Hostname: page-component-8448b6f56d-sxzjt Total loading time: 0 Render date: 2024-04-24T14:48:10.767Z Has data issue: false hasContentIssue false

Kinematic analysis of rock flow and deformation temperature of the Sirjan thrust sheet, Zagros Orogen, Iran

Published online by Cambridge University Press:  04 February 2016

KHALIL SARKARINEJAD
Affiliation:
Department of Earth Sciences, College of Sciences, Shiraz University, Shiraz, Iran
SAEEDE KESHAVARZ*
Affiliation:
Department of Earth Sciences, College of Sciences, Shiraz University, Shiraz, Iran
ALI FAGHIH
Affiliation:
Department of Earth Sciences, College of Sciences, Shiraz University, Shiraz, Iran
BABAK SAMANI
Affiliation:
Faculty of Earth Sciences, Shahid Chamran University, Ahvaz, Iran
*
Author for correspondence: s.keshavarz2007@yahoo.com

Abstract

Microstructural, finite strain and vorticity analyses of quartz-rich mylonites were used in order to investigate kinematics of rock flow and deformation temperature in the Sirjan thrust sheet exposed in a structural window within the Sanandaj–Sirjan High Pressure – Low Temperature (HP–LT) metamorphic belt that forms part of the hinterland of the Zagros orogenic belt of Iran. A dominant top-to-the-SW sense of shear in the study area is indicated by several shear sense indicators such as asymmetric boudins, rotated porphyroclasts, mica fish and S/C fabrics. Quantitative analyses reveal approximately plane strain deformation conditions with Rxz values ranging from 2.5 to 4.3 and increasing towards the Sirjan thrust. Opening angles of quartz c-axis fabrics and recrystallization regimes suggest deformation temperatures vary from 430 to 625 ± 50°C in the hanging wall rocks. Oblique grain shape and quartz c-axis fabrics were used to estimate the degree of non-coaxiality during deformation. The obtained vorticity profile indicates a down-section increase in kinematic vorticity number (Wm) from 0.6 to 0.89. This range of vorticity numbers confirms contributions of both simple (41–68 %) and pure shear (32–59 %) deformation components. The structural characteristics of the study area ultimately were controlled by oblique motion of the Afro-Arabian plate relative to the Iranian plate.

Type
Original Articles
Copyright
Copyright © Cambridge University Press 2016 

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

Agard, P., Omrani, J., Jolivet, L. & Mouthereau, F. 2005. Convergence history across Zagros (Iran): constraints from collisional and earlier deformation. International Journal of Earth Sciences 94, 401–19.Google Scholar
Alavi, M. 1994. Tectonics of the Zagros orogenic belt of Iran: new data and interpretations. Tectonophysics 229, 211–38.Google Scholar
Alizadeh, A., López-Martínez, M. & Sarkarinejad, K. 2010. 40Ar–39Ar geochronology in a gneiss dome within the Zagros Orogénic Belt. Comptes Rendus Geoscience 342, 837–84.Google Scholar
Allen, M., Jackson, J. & Walker, R. 2004. Late Cenozoic reorganization of the Arabia–Eurasia collision and the comparison of short-term and long-term deformation rates. Tectonics 23, 116.Google Scholar
Bailey, C. M. & Eyster, E. L. 2003. General shear deformation in the Pinaleno Mountains metamorphic core complex, Arizona. Journal of Structural Geology 25, 1883–92.Google Scholar
Bailey, C. M., Francis, B. E. & Fahrney, E. E. 2004. Strain and vorticity analysis of transpressional high-strain zones from the Virginia Piedmont, USA. In Flow Processes in Faults and Shear Zones (eds Alsop, G. I., Holdsworth, R. E., McCaffrey, K. J. W. & Hand, M.), pp. 249–64. Geological Society of London, Special Publication no. 224.Google Scholar
Berberian, M. & King, G. C. P. 1981. Towards a paleogeography and tectonic evolution of Iran. Canadian Journal of Earth Sciences 18, 210–65.Google Scholar
