1 Ivrea mantle wedge, arc of the Western ,

2 and kinematic evolution of the Alps-Apennines

3 orogenic system

4 Stefan M. Schmid1, Eduard Kissling1, Tobias Diehl1, Douwe J.J vanHinsbergen2,

5 Giancarlo Molli3

6 1Institute of Geophysics, ETH Zürich, Sonneggstrasse 5, 8092 Zürich, Switzerland

7 2Department of Earth Sciences, Utrecht University, Heidelberglaan 2, 3584 CS, Utrecht, The

8 Netherlands

9 3Dipartimento di Scienze della Terra, Università di Pisa, Pisa,

10 Manuscript submitted to Swiss Journal of Geosciences, version October 2016

11 Abstract

12 The construction of five crustal-scale profiles across the and the Ivrea

13 mantle wedge integrates up-to-date geological and geophysical information and

14 reveals important along strike changes in the overall structure of the crust of the

15 Western Alpine arc. Tectonic analysis of the profiles, together with a review of the

16 existing literature allows for proposing the following multistage evolution of the arc

17 of the Western Alps: (1) Exhumation of the mantle beneath the Ivrea Zone to shallow

18 crustal depths during Mesozoic is a prerequisite for the formation of a strong Ivrea

19 mantle wedge whose strength exceeds that of surrounding mostly quartz-bearing

20 units, and consequently allows for indentation of the Ivrea mantle wedge and

21 eastward back-thrusting of the western Alps during Alpine orogeny. (2) A first early

22 stage (pre-35 Ma) of the West-Alpine orogenic evolution is characterized by top-

1 23 NNW thrusting in sinistral transpression causing at least some 260km displacement

24 of internal Western Alps and E-W-striking Alps farther east, together with the Adria

25 micro-plate, towards N to NNW with respect to stable Europe. (3) The second stage

26 (35-25 Ma), further accentuating the arc, is associated with top-WNW thrusting in the

27 external zones of the central portion of the arc and is related to the lateral indentation

28 of the Ivrea mantle slice towards WNW by some 100-150km. (4) The final stage of

29 arc formation (25-0 Ma) is associated with orogeny in the Apennines leading to

30 oroclinal bending in the southernmost Western Alps in connection with the 50°

31 counterclockwise rotation of the Corsica-Sardinia block and the .

32 Analysis of existing literature data on the Alps-Apennines transition zone reveals that

33 substantial parts of the Northern Apennines formerly suffered Alpine-type shortening

34 associated with an E-dipping Alpine subduction zone and were backthrusted to the

35 NE during Apenninic orogeny that commences in the Oligocene.

36

37

38 A. Introduction

39 The Western Alps represent a prime example of a very arcuate mountain belt

40 curving around from an E-W strike with dominant top-N nappe stacking in the

41 to a N-S strike in the northern Western Alps and finally into an E-W

42 strike with top-S nappe stacking in the Ligurian Alps straddling the Mediterranean

43 coast (Fig. 1). The question after the mechanisms of arc formation is an old and much

44 debated issue in many orogens (Marshak 1988) involving, in the case of the Alps,

45 elements of oroclinal bending of an initially straight orogen (Carey 1955, van der

46 Voo 2004), induced by rigid body rotation of the hinterland (Vialon et al. 1989)

47 and/or by W-directed indentation of the Adria microplate (Laubscher 1971; Schmid

2 48 & Kissling 2000). Transpressional movements associated with large top NNW

49 displacements of portions of the inner zone of the Western Alps that occur at a small

50 angle to the present-day N-S strike of the Western Alps were also invoked for being

51 responsible for explaining the curvature from the Eastern Alps into the N-S striking

52 Western Alps (Ricou & Siddans 1986). There have also been attempts to explain

53 diverging kinematic indicators along the arc of the Western Alps ranging from top-N

54 Western Switzerland to top-SW in southern France without taking resort to a multi-

55 stage evolution allowing for strike-slip movements and/or large rotations (Platt et al.

56 1989; their group 3 kinematic indicators formed in the 40-15 Ma time interval).

57 However, as pointed out by Butler (1986; his Fig. 2), strongly divergent thrust

58 transport directions over large distances inevitably result in excessive orogen-parallel

59 stretching.

60 In the Western Alps many previous workers (e.g. Laubscher 1971; Butler 1983)

61 recognized the crucial role of the Ivrea mantle wedge residing at relatively shallow

62 crustal levels and locally reaching the surface. Novel geophysical methods such as

63 teleseismic tomography aiming at the structure of the earth’s mantle (Lippitsch et al.

64 2003) and crustal tomography based on local earthquake data (Diehl et al. 2009),

65 together with older information based on controlled source seismology (Blundell et

66 al. 1992), led to substantial recent improvements in revealing the deep structure of the

67 arc of the Western Alps. Th new geophysical data set of unprecedented quality calls

68 for integration with a wealth of geological data collected over many decades to

69 elucidate the formation of the arc of the Western Alps that also demands to analyze

70 the Alps-Apennines transition area located in the Ligurian Alps and the Northern

71 Apennines (Molli et al. 2010).

72 This contribution integrates the new results of crustal tomography and

3 73 improved knowledge on Moho depth (Spada et al. 2013) with a geological

74 interpretation of five crustal-scale transects across the Western Alps. Thereby we

75 focus on the role of the Ivrea mantle wedge. Another important aim of our

76 geological-geophysical interpretations is to put new constraints on the question as to

77 how much continental crust was subducted in the Alps and what the variations in pre-

78 orogenic crustal thickness might have been (see discussions by, e.g., Ménard &

79 Molnar 1991; Butler et al. 2013; Mohn et al. 2014).

80 Two transects are revised versions of previously published profiles (Schmid et

81 al. (2004) while three of them were newly constructed. Special emphasis is given to

82 improved mapping of the outlines of the Ivrea mantle wedge based on geophysical

83 data. This is then followed by an analysis of the complex multi-stage evolution of the

84 arc of the Western Alps and an attempt to map the associated displacement history of

85 geological units of the Western Alps-Northern Apennines system using plate

86 reconstruction methodology (GPlates free software (Boyden et al., 2011) embedded

87 in the global plate kinematic reconstruction of Seton et al. 2012). Finally, we provide

88 a review-type discussion of the Alps-Apennines transition area and discuss

89 controversial issues concerning questions of multi-stage vs. single subduction

90 systems in the Apennines.

91

92

93 B. Crustal-scale profiles across the Western Alps and the Ivrea mantle wedge

94 integrating geological and geophysical information

95 To constrain the geometry of the Ivrea mantle wedge, and as a basis for

96 discussing its role during the formation of the arc of the Western Alps, we

97 constructed a series of crustal scale profiles along the arc of the Western Alps (Fig. 1).

4 98 These profiles integrate geological data with recent geophysical data focusing on the

99 deep structure. The construction of the profiles is based on a series of assumptions,

100 the most important one being that advocated by Schmid & Kissling (2000), namely

101 that the European lower crust is decoupled from the upper crust and goes into

102 subduction together with the underlying mantle for considerable distance below the

103 Ivrea mantle wedge. This invokes a basal detachment of the external thrust sheets at

104 the interface between upper and lower crust at a depth of 20-30km. While this

105 assumption is compatible with the velocity structure (see below) and recent

106 geophysical findings of Zhao et al. (2015) it is by no means generally accepted (see

107 discussion in Butler 2013). However, there are considerable uncertainties about the

108 exact pre-orogenic thickness of the European and Briançonnais continental crust in

109 more internal parts of the profiles and this is one reason why we restrain from area

110 balancing over the entire length of our transects. The main reason is that transport

111 directions changed with time during the formation of the arc of the Western Alps,

112 making 2D area balance along the transects impossible.

113 The geological literature used for constructing the individual profiles varies

114 and hence will be cited when presenting the profiles. However, we used one and the

115 same set of geophysical data for constraining the deep structure in all the profiles.

116 Firstly, we used the results of a tomographic high-resolution 3-D P-wave model of

117 the Alpine crust obtained by Diehl et al. (2009). This 3-D velocity model was used

118 for constructing parts of the sections of Fig. 2 that present contours of equal P-wave

119 velocity superimposed with a geological interpretation. When comparing the velocity

120 contours with the geological interpretation it has to be kept in mind, however, that

121 these contours have a low spatial resolution since they are based on assigning a given

122 velocity to a sizable cell volume of 20x20 km (horizontal) and 15km (vertical).

5 123 Secondly, we complemented these tomographic data by geophysical data better

124 constraining Moho-depth. Where available we took Moho depths from Spada et al.

125 (2013; their Fig. 11; see also Wagner et al. 2012 regarding the methodology) who

126 combined results of controlled-source seismology with local receiver functions Moho

127 depth maps and local earthquake tomography-derived Moho maps.

128 Comparison between Moho depth compiled by Spada et al. (2013),

129 characterized by a P-wave velocity jump, and the velocity model of Diehl et al.

130 (2009) that defines Moho depth in terms a high velocity gradient in the range of 6.5–

131 8.0 km s−1, noticeable over large parts of the profiles, shows that on average the

132 Moho as defined by Spada et al. (2013) best corresponds with the velocity contour of

133 7.25 km s−1 in areas characterized by a subhorizontal or only weakly inclined Moho

134 (Diehl et al. 2009). In order to define the strongly inclined boundaries of the Ivrea

135 mantle wedge at shallower depth <30km, not well constrained by the compilation of

136 Spada et al. (2013), we followed the 7.0 km s−1 contour in all profiles.

137 In the profiles ECORS-CROP, NFP20-West and Ticino rather thin slices of the

138 Ivrea wedge and/or the lower crust of the Adria microplate, geologically speaking the

139 Ivrea Zone (Handy & Zingg 1991), reach the surface. In these profiles the outlines of

140 the Ivrea mantle wedge are too thin to be resolved by the P-wave velocity model of

141 Diehl et al. (2009) given its limited spatial resolution. The geometry of the very thin

142 and strongly dismembered Ivrea mantle wedge and Adriatic lower crust is only

143 tentatively drawn, being only constrained by the surface exposures.

144 Another problem one faces when determining the shape of the Ivrea mantle

145 wedge is that the velocities >7.25 km s−1 expected for the Ivrea mantle rocks that are

146 part of the Adriatic microplate and the adjacent European mantle become

147 indistinguishable from velocities expected for potential remnants of the subduction

6 148 channel separating European and Ivrea (=Adriatic) mantle marking the present-day

149 plate boundary at great depth. At a depth of 30-45 km and temperatures around

150 400°C- 500°C, we expect to reach blueschist facies conditions, the exact depth

151 depending on temperature (Oberhänsli et al. 2004). Based on thermal modeling

152 (Bousquet et al. 1997) the expectation of temperatures between 450° and 500°C at a

153 depth of 30km is realistic. The P-wave velocity of blueschist rocks at 400°C and 1

154 GPa is around 7.25 km s−1 on average according to Fujimoto et al. (2010). From a

155 depth of around 50km onward, again depending on temperature, we expect eclogite

156 facies conditions associated with even higher velocities around 8.0-8.5 km s−1

157 (Worthington et al. 2013). Hence, the contours of velocities >7.25 km s-1 are expected

158 to cross over from the European to the Ivrea mantle rocks traversing the subduction

159 channel formed by high-P metamorphic rocks. Such crossover is actually seen in all

160 the profiles of Fig 2, also for velocity contours at 7.0, 6.5 and occasionally at 6.5km

161 s−1. This is because velocities expected for the subducted European lower crust

162 (around 6.5 km s−1; Ye et al. 1995) are similar to those expected for the stack of

163 nappes that is northerly adjacent to the Ivrea mantle wedge, nappes that are only

164 partially overprinted by high-P metamorphism (e.g. Henry et al. 1993). All this,

165 together with the limited resolution of the 3-D velocity model, puts severe limits to a

166 straightforward correlation of velocities with geological-tectonic entities, particularly

167 at greater depths.

168

169 1. Argentera profile (Fig. 2a)

170 The external SW part of this profile is characterized by minor shortening of

171 around 21km in Mesozoic sediments (Chaînes Subalpines) of the European foreland

172 (Lickorish & Ford 1998), associated with rather modest late-stage up-thrusting of the

7 173 Argentera external massif (Malaroda et al. 1970) during the Neogene. In this external

174 part of the profile the European Moho coincides with a strong velocity gradient

175 between 6.5 and 7.5 s-1 in the velocity model of Diehl et al. (2009).

176 The still unfolded Chaînes Subalpines were thrust by sedimentary units

177 attributed to the Subbriançonnais, together with non-metamorphic Helminthoid flysch

178 detached from the - paleogeographical domain above ophiolite-

179 bearing mélanges (Autapie and Parpaillon nappes; Merle & Brun 1984). Final

180 emplacement of these higher nappes occurred during the Late Eocene (Kerckhove

181 1969), i.e. before thrusting along the late Alpine (Early Oligocene) out-of-sequence

182 thrust called “Penninic Front” and Neogene shortening in the Chaînes Subalpines.

183 Since the Autapie and Parpaillon nappes lack metamorphic overprint they are

184 considered as units derived from a part of the Piedmont-Liguria Ocean that always

185 remained in an upper plate position. They form relics of the front of the overriding

186 oceanic upper plate (Piedmont-Liguria Upper Plate units of Fig. 2a) that started to

187 thrust the oceanic lower plate (Piedmont-Liguria Lower Plate units of Fig. 2a) from

188 the late Cretaceous onwards, reaching the front of the Alps at the end of the Eocene,

189 together with the underlying Subbriançonnais units.

190 There is a rather abrupt lateral change of velocities at a given depth across the

191 boundary between external and internal Alps. This change coincides with the

192 downward projection of the southern end of a very steeply NNE dipping late Alpine

193 out-of-sequence thrust well documented in the Pelvoux section and known as the

194 Penninic Front (Ceriani et al. 2001; Lardeaux et al. 2006; Loprieno et al. 2011; see

195 description of the Pelvoux section). The term “Penninic Front”, synonymous with the

196 term “Briançonnais Frontal Thrust” coined by Tricart (1986) who first recognized the

197 importance of this Early Oligocene out-of-sequence thrust in marking the boundary

8 198 between internal and external parts of the Western Alps, is a misnomer in that earlier

199 emplaced Penninic units are occasionally also found in front of this out-of-sequence

200 thrust (Ubaye-Embrunais of profile in Fig. 2; Préalpes Romandes of Fig. 2d; see also

201 Fig. 1). We use the term Penninic Front nevertheless because it became deeply

202 entrenched in the literature by now. This top WMW Penninic Frontal Thrust turns

203 into the sinistral Acceglio-Cuneo Line (Fig. 1) and is characterized by a sinistrally

204 transpressive component as it swings around into the WNW-ESE strike of the

205 Argentera-Cuneo line whose western end is traversed by the Argentera profile. The

206 sinistral Acceglio-Cuneo Line allows for the WNW-directed indentation of the Ivrea

207 mantle wedge along its southern boundary located near Cuneo (Trullenque 2005;

208 Ceriani et al. 2001, see their Fig. 1). Laubscher (1971) first proposed the existence of

209 a sinistral strike slip zone (his “Ligurian shear zone”) allowing for westward

210 indentation of the Ivrea mantle wedge for kinematic reasons (see also Laubscher

211 1991). Based on the interpretation of fission track data (Fügenschuh & Schmid 2003)

212 and radiometric dating of mylonites (Simon-Labric et al. 2009) the thrust marking the

213 Penninic Front (Roselend thrust of Loprieno et al. 2011) was active during the Early

214 Oligocene (34-30 Ma). By downward extrapolation the steep dip of this crustal scale

215 transpressive fault projects (Malaroda et al. 1970) into parallelism with the steeply

216 overturned 6.5 s-1 velocity contour that marks the SW edge of a high-velocity area

217 with > 6.5 s-1 located within the high-pressure Briançonnais units at a depth of

218 >10km. This high velocity area is interpreted to consist of the high-pressure

219 Briançonnais units that include the Dora Maira “massif” exposed at the surface at the

220 eastern margin of the Internal Alps (Henry et al. 1993). We are aware that not all

221 exposed parts of the Dora Maira massif constituting the high pressure Briançonnais

222 show blueschist facies or eclogitic parageneses (Bousquet et al. 2012a). However, the

9 223 high velocities at depth require that most of the volume taken up by the Briançonnais

224 at depth must have undergone high-pressure metamorphism.

