Está en la página 1de 35

1

  Thermal-mechanical modeling of salt-based mountain belts with pre-


2   existing basement faults: Application to the Zagros fold and thrust
3   belt, southwest Iran
4   Faramarz Nilfouroushan1, 2, Russell Pysklywec2, Alexander Cruden3, Hemin Koyi1
5  
6   1) Hans Ramberg Tectonic Laboratory, Department of Earth Sciences, Uppsala University,
7   Villavägen 16, 75236, Uppsala, Sweden,
8   2) Department of Earth Sciences, University of Toronto, 22 Russell St, Toronto, ON M5S 3B1,
9   Canada
10   3) School of Geosciences, Monash University, Melbourne, VIC 3800, Australia
11  

12   Abstract

13   Two-dimensional thermal-mechanical models of thick-skinned, salt-based fold and thrust belts,


14   such as the Zagros, SW Iran, are used to address: 1) the degree of deformation and decoupling
15   between cover and basement rocks due to the presence of a weak salt detachment; 2) the
16   reactivation potential of pre-existing basement normal faults due to brittle or ductile behavior of
17   the lower crust (as related to cold or hot geothermal gradients); and 3) variations in deformation
18   style and strain distribution. The geometry and kinematics of the orogenic wedge and the activity
19   of pre-existing basement faults are strongly influenced by the geothermal gradient (defined by
20   the Moho temperature, MT) and basement rheology. We infer that the MT plays a major role in
21   how the lower and upper crust transfer deformation towards the foreland. In relatively hot
22   geotherm models (MT = 600°C at 36 km depth), the lowermost basement deforms in a ductile
23   fashion while the uppermost basement underlying the sedimentary cover deforms by folding,
24   thrusting, and displacements along pre-existing basement faults. In these models, cover units
25   above the salt detachment occur within a less deformed, wide plateau in the hinterland. In
26   relatively cold geotherm models (MT = 400°C at 36 km depth), deformation is mainly restricted
27   to basement imbricate thrusts that form within the orogenic hinterland. Detachment folding,
28   thrusting and gravity gliding occur within cover sediments above uplifted basement blocks.
29   Gravity gliding contributes to a larger amount of shortening in the cover compared to the
30   basement.

31   Keywords: Thermal-mechanical models, Thick-skinned deformation, Salt detachment,


32   Basement faults, Zagros fold and thrust belt

33   Citation: Nilfouroushan, F., R. Pysklywec, A. Cruden, and H. Koyi (2013), Thermal-mechanical


34   modeling of salt-based mountain belts with pre-existing basement faults: Application to the
35   Zagros fold and thrust belt, southwest Iran, Tectonics, 32, doi:10.1002/tect.20075.
36  

37   1. Introduction

38   Several studies have physically (Graveleau et al., 2012 and references therein) and numerically
39   (Ellis et al., 2004; Stockmal et al., 2007; Yamato et al., 2011; Buiter, 2012) simulated thin-
40   skinned deformation at convergent plate boundaries and illustrated how different rheological and

  1  
41   mechanical parameters change the mechanism and style of deformation of sedimentary cover
42   rocks. Some studies have gone further and modeled the cover deformation above basement
43   normal faults (e.g. Schedl and Wiltschko 1987; Koyi et al. 1991 and 1993; Hardy, 2011) and
44   others have investigated the influence of cover strength on basement-involved fault propagation
45   folding in cover sediments (e.g. Hardy and Finch, 2006). A few studies have modeled thick-
46   skinned deformation in which both brittle cover sediments and deeper brittle-ductile basement
47   rocks are involved (e.g. Barr and Dahlen 1989; Buiter and Torsvik 2007). However, scaled
48   laboratory modeling of thick-skinned deformation is not easy and more sophisticated materials
49   and setups are required to handle temperature-dependent ductile rheologies at depth (Davy and
50   Cobbold 1991; Boutelier and Oncken 2011). Thermal-mechanical numerical experiments are
51   therefore favored for modeling of thick-skinned deformation involving both brittle upper crust and
52   ductile lower crustal rheologies (e.g. Bird 1978; Buiter et al. 2009). In thick-skinned deformation,
53   weak zones such as pre-existing basement faults and their reactivation by inversion tectonics
54   (Lacombe and Mouthereau, 2002), mechanically weak layers such as salt or shale within the
55   cover sediments or above the basement acting as weak detachments (Yamato et al. 2011), and
56   the depth- and temperature-dependent viscosity of the mid- to lower crust, are important
57   parameters in crustal shortening studies (Buiter et al. 2009).

58   Natural examples of basement-involved fold and thrust belts include the Andes (Kley et al.,
59   1999; Cristallini and Ramos, 2000), Urals (Brown et al., 1999), Taiwan (Mouthereau and
60   Lacombe 2006), Rockies (Dechesne and Mountjoy, 1992), the Alps (Rostein and Schaming
61   2004) and the Zagros (Berberian, 1995; Molinaro et al. 2005; Mouthereau et al. 2007). Some
62   fold and thrust belts contain single or multiple weak detachments (salt or shale) such as the
63   Zagros, SW Iran (Davis and Engelder, 1985; Talbot and Alvai, 1996; Bahroudi and Koyi, 2003
64   Sherkati et al., 2006), the Northern Apennines, Italy (Massoli et al., 2006), the southern
65   Canadian Rocky Mountains (Stockmal et al., 2007) and the Jura of the Alps (Davis and
66   Engelder 1985). Mechanically weak layers at the basement-cover interface or between
67   stratigraphic levels of fold and thrust belts influence the deformation style and kinematics of an
68   orogenic system (Davis and Engelder 1985; Koyi 1988; Cotton and Koyi 2000; Costa and
69   Vendeville 2002; Nilforoushan et al., 2008; Nilfouroushan et al. 2012; Ruh et al. 2012). In
70   addition, pre-existing faults in the upper brittle crust, acting as weak zones in the basement, can
71   be reactivated and change the localization and distribution of deformation in overlying
72   sedimentary rocks. It is currently not clear how effectively ductile flow of lower crust can
73   reactivate pre-existing basement faults in the upper brittle crust when there is a layer of salt
74   between the cover and basement rocks.
75   Using a series of 2-D thermal-mechanical finite element models we explore how thick-skinned
76   deformation and the interactions between basement and cover rocks are influenced by different
77   parameters such as the geothermal gradient, expressed as Moho temperature, and the
78   presence of weak basal detachments like salt in the basal succession of the sediment cover,
79   and weak zones in the basement (e.g., pre-existing normal faults). Owing to the inherent
80   approximation of numerical modeling to the complex real Earth, we do not aim to model exactly
81   the history and present-day deformation of any orogenic system. Rather, we relate our modeling
82   approach to nature by considering the available thermal-mechanical parameters of the Zagros

  2  
83   fold and thrust belt, as an opportunity to examine thermal-mechanical rheological and structural
84   interactions with constraints from an active orogen. Although the experiments focus on the
85   Zagros fold and thrust belt, the results can be applied to other fold and thrust belts with similar
86   tectonic configurations.

87   1.1. The Zagros fold and thrust belt

88   The Zagros fold and thrust belt (Fig. 1A) is the result of active convergence between the rifted
89   continental margin of the Arabian plate and the Iranian continental blocks following the closure
90   of the Neo-Tethys Ocean during the Late Eocene (Frizon de Lamotte et al., 2011). The ongoing
91   deformation of this convergence is not distributed homogenously in Iran and is mainly taken up
92   by mountain belts like the Zagros. Central Iran is undergoing relatively little deformation and
93   acts as the rigid backstop to the Zagros (Vernant and Chery, 2006). The present-day active
94   deformation within the Zagros has been defined by several GPS measurement campaigns (e.g.
95   Nilforoushan et al., 2003; Vernant et al., 2004; Hessami et al., 2006; Walpersdorf et al., 2006;
96   Tavakoli et al., 2008). These studies show that the cover sequence of the Zagros is shortening
97   relatively slowly (5±3 mm/yr) above a high-frictional basal detachment in the NW Zagros,
98   whereas in the southeast of the Zagros the shortening rate is higher (8±3 mm/yr)(Fig. 1). Total
99   shortening since ~5 Ma (Blanc et al. 2003; Allen et al. 2004) in the Zagros is estimated to range
100   between 45 and 65 km (Blanc et al., 2003; McQuarrie, 2004; Oveisi et al., 2007), consistent with
101   GPS-derived shortening rates (Hessami et al., 2006; Walpersdorf et al., 2006; Tavakoli et al.,
102   2008).

103   Based on stratigraphic data the total thickness of the cover sediments in the Zagros region
104   varies between 7 km in the Fars area to 9 km in the Dezful area (Fig. 1) (Alavi, 2007). The
105   southeastern cover sediments overlie a 1-2 km thick layer of weak Hormuz (Neoproterozoic–
106   Cambrian) salt that partially decouples the Phanerozoic cover from its Precambrian crystalline
107   basement (Koyi 1988; Talbot and Alavi 1996; Bahroudi and Koyi 2003). The thick Hormuz salt
108   layer, not only acts as an efficient detachment, but also by feeding salt diapirs, changes the
109   deformation mechanism within the cover units mainly in the eastern Zagros. The numerous
110   emerged or buried salt diapirs rising from this salt detachment have influenced the shape,
111   localization, propagation and orientation of the folds in the eastern Zagros (Talbot and Alavi
112   1996; Jahani et al. 2009).
113   By opening of the Neo-Tethys ocean during Permian-Early Triassic (Berberian and King 1981;
114   Verges et al. 2011) a passive margin and several extensional faults formed in the Arabian basin
115   that were then covered by sediments (Berberian and King 1981; Bahroudi and Talbot 2003).
116   The exact locations of these pre-existing rift-related extensional faults which have been inverted
117   during the collision stage are not clear but their existence and approximate locations have been
118   inferred from analysis of surface geomorphologic features, topographical sections and
119   earthquake spatial distribution (Berberian 1995; Mouthereau et al., 2006; Alavi 2007). The
120   vertical distribution of earthquakes in the Zagros ranges from 4 to 30 km in depth with the
121   majority between 7-20 km (Maggi et al., 2000; Tatar, et al., 2004; Nissen et al., 2011;
122   Yaminfard, 2012a and b). This indicates that faults are active in both, the cover and basement