Bhattacharya, A. R. & Weber, K. 2004. Fabric development during shear deformation in the Main Central Thrust Zone, NW-Himalaya, India. Tectonophysics 387, 2346.Google Scholar
Blanc, E. J.-P., Allen, M. B., Inger, S. & Hassani, H. 2003. Structural styles in the Zagros simple folded zone, Iran. Journal of Structural Geology 160, 401–12.Google Scholar
Bouchez, J. L. 1977. Plastic deformation of quartzites at low temperatures in an area of natural strain gradient. Tectonophysics 39, 2550.CrossRefGoogle Scholar
Bouchez, J. L. & Duval, P. 1982. The fabric of polycrystalline ice deformed in simple shear: experiments in torsion, natural deformation and geometrical interpretation. Textures and Microstructures 5, 171–90.Google Scholar
Bouchez, J. L. & Pecher, A. 1981. The Himalayan Main Central Thrust pile and its quartz-rich tectonites in central Nepal. Tectonophysics 78, 2350.Google Scholar
Burg, J. P. 1986. Quartz shape fabric variations and c-axis fabrics in a ribbon mylonite: arguments for an oscillating foliation. Journal of Structural Geology 8, 123–31.Google Scholar
Cavalcanti De Araújo, M. N., Alves Da Silva, F. C., Jardim De Sá, E. F., Holcombe, R. J. & De Vasconcelos, P. M. 2003. Microstructural evolution of the Seridó Belt, NE Brazil: the effect of two tectonic events on development of c-axis preferred orientation in quartz. Journal of Structural Geology 25, 2089–107.Google Scholar
Célérier, J., Harrison, T. M., Beyssac, O., Herman, F., Dunlap, W. J. & Webb, A. G. 2009. Evolution of the Kumaun and Garwhal Lesser Himalaya, India. Part 2: thermal and deformation histories. Geological Society of American Bulletin 121, 1281–97.Google Scholar
Chew, D. M. 2003. An Excel spreadsheet for finite strain analysis using the Rf/Φ technique. Computers and Geosciences 29, 795–99.Google Scholar
Dewey, J. F., Pitman, W. C. III, Ryan, W. B. F. & Bonini, J. 1973. Plate tectonics and the evolution of the Alpine System. Geological Society of America Bulletin 84, 3137–80.2.0.CO;2>CrossRefGoogle Scholar
Escher, A. & Beaumont, C. 1997. Formation, burial and exhumation of basement nappes at crustal scale: a geometric model based on the Western Swiss-Italian Alps. Journal of Structural Geology 19, 955–74.Google Scholar
Etchecopar, A. & Vasseur, G. 1987. A 3D kinematic model of fabric development in polycrystalline aggregates: comparisons with experimental and natural examples. Journal of Structural Geology 9, 705–17.Google Scholar
Faghih, A. & Soleimani, M. 2015. Quartz c-axis fabric development associated with shear deformation along an extensional detachment shear zone: Chapedony Metamorphic Core Complex, Central-East Iranian Microcontinent. Journal of Structural Geology 70, 111.Google Scholar
Fitz Gerald, J. D. & Stünitz, H. 1993. Deformation of granitoids at low metamorphic grades. I: reactions and grain size reduction. Tectonophysics 221, 269–97.Google Scholar
Frassi, C., Carosi, R., Montomoli, C. & Law, R. D. 2009. Kinematics and vorticity of flow associated with post-collisional oblique transpression in the Variscan Inner Zone of northern Sardinia (Italy). Journal of Structural Geology 31, 1458–71.Google Scholar
Ghasemi, A. & Talbot, C. J. 2006. A new tectonic scenario for the Sanandaj–Sirjan Zone (Iran). Journal of Asian Earth Science 26, 683–93.Google Scholar
Gomez-Rivas, E., Bons, P. D., Griera, A., Carreras, J., Druguet, E. & Evans, L. 2007. Strain vorticity analysis using small-scale faults and associated drag folds. Journal of Structural Geology 29, 1882–99.Google Scholar
Goscombe, B. & Passchier, C. W. 2003. Asymmetric boudins as shear sense indicators—an assessment from field data. Journal of Structural Geology 25, 575–89.Google Scholar