225 The shallow crustal parts of the metamorphic core of the Alps in this profile are

226 characterized by large-scale back-folds that formed between 35-30 Ma (Bucher et al.

227 2004) and which affected pre-existing nappe contacts between: (1) the high-pressure

228 Briançonnais units, (2) overlying so-called Acceglio and/or Pre-Piémontais units

229 (Lemoine & Tricart 1986; Caby 1996), representing the former transition of the

230 Briançonnais micro-continent into the Piedmont-Liguria Ocean, (3) Piedmont-Liguria

231 high pressure units and, finally, (4) non-metamorphic Piedmont-Liguria units that are

232 only preserved along strike in the area of the Chenaillet near Briançon (Manatschal et

233 al. 2011). We attribute the Chenaillet ophiolites to that part of the Piedmont-Liguria

234 Unit that occupied an upper plate position in the internal Western Alps (Schwartz et

235 al. 2007) where, in our interpretation, the Austroalpine nappe system is totally

236 lacking due to a westward lateral transition of the active margin of the Late

237 Cretaceous to Paleogene Alpine subduction zone from the northern edge of the Adria

238 continent into an intra-oceanic location (Molli, 2008; Handy et al. 2010). The

239 geometry of the shallow crustal parts of the Internal Alps depicted in this profile

240 closely follows the geometries shown by the works of Henry et al. (1993), Michard et

241 al. (2004), Lardeaux et al. (2006), and Ford et al. (2006), the surface-near parts of the

242 Penninic Front exposing the Subbriançonnais and non-metamorphic Briançonnais

243 units was drawn after a series of detailed profiles given by Gidon (1972).

244 The 7.0 s-1 velocity contour that runs parallel and near the Moho depicted by

245 controlled-source seismology in the Adria microplate rises SW-ward to a depth of

246 some 15 km in the NE part of the profile beneath the HP Briançonnais. It is taken as

247 marking the outline of the top of the Ivrea mantle wedge. The inferred overall

10 248 geometry demonstrates that crust and mantle of the Adriatic microplate became

249 wedged into the Alpine edifice and underplated shallow crustal layers in the internal

250 parts of the Alps. The dip of the Periadriatic line is constrained to a depth of some 30

251 km by connecting the southern prolongation of the Periadriatic Line (Fig. 1),

252 delimiting the eastern edge of the root of the Piedmont-Liguria units near the eastern

253 margin of the Dora Maira massif at the surface, with the western front of the Ivrea

254 mantle wedge paralleling the 7.0 s-1 contour. The exact geometry shown for the

255 downward narrowing deepest parts of internal Alps, i.e. the subduction channel

256 separating the European and Adriatic (micro) plates, is basically unconstrained by

257 geophysical data and the details shown are merely conceptual. The downward

258 bending of the 7.0 and 7.5 s-1 velocity contours across the suspected subduction

259 channel is unable to depict the geometry of the subduction channel, due to the high

260 velocities expected for the rocks within the subduction channel undergoing high-

261 pressure metamorphism, velocities expected to be similar to those of the European

262 lower crust, and at greater depth, similar to those of the adjacent mantle rocks.

263 The Oligocene age of formation of the steeply dipping root of the Piedmont-

264 Liguria Ocean, associated with rock uplift associated with exhumation (Fox et al.

265 2016) of these high-pressure units along the front of the adjacent Adria microplate

266 can be inferred from the cooling history of the Dora Maira high pressure unit

267 evidencing rapid cooling from 550°C to some 260°C within an extremely short time

268 interval between 32 and 30 Ma according to the pressure-time path constrained by

269 petrological data in combination with cooling ages derived from isotopic dating and

270 fission track analysis (Gebauer et al. 1997; Rubatto & Herrmann 2001; Berger &

271 Bousquet 2008). The earlier stages of this exhumation are possibly related to

272 exhumation within the subduction channel, triggered by the buoyancy of continental

11 273 rocks at mantle depth (Chemenda et al. 1995). These time constraints indicate that

274 both the formation of the steeply hinterland-dipping Penninic Front and the wedging

275 of the Adria mantle and crust associated with backthrusting of the internal Alps over

276 the Tertiary Piedmont Basin are contemporaneous. Both appear to be associated with

277 the WNW-directed indentation of the Ivrea mantle during the earliest Oligocene (34-

278 30Ma).

279 The shallow NE crustal parts of the profile are covered by the Oligo-Miocene

280 of the Tertiary Piedmont Basin and younger Plio-Pleistocene sediments. The

281 construction of the profile is based on subsurface data acquired by AGIP (Pieri &

282 Cropi 1981, Cassano et al. 1986, Schumacher & Laubscher 1996) that only vaguely

283 constrain the geometry depicted. The 5.5 to 7.0 velocity contours, as well as the

284 Moho depth indicated by controlled-source seismology, indicate that within the Adria

285 microplate Moho, lower crust and basement-sediment interface are E-dipping.

286 In the immediate vicinity of the Periadriatic line the thickness of the Tertiary

287 Piedmont Basin along the profile is considerable. Carrapa & Garcia-Castellanos

288 (2005) showed that the subsidence of the Tertiary Piedmont Basin is controlled by

289 far-field compression and thrust loading rather than by extensional tectonics. Hence,

290 the early stages of thrust loading are probably due to the Early Oligocene backthrust

291 of the metamorphic units of the Western Alps onto the Adria microplate.

292 However, at the eastern end of the profile the Tertiary Piedmont Basin overlies

293 “Apenninic” units (Fig. 2a). Compressional tectonics continued into the Miocene

294 when the Ligurian Alps became involved in N-directed thrusting, together with

295 Apenninic units exposed in the northern Monferrato hill (Fig. 1) west of Torino

296 (Piana 2000), first during the so-called Paleo-Apenninic phase (Schumacher &

297 Laubscher 1996) sealed by a Mid-Miocene unconformity. Later, Neo-Apenninic

12 298 thrusting formed the morphologically visible thrust front north of the Torino and

299 Monferrato hills. It is this N-directed Apenninic thrusting over the Adria microplate

300 that involved parts of the former Alps referred to as “Apenninic” units in the profile

301 of Fig. 2a (see later discussion on the Alps-Apennines transition). Differential N-

302 directed thrusting of joint Apenninic and Ligurian Alpine units calls for a sinistral

303 strike slip zone that separates the joint Apennines-Ligurian Alps unit moving to the

304 north from those parts of the Piedmont-Liguria units that are immediately adjacent to

305 the Dora Maira backthrust, such as depicted near the eastern end of the section (Fig.

306 1a). The exact 3D geometry of this complicated Early Miocene to recent scenario still

307 remains largely unresolved and is known as the problem of the “Ligurian knot”

308 (Laubscher et al. 1992).

309

310 2. Pelvoux profile (Fig. 2b)

311 The western part of this profile crosses the Vercors massif (Chaînes

312 Subalpines) and the more internal Belledonne–Grandes Rousses–Oisans external

313 massifs and their Mesozoic cover (Philippe et al. 1998; Dumont et al. 2008) (Fig. 1).

314 Profile balancing by Bellahsen et al. (2012), taking into account the effects of normal

315 faulting preceding Alpine orogeny (e.g. de Graciansky et al. 1989) related to the

316 opening of Alpine Tethys, the total amount of Miocene shortening in these external

317 parts of the Alps (Dauphinois) amounts to some 28km.

318 In the external parts of the profile the P-wave velocity contours within the

319 European lithosphere again exhibit a higher velocity gradient between 6.5 and 7.5 s-1

320 near the Moho. However, the 6.0 s-1 contour rises to an unusually shallow depth of

321 <10km twice, a feature that is considered robust and points to the presence of high-

322 velocity material at shallow depth. The more external amongst the two anomalies is

13 323 found underneath the Belledonne and Grandes Rousses massifs, an area that is known

324 for the occurrence of the pre-Alpine Chamrousse ophiolitic complex (Guillot &

325 Ménot 2009), associated with ophiolites and eclogites being part of a former Variscan

326 suture zone. This suture zone evolved into a major SW-NE striking and ca. 1500 km

327 long dextral wrenching zone in Late Carboniferous times, marking the SE margin of

328 the European Variscides with Gondwana. This broad mega shear zone is known as

329 the External Massifs Shear Zone (Guillot & Ménot 2009) or East Variscan Shear

330 Zone (Padovano et al. 2012). The interpretation of this anomaly in terms of high-

331 velocity material constituting this steeply inclined mega shear zone is admittedly

332 speculative. The origin of the second more internal anomaly is unknown but possibly

333 also related to velocity inhomogeneities related to pre-Alpine structures.

334 At the surface the Penninic Front is exposed as a spectacular thrust zone behind

335 the Pelvoux massif and in the footwall of the Massif du Combeynot, dipping with ca.

336 35° to the E and associated with top-WNW out-of-sequence thrusting (Butler 1992;

337 Ceriani et al. 2001; Ceriani & Schmid 2004, Trullenque et al. 2006). The linear

338 extrapolation of this dip observed at the surface to depth is unconstrained by currently

339 available geophysical data along this profile. Fig. 2b opts for an extrapolation of the

340 Penninic Front to great depth such as to join the subduction channel at the base of the

341 Ivrea mantle wedge, consistent with what was observed along the Argentera profile

342 described above. The 6.5 and 7.0 s-1 contours that followed the lower crust of the

343 European lithosphere in the external Western Alps start to rise to shallower levels at

344 an angle of around 45° after having crossed the Penninic Front, indicating that

345 westward they enter high-pressure Briançonnais units.

346 The shallow crustal parts of the metamorphic core of the Alps are drawn

347 according to data published in Barféty et al (1995), Caby (1996), Ganne et al. (2006),

14 348 and Lardeaux et al. (2006). In the Briançonnais units sedimentation continued until

349 the Midle (internal Briançonnais) to Late Eocene (external Briançonnais) (Jaillard

350 1999) and hence predates thrusting at the Early Oligocene Penninic Front. They very

351 much resemble those found along the Argentera profile. Again, almost the entire

352 section across the inner Western Alps, except the very frontal part, is characterized by

353 late-stage backfolding of pre-existing nappe contacts. The geometry of the top of the

354 Ivrea mantle wedge is constrained by the 7.0 s-1 velocity contour and controlled

355 seismology data. Exact location and width of the subduction channel remain

356 unknown, but an inverse velocity gradient of the 7.5 s-1 contour associated with a

357 down-bending of the same contour line points to the existence of a subduction

358 channel below the Ivrea mantle wedge. A similar, although not identical geometry of

359 a geological transect located near the Pelvoux profile was proposed by Zhao et al.

360 (2015), who also postulated subduction of European continental lower crust all the

361 way below the Ivrea mantle wedge, based on their interpretation of seismic imaging

362 by P-wave receiver function analysis. In comparison, their model proposes a much

363 shallower depth for the top of the Ivrea mantle wedge (at around 10km) and a much

364 wider (about 30km) subduction channel below the Ivrea mantle wedge, features that

365 are not compatible with the 3-D velocity model of Diehl et al. (2009) and the Moho

366 depth as compiled by Spada et al. (2013).

367 The eastern parts of the profile are very similar to what was described for the

368 Argentera section. They cross the northern parts of the Dora Maira massif and exhibit

369 the same overturned root of the ophiolitic series of the Piedmont Lower Plate unit,

370 juxtaposed with the front of the Adria microplate including the Ivrea mantle wedge

371 along a backthrust onto the Tertiary Piedmont Basin, formed in the earliest

372 Oligocene, i.e. contemporaneous with the formation of the Penninic Front. At its

15 373 easternmost end the profile also reaches the Apenninic units thrust northward and

374 perpendicular to the trace of the profile towards the Torino Hills in the Late Miocene.

375

376 3. ECORS-CROP profile (Fig. 2c)

377 This transect follows the trace of the geophysical campaign launched in 1985

378 by the French-Italian Étude Continentale et Océanique par Reflexion et Refraction

379 Sismique - Progetto Strategico Crosta Profonda (ECORS-CROP) geophysical

380 campaign in the Western Alps involving high-resolution deep seismic sounding

381 Roure et al. (1996). The geological interpretation of this profile is a modified version

382 of the profile previously published by Schmid & Kissling (2000).

383 The external Alps exhibit 27km shortening, an amount very similar to that

384 along the Pelvoux profile, calculated on the basis an area balance within an originally

385 17km thick upper crust, assuming decoupling at the interface with the lower

386 European crust. This shortening is of Miocene age and postdates thrusting at the

387 Penninic Front (Fügenschuh & Schmid 2003). Note that Butler (1983) postulated

388 very substantially higher amounts of shortening (>50km) since his line balance

389 assumes detachment along a floor thrust located only 1km below base Triassic. The

390 6.5 km s−1 velocity contour beneath the external massifs, in the case of this profile the

391 Belledonne Massif, indicates the presence of a steep tabular high-velocity body in the

392 lower part of the upper crust that we again (see description Pelvoux profile)

393 tentatively attribute to the East Variscan Shear Zone of Late Variscan age (Padovano

394 et al. 2012). The velocities within the rest of the European upper crust are, apart from

395 this anomaly, < 6.5 km s−1 all the way back to the very internal parts of the Alps

396 below the Gran Paradiso massif, built up the high pressure Briançonnais unit. This

397 new finding about the P-wave velocity in the upper crust of the external part of the

16 398 Western Alps (Diehl et al. 2009) is incompatible with the presence of an

399 allochthonous wedge of lower crust underneath the external Western Alps as

400 postulated by Schmid & Kissling (2000) on the basis of gravity arguments (Bayer et

401 al. 1989) and W-dipping seismic reflectors (Thouvenot et al. 1996) at the front of that

402 putative wedge. Below the Gran Paradiso massif the 6.5 km s−1 contour steeply raises

403 to a depth of only some 10 km, across the Penninic Front and into the area of the

404 high-pressure part of the Briançonnais units, indicating a high-pressure metamorphic

405 overprint also at great depth.

406 At the surface the profile crosses the type locality of the top-WNW Penninic

407 Front (“Roselend Thrust” of Loprieno et al. 2011). This thrust, together with a 10 km

408 wide belt of units attributed to the Valaisan paleogeographic domain, containing,

409 amongst other lithologies, black schists interlayered with mafic sills and low-T

410 eclogites (Bousquet et al. 2002, 2012a; Loprieno et al. 2011) forms a planar band of

411 strong reflectors that can be traced down to a depth of some 15 km along the

412 ECORS-CROP reflection seismic profile (Thouvenot et al. 1996). Following Schmid

413 & Kissling (2000), and in analogy with the Pelvoux profile discussed above, we

414 kinematically link the Penninic Front with the base of the Adria mantle wedge.

