Hostname: page-component-848d4c4894-p2v8j Total loading time: 0 Render date: 2024-06-09T07:05:50.991Z Has data issue: false hasContentIssue false

Insights into the velocity-dependent geometry and internal strain in accretionary wedges from analogue models

Published online by Cambridge University Press:  25 January 2017

BIN DENG*
Affiliation:
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation/Chengdu University of Technology, Chengdu, China, 610059 Department of Geology, Trinity College, Dublin
LEI JIANG
Affiliation:
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation/Chengdu University of Technology, Chengdu, China, 610059
GAOPING ZHAO
Affiliation:
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation/Chengdu University of Technology, Chengdu, China, 610059
RUI HUANG
Affiliation:
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation/Chengdu University of Technology, Chengdu, China, 610059
YUANBO WANG
Affiliation:
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation/Chengdu University of Technology, Chengdu, China, 610059
SHUGEN LIU
Affiliation:
State Key Laboratory of Oil and Gas Reservoir Geology and Exploitation/Chengdu University of Technology, Chengdu, China, 610059
*
Author for correspondence: dengb@tcd.ie

Abstract

Although the brittle material in analogue models is characterized by a linear Navier-Coulomb behaviour and rate-independent deformation, the geometry and style of deformation in accretionary wedges is sensitive to shortening velocity. In this study we have constructed a series of analogue models with various shortening velocities in order to study the influence of shortening velocity on the geometry and kinematics of accretionary wedges. Model results illustrate how shortening velocity has an important influence on the geometry and kinematics of the resulting wedge. In general, for models having similar bulk shortening, the accretionary wedges with higher velocities of shortening are roughly steeper, higher and longer, as well as having larger critical wedge angles and height. It accommodates a number of foreland-vergent thrusts, larger fault spacing and displacement rates than those of low- to medium-velocity shortening, which indicates a weak velocity-dependence in geometry of the wedge. Moreover, models with a high velocity of shortening undergo larger amounts of volumetric strain and total layer-parallel shortening than models with low- to medium-velocity shortening. The former accommodate a greater development of back thrusts and asymmetric structures; a backwards-to-forwards style of wedge growth therefore occurs in the frontal zone under high-velocity shortening.

