*Manuscript Click here to view linked References

1 Mantle convection in the Middle East:

2 Reconciling Afar plume, Arabia collision and Aegean trench rollback

3

4 Claudio Faccenna1, Thorsten W. Becker2, Laurent Jolivet3 and Mehmet Keskin4

5

6 1 LET, Laboratory of Experimental Tectonics, Università Roma Tre, Rome, Italy

7 2 Department of Earth Sciences, University of Southern California, Los Angeles CA 90089,

8 USA

9 3Institut des Sciences de la Terre d'Orléans - Université Orleans

10 4 Dept. of Geological Engineering, University, 34320 Avcılar, Istanbul,

11

12 Abstract

13 The Middle East region represents a key site within the Tethyan domain where continental

14 break-up, collision, backarc extension and escape tectonics are linked together. We perform global

15 mantle circulation computations to test the role of slab pull and mantle upwellings as driving forces

16 for the kinematics of the Arabia-Anatolia-Aegean (AAA) system, evaluating different boundary

17 conditions and mantle density distributions as inferred from seismic tomography or slab models.

18 Model results are compared with geodesy, residual topography, and shear wave splitting

19 observations. The AAA velocity field with respect to Eurasia shows an anti-clockwise toroidal

20 pattern, with increasing velocities toward the Aegean trench. The best match to these crustal

21 motions can be obtained by combining the effect of slab pull exerted in the Aegean with a mantle

22 upwelling underneath Afar and, more generally, with the large-scale flow associated with a whole

23 mantle, Tethyan convection cell. Neogene volcanism for AAA is widespread, not only in the

24 extensional or subduction settings, but also within plates, such as in Syria-Jordan-Israel and in

25 Turkey, with geochemical fingerprints similar of those of the Afar lava. In addition, morphological

26 features show large uplifting domains far from plate boundaries. We speculate that the tectonic

1 27 evolution of AAA is related to the progressive northward entrainment of upwelling mantle material

28 which is itself associated with the establishment of the downwelling branch of a convection cell

29 through the segmented Tethyan slab below the northern Zagros and Bitlis collision zone. The newly

30 established westward flow dragged Anatolia and pushed the Aegean slab south-westward, thus

31 accelerating backarc extension. Our model reconciles Afar plume volcanism, the collision in the

32 Bitlis mountains and northern Zagros, and the rapid increase of Aegean trench rollback in a single

33 coherent frame of large scale mantle convection, initiated during the last ~40 Ma.

34

35

36

37

38 1. Introduction

39 The Arabia-Anatolia-Aegean (AAA) system represents a key regional setting within the Tethyan

40 domain where continental break-up, collision, backarc extension, and escape tectonics are

41 intricately linked. This interacting plate system with a wealth of available geological and

42 geophysical constraints offers a chance to evaluate the dynamics behind the motions and

43 deformation of the plates.

44 The geodetically imaged crustal motions are spectacular; they show marked, counter-clockwise

45 rotation of the velocity field with respect to Eurasia (Fig. 1a; Kreemer & Chamot-Rooke 2004,

46 Reilinger et al. 2006; Le Pichon and Kreemer, 2010). Arabia behaves as a rigid plate moving NE to

47 NNE along a rotation pole positioned in Libya, producing spreading in the Red Sea and Gulf of

48 Aden and collision against Eurasia along the Bitlis-Zagros. While convergence in Iran is

49 accommodated by distributed northward decrease in the relative velocity, toward the west, the

50 Anatolian plate moves coherently at an average rate of ~2 cm/yr, separated from Eurasia by the

51 North Anatolian Fault (Reilinger et al., 2006). In the west, the Aegean plate is moving SW at

52 ~3.3 cm/yr following the retreat of the Aegean trench (Kreemer & Chamot-Rooke 2004, Reilinger

2 53 et al. 2006). Overall, the collisional AAA zone shows a circular velocity pattern, accelerating

54 toward the Aegean trench, with a rotation pole for Anatolia-Eurasia positioned north of the Nile

55 delta (Fig. 1a, see also Le Pichon and Kreemer, 2010).

56 The dynamics of the AAA system are not only reflected in horizontal motions, but also seen in

57 the remarkable geomorphic features indicating rapid and recent vertical motions. High plateaux are

58 diffuse within the system from the Ethiopia-Arabian dome to the Turkish-Iranian plateau, both

59 standing at ~2 km elevation (Sengör et al., 2003). In contrast, subsiding regions are found behind

60 the Aegean trench and in the Makran-Persian Gulf. The vertical motion of these areas is Neogene to

61 Recent and its origins are still poorly understood. Residual topography, computed by removing the

62 crustal isostatic topography component from the actual present-day topography, shows positive

63 features over the Ethiopia - Kenya swells extending northward along the Red Sea and Anatolia-Iran,

64 while a negative signal is seen along the Hellenic trench-Persian Gulf (Figs. 1b and c). This

65 indicates that at least part of the topography could be related to mantle dynamics (Daradich et al.,

66 2003; Moucha and Forte, 2011; Becker and Faccenna, 2011).

67 Volcanism is diffuse over the region and it is mainly focused on the large Ethiopian-Yemen

68 igneous province (Fig. 1b). Intraplate volcanism is also present in Syria-Jordan and Israel far from

69 the plate boundary with large basaltic effusions and within the collisional zone, i.e. Eastern

70 Anatolia, where the volcanic signature gets complicated by the mixed contributions of different

71 geochemical sources. Figure 1b shows a general agreement between volcanism and positive

72 residual topography, suggesting a link between mantle dynamics, volcanism, and dynamic

73 topography.

74 Figures 1c and d show shear-wave (mainly SKS) splitting based estimates of azimuthal seismic

75 anisotropy of the upper mantle (from the compilation of Wuestefeld et al., 2009, 05/2012 update),

76 on top of dynamic topography, now as estimated directly from free air gravity data using the same

77 conversion and filtering as in Craig et al. (2011). The fast polarization axes of SKS splitting

78 correlate poorly with the mantle shear that may be inferred from plate motions both in Anatolia and

3 79 in Arabia (e.g. Hansen et al., 2006, 2012), illustrating the complexity of the system. The large-scale

80 mantle structure as imaged by seismic tomography shows low velocity anomalies rooted beneath

81 the Afar system as well as in northern Arabia-Syria-Jordan, and high velocity anomalies beneath the

82 Aegean and the Zagros-Makran trenches (Fig. 1d, also see Hansen et al., 2012). This pattern

83 suggests that the deeper mantle component of the AAA system is dominated by strong seismic

84 velocity contrasts and velocity anomalies that are indicative of large temperature differences, and

85 hence vigorous convection.

86 Several prior studies analyzed the kinematics and the dynamics of the region, addressing such

87 problems as to why Anatolia is moving westward, what is driving Arabia toward the north, or why

88 the Hellenic trench retreats so fast, and different models have been proposed. The presence of an

89 active subduction system and the westward acceleration of the geodetic field argue in favour of the

90 action of the pull of the Aegean slab system as the main engine, sucking Anatolia - Aegea toward

91 the trench (Le Pichon, 1992; Faccenna et al., 2006; Le Pichon and Kreemer, 2010). The motion of

92 Arabia is more problematic. It has been related to the slab-pull along the Makran trench (Bellahsen

93 et al., 2003; Reilinger and McKlusky, 2011), even if there is clear tomographic evidence that the

94 slab is broken beneath the Bitlis suture zone (Keskin, 2003, 2007; Şengör et al., 2003, 2008;

95 Faccenna et al., 2006). Another possible explanation is the Afar plume pushing Arabia northward,

96 even if the constant velocity attained by Arabia during the last ~30 Ma (Reilinger and McKlusky,

97 2011) contrasts with the expected progressive decrease in the velocity from the main upwelling

98 plume phase, ~32 Ma (van Hinsbergen et al., 2011b). More recently, it has been proposed that

99 mantle drag related to a large-scale convection cell could efficiently pull continental plates toward

100 collisional zones, thus favouring indentation (Alvarez, 2010; Becker and Faccenna, 2011). This

101 model also agrees with previous studies which suggested that Arabia is dynamically tilted by mantle

102 convection toward the Persian Gulf (Daradich et al., 2003; Moucha and Forte, 2011).

103 However, basic questions still remain unsolved. For example, it is unclear how the large-scale

104 convection cell associated with the African deep mantle plume relates to the smaller-scale one

4 105 associated with the pull of the Aegean slab, and what causes intraplate volcanism in the Middle

106 East. Moreover, it is also unclear if the high plateaux in the collisional belt are caused by active

107 mantle upwellings, lithospheric delamination, or by crustal thickening during collision.

108 The aim of this paper is to test different hypotheses about the driving forces of the Arabia-

109 Anatolia-Aegean system. Our results show the importance of mantle flow for driving the system,

110 reconciling Afar plume volcanism, the collision within the Bitlis, and the rapid backward

111 acceleration of the Aegean trench.

112

113 2. Mantle dynamics inferred from tomography and anisotropy

114 Several structural models have been proposed for the mantle underneath the Africa-Middle East-

115 Aegean region. Most seismological models consistently show a marked low velocity region over the

116 Ethiopian-Yemen dome and expanding north toward the Red Sea up to Arabia and Middle East

117 (Figs. 2 and S1). Whether the Afar and Arabian low velocity zone is rooted or not in the lower

118 mantle, and, more specifically, whether it is connected or not to the African large shear velocity

119 zone are debated questions. 3D shear wave tomography first showed the linkage between the deep

120 mantle structure and the Ethiopian province (Ritsema et al., 1999; Mégnin and Romanowicz, 2000;

121 Grand, 2002; Simmons et al., 2009; 2010; Ritsema et al. 2011). Similarly, global P-wave

122 tomography models explored the connection between the Afar hot region and deep structures

123 (Montelli et al., 2006; Li et al., 2008; Hansen et al., 2012). Regional models explored the presence

124 and the position of the low velocity anomalies in some detail (Benoit et al., 2006; Bastow et al.,

125 2005; Bastow, 2008) also showing a deep deflection of the low velocity anomaly towards the lower

126 mantle one (Park et al., 2007, 2008; Hansen et al., 2012).

127 In northern Arabia, upper mantle, low velocity anomalies have been imaged by both P-wave and

128 S-wave models (Benoit et al., 2006; Park et al., 2007, 2008). Recently, regional S-velocity structure

129 for Arabia and East Africa has been inferred by jointly inverting regional waveforms, group

130 velocities, and travel times (Chang et al., 2010; Chang and van der Lee, 2011), indicating that the

5 131 low velocity zone beneath Afar is perhaps distinct from the low velocity zone beneath northern

132 Arabia, implying a two plume origin. Conversely, other recent models favour a shallow upper

133 mantle connection of the Arabia low velocity anomaly to the Afar plume, i.e. a single plume,

134 perhaps dragged northward (Figure 2; Hansen et al., 2012).

135 Regional and global studies also show low velocity anomalies in the uppermost parts of the

136 upper mantle below Anatolia, above what can be inferred to be a deformed slab that is stagnating

137 over the transition zone, partially detached from the lithosphere (Faccenna et al., 2006; Lei and

138 Zhao, 2007; van Hinsbergen et al., 2010; Biryol et al., 2011). The Hellenic subduction zone shows a

139 40°-50° north-dipping, high velocity anomaly marked by seismicity down to intermediate depth (~

140 200 km). Lower mantle, high velocity anomalies have also been detected beneath the Hellenic

141 trench down to ~1400 km (Piromallo and Morelli, 2003; Faccenna et al., 2003; Bijward et al., 1998;

142 van Hinsbergen et al., 2005). High velocity anomalies are present also beneath the Zagros-Makran

143 collision zone (Agard et al., 2011) but the resolution over this area is relatively low.

144 Figure 2 shows a comparison between structure seen in two global S-wave tomography

145 approaches, the composite model SMEAN (Becker and Boschi, 2002), and the more recent model

146 TX2008 (Simmons et al., 2009), as well as a P-wave model, HNB12 (Hansen et al., 2012), which

147 includes more regional data than the global study by Li et al. (2008), on which it is based. It can be

148 seen that: i) Low-velocity anomalies are well defined beneath Afar, connect to the lower mantle,

149 and extend north toward the Red Sea; ii) low-velocity anomalies are also found beneath the

150 northern Arabia-Middle East region but there, they do not seem to connect to the lower mantle. The

151 recent P-wave model (Hansen et al., 2012) shows a more pronounced slow anomaly compared to

152 the other global models, widespread at shallow levels also beneath Anatolia, and Iran; iii) high

153 velocity anomalies are located beneath the Aegean and the Persian Gulf; they get more pronounced

154 and enlarge downward at mid-mantle level, as part of the larger Tethyan anomaly.

