Você está na página 1de 50

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  

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


33   Basement faults, Zagros fold and thrust belt

34  

35   1- Introduction

  1  
36   Several studies have physically (Koyi 1997; Mulugeta 1988; Graveleau et al., 2012) and

37   numerically (Ellis et al., 2004; Stockmal et al., 2007; Yamato et al., 2011; Buiter, 2012)

38   simulated thin-skinned deformation at convergent plate boundaries and illustrated how different

39   rheological and mechanical parameters change the mechanism and style of deformation of

40   sedimentary cover rocks. Some studies have gone further and modeled the cover deformation

41   above basement normal faults (e.g. Schedl and Wiltschko 1987; Koyi et al. 1991 and 1993;

42   Hardy, 2011) and others have investigated the influence of cover strength on basement-involved

43   fault propagation folding in cover sediments (e.g. Hardy and Finch, 2006). A few studies have

44   modeled thick-skinned deformation in which both brittle cover sediments and deeper brittle-

45   ductile basement rocks are involved (e.g. Barr and Dahlen 1989; Buiter and Torsvik 2007).

46   However, scaled laboratory modeling of thick-skinned deformation is not easy and more

47   sophisticated materials and setups are required to handle temperature-dependent ductile

48   rheologies at depth (Davy and Cobbold 1991; Boutelier and Oncken 2011). Thermal-mechanical

49   numerical experiments are therefore favored for modeling of thick-skinned deformation involving

50   both brittle upper crust and ductile lower crustal rheologies (e.g. Bird 1978; Buiter et al. 2009). In

51   thick-skinned deformation, weak zones such as pre-existing basement faults and their

52   reactivation by inversion tectonics (Lacombe and Mouthereau, 2002), mechanically weak layers

53   such as salt or shale within the cover sediments or above the basement acting as weak

54   detachments (Yamato et al. 2011), and the depth- and temperature-dependent viscosity of the

55   mid- to lower crust, are important parameters in crustal shortening studies (Buiter et al. 2009).

56   Natural examples of basement-involved fold and thrust belts include the Andes (Kley et al.,

57   1999; Cristallini and Ramos, 2000; Zapata and Almendinger, 1996), Urals (Brown et al., 1999),

58   Taiwan (Mouthereau and Lacombe 2006), Rockies (Dechesne and Mountjoy, 1992), the Alps

59   (Rostein and Schaming 2004) and the Zagros (Jackson, 1980; Berberian, 1995; Sherkati et al.

  2  
60   2004; Molinaro et al. 2005; Mouthereau et al. 2007). Some fold and thrust belts contain single or

61   multiple weak detachments (salt or shale) such as the Zagros, SW Iran (Davis and Engelder

62   1985; Talbot and Alvai 1996; Koyi et al. 2003; Sherkati et al., 2005 and 2006), the Northern

63   Apennines, Italy (Massoli et al., 2006), the southern Canadian Rocky Mountains (Stockmal et

64   al., 2007) and the Jura of the Alps (Davis and Engelder 1985). Mechanically weak layers at the

65   basement-cover interface or between stratigraphic levels of fold and thrust belts influence the

66   deformation style and kinematics of an orogenic system (Davis and Engelder 1985; Koyi 1988;

67   Cotton and Koyi 2000; Costa and Vendeville 2002; Nilfouroushan et al. 2012; Ruh et al. 2012).

68   In addition, pre-existing faults in the upper brittle crust, acting as weak zones in the basement,

69   can be reactivated and change the localization and distribution of deformation in overlying

70   sedimentary rocks. It is currently not clear how effectively ductile flow of lower crust can

71   reactivate pre-existing basement faults in the upper brittle crust when there is a layer of salt

72   between the cover and basement rocks.

73   Using a series of 2-D thermal-mechanical finite element models we explore how thick-skinned

74   deformation and the interactions between basement and cover rocks are influenced by different

75   parameters such as the geothermal gradient, expressed as Moho temperature, and the

76   presence of weak basal detachments like salt in the basal succession of the sediment cover,

77   and weak zones in the basement (e.g., pre-existing normal faults). Owing to the inherent

78   approximation of numerical modeling to the complex real Earth, we do not aim to model exactly

79   the history and present-day deformation of any orogenic system. Rather, we relate our modeling

80   approach to nature by considering the available thermal-mechanical parameters of the Zagros

81   fold and thrust belt, as an opportunity to examine thermal-mechanical rheological and structural

82   interactions with constraints from an active orogen. Although the experiments focus on the

83   Zagros fold and thrust belt, the results can be applied to other fold and thrust belts with similar

84   tectonic configurations.

  3  
85   1.1 The Zagros fold and thrust belt

86   The Zagros fold and thrust belt (Fig. 1A) is the result of active convergence between the rifted

87   continental margin of the Arabian plate and the Iranian continental blocks following the closure

88   of the Neo-Tethys Ocean during the Late Eocene (Frizon de Lamotte et al., 2011). The ongoing

89   deformation of this convergence is not distributed homogenously in Iran and is mainly taken up

90   by mountain belts like the Zagros. Central Iran is undergoing relatively little deformation and

91   acts as the rigid backstop to the Zagros (Vernant and Chery, 2006). The present-day active

92   deformation within the Zagros has been defined by several GPS measurement campaigns (e.g.

93   Nilforoushan et al., 2003; Vernant et al., 2004; Hessami et al., 2006; Walpersdorf et al., 2006;

94   Tavakoli et al., 2008). These studies show that the cover sequence of the Zagros is shortening

95   relatively slowly (5±3 mm/yr) above a high-frictional basal detachment in the NW Zagros,

96   whereas in the southeast of the Zagros the shortening rate is higher (8±3 mm/yr)(Fig. 1). Total

97   shortening since ~5 Ma (Blanc et al. 2003; Allen et al. 2004) in the Zagros is estimated to range

98   between 45 and 65 km (Blanc et al., 2003; McQuarrie, 2003; Oveisi et al., 2007), consistent with

99   GPS-derived shortening rates (Hessami et al., 2006; Walpersdorf et al., 2006; Tavakoli et al.,

100   2008).

101   Based on stratigraphic data the total thickness of the cover sediments in the Zagros region

102   varies between 7 km in the Fars area to 9 km in the Dezful area (Fig. 1) (Alavi, 2004). The

103   southeastern cover sediments overlie a 1-2 km thick layer of weak Hormuz (Neoproterozoic–

104   Cambrian) salt that partially decouples the Phanerozoic cover from its Precambrian crystalline

105   basement (Koyi 1988; Talbot and Alavi 1996; Bahroudi and Koyi 2003). The thick Hormuz salt

106   layer, not only acts as an efficient detachment, but also by feeding salt diapirs, changes the

107   deformation mechanism within the cover units mainly in the eastern Zagros. The numerous

108   emerged or buried salt diapirs rising from this salt detachment have influenced the shape,

  4  
109   localization, propagation and orientation of the folds in the eastern Zagros (Talbot and Alavi

110   1996; Jahani et al. 2009).

111   By opening of the Neo-Tethys ocean during Permian-Early Triassic (Berberian and King 1981;

112   Verges et al. 2011) a passive margin and several extensional faults formed in the Arabian basin

113   that were then covered by sediments (Berberian and King 1981; Bahroudi and Talbot 2003).

114   The exact locations of these pre-existing rift-related extensional faults which have been inverted

115   during the collision stage are not clear but their existence and approximate locations have been

116   inferred from analysis of surface geomorphologic features, topographical sections and

117   earthquake spatial distribution (Berberian 1995; Mouthereau et al., 2006; Alavi 2007). The

118   vertical distribution of earthquakes in the Zagros ranges from 4 to 30 km in depth with the

119   majority between 7-20 km (Maggi et al., 2000; Tatar, et al., 2004; Nissen et al., 2011;

120   Yaminfard, 2012a and b). This indicates that faults are active in both, the cover and basement

121   sequence, suggesting that the shortening of the Zagros fold and thrust belt is not taken up only

122   by the cover sediments but also by the basement (Jackson 1980; Ni and Barazangi 1986;

123   Berberian 1995; Hessami et al., 2001; Talebian and Jackson, 2004; Oveisi et al., 2009). Further,

124   the spatial distribution of earthquakes is limited to the upper 20 km of the crust and implies that

125   the Brittle Ductile Transition (BDT) zone is located at a depth of around 20 km or deeper and

126   faulting is limited to the upper 20km crust of the Zagros (Figs. 1 and 2) (Nissen et al., 2011).

127   The Moho depth beneath the Zagros inferred from geophysical measurements using receiver

128   functions is estimated to be between 40-50 km (Fig. 1) (Hatzfeld et al., 2003, Paul et al 2006,

129   2010, Manaman et al., 2011; Yaminfard et al., 2012). The hinterland-dipping Moho (β=0.5°)

130   makes the crust slightly thicker under the Sanandaj-Sirjan zone relative to the foreland in the

131   simply folded zone (Fig. 1).