Type
Original Article
Copyright
Copyright © Cambridge University Press 2017 

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

Adam, J., Urai, J.L., Wieneke, B., Oncken, O., Pfeiffer, K., Kukowski, N., Lohrmann, J., Hoth, S., Van der zee, W. & Schmatzm, J. 2005. Shear localization and strain distribution during tectonic faulting-new insights from granular-flow experiments and high-resolution optical image correlation techniques. Journal of Structural Geology 27, 283301.Google Scholar
Ahmad, M.I., Dubey, A.K., Toscani, G., Bonini, L. & Seno, S. 2014. Kinematic evolution of thrusts wedge and erratic line length balancing: insights from deformed analogue models. International Journal of Earth Sciences 103, 329–47.Google Scholar
Asensio, E., Khazaradze, G., Echeverria, A., King, R.W. & Vilajosana, I. 2012. GPS studies of active deformation in the Pyrenees. Geophysics Journal International 190, 913–21.Google Scholar
Beaumont, C., Fullsack, P. & Hamilton, J. 1994. Styles of crustal deformation in compressional orogens caused by subduction of the underlying lithosphere. Tectonophysics 232, 119–32.CrossRefGoogle Scholar
Blanc, E.J.P., Allen, M.B., Inger, S. & Hassani, H. 2003. Structural styles in the Zagros Simple Folded Zone, Iran. Journal of the Geological Society, London 160, 401–12.Google Scholar
Bonini, M. 2001. Passive roof thrusting and forelandward fold propagation in scaled brittle-ductile physical models of thrust wedges. Journal of Geophysical Research 106 (B2), 2291–311.Google Scholar
Cloetingh, S., Ziegler, P.A. & Bogaard, P. 2007. Topo-Europe: The geoscience of coupled deep Earth-surface processes. Global and Planetary Change 2007, 58, 1118.CrossRefGoogle Scholar
Contardo, X.J., Kukowski, N. & Cembrano, J.M. 2011. Material transfer and its influence on the formation of slope basins along the South Central Chilean convergent margin: Insights from scaled analogue experiments. Tectonophysics 513, 2036.Google Scholar
Costa, E. & Vendeville, B.C. 2002. Experimental insights on the geometry and kinematics of fold-and-thrust belts above weak, viscous evaporitic decollecment. Journal of Structural Geology 24, 1729–39.CrossRefGoogle Scholar
Couzens-Schultz, B.A., Vendeville, B.C. & Wiltschko, D.V. 2003. Duplex style and triangle zone formation: insights from physical modeling. Journal of Structural Geology 25, 1623–44.Google Scholar
Cruz, L., Teyssier, C., Perg, L., Take, A. & Fayon, A. 2008. Deformation, exhumation, and topography of experimental doubly-vergent orogenic wedges subjected to asymmetric erosion. Journal of Structural Geology 30, 98115.Google Scholar
Cubas, N., Maillot, B. & Barnes, C. 2010. Statistical analysis of an experimental compressional sand wedge. Journal of Structural Geology 32, 818–31.Google Scholar
Dahlen, F. A. 1990. Critical taper model of fold-and-thrust belts and accretionary wedges. Annual Review of Earth and Planetary Sciences 18, 5599.Google Scholar
Davis, D. & Engelder, T. 1985. The role of salt in fold-and-thrust belts. Tectonophysics 119, 6788.Google Scholar
Davis, D., Suppe, J. & Dahlen, F. A. 1983. Mechanics of fold-and-thrust belts and accretionary wedges. Journal of Geophysical Research 88 (B12), 1153–72.Google Scholar
Davies, R.K. & Fletcher, R.C. 1990. Shear bands in a plastic layer at yield under combined shortening and shear: a model for the fault array in a duplex. In Deformation Mechanisms, Rheology and Tectonics (eds Knipe, R. J. & Rutter, E. H.), pp. 123–32. Geological Society of London, Special Publication no. 54.Google Scholar
Goodman, R.E. 1988. Introduction to Rock Mechanics. New York: Wiley.Google Scholar
Graveleau, F., Malavieille, J. & Dominguez, S. 2012. Experimental modelling of orogenic wedges: A review. Tectonophysics 538–40, 166.Google Scholar
Gutscher, M., Klaeschen, D., Flueh, E. & Malavieille, J. 2001. Non-Coulomb wedges, wrong-way thrusting, and natural hazards in Cascadia. Geology 29, 379–82.Google Scholar
Gutscher, M., Kukowski, N., Malavieille, J. & Lallemand, S. 1996. Cyclical behavior of thrust wedges: insights from high basal friction analogue experiments. Geology 24, 135–8.2.3.CO;2>CrossRefGoogle Scholar
Hatzfeld, D., Authemayou, C., Van der beek, P., Bellier, O., Lave, J., Oveisi, B., Tatar, M., Tavakoli, F., Walpersdorf, A. & Yamini-Fadr, F. 2010. The kinematics of the Zagros Mountains (Iran). In Tectonic and Stratigraphic Evolution of Zagros and Makran during the Mesozoic–Cenozoic (eds Leturmy, P. & Robin, C.), pp. 1942. Geological Society of London, Special Publication no. 330.Google Scholar
Hatzfeld, D. & Monlar, P. 2010. Comparisons of the kinematics and deep structures of the Zagros and Himalaya and of the Iranian and Tibetan Plateaus and Geodynamic implications. Reviews of Geophysics 48, RG2005, doi:10.1029/2009RG000304.Google Scholar