155 Mantle anisotropy as reflected in shear wave splitting shows the following features (Fig. 1d):

156 roughly NE-SW trending apparent fast polarization axes north of the East African rift in Ethiopia

6 157 (e.g. Gashawbeza et al., 2004) which rotate to mainly N-S “fast orientations” in Arabia and Israel

158 (e.g. Hansen et al., 2006; Kaviani et al., 2009), a mainly NE-SW trending in Anatolia and in central

159 Iran (e.g. Biryol et al., 2011), turning to a more northerly trend in the Aegean, and a more NW-SE

160 direction below the Zagros. Regional-scale patterns of relatively simple azimuthal anisotropy

161 alternate with more complex signatures, such as a mix of null observations and large spatial

162 gradients in fast orientations within Iran (e.g. Kaviani et al., 2009).

163 The interpretation of anisotropy in terms of mantle flow is not straightforward (see, e.g., Long

164 and Becker, 2010, for a recent review), and both lithospheric and a mantle origins of the observed

165 patterns have been suggested for the region (e.g., Hansen et al., 2006; Biryol et al., 2011). Below,

166 we will explore the predictions for anisotropy from mantle flow modeling; those can provide a

167 baseline to evaluate the feasibility of suggested asthenospheric flow scenarios such as northward

168 transport of a plume.

169

170 3. Quantitative mantle flow and anisotropy modeling

171 To quantify the contribution of mantle heterogeneity such as due to plumes and slabs to mantle

172 circulation and plate motions, we compute instantaneous mantle flow from seismic tomography by

173 converting velocity anomalies to temperature (e.g., Hager and Clayton, 1989; Lithgow-Bertelloni

174 and Silver, 1998; Moucha and Forte, 2011). We solve the Stokes equation for incompressible flow

175 in a global, spherical shell using a finite-element approach (Zhong et al., 2000). The basic approach

176 is discussed in Becker and Faccenna (2011) and Miller and Becker (2012). We test several,

177 geologically-motivated models of lateral viscosity variations as well as several density models to

178 explore the effects of the different components of the AAA convection system. The goal is not so

179 much to find the best model that matches all data at this point, but rather to identify some of key

180 regional convective processes as reflected in geological and geophysical observables.

181 The boundary conditions of the mantle flow computations are shear-stress free at the surface,

182 allowing for dynamically-consistent plate motions guided by weak zones prescribed at plate

7 183 boundaries (Ricard and Vigny, 1989; Zhong et al., 2000). Results are presented for a regional

184 zoom-in of the global models around the Arabia-Middle East, with all velocities rotated into a

185 Eurasia-fixed reference frame whose corresponding best-fit Euler pole is obtained from the surface

186 velocities within that plate.

187 The reference density structure for the mantle is based on SMEAN, with fast anomalies within

188 inferred continental keels removed above 300 km to correct for shallow compositional effects. We

189 use a Rayleigh number of 3.4·108 (defined as in Zhong et al., 2000) with temperature scaled such

190 density anomalies go with shear wave velocity anomalies as dln ρ /dln VS = 0.2. This constant

191 scaling is adopted for simplicity; it does not take depth-dependent mineral physics relationships into

192 account (e.g., Forte, 2007), nor do we allow for compositional anomalies besides keels. However,

193 such complications should mainly affect the amplitude but not the geometry of mantle flow (cf.

194 Gurnis et al., 2000).

195 Density models can also be inferred for slabs alone, based on seismicity in Wadati–Benioff

196 zones and/or inferred past subduction (e.g. Hager, 1984; Ricard et al., 1993; Lithgow-Bertelloni and

197 Richards, 1998), or by mixed models that explore the respective roles of slabs versus tomography

198 and edge forces (e.g. Becker and O'Connell, 2001; Conrad and Lithgow-Bertelloni, 2002; Ghosh et

199 al., 2010; Stadler et al., 2010). Here, we consider upper mantle models that are entirely or partially

200 based on slabs inferred from the Engdahl et al. (1998) seismicity based, RUM slab geometry of

201 Gudmundsson and Sambridge (1998), following Ghosh et al. (2010). No crustal density anomalies

202 or other effects leading to lithospheric driving forces due to gravitational potential energy

203 variations, besides those imaged by tomography, are considered.

204 We use a Newtonian viscous rheology with background viscosities for the lithosphere (depth z ≤

205 100 km), upper mantle (100 < z ≤ 660 km), and lower mantle (z > 660 km) of 5 ⋅ 1022, 1021, and

206 5 ⋅1022 Pas, respectively. Within the lithosphere, relative viscosities are 0.01 for weak zones along

207 plate boundaries, defined as in Becker and Faccenna (2011) and outlined in Fig. 1 for the reference

208 model). This simple model achieves a good general match to global plate velocities and the geoid

8 209 (Becker, 2006; Miller and Becker, 2012), and we explore additional lateral viscosity variations as

210 described below.

211 The mantle flow model results for the region are then evaluated in the light of the present-day

212 geodetic velocity fields by comparing predicted horizontal surface motions. Additionally, we

213 consider geomorphological information concerning the vertical motion and predictions of “dynamic

214 topography”. For the latter, the normal stresses generated by viscous flow at the surface are used to

215 infer the instantaneous surface deflection that would be expected because of convection (Ricard et

216 al., 1984; Richards and Hager, 1984). Assuming that all non-crustal topography is due to active

217 flow, this signal can be compared to observed topography once the effect of crustal isostatic

218 adjustment is removed to arrive at “residual topography” as shown in Fig. 1b (e.g. Daradich et al.,

219 2003; Gurnis et al., 2000).

220 Another constraint for mantle flow is provided by seismic anisotropy which in the upper mantle

221 most likely reflects the alignment of intrinsically-anisotropic, single crystals of olivine into lattice-

222 preferred orientation (LPO) under mantle flow (e.g., Long and Becker, 2010). Our approach

223 consists of assuming that flow is steady-state for the few Myr needed to achieve saturated textures,

224 following tracer particles in the mantle circulation until stable LPOs, as predicted by the mineral

225 physics theory of Kaminski et al. (2004) are reached at each desired location. Then, we assign

226 pressure and temperature-dependent elastic tensors to the resulting 3D model of anisotropy from 50

227 to 410 km depth (Becker et al., 2006a). From this elastic anisotropy, we compute synthetic

228 seismograms assuming lateral homogeneity on Fresnel zone lengths (~100 km), from which the

229 expected splitting can be computed for arbitrary back-azimuths (Becker et al., 2006b). However,

230 here we will only consider station-averaged splits (cf. Miller and Becker, 2012) and mainly discuss

231 our models in terms of the average angular misfit, 〈Δα〉, of apparent fast axes for all available non-

232 null splitting measurements (334 for the study region; Fig. 1d).

233

234 4. Results

9 235 We next illustrate the dynamics of the system using a few selected models to highlight the main

236 driving processes. To isolate the role of suction produced by the subducting slab, we first show

237 results from a model where a negative (slab) density anomaly is localized along the Aegean Wadati-

238 Benioff subduction zone only (Fig. 3a). The effect of the subducting Aegean slab produces mantle

239 suction, driving the Aegean and Anatolia toward SW and WSW, respectively. The Aegean is

240 moving rapidly, similar to what is observed, and Anatolia is moving at ~1 cm/yr, accounting for

241 ~half of its geodetic motion. The suction exerted by the slab produces an unrealistic westward

242 motion of both Arabia and Africa, giving an overall poor regional correlation of plate velocities.

243 The pattern of motion produced by the slab produces a rather good fit with the anisotropy in

244 Anatolia and eastern Aegean, but not in Arabia nor in the central Aegean. This result shows that

245 part of the mantle motion beneath Aegea and Anatolia is likely related to the subduction of the

246 Aegean slab.

247 The influence of the large-scale convection patterns induced by lower mantle density anomalies

248 as inferred from tomography when added to that of the Aegean upper mantle slab is shown on

249 Fig. 3b. Compared to the previous model, the velocity of Arabia is faster, on average ~1 cm/yr,

250 matching about half of its actual speed, and moving in the right direction. Africa's motion is also

251 faster, faster than it should be, but also with a more northerly direction. In addition, the velocity

252 field of Fig. 3b matches the gradual velocity decrease observed across the Zagros and northern

253 Iranian plateau, then fading out north of Kopet Dagh and Alborz. This model also provides a rather

254 good match for the mantle anisotropy (〈Δα〉 ~ 30°) over the entire area, with the exception of the

255 Zagros and south Arabia. The results from the model of Fig. 3b illustrate how superimposing a

256 large-scale convection cell, driven by the superswell beneath Africa and the downwelling over the

257 Tethyan slab (Becker and Faccenna, 2011) to the small scale convection related to the Aegean slab

258 improves the solution both in terms of plate velocity direction and mantle flow, although Arabia and

259 Anatolia velocities are underestimated. Dynamic topography shows a large scale tilting of Arabia

1 0 260 towards the NE, in agreement with previous predictions (Daradich et al., 2003) with maximum

261 uplift centred in southern Ethiopia.

262 To evaluate the influence of the upper mantle related to the Afar-Red Sea-Anatolia system, the

263 slow (“hot”) upper mantle anomalies are added to the previous model in Fig. 3c. This leads to an

264 overall increase of the plate velocities, as expected, produced by the vigorous upwelling centred on

265 the Afar and flowing toward the Aegean subduction zone. In this model, Arabia and Anatolia both

266 move in the right direction, while Arabia moves faster and Anatolia slightly slower than observed

267 and overall correlation for regional velocities is quite good. As expected, the contribution of the

268 upper mantle anomaly within the system increases the dynamic topography showing a very strong

269 signature in the Afar and Carlsberg Ridge (~2 km of topography) and some hundreds of meters

270 positive elevation in eastern Anatolia and the Caucasus, as discussed earlier by Boschi et al. (2010).

271 The faster Arabia motion degrades slightly the correlation of the mantle anisotropy (〈Δα〉 = 35°)

272 with respect to the model of Fig. 3b. This is particularly evident in southern Arabia, where

273 upwelling produces radial mantle flow striking perpendicularly to the observed NS trending

274 anisotropy.

275 The effect of lateral strength variations in modifying seismic anisotropy and flow is explored by

276 adding a low viscosity channel to the model of Fig. 3b. The low viscosity channel runs from the

277 Afar to the Red Sea, where mantle lithosphere is imposed to be thinner than the surrounding. This

278 modification increases the match with anisotropy 〈Δα〉 to 28°, and particularly improves the match

279 in Arabia (Fig. 3d). This exploration of the low viscosity channel provides a quantitative test of

280 previous suggestions that the Afar upwelling is guided within a low viscosity mantle north beneath

281 Red Sea and western Arabia (Hansen et al., 2006) or towards the north beneath Arabia (Chang et

282 al., 2011).

283 The geometry of typical mantle convection patterns for the model of Fig. 3c can be better shown

284 on map view at different depths and cross sections (Fig. 4). At shallow upper mantle depth, the

285 sinking of the Aegean slab and the upwelling of the mantle in Afar generates a large-scale current,

1 1 286 upwelling in the Afar and downwelling in the Aegean, leading to some toroidal motion when

287 interacting with the plate geometry. The convection cell is well represented by the cross sections

288 running from Afar to the Persian Gulf and to the Aegean trench (Figs. 4e, 4f). In addition, a shallow

289 upwelling in eastern Anatolia generates a smaller-scale convection feature from the Aegean trench

290 to the Anatolian plateau. Much of these convection patterns are related to upper mantle motions,

291 which efficiently drag plates and produce pronounced positive dynamic topography at the surface.

292 At the base of the upper mantle, these perturbations flatten out, generating a more regular pattern of

293 mantle flow trending toward WNW.

294 We computed mantle flow and seismic anisotropy for a large number of models, and a

295 representative selection of results for different density models and different choices for lateral

296 viscosity variations are shown in Fig. 5. Such a representation in misfit matrix form allows

297 assessing the combined effects of the different parameters that were illustrated in Fig. 3 (Miller and

298 Becker, 2012). Moreover, the role of density models and the relevance of different model

299 assumptions, such as on the geometry of weak zones, and regional plate coupling, can be assessed.

300 Fig. 5 shows misfits as measured by linear correlation coefficient for global plate velocities (a),

301 regional plate velocities (b), and for the average misfit of the fast axes from seismic anisotropy (c).