415 The near-surface geometry of the upper crustal structures within the

416 Briançonnais units, namely the non-metamorphic Zone Houllière and the High-

417 Pressure Briançonnais units (Ruitor, Zona Interna and Gran Paradiso nappes), as well

418 as the geometries of the Piedmont-Liguria Units (Zermatt-Saas and Combin Zones)

419 are drawn following Schmid & Kissling (2000) and Bucher et al. (2003, 2004). They

420 are characterized by large-scale post-nappe backthrusts and backfolds formed

421 between 35 and 31 Ma, i.e. contemporaneous with thrusting along the Penninic Front.

422 High-pressure metamorphism related to subduction was dated at 48-47 Ma in the

17 423 Briançonnais units, overprinted by greenschist facies conditions related to nappe

424 stacking at around 43-39 Ma (Villa et al. 2014). Pressures associated with the high-

425 pressure event increase eastward and reach eclogite facies conditions in parts of the

426 Gran Paradiso internal massif, characterized by P-wave velocities around 6.0 to

427 almost 7.0 km s−1 at greater depth according to the velocity model of Diehl et al.

428 (2009).

429 Compared to the Briançonnais units high-pressure metamorphism within the

430 more internal Zermatt-Saas-Fee unit (Piedmont-Liguria Lower Plate of Fig. 2c) and

431 within small continental slices at its top indicates an earlier onset of a long and

432 protracted high P overprint (62-42 Ma) according to age dating (e.g. Dal Piaz et al.

433 2001; Skora et al. 2015; Weber et al. 2015; Fassmer et al. 2016). The most internal

434 parts of the ECORS-CROP profile cross the Sesia Zone, which together with the

435 Dent Blanche Klippe, was classically considered as an “Austroalpine” unit. This is

436 because the protoliths forming the Sesia Zone are undoubtedly derived from the

437 Adria microplate in a paleogeographical sense (Compagnoni et al. 1977). However,

438 in a plate tectonic context, the Sesia Zone that reached eclogite facies conditions with

439 up to 2 GPa pressures in the period 85-60 Ma (Regis et al. 2014) and consisting of

440 thin crustal sheets (Giuntoli & Engi 2016) is interpreted as having been accreted as a

441 part of the former Lower Plate to the Upper Plate Units (i.e., the Canavese, Ivrea

442 Zone and Southern Alps) during Late Cretaceous to early Cenozoic orogeny in the

443 Western Alps, associated with the subduction of the Piedmont-Liguria ocean (Babist

444 et al. 2006, Handy et al. 2010). The situation is different in the Austroalpine nappes

445 proper where high-pressure metamorphism is significantly older and related to

446 Cretaceous intra-continental subduction that evolved within the future Austroalpine

447 units that formed the upper plate during Cenozoic orogeny (Schmid et al. 2004). In

18 448 summary, the Sesia Zone is regarded as a lower plate unit following Froitzheim et al.

449 (1996; their Fig. 2) and Pleuger et al. (2007; their Fig. 14) who proposed that the

450 Sesia Zone represents a continental fragment floating within the Piedmont-Liguria

451 oceanic domain that rifted and drifted away from Adria during the opening of Alpine

452 Tethys in the Mesozoic. Two important geological observations support this finding.

453 Firstly, the southern edge of the mostly continental Sesia high-pressure unit,

454 immediately adjacent to the Insubric Line, contains a high-pressure slice containing

455 omphacite-bearing mantle peridotites, which are testimony of a former oceanic tract

456 SE of the Sesia Zone that underwent high-pressure metamorphism (Pognante 1989,

457 Barnes et al. 2014). Secondly, the non-metamorphic Canavese Zone, immediately

458 south of the Insubric Line and north of the Ivrea Zone contains, amongst other

459 lithologies, outcrops of exhumed mantle rocks overlain by Jurassic age radiolarites

460 (Beltrando et al. 2015a), which points to the existence of an oceanic tract between the

461 largely continental Sesia high-pressure unit and the Ivrea Zone. It is only the

462 Canavese Zone, together with the Ivrea Zone and the Southern Alps that form the

463 upper plate of the Alpine orogen in the internal parts of this transect. As demonstrated

464 in the pioneering work by Elter et al. (1966) the Canavese Zone has to be considered

465 the root of the highest nappe exposed in the Préalpes Romandes, the ophiolite-bearing

466 nappe des Gets (see profile of Fig. 2d; unit labeled as Piedmont-Liguria Upper Plate).

467 Both these units represent partly oceanic series that once formed the presently largely

468 eroded upper plate of the Alpine edifice.

469 In contrast to the more southerly-located profiles, very thin slices of the Ivrea

470 mantle wedge, accompanied by lower crustal lithologies of the Ivrea Zone, reach the

471 surface of the earth, slices that are too thin to have a notable effect on the velocity

472 contours. The main part of the Ivrea mantle wedge is confined to a depth >10km as

19 473 indicated by the 7.0 km s−1 contour line. Along this profile the 7.25 km s−1 contour

474 line reveals an inverted velocity gradient, allowing for localizing the base of the Ivrea

475 mantle wedge above the subduction channel.

476 The Adria lower and upper crustal units are affected by backthrusts

477 representing the easternmost extension of the Southern Alps. Details are drawn

478 according to the interpretation given by Roure et al. (1990). This thrusting, detected

479 in the subsurface of the Po plain in the course of the ECORS-CROP reflection

480 seismic campaign, is of pre-Burdigalian (pre-20.5 Ma) age and affects part of the

481 Gonfolite Lombarda sediments, providing an age range for thrusting of Late

482 Oligocene to Early Miocene. Note that this profile does not reach the front of the N-

483 directed Apenninic thrusts, which stays south of the SE end of this section.

484

485 4. NFP-20 West profile (Fig. 2d)

486 This transect follows the trace of a Swiss geophysical campaign (NFP 20 “Deep

487 Structure of Switzerland” whose results were published in Pfiffner et al. (1997).

488 Earlier geological interpretation of the deep structure revealed by controlled source

489 reflection and refraction seismology such as those by Escher et al. (1997) and Schmid

490 et al. (2004) have been revised in the light of the 3D velocity model of Diehl et al.

491 (2009).

492 Along this profile the effects of the foreland propagation of Alpine shortening

493 induced the decollement of the entire Molasse basin along Triassic evaporite horizons,

494 which led to the formation of the Jura Fold-and-Thrust Belt. The Alpine thrust front

495 jumped by some 100km to the north after about 12 Ma ago; this decollement is

496 associated with some 26km of shortening and roots in the external massifs (Aiguille

497 Rouge and M. Blanc massifs) of the Alps (Burkhard & Sommaruga 1998). However,

20 498 the total amount of shortening affecting the upper European crust, which along this

499 transect also includes the lowermost Penninic nappes behind the Penninic Front,

500 amounts to some 50km according to an area balance along this profile, again

501 assuming decoupling at the interface between upper and lower crust. This is almost

502 twice as much as deduced for the ECORS-CROP profile but, in this case, also

503 includes all the shortening that occurred since 35 Ma ago. The deep structure of the

504 External Alps along this profile is characterized very marked undulations of the

505 velocity contours between 7.5 and 6.0 km s−1. Some of these undulations may again

506 have been caused by velocity gradients related to pre-Alpine structures. However,

507 one of these undulations, an antiformal structure located underneath the external

508 massifs, appears to also affect the Moho; the closely spaced 7.5, 7.0 and 6.5 km s−1

509 velocity contours all consistently indicate an antiformal structure affecting the Moho

510 and lower crust. A local up doming of the Moho by about 5km correlates well with

511 earlier observations in reflectivity pattern along the NFP20West near-vertical

512 reflection profile (Pfiffner et al. 1997) that at the time were not further interpreted. As

513 this Moho geometry has now been consistently observed by two independent seismic

514 data sets, we assume this antiformal Moho-structure to be real and possibly

515 representing a pre-Alpine feature.

516 For constructing the near-surface structures in the most external parts of the

517 profile we mainly used Sommaruga (1997) for the Jura Mountains and Sommaruga et

518 al. (2012) for the Molasse basin. In the case of the External Massifs, the Helvetic

519 nappes and the Préalpes Romandes we based our interpretation of data provided by

520 Burkhard (1988), Marchant (1993), Steck et al. (1997), Escher et al. (1997) and

521 Burkhard & Sommaruga (1998).

522 The Penninic Front along this profile coincides with the south dipping Rhone-

21 523 Simplon line, a wide dextral shear zone exhibiting initially ductile deformation

524 starting to be active at around 35 Ma (Steck & Hunziker 1994; Schmid & Kissling

525 2000), which graded into brittle deformation from the Miocene onwards (Cardello &

526 Mancktelow 2015). The Simplon normal fault located immediately east of this

527 transect (Mancktelow 1992; Grasemann & Mancktelow 1993) represents a tensile

528 bridge that accommodates orogen-parallel extension during the backthrusting over

529 the Ivrea mantle wedge towards ESE and associated top WNW thrusting at the

530 Penninic Front of the Western Alps (Keller et al. 2006). The downward projection of

531 the Rhone-Simplon line to some 15km follows that proposed by Escher et al. (1997).

532 At a >15km depth any interpretation of the geometry of this fault and adjacent units

533 is only loosely constrained by migrated reflection seismic data (Marchant 1993) and

534 strike-parallel lateral projection of near surface structures. Our interpretation differs

535 from that of Escher et al. (1997) at a depth > 35km because we conceive the Rhone-

536 Simplon line as delimiting the northern border that acted as a lateral ramp for that

537 part of the Alpine upper crust that moved towards WNW together with the Ivrea

538 mantle wedge during its indentation.

539 The internal units of the Alps SE of the Rhone-Simplon line are characterized

540 by spectacular backfolds within the Briançonnais units whose geometry down to a

541 depth of about 20km is constrained by reflection seismology (Marchant 1993; Steck

542 et al. 1997) and orogen-parallel axial projections of surface structures (Escher et al.

543 1997; Keller & Schmid 2001; Kramer 2002). At a depth >20km the P-wave velocity

544 contours of Diehl et al. (2009) served as a guideline for constraining the geometry.

545 The 6.0 km s−1contour indicates a large-scale velocity inversion that follows the SE

546 limb of the Monte Rosa backfold down to a depth of some 30km. There the 6.0 km

547 s−1 contour swings back into a region characterized by a normal velocity gradient and

22 548 runs parallel to the 6.5 and 7.0 km s−1 velocity contours that are typical for the

549 European lower crust. As a consequence we limit the area characterized by such

550 backfolding to the uppermost 35km of the crust.

551 Farther to the SE the 6.0 and 6.5 km s−1 contours rise to a shallow depth of only

552 some 10 km and 15km, respectively. This rise occurs within the Sesia Unit that is

553 known to exhibit eclogite facies conditions in its most internal parts only (Bousquet

554 et al. 2012a). Because this eclogite facies part of the Sesia Zone is immediately

555 adjacent to the Ivrea mantle rocks, the limit between the Adria mantle wedge and the

556 under-plated Sesia zone is ill constrained. The near-surface structures of the Southern

557 Alps are drawn after Cassano et al. (1986).

558

559 5. Ticino profile (Fig. 2e)

560 The Ticino profile runs parallel to the reflection seismic lines C1, C2 and C3 of

561 NFP 20 project (Pfiffner & Heitzmann 1997); the southernmost parts of this profile

562 are located near lines S4 and S5 of the same project (Schumacher 1997). At depth,

563 the profile crosses the NE-most part of the Ivrea mantle wedge, very near its

564 termination (see Fig. 3). Note also that this profile crosses a pronounced along-strike

565 axial culmination (the core of the Lepontine dome; Berger et al. 2005; Berger &

566 Mercolli 2006; Steck et al. 2013) at a location where the distance between external

567 massifs and Periadriatic line is near a minimum (Fig. 1), going hand in hand with a

568 maximum amount of shortening and exhumation. Since this transect is located east of

569 the Rhone-Simplon line, dextral transpression allowing for top WNW thrusting along

570 the Penninic Front is taken up by the Periadriatic line only.

571 According to an area balance, again assuming decoupling at the interface

572 between upper and lower crust, the total amount of shortening affecting the upper

23 573 European crust (including the Penninic nappes below the Valaisan and Briançonnais

574 units) since around 35 Ma increases by as much as some 115 km when compared to

575 the 50km determined along the NFP20-West traverse. Rosenberg & Kissling (2013)

576 measured some 71 km shortening along a profile located between our NFP20-West

577 and Ticino profiles (their Simplon profile). This indicates a steady increase in the

578 amount of post-collisional (post 35 Ma) shortening within the accreted European

579 lower plate towards northeast into the Ticino transect.

580 In the external part of the profile the 6.0 km s−1 contour rises to shallow depth

581 underneath the frontal part of the external massifs, hinting at the existence of a pre-

582 Alpine structure (External Massifs Shear Zone of Guillot & Ménot 2009?) also along

583 this profile. The near-surface features in the foreland of the Molasse Basin are drawn

584 according to Sommaruga et al. (2012), the geometry of the external Aar and Gastern

585 massifs is modified after Müller (1938) and Pfiffner (2011), the structures in the

586 Helvetic nappes, overlying Penninic nappes and the Gotthard nappe (formerly

587 considered a “massif”) are drawn after Menkveld (1995) and Pfiffner et al. (2010).

588 The internal parts of the transect, depicting the nappe pile within and above the

589 Lepontine dome, are characterized by a large-scale backfold associated with the

590 Northern Steep Belt located within the southern border of the Subpenninic Gotthard

591 nappe (Schmid et al. 2004; Berger et al. 2005), separated by a flat-lying nappe pile

592 from the Southern Steep Belt as defined by Milnes (1974), who was first in realizing

593 the importance of post-nappe recumbent mega-folds affecting a previously existing

594 nappe stack consisting of Briançonnais, Penninic and Subpenninic units. The

595 formation of the Southern Steep Belt, the southern limb of a second more southerly

596 located backfold, is closely associated with backthrusting of the Alpine nappe stack

597 over the Ivrea mantle wedge and overlying crust in the area of the Southern Alps and

24 598 along the Periadriatic Line that initiated between 35 and 30 Ma in a dextrally

599 transpressive scenario (Schmid et al.1987 & 1989; Steck & Hunziker 1994). Note

600 that Miocene thrusting in the Southern Alps is younger, i.e. post-20Ma (Schönborn

601 1992). Based on the offset of the Alpine grade of metamorphism across the

602 Periadriatic Line, the vertical component of dextral transpression amounts to some

603 20km along the trace of this profile. The amount of the simultaneously occurring

604 horizontal displacement across the transpressive Periadriatic Line remains unknown

605 due to the lack of markers; estimates vary between 100km (Schmid & Kissling 2000)

606 and 240km (Handy et al. 2010). Details concerning the shallower parts of this central

607 portion of the profile are drawn based on structural information by Milnes (1974),

608 Huber et al. (1989), Leu (1987), Berger et al. (2005), Steck et al. (2013) and own

609 observations. Reflection seismic data along line C1 and C2 (Pfiffner & Heitzmann

610 1997) were used for placing nappe contacts at the base of the Antigorio nappe and the

611 Gotthard nappe, respectively, and for estimating the depth of the top of the European

612 lower crust.

613 At a depth >20km the NW edge of the Ivrea mantle wedge is defined by the 7.0

614 km s−1 contour. At a depth of <20km the Periadriatic fault, whose dip is evaluated on

615 the basis of surface observations and a series of reflection seismic profiles (Schmid &

616 Kissling 2000; their Fig. 5) delimits the NW edge of the Ivrea mantle wedge,

617 including the Ivrea Zone that represents the adjacent Adriatic lower crust. The Ivrea

618 Zone, including some ultramafic bodies at its northern rim, has been rotated along a

619 near-horizontal axis into a vertical orientation due to Alpine deformation under low-

620 grade metamorphic conditions (Handy 1987; Schmid et al. 1987; Zingg et al. 1990).