132   Geothermal gradient contours in Motiei (1990) and Bordenave (2008), both based on well data

133   from Orbell (1977), indicate a variation in geothermal gradient across and along the belt from

  5  
134   10°C/km to 28°C/km. The increasing geothermal gradients across the belt from the High Zagros

135   to Persian Gulf were partly assigned to tectonically thickened crust near the suture zone (Bird

136   1978). The anomalies in the contours of geothermal gradients along the belt were correlated

137   with north-south trending reactivated old basement faults in the Zagros (Bahroudi and Talbot

138   (2003). For geodynamic modeling of the Zagros, Mouthereau et al., (2006) used a geothermal

139   gradient of 10-15°C/km consistent with a Moho temperature (MT) of 450-675°C. The average

140   surface heat flow of 40 mWm-2 used by Bird (1978) for thermal-mechanical modeling of the

141   Zagros also corresponds to a MT of about 500°C at a Moho depth of 40 km.

142   Balanced cross-sections (e.g. Sherkati and Letouzey 2004) and analysis of topographic profiles

143   across the Zagros (Mouthereau et al., 2006) also show that basement faulting is required to

144   explain the present-day topography. However, in the Zagros, where a relatively thick salt layer

145   covers active basement faults, seismic reflection data fail to image basement structures (Blanc

146   et al., 2003; Alavi, 2004; McQuarrie 2004; Sherkati and Letouzey 2004; Paul et al., 2006) (Fig.

147   1). Moreover, due to the presence of this thick salt detachment, the seismicity has a diffuse

148   pattern (Fig. 1)(Koyi et al., 2000; Nissen et al 2011) and therefore locating seismogenic

149   basement faults is elusive (Berberian 1995).

150  

151   2. Numerical experiments

152  

153   In our experiments, we used a two-dimensional, thermal-mechanical finite element code

154   (SOPALE) which can model high finite strain based on an Arbitrary Lagrangian Eulerian (ALE)

155   method (Fullsack, 1995). This code has been extensively used in a range of different

156   geodynamic modeling applications (Pysklywec and Shahnas, 2003; Beaumont et al., 2004;

157   Pysklywec and Cruden 2004; Cruden et al., 2006; Buiter and Torsvik, 2007; Stockmal et al.,

  6  
158   2007; Beaumont et al 2009; Buiter et al., 2009; Gray and Pysklywec, 2010; Nilfouroushan et al.,

159   2012).

160  

161   In the ALE numerical technique, SOPALE simultaneously uses Eulerian and Lagrangian grids:

162   Finite element computations are performed on a Eulerian grid whose elements are only

163   stretched vertically to accommodate the evolution of topography on the free upper surface; the

164   fully deforming Lagrangian grid tracks the migrating interfaces and material properties. Each of

165   these grids is made up of initially rectangular elements.

166   The models assume incompressibility of materials. While studies show that compressibility of

167   rock may have an effect for deep mantle convection (e.g., Jarvis and Mckenzie, 1980), at the

168   crustal scale the approximation of incompressibility will not influence the behaviour of the

169   models. The governing equations for the models are

170  

171   ∇. ! = 0, (1)

172   ∇. σ!" + +!" = 0, (2)

!"
173   !!! + !. ∇! = !∇! ! + ! , (3)
!"

174   !(!) = !! (1 − ! ! − !! ), (4)

175  

176   where !, σ!" , !, !  , !! , !, !, !  and  t    are velocity, stress tensor, density, gravitational acceleration,

177   specific heat capacity, temperature, thermal conductivity, volumetric rate of internal heat

178   production and time respectively. The other variables, !, !! and !! are thermal expansivity,

179   reference material density and reference temperature respectively.

180   Similar to Mohr-coulomb failure, the brittle deformation for frictional-plastic materials is specified

181   by a pressure-dependent incompressible Drucker-Prager yield criterion:

  7  
!/!
182   !′ ! = ! 1 − ! sin ! + ! cos ! , (5)

183   where !′! is the second invariant of deviatoric stress, ! is is the pressure, ! is the pore fluid

184   factor, ! is the angle of internal friction and ! is the cohesion. The ! value can decrease

185   linearly in a range of specified strain (frictional weakening).

186   The viscous deformation of materials follows either linear (Newtonian) viscous behavior
!
( )
187   (!′! = 2!!! )
!

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

!
!(!!!) !!! !! ( !!) !!
189   !!! = (3 !! 2 ! )! ! !′!! ! !"# (6)

190   where A is the material constant, n is power-law exponent, !′! is second invariant of the

191   deviatoric strain-rate tensor, Q is thermal activation energy, R is gas constant and T is

192   temperature.

193  

194   2.1 Model setup

195   A series (Table 1) of 2-D thermal-mechanical shortening experiments was run to investigate the

196   thick-skinned deformation of a salt-based fold and thrust belt with pre-existing basement faults.

197   The model parameters (Table 2) were set based on the Zagros as a natural prototype. Our aim

198   is not to simulate the full deformational history of the Zagros and its structure in detail. Instead,

199   we use the best available observational constraints for parameters in the models and assess

200   the interaction between the cover and basement, localization and distribution of deformation,

201   and how variations in the geothermal gradient change the strain distribution when there is a

202   weak salt detachment and pre-existing weak zones present in the basement.

  8  
203   To set up our experiments, we started with a rectangular box of 300 km by 36 km that consisted

204   of 8 km sedimentary rocks overlying a stepped basement (inherited from a rifting episode) and

205   separated by a 0.5 km-2.5 km thick salt layer (Fig. 3A). The arbitrary three basement steps (with

206   heights of 0.5km, 1km and 1km respectively from left to right in Fig. 3) are not considered as

207   faults and they only introduce velocity discontinuities. Since the current Moho depth is ~40-50

208   km in the southeastern part of Zagros (Fig. 1) (Hatzfeld et al., 2003, Paul et al., 2006, 2010,

209   Yaminfard et al. 2012) we considered the initial Moho depth of 36 km (Vergas et al. 2011) to

210   take into account crustal thickening after 50 km shortening in about 5-6 million years (Blanc et

211   al. 2003; McQuarrie, 2003, Oveisi et al. 2007, Verges et al. 2011). As stated before, geophysical

212   studies of the current geometry of the Moho under the Zagros (e.g. Hatzfeld et al. 2003; Paul et

213   al., 2006) indicate a gentle hinterland-dipping geometry (β=0.5°) across the belt (Fig. 1).

214   However, this is only an approximation of the current geometry of the Moho, and the initial Moho

215   geometry at the onset of the Zagros shortening is still debated (Vergas et al. 2011). Therefore,

216   to simplify our models we assumed a typical horizontal Moho (i.e. β=0°). The models were

217   shortened orthogonally and continuously from one side by pushing a strong indenter into the

218   material box. The typical orthogonal indentor setup used in previous analytical, analogue and

219   numerical wedge simulations (e.g. Davis et al. 1983; Buiter and Torsvik 2007; Nilfouroushan et

220   al. 2012; Ruh et al. 2012) simplifies the model setup and the interpretations of the results, and

221   facilitates comparisons between different methods and results. All models were shortened at a

222   rate of 8 mm/yr, a rate deduced from present-day GPS measurements (Fig. 1) and consistent

223   with geological rates (Hessami et al. 2006, Walpersdorf et al. 2006). A high viscosity of the

224   indenter (1030 Pas) relative to the other materials in the box meant that the indenter did not

225   deform as it was pushed into the solution space. The relative shortening between the indenter

226   and the material in the box in our models is similar to the Central-Iranian block (Fig. 1) as a rigid

227   indenter pushing against the southeastern part of Zagros. Central Iran is presently undergoing

  9  
228   relatively little deformation (±2 mm/yr) and can be considered a relatively rigid backstop to the

229   Zagros (e.g. Vernant and Chery, 2006).

230   We used a rectangular node resolution of 601 x 121 (equal to 0.5 km for horizontal and 0.3 km

231   for vertical resolution) for the Eulerian and 1801 x 361 for the Lagrangian grids. The upper

232   surface is a free surface, the sidewalls are free-slip, and the bottom surface, the Moho, is a no-

233   slip boundary (Fig. 2). A no-slip basal boundary assumption implies no horizontal movements

234   occur right at the bottom surface (Moho). Due to the temperature-dependent rheology used to

235   model lower crustal rocks, the ductile behavior of the lower crust permits the materials just

236   above the Moho to deform readily, thereby minimizing the effect of the no-slip lower boundary. A

237   similar no-slip assumption for the Moho boundary has been used in other modeling studies

238   (e.g., Mothereau et al. 2006; Buiter and Torsvik 2007).

239   We set the angle of internal friction to 15° for the overburden sediments and 20° for the

240   basement rocks and let the frictional strength of the brittle crust decrease linearly by 50% across

241   a strain range of 0.5 to 1.5 to approximate material weakening due to, for example, an increase

242   in pore-fluid pressure in nature (e.g. Buiter and Torsvik 2007; Gray and Pysklywec, 2012a).