302 At a regional scale, the best predictions of plate velocities when measured in terms of correlation

303 are obtained by adopting the SMEAN composite model with upper mantle slabs added. Overall, the

304 presence of a sub-oceanic plate asthenosphere improves the fit for velocity (e.g. Becker 2006). This

305 is also true for the presence of a low viscosity channel beneath Red Sea (cf. Hansen et al., 2006),

306 and Fig. 5 provides an estimate of the sensitivity of details of the viscosity variations underneath

307 Arabia on shear wave splitting estimates.

308 The delay time amplitudes of splitting measurements are a measure of the strength of anisotropy,

309 yet typically less stable than the “fast azimuth”. In our modeling approach, delay times are mainly

310 controlled by the degree of LPO saturation. Guided by the good fit to SKS splitting delay times

311 achieved in other regions (Becker et al., 2006b, Miller and Becker, 2012), we here chose a

1 2 312 logarithmic strain of 0.75. For the AAA system, this choice leads to an overprediction of delay

313 times by ~0.4 s (Fig. 3). We mainly take this an indication for the heterogeneity and complexity of

314 seismic anisotropy in the study region (Figs. 1c and d). This may indicate the role of lithospheric,

315 frozen-in anisotropy, in obscuring the asthenospheric signal in some regions such as in Iran or the

316 Hellenic, but delay times will not be considered further in this study.

317 Overall, we find that the match to the shear wave splitting fast axes is very sensitive to details of

318 the lateral viscosity variations, and in regions such as Iran we are close to the limit of what can

319 usefully be explored with our global mantle flow models with homogeneous lithospheric rheology.

320 Further model refinement may thus allow a better simultaneous fit to crustal velocities, dynamic

321 topography, and seismic anisotropy. However, for our purposes, the results in terms of the major

322 drivers of horizontal and vertical motions of the AAA plate system dynamics are robust.

323 In summary, the global circulation models show the importance of the combined effect of slab

324 pull and plume push to match the velocity field of microplates in the Middle East. Model results

325 indicate that slab pull is essential to drive Anatolia and the Aegean microplate toward the west and

326 southwest, respectively. A large-scale upwelling in the Afar can be channelled into the upper mantle

327 beneath the Red Sea, and such flow is necessary to match the motion of Arabia correctly. The

328 overall convection patterns are dominated on different scales by convection systems that link the

329 upwelling areas in Afar up to northern Arabia, to the shallow Aegean slab and, at larger depth, with

330 the Tethyan slab.

331

332 5. Geological constraints on mantle convection

333 Geological evidence implies that the convection patterns reconstructed from our circulation

334 models for the present-day were activated in the Oligocene-late Eocene. The first trace of volcanism

335 in Ethiopia dates back to ~45 Ma (Ebinger et al., 1993). It is probably at that time that a large

336 volume of mantle emplaced at the base of the lithosphere (Sleep and Ebinger, 1998) while Africa

337 was still a coherent plate, slowly moving NE toward, Eurasia consuming the Tethyan ocean along

1 3 338 the Aegean-Bitlis trench system (Dercourt et al., 1986; Barrier and Vrielynck, 2010). The deep

339 mantle structure associated with the evolution of the Aegean subduction has previously been

340 reconstructed from the seismically imaged high velocity anomalies, restored accounting for time-

341 dependent convergence kinematics (Faccenna et al., 2003; van Hinsbergen et al., 2005). The match

342 between the absolute motion of the trench and the deep position of the high velocity anomalies into

343 the lower mantle gives a maximum age for the entrance of the Aegean slab into the lower mantle of

344 ~50+10 Ma, probably after a stagnation phase of ~10-20 Myr duration on the 660 km discontinuity

345 (Faccenna et al., 2003; 2012). This tectonic episode is extending regionally toward the Himalayan

346 trench (Replumaz et al., 2004). The episode of slab avalanching at ~45-50 Ma in the Hellenic

347 domain has been interpreted as the cause of stretching of the Sirte basin in north Africa (Capitanio

348 et al., 2009). We infer that the large-scale, Tethyan convective cell that is active today (Becker and

349 Faccenna, 2011) between Afar and the Tethyan subduction zone was already operating at ~45 Ma.

350 The most apparent expression of the deep mantle upwelling is the outpouring of the Ethiopian

351 large igneous province at ~33-30 Ma (Hoffman et al. 1997, Fig. 6a). The Red Sea and the Gulf of

352 Suez (Fig. 1) rifted at the beginning of the Oligocene (Ghebread, 1998). In the Red Sea, dykes and

353 basaltic floods were emplaced during the Oligocene (Ghebread, 1998; Bosworth et al., 2005) and

354 syn-rift deposits accumulated from the Oligocene-Miocene (~30-25 Ma) to Miocene (14 Ma)

355 (Bayer et al., 1988; Garfunkel and Beyth 2006; ArRajehi et al. 2010). Thermochronology shows

356 that uplift of the Red Sea rift shoulders commenced at ~27 Ma and peaked at ~20 Ma (Bosworth et

357 al., 2005) and the erosion of the Blue Nile gorges commenced at 25-29 Ma (Pik et al., 2003).

358 Seismic data show a marked change in sedimentation in the Nile delta both offshore and onshore

359 and both in the Levantine and Herodotus Basin at the Eocene-Oligocene boundary. This probably

360 marks the onset of the proto-Nile drainage (Steinberg et al., 2011). Stratigraphic and morphological

361 studies show the presence of a regional truncation surface developed during the Oligocene from the

362 northern Red Sea to southern Levant related to the formation of an ~3000 km × 1500 km crustal

363 dome above the Afar plume, with a widespread fluvial erosion that truncated the northern edge of

1 4 364 this dome (Avni et al., 2012). Evidence of episodic Miocene to Pliocene regional uplift are also

365 found in Isreal (Gvirtzman et al., 2011). First syn-rift sediments in the Red Sea are dated there to

366 the late Eocene-early Oligocene, ~35 Ma (Roger et al, 1989; Watchorn et al. 1998). The eastern

367 arm of the rift system is represented by the Gulf of Aden (Fig. 6b). Extension occurred in two

368 directions, first N20°E and then N160°E (Huchon et al., 1991; Lepvrier er al., 2002; Huchon and

369 Khanbari, 2003). The absolute northward motion of Africa progressively slowed down after 35 Ma

370 (Jolivet and Faccenna 2000; Reilinger and McKlusky, 2011) while Arabia-Eurasia convergence

371 kept a constant velocity of ~2-3 cm/yr with respect to Eurasia from 56 Ma to the present, and

372 Africa-Eurasia convergence slowed down to ~1 cm/yr after ~25 Ma once the opening of the Red

373 Sea had started (McQuarrie et al., 2003).

374 The timing of the onset of collision in the Bitlis-Zagros is debated. Some authors proposed that

375 continental subduction initiated around the Oligocene (Jolivet and Faccenna, 2000; Allen and

376 Armstrong 2008; Agard et al. 2005; 2011) whereas others argue that orogenic structures were built

377 in the Early Miocene (Berberian and Berberian, 1981; Okay et al., 2010). This latter hypothesis is

378 supported by the timing of the closure of the Tethyan gateway ~18 Ma (Reuter et al., 2009),

379 allowing the deposition of detritus issued from the Eurasian margin onto Arabia. In addition, in

380 Eastern Anatolia last marine deposits date back to the Early Burdigalian (~12 Ma, Gelati, 1975)

381 whereas in the central Anatolia plateau they post-date the Messinian (Cosentino et al., 2012). The

382 stratigraphic record thus indicates that thrust systems in the Bitlis emerged from shallow water ~16

383 Ma. Fission track data indicate a period of increased exhumation around the mid-late Miocene

384 (Okay et al., 2010). Most probably, the onset of continental subduction occurred in the Oligocene,

385 while the final closure of the Tethyan gateway and creation of an orogenic relief later, around the

386 Middle Miocene. Volcanism initiated at HAS and in Northwest Syria ~ 25 Ma ago. The occurrence

387 of volcanism at HAS and in NW Syria has been interpreted as due to the arrival of hot material with

388 melting and magma ascent, enhanced by or enhancing crustal extension (Krienitz et al., 2007;

389 2009).

1 5 390 Around 11+2 Ma the velocity of Africa slowed down by ~50% for most velocity models

391 (Reilinger and McKlusky, 2011) while the motion of Arabia was not strongly perturbed keeping a

392 constant velocity at ~2 cm/yr (Burke, 1996; McQuarrie et al., 2003; Muller̈ et al., 2008). McQuarrie

393 et al. (2003) and Reilinger and McKlusky (2011) described a major re-organization of plate

394 boundaries at 9+4 Ma (Fig. 6c). At that time, Arabia was fully separated from Africa with the

395 complete opening of the Gulf of Aden (Reilinger and McKlusky, 2011). During Miocene times, the

396 activity of the Gulf of Suez decreased, transferred to the Gulf of Aqaba with the consequent

397 formation of the Levant fault, while collisional processes in the Bitlis- Pötürge massif are already

398 well developed. Contemporaneously, a surge of volcanism occurred in NW Syria in the Aleppo

399 region (Krienits et al., 2007; 2009) which has been related to asthenosphere upwelling during slab

400 break-off (Keskin, 2003, 2007; Krienitz et al., 2009; Şengör et al., 2003, 2008).

401 In Eastern Anatolia, the geochemical character of volcanism changed at around 10 Ma from an

402 original calc-alkaline, subduction-related type towards a more alkaline one (Keskin et al., 2007). In

403 particular, the the presence of calc-alkaline volcanism with a distinct subduction signature across

404 Eastern Anatolia has been related to a first episode of slab-steepening event between ~15 and 10

405 Ma, before the slab break off at ~10 Ma (Keskin, 2003). This slab steepening episode likely

406 exposed the base of the Eastern Anatolian accretionary complex to asthenospheric temperatures and

407 dragged subduction-related mantle material southward from beneath the Pontides toward the mantle

408 wedge (Keskin, 2003, 2007). Petrological studies also show that the slab steepening episode was

409 likely followed by slab rupture at ~10 Ma, resulting in the formation of a slab window. This is

410 attested to the eruption of the first alkaline basanitic lavas in the Eastern Anatolian volcanic

411 province at ~10 Ma (Lebedev et al., 2010) in the north of Lake Van (Keskin et al., 2012c). This

412 implies that enriched and fertile asthenospheric, hotter material once located beneath the slab (i.e.,

413 underneath the Arabian continent) moved northward to shallow depths through a slab window

414 (Keskin, 2007). I.e., the convective flow possibly coming from the Afar plume may be the source of

415 the volcanism across SE Anatolia (Keskin et al, 2012b) (Fig. 5). The North Anatolian Fault

1 6 416 probably initiated in east Anatolia (~10-12 Ma) (Sengör, 1979), propagating westward at ~7-6 Ma

417 to central Anatolia, and arriving in the Aegean in the Pliocene. Keskin et al. (2012a, b) show that

418 the extensive volcanic field of the Karacadağ volcanic area of SE Anatolia was sourced by a garnet-

419 bearing, deep asthenospheric mantle which is similar to that of Afar in terms of its Sr, Nd, and Pb

420 isotopic ratios (Fig. 7).

421 The evolution of the volcanic system displays a progressive decrease in the degree of

422 involvement of a shallow, subcontinental lithospheric mantle source toward a deeper,

423 asthenospheric one, with isotopic composition similar to Afar lavas. From the Pliocene onward,

424 widespread volcanism occurred in Syria along the Dead Sea Fault but also in Harrat Ash Shamah

425 (HAS), NW and NE Syria, and SE Turkey (e.g. Karacadağ: Lustrino et al., 2010, 2012; Keskin, et

426 al., 2012a, 2012b; Ekici et al., 2012) contemporaneous with the activity of large volcanic fields

427 along the Red Sea coast (e.g., Giannerini et al., 1988) and during the emplacement of ocean floor in

428 the Red Sea (Girdler, 1990). In the HAS, effusion rates are similar to Ethiopia (Hofmann et al.,

429 1997), with ~700 m thick of plateau basalts erupted in less than 1 Ma (Krienitz et al., 2009). As for

430 Karacadağ, the HAS intraplate volcanic is also derived from a mixed lithosphere-asthenosphere

431 source, activated by lithosphere extension (Shaw et al., 2003).

432 The similarities in isotopic composition between the lavas of the Afar region and those of the

433 Arabian province and the expected increase of the mantle temperature during the diffuse phase of

434 volcanism has been related to a sub-lithosphere source related to the Afar plume (Camp and

435 Roobol, 1992; Krienitz et al., 2007, 2009), supporting a model of northward directed mantle flow

436 (Ershov and Nikishin, 2004; Yarmolyuk et al., 2004) that at ~10 Ma reached the Eastern Anatolian

437 Collision Zone after the slab steepening and breakoff (Keskin, 2003, 2007; Keskin et al., 2012c).

438 The temporal increase in the alkalinity coupled with a gradual decrease in the subduction

439 component starting at around 10 Ma in the collision-related volcanic units of Eastern Anatolia

440 supports this interpretation (Keskin et al., 2012c).