Goscombe, B., Passchier, C. W. & Hand, M. 2004. Boudinage classification: end member boudin types and modified boudin structures. Journal of Structural Geology 26, 739–63.Google Scholar
Grasemann, B., Fritz, H. & Vannay, J. C. 1999. Quantitative kinematic flow analysis from the Main Central Thrust Zone (NW-Himalaya, India): implications for a decelerating strain path and the extrusion of orogenic wedges. Journal of Structural Geology 21, 837–53.Google Scholar
Grujic, D., Casey, M., Davidson, C., Hollister, L. S., Kundig, R., Pavlis, T. & Schmid, S. 1996. Ductile extrusion of the Higher Himalayan Crystalline in Bhutan: evidence from the quartz microfabrics. Tectonophysics 260, 2143.Google Scholar
Hanmer, S. & Passchier, C. 1991. Shear-sense indicators: a review. Geological Survey of Canada, Paper 90–17, 72 pp.Google Scholar
Herwegh, M. & Handy, M. R. 1996. The evolution of high temperature mylonitic microfabrics: evidence for simple shearing of a quartz analogue (norcamphor). Journal of Structural Geology 18, 689710.CrossRefGoogle Scholar
Herwegh, M., Handy, M. R. & Heilbronner, R. 1997. Temperature- and strain rate-dependent microfabric evolution in monomineralic mylonite: evidence from in-situ deformation of norcamphor. Tectonophysics 280, 83106.CrossRefGoogle Scholar
Hirth, G., Teyssier, C. & Dunlap, W. J. 2001. An evaluation of quartzite flow laws based on comparisons between experimentally and naturally deformed rocks. International Journal of Earth Sciences 90, 7787.Google Scholar
Hirth, G. & Tullis, J. 1992. Dislocation creep regimes in quartz aggregates. Journal of Structural Geology 14, 145–59.Google Scholar
Holcombe, R. J. & Little, T. A. 2001. A sensitive vorticity gauge using rotated prophyroblasts, and its application to rocks adjacent to the Alpine fault, New Zealand. Journal of Structural Geology 23, 979–89.Google Scholar
Hubbard, M. 1996. Ductile shear as a cause of inverted metamorphism: example from the Nepal Himalaya. Journal of Geology 104, 493–99.Google Scholar
Hubbard, M. 1997. Ductile shear as a cause of inverted metamorphism: example from the Nepal Himalaya: a reply. Journal of Geology 105, 513–4.Google Scholar
Jackson, J. A. & Mckenzie, D. P. 1984. Active tectonics of Alpine–Himalayan belt between western Turkey and Pakistan. Geophysical Journal of the Royal Astronomical Society 77, 185264.Google Scholar
Jain, A. K. & Manickavasagam, R. M. 1993. Inverted metamorphism in the intercontinental ductile shear zone during Himalayan collision tectonics. Geology 21, 407–10.Google Scholar
Jain, A. K. & Manickavasagam, R. M. 1997. Ductile shear as a cause of inverted metamorphism: example from the Nepal Himalaya: a discussion. Journal of Geology 105, 511–3.CrossRefGoogle Scholar
Jessell, M. W. 1987. Grain-boundary migration microstructures in a naturally deformed quartzite. Journal of Structural Geology 9, 1007–14.Google Scholar
Jessup, M. J., Law, R. D. & Frassi, C. 2007. The Rigid Grain Net (RGN): an alternative method for estimating mean kinematic vorticity number (Wm). Journal of Structural Geology 29, 411–21.Google Scholar
Jessup, M. J., Law, R. D., Searle, M. P. & Hubbard, M. S. 2006. Structural evolution and vorticity of flow during extrusion and exhumation of the Greater Himalayan Slab, Mount Everest massif, Tibet/Nepal: implications for orogen-scale flow partitioning. In Channel Flow, Ductile Extrusion and Exhumation in Continental Collision Zones (eds Law, R., Searle, M. P. & Godin, L.), pp. 379414. Geological Society of London, Special Publications no. 268.Google Scholar