621 The Ivrea Zone is well known for representing a map-view section across lower

622 continental crust (Fountain 1976). It is likely that the topmost parts of the Ivrea Zone

25 623 and mantle became highly tectonized under semi-brittle conditions, as is

624 schematically shown in Fig. 2e, allowing for this rotation.

625 The 6.5 km s−1 contour rises to a shallow depth of around 10 km above the

626 Ivrea mantle wedge also in this section, crossing from the European lower crust

627 across the subduction channel over the presumed top of the Ivrea mantle wedge. The

628 northern limit of the Ivrea mantle wedge and the adjacent Ivrea Zone are constrained

629 by downward extrapolation of the Periadriatic Line to a depth of 20km; at greater

630 depth the Periadriatic line delimits the northern limit of the Ivrea mantle wedge.

631 The profile across the upper crust of the Southern Alps is partly based on the

632 geological interpretation of reflection seismic data gathered during the NFP-20 and

633 CROP campaigns (Schumacher 1997; Schumacher et al. 1997; lines C3-south, S4, S5,

634 S6, S7& CROP 88-1) and partly on the subsurface information provided by Cassano

635 et al. (1986). Furthermore, we based our interpretation also on the work of Bertotti

636 (1991) and Schönborn (1992). The total amount of shortening within the Southalpine

637 retro-belt amounts to 74km according to section balancing in the area crossed by this

638 transect (Schumacher et al. 1997). A large, although not well constrained part

639 (between 24km and 43km along a section east of Lake Como according to Schönborn

640 1992) of this shortening pre-dates the Late Eocene. This leaves between 31 and 50

641 km of shortening that occurred in the Middle to Late Miocene. The lower bound of

642 this shortening estimate is more realistic in view of the overall geometry of the

643 transect depicted in Fig. 2e. The reason for the upward excursion of the 6.5 km s−1

644 contour inside the Adriatic upper crust remains unknown; possibly it is related to

645 heterogeneities formed during Variscan tectonics.

646

647 6. Horizontal section (Fig. 3)

26 648

649 To discuss the overall shape of the Ivrea mantle wedge and lateral variations in

650 its geometry the five profiles described above are complemented with a horizontal

651 section across the 3-D P-wave tomographic model of Diehl et al. (2009) at a depth of

652 20km (Figure 3). This horizontal section shows (1) a curved area of positive P-wave

653 anomalies >10% nicely mimicking the well-known positive Bouguer anomaly

654 associated with the Ivrea mantle wedge (Kissling 1984; Serva 2005), (2) an extension

655 of the Ivrea mantle wedge across the Periadriatic Fault visible north of Locarno (in

656 reality the wedge underlies the N-dipping Periadriatic Fault, a backthrust in its

657 hangingwall see Fig. 2e), and, (3) a possible dextral offset of the central axis of the

658 Ivrea anomaly between profiles ECORS-CROP and NFP-20 West that leaves no trace

659 at the surface of the earth when consulting the geological maps of the area.

660 Interestingly, the central axis of the high-velocity area is located east of the

661 Periadriatic Fault in the north and gradually changes over to lying west of the

662 Periadriatic line in the south. This indicates increasing amounts of underthrusting of

663 the mantle wedge beneath the Internal Western Alps towards south, culminating in

664 what is seen in the Argentera section (Fig. 2a): a Periadriatic Fault that acted as a

665 huge backthrust and severe modification of the suture between Europe and Adria

666 (lower plate Piedmont-Liguria ophiolites) in the form of a huge backfold with the

667 high-pressure Briançonnais Units in its core (Fig. 2a).

668

669 C. Multi-stage evolution of the arc of the Western Alps

670 The following discussion on the evolution of the arc of the Western Alps

671 integrates the findings described above with a review of existing knowledge on the

672 multi-stage evolution of the Western Alps concerning kinematic evolution and time

27 673 constraints derived from field data. In view of the considerable difficulties in exactly

674 dating metamorphic events using isotopic ages, particularly in the case of HP

675 metamorphism (see review by Berger & Bousquet (2008), we also heavily rely on

676 biostratigraphical constraints. We also present tectonic reconstructions using Gplates

677 reconstruction software (Boyden et al. 2011) that have already been applied to parts

678 of the Western Mediterranean and eastern realm adjacent of the Alps-Apennines

679 orogenic system (van Hinsbergen & Schmid 2012; van Hinsbergen et al 2014a;

680 Advokaat et al. 2014).

681

682 1. Exhumation of Ivrea Zone and underlying lithospheric mantle in the

683 Mesozoic

684

685 The Ivrea Zone is well known for exposing a map-view section across lower

686 continental crust of the Southern Alps that formerly was part of the Adriatic

687 microcontinent (e.g. Fountain 1976, Handy 1987, Handy & Zingg 1991). It includes a

688 series of small peridotite lenses at its northern rim, due to later tilting in the context of

689 Alpine orogeny. Restoration of the pre-Alpine setting via backrotation of Ivrea Zone

690 and adjacent parts of the Southern Alps adjacent to the Periadriatic Line reveals that

691 the Ivrea Zone was part of the distal passive continental margin that formed in

692 connection with Tethyan rifting culminating at around 180-170 Ma ago (e.g. Bertotti

693 et al. 1993). The Ivrea Zone and underlying Ivrea mantle wedge were located near the

694 western margin of the distal continental margin of the Adria microplate that had a

695 strike (Weissert & Bernoulli 1985) almost parallel to the present-day NNE-SSW

696 strike of the Ivrea mantle wedge (Fig. 3). Based on a combination of P-T estimates of

697 pre-Alpine parageneses and radiometric dating it was inferred that crustal thickness

28 698 was locally reduced to 10km or less in the area of the Ivrea Zone and adjacent upper

699 crustal units (Handy et al. 1999). Hence the present-day slope of the Moho, rising

700 from some 40km to about 20km revealed by the Moho-map of Spada et al. (2013),

701 and, rising to some 15km in Figs. 2c&d, is interpreted as inherited from passive

702 margin formation in Mesozoic times. The Canavese area, coinciding with the ocean-

703 continent boundary, from which the Sesia extensional allochthon separated during

704 rifting (Froitzheim et al. 1996; Beltrando et al. 2015a,b), is only some 50km away

705 from the Ivrea Zone.

706 The shallow depth of the top of the Ivrea mantle wedge inherited from

707 Mesozoic rifting has important consequences regarding rheology and strength of the

708 Ivrea mantle wedge during Alpine orogeny. The brittle to crystal plastic transition of

709 peridotite is estimated to be at around 600° according to most workers (e.g.

710 Druiventak et al. 2011). Temperatures in the depth range of the Ivrea mantle wedge

711 (15-40km) are between 400° and 650°C in the central area of the Alpine orogen

712 according to the steady state kinematic model of Bousquet et al. (1997). Hence, most

713 of the Ivrea mantle wedge is expected to be controlled by frictional strength around

714 1000 MPa at 15 km depth in a compressional regime under dry conditions, reaching

715 excessive values at greater depth (Byerlee 1978). At its front the Ivrea mantle wedge

716 is surrounded by upper crustal nappes (Fig. 2) whose rheology is likely to be

717 controlled by viscous creep of large quartz-bearing volumes of rock. Extrapolation of

718 flow stress data to geological strain rates, combined with paleopiezometry using

719 recrystallized grain size, predict flow stresses between of 10-100 MPa within the

720 same 400° and 650°C temperature range (Stipp et al. 2002). This indicates that the

721 strength of the Ivrea mantle wedge exceeds that of quartz-bearing units at its front by

722 at least one order of magnitude. This suggests that indentation of the Ivrea mantle

29 723 wedge into the Alpine nappe stack probably played an important role in the formation

724 of the arc of the Western Alps, as will be further discussed below.

725

726 2. Top-NNW thrusting during sinistral transpression (Stage 1, pre-35 Ma)

727

728 The orientation of stretching lineations in combination with associated sense of

729 shear criteria have been applied to deduce tectonic transport directions in highly

730 sheared rocks from the Alps (e.g. Malavieille et al. 1984; Platt et al. 1989). Along the

731 arc of the Western Alps the orientation of the lineations is dominantly transverse to

732 the strike of the orogen. The inferred tectonic transport smoothly changes from top

733 NW in the area of Monte Rosa to top WSW in the area of the Dora Maira area

734 (Malavieille et al. 1984). This apparently simple arrangement, if taken to indicate

735 sense of movement, leads to obvious difficulties in geometrical and plate kinematic

736 terms (Platt et al. 1989). The major difficulties in interpreting kinematic data based

737 on stretching lineation analysis is that (1) stretching lineations may indicate the

738 orientations of finite strain (e.g. Ramsay 1980) rather than shearing during a

739 particular stage of a complex orogenic evolution and (2) that orientations of older

740 lineations may have been reoriented by younger deformations (e.g. Walcott and

741 White, 1998; Keller & Schmid 2001; van Hinsbergen and Schmid, 2012). Hence,

742 only careful local studies addressing the polyphase structural evolution of particular

743 areas can reveal a more complex kinematic evolution. The results of such studies are

744 compiled in Fig. 4.

745 A first generation of lineations indicating top north transport is best

746 documented in areas of the Western Alps located behind, i.e. east, of the Penninic

747 Front (areas 1, 2 and 3 in Fig. 4 and references therein). In these areas the orientations

30 748 of this first generation of inferred transport directions is related to pre-35Ma nappe

749 stacking. According to detailed analyses of structural superposition this first

750 generation of lineations pre-dates thrusting along the Penninic Front initiating at

751 around 35 Ma and is hence is of Eocene or even earlier age (e.g. Bucher et al. 2003;

752 Ceriani & Schmid 2004; Loprieno et al. 2011). However the orientations of these first

753 generation lineations scatter considerably between top-NW and top-NE due to later

754 re-orientation. In the more internal parts of the internal Western Alps the preservation

755 potential of the orientation of such early stretching lineations is low. In most places

756 they are not discernable any more because of intense shearing during subsequent

757 deformation during stage 2 (see below). In the more internal units of the Aosta

758 Valley, for example, they are seen to scatter too much for inferring transport

759 directions of this earliest phase (Bucher et al 2004). Only in some parts of the

760 eclogitic lower plate Piedmont-Liguria units is a former top NNW eclogite facies

761 lineation preserved (area 5 in Fig. 4; Philippot 1988, 1990). Interestingly, also parts

762 of the External Western Alps (Pelvoux massif and its cover) are known to have been

763 affected by stage 1 to N to NW-directed deformation of pre-Priabonian age (Dumont

764 et al. 2011).

765 Large-scale structural considerations (Schmid & Kissling 2000; Handy et al.

766 2010) and plate kinematic constraints (Fig. 5) indicate that thrusting and plate

767 convergence were top N to NNW before 35 Ma. This implies a scenario of sinistral

768 transpression in the Western Alps during the early stages of orogeny when the upper

769 plate tectonic units below which the Alpine nappe stack formed (Adria lithosphere;

770 oceanic in case of the Western Alps, continental in case of the Eastern Alps) moved

771 towards N to NNW, presumably past that part of the European plate presently located

772 in SE France, as first proposed by Ricou & Siddans (1986). Pre-35Ma plate

31 773 convergence was associated with nappe stacking and early stages of exhumation of

774 parts of nappes buried to high-P (e.g. Schmid et al. 1996, their Fig. 8). A point

775 located at the northern tip of Adria (Monti Lessini) moved some 260km to the NNW

776 in respect to stable Europe between 55 and 35 Ma go (1.3 cm/y) according to the

777 reconstruction shown in Fig. 5d-f, which assumes that Adria was part of Africa at this

778 time, and used the Europe-Africa plate circuit of Seton et al. (2012). This is more

779 than the 195km (0.98 cm/y) of shortening calculated by Schmid & Kissling (2000)

780 for the same time interval but considerably less than the estimate of Handy et al.

781 (2010) who inferred 465km shortening for the interval 35-67Ma (1.45cm/y), both

782 these estimates being solely based on interpretations of the geological structure and

783 metamorphic history of the Alps. According to Handy et al. (2010) top-N thrusting

784 started as early as some 84Ma ago, consistent with the plate circuit estimates for the

785 onset of convergence based on Seton et al. (2012), and first led to high-pressure

786 metamorphism in the Sesia Zone (Manzotti et al. 2014), predating high-pressure

787 metamorphism within the lower plate units derived from the Piedmont-Liguria Ocean

788 that started at around 60Ma ago (Berger & Bousquet 2008).

789 In summary, peak pressures within the Western Alps were reached in a scenario

790 of sinistral transpression along an original approximately N-S striking segment of the

791 western Alps (Fig. 5d&e) during this first stage. Most of the exhumation of the

792 eclogites to moderate crustal levels very probably also took place during this first

793 stage of the evolution of the arc of the Western Alps.

794

795 3. Top-WNW thrusting and WNW-directed indentation of the Ivrea mantle slice

796 (Stage 2, 35-25 Ma)

797

32 798 A second population of stretching lineations is associated with deformation that

799 post-dates stage 1 top-NNW thrusting according to structural overprinting criteria.

800 The top-WNW lineations formed during this stage can reliably be attributed directly

801 to thrusting along the Penninic Front in the case of locations 1,3,4 and 7 in Fig. 4

802 (references cited therein). There the inferred thrusting directions vary between

803 azimuth 260° and 345° with a mean at around azimuth 300°. The measurements at

804 locations 2 and 5 were taken in more internal positions with respect to the Penninic

805 Front while those at location 6 are from the footwall of the Penninic Front (lineations

806 in the “schistes à bloc formation” immediately below the base of the Embrunais-

807 Ubaye nappe stack; Trullenque 2005). Given the similar spread in orientations

808 shearing during stage 2 top-WNW thrusting is likely also for locations 2, 5 and 6.

809 Several independent sets of data provide dating constraints. 40Ar-39Ar dating of

810 phengites that formed syn-kinematically with shearing of basement rocks of the

811 Pelvoux massif in shear zones along the Penninic Front or in its immediate footwall

812 vary between 35 and 30 Ma with one younger age at 27Ma (Simon-Labric et al.

813 2009; Dumont et al. 2012). This dating confirms earlier evidence for the onset of top

814 WNW thrusting at around 32 Ma based on fission track dating (Fügenschuh &

815 Schmid 2003). Since the Penninic Front is associated with dextral shearing along the

816 Rhone-Simplon line (Fig. 1) and dextral shearing in the Insubric mylonites (part of

817 PL of Fig. 1 around Locarno) the dating of such dextral shearing provides further

818 timing constraints. The “ductile Simplon shear zone” of Steck et al. (2015), predating

819 late stages of normal faulting at the Simplon normal fault (Mancktelow 1992), was

820 active between 34 and 25 Ma ago as inferred from the ages of the mantle-derived

821 andesite and tonalite (32–29 Ma) and synkinematic crustal aplite and pegmatite

822 intrusions in the southern steep belt between Domodossola and Locarno (Steck et al.