441

1 7 442 6. Discussion: Geodynamic evolution of the AAA system

443 In the context of geological and geophysical constraints on the tectonic history of the region, our

444 geodynamic modeling results indicate that the upwelling mantle underneath Afar, the northward

445 motion of Arabia with consequent collision in the Bitlis, the westward escape of Anatolia, and

446 backarc extension in the Aegean may all be the result of the progressive evolution of the Tethyan

447 convection cell (Figs. 6 and 8).

448 Combined effects of an upwelling in Afar and subduction in the Hellenic drag Arabia toward

449 collision. The upwelling in the Afar likely commenced at ~40 Ma but massive volcanic outpouring

450 and the formation of a large igneous province is dated at ~33 - 30 Ma. The volcanic record shows

451 that magmatism propagated northward from Ethiopia-Yemen toward the Arabia-Anatolia

452 collisional zone. The spatio-temporal distribution of volcanism, and the similarities of the isotopic

453 compositions of the late stage lavas, suggest that the origin of these melts is most likely related to a

454 sub-lithosphere source resembling the one erupted in Arabia, and connected with the increased

455 temperature due to the emplacement of the Afar plume (Krienitz et al., 2009). The isotopic

456 signature of the Afar plume can be traced up to the Karacadağ neovolcanic area on the northernmost

457 part of the Arabian continent in Turkey (Fig. 7).

458 Those data then support a model of northward directed mantle flow (Ershov and Nikishin, 2004;

459 Yarmolyuk et al., 2004), channelling the Afar plume head beneath Arabia (Camp and Roobol, 1992;

460 Ebinger and Sleep, 1998; Hansen et al., 2006) and further north beneath Eastern Anatolia after slab

461 breakoff (Keskin, 2007). Ebinger and Sleep (1998) proposed that rifts may represent natural

462 channels for hot material from mantle plumes to flow horizontally beneath thin lithosphere. Ershov

463 and Nikishin (2004) argue that the lithosphere is relatively thin along a south-north line extending

464 from Afar to Anatolia due to a previous orogenic event. Alternatively, Chang and van der Lee

465 (2011) suggested that the low velocity anomaly beneath the Middle East is distinct from the Afar

466 one and is rooted in the lower mantle. However, other tomography models indicate that the slow

467 velocity beneath Arabia is limited to the upper mantle (Figs. 2, S1). For example, Hansen et al.

1 8 468 (2012) find that the Arabia low velocity zone is connected to the Afar plume in the upper mantle,

469 eventually connecting down to the lower mantle African low shear wave velocity anomaly. We

470 consider this single plume scenario the simpler explanation of deep mantle structure, on which local

471 complexities and volcanic trends may well be superimposed.

472 Shear-wave splitting and tomographic results of Hansen et al. (2006) and Park et al. (2007, 2008)

473 and the simplified models of Ershov and Nikishin (2004) suggest channelled, horizontal mantle

474 flow from the Afar hotspot. Our modelling substantiates these suggestions and shows that shear

475 wave splitting and mantle anisotropy below Arabia can be fit best when introducing a low viscosity

476 zone, running along the Red Sea which would channel hot mantle material northward. The spatio-

477 temporal distribution of within plate alkaline volcanism in Arabia also suggests that the deep source

478 migrated northward from Yemen (31 Ma) to northern Syria (19 Ma) at a high, average speed of ~

479 22 cm/yr (Krienitz et al., 2009) even though this estimate could be significantly reduced by

480 considering a larger initial extent of the Afar plume (Ebinger and Sleep, 1998). Our models show

481 that the velocity of sub-lithosphere material can reach velocity around 10 to 20 cm/yr for models

482 accounting for a low viscosity asthenosphere, or temperature dependent viscosity, and this

483 asthenospheric flow could be even faster if channelled along the NS corridor.

484 Our flow computations also show that the Afar - Arabia upwellings drive Arabia northward,

485 while dynamically tilting the entire plate toward the Persian Gulf (Daradich et al., 2003). The

486 asthenospheric mantle is “leading the plate”, and the present collision with Eurasia is inferred to be

487 driven by a mantle “conveyor belt” (Becker and Facenna, 2011). The steady, ~2 cm/yr northward

488 motion of Arabia prior and after its separation with Africa has been taken as an indication of the

489 role of slab-pull from the Zagros-Bitlis to move the plate northward (Reilinger and McKlusky,

490 2011). Conversely, we take the relatively constant plate speed, irrespectively of the collision and

491 break-off episodes (e.g. Keskin, 2003, 2007; Şengör et al., 2003, 2008Faccenna et al., 2006), as a

492 sign that other forces rather than slab-pull drive Arabia northward. We conclude that mantle

1 9 493 upwellings in the Afar-Arabia first separated Arabia from Africa, and then propelled Arabia

494 northward.

495 Following Burke (1996) and Jolivet and Faccenna (2000) we suggest that the reduction of the

496 absolute velocity of Africa after ~30-35 Ma to the present is a consequence of the collision in the

497 northern Zagros and Bitlis suture zone and that Arabia was detached from the slowing main African

498 plate driven by the continuous basal drag imposed by the asthenospheric flow underneath. Arabia

499 then rotated counterclockwise as a consequence of the asymmetric conditions imposed by the new

500 plate boundary geometry, as suggested in Jolivet and Faccenna (2000) and Bellahsen et al. (2003).

501 The massive outpouring of volcanic material in the Afar triple junction area is then allowed by the

502 extension recorded in this area.

503 Downwelling system in the Aegean and the extensional system. Subduction in the Aegean-

504 Tethyan system extended from the Mediterranean to Asia (van der Voo et al., 1999), but the timing

505 of the tectonic evolution of the Tethyan margin appears to also reflect temporal fluctuations in

506 mantle upwellings, encouraging the view of a coupled system. In the Aegean, the penetration of a

507 slab into the lower mantle probably occurred at ~50-40 Ma (Faccenna et al., 2003; Capitanio et al.,

508 2009), contemporaneously with the onset of Ethiopian volcanism. This indicates that those two

509 processes overlap in time, producing the onset of a large-scale, whole mantle convection cell.

510 Subduction in the Aegean exerts a first order control on the evolution of the AAA system. The

511 efficiency of slab pull on the Middle east system is illustrated by the model where only slab pull is

512 at work, matching shearing deformation in the mantle as revealed by mantle anisotropy data

513 (Fig. 3a). Geological data show that fast back-arc extension in the area commenced at ~30-35 Ma

514 (Jolivet and Faccenna, 2000; Jolivet et al., 2012) even though slower extension in the Rhodope

515 region could have started earlier (Brun and Sokoutis, 2007). The total amount of extension in the

516 Aegean has been estimated to be of the order of 300-400 km. But paleomagnetic data show that the

517 present day curvature of the Aegean arc occurred mainly after 15 Ma, indicating that extensional

518 rate accelerates with time.

2 0 519 The interaction between mantle upwelling and collisional processes. Efficient subduction, under

520 the influence of vigorous slab pull, does not favour collision when the trench is moving backward

521 (Molnar and Atwater, 1971; Faccenna et al., 2012). The trench northeast of Arabia is, however,

522 moving northward, indenting the overriding plate. This advancing motion of the trench can be

523 favoured by the entrance of continental lithosphere in the subduction zone (Bellahsen et al., 2003),

524 although numerical models suggested that this motion is limited to less than 100 km (Magni et al.,

525 2012). The Bitlis convergence zone advanced toward the upper plate by 400-500 km during the last

526 25 Ma. This cannot be easily reconciled with the entrance of continental material alone, and an

527 extra push force related to mantle drag is thus required. The onset of continental subduction

528 probably occurred at ~30 Ma. Remarkably, only a few hundreds of kilometers south of the collision

529 zone, effusion of alkali basalts was taking place some 10 Ma later while the collision process was

530 ongoing (Fig. 6). The proximity of these two contrasting processes suggests local interactions

531 between mantle upwellings and collision. The orogenic build up as revealed by rapid exhumation

532 and continental stratigraphy (Okay et al., 2010) occurred in the late early Miocene (~18 to 13 Ma).

533 Tomographic images and volcanological evolution reveals that the slab is broken beneath the Bitlis

534 suture zone (Faccenna et al., 2006). It is then possible that slab first acted as a “sail”, advancing

535 northward under the action of the flowing mantle and was then eroded and perhaps segmented,

536 facilitated by a thermal, weakening effect of the plume, which then allowed for the entrance of

537 mantle material within the Mediterranean region (Fig. 6) (Keskin, 2007).

538 Collision, Anatolia motion and Aegean extension. One of the main geodynamic problems of the

539 Middle East is represented by the escape of Anatolia. This motion probably initiated during the Late

540 Miocene, from ~ 10-12 Ma with a progressive localisation of the North Anatolian Fault at 5-6 Ma

541 and onward (Sengör et al., 2005; Faccenna et al., 2006), when both collision in the Bitlis mountains

542 and backarc extension in the Aegean were well under way. The lateral escape of Anatolia probably

543 occurred contemporaneously with the rise up of the plateau, which in Central Anatolia occurred

544 after ~6 Ma (Cosentino et al., 2012), while in eastern Anatolia it could have started earlier after

2 1 545 ~12 Ma (Gelati, 1972). Our results show that more than ~50% of the motion of Anatolia is

546 explained by the upwelling beneath Arabia-Middle East. As shown in the cross sections of Fig. 6,

547 for this flow to be efficient in Anatolia or Aegean, it has to bypass the subduction system plunging

548 northward. The formation of a slab window, probably at ~8-10 Ma and progressively enlarging

549 westward (Faccenna et a., 1996; Cosentino et al., 2012) could have triggered and favoured by the

550 arrival of hot material from Arabia towards the collision zone (Keskin, 2007). The onset of a surge

551 of volcanism over the central Anatolia plateau attests to the arrival of hot material in the system at

552 ~5 Ma. We may further speculate that the hot, northward flowing material could have even

553 triggered and favoured the formation of this slab window, progressively pushing against the slab

554 itself. If this is true, we are facing the evolution of a convection cell that was invaded by the hot-

555 upwelling material because the downwelling branch of the cell below Bitlis vanished due to the

556 closure of the oceanic basin.

557 None of our models succeeds to fully match the complexity of the region as attested by non-

558 isostatic topography, surface velocities, and the complex patterns of seismic anisotropy (Fig. 5), and

559 further adjustments of models may yield locally refined dynamic scenarios. However, our models

560 give a robust, first order view on the dynamics of the system, where the coupling between

561 upwelling mantle in the Afar-Arabia system and the downwelling of the Aegean - Tethyan slab can

562 generate the general patterns of deformation observed at the surface.

563

564 Conclusion

565 We reconstruct the patterns of mantle convection beneath the Afar-Anatolia-Aegean system and

566 compare our present-day results to geodetically measured crustal motions and constraints from

567 seismic anisotropy. Our mantle flow models highlight the importance of coupling the pull exerted

568 by the Hellenic slab with the upwelling below Afar-Arabia; this explains the northward indenter

569 motion of Arabia, the northward progression of volcanism along Arabia, the dynamic uplift of

2 2 570 Anatolia, its westward motion, and the Aegean extension in a single convection system, established

571 episodically over the last ~40 Ma.

572 Jointly interpreting the present-day geodynamic models with geological constraints, we suggest

573 that it may be possible to backtrack the geometry and interactions of the AAA lithosphere-mantle

574 system throughout the Tertiary. The upwelling material associated with the deep mantle Africa-Afar

575 plume flowed northward in the asthenosphere during the closure of the ocean, assisted in

576 continental breakup, and propelled Arabia to induce collision. Several slab segmentation episodes

577 might have opened mantle flow windows, facilitating plume northward transport. Eventually, this

578 lead to asthenospheric flow inside the Aegean region, and accelerating of the Hellenic backarc

579 extension to the rates we see today. Our findings emphasize the role of the deep mantle in affecting

580 microplate motions and controlling the geological expression of upper mantle convection.

581

582 Acknowledgments

583

584 Discussions with Zohar Gvirtzman and Jean Pierre Brun contributed to clarify some of the

585 concepts expressed here. We thank CitcomS authors including S. Zhong, E. Tan, A. McNamara,

586 and L. Moresi for their contributions, CIG (geodynamics.org), E. Kaminski for sharing DREX, and

587 all seismologists who share their tomographic models and shear wave splitting measurements in

588 electronic form. Computations were performed on USC's High Performance Computing Center, and

589 research was partially supported by NSF grants EAR-0809023 and EAR-0643365. M. Keskin

590 thanks to TUBITAK for their support (project no. 108Y222).

591

592 References

593 Agard, P., Omrani, J., Jolivet, L., Mouthereau, F., 2005. Convergence history across Zagros

594 (Iran): constraints from collisional and earlier deformation. International Journal of Earth Sciences

595 94, 401–419.

