogle Scholar
Wernicke, B., Friedrich, A. M., Niemi, N. A., Bennett, R. A., and Davis, J. L. (2000). Dynamics of plate boundary fault systems from Basin and Range Geodetic Network (BARGEN) and geological data. GSA Today, 10: 1–7.Google Scholar
Wernicke, B. P. (1992). Cenozoic extensional tectonics of the U.S. Cordillera. In The Cordilleran Orogen: Conterminous U.S., ed. Burchfiel, B. C., Zoback, M. L. and Lipman, P. W.. Vol. G-3. Boulder, CO: Geological Societyof America, pp. 553–581.Google Scholar
Xing, H. L., Mora, P., and Makinouchi, A. (2004). Finite element analysis of fault bend influence on stick-slip instability along an intra-plate fault. Pure and Applied Geophysics, 161: 2091–2102.CrossRefGoogle Scholar
Yang, Y. and Liu, M. (2009). Algorithms for optimizing rheology and loading forces in finite element models of lithospheric deformation. In Advances in Geocomputing, ed. Xing, H.. Lecture Notes in Earth Sciences, Vol. 119. Berlin: Springer-Verlag, pp. 119–138.CrossRefGoogle Scholar
Yang, Y. and Liu, M. (2010). What drives short- and long-term crustal deformation in the southwestern United States?Geophysical Research Letters, 37: doi:10.1029/2009GL041598.CrossRefGoogle Scholar
Yang, Y., Liu, M., and Stein, S. (2003). A 3D geodynamic model of lateral crustal flow during Andean mountain building. Geophysical Research Letters, 30: doi:10.1029/2003GL18.CrossRefGoogle Scholar
Zhang, H., Liu, M., Shi, Y.et al. (2007). Toward an automated parallel computing environment for geosciences. Physics of the Earth and Planetary Interiors, 163: 2–22.CrossRefGoogle Scholar
Zoback, M. D., Anderson, R. E., and Thompson, G. A. (1981). Cenozoic evolution of the state of stress and style of tectonism of the Basin and Ranges Province of the western United States. Royal Societyof London Philosophical Transactions, Serial A 300: 407–434.
Zoback, M. D. and Zoback, M. L. (1981). State of stress and intraplate earthquakes in the United States. Science, 213: 96–104.CrossRefGoogle ScholarPubMed

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
×