Hessami, K., Nilforoushan, F. & Talbot, C.J. 2006. Active deformation within the Zagros Mountains deduced from GPS measurements. Journal of the Geological Society, London 163, 143–48.CrossRefGoogle Scholar
Hoshino, K., Koide, H., Inami, K., Iwamura, S. & Mitsui, S. 1972. Mechanical properties of Japanese Tertiray sedimentary rocks under high confined pressure. Geological Survey of Japan, Report no. 244.Google Scholar
Hoth, S., Adam, J., Kukowski, N. & Oncken, O. 2006. Influence of erosion on the kinematics of bivergent orgens, Results from scaled analogue simulations. In Tectonics, Climate, and Landscape Evolution (eds Willett, S.D., Hovius, N., Brandon, M. T. & Fisher, D. M.), pp. 201–25. Geological Society of America, Special Paper no. 398.Google Scholar
Hoth, S., Hoffmann-Rothe, A. & Kukowski, N. 2007. Frontal accretion: An internal clock for bivergent wedge deformation and surface uplift. Journal of Geophysical Research 112, doi:10.1029/2006JB004357.Google Scholar
Huang, J.H., Wiltschko, D.V., Lin, H.C., Hichman, J.B., Fang, P. & Bock, Y. 1999. Structure and motion of the Southwestern Taiwan Fold and Thrust Belt. Tao 10, 543–68.Google Scholar
Koyi, H. A. 1988. Experimental modeling of the role of gravity and lateral shortening in the Zagros mountain belt. AAPG Bulletin 72, 1381–94.Google Scholar
Koyi, H. A. 1995. Mode of internal deformation in sand wedges. Journal of Structural Geology 17, 293300.Google Scholar
Koyi, H. A. & Vendeville, B. C. 2003. The effect of décollement dip on geometry and kinematics of model accretionary wedges. Journal of Structural Geology 25, 1445–50.Google Scholar
Koyi, H. A., Sans, M., Teixell, A., Cotton, J. & Zeyen, H. 2004. The significance of penetrative strain in the restoration of shortened layers-insights from sand models and the Spanish Pyrenees. In Thrust Tectonics and Hydrocarbon Systems (ed. McClay, K. R.), pp. 207–22. American Association of Petroleum Geologists, Memoir no. 82.Google Scholar
Liu, H., McClay, K. R. & Powell, D. 1992. Physical models of thrust wedges. In Thrust Tectonics (eds McClay, K. R.), pp. 7181. London: Chapman and Hall.Google Scholar
Lohrmann, J., Kukowski, N., Adam, J. & Oncken, O. 2003. The impact of analogue material properties on the geometry, kinematics, and dynamics of convergent sand wedges. Journal of Structural Geology 25, 1691–711.Google Scholar
Makel, G. & Walters, J. 1993. Finite-element analysis of thrust tectonics: Computer simulation of detachment phase and development of thrust faults. Tectonophysics 226, 167–85.Google Scholar
Marshak, S. & Wilkerson, M.S. 1992. Effect of overburden thickness on thrust belt geometry and development. Tectonics 11, 560–6.Google Scholar
McClay, K.R. 1990. Extensional fault systems in sedimentary basins. A review of analogue model studies. Marine and Petroleum Geology 7, 206–33.Google Scholar
McClay, K. R. & Whitehouse, P. S. 2004. Analog modeling of doubly vergent thrust wedges. In Thrust Tectonics and Hydrocarbon Systems (ed. McClay, K. R.), pp. 184206. American Association of Petroleum Geologists, Memoir no. 82.Google Scholar
McQuarrie, N. 2004. Crustal scale geometry of the Zagros fold-thrust belt, Iran. Journal of Structural Geology 26, 519–35.Google Scholar
Mouthereau, F., Deffontaines, B., Lacombe, O. & Angelier, J. 2002. Variations along the strike of the Taiwan thrust belt: Basement control on structural style, wedge geometry, and kinematics. In Geology and Geophysics of an Arc-Continent Collision, Taiwan, Republic of China (eds Byrne, T. B. & Liu, C. S.), pp. 3558. Geological Society of America, Special Paper no. 358.Google Scholar
Mouthereau, F. & Lacombe, O. 2006. Inversion of the Paleogene Chines continental margin and thick-skinned deformation in the Wstern Foreland of Taiwan. Journal of Structural Geology 28, 1977–93.Google Scholar
Mouthereau, F., Lacombe, O., Deffontaines, B., Angelier, J. & Brusset, S. 2001. Deformation history of the southwestern Taiwan foreland thrust belt: insights from tectono-sedimentary analyses and balanced cross-section. Tectonophysics 333, 293322.Google Scholar
Mulugeta, G. 1988. Squeeze-box in a centrifuge. Tectonophysics 148, 323–35.Google Scholar
Mulugeta, G. & Koyi, H. 1987. Three-dimensional geometry and kinematics of experimental piggyback thrusting. Geology 15, 1052–6.Google Scholar
Mulugeta, G. & Koyi, H. 1992. Episodic accretion and strain partitioning in a model sand wedge. Tectonophysics 202, 319–33.CrossRefGoogle Scholar
Nihei, K.T., Hilbert, L.B. Jr., Cook, N.G.W., Nakagawa, S. & Myer, L.R. 2000. Frictional effects on the volumetric strain of sandstone. International Journal of Rock Mechanics and Mining Sciences 37, 121–32.CrossRefGoogle Scholar
Nilforoushan, F., Koyi, H.A., Swantesson, J.O.H. & Talbot, C.J. 2008. Effect of basal friction on surface and volumetric strain in models of convergent settings measured by laser scanner. Journal of Structural Geology 30, 366–79.Google Scholar