2 3 596 Agard, P., Omrani, J. Jolivet, L., Whitechurch, H., Vrielynck, B., Spakman, W., Monié, P.,

597 Meyer, B., Wortel, R., 2011. Zagros orogeny: a subduction-dominated process. Geological

598 Magazine, 1-34.

599 Allen M. B., Howard, A., 2008. Armstrong Arabia–Eurasia collision and the forcing of mid-

600 Cenozoic global cooling, Palaeogeography, Palaeoclimatology, Palaeoecology 265, 52–58.

601 Alvarez, W., 2010. Protracted continental collisions argue for continental plates driven by basal

602 traction. Earth Planet. Sci. Lett. 296, 434-442.

603 ArRajehi, A., McClusky, S., Reilinger, R., Daoud, M., Alchalbi, A., Ergintav, S., Gomez, F.,

604 Sholan, J., Bou-Rabee, F., Ogubazghi, G., Haileab, B., Fisseha, S., Asfaw, L., Mahmoud, S., Rayan,

605 A., Bendik, R., Kogan, L., 2010. Geodetic constraints on present day motion of the Arabian Plate:

606 implications for Red Sea and Gulf of Aden rifting. Tectonics 29, TC3011.

607 doi:10.1029/2009TC002482.

608 Avni, Y., Segev, A., Ginat, H., 2012.Oligocene regional denudation of the northern Afar dome:

609 Pre- and syn-breakup stages of the Afro-Arabian plate. Geological Society of America Bulletin 124,

610 11-12, 1871-1897, doi:10.1130/B30634.1

611 Barrier, E., Vrielynck, B., 2010. Palaetectonic map of the Midlle East. Middle East basin

612 evolution Programme. Scale 1/18 500 000

613 Bassin, C., Laske, G., Masters, G., 2000. Current limits of resolution for surface wave

614 tomography in North America. Eos Trans. AGU, 81, F897.

615 Bastow, I.D., et al., 2008. Upper mantle seismic structure beneath the Ethiopian hot spot: Rifting

616 at the edge of the African low-velocity anomaly. Geochem. Geophys. Geosyst. 9, Q12022.

617 Bastow, I., Stuart, G., Kendall, J.-M., Ebinger, C., 2005. Upper-mantle seismic structure in a

618 region of incipient continental breakup: northern Ethiopian rift. Geophys. J. Int. 162(2), 479–493.

619 Bayer, H.J., HoÅstzl, G., Jado, A.R., Roscher, B., Voggenreiter, W. 1998. Sedimentary and

620 structural evolution of the northwest Arabian Red Sea margin. Tectonophysics 153, 137-151.

2 4 621 Becker, T. W., 2006. On the effect of temperature and strain-rate dependent viscosity on global

622 mantle flow, net rotation, and driving forces. Geophys. J. Int. 167, 943-957.

623 Becker, T. W., Faccenna, C., 2011. Mantle conveyor beneath the Tethyan collisional belt. Earth

624 Planet. Sci. Lett. 310, 453-461.

625 Becker, T. W., O'Connell, R. J., 2001. Predicting plate velocities with geodynamic models.

626 Geochem. Geophys. Geosys. 2, 12.

627 Becker, T. W., Chevrot, S., Schulte-Pelkum, V., and Blackman, D. K., 2006a. Statistical

628 properties of seismic anisotropy predicted by upper mantle geodynamic models. J. Geophys. Res.

629 111, B08309, doi:10.1029/2005JB004095.

630 Becker, T. W., Schulte-Pelkum, V., Blackman, D. K., Kellogg, J. B., and O'Connell, R. J.,

631 2006b. Mantle flow under the western United States from shear wave splitting. Earth Planet. Sci.

632 Lett. 247, 235-251,

633 Becker, T.W., Boschi, L., 2002. A comparison of tomographic and geodynamic mantle models.

634 Geochem. Geophys. Geosys. 3, 2001GC000168.

635 Bellahsen, N., Faccenna, C., R., F., J.M., D., Jolivet, L., 2003. Why did Africa separate from

636 Africa? Insight from 3D laboratory experiments. Earth Planet. Sci. Lett. 216, 365–381.

637 Benoit, M., Nyblade, A., VanDecar, J., 2006. Upper mantle P wavespeed variations beneath

638 Ethiopia and the origin of the Afar hotspot. Geology 34(5), 329–332.

639 Berberian, F., Berberian, M., 1981. Tectono-plutonic episodes in Iran, in: H.K. Gupta, F.M.

640 Delany (Eds.), Zagros, Hindu Kush, Himalaya. Geodynamic Evolution, Geodynamics Ser. 3,

641 Washington, DC, pp. 5-32.

642 Bijwaard, H., Spakman, W., Engdahl, E., 1998. Closing the gap between regional and global

643 travel time tomography, J. Geophys. Res. 103, 30055-30078.

644 Bird, P., 2003. An updated digital model of plate boundaries. Geochemistry Geophysics

645 Geosystems 4(3), 1027.

2 5 646 Biryol, C.B., Beck, S.L., Zandt, G., Özacar, A.A., 2011. Segmented African lithosphere beneath

647 the Anatolian region inferred from teleseismic P-wave tomography. Geophysical Journal

648 International. 184, 1037– 1057.

649 Boschi, L., Faccenna, C., Becker, T.W., 2010. Mantle structure and dynamic topography in the

650 Mediterranean Basin. Geophys. Res. Lett. 37, L20303.

651 Bosworth, W., Huchon, P., McClay, K., 2005. The Red Sea and Gulf of Aden Basins. Journal of

652 African Earth Sciences 43, 334–378.

653 Brun, J.P., Sokoutis, D., 2007. Kinematics of the Southern Rhodope Core Complex (North

654 Greece). International Journal of Earth Science, DOI 10.1007/s00531-007-0174-2.

655 Burke, K., 1996. The African Plate. S. African J. Geology 99, 341–409.

656 Camp, V. E., and M. J. Roobol (1992), Upwelling asthenosphere beneath western Arabia and its

657 regional implications. J. Geophys. Res. 97, 15,255 – 15,271, doi:10.1029/ 92JB00943.

658 Capitanio, F.A., Faccenna, C., Funiciello, R., 2009. The opening of Sirte basin: Result of slab

659 avalanching?. Earth Planet. Sci. Lett. 285, 210-216.

660 Chang, S.-J., van der Lee, S. , Flanagan, M. P., Bedle, H., Marone, F., Matzel, E. M. , Pasyanos,

661 M. E., Rodgers, A.J., Romanowicz, B., Schmid, C. . 2010. Joint inversion for three-dimensional S

662 velocity mantle structure along the Tethyan margin. J. Geophys. Res. 115, B08309,

663 doi:10.1029/2009JB007204.

664 Chang, S.-J., Van der Lee, S., 2011. Mantle plumes and associated flow beneath Arabia and East

665 Africa. Earth Planet. Sci. Lett. doi:10.1016/j.epsl.2010.12.050.

666 Chang, S.-J., Merino, M., Van der Lee, S., Stein, S., Stein, C. A., 2011. Mantle flow beneath

667 Arabia offset from the opening Red Sea. Geophys. Res. Lett. 38, L04301,

668 doi:10.1029/2010GL045852.

669 Conrad, C.P., Lithgow-Bertelloni, C., 2002. How mantle slabs drive plate tectonics. Science 298,

670 207-209.

2 6 671 Cosentino, D.; Schildgen, T. F.; Cipollari, P., et al. 2012. Late Miocene surface uplift of the

672 southern margin of the Central Anatolian Plateau, Central Taurides, Turkey. Geological Society of

673 America Bulletin 124, 1-2, 133-145.

674 Daradich, A., Mitrovica, J.X., Pysklywec, R.N., Willett, S.D., Forte, A.M., 2003. Mantle flow,

675 dynamic topography, and rift-flank uplift of Arabia. Geology 31, 901–904.

676 DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1994. Effect of recent revisions to the

677 geomagnetic reversal time scale on estimates of current plate motions. Geophys. Res. Lett. 21,

678 2191–2194.

679 Dercourt, J., et al., 1986. Geological evolution of the Tethys belt from the Atlantic to the Pamirs

680 since the Lias. Tectonophysics 123, 241–315.

681 Ebinger, C. J., Sleep, N. H., 1998. Cenozoic magmatism throughout east Africa resulting from

682 impact of a single plume. Nature 395, 788–791.

683 Ebinger, C. J., Yemane, T., Wolde Gabriel, G., Aronson, J., 1993. Late Eocene-Recent

684 volcanism and faulting in the southern Main Ethiopian rift system. J. Geol. Soc. Lond. 150, 99–108.

685 Ekici, T., Macpherson, C.G., Otlu, N., 2012. Polybaric melting of a single mantle source during

686 the Neogene Siverek phase of the Karacadağ Volcanic Complex, SE Turkey. Lithos 146–147, 152–

687 163.

688 Engdahl, E.R., van der Hilst, R.D., Buland, R., 1998. Global teleseismic earthquake relocation

689 with improved travel times and procedures for depth determination. Bull. Seismol. Soc. Am. 88,

690 722–743.

691 Ershov, A. V., Nikishin, A. M., 2004. Recent geodynamics of the Caucasus-Arabia-East Africa

692 Region. Geotectonics, Engl. Transl., 38(2), 123–136.

693 Faccenna C., Bellier, O., Martinod, J., Piromallo, C., Regard, V., 2006. Slab detachment beneath

694 eastern Anatolia: A possible cause for the formation of the North Anatolian fault. Earth and

695 Planetary Science Letters 242, 85– 97.

2 7 696 Faccenna, C., Jolivet, L., Piromallo, C., Morelli, A., 2003. Subduction and the depth of

697 convection in the Mediterranean mantle. Journal Geophysical Research 108, (B2), 2099.

698 Faccenna, C., Becker, T. W., Conrad, C., Husson, L., 2012. Mountain abuilding and mantle

699 dynamics. Tectonics, doi:10.1029/2012TC003176, in press.

700 Forte, A.M., 2007. Constraints on seismic models from other disciplines—implications for

701 mantle dynamics and composition. In Schubert, G., Bercovici, D. (Eds.), Treatise on Geophysics.

702 Elsevier, Amsterdam, pp. 805–858.

703 Garfunkel, Z., Beyth, M., 2006. Constraints on the structural development of Afar imposed by

704 the kinematics of the major surrounding plates, in The Afar Volcanic Province within the East

705 African Rift System, eds Yirgu, G., Ebinger, C.J.& Maguire, P.K.H., Geological Society of

706 London, Special Pub. 259, 23-42.

707 Gashawbeza, E.M., Klemperer, S.L., Nyblade, A.A., Walker, K.T., Keranen, K.M., 2004. Shear-

708 wave splitting in Ethiopia: Precambrian mantle anisotropy locally modified by Neogene rifting,

709 Geophys. Res. Lett. (31), L18602.

710 Gelati, R., 1975, Miocene marine sequence from Lake Van: Eastern Turkey. Rivista Italiana di

711 Paleontologia e Stratigrafia 81, 477–490.

712 Ghebreab W., 1998. Tectonics of the Red Sea region reassessed. Earth Sci. Rev. 45, 1-44.

713 Ghosh, A., Becker, T. W., Zhong, S., 2010. Effects of lateral viscosity variations on the geoid.

714 Geophys. Res. Lett., 37, L01301.

715 Giannérini, G., Campredon, R., Féraud, G., Abou Zakhem, B., 1988. Déformations intraplaques

716 et volcanisme associé: example de la bordure NW de la plaque Arabique au Cénozoique. Bull. Soc.

717 Géol. France IV, 6, 937-947.

718 Grand, S. P., 2002. Mantle shear wave tomography and the fate of subducted slabs. Phil. Trans.

719 R. Soc. Lond. A, 360, pp. 2475-2491.

720 Gudmundsson, O., Sambridge, M., 1998. A regionalized upper mantle (RUM) seismic model. J.

721 Geophys. Res. 103, 7121–7136.

2 8 722 Gurnis, M., Mitrovica, J.X., Ritsema, J., van Heijst, H.-J., 2000. Constraining mantle density

723 structure using geological evidence of surface uplift rates: the case of the African superplume.

724 Geochem. Geophys. Geosyst., 1.

725 Gurnis, M., Turner, M., Zahirovic, S., DiCaprio, L., Spasojevic, S., Müller, R. D., Boyden, J.,

726 Seton, M., Manea, V. C., Bower, D. J., 2012. Plate tectonic reconstructions with continuously

727 closing plates, Computers & Geosciences 38, 35-42.

728 Gvirtzman, Z. , J. Steinberg, J., Bar, O., Buchbinder, B., Zilberman, E., Siman-Tov, R., Calvo,

729 R., Grossowicz, L., Almogi-Labin, A., Rosensaft, M., 2011. Retreating Late Tertiary shorelines in

730 Israel: Implications for the exposure of north Arabia and Levant during Neotethys closure,

731 Lithosphere 3, 2, 95–109.

732 Hager, B. H. and R. W. Clayton, (1989), Constraints on the structure of mantle convection using

733 seismic observations, flow models, and the geoid, in: R. W. Peltier (ed.) Mantle Convection; plate

734 tectonics and global dynamics, Gordon and Breach Science Publishers, New York, 657-763.