243   Similar to previous modeling studies (Pysklywec and Beaumont, 2004; Gray and Pysklywec,

244   2012a), we use an "effective angle of internal friction" for ! with pore fluid factor ! = 0. This

245   factors in the pore fluid pressure implicitly with the assumed (lower) angle of internal friction of

246   15°-7.5° for overburden sediments (Table 2).

247   Following previous studies we model the weak Hormuz salt in Zagros by a Newtonian viscous

248   rheology with an effective viscosity of 1018 Pas and density of 2200 kgm-3 (Table 2) (e.g.

249   Mouthereau et al. 2006; Yamato et al. 2011). The extension of the salt detachment is varied in

250   our models to study its effect on orogenic wedge deformation (Table 1). However, we limit the

  10  
251   frontal extension of the salt detachment to a distance of 240km (in initial setups) from the

252   indenter (Fig. 2) to avoid any frontal boundary effect on cover deformation. We used density

253   values of 2600 kg m−3 for the cover sediments and 2900 kg m−3 for the crystalline crust (Snyder

254   and Barazangi 1986; Paul et al., 2006).

255   The composition of the basement of the Zagros is poorly known (e.g., Bahroudi & Talbot, 2003).

256   The only observed basement rocks are blocks of orthogneiss, metasediments, amphibolites and

257   serpentinites intruded by granite, gabbro and basalt brought to the surface in salt diapirs

258   (Haynes and McQuillan, 1974; Kent, 1979). Following Mouthereau et al. (2006), we assumed

259   quartz diorite and diabase to be suitable compositions for the basement rocks of the Zagros

260   basement. Hence, we evaluated the available temperature-dependent power-law creep laws for

261   these compositions in our numerical experiments (Tables 1 and 2). As stated above,

262   earthquakes in the southeastern Zagros occur at depths between 4-30 km with the majority

263   between 7-20 km (Maggi et al., 2000; Talebian and Jackson, 2004, Nissen et al., 2011;

264   Yaminifard et al. 2012a and b), which suggests that ductile deformation should occur below ~20

265   km. To demonstrate the strength profiles that result from using the flow laws for diabase and

266   quartz diorite, we used the same flow laws as Mouthereau et al., (2006) (Fig. 4). The input

267   parameters for the brittle and ductile deformation of the model crust are given in Table 2, using

268   parameters suggested by Vernant and Chery (2006), Mouthereau et al., (2006) and references

269   therein (i.e. Goetze 1978; Hansen and Carter 1982; Wilks and Carter 1990). The selection of

270   diabase and quartz diorite flow laws to represent the rheology of the ductile crust is reasonable

271   as their strength profiles result in a brittle ductile transition (BDT) depth of ≥ 20 km depth for MTs

272   of 400-600°C (Fig. 4).

273   The number, location, and geometry of the basement faults in the Zagros are poorly known,

274   hence we introduce them arbitrarily in our models. In experiments with pre-existing basement

  11  
275   faults, we consider the faults to be 1500 meters wide and dipping at 60° (as such, the faults are

276   resolved by three Eulerian elements, 3x500=1500m). Following Buiter and Torsvik (2007), the

277   faults are filled with a Newtonian material with a viscosity of 1020 Pa s to mimic weak inherited

278   normal faults in the basement. The faults extend down to 20 km depth where the brittle

279   deformation depth has been inferred from earthquakes studies (Tatar et al., 2004; Yaminfard et

280   al., 2012a and b).

281  

282   We ran 19 shortening experiments (Table 2) with or without basement faults, changing the

283   basement rheology, salt distribution and increasing Moho temperature (MT) systematically from

284   400°C to 500°C and 600°C (assuming geothermal gradient of 11°-17°C/km for a 36km thick

285   crust) which covers the reported MT range in the previous studies (Bird 1978; Mouthereau et al.,

286   2006; Vernant and Chery 2006).

287   We simplified our models by ignoring the effect of isostatic adjustment and thermal subsidence

288   of the underlying lithosphere; also, erosional and depositional processes were not included,

289   although these can have an influence on the behavior of such models (Pysklywec 2006; Gray

290   and Pysklywec, 2012b). As described in Buiter and Torsvik (2007), in orogenic wedge models

291   with this type of configuration the ductile/viscous lower crust in the models allows a “simple

292   form” of effective isostatic compensation in the crust to occur. With a deeper (i.e., mantle)

293   isostatic compensation, there may be some modification to details of the structural geometry of

294   the crust, but this is beyond the scope of the modelling code at this scale of lithospheric

295   investigation.

296  

297   3. Results

298   3.1 Rheology of basement rocks

  12  
299   Using a diabase composition and rheology results in a higher effective viscosity lower crust

300   compared to quartz diorite (Fig. 4). The BDT depths are 24, 20 and 16 km for quartz diorite and

301   31, 26 and 22 km for diabase, for MTs of 400°C, 500°C and 600°C, respectively. This means

302   that diabase has a deeper BDT than quartz diorite under the same thermal and mechanical

303   conditions.

304   We tested basement comprising both diabase and quartz diorite in our models and observe

305   composition has a strong influence on deformation behavior (Fig. 5). In the early stages of

306   shortening, deformation in the cover units starts by formation of shear zones in the frontal part of

307   the cover where the salt is pinched out. With further shortening deformation in the cover units

308   propagates backwards (Fig. 5). Due to the shallower BDT depth in experiments using quartz

309   diorite, (Fig. 4), flow in the more ductile lower crust suppresses significant deformation in the

310   cover above the salt detachment (Fig. 5B-D). In all three models using quartz diorite with

311   different MTs, the upper crust deforms similarly and no shear zones (except one near the

312   indentor) are localized in the 200 km long sedimentary cover (Fig. 5B-D). However, the

313   basement is folded in the case of all MTs. In the hottest model, MT = 600°C, the basement is

314   also affected by the load of the especially thickened cover at the second basement step

315   location. Due to the flow of hot ductile lower crust in this model, the upper brittle crust including

316   the cover subside significantly at this location (Fig. 5D).

317   The cover units in models with diabase deformed more, with faulting and development of pop-up

318   structures especially in the coldest model (MT = 400°C) (Fig. 5E). In the case of the Zagros,

319   with reported earthquakes depths to around 20km, diabase seems to be a better choice for the

320   composition and rheology of the basement rocks. This is in agreement with a recent study on

321   lithospheric strength of the Zagros (Nankali 2012) that suggests a relatively cold geothermal

322   gradient and a diabase or granulitic composition with a BDT located around 21-28 km We

  13  
323   therefore focus our modeling using diabase as the preferred composition and rheology for the

324   basement rocks.

325   3.2 Salt detachment

326   The viscous “salt” layer in our experiments decouples the cover sediments from the basement

327   and causes rapid propagation of deformation in the cover towards the distal pinch out of the salt

328   layer in the foreland during early stages of deformation (Figs 5A-G). To better illustrate this

329   decoupling effect and how its spatial distribution above the basement can influence wedge

330   deformation, we ran three more models with three different MTs but the same diabase rheology

331   and removed the salt between the cover and basement in the hinterland near the indenter

332   (hereafter called partial-salt models, Figs. 3B and 6A-D). After 50km shortening, wedge

333   deformation, topography, and the localization of shear zones are very different in the hinterland

334   in these partial-salt models compared to salt-models, especially for MT = 400°C (c.f., Figs 5E-G

335   and 6B-D). For example, in salt-models with MT = 400°C, the basement is highly deformed into

336   stacks of thrust sheets and the cover sediments are extended and thinned in the hinterland (Fig.

337   5E). In contrast, for the partial-salt model with the same MT, the cover sediments do not localize

338   many shear zones and are less uplifted in the hinterland near the indenter (Fig. 6B). This

339   indicates that the basement and cover deformation is strongly affected by the salt detachment

340   near the indenter in these “cold” models (MT = 400°C). However, the foreland deformation in

341   both the cover sediments and basement is very similar in both salt-models and partial-salt

342   models with equal MTs, where pop-up structures above the salt layer are similarly developed

343   (Figs 5E-G and 6B-D). By increasing the MT, the basement deforms in a more ductile manner

344   and deformation is less localized near the indenter. As a result, due to the lower degree of

345   deformation of the hinterland basement in hotter experiments, the cover deformation is similar in

346   the salt- and partial-salt models (Figs 5F and G and 6C and D). Boundary effects near the

  14  
347   indenter also contribute to the local model deformation, but as model results show (Figs 5E-G

348   and 6B-D), the influence of the salt detachment near the indenter is the more dominant effect

349   and completely changes the mechanism of deformation especially in cold models.

350   In the following section, we present models with no salt detachment and discuss the role of

351   coupling between cover and basement deformation in the presence of pre-existing weak zones

352   (faults) in the basement.