  3  
123   sequence, suggesting that the shortening of the Zagros fold and thrust belt is not taken up only
124   by the cover sediments but also by the basement (Jackson 1980; Berberian 1995; Hessami et
125   al., 2001; Talebian and Jackson, 2004; Oveisi et al., 2009). Further, the spatial distribution of
126   earthquakes is limited to the upper 20 km of the crust and implies that the Brittle Ductile
127   Transition (BDT) zone is located at a depth of around 20 km or deeper and faulting is limited to
128   the upper 20km crust of the Zagros (Figs. 1 and 2) (Nissen et al., 2011).
129   The Moho depth beneath the Zagros inferred from geophysical measurements using receiver
130   functions is estimated to be between 40-50 km (Fig. 1) (Hatzfeld et al., 2003, Paul et al 2006,
131   2010, Manaman et al., 2011; Yaminifard et al., 2012a). The hinterland-dipping Moho (β=0.5°)
132   makes the crust slightly thicker under the Sanandaj-Sirjan zone relative to the foreland in the
133   simply folded zone (Fig. 1).
134   Geothermal gradient contours in Motiei (1990) and Bordenave (2008), both based on well data
135   from Orbell (1977), indicate a variation in geothermal gradient across and along the belt from
136   10°C/km to 28°C/km. The increasing geothermal gradients across the belt from the High Zagros
137   to Persian Gulf were partly assigned to tectonically thickened crust near the suture zone (Bird
138   1978). The anomalies in the contours of geothermal gradients along the belt were correlated
139   with north-south trending reactivated old basement faults in the Zagros (Bahroudi and Talbot
140   (2003). For geodynamic modeling of the Zagros, Mouthereau et al., (2006) used a geothermal
141   gradient of 10-15°C/km consistent with a Moho temperature (MT) of 450-675°C. The average
142   surface heat flow of 40 mWm-2 used by Bird (1978) for thermal-mechanical modeling of the
143   Zagros also corresponds to a MT of about 500°C at a Moho depth of 40 km.
144   Balanced cross-sections (e.g. Sherkati and Letouzey 2004) and analysis of topographic profiles
145   across the Zagros (Mouthereau et al., 2006) also show that basement faulting is required to
146   explain the present-day topography. However, in the Zagros, where a relatively thick salt layer
147   covers active basement faults, seismic reflection data fail to image basement structures (Blanc
148   et al., 2003; Alavi, 2004; McQuarrie 2004; Sherkati and Letouzey 2004; Paul et al., 2006) (Fig.
149   1). Moreover, due to the presence of this thick salt detachment, the seismicity has a diffuse
150   pattern (Fig. 1)(Koyi et al., 2000; Nissen et al 2011) and therefore locating seismogenic
151   basement faults is elusive (Berberian 1995).
152  

153   2. Numerical experiments

154  
155   In our experiments, we used a two-dimensional, thermal-mechanical finite element code
156   (SOPALE) which can model high finite strain based on an Arbitrary Lagrangian Eulerian (ALE)
157   method (Fullsack, 1995). This code has been extensively used in a range of different
158   geodynamic modeling applications (Pysklywec and Shahnas, 2003; Beaumont et al., 2004;
159   Pysklywec and Cruden 2004; Cruden et al., 2006; Buiter and Torsvik, 2007; Stockmal et al.,
160   2007; Beaumont et al 2009; Buiter et al., 2009; Gray and Pysklywec, 2010; Nilfouroushan et al.,
161   2012).
162  

  4  
163   In the ALE numerical technique, SOPALE simultaneously uses Eulerian and Lagrangian grids:
164   Finite element computations are performed on a Eulerian grid whose elements are only
165   stretched vertically to accommodate the evolution of topography on the free upper surface; the
166   fully deforming Lagrangian grid tracks the migrating interfaces and material properties. Each of
167   these grids is made up of initially rectangular elements.
168   The models assume incompressibility of materials. While studies show that compressibility of
169   rock may have an effect for deep mantle convection (e.g., Jarvis and Mckenzie, 1980), at the
170   crustal scale the approximation of incompressibility will not influence the behaviour of the
171   models. The governing equations for the models are
172  
173   ∇. ! = 0, (1)
174   ∇. σ!" + +!" = 0, (2)
!"
175   !!! + !. ∇! = !∇! ! + ! , (3)
!"
176   !(!) = !! (1 − ! ! − !! ), (4)
177  
178   where !, σ!" , !, !  , !! , !, !, !  and  t    are velocity, stress tensor, density, gravitational acceleration,
179   specific heat capacity, temperature, thermal conductivity, volumetric rate of internal heat
180   production and time respectively. The other variables, !, !! and !! are thermal expansivity,
181   reference material density and reference temperature respectively.
182   Similar to Mohr-coulomb failure, the brittle deformation for frictional-plastic materials is specified
183   by a pressure-dependent incompressible Drucker-Prager yield criterion:

!/!
184   !′ ! = ! 1 − ! sin ! + ! cos ! , (5)

185   where !′! is the second invariant of deviatoric stress, ! is is the pressure, ! is the pore fluid
186   factor, ! is the angle of internal friction and ! is the cohesion. The ! value can decrease
187   linearly in a range of specified strain (frictional weakening).

188   The viscous deformation of materials follows either linear (Newtonian) viscous behavior
!
( )
189   (!′!! = 2!!! )

190   or power-law creep where the effective viscosity,  !!! , is:


!
!(!!!) !!! !! ( !!) !!
191   !!! = (3 !! 2 ! )! ! !′!! ! !"# (6)

192   where A is the material constant, n is power-law exponent, !′! is second invariant of the
193   deviatoric strain-rate tensor, Q is thermal activation energy, R is gas constant and T is
194   temperature.

195  

196   2.1. Model setup

  5  
197   A series (Table 1) of 2-D thermal-mechanical shortening experiments was run to investigate the
198   thick-skinned deformation of a salt-based fold and thrust belt with pre-existing basement faults.
199   The model parameters (Table 2) were set based on the Zagros as a natural prototype. Our aim
200   is not to simulate the full deformational history of the Zagros and its structure in detail. Instead,
201   we use the best available observational constraints for parameters in the models and assess
202   the interaction between the cover and basement, localization and distribution of deformation,
203   and how variations in the geothermal gradient change the strain distribution when there is a
204   weak salt detachment and pre-existing weak zones present in the basement.

205   To set up our experiments, we started with a rectangular box of 300 km by 36 km that consisted
206   of 8 km sedimentary rocks overlying a stepped basement (inherited from a rifting episode) and
207   separated by a 0.5 km-2.5 km thick salt layer (Fig. 3A). The arbitrary three basement steps (with
208   heights of 0.5km, 1km and 1km respectively from left to right in Fig. 3) are not considered as
209   faults and they only introduce velocity discontinuities. Since the current Moho depth is ~40-50
210   km in the southeastern part of Zagros (Fig. 1) (Hatzfeld et al., 2003, Paul et al., 2006, 2010,
211   Yaminfard et al. 2012) we considered the initial Moho depth of 36 km (Vergas et al. 2011) to
212   take into account crustal thickening after 50 km shortening in about 5-6 million years (Blanc et
213   al. 2003; Sherkati et al., 2006; Oveisi et al. 2007, Verges et al. 2011). As stated before,
214   geophysical studies of the current geometry of the Moho under the Zagros (e.g. Hatzfeld et al.
215   2003; Paul et al., 2006) indicate a gentle hinterland-dipping geometry (β=0.5°) across the belt
216   (Fig. 1). However, this is only an approximation of the current geometry of the Moho, and the
217   initial Moho geometry at the onset of the Zagros shortening is still debated (Vergas et al. 2011).
218   Therefore, to simplify our models we assumed a typical horizontal Moho (i.e. β=0°). The models
219   were shortened orthogonally and continuously from one side by pushing a strong indenter into
220   the material box. The typical orthogonal indentor setup used in previous analytical, analogue
221   and numerical wedge simulations (e.g. Davis et al. 1983; Buiter and Torsvik 2007;
222   Nilfouroushan et al. 2012; Ruh et al. 2012) simplifies the model setup and the interpretations of
223   the results, and facilitates comparisons between different methods and results. All models were
224   shortened at a rate of 8 mm/yr, a rate deduced from present-day GPS measurements (Fig. 1)
225   and consistent with geological rates (Hessami et al. 2006, Walpersdorf et al. 2006). A high
226   viscosity of the indenter (1030 Pas) relative to the other materials in the box meant that the
227   indenter did not deform as it was pushed into the solution space. The relative shortening
228   between the indenter and the material in the box in our models is similar to the Central-Iranian
229   block (Fig. 1) as a rigid indenter pushing against the southeastern part of Zagros. Central Iran is
230   presently undergoing relatively little deformation (±2 mm/yr) and can be considered a relatively
231   rigid backstop to the Zagros (e.g. Vernant and Chery, 2006).

232   We used a rectangular node resolution of 601 x 121 (equal to 0.5 km for horizontal and 0.3 km
233   for vertical resolution) for the Eulerian and 1801 x 361 for the Lagrangian grids. The upper
234   surface is a free surface, the sidewalls are free-slip, and the bottom surface, the Moho, is a no-
235   slip boundary (Fig. 2). A no-slip basal boundary assumption implies no horizontal movements
236   occur right at the bottom surface (Moho). Due to the temperature-dependent rheology used to
237   model lower crustal rocks, the ductile behavior of the lower crust permits the materials just

  6  
238   above the Moho to deform readily, thereby minimizing the effect of the no-slip lower boundary. A
239   similar no-slip assumption for the Moho boundary has been used in other modeling studies
240   (e.g., Mothereau et al. 2006; Buiter and Torsvik 2007).

241   We set the angle of internal friction to 15° for the overburden sediments and 20° for the
242   basement rocks and let the frictional strength of the brittle crust decrease linearly by 50% across
243   a strain range of 0.5 to 1.5 to approximate material weakening due to, for example, an increase
244   in pore-fluid pressure in nature (e.g. Buiter and Torsvik 2007; Gray and Pysklywec, 2012a).
245   Similar to previous modeling studies (Pysklywec and Beaumont, 2004; Gray and Pysklywec,
246   2012a), we use an "effective angle of internal friction" for ! with pore fluid factor ! = 0. This
247   factors in the pore fluid pressure implicitly with the assumed (lower) angle of internal friction of
248   15°-7.5° for overburden sediments (Table 2).