Jones, R. R., Holdsworth, R. E., Hand, M. & Goscombe, B. 2006. Ductile extrusion in continental collision zones: ambiguities in the definition of channel flow and its identification in ancient orogens. In Channel Flow, Ductile Extrusion and Exhumation in Continental Collision Zones (eds Law, R. D., Searle, M. P. & Godin, L.), pp. 201–19. Geological Society of London, Special Publications no. 268.Google Scholar
Klepeis, K. A., Daczko, N. R. & Clarke, G. L. 1999. Kinematic vorticity and tectonic significance of superposed mylonites in a major lower crustal shear zone, northern Fiordland, New Zealand. Journal of Structural Geology 21, 1385–406.Google Scholar
Kruhl, J. H. 1996. Prism- and basal-plane parallel subgrain boundaries in quartz: a microstructural geothermobarometer. Journal of Metamorphic Geology 14, 581–9.Google Scholar
Kruhl, J. H. 1998. Reply: Prism- and basal-plane parallel subgrain boundaries in quartz: a microstructural geothermobarometer. Journal of Metamorphic Geology 16, 142–6.Google Scholar
Langille, J., Jessup, M. J., Cottle, J. M. & Newell, D. L. 2010. Kinematics of the Ama Drime Detachment: insights into orogen-parallel extension and exhumation of the Ama Drime Massif, Tibet–Nepal. Journal of Structural Geology 32, 900–19.Google Scholar
Law, R. D. 1987. Heterogeneous deformation and quartz crystallographic fabric transitions: natural examples from the Moine thrust zone at the Stack of Glencoul, northern Assynt. Journal of Structural Geology 9, 819–34.CrossRefGoogle Scholar
Law, R. D. 1990. Crystallographic fabrics: a selective review of their applications to research in structural geology. In Deformation Mechanisms, Rheology and Tectonics (eds Knipe, R. J. & Rutter, E. H.), pp. 335–52. Geological Society of London, Special Publication no. 54.Google Scholar
Law, R. D. 2010. Moine Thrust zone mylonites at the Stack of Glencoul: II – results of vorticity analyses and their tectonic significance. In Continental Tectonics and Mountain Building (eds Law, R. D., Butler, R., Holdsworth, B., Krabbendam, M. & Strachan, R.), pp. 579602. The Legacy of Peach and Horne; Geological Society of London, Special Publication no. 335.Google Scholar
Law, R. D. 2014. Deformation thermometry based on quartz c-axis fabrics and recrystallization microstructures: a review. Journal of Structural Geology 66, 129–61.Google Scholar
Law, R. D., Casey, M. & Knipe, R. J. 1986. Kinematic and tectonic significance of microstructures and crystallographic fabrics within quartz mylonites from the Assynt and Eriboll regions of the Moine thrust zone, NW Scotland. Transactions of the Royal Society of Edinburgh: Earth Sciences 77, 99126.Google Scholar
Law, R. D., Knipe, R. J. & Dayan, H. 1984. Strain path partitioning within thrust sheets: microstructural and petrofabric evidence from the Moine thrust zone at Loch Eriboll, NW Scotland. Journal of Structural Geology 6, 477–97.Google Scholar
Law, R. D., Mainprice, D. H., Casey, M., Lloyd, G. E., Knipe, R. J., Cook, B. & Thigpen, J. R. 2010. Moine thrust zone mylonites at the Stack of Glencoul: I – microstructures, strain and influence of recrystallization on quartz crystal fabric development. In Continental Tectonics and Mountain Building (eds Law, R. D., Butler, R., Holdsworth, B., Krabbendam, M. & Strachan, R.), pp. 543–77. The Legacy of Peach and Horne; Geological Society of London, Special Publication no. 335.Google Scholar
Law, R. D., Morgan, S. S., Casey, M., Sylvester, A. G. & Nyman, M. 1992. The Papoose Flat pluton of eastern California: a re-assessment of its emplacement history in the light of new microstructural and crystallographic fabric observations. Transactions of the Royal Society of Edinburgh: Earth Sciences 83, 361–75.CrossRefGoogle Scholar
Law, R. D., Searle, M. P. & Simpson, R. L. 2004. Strain, deformation temperatures and vorticity of flow at the top of the Greater Himalayan Slab, Everest Massif. Tibet. Journal of Geological Society, London 161, 305–20.Google Scholar