353   3.3 Pre-existing basement faults and Moho temperature

354   In order to study the influence of basement faults during thick-skinned deformation, we ran three

355   partial-salt experiments (with three different MTs) containing three basement faults located in

356   the same position as the initial basement steps inherited from continental rifting (Figs. 3 and 6E-

357   H). After 50 km shortening, the partial-salt models with basement faults are strongly sensitive to

358   the temperature in the basement and deform differently from the models without any basement

359   faults (c.f., Figs. 6B-D and 6F-H). In the MT = 400°C experiment, the frontal part deforms

360   similarly to the partial-salt model without any basement faults because there is little ductile flow

361   of lower crust (Figs. 6B and 6F). In this model, deformation mainly occurs in the cover

362   sediments and basement is mainly involved near the indenter where the first pre-existing

363   basement fault shows the most reactivation (Fig. 6F). By increasing the Moho temperature (i.e.

364   MT=500°C and MT=600°C), the basement faults far from the indenter are reactivated and take

365   up significant displacement (Figs. 6G and 6H). In these “hotter” models the basement is folded

366   and pre-existing faults localize the large amount of deformation in the basement owing to the

367   ductile lower crust that transfers the deformation forward towards the foreland. In these hotter

368   models, larger faulted blocks form in the cover relative to the cold models where small pop-up

369   structures are developed (Figs 6F-H). This implies that in colder salt-based fold and thrust belts

  15  
370   the displacement related to reactivation of pre-existing basement faults is greater near the

371   indenter. Increasing the MT increases ductile flow in the basement, and consequently the distal

372   basement faults in the middle and frontal parts of fold and thrust belt are also reactivated.

373   In order to illustrate systematically how pre-existing basement faults are reactivated sequentially

374   from the hinterland towards the foreland, and how deformation is taken up by displacements

375   along these faults, we changed the spacing of the basement faults to an arbitrary equal distance

376   of 30 km, and ran three more salt models (Figs. 7A-D). The MT was varied from 400°C to 500°C

377   and 600°C and the models were shortened up to 50 km from one side. We clearly observe that

378   by increasing the MT, the extent of basement deformation and basement fault reactivation are

379   increased while the amount of cover deformation is decreased in the hinterland (Figs. 7B-D). In

380   the cold model (MT = 400°C), the basement deformation is mostly localized near the indenter

381   and the amount of displacement along the basement faults decreases towards the foreland (Fig.

382   7B). Imbrication of the basement blocks and gravity gliding of the cover sediments above the

383   salt layer in the cold experiment introduce significant extension in the cover (see Section 4.2). In

384   contrast, in models with hotter MTs (500°C and 600°C) the cover sediments above the

385   basement faults are less deformed. This happens because the basement blocks can easily

386   rotate and displace along the pre-existing basement faults and take up more deformation than

387   the cover sediments. Basement blocks rotate more in models with hotter MTs due to the

388   increased ductility of the lower crust (Figs. 7A-D). Consequently, the salt detachment is

389   segmented into triangular salt zones and salt flows towards the hinterland.

390   The rotation of basement blocks in the hotter models steepens the dip of the pre-existing

391   basement faults. In these hotter models, all pre-existing faults are reactivated as blind faults that

392   do not cut through the cover, which is decoupled by the weak salt detachment.

  16  
393  

394   3.4 Pre-existing basement faults and salt detachment

395  

396   To emphasize the effect of salt on deformation decoupling we ran three experiments without a

397   salt detachment but containing five equally spaced basement faults (Figs. 7E-H). The MT was

398   varied as before and the models were shortened by 50 km from one side. In these experiments

399   and for all MTs the pre-existing basement faults were all reactivated during shortening and they

400   cut through the cover units to emerge to the surface (Figs. 7F-H). Pop-up structures were

401   localized in the hanging walls of pre-existing basement faults and developed more in colder

402   models (i.e., MT = 400°C and 500°C; Figs. 7E-G). In the hottest model (MT = 600°C), the

403   basement blocks display the greatest amount of clockwise rotation (about 20° clockwise relative

404   to about 15° in salt present models measured from dip change in basement faults) and pop-up

405   structures did not develop. The deformation mechanism is significantly different in these no-salt

406   models compared to models with salt (Figs. 7A-H). Deformation is mainly localized across half

407   of the model in the no-salt models, whereas in models with salt deformation is distributed over a

408   wider area and cover sediments are deformed far into the foreland. Due to coupling of the

409   basement and cover in the hotter no-salt models, the cover units rotate as a consequence of

410   basement block rotation (Figs. 7G and H). In contrast, the cover units in hotter salt-present

411   models dip slightly towards the foreland.

412   3.5 Deformation localization and distribution

413  

  17  
414   We illustrate strain-rate localization and distribution in different stages of model shortening by

415   including one more partial-salt model with five equally spaced (30 km) pre-existing basement

416   faults and MT = 400°C (Fig. 8). After 2 km shortening, weak zones, which coincide with pre-

417   existing faults and the viscous basal detachment, record relatively higher strain-rates than other

418   deformation zones in the cover and the basement (Fig. 8B). Resistance along the no-slip

419   boundary in the bottom of the model also accommodates higher strain-rates. At the 2 km

420   shortening stage, we observe a high strain-rate zone associated with the pre-existing basement

421   fault located closest to the indenter where salt is missing, that extends all the way from the

422   basement to the cover. This basement fault has therefore localized a relatively high strain rate

423   and propagated upward into sediments in the no-salt zone. The second and the third pre-

424   existing faults closest to the indenter also localize high strain rates. At this early stage,

425   deformation in the cover sediments extends to about 240 km away from the indenter whereas

426   basement deformation is confined to about 100 km from the indenter. This shows that

427   deformation is transferred quickly forward to the distal end of salt layer and indicates that

428   deformation does not take place simultaneously in the cover and basement. In the distal part of

429   the system, the involvement of the basement is preceded by a phase during which only the

430   cover is deformed (thin-skinned phase).

431   By further shortening to 8.1 km and 16.2 km, pop-up structures develop in the cover units above

432   the salt detachment. The steps in the basement (Figs. 3B, 8B and C) and the distal end of the

433   viscous layer initiate relatively higher strain-rate zones resulting in shear zones and pop-up

434   structures in the cover at early stages of shortening. The relatively hotter lower crust below 30

435   km also records accommodation and transfer of the high strain-rates zones in the first half of the

436   model. The strain rate plots also illustrate that the higher strain-rate zones in the lower, ductile

  18  
437   part of the experiments (below 30km depth) are linked by higher strain rate zones that are

438   coincident with pre-existing faults in the upper crust (Fig. 8).

439   We observe that after 24.3 km of shortening (Fig. 7E) the zone near the indenter does not

440   record high strain rates, indicating that the first basement fault becomes inactive and the whole

441   block near the indenter is pushed forward as an almost rigid block. Moving away from the

442   indenter, basement faults are reactivated sequentially with progressive shortening and as

443   horizontal stress is transferred forwards. With further shortening pre-existing weak zones in the

444   basement are preferred zones of localized shear strain accumulation and only one major back

445   thrust is formed in the basement at a late stage (Figs. 8G and H, 50 km).

446   After 24.3 km shortening (Figs. 8E-H), higher strain rates in the cover are partly associated with

447   extension in the cover due to the gravity gliding above the salt detachment above and after the

448   2nd basement fault. In next section, we explain in detail how gravity gliding increases the amount

449   of cover shortening relative to the basement in the salt-based models.

450   3.6. Shortening velocities

451   All experiments were shortened at a constant horizontal velocity of 8 mm/yr. In general, the local

452   shortening velocity within the models decreases from the hinterland towards the foreland but

453   also increases in cover sediments in the salt-present models. To illustrate these velocity

454   variations, we plot in Figure 9 the horizontal velocities of Eulerian grid points for three different

455   models: no-salt, partial-salt and salt models, each having five pre-existing basement faults and

456   MT = 400°C at 36 km depth. In the no-salt model, horizontal velocities decrease gradually from

457   the hinterland towards the foreland and cover and basement rocks shorten simultaneously at

458   almost the same rate. As stated before, in the no-salt models, the basement and cover are

459   coupled, and shortening is taken up by formation of structures in the cover and basement.

  19  
460   Shortening velocities gradually decrease in partial-salt and salt models from the indenter

461   towards the frontal part of the model, but because of the imbrication of basement blocks and

462   uplift of the cover units, the sediments above the salt glide, due to gravity, towards the foreland,

463   resulting in higher velocities. Gravity gliding is defined as downslope movement of a rock mass

464   above a weak detachment surface or zone (Schultza-Ela 2001). The gravity gliding in our

465   models locally increases horizontal velocity so that horizontal shortening rates can be higher in

466   the cover than in the basement (Fig. 9). In these models, horizontal velocities reach up to 16

467   mm/yr, almost double the indenter velocity, and in salt and partial-salt models their distribution is

468   heterogeneous (Fig. 9). In the salt and partial-salt models, the cover units deform faster and

469   they are decoupled from the basement.