249   Following previous studies we model the weak Hormuz salt in Zagros by a Newtonian viscous
250   rheology with an effective viscosity of 1018 Pas and density of 2200 kgm-3 (Table 2) (e.g.
251   Mouthereau et al. 2006; Yamato et al. 2011). The extension of the salt detachment is varied in
252   our models to study its effect on orogenic wedge deformation (Table 1). However, we limit the
253   frontal extension of the salt detachment to a distance of 240km (in initial setups) from the
254   indenter (Fig. 2) to avoid any frontal boundary effect on cover deformation. We used density
255   values of 2600 kg m−3 for the cover sediments and 2900 kg m−3 for the crystalline crust (Snyder
256   and Barazangi 1986; Paul et al., 2006).

257   The composition of the basement of the Zagros is poorly known (e.g., Bahroudi & Talbot, 2003).
258   The only observed basement rocks are blocks of orthogneiss, metasediments, amphibolites and
259   serpentinites intruded by granite, gabbro and basalt brought to the surface in salt diapirs
260   (Haynes and McQuillan, 1974; Kent, 1979). Following Mouthereau et al. (2006), we assumed
261   quartz diorite and diabase to be suitable compositions for the basement rocks of the Zagros
262   basement. Hence, we evaluated the available temperature-dependent power-law creep laws for
263   these compositions in our numerical experiments (Tables 1 and 2). As stated above,
264   earthquakes in the southeastern Zagros occur at depths between 4-30 km with the majority
265   between 7-20 km (Maggi et al., 2000; Talebian and Jackson, 2004, Nissen et al., 2011;
266   Yaminifard et al. 2012a and b), which suggests that ductile deformation should occur below ~20
267   km. To demonstrate the strength profiles that result from using the flow laws for diabase and
268   quartz diorite, we used the same flow laws as Mouthereau et al., (2006) (Fig. 4). The input
269   parameters for the brittle and ductile deformation of the model crust are given in Table 2, using
270   parameters suggested by Vernant and Chery (2006), Mouthereau et al., (2006) and references
271   therein (i.e. Goetze 1978; Hansen and Carter 1982; Wilks and Carter 1990). The selection of
272   diabase and quartz diorite flow laws to represent the rheology of the ductile crust is reasonable
273   as their strength profiles result in a brittle ductile transition (BDT) depth of ≥ 20 km depth for MTs
274   of 400-600°C (Fig. 4).

275   The number, location, and geometry of the basement faults in the Zagros are poorly known,
276   hence we introduce them arbitrarily in our models. In experiments with pre-existing basement
277   faults, we consider the faults to be 1500 meters wide and dipping at 60° (as such, the faults are

  7  
278   resolved by three Eulerian elements, 3x500=1500m). Following Buiter and Torsvik (2007), the
279   faults are filled with a Newtonian material with a viscosity of 1020 Pa s to mimic weak inherited
280   normal faults in the basement. The faults extend down to 20 km depth where the brittle
281   deformation depth has been inferred from earthquakes studies (Tatar et al., 2004; Yaminfard et
282   al., 2012a and b).
283  
284   We ran 19 shortening experiments (Table 2) with or without basement faults, changing the
285   basement rheology, salt distribution and increasing Moho temperature (MT) systematically from
286   400°C to 500°C and 600°C (assuming geothermal gradient of 11°-17°C/km for a 36km thick
287   crust) which covers the reported MT range in the previous studies (Bird 1978; Mouthereau et al.,
288   2006; Vernant and Chery 2006).
289   We simplified our models by ignoring the effect of isostatic adjustment and thermal subsidence
290   of the underlying lithosphere; also, erosional and depositional processes were not included,
291   although these can have an influence on the behavior of such models (Pysklywec 2006; Gray
292   and Pysklywec, 2012b). As described in Buiter and Torsvik (2007), in orogenic wedge models
293   with this type of configuration the ductile/viscous lower crust in the models allows a “simple
294   form” of effective isostatic compensation in the crust to occur. With a deeper (i.e., mantle)
295   isostatic compensation, there may be some modification to details of the structural geometry of
296   the crust, but this is beyond the scope of the modelling code at this scale of lithospheric
297   investigation.
298  
299   3. Results

300   3.1. Rheology of basement rocks

301   Using a diabase composition and rheology results in a higher effective viscosity lower crust
302   compared to quartz diorite (Fig. 4). The BDT depths are 24, 20 and 16 km for quartz diorite and
303   31, 26 and 22 km for diabase, for MTs of 400°C, 500°C and 600°C, respectively. This means
304   that diabase has a deeper BDT than quartz diorite under the same thermal and mechanical
305   conditions.

306   We tested basement comprising both diabase and quartz diorite in our models and observe
307   composition has a strong influence on deformation behavior (Fig. 5). In the early stages of
308   shortening, deformation in the cover units starts by formation of shear zones in the frontal part of
309   the cover where the salt is pinched out. With further shortening deformation in the cover units
310   propagates backwards (Fig. 5). Due to the shallower BDT depth in experiments using quartz
311   diorite, (Fig. 4), flow in the more ductile lower crust suppresses significant deformation in the
312   cover above the salt detachment (Fig. 5B-D). In all three models using quartz diorite with
313   different MTs, the upper crust deforms similarly and no shear zones (except one near the
314   indentor) are localized in the 200 km long sedimentary cover (Fig. 5B-D). However, the
315   basement is folded in the case of all MTs. In the hottest model, MT = 600°C, the basement is
316   also affected by the load of the especially thickened cover at the second basement step

  8  
317   location. Due to the flow of hot ductile lower crust in this model, the upper brittle crust including
318   the cover subside significantly at this location (Fig. 5D).

319   The cover units in models with diabase deformed more, with faulting and development of pop-up
320   structures especially in the coldest model (MT = 400°C) (Fig. 5E). In the case of the Zagros,
321   with reported earthquakes depths to around 20km, diabase seems to be a better choice for the
322   composition and rheology of the basement rocks. This is in agreement with a recent study on
323   lithospheric strength of the Zagros (Nankali 2012) that suggests a relatively cold geothermal
324   gradient and a diabase or granulitic composition with a BDT located around 21-28 km We
325   therefore focus our modeling using diabase as the preferred composition and rheology for the
326   basement rocks.

327   3.2. Salt detachment

328   The viscous “salt” layer in our experiments decouples the cover sediments from the basement
329   and causes rapid propagation of deformation in the cover towards the distal pinch out of the salt
330   layer in the foreland during early stages of deformation (Figs 5A-G). To better illustrate this
331   decoupling effect and how its spatial distribution above the basement can influence wedge
332   deformation, we ran three more models with three different MTs but the same diabase rheology
333   and removed the salt between the cover and basement in the hinterland near the indenter
334   (hereafter called partial-salt models, Figs. 3B and 6A-D). After 50km shortening, wedge
335   deformation, topography, and the localization of shear zones are very different in the hinterland
336   in these partial-salt models compared to salt-models, especially for MT = 400°C (c.f., Figs 5E-G
337   and 6B-D). For example, in salt-models with MT = 400°C, the basement is highly deformed into
338   stacks of thrust sheets and the cover sediments are extended and thinned in the hinterland (Fig.
339   5E). In contrast, for the partial-salt model with the same MT, the cover sediments do not localize
340   many shear zones and are less uplifted in the hinterland near the indenter (Fig. 6B). This
341   indicates that the basement and cover deformation is strongly affected by the salt detachment
342   near the indenter in these “cold” models (MT = 400°C). However, the foreland deformation in
343   both the cover sediments and basement is very similar in both salt-models and partial-salt
344   models with equal MTs, where pop-up structures above the salt layer are similarly developed
345   (Figs 5E-G and 6B-D). By increasing the MT, the basement deforms in a more ductile manner
346   and deformation is less localized near the indenter. As a result, due to the lower degree of
347   deformation of the hinterland basement in hotter experiments, the cover deformation is similar in
348   the salt- and partial-salt models (Figs 5F and G and 6C and D). Boundary effects near the
349   indenter also contribute to the local model deformation, but as model results show (Figs 5E-G
350   and 6B-D), the influence of the salt detachment near the indenter is the more dominant effect
351   and completely changes the mechanism of deformation especially in cold models.

352   In the following section, we present models with no salt detachment and discuss the role of
353   coupling between cover and basement deformation in the presence of pre-existing weak zones
354   (faults) in the basement.

355   3.3. Pre-existing basement faults and Moho temperature

  9  
356   In order to study the influence of basement faults during thick-skinned deformation, we ran three
357   partial-salt experiments (with three different MTs) containing three basement faults located in
358   the same position as the initial basement steps inherited from continental rifting (Figs. 3 and 6E-
359   H). After 50 km shortening, the partial-salt models with basement faults are strongly sensitive to
360   the temperature in the basement and deform differently from the models without any basement
361   faults (c.f., Figs. 6B-D and 6F-H). In the MT = 400°C experiment, the frontal part deforms
362   similarly to the partial-salt model without any basement faults because there is little ductile flow
363   of lower crust (Figs. 6B and 6F). In this model, deformation mainly occurs in the cover
364   sediments and basement is mainly involved near the indenter where the first pre-existing
365   basement fault shows the most reactivation (Fig. 6F). By increasing the Moho temperature (i.e.
366   MT=500°C and MT=600°C), the basement faults far from the indenter are reactivated and take
367   up significant displacement (Figs. 6G and 6H). In these “hotter” models the basement is folded
368   and pre-existing faults localize the large amount of deformation in the basement owing to the
369   ductile lower crust that transfers the deformation forward towards the foreland. In these hotter
370   models, larger faulted blocks form in the cover relative to the cold models where small pop-up
371   structures are developed (Figs 6F-H). This implies that in colder salt-based fold and thrust belts
372   the displacement related to reactivation of pre-existing basement faults is greater near the
373   indenter. Increasing the MT increases ductile flow in the basement, and consequently the distal
374   basement faults in the middle and frontal parts of fold and thrust belt are also reactivated.