735 Hager, B.H., 1984. Subducted slabs and the geoid: constraints on mantle rheology and flow. J.

736 Geophys. Res. 89, 6003–6015.

737 Hansen, S., Schwartz, S., Al-Amri, A., Rodgers, A., 2006. Combined plate motion and density

738 driven flow in the asthenosphere beneath Saudi Arabia: Evidence from shear-wave splitting and

739 seismic anisotropy. Geology 34 (10), 869-872.

740 Hansen, S.E.; Nyblade, A.A., Benoit, M.H., 2012. Mantle structure beneath Africa and Arabia

741 from adaptively parameterized P-wave tomography: Implications for the origin of Cenozoic Afro-

742 Arabian tectonism. Earth and Planetary Science Letters 319-320, 23–34.

743 Hofmann, C., Courtillot, V., Féraud, G., Rochette, P., Yirgu, G., Ketefo, E., Pik, R., 1997.

744 Timing of the Ethiopian flood basalt event and implications for plume birth and global change.

745 Nature 389, 838–841.

746 Huchon, P. Khanbari, K., 2003. Rotation of the syn-rift stress Åëeld of the northern Gulf of

747 Aden margin, Yemen. Tectonophysics 364, 147-166.

2 9 748 Huchon, P. Jestin, F., Cantagrel, J.M., Gaulier, J.M., Khirbash, S.A., Gafaneh, A., 1991.

749 Extensional deformation in Yemen since Oligocene and the Africa-Arabia-Somalia triple junction,

750 Ann. Tecton. 5, 141-163.

751 Jolivet, L., Faccenna, C., 2000. Mediterranean extension and the Africa-Eurasia collision.

752 Tectonics 19(6), 1095-1106.

753 Kaminski, E., Ribe, N. M., Browaeys, J. T. , 2004. D-Rex, a program for calculation of seismic

754 anisotropy due to crystal lattice preferred orientation in the convective upper mantle. Geophys. J.

755 Int. 157, 1–9.

756 Kaviani, A., Hatzfeld, D., Paul, A., Tatar, M. and Priestley, K., 2009. Shear-wave splitting,

757 lithospheric anisotropy, and mantle deformation beneath the Arabia-Eurasia collision zone in Iran.

758 Earth Planet. Sci. Lett. 286, 371-378.

759 Keskin, M., Chugaev, A.V., Lebedev, V.A., Sharkov, E.V., Oyan, V., Kavak O., 2012a. 760 Geochronology and Origin of Mantle Sources for Late Cenozoic Intraplate Volcanism in the Frontal 761 Part of the Arabian Plate in the Karacadağ Neovolcanic Area of Turkey. Part 1. The Results of 762 Isotope-Geochronologic Studies, Journal of Volcanology and Seismology, 6, 31–42. DOI: 763 10.1134/S0742046312060036 764 Keskin, M., Chugaev, A.V., Lebedev, V.A., Sharkov, E.V., Oyan, V., Kavak, O., 2012b.

765 Geochronology and Origin of Mantle Sources for Late Cenozoic Intraplate Volcanism in the Frontal

766 Part of the Arabian Plate, in the Karacadağ Neovolcanic Area of Turkey. Part 2. The Results of

767 Geochemical and Isotope (Sr--Nd--Pb) Studies. Journal of Volcanology and Seismology, 6, 43–70..

768 DOI: 10.1134/S0742046312060048.

769 Keskin, M., Sharkov, E.V., Lebedev, V.A., Chugaev, A.V., Oyan, V., Genç, Ş.C., Ünal, E. and

770 Aysal, N. (2012c). Miocene to Recent Magmatism and Geodynamics of Eastern Turkey, European

771 Union of Geosciences,Vienna, Session: TS6.1/GD5.5 The Alpine-Himalayan convergence zone:

772 from the Mediterranean to SE Asia. http://meetingorganizer.copernicus.org/EGU2012/EGU2012-

773 12748-3.pdf

3 0 774 Keskin, M., 2003. Magma generation by slab steepening and breakoff beneath a subduction–

775 accretion complex: an alternative model for collision-related volcanism in eastern Anatolia, Turkey.

776 Geophys. Res. Lett., 30(24), 8046.

777 Keskin, M., 2007. Eastern Anatolia: A hotspot in a collision zone without a mantle plume, in

778 Foulger, G.R., and Jurdy, D.M., eds., Plates, plumes, and planetary processes. Geological Society of

779 America Special Paper 430, 693–722.

780 Keskin, M., Pearce, J.A. and Mitchell, J.G., 1998. Volcano-stratigraphy and geochemistry of

781 collision-related volcanism on the Erzurum-Kars Plateau, North Eastern Turkey. Journal of

782 Volcanology and Geothermal Research 85/1-4, 355-404.

783 Kreemer, C, Chamot-Rooke, N., 2004. Contemporary kinematics of the southern Aegean and the

784 Mediterranean Ridge. Geophys. J. Int. 157, 1377–92.

785 Krienitz, M.-S., Haase, K. M., Mezger, K., van den Bogaard, P., Thiemann, V., Shaikh-Mashail,

786 M. A., 2009. Tectonic events, continental intraplate volcanism, and mantle plume activity in

787 northern Arabia: constraints from geochemistry and Ar-Ar dating of Syrian lavas. Geochemistry,

788 Geophysics, Geosystems 10, doi:10.1029/2008GC002254.

789 Krienitz, M.-S., Haase, K. M., Mezger, K., Shaikh- Mashail, M. A., 2007. Magma genesis and

790 mantle dynamics at the Harrat Ash Shamah Volcanic Field (southern Syria). J. Petrol. 48, 1513–

791 1542, doi:10.1093/petrology/egm028.

792 Le Pichon, X., 1982. Land-locked oceanic basins and continental collision: the eastern

793 Mediterranean as a case example. In: Mountain Building Processes, ed. K Hsu, pp. 129–46.

794 London: Academic

795 Le Pichon, X., Kreemer, C., 2010. The Miocene-to-Present Kinematic Evolution of the Eastern

796 Mediterranean and Middle East and Its Implications for Dynamics. Annu. Rev. Earth Planet. Sci.

797 38, 323–351.

798 Lebedev, S., Van der Hilst, R.D., 2008. Global upper-mantle tomography with the automated

799 multi-mode surface and S waveforms. Geophysical Journal International 173, 3, 827-843

3 1 800 Lebedev, V.A., Sharkov, E.V., Keskin, M., Oyan, V., 2010. Geochronology of the Late

801 Cenozoic volcanism in the area of Van Lake (Turkey): an example of the developmental dynamics

802 for magmatic processes. ISSN 1028-334X, Doklady Earth Sciences, 2010, 433, 1031–1037, doi:

803 10.1134/S1028334X1008009X.

804 Lei, J., Zhao, D., 2007. Teleseismic evidence for a break-off subducting slab under Eastern 805 Turkey. Earth and Planetary Science Letters 257, 14–28. 806 Lepvrier, C., Fournier, M., Berard, T., Roger, J., 2002. Cenozoic extension in coastal Dhofar

807 (southern Oman): Implications on the oblique rifting of the Gulf of Aden. Tectonophysics 357, 279-

808 293.

809 Li, C., van der Hilst, R. D., Meltzer, A. S., Engdahl, E. R., 2008. Subduction of the Indian

810 lithosphere beneath the Tibetan Plateau and Burma. Earth Planet. Sci. Lett. 274, 157–168.

811 Lithgow-Bertelloni, C., Silver, P. G., 1998. Dynamic topography, plate driving forces and the

812 African superswell. Nature 395, 269-272.

813 Lithgow-Bertelloni, C., Richards, M.A. 1998. The dynamics of Cenozoic and Mesozoic plate

814 motions. Rev. Geophys. 36, 27–78.

815 Long M.D., Becker, T. W., 2010. Mantle dynamics and seismic anisotropy. Earth and Planetary

816 Science Letters 297, 341–354.

817 Lustrino, M., Keskin, M., Mattioli, M., Kavak, O., 2012. Heterogeneous mantle sources feeding

818 the volcanic activity of Mt. Karacadağ (SE Turkey). Journal of Asian Earth Sciences 46, 120–139.

819 doi:10.1016/j.jseaes.2011.11.016.

820 Lustrino, M., Keskin, M., Mattioli, M., Lebedev, V.A., Chugaev, A., Sharkov, E. and Kavak, O.,

821 2010. Early activity of the largest Cenozoic shield volcano in the circum-Mediterranean area: Mt.

822 Karacadağ, SE Turkey. European Journal Mineralogy 22, 343–362, doi: 10.1127/0935-

823 1221/2010/0022-2024.

824 Ma, G.S.K., Malpas, J., Xenophontos, C. and Chan, G.H.N. (2011). Petrogenesis of Latest 825 Miocene-Quaternary Continental IntraplateVolcanism along the Northern Dead Sea Fault System

3 2 826 (Al Ghab-Homs Volcanic Field), Western Syria: Evidence for Lithosphere-Asthenosphere 827 Interaction, Journal of Petrology, 52(2), 401-430. doi:10.1093/petrology/egq085. 828 Magni, V., van Hunen, J., Funiciello, F., Faccenna, C., 2012. Numerical models of slab

829 migration in continental collision zones. Solid Earth 3, 1–14.

830 McQuarrie, N, Stock, JM, Verdel C, Wernicke, BP., 2003. Cenozoic evolution of Neotethys and

831 implications for the causes of plate motions. Geophys. Res. Lett. 30, 2036.

832 Mégnin, C., B. Romanowicz, B. 2000. The 3D shear velocity structure of the mantle from the

833 inversion of body, surface and higher mode waveforms. Geophys. J. Inter., 143, 709-728.

834 Miller, M. S., Becker, T. W., 2012. Mantle flow deflected by interactions between subducted

835 slabs and cratonic keels. Nature Geoscience 5, 726–730, doi:10.1038/ngeo1553

836 Molnar, P., Atwater, T., 1979. Interarc spreading and cordilleran tectonics as alternates related to

837 the age of subducted oceanic lithosphere. Earth Planet. Sci. Lett. 41, 330-340, 1979.

838 Montelli, R., Nolet, G., Dahlen, F. A., Masters, G., 2006. A catalogue of deep mantle plumes:

839 New results from finite-frequency tomography., Geochem. Geophys. Geosyst. 7111, 2006

840 Moucha, R., Forte, A., 2011. Changes in African topography driven by mantle convection.

841 Nature Geoscience 4, 10, 707-712.

842 Müller, R. D., Sdrolias, M., Gaina, C., Steinberger, B., Heine, C., 2008. Long-term sea-level

843 fluctuations driven by ocean basin dynamics. Science 319, 1357-1362.

844 Nataf, H-C., Ricard, Y., 1996. 3SMAC : an a priori tomographic model of the upper mantle

845 based on geophysical modeling. Phys. Earth Planet. Inter. 95, 101-122.

846 Okay, A.I., Zattin, M., Cavazza, W., 2010. Apatite fission-track data for the Miocene Arabia-

847 Eurasia collision. Geology 38, 1, 35–38;

848 Park, Y., Nyblade, A. A., Rodgers, A., Al-Amri, A., 2007. Upper mantle structure beneath the

849 Arabian Peninsula and northern Red Sea from teleseismic body wave tomography: Implications for

850 the origin of Cenozoic uplift and volcanism in the Arabian Shield. Geochem. Geophys. Geosyst. 8,

851 Q06021, doi:10.1029/2006GC001566.

3 3 852 Park, Y., Nyblade, A. A., Rodgers, A., Al-Amri, A., 2008. S wave velocity structure of the

853 Arabian Shield upper mantle from Rayleigh wave tomography. Geochem. Geophys. Geosyst. 9,

854 Q07020, doi:10.1029/2007GC001895.

855 Partridge, T.C, Maud, R.R., 1987. Geomorphic evolution of southern Africa since the Mesozoic.

856 S. Afr. J. Geol., 90, 165–184.

857 Pik, R., Marty, B., Carignan, J., Lavé, J., 2003. Stability of the Upper Nile drainage network

858 (Ethiopia) deduced from (U-Th)/He thermochronometry: implications for uplift and erosion of the

859 Afar plume dome. Earth and Planetary Science Letters 215, 73-88.

860 Piromallo, C., Morelli, A., 2003. P-wave tomography of the mantle under the Alpine-

861 Mediterranean area. J. Geophys. Res. 108, 2065, doi:10.1029/2002JB001757.

862 Reilinger, R., McClusky, S., Vernant, P., Lawrence, S., Ergintav, S., Cakmak, R., Ozener, H.,

863 Kadirov, F., Guliev, I., Stepanyan, R., Nadariya, M., Hahubia, G., Mahmoud, S., Sakr, K.,

864 ArRajehi, A., Paradissis, D., Al-Aydrus, A., Prilepin, M., Guseva, T., Evren, E., Dmitrotsa, A.,

865 Filikov, S.V., Gomez, F., Al-Ghazzi, R., Karam, G., 2006. GPS constraints on continental

866 deformation in the Africa–Arabia–Eurasia continental collision zone and implications for the

867 dynamics of plate interactions. J. Geophys. Res. 111, B05411. doi:10.1029/2005JB004051.