470   4. Discussion and Conclusions

471   4.1. Implications for the Zagros

472   The geometrical, mechanical and thermal parameters (i.e. cover thickness, Moho depth, salt

473   distribution, Moho temperature, shortening rate, total shortening, etc.) are different throughout

474   the crust underling the Zagros, which must change the geometry, kinematics and dynamics of

475   deformational structures across and along the belt (Fig. 1)(e.g. Sherkati and Letouzey 2004,

476   Jahani et al. 2009, Mouthereau et al. 2012). Therefore, we avoid selecting any specific model to

477   represent deformation in the Zagros. Rather, we discuss possible applications of our model to

478   the Zagros fold and thrust belt in order to highlight processes that have influenced the evolution

479   of this active orogen.

480   From our simplified modelling results (e.g. Figs. 6-8), we observe both imbrication of the

481   basement that is decoupled from the cover (Molinaro et al. 2005; Mouthereau et al., 2006 and

482   2007; Sherkati, et al., 2006) and large displacements on pre-existing basement faults that cut

  20  
483   through the cover, as proposed by Blanc et al. (2003) and Alavi (2007). Our modelling results

484   find that salt distribution and geothermal gradient are key factors for controlling the crustal-scale

485   deformation of the Zagros. The presence of a relatively thick (1-2 km) salt layer at the

486   basement-cover interface together with a hot geotherm (higher Moho temperature) can prevent

487   reactivated basement faults from propagating into the cover units. Such hidden (“blind”) faults

488   have been discussed, for example, by Berberian et al. (1995) and Bahroudi and Talbot (2003)

489   and they are very important for earthquake studies of the Zagros (e.g. Nissen et al. 2011).

490   Moreover, future studies of these hidden basement faults might aid in improving balanced cross

491   sections and determining the total shortening of the cover and basement across the Zagros

492   more reliably. However, recent research by Mouthereau et al. (2006) explored differential

493   topographic uplift due to displacements on some of these basement faults in the Fars region

494   (Fig. 1). This means that although basement faults can be hidden beneath cover units, their

495   reactivation can be observed in surface topography data.

496   The wider extent of cover deformation in salt-based models is in agreement with previous

497   research (Davis and Engelder 1985; Cotton and Koyi 2000; Bahroudi and Koyi 2003;

498   Nilforoushan and Koyi 2007; Nilfouroushan et al. 2012) and indicates that salt distribution has a

499   major influence on the structural and final wedge geometry of fold and thrust belts like the

500   Zagros. Different distributions of salt beneath the cover sediments in the northwest and

501   southwest regions of the Zagros can partly explain faster GPS-based shortening rates observed

502   in the southwest compared to the northwest (Hessami et al., 2006; Nilforoushan and Koyi, 2007;

503   Walpersdorf et al., 2006).

504   Our modelling results also find that the rate of shortening of the cover and the basement can

505   vary considerably. In cold models (MT=400°C), imbrication of basement blocks occurs in the

506   hinterland near the indenter, which can cause uplift of the cover units and consequently their

  21  
507   gravity gliding above a relatively thick salt layer. In the Zagros, with a salt detachment thickness

508   of 1-2 km, our modelling results support Molinaro et al.ʼs (2005) proposal that more shortening

509   has occurred in the cover than the basement if a cool geotherm is assumed (Fig. 9). Therefore,

510   different amounts of shortening and styles of deformation of cover and basement rocks in salt-

511   based fold and thrust belts like the Zagros can be expected (e.g. Molinaro et al., 2005). Molinaro

512   et al. (2005), however, suggested that multiple phases of deformation occurred in the Zagros in

513   which thin-skinned cover deformation started first, subsequently followed by thick-skinned

514   deformation expressed as out-of sequence faulting in the cover and reactivation of basement

515   faults. Similarly, our model results also indicate that in the distal part of the system, the

516   involvement of the basement is preceded by a phase during which only the cover is deformed

517   (thin-skinned phase). This supports the kinematic scenario proposed by Molinaro et al. (2005) or

518   Sherkati et al. (2006) and contrasts Mouthereau et al. 's (2006), who suggested that the

519   basement deformation is activated early, even at the deformation front. Further investigation

520   using thermo-mechanical models can potentially resolve the issue of multi-phase shortening in

521   the Zagros.

522  

523   4.2 Application to other fold and thrust belts (the Jura Mountains)

524   Although we selected our model parameters for the Zagros, the models presented here also

525   have implications for other mountain belts that are tectonically similar to the Zagros fold and

526   thrust belt. For example, the Jura Mountains are a salt-based fold and thrust belt formed over a

527   younger and hotter basement (Sommaruga 1997 and 1999 Mosar 1999). The Jura Mountains

528   and the Swiss molasse basin represent the youngest deformation zone of the northwestern Alps

529   (Sommaruga 1999). Here, the Mesozoic and Cenozoic cover units were deformed above a

  22  
530   weak basal detachment comprising Triassic evaporites (Smith et al. 2003). The thickness of

531   evaporates reaches 1km, decreasing toward the frontal part of the orogenic wedge (Sommaruga

532   1999). The temperature at the brittle-ductile transition zone (BDT) is estimated to be around

533   450°C, basement faults extend to 15-20 km depth, and the Moho depth is around 25km (Mosar

534   1999). Deformation in the Jura Mountains is distributed in several contrasting domains: long

535   wavelength, low amplitude folding in the Molasse basin; thrusting and box folding in the High

536   Jura; a mostly undeformed Jura Plateau; and imbrication in the frontal Faisceau zone (Smith et

537   al. 2003) (See Figure 17a in Smith et al. 2003 or Figure 3 in Sommaruga 1999). The mostly

538   undeformed Jura Plateau is comparable to the results of our experiments. For example,

539   experiments 5B and 5C, which have a similar thermal signature to the Jura mountains, are

540   characterized by a less-deformed wide plateau in the cover sediments in hinterland. In these

541   models, cover deformation is mainly observed in the foreland, near the salt pinch out. Compared

542   to our experiments, salt detachment and hot basement in the Jura fold and thrust belt probably

543   contributes to formation of a less-deformed plateau.

544  

545   4.3 Conclusions

546   A series of 2-D thermo-mechanical numerical experiments that focus on the Zagros fold and

547   thrust belt, evaluated the possible interaction between pre-existing faults in Precambrian

548   crystalline basement and its sedimentary cover containing a 1-2 km thick intervening layer of

549   weak Hormuz salt. The results find that the degree to which pre-existing basement faults are

550   reactivated is correlated to temperature-dependent ductile flow of the lower crust. A cooler lower

551   crust prevents the transfer of deformation in the basement towards the foreland and only

552   reactivates the basement faults in the hinterland. In relatively warmer models, the lower crust

  23  
553   deforms by ductile flow, allowing the basement blocks to rotate and segment the salt layer. In

554   general, salt-based experiments with and without pre-existing basement faults suggest that a

555   cold rheology model simulates better the present structure of the Zagros, in which many

556   detachment folds and thrust faults are observed in the cover. In hotter models with and without

557   pre-existing basement faults, the cover is much less deformed owing to lower-crustal ductile

558   flow and most of the deformation occurs in the basement by folding, thrusting, or displacements

559   along pre-existing faults. However, itʼs worth noting that other factors, including the geometry of

560   pre-existing faults and the magnitude of the imposed strain rates, are likely also important for

561   studies of hidden basement faults beneath a salt-based fold and thrust belt, and should be

562   included in future investigations.

563   The presence of a salt detachment layer near the indenter favors the uplift of basement blocks,

564   resulting in a large amount of cover extension due to gravity gliding, which in turn drives

565   shortening in the foreland. Our results indicate that the amount and style of tectonic deformation

566   in the cover and basement and the degree of decoupling between them are strongly governed

567   by the presence and distribution of the salt detachment in the Zagros.

568   The thermal and mechanical parameters and the crustal configuration we employed for our

569   numerical modeling were selected to study systematically the thick-skinned deformation of an

570   idealized salt-based fold and thrust belt like the Zagros. We did not attempt to make our

571   experiments simulate fully the complex tectonic evolution of the Zagros itself. However, the

572   results provide insights into “Zagros-like” thick-skinned deformation and are a step further to

573   understanding the interaction of cover and basement rocks by including a salt detachment,

574   temperature effects, and pre-existing basement faults.

575  

  24  
576   Acknowledgements

577   Research Council of Sweden (VR) funds FN and HK. GMT free software was used for making

578   Figure 1. RP and ARC were funded by Natural Sciences and Engineering Research Council of

579   Canada (NSERC) Discovery Grants. We are grateful for detailed and constructive reviews of

580   Jürgen Adam and Dominique Frizon de Lamotte. We also acknowledge Onno Oncken and

581   anonymous editor for handling our manuscript.