375   In order to illustrate systematically how pre-existing basement faults are reactivated sequentially
376   from the hinterland towards the foreland, and how deformation is taken up by displacements
377   along these faults, we changed the spacing of the basement faults to an arbitrary equal distance
378   of 30 km, and ran three more salt models (Figs. 7A-D). The MT was varied from 400°C to 500°C
379   and 600°C and the models were shortened up to 50 km from one side. We clearly observe that
380   by increasing the MT, the extent of basement deformation and basement fault reactivation are
381   increased while the amount of cover deformation is decreased in the hinterland (Figs. 7B-D). In
382   the cold model (MT = 400°C), the basement deformation is mostly localized near the indenter
383   and the amount of displacement along the basement faults decreases towards the foreland (Fig.
384   7B). Imbrication of the basement blocks and gravity gliding of the cover sediments above the
385   salt layer in the cold experiment introduce significant extension in the cover (see Section 3.6). In
386   contrast, in models with hotter MTs (500°C and 600°C) the cover sediments above the
387   basement faults are less deformed. This happens because the basement blocks can easily
388   rotate and displace along the pre-existing basement faults and take up more deformation than
389   the cover sediments. Basement blocks rotate more in models with hotter MTs due to the
390   increased ductility of the lower crust (Figs. 7A-D). Consequently, the salt detachment is
391   segmented into triangular salt zones and salt flows towards the hinterland.

392   The rotation of basement blocks in the hotter models steepens the dip of the pre-existing
393   basement faults. In these hotter models, all pre-existing faults are reactivated as blind faults that
394   do not cut through the cover, which is decoupled by the weak salt detachment.

395  

  10  
396   3.4. Pre-existing basement faults and salt detachment

397  

398   To emphasize the effect of salt on deformation decoupling we ran three experiments without a
399   salt detachment but containing five equally spaced basement faults (Figs. 7E-H). The MT was
400   varied as before and the models were shortened by 50 km from one side. In these experiments
401   and for all MTs the pre-existing basement faults were all reactivated during shortening and they
402   cut through the cover units to emerge to the surface (Figs. 7F-H). Pop-up structures were
403   localized in the hanging walls of pre-existing basement faults and developed more in colder
404   models (i.e., MT = 400°C and 500°C; Figs. 7E-G). In the hottest model (MT = 600°C), the
405   basement blocks display the greatest amount of clockwise rotation (about 20° clockwise relative
406   to about 15° in salt present models measured from dip change in basement faults) and pop-up
407   structures did not develop. The deformation mechanism is significantly different in these no-salt
408   models compared to models with salt (Figs. 7A-H). Deformation is mainly localized across half
409   of the model in the no-salt models, whereas in models with salt deformation is distributed over a
410   wider area and cover sediments are deformed far into the foreland. Due to coupling of the
411   basement and cover in the hotter no-salt models, the cover units rotate as a consequence of
412   basement block rotation (Figs. 7G and H). In contrast, the cover units in hotter salt-present
413   models dip slightly towards the foreland.

414   3.5. Deformation localization and distribution

415  

416   We illustrate strain-rate localization and distribution in different stages of model shortening by
417   including one more partial-salt model with five equally spaced (30 km) pre-existing basement
418   faults and MT = 400°C (Fig. 8). After 2 km shortening, weak zones, which coincide with pre-
419   existing faults and the viscous basal detachment, record relatively higher strain-rates than other
420   deformation zones in the cover and the basement (Fig. 8B). Resistance along the no-slip
421   boundary in the bottom of the model also accommodates higher strain-rates. At the 2 km
422   shortening stage, we observe a high strain-rate zone associated with the pre-existing basement
423   fault located closest to the indenter where salt is missing, that extends all the way from the
424   basement to the cover. This basement fault has therefore localized a relatively high strain rate
425   and propagated upward into sediments in the no-salt zone. The second and the third pre-
426   existing faults closest to the indenter also localize high strain rates. At this early stage,
427   deformation in the cover sediments extends to about 240 km away from the indenter whereas
428   basement deformation is confined to about 100 km from the indenter. This shows that
429   deformation is transferred quickly forward to the distal end of salt layer and indicates that
430   deformation does not take place simultaneously in the cover and basement. In the distal part of
431   the system, the involvement of the basement is preceded by a phase during which only the
432   cover is deformed (thin-skinned phase).

  11  
433   By further shortening to 8.1 km and 16.2 km, pop-up structures develop in the cover units above
434   the salt detachment. The steps in the basement (Figs. 3B, 8B and C) and the distal end of the
435   viscous layer initiate relatively higher strain-rate zones resulting in shear zones and pop-up
436   structures in the cover at early stages of shortening. The relatively hotter lower crust below 30
437   km also records accommodation and transfer of the high strain-rates zones in the first half of the
438   model. The strain rate plots also illustrate that the higher strain-rate zones in the lower, ductile
439   part of the experiments (below 30km depth) are linked by higher strain rate zones that are
440   coincident with pre-existing faults in the upper crust (Fig. 8).

441   We observe that after 24.3 km of shortening (Fig. 7E) the zone near the indenter does not
442   record high strain rates, indicating that the first basement fault becomes inactive and the whole
443   block near the indenter is pushed forward as an almost rigid block. Moving away from the
444   indenter, basement faults are reactivated sequentially with progressive shortening and as
445   horizontal stress is transferred forwards. With further shortening pre-existing weak zones in the
446   basement are preferred zones of localized shear strain accumulation and only one major back
447   thrust is formed in the basement at a late stage (Figs. 8G and H, 50 km).

448   After 24.3 km shortening (Figs. 8E-H), higher strain rates in the cover are partly associated with
449   extension in the cover due to the gravity gliding above the salt detachment above and after the
450   2nd basement fault. In next section, we explain in detail how gravity gliding increases the amount
451   of cover shortening relative to the basement in the salt-based models.

452   3.6. Shortening velocities

453   All experiments were shortened at a constant horizontal velocity of 8 mm/yr. In general, the local
454   shortening velocity within the models decreases from the hinterland towards the foreland but
455   also increases in cover sediments in the salt-present models. To illustrate these velocity
456   variations, we plot in Figure 9 the horizontal velocities of Eulerian grid points for three different
457   models: no-salt, partial-salt and salt models, each having five pre-existing basement faults and
458   MT = 400°C at 36 km depth. In the no-salt model, horizontal velocities decrease gradually from
459   the hinterland towards the foreland and cover and basement rocks shorten simultaneously at
460   almost the same rate. As stated before, in the no-salt models, the basement and cover are
461   coupled, and shortening is taken up by formation of structures in the cover and basement.
462   Shortening velocities gradually decrease in partial-salt and salt models from the indenter
463   towards the frontal part of the model, but because of the imbrication of basement blocks and
464   uplift of the cover units, the sediments above the salt glide, due to gravity, towards the foreland,
465   resulting in higher velocities. Gravity gliding is defined as downslope movement of a rock mass
466   above a weak detachment surface or zone (Schultza-Ela 2001). The gravity gliding in our
467   models locally increases horizontal velocity so that horizontal shortening rates can be higher in
468   the cover than in the basement (Fig. 9). In these models, horizontal velocities reach up to 16
469   mm/yr, almost double the indenter velocity, and in salt and partial-salt models their distribution is
470   heterogeneous (Fig. 9). In the salt and partial-salt models, the cover units deform faster and
471   they are decoupled from the basement.

  12  
472   4. Discussion and Conclusions

473   4.1. Implications for the Zagros

474   The geometrical, mechanical and thermal parameters (i.e. cover thickness, Moho depth, salt
475   distribution, Moho temperature, shortening rate, total shortening, etc.) are different throughout
476   the crust underling the Zagros, which must change the geometry, kinematics and dynamics of
477   deformational structures across and along the belt (Fig. 1)(e.g. Sherkati and Letouzey 2004,
478   Jahani et al. 2009, Mouthereau et al. 2012). Therefore, we avoid selecting any specific model to
479   represent deformation in the Zagros. Rather, we discuss possible applications of our model to
480   the Zagros fold and thrust belt in order to highlight processes that have influenced the evolution
481   of this active orogen.

482   From our simplified modelling results (e.g. Figs. 6-8), we observe both imbrication of the
483   basement that is decoupled from the cover (Molinaro et al. 2005; Mouthereau et al., 2006 and
484   2007; Sherkati, et al., 2006) and large displacements on pre-existing basement faults that cut
485   through the cover, as proposed by Blanc et al. (2003) and Alavi (2007). Our modelling results
486   find that salt distribution and geothermal gradient are key factors for controlling the crustal-scale
487   deformation of the Zagros. The presence of a relatively thick (1-2 km) salt layer at the
488   basement-cover interface together with a hot geotherm (higher Moho temperature) can prevent
489   reactivated basement faults from propagating into the cover units. Such hidden (“blind”) faults
490   have been discussed, for example, by Berberian et al. (1995) and Bahroudi and Talbot (2003)
491   and they are very important for earthquake studies of the Zagros (e.g. Nissen et al. 2011).
492   Moreover, future studies of these hidden basement faults might aid in improving balanced cross
493   sections and determining the total shortening of the cover and basement across the Zagros
494   more reliably. However, recent research by Mouthereau et al. (2006) explored differential
495   topographic uplift due to displacements on some of these basement faults in the Fars region
496   (Fig. 1). This means that although basement faults can be hidden beneath cover units, their
497   reactivation can be observed in surface topography data.

498   The wider extent of cover deformation in salt-based models is in agreement with previous
499   research (Davis and Engelder 1985; Cotton and Koyi 2000; Bahroudi and Koyi 2003;
500   Nilforoushan and Koyi 2007; Nilfouroushan et al. 2012) and indicates that salt distribution has a
501   major influence on the structural and final wedge geometry of fold and thrust belts like the
502   Zagros. Different distributions of salt beneath the cover sediments in the northwest and
503   southwest regions of the Zagros can partly explain faster GPS-based shortening rates observed
504   in the southwest compared to the northwest (Hessami et al., 2006; Nilforoushan and Koyi, 2007;
505   Walpersdorf et al., 2006).