868 Reilinger, R., McClusky, S., 2011. Nubia–Arabia–Eurasia plate motions and the dynamics of

869 Mediterranean and Middle East tectonics. Geophys. J. Int. 186, 971–979.

870 Replumaz, A., Karason, H. , van der Hilst, R.D., Besse, J., Tapponnier, P., 2004. 4-D evolution

871 of SE Asia’s mantle from geological reconstructions and seismic tomography. Earth and Planetary

872 Science Letters 221, 103-115.

873 Reuter M., Piller, W. E., Harzhauser, M., Mandic, O., Berning, B., Rogl, F., Kroh, A., Aubry,

874 M.-P., Wielandt-Schuster, U., Hamedani, A., 2009. The Oligo-/Miocene Qom Formation (Iran):

875 evidence for an early Burdigalian restriction of the Tethyan Seaway and closure of its Iranian

876 gateways. Int J Earth Sci (Geol Rundsch) 98, 627–650.

3 4 877 Ricard, Y., Vigny, C., 1989. Mantle dynamics with induced plate tectonics. J. Geophys. Res. 94,

878 17 543–17,559..

879 Ricard, Y., Fleitout, L., Froidevaux, C., 1984. Geoid heights and lithospheric stresses for a

880 dynamic Earth. Ann. Geophys. 2, 267–286.

881 Ricard, Y., M.A. Richards, C. Lithgow-Bertelloni, and Y. LeStunff (1993), Geodynamic model

882 of mantle density heterogeneity, J. Geophys. Res. 98, 21895–21909.

883 Richards, M., Hager, B.H., 1984. Geoid anomalies in a dynamic Earth. J. Geophys. Res. 89,

884 5987–6002.

885 Ritsema J, van Heijst HJ, Deuss A, Woodhouse JH, S40RTS: a degree-40 shear-velocity model

886 for the mantle from new Rayleigh wave dispersion, teleseismic traveltimes, and normal-mode

887 splitting function measurements, Geophys. J. Int., 184, 10.1111/j.1365–246X.2010.04884.x, 2011.

888 Ritsema, H.J., J.H. van Heijst, and J.H. Woodhouse (1999), Complex shear velocity structure

889 beneath Africa and Iceland, Science, 286, 1925–1928.

890 Roger, J., Platel, J.P., Cavelier, C., Grisac, C.B.D., 1989. Donnees nouvelles sur la stratigraphie

891 et l’histoire geologique du Dhofar (Sultanat d’Oman). Bull. Soc. Geol. France 5, 265-277.

892 Şengör A.M.C AMC, Özeren S, Genç T, Zor E. 2003. East Anatolian high plateau as a mantle-

893 supported, north-south shortened domal structure. Geophys. Res. Lett. 30:2003

894 Şengör, A.M.C., Özeren, M.S., Keskin, M., Sakınç, M., Özbakır, A.D. and Kayan, I. (2008),

895 Eastern Turkish high plateau as a small Turkic-type orogen: implications for post-collisional crust-

896 forming processes in Turkic-type orogens, Earth Science Reviews, 90(1-2), 1-48. doi:

897 10.1016/j.earscirev.2008.05.002.

898 Şengör, A.M.C., 1979. The North Anatolian transform fault: its age, offset and tectonic

899 significance, J. Geol. Soc. (Lond.) 136 , 269– 282.

900 Shaw, J.A., Baker, J.A., Menzies, M.A., Thirlwall, M.F. and Ibrahim, K.M. (2003), Petrogenesis

901 of the largest intraplate volcanic field on the Arabian Plate (Jordan): a mixed lithosphere-

3 5 902 asthenosphere source activated by lithosphere extension, Journal of Petrology, 44(9), 1657-1679.

903 doi: 10.1093/petrology/egg052

904 Siebert, L. and Simkin, T. Volcanoes of the World: an Illustrated Catalog of Holocene

905 Volcanoes and their Eruptions, Smithsonian Institution, 2002-, GVP-3, Global Volcanism Program

906 Digital Information Series, Available online at www.earthchem.org.

907 Simmons, N.A., Forte, A.M., Boschi, L., Grand, S.P., 2010. GyPSuM: a joint tomographic

908 model of mantle density and seismic wave speeds. J. Geophys. Res. 115, B12310.

909 doi:10.1029/2010JB007631.

910 Simmons, N.A., Forte, A.M., Grand, S.P., 2007. Thermochemical structure and dynamics of the

911 African superplume. Geophys. Res. Lett. 34(2), L02301, doi: 10.1029/2006GL028009.

912 Simmons, N., Forte, A., Grand, S. P., 2009. Joint seismic, geodynamic and mineral physical

913 constraints on three-dimensional mantle heterogeneity: Implications for the relative importance of

914 thermal versus compositional heterogeneity. Geophys. J. Int., 177, 1284-1304, doi:10.1111/j.1365-

915 246X.2009.04133.x.

916 Stadler, G., Gurnis, M., Burstedde, C., Wilcox, L. C., Alisic, L., Ghattas, O., 2010. The

917 dynamics of plate tectonics and mantle flow: from local to global scales. Science 329, 1033-1038.

918 Steinberg, J., Gvirtzman, Z., Folkman, Y. Garfunkel, Z., 2011. Origin and nature of the rapid late

919 Tertiary filling of the Levant Basin. Geology 39, 4, 355-358

920 van der Voo, R., Spakman, W., Bijwaard, H., 1999. Tethyan subducted slabs under India. Earth

921 Plan. Sci. Lett. 171, 7–20.

922 van Hinsbergen, D.J.J., Steinberger, B. Doubrovine, P., Gassmöller, R., 2011. Acceleration-

923 deceleration cycles of India-Asia convergence: roles of mantle plumes and continental collision, J.

924 Geophys. Res., 116, B06101, doi:10.1029/2010JB008051.

925 van Hinsbergen, D.J.J., Hafkenscheid, E., Spakman, W., Meulenkamp, J.E., Wortel, R., 2005.

926 Nappe stacking resulting from subduction of oceanic and continental lithosphere below Greece.

927 Geology 33(4): 325-328, doi: 10.1130/G20878.1.

3 6 928 van Hinsbergen, D.J.J., Kaymakcı, N., Spakman, W., Torsvik, T.H., 2010. Reconciling the

929 geological history of western Turkey with plate circuits and mantle tomography. Earth and

930 Planetary Science Letters 297, 674-686.

931 Watchorn, F., Nichols, G.J., Bosence, D.W.J., 1998. Rift related sedimentation and stratigraphy,

932 southern Yemen (Gulf of Aden). In: B.H. Purser, D.W.J. Bosence (Eds.), Sedimentation and

933 Tectonics of Rift Basins: Red Sea-Gulf of Aden, Chapman and Hall, London, 165-189.

934 Wüstefeld, A., Bokelmann, G. H. R., Barruol, G., Montagner, J. P., 2009. Identifying global

935 seismic anisotropy patterns by correlating shear-wave splitting and surface-wave data. Phys. Earth

936 Planet. Int., 176, 198-212. Database available updated online at http://www.gm.univ-

937 montp2.fr/splitting/DB/.

938 Yarmolyuk, V.V., Bogatikov, O.A., Kovalenko, V.I., 2004. Late Cenozoic Transcontinental

939 Features and the Magmatism in the Earth's Euro-African Segment and the Geodynamics of Their

940 Evolution. DAN 395, 1, 91-95.

941 Zhong, S., Zuber, M. T., Moresi, L., Gurnis, M., 2000. Role of temperature-dependent viscosity

942 and surface plates in spherical shell models of mantle convection. J. Geophys. Res. 105, 11063-

943 11082.

944

945

946

3 7 946 References

947 Figure and Table Captions

948

949 Figure 1: Topography and deformation indicators for the Arabia-Anatolia-Aegean system. a)

950 Geodetic velocity field (ArRajehi et al., 2010) in a Eurasia fixed reference frame (orange vectors)

951 compared to geological plate motions from NUVEL1A (de Mets et al., 1994; dark gray vectors),

952 topography from ETOPO1, and seismicity (Engdahl et al., 1998), color coded by depth, z, with

953 major geographic features labelled. b) Residual topography, zres, estimated by correcting smoothed

954 actual topography for crustal isostatic adjustment using the CRUST2.0 model (Bassin et al., 2000)

955 using the same parameters as in Becker and Faccenna (2011). Open triangles are Holocene

956 volcanoes from Siebert and Simpkin (2002), and colored inverted triangles show all volcanic rocks

957 younger than 50 Ma in the EarthChem database compilation available at www.earthchem.org as of

958 05/2012. Plate boundaries (dark green) are modified from NUVEL1 (de Mets et al., 1994) to

959 include an Anatolian plate, as in Becker and Faccenna (2011). c) Dynamic topography estimated by

960 converting free-air gravity anomalies using a 50 mGal/km conversion, assuming a mantle density of

961 3300 kg/m3 (Craig et al., 2011). d) SMEAN composite tomography model (Becker and Boschi,

962 2002) averaged over uppermost mantle depths from 150 to 400 km (see also Figs. 2 and S1). Both

963 c) and d) are superimposed with SKS splitting observations (from the compilation of Wuestefeld et

964 al., 2009, updated 05/2012) where yellow stick orientation shows the “fast axes” of azimuthal

965 anisotropy, c) for all available splits, and d) for station-averaged estimates.

966

967 Figure 2: Comparison of tomographically imaged structure underneath the study region (also see

968 Fig. S1). Maps a-c show 250 km depth layer of models HNB12 (Hansen et al., 2012), SMEAN

969 (Becker and Boschi, 2002), and TX2008 (Simmons et al., 2007), respectively. Smoothed elevation

970 contours are overlaing in 2000 m contours for positive topography, plate boundaries (dark green)

971 from Bird (2002). Dark cyan profiles are plotted in the underneath (d-f), correponding to the model

3 8 972 shown in map above; dashed line indicates 660 km.

973

974 Figure 3: Mantle flow solutions for four, illustrative example models for the AAA system. Left

975 column: surface velocity field (white arrows) compared with geodetic velocity (orange arrows,

976 averaged from the original vectors shown in Fig. 1a) with respect to Eurasia reference frame, with

977 background indicating dynamic topography estimates. Right column: Shear wave splitting

978 measurements indicated by sticks, colored by angular misfit of “fast axes” with model predictions

979 (gray) whose back-azimuthal variations are denoted by wedges and two different stick lengths for

980 fast axes and delay times, respectively (see Becker et al., 2006b, and Miller and Becker, 2012, for

981 details). Inset shows the regional arithmetic averages for the angular misfit, <Δα>, and the delay

982 time misfit, <Δ(δt)>, for all 334 non-zero measurements of Fig. 1d. a) Flow produced by upper

983 mantle density anomaly assigned to the Wadati-Benioff zones only (slab driven flow); b) Flow

984 produced by density anomaly within Wadati-Benioff zones and lower mantle convection produced

985 by density anomaly scaled from the SMEAN model; c) Flow produced by density anomaly within

986 the Wadati-Benioff zones, lower mantle anomalies, and convection produced by slow velocity

987 anomaly only (from SMEAN) in the upper mantle. d) As in b) but with a low viscosity channel

988 extending northward from Afar to Israel, sensu Hansen et al. (2006).

989

990 Figure 4: Mantle flow predictions at three different depths: a) 210, b) 345, and, c), 570 km for the

991 model of Fig. 3c, with vectors and background shading denoting horizontal and radial velocities,

992 respectively, and velocities within three cross sections (d-f) as marked on a), with background

993 showing non-dimensional temperature.

994

995 Figure 5: Misfit matrices for a subset of all global flow models considered (cf. Miller and Becker,

996 2012), illustrating the effect of different choices of density (columns) and lateral viscosity variation

997 (rows) models for, a), global plate motion correlations, b), regional plate motion correlations, and,

3 9 998 c), average angular misfits of shear wave splitting predictions . White boxes show the models

999 illustrated in Fig. 3. For legend, see Table 1.