582  

583   References

584   Allen, M., J. Jackson, and R. Walker (2004), Late Cenozoic reorganization of the Arabia‐Eurasia

585   collision and the comparison of short‐term and long‐term deformation rates, Tectonics, 23,

586   TC2008, doi:10.1029/ 2003TC001530.

587   Allen, M. B., C. Saville, E. J.-P. Blanc, M. Talebian, and E. Nissen (2013), Orogenic plateau

588   growth: Expansion of the Turkish-Iranian Plateau across the Zagros fold-and-thrust belt,

589   Tectonics, 32, doi:10.1002/tect.20025.

590   Alavi, M. (2007), Structures of the Zagros fold-thrust belt in Iran. American Journal of Science

591   307, 1064–95.

592   Bahroudi, A., Koyi, H. A., (2003), The effect of spatial distribution of Hormuz salt on deformation

593   style in the Zagros fold and thrust belt: an analogue modeling approach. Journal of the

594   Geological Society, London 160, 719-733.

595   Bahroudi, A., Talbot, C. J. (2003), The configuration of the basement beneath the Zagros Basin.

596   Journal of Petroleum Geology 26, 257–82.

  25  
597   Barr, T. D., Dahlen, F. A., (1989), Brittle frictional mountain builiding, 2: thermal structure and

598   heat budget, Journal of geophysical Research, 94, 3923–3947.

599   Blanc, E. J.P., Allen, M. B., Inger, S., Hassani, H. (2003), Structural styles in the Zagros Simple

600   Folded Zone, Iran. Journal of the Geological Society, London 160, 401–12.

601   Beaumont, C., Jamieson, R. A., Nguyen, M. H., Lee, B., (2001), Himalayan tectonics explained

602   by extrusion of a low-viscosity crustal channel coupled to focused surface denudation. Nature,

603   414, 738-742.

604   Beaumont, C., Jamieson, R. A., Nguyen, M. H., Medvedev S., (2004), Crustal channel flows: 1.

605   Numerical models with applications to the tectonics of the Himalayan-Tibetan orogen, Journal of

606   Geophysical Research, 109, doi:10.1029/2003JB002809.

607   Beaumont, C., Jamieson R. A., Butler JP, Warren CJ, (2009), Crustal structure: A key constraint

608   on the mechanism of ultra-high-pressure rock exhumation, Earth and Planetary Science Letters

609   287, 116–129.

610   Berberian, M., (1995), Master ʻblindʼ thrust faults hidden under the Zagros folds: active

611   basement tectonics and surface tectonics surface morphotectonics, Tectonophysics, 241, 193–

612   224.

613   Berberian, M., King, G. C. P., (1981), Towards a paleogeography and tectonic evolution of Iran;

614   Canadian Journal of Earth Sciences, 18, 210-265.

615   Bird, P., (1978), Finite element modeling of the lithospheric deformation: the Zagros collision

616   orogeny. Tectonophysics, 50, 307–336.

617   Buiter, S.J.H., (2012), A review of brittle compressional wedge models, Tectonophysics, 530–

618   531, 1-17, doi:10.1016/j.tecto.2011.12.018.

  26  
619   Buiter, S.J.H., Pfiffner O.A., Beaumont C., (2009), Inversion of extensional sedimentary basins:

620   A numerical evaluation of the localisation of shortening, Earth and Planetary Science Letters,

621   288, 492–504.

622   Buiter, S.J.H. and Torsvik T.H., (2007), Horizontal movements in the eastern Barents Sea

623   constrained by numerical models and plate reconstructions, Geophysical Journal International,

624   171, 1376-1389.

625   Buiter, S.J.H., Babeyko, A.Yu., Ellis, S., Gerya, T.V., Kaus, B.J.P., Kellner, A., Schreurs,

626   G.,Yamada,Y., (2006), The numerical sandbox: comparison of model results for a shortening

627   and an extension experiment, in Analogue and Numerical Modelling of Crustal-Scale Processes,

628   253, pp. 29–64, eds Buiter, S.J.H. & Schreurs, G., Geol. Soc., Spec. Publ., London.

629   Boutelier, D., Oncken O., (2011), 3-D thermo-mechanical laboratory modeling of plate-tectonics:

630   modeling scheme, technique and first experiments, Solid Earth, doi:10.5194/se-2-35-2011.

631   Costa, E., Vendeville B.C., (2002), Experimental insights on the geometry and kinematics of

632   fold-and-thrust belts above weak, viscous evaporitic décollement, Journal of Structural Geology,

633   24, 1729–1739.

634   Brown D., J. Alvarez-Marron, A. Perez-Estaun, V. Puchkov, and C. Ayala, Basement influence

635   on foreland thrust and fold belt development: An example from the southern Urals,

636   Tectonophysics, 308, 459 – 472, 1999.

637   Cotton, J., Koyi, H.A., (2000), Modelling of Thrust Fronts above Ductile and Frictional

638   Décollements; Examples from The Salt Range and Potwar Plateau, Pakistan. Geological

639   Society of America Bull., 112, 351-363.

640   Cristallini E. O., and V. A. Ramos, (2000), Thick-skinned and thin-skinned thrusting in the La

641   Ramada fold and thrust belt: Crustal evolution of the High Andes of San Juan Argentina (32°SL),

  27  
642   Tectonophysics, 317, 205 – 235.

643   Cruden, AR, Naseri F., Pysklywec RN, (2006), Surface topography and internal strain variation

644   in wide hot orogens from three-dimensional analogue and two-dimensional numerical vice

645   models; Analogue and numerical modelling of crustal-scale processes, Geological Society

646   Special Publications, 253:79-104.

647   Davis, D.M., Engelder T., (1985), The role of salt in fold-and-thrust belts, Tectonophysics, 119

648   (1-4), 67-88, doi:10.1016/0040-1951(85)90033-2.

649   Davis, D., J. Suppe, and F. A. Dahlen (1983), Mechanics of fold-and-thrust belts and

650   accretionary wedges, J. Geophys. Res., 88, 1153–1172.

651   Davy, P., Cobbold, P.R., (1991), Experiments on shortening of 4-layer model of continental

652   lithosphere. Tectonophysics 188, 1–25.

653   Ellis S., Schreurs G., Panien M., (2004), Comparisons between analogue and numerical models

654   of thrust wedge development, Journal of Structural Geology, Volume 26, 1659-1675.

655   Fullsack, P., (1995), An arbitrary Lagrangian-Eulerian formulation for creeping flows and

656   applications in tectonic models, Geophysical Journal International, 120, 1–23.

657   Frizon de Lamotte, D., Raulin, C., Mouchot, N., Wrobel-Daveau, J.C., Blanpied, C., Ringenbach,

658   J.C., (2011), The southernmost margin of the Tethys realm during the Mesozoic and Cenozoic:

659   initial geometry and timing of the inversion processes. Tectonics, 30, TC3002,

660   doi:10.1029/2010TC002691.

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

662   119.

  28  
663   Gray, R., Pysklywec, R. N., (2010), Geodynamic models of Archean continental collision and the

664   formation of mantle lithosphere keels, Geophysical Research Letters, 37, 4 pp.

665   DOI: 10.1029/2010GL043965

666   Gray, R., Pysklywec, R. N., (2012a), Geodynamic models of mature continental collision:

667   Evolution of an orogen from lithospheric subduction to continental retreat/delamination, Journal

668   of Geophysical Research, 117, 14 pp, DOI: 10.1029/2011JB008692

669   Gray, R., Pysklywec R. N. (2012b), Depositional controls on the evolution of the deep

670   lithosphere during continental collision, Geophysical Research Letters, 39, L11312,

671   doi:10.1029/2012GL051947.

672   Hansen, F.D., Carter, N.L., (1982), Creep of selected crustal rocks at 1000 MPa, EOS, Trans.

673   Am. geophys. Un., 63, 437.

674   Hatzfeld, D., Tatar, M., Priestly, K., Ghafory-Ashtiany, M., 2003. Seismological con- straints on

675   the crustal structure beneath the Zagros Mountain belt (Iran). Geophysical Journal International

676   155, 403–410.

677   Haynes, S.J., Mcquillan H., (1974), Evolution of the Zagros Suture Zone, Southern Iran,

678   Geological Society of America Bulletin, 85, 739-744.

679   Hessami, K., Koyi, H.A., Talbot, C.J., (2001), The significance of strike-slip faulting in the

680   basement of the Zagros fold and thrust belt. Journal of Petroleum Geology, 24, 5–28.

681   Hessami, K., Nilforoushan F., Talbot, C.J., (2006), Active deformation within the Zagros

682   Mountains deduced from GPS measurements. Journal of the Geological Society London 163,

683   143-148.

  29  
684   Jahani, S., Callot, J.-P., Letouzey, J., Frizon de Lamotte, D., (2009). The eastern termination of

685   the Zagros Fold-and-Thrust Belt, Iran: structures, evolution, and relationships between salt

686   plugs, folding, and faulting. Tectonics 28, TC6004. doi:10.1029/ 2008TC002418.

687   Jarvis, G. T., and D. P. McKenzie (1980), Convection in a compressible fluid with infinite Prandtl

688   number, J. Fluid Mech., 96, 515–583.