33 823 2013). Dextral shearing in the mylonite belt of the Insubric line, a segment of the

824 Periadriatic Line (PL in Fig.1) initiated at around 32 Ma ago based on radiometric

825 age data from the eastern margin of the Bergell pluton (32-30 Ma; von Blanckenburg,

826 1992). The deep-seated Bergell intrusion including its subsequent fast exhumation is

827 closely related to a combination of backthrusting and dextral strike-slip movements

828 accommodated by the greenschist mylonites of the Insubric line (Schmid et al., 1989;

829 Berger et al., 1996). The top-WNW senses of shear in the “schistes à bloc” formation

830 (Kerckhove 1969) at location 6 of Fig. 4, affected by top-WNW shearing during a

831 first phase of deformation (Trullenque 2005), provides a stratigraphic age constraint;

832 this chaotic mélange overlies the Priabonian-age flysch de Champsaur (Bürgisser &

833 Ford 1998), and hence, its age of formation is considered to be latest Priabonian to

834 earliest Oligocene (around 34 Ma). A similar age constraint is provided for the

835 sinistral senses of shear observed at the western termination of the Argentera-Cuneo

836 Line at location 7 of Fig. 4 that is kinematically connected with the Penninic Front

837 (Trullenque 2005); these mylonites are found in immediate tectonic contact with the

838 grès d’Annot of Late Eocene to Early Oligocene age that terminate the evolution of

839 the Eocene Northern Alpine foreland basin in the southern Subalpine chains (Ford &

840 Lickorish 2004).

841 The end of stage 2 shearing is not equally well constrained but certainly pre-

842 dates the activity of the left-lateral Giudicarie transpressive belt starting to offset the

843 Periadriatic line at around 23-21 Ma ago (Scharf et al. 2013). An end of stage 2 at

844 around 25 Ma ago is compatible with the onset of the Paleo-Apenninic phase

845 (Schumacher & Laubscher 1996) at about this time. Stage 2 certainly pre-dates the

846 onset of rapid counterclockwise rotation of the Corsica-Sardinia block at 21 Ma ago

847 (Gattacecca et al. 2007) that we link to stage 3 oroclinal bending discussed in the next

34 848 chapter.

849 At the large scale thrusting at the Penninic Front is interpreted to be

850 kinematically connected to dextral strike slip along the Rhone Simplon and

851 Periadriatic Lines on the one hand, and to sinistral strike slip along the Argentera-

852 Cuneo Line on the other hand. The orogen-perpendicular lineations in the Western

853 Alps compiled by Malavieille et al. (1984) are probably related to stage 2 top-WNW

854 shearing and associated backfolding that is pervasive within entire volume of rocks

855 between Penninic Front and the Periadriatic Line between Ivrea and Cuneo; this

856 shearing formed new lineations and/or reoriented older ones (Fig. 1). It is proposed

857 that this pervasive shearing, linked to WNW-directed indentation of the Ivrea mantle

858 wedge, first proposed by Laubscher (1971), also took place during stage 2 (35 and 25

859 Ma). The amount of top WNW displacement along the Penninic Front is not well

860 constrained. Schmid & Kissling (2000) estimated WNW-directed displacement of the

861 Ivrea Zone to amount to around 100km, (Handy et al. 2010) proposed some 200km.

862 The reconstruction in Fig. 5c yields 93km and 150km for the SW and NE ends of the

863 Ivrea mantle wedge, respectively.

864 The reconstruction shown in Fig. 5c depicts displacement vectors that diverge

865 between Western Alps and Eastern Alps and hence are bound to induce orogen-

866 parallel extension in the Alps that is known to also have started at around 35 Ma ago

867 (Pleuger et al. 2008; Beltrando et al. 2010). The backfolding and backthrusting,

868 which is so typical for the Western Alps (see profiles of Fig. 2 and accompanying

869 text) and which affected a pre-existing nappe pile, also occurred during stage 2 (e.g.

870 Bucher et al. 2003). Hence, it is also associated with the indentation of the Ivrea

871 mantle wedge that underthrusted much of the internal zones of the Alps (see Fig 2a-

872 c). The spectacular backthrust over the Cenozoic cover of the Tertiary Piedmont

35 873 Basin shown at the eastern end the profiles shown in Fig. 2 a & b also formed during

874 stage 2.

875

876 4. Oroclinal bending and top SW thrusting in the southern part of the Western

877 Alps (Stage 3, 25-0 Ma)

878

879 The external Alps south of the Pelvoux massif are well known for late-stage

880 top-SW thrusting (e.g. Fry 1989, Lickorish & Ford 1998, Bürgisser & Ford 1998) that

881 is not seen farther to the north along the Western Alpine arc. By studying

882 overprinting relationships south of the Pelvoux massif (area at location 4 in Fig. 4)

883 Trullenque (2005) discovered that this top-SW thrusting actually post-dates stage 2

884 top WNW thrusting. Lickorish & Ford (1998) have shown that 10.5 km out of a total

885 of around 21 km shortening, associated with minor thrusting rooting in the basement

886 of the Argentera Massif, occurred in Late Miocene to Pliocene times. Ritz (1992),

887 based on a paleostress analysis in the Digne, Castellane and Nice thrust systems,

888 found that during Middle Miocene to recent times the σ1 trajectories varied from NE-

889 SW in the Digne area, gradually turning into a N-S orientation in the Castellane arc

890 and eventually to a NW-SE direction in the case of the Nice thrust sheet. This same

891 radial arrangement was found for the p-axes derived from stress inversion of

892 earthquake focal mechanism data by Delacou et al. (2004, their Fig. 5). This radial

893 pattern is very likely related to rotation around a pivot located east of the Argentera

894 massif in the area of the “Ligurian knot” (Laubscher et al. 1992), i.e. in an area where

895 the 3 Moho’s beneath Alps, Apennines and northern Tyrrhenian Sea meet (Spada et

896 al. 2013).

897 Three largely independent lines of evidence support the idea that oroclinal

36 898 bending in the southernmost Western Alps is to be seen in connection with the

899 structural observations mentioned above. Firstly, Collombet et al. (2002) reported

900 counterclockwise rotations about a subvertical axis in respect to stable Europe that

901 increase from 68° in the Ubaye region (area between profiles 1 and 2 of Figure 1) to

902 117° in the Ligurian Alps SE of Cuneo (Fig. 1). The data were collected in

903 sedimentary rocks of the Briançonnais tectonic unit. The age constrains for

904 magnetization are rather loose; the authors propose that the rocks were remagnetized

905 after the backthrusting phase (our stage 2) and therefore to be of Late Oligocene age.

906 Secondly, all the major tectonic elements of the internal Alps derived from the

907 Briançonnais and Piedmont-Liguria paleogeographical domains can be followed into

908 western Liguria and all the way to Genova and slightly beyond (Fig. 1). In the

909 Ligurian Alps all the pre-Oligocene Alpine structures, sealed by the Tertiary

910 Piedmont Basin, are top-S in present-day coordinates. Since the counterclockwise

911 rotation of stable Adria only amounts to 10±10° (van Hinsbergen et al., 2014b) or 20°

912 (Marton et al. 2011) at least a part of the 49° differential rotation of Liguria in respect

913 to the Ubaye region (117°-68°= 49°), namely 39±10°, is probably related to oroclinal

914 bending during stage 3. It is likely that oroclinal bending during stage 3 was partially

915 driven by eastward trench retreat of the Apennines slab. However, the rotation in the

916 Ubaye region itself, located north of the Argentera-Cuneo Line is probably also due

917 to stage 2 indentation of Adria. Thirdly, a 50° counterclockwise rotation phase of the

918 Corsica-Sardinia block occurred between 20.5 and 16 Ma ago (Gattacceca et al.

919 2007) around a pivot near the N tip of Corsica, as Corsica and Sardinia drifted away

920 from their pre-Miocene location south of southern France (Advokaat et al. 2014); this

921 virtually demands substantial counterclockwise rotation in Liguria associated with

922 oroclinal bending. Note that such counterclockwise rotation is directly demonstrated

37 923 by the ~45-50° anticlockwise rotation of mostly Oligocene sediments of the Tertiary

924 Piedmont Basin sealing pre-Oligocene Alpine structures evidenced by paleomagnetic

925 data (Maffione et al. 2008).

926 Oroclinal bending in the southernmost Western Alps during stage 3 is evidently

927 associated with orogeny in the Apennines. Slab rollback of the Adria plate towards N

928 in the northernmost Apennines is the underlying driving force for both oroclinal

929 bending and the rotation of Corsica and Sardinia. The discussion of the Alps-

930 Apennines transition, discussed in the next chapter, will shed more light to the

931 Miocene history of the Ligurian Alps that cannot be understood without discussing

932 orogeny in the Apennines and its spatial and temporal relation to orogeny in the Alps.

933

934 D. The Alps-Apennines transition in space and time

935 1. Tertiary Piedmont Basin and Epiligurian basins

936

937 Late Eocene to Neogene sedimentary basins found in the interior of the

938 Western Alpine arc, the Tertiary Piedmont Basin, and similar basins found on top of

939 the External Ligurides of the Northern Apennines, referred to as “Epiligurian” basins,

940 play a crucial role for understanding the spatial and temporal transitions between the

941 west- and north-verging Alps and the east- and north-verging Apennines (Fig. 6).

942 The southern limit of the outcropping pre-Pliocene parts of the Tertiary

943 Piedmont Basin (Fig. 1) transgresses the high-pressure lower plate units of the

944 Ligurian Alps (i.e. the high pressure Briançonnais and Piedmont-Liguria units), as

945 well as adjacent upper plate units of the Alps derived from the Piedmont-Liguria

946 Ocean found immediately east of Genova (i.e. the Antola nappe). As first pointed by

947 Elter & Pertusati (1973) the Antola nappe, predominantly consisting of Helminthoid

38 948 Flysch, is best considered as a former part of the Alps; they wrote: “La falda del

949 Monte Antola e le altre formazioni liguri comprese nel triangolo Genova-Val

950 Staffora-Levanto sono da considerarsi appartenenti alle Alpi in quanta sono saldate a

951 queste dai terreni trasgressivi del Bacino terziario Piemontese fin dall'Eocene”).

952 While earlier works suggested that this basin initiated by subsidence in a scenario of

953 crustal stretching (Mutti et al. 1995) there seems to be some consensus in more recent

954 publications that the Oligocene to Lower Miocene (Aquitanian) sediments of this

955 basin were laid down in a compressive to transpressive setting (e.g. Carrapa &

956 Garcia-Castellanos 2005; Mosca et al. 2010).

957 As shown in the profiles of Fig. 2 a & b the westernmost parts of the Tertiary

958 Piedmont Basin were installed onto the upper crust of the Adria plate, which in turn

959 overlies the Ivrea mantle wedge. It formed the foredeep of the backthrusted and -

960 folded Western Alps during the Oligocene, which later became buried under post-

961 Messinian cover. Such syn-sedimentary backthrusting to the east to southeast is seen

962 in the seismic sections of lines 3 and 4 of Mosca et al. (2010, their Fig. 4). Thrusting

963 is sealed by the Burdigalian unconformity that is ubiquitous in seismic sections of the

964 subsurface of the Po plain (e.g. Schumacher & Laubscher 1996 and references

965 therein). Such pre-Burdigalian thrusting is also seen along the front of the Southern

966 Alps in western Lombardy where it affects sediments of Oligocene to Lower

967 Miocene age, the Gonfolite Lombarda Group sensu lato (e.g. Bernoulli et al. 1989,

968 1993) that are part of the Southalpine foredeep (Garzanti & Malusa 2008) and

969 laterally connect with the Tertiary Piedmont Basin (Mosca et al. 2010). Subsidence in

970 the Piemonte region continued in Burdigalian and younger times, when this basin

971 started to become involved in N-directed thrusting in connection with the Apenninic

972 orogeny. The Torino and northern Monferrato hills immediately east of Torino (Fig.

39 973 1) are also considered as a part of the Tertiary Piedmont Basin by most authors

974 although they would better be considered a part of the Epiligurian basins (see below).

975 They morphologically mark the northern front of the Apenninic orogen that formed

976 in Late Miocene to Pliocene times. These two northernmost parts of the Tertiary

977 Piedmont Basin transgressed an “Alpine-type” basement in case of the Torino hills

978 (Piana 2000) and a part of the External Ligurides in the case of the northern

979 Monferrato hills (Laubscher 1971; Mosca et al. 2010; Festa & Codegone 2013).

980 Together with the rest of the Tertiary Piedmont Basin that transgressed Alpine

981 structures they became thrusted northwards as wedge-top basins, much like the so

982 called Epiligurian basins of the External Ligurides, described below, along-strike

983 farther to the east and southeast.

984 The term Epiligurian basins is used for wedge-top basins overlying the External

985 Ligurides of the northern Apennines (Fig. 6) whose age-range and sedimentary

986 characteristics are similar to those of the Tertiary Piedmont Basin (Molli et al. 2010

987 and references therein). The sedimentary sequence starts with the deposition of the

988 deep marine Middle to Upper Eocene Monte Piano marls above a regional

989 unconformity that indicates some pre-Apenninic deformation of parts of the External

990 Ligurides that represent a fossil ocean-continent transition (Marroni et al. 2001)

991 within the Adria plate that was part of the upper plate of the Alpine orogen. The Early

992 Oligocene Ranzano formation was laid down above a second unconformity and

993 received detritus from the metamorphic units of the Central Alps (Lepontine dome)

994 that started to become exhumed north of the Periadriatic line (Garzanti & Malusa

995 2008). In the Early Oligocene the External Ligurides were part of the Adriatic

996 foredeep in respect to the Southern Alps that formed due to “retroshearing”

997 (Beaumont et al 1994) within the Alpine orogen (Schmid et al. 1996). Subsidence and

40 998 siliciclastic sedimentation in the Epiligurian basins continued during the Late

999 Oligocene when these basins and the underlying units became involved in the

1000 Apenninic orogeny during its earliest stages. From ca. 23 Ma onward the Epiligurian

1001 basins evolved into wedge-top basins as the External Ligurides became underplated

1002 by the accretion of Subligurian and Tuscanide units within the evolving Apenninic

1003 orogen (Molli et al. 2010 and references therein). Very recently Piazza et al. (2016)

1004 provided evidence for intense syn-depositional sinistral transpression during

1005 Oligocene to Early Miocene times sealed by a Burdigalian unconformity within one

1006 of the Epiligurian basins. This sinistral transpression is contemporaneous with

1007 sinistral strike slip along the E-W striking Villalvernia-Varzi line (Elter & Pertusati

1008 1973) that defines the boundary between the Antola nappe often considered as a part

1009 of the Alps in the south and the External Ligurides considered as a part of the

1010 Apennines in the north (stippled line in Fig. 1 north of Genova). This sinistral strike

1011 slip movement was active in Late Oligocene to Early Miocene times as well (Di

1012 Giulio & Galabati 1995; Schumacher & Laubscher 1996). Laubscher (1991)

1013 considered the Villalvernia-Varzi line as accommodating the WNW-directed

1014 indentation of the Ivrea mantle wedge previously discussed (stage 2, during the

1015 evolution of the arc of the Western Alps). Although we consider the direct link

1016 between the sinistral Argentera-Cuneo line and the Villalvernia-Varzi line proposed

1017 by Laubscher (1991) as unlikely (see Figs. 1 & 6) sinistral strike slip motion along

1018 the Villalvernia-Varzi line and sinistral transpression in the Epiligurian basins were

1019 probably kinematically connected with stage 2 WNW-directed indentation of the

1020 Ivrea mantle wedge during the Oligocene.

1021 In summary, the Tertiary Piedmont Basin seals very substantial pre-Late

1022 Eocene Alpine structures. The Epiligurian basins are of a similar type and age but

41 1023 they were deposited above an angular unconformity onto a unit that never underwent

1024 substantial Alpine deformation, namely the External Ligurides of the Northern

1025 Apennines (Fig. 6). Later, i.e. from about 23 Ma onwards, the latter represented

1026 wedge-top basins riding on the External Ligurides, which thrusted over more external

1027 units of the Northern Apennines towards NE (Molli 2008; Molli et al. 2010, and

1028 references therein). In essence, the stratigraphic base of these basins post-dates the

1029 pre-Late Eocene top-NNW thrusting during stage 1 sinistral transpression in the

1030 Western Alps described earlier while it pre-dates the onset of top-NE thrusting during

1031 the Apenninic orogeny, which is contemporaneous with stage 3 oroclinal bending and

1032 top SW thrusting in the southern part of the Western Alps. It follows from this that

1033 substantial parts of these basins probably represented syn-orogenic deposits during

1034 stage 2 top-WNW thrusting and indentation of the Ivrea mantle wedge into the

1035 Western Alps at 35-25 Ma ago.

1036

1037 2. The Alps-Apennines transition in map view

1038

1039 An interpretative map sketch of the Alps-Apennines transition is presented in

1040 Fig. 6. Considerable parts of this area are either buried by deposits of the Tertiary

1041 Piedmont Basin and very young Cenozoic cover of the Po plain or by the waters of

1042 the Ligurian and North Tyrrhenian Sea; hence this compilation in Fig. 6 remains

1043 speculative in many respects.