506   Our modelling results also find that the rate of shortening of the cover and the basement can
507   vary considerably. In cold models (MT=400°C), imbrication of basement blocks occurs in the
508   hinterland near the indenter, which can cause uplift of the cover units and consequently their
509   gravity gliding above a relatively thick salt layer. In the Zagros, with a salt detachment thickness
510   of 1-2 km, our modelling results support Molinaro et al.ʼs (2005) proposal that more shortening

  13  
511   has occurred in the cover than the basement if a cool geotherm is assumed (Fig. 9). Therefore,
512   different amounts of shortening and styles of deformation of cover and basement rocks in salt-
513   based fold and thrust belts like the Zagros can be expected (e.g. Molinaro et al., 2005). Molinaro
514   et al. (2005), however, suggested that multiple phases of deformation occurred in the Zagros in
515   which thin-skinned cover deformation started first, subsequently followed by thick-skinned
516   deformation expressed as out-of sequence faulting in the cover and reactivation of basement
517   faults. Similarly, our model results also indicate that in the distal part of the system, the
518   involvement of the basement is preceded by a phase during which only the cover is deformed
519   (thin-skinned phase). This supports the kinematic scenario proposed by Molinaro et al. (2005) or
520   Sherkati et al. (2006) and contrasts Mouthereau et al. 's (2006), who suggested that the
521   basement deformation is activated early, even at the deformation front. Further investigation
522   using thermo-mechanical models can potentially resolve the issue of multi-phase shortening in
523   the Zagros.

524  

525   4.2. Application to other fold and thrust belts (the Jura Mountains)

526   Although we selected our model parameters for the Zagros, the models presented here also
527   have implications for other mountain belts that are tectonically similar to the Zagros fold and
528   thrust belt. For example, the Jura Mountains are a salt-based fold and thrust belt formed over a
529   younger and hotter basement (Sommaruga 1997 and 1999 Mosar 1999). The Jura Mountains
530   and the Swiss molasse basin represent the youngest deformation zone of the northwestern Alps
531   (Sommaruga 1999). Here, the Mesozoic and Cenozoic cover units were deformed above a
532   weak basal detachment comprising Triassic evaporites (Smith et al. 2003). The thickness of
533   evaporates reaches 1km, decreasing toward the frontal part of the orogenic wedge (Sommaruga
534   1999). The temperature at the brittle-ductile transition zone (BDT) is estimated to be around
535   450°C, basement faults extend to 15-20 km depth, and the Moho depth is around 25km (Mosar
536   1999). Deformation in the Jura Mountains is distributed in several contrasting domains: long
537   wavelength, low amplitude folding in the Molasse basin; thrusting and box folding in the High
538   Jura; a mostly undeformed Jura Plateau; and imbrication in the frontal Faisceau zone (Smith et
539   al. 2003) (See Figure 17a in Smith et al. 2003 or Figure 3 in Sommaruga 1999). The mostly
540   undeformed Jura Plateau is comparable to the results of our experiments. For example,
541   experiments 5B and 5C, which have a similar thermal signature to the Jura mountains, are
542   characterized by a less-deformed wide plateau in the cover sediments in hinterland. In these
543   models, cover deformation is mainly observed in the foreland, near the salt pinch out. Compared
544   to our experiments, salt detachment and hot basement in the Jura fold and thrust belt probably
545   contributes to formation of a less-deformed plateau.

546  

547   4.3. Conclusions

  14  
548   A series of 2-D thermo-mechanical numerical experiments that focus on the Zagros fold and
549   thrust belt, evaluated the possible interaction between pre-existing faults in Precambrian
550   crystalline basement and its sedimentary cover containing a 1-2 km thick intervening layer of
551   weak Hormuz salt. The results find that the degree to which pre-existing basement faults are
552   reactivated is correlated to temperature-dependent ductile flow of the lower crust. A cooler lower
553   crust prevents the transfer of deformation in the basement towards the foreland and only
554   reactivates the basement faults in the hinterland. In relatively warmer models, the lower crust
555   deforms by ductile flow, allowing the basement blocks to rotate and segment the salt layer. In
556   general, salt-based experiments with and without pre-existing basement faults suggest that a
557   cold rheology model simulates better the present structure of the Zagros, in which many
558   detachment folds and thrust faults are observed in the cover. In hotter models with and without
559   pre-existing basement faults, the cover is much less deformed owing to lower-crustal ductile
560   flow and most of the deformation occurs in the basement by folding, thrusting, or displacements
561   along pre-existing faults. However, itʼs worth noting that other factors, including the geometry of
562   pre-existing faults and the magnitude of the imposed strain rates, are likely also important for
563   studies of hidden basement faults beneath a salt-based fold and thrust belt, and should be
564   included in future investigations.

565   The presence of a salt detachment layer near the indenter favors the uplift of basement blocks,
566   resulting in a large amount of cover extension due to gravity gliding, which in turn drives
567   shortening in the foreland. Our results indicate that the amount and style of tectonic deformation
568   in the cover and basement and the degree of decoupling between them are strongly governed
569   by the presence and distribution of the salt detachment in the Zagros.

570   The thermal and mechanical parameters and the crustal configuration we employed for our
571   numerical modeling were selected to study systematically the thick-skinned deformation of an
572   idealized salt-based fold and thrust belt like the Zagros. We did not attempt to make our
573   experiments simulate fully the complex tectonic evolution of the Zagros itself. However, the
574   results provide insights into “Zagros-like” thick-skinned deformation and are a step further to
575   understanding the interaction of cover and basement rocks by including a salt detachment,
576   temperature effects, and pre-existing basement faults.

577  

578   Acknowledgements

579   Research Council of Sweden (VR) funds FN and HK. GMT free software was used for making
580   Figure 1. RP and ARC were funded by Natural Sciences and Engineering Research Council of
581   Canada (NSERC) Discovery Grants. We are grateful for detailed and constructive reviews of
582   Jürgen Adam and Dominique Frizon de Lamotte. We also acknowledge Onno Oncken and
583   Paola Vannucchi for handling our manuscript.

584  

585  

  15  
586   References

587   Allen, M., J. Jackson, and R. Walker (2004), Late Cenozoic reorganization of the Arabia‐Eurasia
588   collision and the comparison of short‐term and long‐term deformation rates, Tectonics, 23,
589   TC2008, doi:10.1029/ 2003TC001530.

590   Allen, M. B., C. Saville, E. J.-P. Blanc, M. Talebian, and E. Nissen (2013), Orogenic plateau
591   growth: Expansion of the Turkish-Iranian Plateau across the Zagros fold-and-thrust belt,
592   Tectonics, 32, doi:10.1002/tect.20025.

593   Alavi, M. (2007), Structures of the Zagros fold-thrust belt in Iran. American Journal of Science
594   307, 1064–95.

595   Bahroudi, A., Koyi, H. A., (2003), The effect of spatial distribution of Hormuz salt on deformation
596   style in the Zagros fold and thrust belt: an analogue modeling approach. Journal of the
597   Geological Society, London 160, 719-733.

598   Bahroudi, A., Talbot, C. J. (2003), The configuration of the basement beneath the Zagros Basin.
599   Journal of Petroleum Geology 26, 257–82.

600   Barr, T. D., Dahlen, F. A., (1989), Brittle frictional mountain builiding, 2: thermal structure and
601   heat budget, Journal of geophysical Research, 94, 3923–3947.

602   Beaumont, C., Jamieson, R. A., Nguyen, M. H., Medvedev S., (2004), Crustal channel flows: 1.
603   Numerical models with applications to the tectonics of the Himalayan-Tibetan orogen, Journal of
604   Geophysical Research, 109, doi:10.1029/2003JB002809.

605   Beaumont, C., Jamieson R. A., Butler JP, Warren CJ, (2009), Crustal structure: A key constraint
606   on the mechanism of ultra-high-pressure rock exhumation, Earth and Planetary Science Letters
607   287, 116–129.

608   Berberian, M., (1995), Master ʻblindʼ thrust faults hidden under the Zagros folds: active
609   basement tectonics and surface tectonics surface morphotectonics, Tectonophysics, 241, 193–
610   224.

611   Berberian, M., King, G. C. P., (1981), Towards a paleogeography and tectonic evolution of Iran;
612   Canadian Journal of Earth Sciences, 18, 210-265.

613   Bird, P., (1978), Finite element modeling of the lithospheric deformation: the Zagros collision
614   orogeny. Tectonophysics, 50, 307–336.

615   Buiter, S.J.H., (2012), A review of brittle compressional wedge models, Tectonophysics, 530–
616   531, 1-17, doi:10.1016/j.tecto.2011.12.018.

617   Buiter, S.J.H., Pfiffner O.A., Beaumont C., (2009), Inversion of extensional sedimentary basins:
618   A numerical evaluation of the localisation of shortening, Earth and Planetary Science Letters,
619   288, 492–504.

620   Buiter, S.J.H. and Torsvik T.H., (2007), Horizontal movements in the eastern Barents Sea
621   constrained by numerical models and plate reconstructions, Geophysical Journal International,
622   171, 1376-1389.

  16  
623   Bordenave, M.L., (2008), The origin of the Permo-Triassic gas accumulations in the Iranian
624   Zagros foldbelt and contiguous offshore area: a review of the Palaeozoic petroleum system.
625   Journal of Petroleum Geology, 31, 3–42.

626   Boutelier, D., Oncken O., (2011), 3-D thermo-mechanical laboratory modeling of plate-tectonics:
627   modeling scheme, technique and first experiments, Solid Earth, doi:10.5194/se-2-35-2011.
628   Costa, E., Vendeville B.C., (2002), Experimental insights on the geometry and kinematics of
629   fold-and-thrust belts above weak, viscous evaporitic décollement, Journal of Structural Geology,
630   24, 1729–1739.
631   Brown D., J. Alvarez-Marron, A. Perez-Estaun, V. Puchkov, and C. Ayala, Basement influence
632   on foreland thrust and fold belt development: An example from the southern Urals,
633   Tectonophysics, 308, 459 – 472, 1999.
634   Cotton, J., Koyi, H.A., (2000), Modelling of Thrust Fronts above Ductile and Frictional
635   Décollements; Examples from The Salt Range and Potwar Plateau, Pakistan. Geological
636   Society of America Bull., 112, 351-363.