1000

1001 Figure 6: Tectonic evolution of the AAA system at four stages at 30-33 (a), 20-23 (b), 10-12 (c),

1002 and 3-5 Ma (d) and relative, idealized cross sections (e-h), cf. Fig. 4 for the present-day. Plate

1003 motions are from GPlates (Muller et al., 2008; Gurnis et al., 2012). Geological features from Barrier

1004 and Vrielynck (2010). Red indicates active volcanic fields (undistinguished); blue indicates deep

1005 basins; red lines indicate active subduction zone; black lines indicate collisional thrust system;

1006 yellowish background area indicate regions under influence of Afar plume; arrows indicate

1007 direction of mantle flow.

1008

1009 Figure 7: Geochemical isotope Diagram for 206Pb/204Pb vs. 207Pb/204Pb for the lavas of the

1010 Karacadağ neovolcanic area (data from Keskin et al., 2012b). Data source for the fields drawn in

1011 this diagram are listed in Ma et al. (2011), except for two basaltic samples from the Erzurum-Kars

1012 Plateau (i.e. EKP in the legend) which are from Keskin et al. (1998).

1013

1014 Figure 8: Conceptual cartoon of mantle flow in the AAA system as inferred from our analysis of

1015 geological and geophysical models.

1016

1017 Table 1: Legend for density and viscosity models considered in the misfit matrix plots of Fig. 5.

1018

4 0 1018 Table 1

1019

1020 Mantle density models.

1021 Columns in Fig. 5 denote density models tested (details of model parameters are described in

1022 Becker and Faccenna, 2011, and Miller and Becker, 2012):

MITP08 Li et al. (2008) (MIT gorup) global P wave model, scaled to temperature in whole mantle HNB12 Hansen et al. (2012) P wave model in whole mantle, building on MITP08, highlighting effect of regional resolution improvement. LH08 Lebedev and van der Hilst (2008) SV wave model in upper mantle, SMEAN (Becker and Boschi, 2002) in lower mantle SMEAN Composite S wave model with anomalies underneath continental keels down to 300 km removed RUM Density anomalies in upper mantle only, as inferred from Wadati-Benioff zone geometry as assigned from the RUM model (Gudmundsson and Sambridge, 1998). RUM+SMEAN_lm RUM plus density anomalies in the lower mantle inferred from SMEAN tomography RUM+SMEAN_neg Like RUM+SMEAN_lm, but additionally slow (“hot”) velocity anomalies in the upper mantle from SMEAN, as in Ghosh et al. (2010) RUM+SMEAN_neg+SMEAN_100 Like RUM+SMEAN_neg, but additionally all velocity anomalies in the uppermost 100 km (“lithospheric anomalies”) from SMEAN TX2008 Global mantle temperature anomalies inferred from the TX2008 S wave tomography model of Simmons et al. (2007) TX2008+RUM RUM upper mantle slabs as in RUM, lower mantle density anomalies from TX2008. 1023 1024 Models of lateral variations in viscosity as shown in different rows of the misfit matrices of Fig. 5 reference Only radial viscosity variations besides weak zone geometry prescribed in the lithosphere with modified NUVEL1 geometry as in Becker and Faccenna (2011), as in Fig. 4. Bird_boundaries Same width and viscosity reduction for weak zones as in reference, but geometry taken from Bird (2003), as in Fig. 2. Bird_wo_Arabia Same as Bird_boundaries, but no weak zone between NW Arabia and Eurasia keels Same as reference, but stiff (factor 500 viscosity increase) continental keels down to 300 km depth as inferred from the

4 1 3SMAC model (Nataf and Ricard, 1996) asthenosphere Sub-oceanic plate reduction of viscosity by a factor of 0.01 in the depths between 100 and 300 km keels+asthenosphere Combined effect of keels and asthenosphere. simple_channel Reference model with factor 0.01 viscosity reduction along a channel from Afar to Isreal between 100 and 300 km depth. channel-LVV Reference model with lateral viscosity variations inferred from scaling the depth-averaged velocity anomalies in LH08 between 100 and 300 km depth to ~four orders of magnitude lateral variations of viscosity (cf. Fig. 1d), yielding a low viscosity channel similar to simple_channel. channel-LVV(z) As channel-LVV but allowing for depth-variations in lateral viscosity variations from 100 to 300 km depth. 1025 1026

4 2 *Highlights (for review)

-The Middle East region represents a key site within the Tethyan domain where continental

break-up, collision, backarc extension and escape tectonics are linked together.

-We perform global mantle circulation computations to test the role of slab pull and mantle

upwellings as driving forces for the kinematics of the Arabia-Anatolia-Aegean (AAA) system and

we teste our result with geodesy, residual topography, and shear wave splitting observations.

- The best match to these crustal motions can be obtained by combining the effect of slab pull

exerted in the Aegean with a mantle upwelling underneath Afar and, more generally, with the large-

scale flow associated with a whole mantle;

-We suggest that the tectonic evolution is related to the progressive northward entrainment of

upwelling mantle material from Afar associated with the downwelling Tethyan slab;

-Our model reconciles Afar plume volcanism, the collision in the Bitlis mountains and northern

Zagros, and the rapid increase of Aegean trench rollback in a single coherent frame of large scale

mantle convection, initiated during the last ~40 Ma.

20˚ 30˚ 40˚ 50˚ 60˚ 20˚ 30˚ 40˚ 50˚ 60˚ Figure1z [km] zres [km] Click here to downloadAnatolia Figure: figure1.pdf 2 2 40˚ 40˚ 1 0 Aegea Bitlis 0 -1 -2 -2

30˚ 30˚ Arabia

d [km] age [Ma] 200 50

40

20˚ 30 20˚ 100 Africa 20 10

0 0 Gulf of Aden Afar 10˚ 10˚ Ethiopian a) 4 cm/yr Rift b) 20˚ 30˚ 40˚ 50˚ 60˚ 20˚ 30˚ 40˚ 50˚ 60˚ zeq [km] δv [%] 2 2

40˚ 40˚ 1 1

0 0

-1 -1

-2 -2

30˚ 30˚

20˚ 20˚

10˚ 10˚ δt = 2 s δ c) SKS d) tSKS = 2 s

Figure 1 a)Figure220˚ 30˚ A 40˚ 50˚ 60˚ b) 20˚ 30˚ A 40˚ 50˚ 60˚ c) 20˚ 30˚ A 40˚ 50˚ 60˚ hnb12Click here to download Figure: figure2.pdf40˚ smean 40˚ tx2008 40˚

δvP [%] δvS [%] δvS [%] 2

2 2 1 30˚ 30˚ 30˚

0 0 0

-1 -2 -2

-2

20˚ 20˚ 20˚

10˚ 10˚ 10˚

B B B d) e) f) x [km] x [km] x [km] 0 0 0 -1000 1000 -1000 1000 -1000 1000 A 0 2000 B A 2000 B A 2000 B -200 -2000 -2000

] 0 6000 600 6000

r [km] r [km r [km]

0 0 550 550 5500

δvS δvS δvS -1 0 1 -2 -1 0 1 2 -2 -1 0 1 2

Figure 2 Figure3 Click here to download Figure: figure3.pdf

20˚ 30˚ 40˚ 50˚ 60˚ δzdyn [km] 2

40˚ 1 2 cm/yr

0

-1

-2

30˚

20˚

a) 10˚ 20˚ 30˚ 40˚ 50˚ 60˚ δzdyn [km] 2

40˚ 1 2 cm/yr

0

-1

-2

30˚

20˚

b) 10˚

20˚ 30˚ 40˚ 50˚ 60˚ δzdyn [km] 2

40˚ 1 2 cm/yr

0

-1

-2

30˚

20˚

c) 10˚ 20˚ 30˚ 40˚ 50˚ 60˚ δzdyn [km] 2

40˚ 1 2 cm/yr

0

-1

-2

30˚

20˚

d) 10˚

Figure 3 Figure4 Click here to download Figure: figure4.pdf 20˚ 30˚ 40˚ 50˚ 60˚ 20˚ 30˚ 40˚ 50˚ 60˚ vr [cm/yr] v [cm/yr] B r 4 4 40˚ 40˚ 2 cm/yr 2 cm/yr 0 z = 210 km z = 345 km F E 0

-4 D -4

30˚ 30˚

20˚ 20˚

smean_al+RUM smean_al+RUM ref ref C 10˚ 10˚ a) A b)

20˚ 30˚ 40˚ 50˚ 60˚ vr [cm/yr]

4 40˚ 2 cm/yr distance along track [km] 0 z = 570 km F E -1000 -500 0 500 1000

-4 0

30˚ -200

-400

-600

20˚ -800 depth [km]

smean_al+RUM -1000

ref -1200 10˚ T c) d) 0.3 0.4 0.5 0.6 0.7 B distance along track [km] A -2000 -1500 -1000 -500 0 500 1000 1500 2000 0

-200

-400

-600 depth [km] -800

-1000 e) -1200 T 0.3 0.4 0.5 0.6 0.7

C distance along track [km] D -1500 -1000 -500 0 500 1000 1500 0

-200

-400

-600 depth [km] -800

-1000 f) -1200 T 0.3 0.4 0.5 0.6 0.7 Figure 4 Figure5 Click here to download Figure: figure5.pdf

MITP08HNB12 LH08 SMEANRUM RUM+SMEAN_lmRUM+SMEAN_negRUM+SMEAN_neg+SMEAN_100TX2008TX2008+RUM a) 0.2 0.3 0.7 0.8 0.7 0.7 0.8 0.8 0.7 0.7 reference 0.6

Bird_boundaries 0.6 0.9 0.8 Bird_w/o_Arabia 0.6 0.7 keels 0.6 0.6 vel

0.5 r asthenosphere 0.8 0.4 keels+asthenosphere 0.8 0.3 0.2 simple_channel 0.8

channel-LVV 0.6

channel-LVV(z) 0.6

0.4 0.7 0.7 0.7 0.7 0.6 0.7 0.7 0.6 0.6 b) reference 0.6

Bird_boundaries 0.6 0.7 Bird_w/o_Arabia 0.6

0.6 keels 0.7

0.5 vel reg

asthenosphere 0.6 r

0.4 keels+asthenosphere 0.6

0.3 simple_channel 0.6

channel-LVV 0.6

channel-LVV(z) 0.6

54.8 44.7 47.6 34.7 40.2 34.2 36.4 39.4 44.5 43.5 c) reference 40.1

Bird_boundaries 46.2 55 Bird_w/o_Arabia 45.3 50 ] keels 41.0 o 45 [

asthenosphere 41.5 40 ∆α keels+asthenosphere 41.7 35 30 simple_channel 39.2

channel-LVV 42.0

channel-LVV(z) 41.0

LH08 RUM MITP08 HNB12 SMEAN TX2008

TX2008+RUM RUM+SMEAN_lm RUM+SMEAN_neg

RUM+SMEAN_neg+SMEAN_100

Figure 5 Figure6 Click here to download Figure: figure6.pdf about 30 Ma Cyprus A a) b) subduction Arabia Red Sea Nubia MER B 0 0

-200 -200

-400 -400

-600 -600

-800 -800

depth [km] -1000 -1000

-1200 -1200 -2000 -1500 -1000 -500 0 500 1000 1500 2000 distance along track [km] e) upper about 20 Ma A Cyprus B mantle subduction Arabia Red Sea Nubia MER ow 0 0 -200 -200

-400 -400

-600 -600

-800 -800

depth [km] -1000 -1000 depth [km]

-1200 -1200 -2000 -1500 -1000 -500 0 500 1000 1500 2000 distance along track [km] 30 - 33 Ma 20 - 23 Ma about 10 Ma Collision - slab break o f) Dead Sea Cyprus Transform A NAF B subduction Arabia Red Sea Nubia MER 0 0 c) d) -200 -200 -400 -400

-600 -600

-800 -800

depth [km] -1000 -1000 depth [km]

-1200 -1200 -2000 -1500 -1000 -500 0 500 1000 1500 2000 distance along track [km] g) present day Dead Sea Cyprus Transform A NAF B Arabia subduction Arabia Red Sea Nubia MER Nubia Arabia 0 0 Nubia -200 -200

-400 -400

-600 -600

-800 -800

depth [km] -1000 -1000 depth [km]

-1200 -1200 -2000 -1500 -1000 -500 0 500 1000 1500 2000 T Somalia 0.3 0.4 0.5 0.6 0.7 10 Ma 3-5 Ma distance along track [km] h)

Figure 6 Figure7 Click here to download Figure: figure7.pdf

15,8 Crust or metasomatised ? mantle Saudi Arabian intraplate 15,7 volcanism NHRL 207Pb/204Pb 15,6

S Syria, Jordan & Israel Red Sea MORB 15,5 Afar plume

206Pb/204Pb 15,4 18,0 18,3 18,6 18,9 19,2 19,5

Figure8 Click here to download Figure: figure8.pdf

10 Ma

20 Ma

35 Ma

Figure 8