689   Kent, P. E., (1979), The emergent Hormuz salt plugs of southern Iran, Journal of Petroleum

690   Geology 2, 117–144.

691   Kley J., C. R. Monaldi, and J. A. (1999) Salfity, Along-strike segmentation of the Andean

692   foreland: Causes and consequences, Tectonophysics, 301, 75 – 94.

693  

694   Koyi, H.A., (1988), Experimental modeling of the role of gravity and lateral shortening in the

695   Zagros mountain belt, AAPG Bulletin 72, 1381-1394.

696   Koyi, H.A., (1991), Gravity overturn, extension and basement fault activation, Journal of

697   Petroleum Geology, 14, 117-142.

698   Koyi H., Jenyon M.K., Petersen K. (1993), The effect of basement faulting on diapirism, Journal

699   of Petroleum Geology, 16, 285 – 312.

700   Koyi, H., (1997), Analogue modelling: from qualitative to quantitative technique - A historical

701   outline. Journal of Petroleum Geology, 20(2): 223-238.

702   Koyi, HA, Hessami K, Teixell A, (2000), Epicentre distribution and magnitude of earthquakes in

703   fold-thrust belts: insights from sandbox modelling. Geophysica Research Letters 27, 273–276.

704   Lacombe O., Mouthereau F., (2002) Basement-involved shortening and deep detachment

705   tectonics in forelands of orogens: Insights from recent collision belts (Taiwan, Western Alps,

  30  
706   Pyrenees), Tectonics, 21, 1030, doi: 10.1029/2001TC901018

707   Maggi, A., Jackson J., Priestley K., Baker C., (2000), A re-assessment of focal depth

708   distributions in southern Iran, the Tien Shan and northern India: Do earthquakes really occur in

709   the continental mantle?, Geophysical Journal International, 143, 629–661.

710   Manaman, N. S., Shomali, H. and Koyi, H. (2011), New constraints on upper-mantle S-velocity

711   structure and crustal thickness of the Iranian plateau using partitioned waveform inversion.

712   Geophysical Journal International, 184: 247–267. doi: 10.1111/j.1365-246X.2010.04822.x.

713   McQuarrie, N. (2004), Crustal scale geometry of the Zagros fold–thrust belt, Iran. Journal of

714   Structural Geology 26, 519–35.

715   McQuarrie, N., Stock, J. M., Verdel, C., Wernicke, B. P. (2003), Cenozoic evolution of Neotethys

716   and implications for the causes of plate motions. Geophysical Research Letters 30, 2036,

717   doi:10.1029/2003GL017992.

718   Mosar J., Present-day and future tectonic underplating in the western Swiss Alps: reconciliation

719   of basement/wrench-faulting and décollement folding of the Jura and Molasse basin in the

720   Alpine foreland, (1999), Earth and Planetary Science Letters, 173, 143-155, ISSN 0012-821X,

721   10.1016/S0012-821X(99)00238-1.

722  

723   Molinaro, M., Leturmy, P., Guezou, J. C., Frizon de Lamotte, D., Eshraghi, S.A., (2005), The

724   structure and kinematics of the south-eastern Zagros foldthrust belt; Iran: from thin-skinned to

725   thick-skinned tectonics. Tectonics, 24, doi:10.1029/2004TC001633.

726   Motiei, H., (1990), The role of diapirism from the stand point of Hydrocarbon reserves in

727   southwest of Iran. Proceeding of Symposium on Diapirism (with special reference to Iran), 23-

728   53.

  31  
729   Mouthereau, F., Lacombe O., Meyer B., (2006), The Zagros folded belt (Fars, Iran): constraints

730   from topography and critical wedge modeling, Geophysical Journal International 165, 336–356.

731   Mouthereau, F., Lacombe, O., Verges, J., (2012), Building the Zagros collisional orogen: Timing,

732   strain distribution and the dynamics of Arabia/Eurasia plate convergence, Tectonophysics 532-

733   535, 27-60.

734   Mouthereau, F., J. Tensi, N. Bellahsen, O. Lacombe, T. De Boisgrollier, and S. Kargar (2007),

735   Tertiary sequence of deformation in a thin-skinned/thick-skinned collision belt: The Zagros

736   Folded Belt (Fars, Iran), Tectonics, 26, TC5006, doi:10.1029/2007TC002098.

737   Nankali H.R., (2012), Brittle-Ductile transition zone, case study (Zagros Iran), Journal of Asian

738   Earth Sciences, 59, 156-166, doi: 10.1016/j.jseaes.2012.07.011.

739   Nilforoushan, F., Vernant, P., Masson, F. et al. (2003), GPS network monitors the Arabia–

740   Eurasia collision deformation in Iran, Journal of Geodesy, 77, 411–422.

741   Nilforoushan, F., Koyi, A.H., (2007), Displacement fields and finite strains in a sandbox model

742   simulating a fold-thrust-belt, Geophysical Journal International 169, 1341-1355,

743   doi:10.1111/j.1365-246X.2007.03341.x.

744   Nilforoushan, F., Koyi H., Swantesson J., Talbot CJ., (2008), Effect of basal friction on surface

745   and volumetric strain in models of convergent settings measured by laser scanner, Journal of

746   Structural Geology 30, 366-379.

747   Nilfouroushan, F., Pysklywec, R. and Cruden, A., (2012), Sensitivity analysis of numerical

748   scaled models of fold-and-thrust belts to granular material cohesion variation and comparison

749   with analog experiments, Tectonophysics; Vol 526-529,196-206;

750   doi:10.1016/j.tecto.2011.06.022

  32  
751  

752   Nissen, E., Tatar, M., Jackson, J. A. and Allen, M. B., (2011), New views on earthquake faulting

753   in the Zagros fold-and-thrust belt of Iran. Geophysical Journal International, 186: 928–

754   944,doi: 10.1111/j.1365- 246X.2011.05119.x.

755   Oveisi B., Lavé J., Peter van der Beek, Carcaillet J., Benedetti L., Aubourg C., (2009), Thick-

756   and thin-skinned deformation rates in the central Zagros simple folded zone (Iran) indicated by

757   displacement of geomorphic surfaces, Geophysical Journal International, 176, 627-654.

758   Oveisi, B., Lavé, J., Van der Beek, P. A. (2007), Rates and processes of active folding

759   evidenced by Pleistocene terraces at the central Zagros front, Iran. In: Lacombe, O., Lavé, J.,

760   Roure, F. & Vergés, J. (eds.) Thrust Belts and Foreland Basins. Springer, Berlin, 267-287.

761   Paul, A., Kaviani, A., Hatzfeld, D., Vergne, J., Mokhtari, M., (2006), Seismological evidence for

762   crustal-scale thrusting in the Zagros Mountain belt (Iran). Geophysical Journal International 166,

763   227–237.

764   Paul, A., Hatzfeld, D., Kaviani, A., Tatar, M., Pequegnat, C., (2010), Seismic Imaging of the

765   Lithospheric Structure of the Zagros Mountain Belt (Iran), 330. Geological Society Special

766   Publications, London, pp. 5–18.

767   Pfiffner, O. A., Ellis S., Beaumont C., (2000), Collision tectonics in the Swiss Alps: Insight from

768   geodynamic modeling, Tectonics, 19, 1065–1094, doi:10.1029/2000TC900019.

769   Pysklywec, R. N., (2006). Surface erosion control on the evolution of the deep lithosphere,

770   Geology, 34, 225-228.

  33  
771   Pysklywec, R.N., Beaumont C., (2004), Intraplate tectonics: Feedback between radioactive

772   thermal weakening and crustal deformation driven by mantle lithosphere instabilities, Earth and

773   Planetary Science Letters, 221, 275-292.

774   Pysklywec, R.N., Beaumont B., Fullsack P., (2000), Modeling the behavior of the continental

775   mantle lithosphere during plate convergence, Geology, 28, 655–658.

776   Pysklywec, R.N., Shahnas, M.H., (2003), Time-dependent surface topography in a coupled

777   crust–mantle convection model. Geophys. J. Int. 154, 268–278.

778   Pysklywec, R.N., Cruden, A.R., (2004), Coupled crust–mantle dynamics and intraplate

779   tectonics: two-dimensional numerical and three-dimensional analogue modeling. G3:

780   Geochemistry, Geophysics, Geosystems 5, 1–22.

781   Rotstein, Y., Schaming M., (2004), Seismic reflection evidence for thick-skinned tectonics in the

782   northern Jura, Terra Nova, 16, 250–256.

783   Ruh, J. B., B. J. P. Kaus, and J. P. Burg, (2012), Numerical investigation of deformation

784   mechanics in fold-and-thrust belts: Influence of rheology of single and multiple décollements,

785   Tectonics, 31, TC3005, doi:10.1029/2011TC003047.

786   Schultz-Ela, D.D., (2001), Excursus on gravity gliding and gravity spreading, Journal of

787   Structural Geology 23, 725-731.

788   Schedl, A., Wiltschko D.V., (1987), Possible effects of pre-existing basement topography on

789   thrust fault ramping, Journal of Structural Geology, 9, 1029-1037.