1044 The continuation of the undeformed or little deformed Alpine foreland has to be

1045 located in Variscan Corsica that became thrusted by Alpine allochthonous units until

1046 the Late Eocene (Malavieille et al. 1998; Molli 2008). Note, however, that the

1047 present-day NW-SE strike of the Late Eocene Alpine front across the Ligurian Sea

42 1048 and Corsica restores to a N-S strike after the retro-deformation of 47° post-Eocene

1049 rotation of the Ligurian Alps (Maffione et al. 2008; see Figs. 5d&e) and 50° rotation

1050 of Corsica-Sardinia (Gattacceca et al. 2007). The Alpine allochthons in Corsica

1051 consist of, from bottom to top, the high-pressure units of the Tenda “massif”

1052 representing formerly subducted and subsequently exhumed parts of Variscan

1053 Corsica that remains devoid of an Alpine overprinted in the west, the high-pressure

1054 units of the Piedmont-Liguria Ocean and finally overlying non-metamorphic

1055 ophiolitic and continental nappes (Balagne and Nebbio nappes). The two lower units

1056 clearly represent lower plate units of the Alps that have to be correlated with the

1057 Briançonnais and Piedmont-Liguria units of the Western Alps, respectively. Although

1058 the origin of the uppermost non-metamorphic units is controversial (see Marroni &

1059 Pandolfi 2003 and references therein) we follow Malavieille et al. (1998) who argue

1060 for an origin of these highest units from the ocean-continent transition at the eastern

1061 margin of the Piedmont-Liguria Ocean. This makes them comparable to the non-

1062 metamorphic upper plate units of the Western Alps mapped in Figs. 1 & 6 (Préalpes

1063 Romandes, Embrunais-Ubaye and Western Liguria Helminthoid flysch). Oligocene-

1064 Miocene extension and transtension localized in the Gulf of Lion and the Ligurian

1065 Sea strongly affected the entire Corsican belt (Fig. 6), reducing its crustal thickness to

1066 normal thickness. In map view the width of the Alpine belt in Corsica is drastically

1067 reduced in comparison with the Western Alpine orogenic belt. We interpret this

1068 reduction in width to be partly related to the switch in subduction polarity between

1069 Alps and Apennines that led to a new west-dipping “Apenninic” subduction zone that

1070 must have evolved offshore eastern Corsica since latest Eocene or earliest Oligocene

1071 times (Fig. 6; Doglioni et al. 1998; Carminati et al. 2004; Molli & Malavieille 2011).

1072 This new plate boundary was located east of the Corsica basin (Mauffret et al.

43 1073 1999; see also CROP-03 and M-12A seismic sections in Finetti et al. 2001 and Finetti

1074 2005). It evolved after cessation of Alpine, southeastward subduction in the

1075 Oligocene, and was followed by the two-stage opening of the Gulf of Lion back-arc

1076 basin. In a first, Oligocene stage, at around 30 Ma Corsica-Sardinia rifted off the

1077 Provence margin of southern France (Séranne, 1999). This was followed by the 21-16

1078 Ma Corsica-Sardinia rotation around a pole located somewhere between the northern

1079 tip of Corsica and the Ligurian coast (Advokaat et al. 2014). The new plate boundary

1080 probably had a ~N-S-direction, slightly oblique to the NNW-SSE present-day strike

1081 of the Alpine orogen in Corsica (Fig. 6). In respect to the Apenninic orogeny both

1082 lower and upper plate units of Alpine Corsica occupied an upper plate position,

1083 remaining largely unaffected by Apenninic shortening, while easterly and northerly

1084 adjacent parts of the former Alps were heavily reworked when backthrusted onto the

1085 Adria continental margin and subsequently stretched as the Adria plate below the

1086 North Apennines rapidly rolled back eastward (Grandjacquet & Haccard 1977;

1087 Doglioni et al 1998; Rosenbaum & Lister 2004; Spakman & Wortel 2004; Marroni et

1088 al. 2010).

1089 The northern margin of the lower plate metamorphic core of the Ligurian Alps

1090 (AFSZ in Fig. 1) is not exposed being buried beneath the Cenozoic cover of the

1091 Tertiary Piedmont Basin and younger cover (Fig.1). In Fig. 6 it is mapped according

1092 to subsurface seismic data (Mosca et al. 2010). Along a transect located west of

1093 Genova (profile A in Mosca et al. 2010) the metamorphic units of the Ligurian Alps

1094 were thrusted northward by about 50 km over the Adria continental autochthonous

1095 during the Apenninic orogeny in Miocene-Pliocene times. This implies that the

1096 western continuation of the older (pre-25Ma) Argentera-Cuneo Line, delimiting the

1097 southern boundary of the indenting Ivrea mantle wedge and overlying internal

44 1098 Western nappe stack during the stage 2 top-WNW thrusting in the arc of the Western

1099 Alps, must be cut off by a N-S striking sinistral shear zone (Figs. 1 & 6). This

1100 putative shear zone, which is only schematically drawn in Fig. 6, delimits the western

1101 edge of the Ligurian Alps affected by these 50 km of Apenninic top N thrusting from

1102 a narrow strip of the Southern Alps retro-belt in the west.

1103 Eastward, i.e. north of Genova, the northern limit of the metamorphic core of

1104 the Ligurian Alps bends into a N-S strike and is exposed at the surface forming the

1105 well-known Sestri-Voltaggio line (or zone; Cortesogno & Haccard 1984) that is often

1106 taken as representing the Alps-Apennines junction (e.g. Castellarin 2001). Although

1107 reworked in later Neogene times (Crispini et al. 2009) this line basically represented

1108 a Late Paleocene to Early Eocene normal fault (Hoogerduijn Strating 1994) sealed by

1109 the deposits of the Tertiary Piedmont Basin. This normal fault juxtaposes the high-

1110 pressure rocks of the Voltri Group in the west (1.5-2.5 GPa), attributed to the Alpine

1111 lower plate, with low metamorphic grade units to the east (Sestri-Voltaggio Zone,

1112 various Ligurian flysch units and finally the Antola nappe; see Fig. 8d in Molli et al.

1113 2010), whose grade of metamorphism progressively decreases eastward and up-

1114 section (see Fig. 2 in Capponi et al. 2009). It finally reaches the diagenetic zone in

1115 terms of illite crystallinity (Ellero et al. 2001) over a short horizontal distance of only

1116 some 5-7 km. Following Elter & Pertusati (1973) who considered the Antola nappe

1117 as an integral part of the former Alps that became backthrusted during Alpine

1118 orogeny, we attribute the units east of the Sestri-Voltaggio normal fault to the Alpine

1119 upper plate (Fig. 6).

1120 The unknown Alpine-type basement of the Torino hills buried under the

1121 deposits of the Tertiary Piedmont Basin and the Antola nappe were mapped together

1122 as parts of a strip of former Alpine upper plate units mostly buried under Cenozoic

45 1123 deposits in Fig. 6. During the Apenninic orogeny, together with the metamorphic core

1124 of the Ligurian Alps, they thrusted the External Liguride units of the Monferrato hills

1125 along the Rio Freddo Fault Zone (Piana 2000) striking NW-SE between the Torino

1126 and Monferrato hills. The Rio Freddo Fault Zone links up with the E-W-striking

1127 Villalvernia-Varzi line (Fig. 1), a Late Oligocene to Early Miocene thrust, active

1128 during the deposition of a part of the Tertiary Piedmont basin fill, with a strong

1129 sinistral strike slip component (Laubscher et al. 1992; Schumacher & Laubscher

1130 1996). This sinistral transpressional fault is responsible for laterally transporting the

1131 External Ligurides exposed in the Monferrato hills in front of the Antola nappe and

1132 the Torino hills Alpine basement. At its eastern termination the Villalvernia-Varzi

1133 line bends into a N-S-strike as it transforms into an Apenninic top NNE thrust of the

1134 Antola nappe over the external Ligurides in the north and over the Internal Ligurides

1135 further south (Fig. 6). The polyphase history of the Antola nappe, with early and

1136 Alpine top-NW thrusting and folding, followed by top-NNE Apenninic thrusting over

1137 the Ligurides, is well documented (e.g. Marroni et al. 1999; 2002a; Levi et al. 2006).

1138 Offshore the continuation of this thrust is only tentatively drawn in Fig. 6; southward

1139 it is interpreted to eventually abut the northern continuation of the west-dipping

1140 upper/lower plate boundary of the Apennines orogen.

1141 The next eastward unit of the Apennines, the Internal Ligurides, has also to be

1142 considered as having represented an integral part of the former Alps. It occupied an

1143 upper plate position before is became backthrusted to the NNE during the Apenninic

1144 orogeny. The internal Ligurides are characterized by the presence of classical

1145 ophiolites and an Upper Jurassic to Lower Cretaceous sedimentary cover (cherts,

1146 Calpionella limestone and Palombini shales) associated with Upper Cretaceous–

1147 Paleocene turbiditic sequences (Marroni & Pandolfi 1996) that represent the pelagic

46 1148 cover of the ophiolitic basement of the Piedmont-Liguria ocean (Decandia & Elter

1149 1972) whose stratigraphy is virtually indistinguishable from that of the ophiolites of

1150 the Alps sensu stricto (Bernoulli et al. 2003). Moreover, since the pioneering work of

1151 Grandjacquet & Haccard (1977) several authors evidenced classical structural

1152 overprinting patterns at all scales, clearly showing that the early tectonic history of

1153 the Internal Ligurides was characterized by a top W to NW Alpine tectonic

1154 fingerprint, associated with low-grade metamorphism reaching some 300°C

1155 (Pertusati & Horrenberger 1975; Van Wamel 1987; Marroni & Pandolfi 1996; Ellero

1156 et al. 2001).

1157 The External Ligurides are far-travelled and detached units that are confined to

1158 the northernmost Apennines between Torino and Bologna (Fig. 6). They overlie thin

1159 slices of Subligurian units and the Cervarola unit representing flysch deposits of the

1160 Adriatic foredeep outcropping in two small windows (Molli et al. 2010). The

1161 External Ligurides dominantly consist of detached Cretaceous-Paleocene flysch

1162 sequences (Helminthoid flysch) overlying basal mélange-type complexes (Marroni et

1163 al. 1998, 2001, 2002b). Two main subgroups of units can be recognized: those

1164 associated with ophiolites and with ophiolite derived debris (Molli 1996), and others

1165 without ophiolites and associated with fragments of Mesozoic sedimentary sequences

1166 and conglomerates with continental Adria affinity often remarkably similar to those

1167 found in the Alpine upper plate units of the Préalpes Romandes (Elter et al. 1966).

1168 Hence, the External Ligurides are regarded as representing the former ocean-

1169 continent transition within the Adria plate. Very commonly the stratigraphic base of

1170 the Oligocene part of the Epiligurian basins only exhibits a modest angular

1171 unconformity up to few tens of degrees (Piazza et al. 2016). Due to the syn-

1172 depositional sinistral transpression during Oligocene to Early Miocene times sealed

47 1173 by a Burdigalian unconformity mentioned earlier, the base of the Burdigalian and

1174 younger deposits of the Epiligurian succession locally overlies overturned older

1175 sequences of Late Cretaceous to Early Miocene successions (Piazza et al. 2016). In

1176 summary, the External Ligurides represent an ocean-continent transition area within

1177 the Adria upper plate of the Alps that was only locally overprinted by substantial late

1178 Alpine deformations of Oligocene to Early Miocene age, largely synchronous with

1179 stage 2 deformation in the Western Alps.

1180 A substantial Alpine overprint has not been documented in the southerly

1181 adjacent Tuscan Ligurides, representing a Piedmont-Liguria ophiolitic sequence

1182 identical to that observed in the Internal Ligurides or the Alps (Principi et al. 2004).

1183 Due to massive extension locally leading to core complex formation (Brunet et al.

1184 2000), Tuscany exposes deep portions of the Apenninic orogen that locally reached

1185 blueschist facies conditions with pressures up to some 1.3 GPa (Rossetti et al. 2002),

1186 mostly affecting the underlying Tuscan nappe that represents the lower plate unit

1187 during Apenninic orogeny. Locally, also the immediately overlying Tuscan Ligurides

1188 (Jolivet et al. 1998; Brunet et al. 2000; Rossetti et al. 2002) became involved in

1189 subduction during the Apenninic orogeny. Hence, it is likely that the Tuscan

1190 Ligurides only became involved in W-directed Apenninic subduction and suffered

1191 penetrative deformation within a W-dipping subduction channel. A significant

1192 ophiolitic complex that is part of the Tuscan Ligurides is exposed on Gorgona Island

1193 (Orti et al. 2002). Metasediments associated with these ophiolites suffered blueschist

1194 facies metamorphism reaching about 1.5 GPa at around 25 Ma (Rossetti et al. 2001).

1195 Gorgona island is located far offshore Tuscany (Fig. 6) and very close to the northern

1196 continuation of the new plate boundary that formed during the earliest stages of

1197 Apenninic orogeny associated with W-directed subduction and subsequent roll back

48 1198 of the mantle slab below the North Apennines (Spakman & Wortel 2004). W-directed

1199 subduction initiated east of the Corsica basin that represents a >8.5 km thick rift basin

1200 filled with undeformed Late Oligocene to Burdigalian sediments (Mauffret et al.

1201 1999; Finetti 2005) deposited as the Adriatic slab started to roll back, causing

1202 extension in the Corsica basin located in the Apenninic upper plate. The range of

1203 radiometric ages obtained for the high-pressure metamorphism in the Tuscan units is

1204 also significantly younger (26-20 Ma; Rossetti et al. 2002) when compared with the

1205 radiometric ages obtained in the ophiolitic units of Corsica where Cretaceous to latest

1206 Eocene ages (Vitale Brovarone & Herwartz 2013, and references therein) have been

1207 obtained. In summary, the Tuscan Ligurides are very likely to only have been

1208 affected by Apenninic orogeny, associated with W-directed subduction in latest

1209 Oligocene to Miocene times.

1210 The Mesozoic Adriatic passive margin stratigraphy that accreted to the

1211 overriding European plate (Corsica) and the Ligurides during the Apenninic orogeny

1212 is exposed in the Tuscan nappe. The Chattian-Aquitanian Macigno turbidites (Di

1213 Giulio 1999) represent the youngest sediments involved. This 2.5-3 km thick

1214 turbiditic sequence is testimony of the deposition of siliciclastics derived from the

1215 Alps, particularly from the Lepontine dome (Garzanti & Malusa 2008). This dome

1216 shed an immense volume of metamorphic basement estimated at 17-20 x 103 km3 (Di

1217 Giulio 1999) as the Lepontine dome was uplifted by retro-thrusting and exhumed by

1218 erosion. This led to deposition of the Macigno pile of turbidites onto the still

1219 undeformed Adriatic foredeep during retro-folding and thrusting of the Alps north of

1220 the Periadriatic Line, predating most of the backthrusting of the Southern Alps in

1221 Lombardy (Schmid et al. 1996). Related exhumation, erosion and re-deposition onto

1222 the still undeformed Adriatic continental margin are largely synchronous with stage 2

49 1223 top-WNW thrusting and WNW-directed indentation of the Ivrea mantle wedge (Stage

1224 2, 35-25 Ma).