Nilforoushan, F., Pysklywec, R. & Cruden, A. 2012. Sensitivity analysis of numerical scaled models of fold-and-thrust belts to granular material cohesion variation and comparison with analog experiments. Tectonophysics 526–9, 196206.Google Scholar
Panian, J. & Wiltschko, D. V. 2004. Ramp initiation in a thrust wedge. Nature 427, 624–7.Google Scholar
Panian, J. & Wiltschko, D. V. 2007. Ramp initiation and spacing in a homogeneous thrust wedge. Journal of Geophysical Research 112, B05417, doi: 10.1029/2004JB003596.Google Scholar
Paola, C., Staub, K., Mohrig, D. & Reinhardt, L. 2009. The “unreasonable effectiveness” of stratigraphic and geomorphic experiments. Earth-Science Reviews 7, 143.Google Scholar
Reiter, K., Kukowski, N. & Ratschbacher, L. 2011. The interaction of two indenters in analogue experiments and implications for curved fold-and-thrust belts. Earth and Planetary Science Letters 302, 132–46.Google Scholar
Rossetti, F., Faccenna, C. & Ranall, I. G. 2002. The influence of backstop dip and convergence velocity in the growth of viscous doubly-vergent orogenic wedges: insights from thermomechanical laboratory experiments. Journal of Structural Geology 24, 953–62.Google Scholar
Rossetti, F., Faccenna, C., Ranall, I. G. & Storti, F. 2000. Convergence rate-dependent growth of experimental viscous orogenic wedges. Earth and Planetary Science Letters 178, 367–72.Google Scholar
Ruh, J. B., Gerya, T. & Burg, J. 2014. 3D effects of strain vs. velocity weakening on deformation patterns in accretionary wedges. Tectonophysics 615–6, 122–41.Google Scholar
Santimano, T., Rosenau, M. & Oncken, O. 2015. Intrinsic versus extrinsic variability of analogue sand-box experiments: Insights from statistical analysis of repeated accretionary sand wedge experiments. Journal of Structural Geology 75, 80100.Google Scholar
Schellart, W. P. 2000. Shear test results for cohesion and friction coefficients for different granular materials: scaling implications for their usage in analogue modelling. Tectonophysics 324, 116.Google Scholar
Sepehr, M. & Cosgrove, J.W. 2004. Structural framework of the Zagros Fold-Thrust Belt, Iran. Marine and Petroleum Geology 21, 829–34.Google Scholar
Sherkati, S., Letouzey, J. & De Lamotte, F. 2006. Central Zagros fold-thrust belt (Iran): New insights from seismic data, field observation, and sandbox modeling. Tectonics 25, doi:10.1029/2004TC001766.Google Scholar
Smit, J. W., Brun, J. P. & Soukoutis, D. 2003. Deformation of brittle-ductile thrust wedges in experiments and nature. Journal of Geophysical Research 108 (B10), doi: 10.1029/2002JB002190.Google Scholar
Sonder, L.J. & England, P. 1986. Vertical averages of rheology of the continental lithosphere: relation to thin sheet parameters. Earth and Planetary Science Letters 77, 8190.Google Scholar
Souloumiac, P., Maillot, B. & Leroy, Y.M. 2012. Bias due to side wall friction in sand box experiments. Journal of Structural Geology 35, 90101.Google Scholar
Strayer, L. M., Hudleston, P. J. & Lorig, L. J. 2001. A numerical model of deformation and fluid-flow in an evolving thrust wedge. Tectonophysics 335, 121–45.Google Scholar
Storti, F. & McClay, K. 1995. Influence of syntectonic sedimentation on thrust wedges in analogue models. Geology 23, 9991002.Google Scholar
Storti, F., Salvini, F. & McClay, K. 2000. Synchronous and velocity-partitioned thrusting and thrust polarity reversal in experimentally produced, doubly-vergent thrust wedges: implications for natural orogens. Tectonics 19, 378–96.Google Scholar
Suppe, J. 2007. Absolute fault and crustal strength from wedge tapers. Geology 35, 1127–30.Google Scholar
Tavokoli, F., Walpersdorf, A., Authemayou, C., Nankali, H.R., Hatzfeld, D., Tatar, M., Djamour, Y., Nilforoushan, F. & Cotte, N. 2008. Distribution of the right-lateral strike-slip motion from the Main Recent Fault to the Kazerun Fault System (Zagros, Iran): Evidence from present-day GPS velocities. Earth and Planetary Science Letters 275 (3–4), 342–7, doi: 10.1016/j.epsl.2008.08.030.Google Scholar
Walpersdorf, A., Baize, S., Calais, E., Tregoning, P. & Nocquet, J. 2006. Deformation in the Jura Mountains (France): First results from semi-permanent GPS measurements. Earth and Planetary Science Letters 245, 365–72.Google Scholar
Wang, Q., Zhang, P.Z., Freymueller, J.T., Bilham, R., Larson, K.M., Lai, X.A., You, X.Z., Niu, Z.J., Wu, J.C., Li, Y.X., Liu, J.N., Yang, Z.Q. & Chen, Q.Z. 2001. Present-day crustal deformation in China constrained by Global Positioning System Measurements. Science 294, 574–7.Google Scholar
Willett, S. D. 1999. Orogeny and orography: the effects of erosion on the structure of mountain belts. Journal of Geophysical Research 104 (B12), 28957–82.Google Scholar
Yu, S.B., Chen, H.Y. & Kuo, L.C. 1997. Velocity field of GPS stations in the Taiwan area. Tectonophysics 274, 4159.Google Scholar
Zar, J.H. 2010. Biostatistical Analysis, fifth edition. New Jersey: Pearson Prentice-Hall.Google Scholar