Law, R. D., Stahr, D. W., Francsis, M., Ashley, K. T., Grasemann, B. & Ahmad, T. 2013. Deformation temperatures and flow vorticities near the base of the Greater Himalayan Series, Sutlej Valley and Shimla Klippe, NW India. Journal of Structural Geology 54, 2153.CrossRefGoogle Scholar
Lin, S., Jiang, D. & Williams, P. F. 1998. Transpression (or transtension) zones of triclinic symmetry: natural example and theoretical modelling. In: Continental Transpressional and Transtensional Tectonics (eds Holdsworth, R. E., Strachan, R. A. & Dewey, J. F.), pp. 4157. Geological Society of London, Special Publication no. 135.Google Scholar
Lloyd, G. E. & Freeman, B. 1994. Dynamic recrystallization of quartz under greenschist conditions. Journal of Structural Geology 16, 867–81.Google Scholar
Lisle, R. J. 1985. Geological Strain Analysis: A Manual for the Rf/Φ Method. New York: Pergamon Press, 99 pp.Google Scholar
Lister, G. S. 1977. Discussion: crossed-girdle c-axis fabrics in quartzites plastically deformed by plane strain and progressive simple shear. Tectonophysics 39, 51–4.Google Scholar
Lister, G. S. & Dornsiepen, U. F. 1982. Fabric transitions in the Saxony granulite terrain. Journal of Structural Geology 4, 8192.Google Scholar
Lister, G. S. & Hobbs, B. E. 1980. The simulation of fabric development during plastic deformation and its application to quartzite: fabric transition. Journal of Structural Geology 1, 99115.Google Scholar
Lister, G. S. & Snoke, A. W. 1984. S–C mylonites. Journal of Structural Geology 6, 616–38.Google Scholar
Lister, G. S. & Williams, P. F. 1979. Fabric development in shear zones: theoretical controls and observed phenomena. Journal of Structural Geology 1, 283–98.CrossRefGoogle Scholar
Mainprice, D., Bouchez, J. L., Blumenfeld, P. & Tubia, J. M. 1986. Dominant c slip in naturally deformed quartz: implications for dramatic plastic softening at high temperature. Geology 14, 819–22.Google Scholar
McCall, G. J. H. & Kidd, R. G. W. 1981. The Makran, southeastern Iran: the anatomy of a converging plate margin active from Cretaceous to Present. In: Trench–Fore Arc Geology (eds Leggett, J.), pp. 387–97. Geological Society of London, Special Publications no. 10.Google Scholar
Means, W. D. 1994. Rotational quantities in homogeneous flow and development of small-scale structures. Geology 16, 437–46.Google Scholar
Mohajjel, M. & Fergusson, C. L. 2000. Dextral transpression in Late Cretaceous continental collision, Sanandaj Sirjan Zone, western Iran. Journal of Structural Geology 22, 1125–39.Google Scholar
Mohajjel, M., Fergusson, C. L. & Sahandi, M. R. 2003. Cretaceous–Tertiary convergence and continental collision, Sanandaj–Sirjan Zone, western Iran. Journal of Asian Earth Sciences 21, 397412.Google Scholar
Morgan, S. S. & Law, R. D. 2004. Unusual transition in quartzite dislocation creep regimes and crystal slip systems in the aureole of the Eureka Valley–Joshua Flat–Beer Creek pluton, California: a case for anhydrous conditions created by decarbonation reactions. Tectonophysics 384, 209–31.Google Scholar
Mouthereau, F., Lacombe, O., Tensi, J., Bellahsen, N., Kargar, S. & Amrouch, K. 2007. Mechanical constraints on the development of the Zagros Folded Belt (Fars). In Thrust Belts and Foreland Basins: From Fold Kinematics to hydrocarbon Systems (eds Lacombe, O., Lavé, J., Roure, F. & Vergés, J.), pp. 245–64. Berlin, Heidelberg: Springer-Verlag.Google Scholar
Nyman, M. W., Law, R. D. & Morgan, S. S. 1995. Conditions of contact metamorphism, Papoose Flat Pluton, eastern California, USA: implications for cooling and strain histories. Journal of Metamorphic Geology 13, 627–43.Google Scholar