1225 The more external Apennines thrust sheets involve flysch deposits of the

1226 foredeep only. The Cervarola thrust sheet is made up of Aquitanian to Early

1227 Langhian turbidites that still received detritus from the Alps. The more external

1228 Umbria-Marche thrust sheets involve the Marnoso Arenacea sequence of Serravallian

1229 to Tortonian age (Di Giulio 1999). The post-Messinian frontal imbricates are buried

1230 below younger deposits of the Po plain (Picotti & Pazzaglia 2008).

1231

1232 3. Lateral or temporal change in subduction polarity between Alps and

1233 Apennines? A discussion

1234

1235 Since the early days of Argand (1924) most authors (e.g. Mattauer et al. 1981;

1236 Durand-Delga 1984; Molli et al. 2006; Molli and Malavieille 2011; Handy et al.

1237 2010) regarded the pre-Oligocene Corsican nappe stack as a continuation of the

1238 Alpine orogen. Some authors (e.g. Principi and Treves 1984; Lahondère, 1996;

1239 Jolivet et al. 1998; Brunet et al., 2000), however, regarded Alpine Corsica as an

1240 integral part of the Apennines, both having formed in connection with a W- dipping

1241 subduction zone since Cretaceous times. They considered Variscan Corsica as a

1242 backstop of the Apennines orogenic wedge onto which the deepest part of the

1243 Apennines accretionary complex, namely Alpine Corsica, was backthrusted. This

1244 view essentially transfers the area of the Alps-Apennines transition more to the north

1245 into the area around Genova, which is rather unlikely in the light of the available data

1246 on Alps and northern Apennines summarized in the previous chapters.

1247 Debate also exists on whether the Alpine subduction zone with Eurasia in a

50 1248 lower plate position has continued beyond Corsica (e.g., Handy et al., 2010), or

1249 whether a lateral change in pre-Oligocene subduction polarity occurred in an area

1250 located somewhere between Corsica and Sardinia active since Late Cretaceous to

1251 Early Eocene times (e.g. Marchant & Stampfli 1997; Faccenna et al. 2003; Lacombe

1252 & Jolivet 2005; Lustrino et al. 2009; Argnani 2012; Advokaat et al., 2014; van

1253 Hinsbergen et al. 2014a). This transition is supposed to have occurred along a

1254 discrete trench-trench transform located SE of Corsica-Sardinia (e.g. Argnani 2012)

1255 or more diffusively, with partly overlapping trenches associated with opposite

1256 subduction polarity somewhere offshore east of Corsica (e.g. Lacombe & Jolivet

1257 2005). An often quoted argument in favor of this view is the existence of Late Eocene

1258 magmatism in Sardinia (Lustrino et al. 2009) within but a tiny fraction (Calabona

1259 subvolcanics) of the dominantly latest Oligocene and younger igneous magmatic

1260 province of Sardinia, starting around 28 Ma and peaking during the 22–18 Myr time

1261 range. This younger magmatic province undoubtedly has to be interpreted as related

1262 to W-directed Apenninic subduction and roll back. However, the Late Eocene

1263 magmatism at the Calabona locality was also interpreted to represent the effect of the

1264 northwest-directed subduction system that started to consume ancient oceanic

1265 lithosphere of the Tethys Ocean. This subduction would have initiated in a time span

1266 prior to the Late Eocene that is long enough to culminate in 100-150 km of

1267 subduction required to generate an arc (Lustrino et al. 2009). Given the very slow

1268 Africa-Europe convergence rates in the pre-Late Eocene (only 250 km of

1269 convergence occurred between Africa and Eurasia prior to 35 Ma at the position of

1270 Sardinia, and even less farther to the west, van Hinsbergen et al. 2014a), this would

1271 suggest that an Alpine-vergent thrusting in Sardinia and farther to the west would

1272 have been a back-thrust above a long-lived Apennines-type subduction. However

51 1273 later, Lustrino et al. (2013) argued, based on geochemical arguments, that the Eocene

1274 manifestations of the subduction related magmatism are due to the partial melting of

1275 Hercynian lower crust rather than to the dehydration processes of subducting Alpine

1276 Tethys oceanic crust. They pessimistically concluded that “whatever subduction dip,

1277 convergence rate and depth of partial melting are chosen, it is not possible to obtain a

1278 geologically sound model able to explain the spatial–temporal relationships of the

1279 western Mediterranean subduction-related igneous districts”.

1280 Since the pioneering work of Boccaletti et al. (1971) another group of authors

1281 (e.g. Guerrera et al. 1993; Doglioni et al. 1999; Michard 2006; Molli 2008; Handy et

1282 al. 2010) favored the idea of a temporal rather than lateral change in subduction

1283 polarity for the entire Apennines area (see discussion in Michard et al. 2006, their

1284 figure 9). The validity of this hypothesis is intimately linked to the question about

1285 kinematics and timing of thrusting, and, paleogeographic origin of the Calabrian-

1286 Peloritanian Belt that consists of high-grade Variscan basement nappes thrusted on

1287 top of Ligurian-type ophiolites (Schenk 1990; Vitale & Ciarcia 2013). It is also

1288 related to the question as to weather the Calabrian belt is part of an ALKAPECA

1289 block (Bouillin et al. 1986; Michard et al. 2006; Handy et al. 2010) rather than a part

1290 of Sardinia, i.e. Europe.

1291 The data and their interpretation presented in the previous chapters favor a

1292 temporal rather than lateral change in subduction polarity at least for the Northern

1293 Apennines and their transition into the Alps. However, a more detailed discussion of

1294 the southern Apennines of Calabria is outside the scope of this contribution. The

1295 controversy around a lateral vs. temporal change in subduction polarity in Calabria

1296 needs be solved by future investigations in Calabria.

1297 Although our data and interpretations support a temporal change in subduction

52 1298 polarity in the Northern Apennines realm and Corsica this interpretation faces a

1299 serious problem that is not yet solved. It is not easy to explain why this change

1300 occurred so shortly after the end of E-directed Alpine subduction in Corsica (about

1301 35-37 Ma; Vitale Bavarone & Herwartz 2013). The switch in subduction polarity

1302 around Corsica is not precisely known but probably occurred in Latest Eocene to

1303 Early Oligocene times. According to kinematic and paleomagnetic reconstructions

1304 (Séranne 1999; Gattacecca et al. 2007; Advokaat et al. 2014) fast roll back of the East

1305 Ligurian oceanic crust and rotation of Corsica-Sardinia commenced at around 30Ma,

1306 followed by rotation of Sardinia between 21-16 Ma. This only leaves some 5-7 Ma

1307 for slab break and change of subduction polarity and sufficient depth of subduction in

1308 order to trigger roll back of the W-dipping slab. Apenninic W-directed subduction

1309 and fast roll-back was certainly triggered by the negative buoyancy of large parts of

1310 the Piedmont-Liguria Ocean that still stayed open when Alpine orogeny in the

1311 Ligurian Alps ended at the end of the Eocene (Handy et al. 2010). This points to a

1312 change from plate convergence-driven Alpine E- to SE-directed subduction in

1313 Western Alps and Corsica toward W-directed subduction almost entirely driven by

1314 slab roll back in the Northern Apennines.

1315

1316 E. Summary and conclusions

1317 The geological interpretation of the P-wave velocity model of Diehl et al.

1318 (2009) allows for depicting unprecedented details concerning the outlines of this

1319 Adriatic mantle wedge and integrates its outline into the overall structure of the

1320 Western Alps. Important along-strike changes are visualized in five radially arranged

1321 crustal-scale transects across the Alps (Fig. 2) and the tectonic maps of Figs. 1 & 6).

1322 The interpretation of these findings reveals an important difference between Eastern

53 1323 and Western Alps. While the upper plate of the Eastern Alps is defined by the

1324 Austroalpine nappe system consisting of crustal imbricates of continental material,

1325 the upper plate of the Western Alps is formed by non-metamorphic klippen of the

1326 Piedmont-Ligurian Oceanic crust and lithosphere (Molli 2008).

1327 Based on this, and combined with a review of existing structural data, a 3-stage

1328 evolution is proposed for the formation of the arc of the Western Alps. A first stage

1329 (Late Cretaceous-35 Ma) is characterized by top-NNW thrusting in sinistral

1330 transpression causing at least some 260km displacement of the E-W-striking Central

1331 and Eastern Alps, together with the Adria micro-plate, towards N to NNW with

1332 respect to stable Europe. A second stage (35-25 Ma) is characterized by lateral

1333 indentation of Adria including the Ivrea mantle wedge at its leading edge towards the

1334 WNW. This culminated in some 100-150km of right-lateral strike-slip of the Ivrea

1335 mantle wedge, Western Alps and Adriatic micro-plate relative to the eastern Alps and

1336 Eurasia. This was accommodated in the Western Alpine nappe pile by top-WNW

1337 thrusting in the external zones of the central portion of the arc along the Penninic

1338 Frontal thrust combined with underthrusting of the Ivrea mantle wedge below the

1339 Western Alpine nappe pile. The third and final stage of arc formation (25-0Ma) was

1340 associated with orogeny in the Apennines and led to oroclinal bending in the

1341 southernmost Western Alps, simultaneously with the 50° counterclockwise rotation

1342 of the Corsica-Sardinia block. Particularly during the second stage of the evolution of

1343 the arc of the Western Alps the Ivrea mantle wedge, exhumed to shallow crustal

1344 depth already during Jurassic rifting at the passive continental margin of the Adria

1345 continent and representing a high strength wedge, played a crucial role by controlling

1346 the lateral ends of indentation near Cuneo and the Rhone-Simplon fault.

1347 The third stage of the evolution of the arc of the Western Alps calls for an

54 1348 analysis of the Alps-Apennines transition zone. A pre-requisite for understanding this

1349 transition is the realization that those parts of the Piemont-Liguria oceanic domain

1350 that form the upper plate in the Western Alps remained open until the Late Eocene. A

1351 combination of the gravitational potential of a southwards enlarging domain of old

1352 and oceanic lithosphere, the arrival of the Corsican continent in the Alpine trench

1353 resisting subduction, and ongoing Africa-Europe convergence induced a change in

1354 subduction polarity from Alpine S to SE-directed subduction towards NW-directed

1355 subduction and roll back of the Apennines slab, pulled by the negative buoyancy of

1356 old oceanic lithosphere large parts of which remained unaffected by Alpine orogeny.

1357

1358

1359 Acknowledgements

1360

1361 We thank Peter Nievergelt for his great help during the preparation of the

1362 figures. The careful and critical reviews of Rob Butler and Niko Froitzheim helped

1363 much in improving an earlier version of the manuscript. We thank our colleagues of

1364 the new research initiative AlpArray for many inspiring discussions, and we look

1365 forward to learn about the new insights and concepts that will certainly result from

1366 this multi-disciplinary program.

1367

1368 References

1369

1370 Advokaat, E.L., van Hinsbergen D.J.J., Maffione, M., Langereis Cor G., Vissers,

1371 R.L.M., Cherchi, A., Schroeder, R., Madani, H. & Columbu, S. (2014). Eocene

1372 rotation of Sardinia, and the paleogeography of the western Mediterranean

55 1373 region. Earth and Planetary Science Letters, 401, 183-195.

1374 Argand, E. (1924). Des Alpes et de l'Afrique. Bulletin de la Société vaudoise des

1375 sciences naturelles, 55, 233–236.

1376 Argnani, A. (2012). Plate motion and the evolution of Alpine Corsica and Northern

1377 Apennines. Tectonophysics, 579, 207–219.

1378 Babist, J., Handy, M.R., Konrad-Schmolke, M. & Hammerschmidt, K. (2006).

1379 Precollisional, multistage exhumation of subducted continental crust: The Sesia

1380 Zone, western Alps. Tectonics, 25, TC6008, doi:10.1029/2005TC001927.

1381 Barféty, J.C., Lemoine. M., de Graciansky. P.C., Tricart. P. & Mercier. D. (1995).

1382 Notice explicative de la Carte géologique de France 1:50’000 Feuille Briançon

1383 (823). BRGM Orléans. 180 pp.

1384 Barnes, J., Beltrando, M., Lee, C.T., Cisneros, M., Loewy, S. & Chin, E. (2014).

1385 Geochemistry of Alpine serpentinites from rifting to subduction: A view across

1386 paleogeographic domains and metamorphic grade. Chemical Geology, 389, 29-

1387 47.

1388 Bayer, R., Carozzo, M.T., Lanza, R., Miletto, M. & Rey D. (1989). Gravity modeling

1389 along the ECORS-CROP vertical seismic reflection profile through the western

1390 Alps. Tectonophysics, 162, 203-218.

1391 Beaumont, C., Fullsack, P.H. & Hamilton, J. (1994). Styles of crustal deformation in

1392 compressional orogens caused by subduction of the underlying lithosphere.

1393 Tectonophysics, 232, 119-132.

1394 Bellahsen N., L. Jolivet L., Lacombe O., Bellanger M., Boutoux A., Garcia S.,

1395 Mouthereau F., Le Pourhiet L. & Gumiaux C. (2012). Mechanisms of margin

1396 inversion in the external Western Alps: Implications for crustal rheology.

1397 Tectonophysics, 560-561, 62-83.

56 1398 Beltrando, M., Compagnoni, R., Ferrando, S., Mohn, G., Frasca, G., Odasso, N. &

1399 Vukmanovič, Z. (2015a). Adriatic margin: Canavese, Italy: Crustal thinning

1400 and mantle exhumation in the Levone area (Southern Canavese zone, Western

1401 Alps) Journal of the Virtual Explorer, 49.

1402 Beltrando, M., Stöckli, D.F., Decarlis, A. & Manatschal, G. (2015b). A crustal-scale

1403 view at rift localization along the fossil Adriatic margin of the Alpine Tethys

1404 preserved in NW Italy. Tectonics, 34, 1927-1951.

1405 Beltrando, M., Lister, G.S., Rosenbaum, G., Richards, S. & Forster, M.A. (2010).

1406 Recognizing episodic lithospheric thinning along a convergent plate margin:

1407 The example of the Early Oligocene Alps. Earth-Science Reviews, 103. 81–98.

1408 Berger, A. & Mercolli, I. (2006). Tectonic and Petrographic Map of the Central

1409 . Carta geologica speciale 127, map sheet 43 Sopra Ceneri.

1410 Federal Office of Topography swisstopo.

1411 Berger, A. & Bousquet, R. (2008). Subduction-related metamorphism in the Alps:

1412 review of isotopic ages based on petrology and their geodynamic consequences.

1413 In S. Siegesmund, B. Fügenschuh & N. Froitzheim (Eds.) Tectonic Aspects of

1414 the Alpine-Dinaride-Carpathian System (p. 117-144). Geological Society

1415 London, Special Publications, 298, 117–144.

1416 Berger, A., Rosenberg, C. & Schmid, S.M. (1996). Ascent, emplacement and

1417 exhumation of the Bergell pluton within the Southern Steep Belt of the Central

1418 Alps. Schweizerische Mineralogische und Petrographische Mitteilungen, 76,

1419 357-382.

1420 Berger, A., Mercolli, I. & Engi, M. (2005). The central Lepontine Alps: Notes

1421 accompanying the tectonic and petrographic map sheet Sopra Ceneri

1422 (1:100’000). Schweizerische Mineralogische und Petrographische

57 1423 Mitteilungen, 85, 109-146.