790   Sherkati, S., Letouzey, J., (2004), Variation of structural style and basin evolution in the central

791   Zagros (Izeh zone and Dezful Embayment), Iran, Mar. Petrol. Geol., 21, 535–554.

  34  
792   Sherkati, S., Letouzey, J., Frizon De Lamotte, D. (2006), Central Zagros fold-thrust belt (Iran):

793   new insights from seismic data, field observation, and sandbox modeling. Tectonics, 25,

794   TC4007, doi:10.1029/2004TC001766.

795   Smit, J.H.W., Brun, J.P., and Sokoutis, D., (2003), Deformation of brittle-ductile thrust wedges in

796   experiments and nature: Journal of Geophysical Research, 108, no. B10, 2480, doi:

797   10.1029/2002JB002190.

798   Sommaruga, A., (1997), Geology of the Central Jura and the Molasse Basin: new insight into an

799   evaporite-based foreland fold and thrust belt, Mém. Soc. Neuchâtel. Sci. Nat., 12 , p. 176.

800   Sommaruga A., Decollement tectonics in the Jura foreland fold-and-thrust belt (1999), Mar. Pet.

801   Geol. 16 111–134.

802   Stockmal, G.S., Beaumont, C., Nguyen, M., Lee, B., (2007), Mechanics of thin-skinned fold-

803   and-thrust belts: insights from numerical models. In: Sears, J.W., Harms, T.A., Evenchick, C.A.

804   (Eds.), Whence the Mountains? Inquiries Into the Evolution of Orogenic Systems: A Volume in

805   Honor of Raymond A. Price: Geological Society of America Special Paper, 433, pp. 63–98.

806   doi:10.1130/2007.2433(04).

807   Stöcklin, J., (1968), Structural history and tectonics of Iran: A review, American Association of

808   Petroleum Geologists Bulletin, 52, 1229– 1258.

809   Snyder, D.B., Barazangi, M., (1986), Deep crustal structure and flexure of the Arabian plate

810   beneath the Zagros collisional mountain belt as inferred from gravity observations, Tectonics, 5,

811   361–373.

812   Talbot, C. J., Alavi, M., (1996), The past of a future syntaxis across the Zagros. In: G. I. Alsop,

813   D., J. Blundell, and I. Davison, (Eds.), Salt Tectonics. Spec. Publ. Geol. Soc. London, 100, 89-

814   109.

  35  
815   Talebian, M., Jackson, J.A., (2004), A reappraisal of earthquake focal mechanisms and active

816   shortening in the Zagros Mountains of Iran, Geophysical Journal International, 156, 506–526.

817   Verges J., Saura E., Casciello E., Fernandez M., Villasenor A., Jimenez-Munt I., Garcia-

818   Castellanos D. (2011), Crustal scale cross sections across the NW Zagros belt; implications for

819   the Arabian margin reconstruction (in The Zagros; geodynamics and overall structure),

820   Geological Magazine, 148, 739-761.

821   Vernant, P., J. Chéry, (2006), Mechanical modelling of oblique convergence in the Zagros, Iran,

822   Geophysical Journal International, 165, 991-1002.

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

824   Martinod, A. Ashtiani, R. Bayer, F. Tavakoli, and J. Chéry, (2004), Contemporary Crustal

825   Deformation and Plate Kinematics in Middle East Constrained by GPS measurements in Iran

826   and Northern Oman, Geophysical Journal International, 157, 381-398.

827  
828   Walpersdorf´A., Hatzfeld D., Nankali H., Tavakoli F., Nilforoushan F., Tatar M., Vernant P.,

829   Chery J., Masson F., (2006), Difference in the GPS deformation pattern of North and Central

830   Zagros (Iran), Geophysical Journal International, 167, 1077-1088.

831  

832   Wilks, K.R., Carter, N.L., (1990), Rheology of some continental lower crust, Tectonophysics,

833   182, 57–77.

834   Yaminifard F., Hassanpour Sedghi M., Gholamzadeh A., Tatar M., Hessami K., (2012a), Active

835   faulting of the southeastern-most Zagros (Iran): Microearthquake seismicity and crustal

836   structure, Journal of Geodynamics, 55, 56–65.

  36  
837   Yaminifard F., Tatar M., Hessami K., Gholamzadeh A., Bergman E. A., (2012b), Aftershock

838   analysis of the 2005 November 27 (Mw 5.8) Qeshm Island earthquake (Zagros-Iran): Triggering

839   of strike-slip faults at the basement, Journal of Geodynamics, 61, 138-147,

840   doi:10.1016/j.jog.2012.04.005.

841   Yamato P., Kaus B.J.P., Mouthereau F. and Castelltort S. (2011), Dynamic constraints on the

842   crustal-scale rheology of the Zagros fold belt, Iran, Geology, v.39, p. 815-818,

843   doi:10.1130/G32136.

844  

845   Figure captions

846   Fig.1. A) Shaded relief map of the Zagros with overlaid simplified tectonic structures and GPS

847   velocity vectors relative to Arabia (modified after Hessami et al. 2006). The inset shows the

848   shaded relief map of Iran with overlaid earthquake distribution from ISC catalogue (magnitude >

849   4, during 1973-2012). MFF, Mountain Front Fault; HZF, High Zagros Fault; MZT, Main Zagros

850   Thrust. The dotted lines show the approximate location of three cross-sections modified from

851   previous studies: B) from Allen et al. 2012, C) and D) from Sherkati et al. 2006. The sections in

852   different locations of Zagros illustrate that the salt distributions and cover and basement

853   deformation are different along and across the belt. As shown in these sections, basement

854   faults dip to the north and they root to about 20km depth.

855   Fig. 2 Histogram shows the centroid depths of teleseismically earthquakes in the Simply Folded

856   Belt of the Zagros (shown in Fig. 1) (modified after Nissen et al. 2011).

857   Fig. 3. Pre-deformation setups for salt (top) and partial-salt (bottom) models. The top surface is

858   a free surface and shortening is imposed by movement of a rigid indenter from the right hand

  37  
859   side. The Moho temperature (MT) at 36 km depth is varied in different models from 400°C to

860   600°C. The stepped basement, inherited from rifting episode, is composed of materials with

861   temperature-dependent power-law creep rheologies and is overlain by weak Newtonian viscous

862   salt of variable thickness and frictional-plastic cover sediments.

863   Fig. 4. Strength envelopes (compressional differential stress vs. depth) calculated for a thrust

864   fault regime for a brittle upper crust and for temperature-dependent dry diabase and quartz

865   diorite rheologies using Moho temperatures of 400°C, 500°C and 600°C at 36km depth.

866   Parameters are summarized in Table 2.

867   Fig. 5. A: initial setup of the experiments using quartz diorite and diabase rheologies to model

868   the basement rocks and lower crust. B-D (quartz diorite rheology) and E-G (with diabase

869   rheology) are deformed models with different MTs of 400°C, 500°C and 600°C after 50 km

870   shortening. Axes are in km. Small-scale strength envelopes (Fig. 4) are used to show the BDT

871   depth for each MT in frontal part of the models.

872   Fig. 6. A and E: initial setup of the partial-salt models without (A) and with (E) salt pre-existing

873   basement faults; B-D and F-H: deformed models with different MTs of 400°C, 500°C and 600°C

874   after 50 km shortening. Axes are in km. Small-scale strength envelopes are used to show the

875   BDT depth for each MT in frontal part of the models.

876   Fig. 7. A and E: initial setup of the models with (A) and without (E) a salt detachment layer and

877   five pre-existing equally spaced basement faults. B-D and F-H: deformed models with different

878   MTs of 400°C, 500°C and 600°C at 36km depth after 50 km shortening. Axes are in km. Small-

879   scale strength envelopes are used to show the BDT depth for each MT in frontal part of the

880   models.

  38  
881   Fig. 8. A: initial setup of a partial-salt model with a salt detachment layer and five pre-existing

882   basement faults. H: deformed partial-salt model after 50 km shortening. B-G: plots of logarithmic

883   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.

884   Axes are in km. The color scale for logarithmic strain rates is the same for all images. Small-

885   scale strength envelopes are used to show the BDT depth for each MT in frontal part of the

886   models.

887   Fig. 9: Interpolated instantaneous horizontal velocities imposed by the movement of the indenter

888   at a rate of 8 mm/yr for three different experiments with no salt (Fig. 7E), partial-salt (Fig. 8) and

889   salt (Fig. 7A). The horizontal velocities gradually decrease from the hinterland towards the

890   foreland in all models. In the salt and partial-salt models, additional movements due to gravity

891   gliding of sediments above the salt detachment result in higher translation velocities in the cover

892   compared to the basement.

  39  
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


number of earthquakes
0 5 10 15
Simply folded belt

depth (km)
10

15

20 reverse faulting
strike-slip
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
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)
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
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
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
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
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
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
4 D 240 0 400° Ref. model 5E
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.

Você também pode gostar