Okudaira, T., Takeshita, T., Hara, I. & Ando, J. 1995. A new estimate of the conditions for transition from basal <a> to prism [c] slip in naturally deformed quartz. Tectonophysics 250, 3146.Google Scholar
Pamplona, J. & Rodrigues, B. 2011. Kinematic interpretation of shearband boudins: new parameters and ratios useful in HT simple shear zones. Journal of Structural Geology 33, 3850.CrossRefGoogle Scholar
Passchier, C. W. 1987. Stable positions of rigid objects in non-coaxial flow: a study in vorticity analysis. Journal of Structural Geology 9, 679–90.Google Scholar
Passchier, C. W. 1988. Analysis of deformation paths in shear zones. Geologische Rundschau 77, 309–18.CrossRefGoogle Scholar
Passchier, C. W. 1998. Monoclinic model shear zone. Journal of Structural Geology 20, 1121–37.Google Scholar
Passchier, C. W. & Simpson, C. 1986. Porphyroclast systems as kinematic indicators. Journal of Structural Geology 8, 831–43.Google Scholar
Passchier, C. W. & Trouw, R. A. J. 2005. Microtectonics. Berlin: Springer, 366 pp.Google Scholar
Platt, J. P. & Behrmann, J. H. 1986. Structures and fabrics in a crustal scale shear zone, Betic Cordilleras, S.E. Spain. Journal of Structural Geology 8, 1534.Google Scholar
Ramsay, J. G. & Huber, M. I. 1983. The Techniques of Modern Structural Geology. Volume 1: Strain Analysis. London: Academic Press, 307 pp.Google Scholar
Ratschbacher, L., Wenk, R. H. & Sintubin, M. 1991. Calcite textures: examples from nappes with strain-path partitioning. Journal of Structural Geology 13, 369–84.Google Scholar
Regard, V., Bollier, O., Thomas, J. C., Abbasi, M. R., Mercier, J., Shabanian, E., Feghhi, K. & Soleymani, S. 2004. Accommodation of Arabia–Eurasia convergence in the Zagros–Makran transfer zone, SE Iran: a transition between collision and subduction through a young deformation system. Tectonics 23, TC4007, doi: 10.1029/2003TC001599.Google Scholar
Ricou, L. E. 1971. Le croissant ophiolitique péri-arabe. Une ceinture de nappes mises en place au Crétacé supérieur. Revue de Géographie Physique et de Géologie Dynamique XIII, 327–50.Google Scholar
Samani, B. 2013. Quartz c-axis evidence for deformation characteristics in the Sanandaj–Sirjan metamorphic belt, Iran. Journal of African Earth Sciences 81, 2834.Google Scholar
Sanderson, D. J. & Marchini, W. R. D. 1984. Transpression. Journal of Structural Geology 6, 449–58.Google Scholar
Sarkarinejad, K. 1999. Tectonic finite strain analysis using Ghouri deformed conglomerate, Neyriz area, Southwestern Iran. Iranian Journal of Science and Technology 23, 352–63.Google Scholar
Sarkarinejad, K. 2005. Structures and microstructures related to steady-state mantle flow in the Neyriz ophiolite, Iran. Journal of Asian Earth Sciences 25, 859–81.Google Scholar
Sarkarinejad, K. 2007. Quantitative finite strain and kinematic flow analyses along the Zagros transpression zone, Iran. Tectonophysics 442, 4965.Google Scholar
Sarkarinejad, K. & Azizi, A. 2008. Slip partitioning and inclined dextral transpression along the Zagros Thrust System, Iran. Journal of Structural Geology 30, 116–36.Google Scholar
Sarkarinejad, K., Faghih, A. & Grasemann, B. 2008. Transpressional deformations within the Sanandaj–Sirjan Metamorphic Belt (Zagros Mountains, Iran). Journal of Structural Geology 30, 818–26.Google Scholar
Sarkarinejad, K. & Ghanbarian, M. A. 2014. The Zagros hinterland fold-and-thrust belt in-sequence thrusting, Iran. Journal of Asian Earth Sciences 85, 6679.CrossRefGoogle Scholar
Sarkarinejad, K., Godin, L. & Faghih, A. 2009. Kinematic vorticity flow analysis and 40Ar/39Ar geochronology related to inclined extrusion of the HP–LT metamorphic rocks along the Zagros accretionary prism, Iran. Journal of Structural Geology 31, 691706.Google Scholar