637   Cristallini E. O., and V. A. Ramos, (2000), Thick-skinned and thin-skinned thrusting in the La
638   Ramada fold and thrust belt: Crustal evolution of the High Andes of San Juan Argentina (32°SL),
639   Tectonophysics, 317, 205 – 235.
640   Cruden, AR, Naseri F., Pysklywec RN, (2006), Surface topography and internal strain variation
641   in wide hot orogens from three-dimensional analogue and two-dimensional numerical vice
642   models; Analogue and numerical modelling of crustal-scale processes, Geological Society
643   Special Publications, 253:79-104.

644   Davis, D.M., Engelder T., (1985), The role of salt in fold-and-thrust belts, Tectonophysics, 119
645   (1-4), 67-88, doi:10.1016/0040-1951(85)90033-2.

646   Davis, D., J. Suppe, and F. A. Dahlen (1983), Mechanics of fold-and-thrust belts and
647   accretionary wedges, J. Geophys. Res., 88, 1153–1172.

648   Davy, P., Cobbold, P.R., (1991), Experiments on shortening of 4-layer model of continental
649   lithosphere. Tectonophysics 188, 1–25.

650   Dechesne, R. G. and Mountjoy, E. W. (1992), Multiple decollements at deep levels of the
651   southern Canadian Rocky Mountain Main Ranges, Alberta and British Columbia. ln: Structural
652   Geology of Fold and Thrust Belts (S. Mitra and G. W. Fisher, eds), John Hopkins University,
653   Baltimore, pp. 225-238.
654  

655   Ellis S., Schreurs G., Panien M., (2004), Comparisons between analogue and numerical models
656   of thrust wedge development, Journal of Structural Geology, Volume 26, 1659-1675.

657   Fullsack, P., (1995), An arbitrary Lagrangian-Eulerian formulation for creeping flows and
658   applications in tectonic models, Geophysical Journal International, 120, 1–23.

659   Frizon de Lamotte, D., Raulin, C., Mouchot, N., Wrobel-Daveau, J.C., Blanpied, C., Ringenbach,
660   J.C., (2011), The southernmost margin of the Tethys realm during the Mesozoic and Cenozoic:
661   initial geometry and timing of the inversion processes. Tectonics, 30, TC3002,
662   doi:10.1029/2010TC002691.

  17  
663   Goetze, C., (1978). The mechanics of creep in Olivine, Phil. Trans. R. Soc. London, 288, 99–
664   119.

665   Graveleau, F., J. Malavieille, and S. Dominguez (2012), Experimental modelling of orogenic
666   wedges: A review, Tectonophysics, 1-66.

667   Gray, R., Pysklywec, R. N., (2010), Geodynamic models of Archean continental collision and the
668   formation of mantle lithosphere keels, Geophysical Research Letters, 37, 4 pp.

669   Gray, R., Pysklywec, R. N., (2012a), Geodynamic models of mature continental collision:
670   Evolution of an orogen from lithospheric subduction to continental retreat/delamination, Journal
671   of Geophysical Research, 117, 14 pp.

672   Gray, R., Pysklywec, R. N., (2012b), Depositional controls on the evolution of the deep
673   lithosphere during continental collision, Geophysical Research Letters, 39, 6 pp.

674   Hansen, F.D., Carter, N.L., (1982), Creep of selected crustal rocks at 1000 MPa, EOS, Trans.
675   Am. geophys. Un., 63, 437.

676   Hardy, S. (2011), Cover deformation above steep, basement normal faults: Insights from 2D
677   discrete element modeling. Marine and Petroleum geology, 28, 966-972.
678  
679   Hardy S, Finch E (2006), Discrete element modelling of the influence of cover strength on
680   basement involved fault-propagation folding. Tectonophysics, 415, 225–238.

681   Hatzfeld, D., Tatar, M., Priestly, K., Ghafory-Ashtiany, M., 2003. Seismological con- straints on
682   the crustal structure beneath the Zagros Mountain belt (Iran). Geophysical Journal International
683   155, 403–410.

684   Haynes, S.J., Mcquillan H., (1974), Evolution of the Zagros Suture Zone, Southern Iran,
685   Geological Society of America Bulletin, 85, 739-744.

686   Hessami, K., Koyi, H.A., Talbot, C.J., (2001), The significance of strike-slip faulting in the
687   basement of the Zagros fold and thrust belt. Journal of Petroleum Geology, 24, 5–28.

688   Hessami, K., Nilforoushan F., Talbot, C.J., (2006), Active deformation within the Zagros
689   Mountains deduced from GPS measurements. Journal of the Geological Society London 163,
690   143-148.

691   Jahani, S., Callot, J.-P., Letouzey, J., Frizon de Lamotte, D., (2009). The eastern termination of
692   the Zagros Fold-and-Thrust Belt, Iran: structures, evolution, and relationships between salt
693   plugs, folding, and faulting. Tectonics 28, TC6004. doi:10.1029/ 2008TC002418.

694   Jarvis, G. T., and D. P. McKenzie (1980), Convection in a compressible fluid with infinite Prandtl
695   number, J. Fluid Mech., 96, 515–583.

696   Kent, P. E., (1979), The emergent Hormuz salt plugs of southern Iran, Journal of Petroleum
697   Geology 2, 117–144.

698   Kley J., C. R. Monaldi, and J. A. (1999) Salfity, Along-strike segmentation of the Andean
699   foreland: Causes and consequences, Tectonophysics, 301, 75 – 94.
700  

  18  
701   Koyi, H.A., (1988), Experimental modeling of the role of gravity and lateral shortening in the
702   Zagros mountain belt, AAPG Bulletin 72, 1381-1394.

703   Koyi, H.A., (1991), Gravity overturn, extension and basement fault activation, Journal of
704   Petroleum Geology, 14, 117-142.

705   Koyi H., Jenyon M.K., Petersen K. (1993), The effect of basement faulting on diapirism, Journal
706   of Petroleum Geology, 16, 285 – 312.

707   Koyi, HA, Hessami K, Teixell A, (2000), Epicentre distribution and magnitude of earthquakes in
708   fold-thrust belts: insights from sandbox modelling. Geophysica Research Letters 27, 273–276.

709   Lacombe O., Mouthereau F., (2002) Basement-involved shortening and deep detachment
710   tectonics in forelands of orogens: Insights from recent collision belts (Taiwan, Western Alps,
711   Pyrenees), Tectonics, 21, 1030, doi: 10.1029/2001TC901018
712   Maggi, A., Jackson J., Priestley K., Baker C., (2000), A re-assessment of focal depth
713   distributions in southern Iran, the Tien Shan and northern India: Do earthquakes really occur in
714   the continental mantle?, Geophysical Journal International, 143, 629–661.

715   Manaman, N. S., Shomali, H. and Koyi, H. (2011), New constraints on upper-mantle S-velocity
716   structure and crustal thickness of the Iranian plateau using partitioned waveform inversion.
717   Geophysical Journal International, 184: 247–267. doi: 10.1111/j.1365-246X.2010.04822.x.

718   Massoli, D., Koyi, H.A., Barchi, M.R., (2006), Structural evolution of a fold and thrust belt
719   generated by multiple decollements: analogue models and natural examples from the Northern
720   Apennines (Italy). Journal of Structural Geology 28, 185-199.

721   McQuarrie, N. (2004), Crustal scale geometry of the Zagros fold–thrust belt, Iran. Journal of
722   Structural Geology 26, 519–35.

723   Mosar J., Present-day and future tectonic underplating in the western Swiss Alps: reconciliation
724   of basement/wrench-faulting and décollement folding of the Jura and Molasse basin in the
725   Alpine foreland, (1999), Earth and Planetary Science Letters, 173, 143-155, ISSN 0012-821X,
726   10.1016/S0012-821X(99)00238-1.
727  
728   Molinaro, M., Leturmy, P., Guezou, J. C., Frizon de Lamotte, D., Eshraghi, S.A., (2005), The
729   structure and kinematics of the south-eastern Zagros foldthrust belt; Iran: from thin-skinned to
730   thick-skinned tectonics. Tectonics, 24, doi:10.1029/2004TC001633.

731   Motiei, H., (1990), The role of diapirism from the stand point of Hydrocarbon reserves in
732   southwest of Iran. Proceeding of Symposium on Diapirism (with special reference to Iran), 23-
733   53.

734   Mouthereau, F., and Lacombe, O., 2006, Inversion of the Paleogene Chinese continental
735   margin and thick-skinned deformation in the Western Foreland of Taiwan: Journal of Structural
736   Geology, 28, 1977-1993.

737   Mouthereau, F., Lacombe O., Meyer B., (2006), The Zagros folded belt (Fars, Iran): constraints
738   from topography and critical wedge modeling, Geophysical Journal International 165, 336–356.

  19  
739   Mouthereau, F., Lacombe, O., Verges, J., (2012), Building the Zagros collisional orogen: Timing,
740   strain distribution and the dynamics of Arabia/Eurasia plate convergence, Tectonophysics 532-
741   535, 27-60.

742   Mouthereau, F., J. Tensi, N. Bellahsen, O. Lacombe, T. De Boisgrollier, and S. Kargar (2007),
743   Tertiary sequence of deformation in a thin-skinned/thick-skinned collision belt: The Zagros
744   Folded Belt (Fars, Iran), Tectonics, 26, TC5006, doi:10.1029/2007TC002098.

745   Nankali H.R., (2012), Brittle-Ductile transition zone, case study (Zagros Iran), Journal of Asian
746   Earth Sciences, doi: http://dx.doi.org/10.1016/j.jseaes.2012.07.011.

747   Nilforoushan, F., Vernant, P., Masson, F. et al. (2003), GPS network monitors the Arabia–
748   Eurasia collision deformation in Iran, Journal of Geodesy, 77, 411–422.

749   Nilforoushan, F., Koyi, A.H., (2007), Displacement fields and finite strains in a sandbox model
750   simulating a fold-thrust-belt, Geophysical Journal International 169, 1341-1355,
751   doi:10.1111/j.1365-246X.2007.03341.x.

752   Nilforoushan, F., Koyi H., Swantesson J., Talbot CJ., (2008), Effect of basal friction on surface
753   and volumetric strain in models of convergent settings measured by laser scanner, Journal of
754   Structural Geology 30, 366-379.

755   Nilfouroushan, F., Pysklywec, R. and Cruden, A., (2012), Sensitivity analysis of numerical
756   scaled models of fold-and-thrust belts to granular material cohesion variation and comparison
757   with analog experiments, Tectonophysics, doi:10.1016/j.tecto.2011.06.022.

758   Nissen, E., Tatar, M., Jackson, J. A. and Allen, M. B., (2011), New views on earthquake faulting
759   in the Zagros fold-and-thrust belt of Iran. Geophysical Journal International, 186: 928–
760   944,doi: 10.1111/j.1365- 246X.2011.05119.x.

761   Orbell G., (1977), The revised geothermal gradient map of SW Iran and its application, Iranian
762   Oil Operating Company, Technical note 19/1977, 5 p.
763  

764   Oveisi B., Lavé J., Peter van der Beek, Carcaillet J., Benedetti L., Aubourg C., (2009), Thick-
765   and thin-skinned deformation rates in the central Zagros simple folded zone (Iran) indicated by
766   displacement of geomorphic surfaces, Geophysical Journal International, 176, 627-654.

767   Oveisi, B., Lavé, J., Van der Beek, P. A. (2007), Rates and processes of active folding
768   evidenced by Pleistocene terraces at the central Zagros front, Iran. In: Lacombe, O., Lavé, J.,
769   Roure, F. & Vergés, J. (eds.) Thrust Belts and Foreland Basins. Springer, Berlin, 267-287.

770   Paul, A., Kaviani, A., Hatzfeld, D., Vergne, J., Mokhtari, M., (2006), Seismological evidence for
771   crustal-scale thrusting in the Zagros Mountain belt (Iran). Geophysical Journal International 166,
772   227–237.

773   Paul, A., Hatzfeld, D., Kaviani, A., Tatar, M., Pequegnat, C., (2010), Seismic Imaging of the
774   Lithospheric Structure of the Zagros Mountain Belt (Iran), 330. Geological Society Special
775   Publications, London, pp. 5–18.

776   Pysklywec, R. N., (2006). Surface erosion control on the evolution of the deep lithosphere,
777   Geology, 34, 225-228.

  20  
778   Pysklywec, R.N., Beaumont C., (2004), Intraplate tectonics: Feedback between radioactive
779   thermal weakening and crustal deformation driven by mantle lithosphere instabilities, Earth and
780   Planetary Science Letters, 221, 275-292.

781   Pysklywec, R.N., Beaumont B., Fullsack P., (2000), Modeling the behavior of the continental
782   mantle lithosphere during plate convergence, Geology, 28, 655–658.

783   Pysklywec, R.N., Shahnas, M.H., (2003), Time-dependent surface topography in a coupled
784   crust–mantle convection model. Geophys. J. Int. 154, 268–278.

785   Pysklywec, R.N., Cruden, A.R., (2004), Coupled crust–mantle dynamics and intraplate
786   tectonics: two-dimensional numerical and three-dimensional analogue modeling. G3:
787   Geochemistry, Geophysics, Geosystems 5, 1–22.

788   Rotstein, Y., Schaming M., (2004), Seismic reflection evidence for thick-skinned tectonics in the
789   northern Jura, Terra Nova, 16, 250–256.

790   Ruh, J. B., B. J. P. Kaus, and J. P. Burg, (2012), Numerical investigation of deformation
791   mechanics in fold-and-thrust belts: Influence of rheology of single and multiple décollements,
792   Tectonics, 31, TC3005, doi:10.1029/2011TC003047.

793   Schultz-Ela, D.D., (2001), Excursus on gravity gliding and gravity spreading, Journal of
794   Structural Geology 23, 725-731.

795   Schedl, A., Wiltschko D.V., (1987), Possible effects of pre-existing basement topography on
796   thrust fault ramping, Journal of Structural Geology, 9, 1029-1037.

797   Sherkati, S., Letouzey, J., (2004), Variation of structural style and basin evolution in the central
798   Zagros (Izeh zone and Dezful Embayment), Iran, Mar. Petrol. Geol., 21, 535–554.

799   Sherkati, S., Letouzey, J., Frizon De Lamotte, D. (2006), Central Zagros fold-thrust belt (Iran):
800   new insights from seismic data, field observation, and sandbox modeling. Tectonics, 25,
801   TC4007, doi:10.1029/2004TC001766.

802   Smit, J.H.W., Brun, J.P., and Sokoutis, D., (2003), Deformation of brittle-ductile thrust wedges in
803   experiments and nature: Journal of Geophysical Research, 108, no. B10, 2480, doi:
804   10.1029/2002JB002190.

805   Sommaruga, A., (1997), Geology of the Central Jura and the Molasse Basin: new insight into an
806   evaporite-based foreland fold and thrust belt, Mém. Soc. Neuchâtel. Sci. Nat., 12 , p. 176.
807  
808   Sommaruga A., Decollement tectonics in the Jura foreland fold-and-thrust belt (1999), Mar. Pet.
809   Geol. 16 111–134.
810  

811   Stockmal, G.S., Beaumont, C., Nguyen, M., Lee, B., (2007), Mechanics of thin-skinned fold-
812   and-thrust belts: insights from numerical models. In: Sears, J.W., Harms, T.A., Evenchick, C.A.
813   (Eds.), Whence the Mountains? Inquiries Into the Evolution of Orogenic Systems: A Volume in
814   Honor of Raymond A. Price: Geological Society of America Special Paper, 433, pp. 63–98.
815   doi:10.1130/2007.2433(04).

  21  
816   Snyder, D.B., Barazangi, M., (1986), Deep crustal structure and flexure of the Arabian plate
817   beneath the Zagros collisional mountain belt as inferred from gravity observations, Tectonics, 5,
818   361–373.

819   Talbot, C. J., Alavi, M., (1996), The past of a future syntaxis across the Zagros. In: G. I. Alsop,
820   D., J. Blundell, and I. Davison, (Eds.), Salt Tectonics. Spec. Publ. Geol. Soc. London, 100, 89-
821   109.

822   Talebian, M., Jackson, J.A., (2004), A reappraisal of earthquake focal mechanisms and active
823   shortening in the Zagros Mountains of Iran, Geophysical Journal International, 156, 506–526.

824   Tatar, M., Hatzfeld, D., Ghafori-Ashtiany, M., (2004), Tectonics of the Central Zagros (Iran)
825   deduced from microearthquakes seismicity. Geophysical Journal International, 156, 255–266.
826  
827   Tavakoli, F., A. Walpersdorf, C. Authemayou, H. R. Nankali, D. Hatzfeld, M. Tatar, Y.
828   Djamour, F. Nilforoushan, and N. Cotte (2008), Distribution of the right-lateral strike-slip motion
829   from the Main Recent Fault to the Kazerun Fault System (Zagros, Iran): Evidence from present-
830   day GPS velocities, Earth Planet Sc Lett, 275(3-4), 342-347.
831  

832   Verges J., Saura E., Casciello E., Fernandez M., Villasenor A., Jimenez-Munt I., Garcia-
833   Castellanos D. (2011), Crustal scale cross sections across the NW Zagros belt; implications for
834   the Arabian margin reconstruction (in The Zagros; geodynamics and overall structure),
835   Geological Magazine, 148, 739-761.

836   Vernant, P., J. Chéry, (2006), Mechanical modelling of oblique convergence in the Zagros, Iran,
837   Geophysical Journal International, 165, 991-1002.

838   Vernant, P., F. Nilforoushan, D. Hatzfeld, M. Abbassi, C. Vigny, F. Masson, H. Nankali, J.


839   Martinod, A. Ashtiani, R. Bayer, F. Tavakoli, and J. Chéry, (2004), Contemporary Crustal
840   Deformation and Plate Kinematics in Middle East Constrained by GPS measurements in Iran
841   and Northern Oman, Geophysical Journal International, 157, 381-398.

842  
843   Walpersdorf´A., Hatzfeld D., Nankali H., Tavakoli F., Nilforoushan F., Tatar M., Vernant P.,
844   Chery J., Masson F., (2006), Difference in the GPS deformation pattern of North and Central
845   Zagros (Iran), Geophysical Journal International, 167, 1077-1088.
846  
847   Wilks, K.R., Carter, N.L., (1990), Rheology of some continental lower crust, Tectonophysics,
848   182, 57–77.

849   Yaminifard F., Hassanpour Sedghi M., Gholamzadeh A., Tatar M., Hessami K., (2012a), Active
850   faulting of the southeastern-most Zagros (Iran): Microearthquake seismicity and crustal
851   structure, Journal of Geodynamics, 55, 56–65.

852   Yaminifard F., Tatar M., Hessami K., Gholamzadeh A., Bergman E. A., (2012b), Aftershock
853   analysis of the 2005 November 27 (Mw 5.8) Qeshm Island earthquake (Zagros-Iran): Triggering
854   of strike-slip faults at the basement, Journal of Geodynamics, doi:10.1016/j.jog.2012.04.005.

  22  
855   Yamato P., Kaus B.J.P., Mouthereau F. and Castelltort S. (2011), Dynamic constraints on the
856   crustal-scale rheology of the Zagros fold belt, Iran, Geology, v.39, p. 815-818,
857   doi:10.1130/G32136.

858  

859   Figure captions

860   Fig.1. A) Shaded relief map of the Zagros with overlaid simplified tectonic structures and GPS
861   velocity vectors relative to Arabia (modified after Hessami et al. 2006). The inset shows the
862   shaded relief map of Iran with overlaid earthquake distribution from ISC catalogue (magnitude >
863   4, during 1973-2012). MFF, Mountain Front Fault; HZF, High Zagros Fault; MZT, Main Zagros
864   Thrust. The dotted lines show the approximate location of three cross-sections modified from
865   previous studies: B) from Allen et al. 2013, C) and D) from Sherkati et al. 2006. The sections in
866   different locations of Zagros illustrate that the salt distributions and cover and basement
867   deformation are different along and across the belt. As shown in these sections, basement
868   faults dip to the north and they root to about 20km depth.

869   Fig. 2 Histogram shows the centroid depths of teleseismically earthquakes in the Simply Folded
870   Belt of the Zagros (shown in Fig. 1) (modified after Nissen et al. 2011).