Sarkarinejad, K. & Keshavarz, S. 2015. Quantitative kinematic analysis of the asymmetric boudins of the Zagros accretionary prism, Iran. Geosciences Journal 19, 415–30.Google Scholar
Sarkarinejad, K., Partabian, A. & Faghih, A. 2013. Variations in the kinematics of deformation along the Zagros inclined transpression zone, Iran: implications for defining a curved inclined transpression zone. Journal of Structural Geology 48, 126–36.Google Scholar
Schmid, S. M. & Casey, M. 1986. Complete fabric analysis of some commonly observed quartz c-axis patterns. In Mineral and Rock Deformation: Laboratory Studies – The Paterson Volume (eds Hobbs, B. E. & Heard, H. C.), pp. 263–86. American Geophysical Union, Geophysical Monograph vol. 36. Washington, DC, USA.Google Scholar
Sengör, A. M. C., Altiner, D., Cin, A., Custaomer, T. & Hsu, K. J. 1988. Origin and assembly of the Tethyside orogenic collage at the expense of Gondwana Land. In Gondwana and Tethys (eds Audley-Charles, M. G. & Hallam, A.), pp. 119–81. Geological Society of London, Special Publication no. 37.Google Scholar
Sheikholeslami, M. R., Pique, A., Mobayen, P., Sabzehei, M., Bellon, H. & Hashem Emami, M. 2008. Tectono-metamorphic evolution of the Neyriz metamorphic complex, Quri-Kor-e-Sefid area (Sanandaj–Sirjan Zone, SW Iran). Journal of Asian Earth Sciences 31, 504–21.Google Scholar
Simpson, C. & De Paor, D. G. 1997. Practical analysis of general shear zones using the porphyroclast hyperbolic distribution method: an example from Scandinavian Caledonides. In Evolution of Geological Structures in Micro to Macro Scales (eds Sengupta, S.), pp. 169–84. London: Chapman and Hall.Google Scholar
Simpson, C. & Schmid, S. M. 1983. An evaluation of criteria to deduce the sense of movement in sheared rocks. Geological Society of America Bulletin 94, 1281–8.Google Scholar
Stipp, M., Stunitz, H., Heilbronner, R. & Schmid, S. M. 2002 a. The eastern Tonale fault zone: a ‘natural laboratory’ for crystal plastic deformation of quartz over a temperature range from 250 to 700°C. Journal of Structural Geology 24, 1861–84.Google Scholar
Stipp, M., Stünitz, H., Heilbronner, R. & Schmid, S. M. 2002 b. Dynamic recrystallization of quartz: correlation between natural and experimental conditions. In Deformation Mechanisms, Rheology and Tectonics: Current Status and Future Perspectives (eds De Meer, S., Drury, M. R., De Bresser, J. H. P. & Pennock, G. M.), pp. 171–90. Geological Society of London, Special Publication no. 200.Google Scholar
Stöcklin, J. 1968. Structural history and tectonics of Iran: a review. American Association of Petroleum Geologists Bulletin, 52, 1229–58.Google Scholar
Stöcklin, J. 1974. Possible ancient continental margins in Iran. In The Geology of Continental Margins (eds Burk, C. A. & Drake, C. L.), pp. 873–88. New York: Springer.Google Scholar
Sullivan, W. A. 2008. Significance of transport-parallel strain variations in part of the Raft River shear zone, Raft River Mountains, Utah, USA. Journal of Structural Geology 30, 138–58.Google Scholar
Sullivan, W. A. & Law, R. D. 2007. Deformation path partitioning within the transpressional White Mountain shear zone, California and Nevada. Journal of Structural Geology 29, 583–98.Google Scholar
Tatar, M., Hatzfeld, D. & Ghafory-Ashtiyani, M. 2004. Tectonics of the Central Zagros (Iran) deduced from microearthquake seismicity. Geophysical Journal International 156, 255–66.Google Scholar
Tikoff, B. & Fossen, H. 1995. The limitation of three-dimensional kinematic vorticity analysis. Journal of Structural Geology 17, 1771–84.Google Scholar
Tikoff, B. & Teyssier, C. 1994. Strain and fabric based on porphyroclast interaction. Journal of Structural Geology 16, 477–91.Google Scholar
Tullis, J. & Yund, R. A. 1992. The brittle-ductile transition in feldspar aggregates; an experimental study. In Fault Mechanics and Transport Properties in Rocks (eds Evans, B. & Wong, T. F.), pp. 89118. New York: Academic Press.Google Scholar
Vernant, P., Nilforoushan, F., Hatzfeld, D., Abbasi, M. R., Vigny, C., Masson, F., Nankali, H., Martinod, J., Ashtiani, A., Bayer, R., Tavakoli, F. & Chery, J. 2004. Present-day crustal deformation and plate kinematics in the Middle East constrained by GPS measurement in Iran and northern Oman. International Journal of Geophysics 157, 381–98.Google Scholar
Vernooij, M. G. C., Brok, B. & Kunze, K. 2006. Development of crystallographic preferred orientations by nucleation and growth of new grains in experimentally deformed quartz single crystals. Tectonophysics 427, 3553.Google Scholar
Wagner, T., Lee, J., Hacker, B. R. & Seward, G. 2010. Kinematics and vorticity in Kangmar Dome, southern Tibet: testing mid-crustal channel-flow models for the Himalaya. Tectonics 29, 126.Google Scholar
Wallis, S. R. 1992. Vorticity analysis in a metachert from the Sanbagawa Belt, SW Japan. Journal of Structural Geology 14, 271–80.Google Scholar
Wallis, S. R. 1995. Vorticity analysis and recognition of ductile extension in the Sanbagawa belt, SW Japan. Journal of Structural Geology 17, 1077–93.Google Scholar
Wallis, S. R., Platt, J. P. & Knott, S. D. 1993. Recognition of syn-convergence extension in accretionary wedges with examples from the Calabrian Arc and the Eastern Alps. American Journal of Science 293, 463–95.Google Scholar
Wang, Y., Zhang, Y., Fan, W. & Peng, T. 2005. Structural signatures and 40Ar/39Ar geochronology of the Indosinian Xuefengshan tectonic belt, South China Block. Journal of Structural Geology 27, 985–98.Google Scholar
Wells, M. L. 2001. Rheological control on the initial geometry of the Raft River detachment fault and shear zone, western United States. Tectonics 20, 435–57.Google Scholar
Xypolias, P. 2009. Some new aspects of kinematic vorticity analysis in naturally deformed quartzites. Journal of Structural Geology 31, 310.Google Scholar
Xypolias, P. 2010. Vorticity analysis in shear zones: a review of methods and applications. Journal of Structural Geology 32, 2072–92.Google Scholar
Xypolias, P. & Doutsos, T. 2000. Kinematics of rock flow in a crustal-scale shear zone: implication for the orogenic evolution of the southwestern Hellenides. Geological Magazine 137, 8196.Google Scholar
Xypolias, P., Chatzaras, V. & Koukouvelas, I. K. 2007. Strain gradients in zones of ductile thrusting: insights from the External Hellenides. Journal of Structural Geology 29, 1522–37.Google Scholar
Xypolias, P. & Kokkalas, S. 2006. Heterogeneous ductile deformation along a mid-crustal extruding shear zone: an example from the External Hellenides (Greece). In Channel Flow, Ductile Extrusion and Exhumation in Continental Collision Zones (eds Law, R. D., Searle, M. P. & Godin, L.), pp. 497516. Geological Society of London, Special Publications no. 268.Google Scholar
Xypolias, P. & Koukouvelas, I. K. 2001. Kinematic vorticity and strain patterns associated with ductile extrusion in the Chelmos shear zone (External Hellenides, Greece). Tectonophysics 338, 5977.Google Scholar
Xypolias, P., Spanos, D., Chatzaras, V., Kokkalas, S. & Koukouvelas, I. 2010. Vorticity of flow in ductile thrust zones: examples from the Attico–Cycladic Massif (Internal Hellenides, Greece). In Continental Tectonics and Mountain Building (eds Law, R. D., Butler, R., Holdsworth, B., Krabbendam, M. & Strachan, R.), pp. 687714. Geological Society of London, Special Publication no. 335.Google Scholar