871   Fig. 3. Pre-deformation setups for salt (top) and partial-salt (bottom) models. The top surface is
872   a free surface and shortening is imposed by movement of a rigid indenter from the right hand
873   side. The Moho temperature (MT) at 36 km depth is varied in different models from 400°C to
874   600°C. The stepped basement, inherited from rifting episode, is composed of materials with
875   temperature-dependent power-law creep rheologies and is overlain by weak Newtonian viscous
876   salt of variable thickness and frictional-plastic cover sediments.

877   Fig. 4. Strength envelopes (compressional differential stress vs. depth) calculated for a thrust
878   fault regime for a brittle upper crust and for temperature-dependent dry diabase and quartz
879   diorite rheologies using Moho temperatures of 400°C, 500°C and 600°C at 36km depth.
880   Parameters are summarized in Table 2.

881   Fig. 5. A: initial setup of the experiments using quartz diorite and diabase rheologies to model
882   the basement rocks and lower crust. B-D (quartz diorite rheology) and E-G (with diabase
883   rheology) are deformed models with different MTs of 400°C, 500°C and 600°C after 50 km
884   shortening. Axes are in km. Small-scale strength envelopes (Fig. 4) are used to show the BDT
885   depth for each MT in frontal part of the models.

886   Fig. 6. A and E: initial setup of the partial-salt models without (A) and with (E) pre-existing
887   basement faults; B-D and F-H: deformed models with different MTs of 400°C, 500°C and 600°C
888   after 50 km shortening. Axes are in km. Small-scale strength envelopes are used to show the
889   BDT depth for each MT in frontal part of the models.

890   Fig. 7. A and E: initial setup of the models with (A) and without (E) a salt detachment layer and
891   five pre-existing equally spaced basement faults. B-D and F-H: deformed models with different
892   MTs of 400°C, 500°C and 600°C at 36km depth after 50 km shortening. Axes are in km. Small-
893   scale strength envelopes are used to show the BDT depth for each MT in frontal part of the
894   models.

  23  
895   Fig. 8. A: initial setup of a partial-salt model with a salt detachment layer and five pre-existing
896   basement faults. H: deformed partial-salt model after 50 km shortening. B-G: plots of logarithmic
897   strain-rate (s-1) after 2, 8.1, 16.2, 24.3, 34.5 and 50 km of shortening of the model shown in A.
898   Axes are in km. The color scale for logarithmic strain rates is the same for all images. Small-
899   scale strength envelopes are used to show the BDT depth for each MT in frontal part of the
900   models.

901   Fig. 9: Interpolated instantaneous horizontal velocities imposed by the movement of the indenter
902   at a rate of 8 mm/yr for three different experiments with no salt (Fig. 7E), partial-salt (Fig. 8) and
903   salt (Fig. 7A). The horizontal velocities gradually decrease from the hinterland towards the
904   foreland in all models. In the salt and partial-salt models, additional movements due to gravity
905   gliding of sediments above the salt detachment result in higher translation velocities in the cover
906   compared to the basement.

  24  
46˚ 50˚ 54˚ 58˚ 62˚
a)
Eurasia
40˚
34˚
Talesh
Kope Dagh
. 1d Alborz
36˚
Fig Sa Central Iran
na
nd Central Iran
aj-
Sir
jan Sistan
32˚
M Za
32˚ Dezful
ZT
(su zo gr
os
ne
tu
re 28˚
)
MMakran

c
akran

.1
Arabia

Fig
HZ Hi 24˚
F gh
Za
30˚ gr .1
b
os i g
F

Fars

28˚ Simply folded belt


Pe
r sia
1 cm/yr nG MFF
ul
f

26˚
48˚ 50˚ 52˚ 54˚ 56˚ 58˚
b)
Persian Gulf Simply folded belt High Zagros Sanandaj-
Sirjan zone
HZF Zagros suture
salt detachment
SW NE

Moho Ductile layer

c) salt detachment
SW NE

Basement Ductile layer


salt detachment NE
d) SW

Basement Ductile layer

Fig. 1
number of earthquakes
0 5 10 15
Simply folded belt

depth (km)
10

15

20 reverse faulting
strike-slip

Fig. 2
A
no salt salt detachment rigid
T=0 °C indenter
8 km cover sediments v=8 mm/yr
28 km basement 24.5 km
steps
0 300 km
TMoho= 400-600°C
0.5 km salt 1.5 km salt 2.5 km salt

B
no salt salt detachment no salt rigid
T=0 °C indenter
8 km cover sediments v=8 mm/yr
28 km basement 24.5 km
steps
0 300 km
TMoho= 400-600 ° C 0.5 km salt 1.5 km salt

Fig. 3
0

cover sediments
5

weak salt detachment


10

brittle
15 quartz-diorite
TMoho=600° diabase
Depth (km)

20 TMoho=500°
° basement
T Moho=600
25 TMoho=400°
°
T Moho=500
30
°
T Moho=400
35 Moho

40
0 200 400 600 800 1000
Compressional differential stress (MPa)

Fig. 4
no salt salt detachment

TMoho=400°, 500°, 600° 50 100 150 200 250 300

0
diff. stress
br i

10 50 km
ttle

20
B
30
ductile
40
km

0
br i

10
ttle

20
ductile
C
30
40

0
bri

10
ttle

20 D
30
ductile
40

0
diff. stress
50 km
10
bri

E
ttle

20
30
40 ductile

0
10
bri
tt

20
F
le

30
ductile
40

0
10
bri
tt

20
G
le

30 ductile
40

Fig. 5
no salt salt detachment no salt

TMoho=400°, 500°, 600° 50 100 150 200 250 300

0
diff. stress
50 km
10
br i
ttle

20
B
30
40 ductile
km
0
10
br i
ttle

20 C
30
ductile
40

0
10
bri
ttle

20
D
30 ductile
40

no salt pre-existing faults (60°) salt detachment no salt

TMoho=400°, 500°, 600° 50 100 150 200 250 300

diff. stress
0 50 km
10
bri

F
ttle

20
30
40 ductile
0
10
bri
ttle

20
G
30
ductile
40

0
10
bri
ttle

20
H
30 ductile
40

Fig. 6
no salt salt detachment pre-existing faults (60°)

TMoho=400°, 500°, 600° 50 100 150 200 250 300

0
diff. stress
10
50 km
br i

B
ttle

20
30
40 ductile
km

0
10
br i

C
ttle

20
30
ductile
40

0
10
bri
ttle

20 D
30 ductile
40

no salt detachment pre-existing faults (60°)

TMoho=400°, 500°, 600° 50 100 150 200 250 300

0
diff. stress
10 50 km
bri

F
ttle

20
30
40 ductile

0
10
bri
tt

20 G
le

30
ductile
40

0
10
bri
ttle

20
H
30 ductile
40

Fig. 7
A no salt salt detachment pre-existing faults no salt
0 diff. stress
10 br i
ttle
20
30
40 ductile T
km Moho
=400° 50 100 150 200 250 300
−13 −14 −15 −16
salt detachment no salt
0
B
10
br i

2 km
ttle

20
30
40 ductile
50 100 150 200 250 300

0
C
10 8.1 km
br i
ttle

20
30
40 ductile 50 100 150 200 250

0 D
10 16.2 km
bri
ttle

20
30
40 ductile
50 100 150 200 250

0
E
10 24.3 km
bri
ttle

20
30
40 ductile
50 100 150 200 250

0 F
10 34.5 km
bri
tt

20
le

30
40 ductile
50 100 150 200 250

0 G
10
50 km
bri
tt

20
le

30
40 ductile
50 100 150 200 250

0
H
10 50 km
bri
ttle

20
30
40 ductile
50 100 150 200 250
Fig. 8
15

10

0
no salt detachment
horizontal velocities (mm/yr)

8 mm/yr

50 km
shortening
50 100 150 200 250

cover deforming above salt detachment


no salt
no salt gravity gliding

8 mm/yr

50 km
shortening
50 100 150 200 250

cover deforming above salt detachment


no salt gravity gliding

0 8 mm/yr

18 50 km
shortening
36 km 50 100 150 200 250

Fig. 9
Table1. List of models.
Salt No. of pre-
Moho temp
Basement detachment existing
Model at 36 km Remark Figures
rheology* extension basement
depth
(km) faults
1 QD 240 0 400° 5B
2 QD 240 0 500° 5C
3 QD 240 0 600° 5D
Ref.
4 D 240 0 400° 5E
model
5 D 240 0 500° 5F
6 D 240 0 600° 5G
7 D 180 0 400° 6B
8 D 180 0 500° 6C
9 D 180 0 600° 6D
10 D 180 3 400° 6F
11 D 180 3 500° 6G
12 D 180 3 600° 6H
13 D 240 5 400° 7B
14 D 240 5 500° 7C
15 D 240 5 600° 7D
16 D 0 5 400° No salt 7F
17 D 0 5 500° No salt 7G
18 D 0 5 600° No salt 7H
19 D 180 5 400° Strain-rate 8

* QD= quartz-diorite, D=diabase

1
Table2. Model and material parameters.
Geometry and kinematics:
Model dimensions (km) 300x36
Eulerian elements 601*121
Lagrangian tracking points 1801*361
Shortening rate (mm/yr) 8
Strain rate (/s) 10-15
Mechanical properties of
Cover Basement Indenter Salta
materials
Angle of internal friction (phi) 15°-7.5° 20°-10° 30°
Cohesion (MPa) 10 10 10
Pore fluid factor (Lambda) 0 0 0
-3
Density (kg m ) 2600 2900 2900 2200
Newtonian Viscosity (Pa s) 1018
Basement rock creep parametersa : quartz diorite diabase
Material constant A (Pa-n s-1) 1.2x10-16 6.31x10-20
Activation energy Q (kJ mol−1) 212 276
Power-law exponent n 2.4 3.05
Thermal parameters:
Heat capacity (m2 s−2 K−1) 750
−1 −1
Thermal conductivity (W m K ) 2.25
Thermal expansivity (K-1) 2x10-5
Heat production (Wm-3) 0
-2
Vertical heat flux (W m ) 0.03375

a) from Mouthereau et al. (2006) and references therein.

También podría gustarte