1424 Bernoulli, D., Bertotti, G. & Zingg, A. (1989). Northward thrusting of the Gonfolite

1425 Lombarda ("South-Alpine Molasse") onto the Mesozoic sequence of the

1426 Lombardian Alps: Implications for the deformation history of the Southern

1427 Alps. Eclogae Geologicae Helvetiae, 82, 841-856.

1428 Bernoulli, D., Giger, M. & Müller, D.W. (1993). Sr-isotope-stratigraphy of the

1429 Gonfolite Lombarda Group ("South-Alpine Molasse", northern Italy) and

1430 radiometric constraints for its age of deposition. Eclogae Geologicae Helvetiae,

1431 86, 751-767.

1432 Bernoulli, D., Manatschal, G., Desmurs, L., and Müntener, O., (2003). Where did

1433 Gustav Steinmann see the trinity? Back to the roots of an Alpine ophiolite

1434 concept. In Y. Dilek, & S. Newcomb, S. (Eds.), Ophiolite concept and the

1435 evolution of geological thought (pp. 93–110). Boulder, Colorado. Geological

1436 Society of America Special Paper 373.

1437 Bertotti, G. (1991). Early Mesozoic extension and alpine shortening in the western

1438 Southern Alps: the geology of the area between Lugano and Menaggio

1439 (Lombardy, Northern Italy). Memorie di Scienze Geologiche Padova, XLIII,

1440 17-123.

1441 Bertotti, G., Picotti, V., Bemoulli, D. & Castellarin, A. (1993). From rifting to

1442 drifting: Tectonic evolution of the South-Alpine upper crust from the Triassic

1443 to the Early Cretaceous. Sedimentary Geology, 86, 53-76.

1444 Blundell, D.J., Freeman, R. & Mueller, S. (1992). A Continent Revealed. The

1445 European Geotraverse, Structure and Dynamic Evolution. Cambridge

1446 University Press.

1447 Boccaletti, M., Elter, P. & Guazzone, G. (1971). Plate tectonic models for the

58 1448 development of the Western Alps and Northern Apennines. Nature Physical

1449 Science, 234/49, 108-111.

1450 Byerlee, J., (1978). Friction of rocks. Pure and Applied Geophysics, 116, 615–626.

1451 Bouillin, J.-P., Durand-Delga, M., Olivier, P. (1986). Betic-Rifian and Tyrrhenian

1452 Arcs: Distinctive features, genesis, and development Stages. In F.C. Wezel

1453 (Ed.), The Origin of Arcs (pp.281-304), Amsterdam, Elsevier.

1454 Bousquet, R., Goffé, B., Henry, P., LePichon, X. & Chopin, C. (1997). Kinematic,

1455 thermal and petrological model of the Central Alps: Lepontine metamorphism

1456 in the upper crust and eclogitization of the lower crust. Tectonophysics, 273,

1457 105-127.

1458 Bousquet, R., Goffé, B., Vidal, O., Oberhänsli, R. & Patriat, M. (2002). The tectono-

1459 metamorphic history of the Valaisan domain from the Western to the Central

1460 Alps: new constraints for the evolution of the Alps. Geological Society of

1461 America Bulletin, 114, 207–225.

1462 Bousquet, R., Oberhänsli, R., Schmid, S.M., Berger, A., Wiederkehr, M., Robert, C.,

1463 Möller, A., Rosenberg, C., Zeilinger, G., Molli, G. & Koller, F. (2012a).

1464 Metamorphic Framework of the Alps. Commission for the Geological Map of

1465 the World; Subcommission for Magmatic and Metamorphic Maps. IUGS and

1466 IUGG, Paris. www.ccgm.org.

1467 Bousquet R., Schmid S.M., Zeilinger G, Oberhänsli R., Rosenberg C., Molli G.,

1468 Robert C., Wiederkehr M., Rossi Ph. (2012b). Tectonic framework of the

1469 Alps. Map 1: 1 000 000. Commission for the Geological Map of the World;

1470 Subcommission for Magmatic and Metamorphic Maps. IUGS and IUGG,

1471 Paris. www.ccgm.org.

59 1472 Boyden, J.A., Müller, R.D., Gurnis, M., Torsvik, T.H., Clark, J.A., Turner, M., Ivey-

1473 Law, H., Watson, R.J. & Cannon, J.S. (2011). Next-generation plate-tectonic

1474 reconstructions using GPlates. In G.R. Keller & C. Baru (Eds),

1475 Geoinformatics (pp. 95-114). Cambridge University Press.

1476 Brunet, C., Monié, P., Jolivet, L. & Cadet, J.P. (2000). Migration of compression and

1477 extension in the Tyrrhenian Sea, insights from 40Ar/39Ar ages on micas

1478 along a transect from Corsica to Tuscany. Tectonophysics, 321, 127–155.

1479 Bucher, S., Schmid, S.M., Bousquet, R. and Fügenschuh, B. (2003). Late-stage

1480 deformation in a collisional orogen (Western Alps): nappe refolding, back-

1481 thrusting or normal faulting? Terra Nova, 15, 109-117.

1482 Bucher S., Ulardic, C., Bousquet R., Ceriani, S., Fügenschuh B. & Schmid, S. M.

1483 (2004). Tectonic evolution of the Briançonnais units along the ECORS-CROP

1484 transect through the Italian-French Western Alps. Eclogae Geologicae

1485 Helvetiae, 97, 321-345.

1486 Bürgisser, J. & Ford, M. (1998). Overthrust shear deformation of a foreland basin;

1487 structural studies south-east of the Pelvoux massif, SE France. Journal of

1488 Structural Geology, 20, 1455-1475.

1489 Burkhard, M. (1988). L'Helvétique de la bordure occidentale du massif de l'Aar

1490 (évolution tectonique et métamorphique) Eclogae Geologicae Helvetiae, 81,

1491 63-114.

1492 Burkhard, M. & Sommaruga A. (1998). Evolution of the western Swiss Molasse

1493 basin: Structural relations with the Alps and the Jura belt. In: Mascle C. et al.

1494 (Eds): Cenozoic Foreland Basins of Western Europe. Special Publications

1495 Geological Society London, 134, 279–298.

1496 Butler, R.W.H. (1983). Balanced cross sections and their implications for the deep

60 1497 structure of the northwest Alps. Journal of Structural Geology, 5, 125-137.

1498 Butler, R.W.H. (1986). Thrust tectonics, deep structure and crustal subduction in the

1499 Alps and Himalayas. Journal of the Geological Society of London, 143, 857-

1500 873.

1501 Butler, R.W.H. (1992). Thrust zone kinematics in a basement-cover imbricate stack;

1502 eastern Pelvoux Massif, French Alps. Journal of Structural Geology, 14, 29–

1503 40.

1504 Butler, R.H.W (2013). Area balancing as a test of models for the deep structure of

1505 mountain belts, with specific reference to the Alps. Journal of Structural

1506 Geology, 52, 2-16.

1507 Caby. R. (1996). Low-angle extrusion of high-pressure rocks and the balance

1508 between outward and inward displacements of Middle Penninic units in the

1509 western Alps. Eclogae Geologicae Helvetiae, 89/1, 229-267.

1510 Cardello, G.L. & Mancktelow, N. (2015). Veining and post-nappe transtensional

1511 faulting in the SW Helvetic Alps (Switzerland). Swiss Journal of Geosciences,

1512 108, 379-400.

1513 Carey, S. W. (1955). The orocline concept in geotectonics, part I. Papers and

1514 proceedings of the Royal Society of Tasmania, 89, 255–288.

1515 Carminati, E., Doglioni, C. & Scrocca, D. (2004). Alps vs Apennines. In U. Crescenti

1516 Geology of Italy: Special Volume of the Italian Geological Society for the IGC

1517 32 Florence 2004 (pp. 141-151). Società geologica italiana, Roma.

1518 Carrapa, B. & Garcia-Castellanos D. (2005). Western Alpine back-thrusting as

1519 subsidence mechanism in the Tertiary Piedmont Basin (Western Po Plain, NW

1520 Italy). Tectonophysic,s 406, 197– 212.

1521 Capponi, G., Crispini, L. & Scambelluri, M. (2009). Discussion: Comment on

61 1522 “Subduction polarity reversal at the junction between the Western Alps and the

1523 Northern Apennines, Italy”, by G. Vignaroli, C. Faccenna, L. Jolivet, C.

1524 Piromallo, F. Rossetti. Tectonophysics, 465, 221–226.

1525 Cassano, E., Anelli, L., Fichera, R. & Cappelli, V. (1986). Pianura Padana.

1526 Interpretazione integrate di dati geofisici e geologici. 73. Congresso Società

1527 Geologica Italiana. Centro Stampa AGIP Roma, pp. 1-28.

1528 Castellarin, A. (2001). Alps-Apennines and Po-Plain – Frontal Apennines

1529 relationships. In G.B. Vai et al., Anatomy of an Orogen. The Apennines and

1530 adjacent Mediterranean Basins (pp. 177-196). London, Kluver.

1531 Chemenda, A.I., Mattauer, M., Malavieille, J. & Bokun, A.N. (1995). A mechanism

1532 for syn-collisional rock exhumation and associated normal faulting: Results

1533 from physical modeling. Earth and Planetary Science Letters, 132, 225-232.

1534 Ceriani, S., Fügenschuh, B. and Schmid, S.M. (2001). Multi-stage thrusting at the

1535 „Penninic Front“ in the Western Alps between Mont Blanc and Pelvoux

1536 massifs. International Journal of Earth Sciences, 90, 685-702.

1537 Ceriani, S. & Schmid, S.M. (2004). From N-S collision to WNW-directed post-

1538 collisional thrusting and folding: Structural study of the Frontal Penninic

1539 Units in Savoie (Western Alps, France). Eclogae geologicae Helvetiae, 97,

1540 347-369.

1541 Chemenda, A.I., Mattauer, M., Malavieille, J. & Bokun, A.N. (1995). A mechanism

1542 for syn-collisional rock exhumation and associated normal faulting: Results

1543 from physical modeling. Earth and Planetary Science Letters, 132, 225-232.

1544 Collombet, M., Thomas, J.C., Chauvin, A., Tricart, P., Bouillin, J.P. & Gratier, J.P.

1545 (2002). Counterclockwise rotation of the western Alps since the Oligocene:

1546 New insights from paleomagnetic data. Tectonics, 21, 10.1029/2001TC901016.

62 1547 Compagnoni, R., Dal Piaz, G. V., Hunziker, J. C., Gosso, G., Lombardo, B. &

1548 Williams, P. F. (1977). The Sesia-Lanzo Zone, a slice of continental crust with

1549 Alpine high pressure-low temperature assemblages in the Western Italian Alps.

1550 Rendiconti della Società Italiana di Mineralogia e Petrologia, 33, 281-334.

1551 Cortesogno, L. & Haccard, D. (1984). Note illustrative alla carta geologica della

1552 Zona Sestri-Voltaggio. Memorie Società Geologica Italiana, 28, 115–150.

1553 Crispini, L., Federico, L., Cappini, G. & Spagnolo, C. (2009). Late orogenic

1554 transpressional tectonics in the «Ligurian Knot». Italian Journal of

1555 Geosciences, 128, 433-441.

1556 Dal Piaz G.V., Cortiana G., Del Moro A., Martin S., Pennacchioni G., Tartarotti P.

1557 2001. Tertiary age and paleostructural inferences of the eclogitic imprint in the

1558 Austroalpine outliers and Zermatt-Saas ophiolite, western Alps. International

1559 Journal of Earth Sciences, 90, 668–684.

1560 Decandia, F.A. & Elter, P. (1972). La ‘zona’ ofiolitifera del Bracco nel settore

1561 compreso tra Levanto e la Val Graveglia (Appennino ligure). Memorie della

1562 Società Geologica Italiana, 11, 503–530.

1563 de Graciansky, P.C., Dardeau, G., Lemoine, M., Tricart, P. (1989). The inverted

1564 margin of the French Alps and foreland basin inversion. Geological Society

1565 London Special Publications, 44, 87–104.

1566 Di Giulio, A. (1999). Mass transfer from the Alps to the Apennines: volumetric

1567 constraints in the provenance study of the Macigno–Modino source–basin

1568 system, Chattian–Aquitanian, northwestern Italy. Sedimentary Geology, 124,

1569 69–80.

1570 Di Giulio, A. & Galbiati, B. (1995). Interaction between tectonics and deposition into

1571 an episutural basin in the Alps-Apennine knot. In R. Polino & R. Sacchi (Eds.),

63 1572 Atti del convegno "Rapporti Alpi-Appennino" (pp. 113-128). Accademia

1573 Nazionale delle Scienze detta dei XL, 14, Roma.

1574 Delacou, B., Sue, C., Champagnac, J.D. & Burkhard, M. (2004). Present-day

1575 geodynamics in the bend of the western and central Alps as constrained by

1576 earthquake analysis. Geophysical Journal International, 158, 753-774.

1577 Diehl, T., Husen, S., Kissling E. & Deichmann, N. (2009). High-resolution 3-D P-

1578 wave model of the Alpine crust. Geophysical Journal International, 179, 1133–

1579 1147.

1580 Doglioni, C., Mongelli, F. & Pialli, G. (1998). Boudinage of the Alpine belt in the

1581 Apenninic back-arc. Memorie della Società Geologica Italiana, 52, 457-468.

1582 Doglioni, C., Fernandez, M., Gueguen, E., & Sabat, F. (1999). On the interference

1583 between the early Apennines–Maghrebides back-arc extension and the Alps-

1584 Betics orogen in the Neogene geodynamics of the Western Mediterranean.

1585 Bolletino Società geologica Italiana, 118, 75–89.

1586 Druiventak, A., Trepmann, C.A., Renner, J. & Hanke, K. (2011). Low-temperature

1587 plasticity of olivine during high stress deformation of peridotite at lithospheric

1588 conditions - An experimental study. Earth and Planetary Science Letters, 311,

1589 199–211.

1590 Dumont, T., Champagnac, J.D., Crouzet, C. & Rochat, P. (2008). Multistage

1591 shortening in the Dauphiné zone (French Alps): the record of Alpine collision

1592 and implications for pre-Alpine restoration. Swiss Journal of Geosciences, 101

1593 Supplement 1, 89–110.

1594 Dumont, T., Simon-Labric, T., Authemayou, C. & Heymes, T. (2011). Lateral

1595 termination of the north-directed Alpine orogeny and onset of westward escape

1596 in the Western Alpine arc: Structural and sedimentary evidence from the

64 1597 external zone. Tectonics, 30, doi:10.1029/2010TC002836.

1598 Dumont, T., Schwartz, S., Guillot, S., Simon-Labric, T., Tricart, P. & Jourdan, S.

1599 (1012). Structural and sedimentary records of the Oligocene revolution in the

1600 Western Alpine arc. Journal of Godynamics, 56-57, 18-28.

1601 Durand-Delga, M. (1984). Principaux traits de la Corse Alpine et corrélations avec

1602 les Alpes Ligures. Memorie della Società Geologica Italiana, 28, 285–329.

1603 Ellero, A., Leoni, L., Marroni, M. & Sartori, F. (2001). Internal Liguride Units from

1604 Central Liguria, Italy: new constraints to the tectonic setting from white mica

1605 and chlorite studies. Schweizerische Mineralogische und Petrographische

1606 Mitteilungen, 81, 39-53.

1607 Elter, P. & Pertusati, P. (1973). Considerazioni sul limite Alpi-Apennino a sulle sue

1608 relazioni con l’arco delle Alpi Occidentali. Memorie della Società Geologica

1609 Italiana, 12, 359-375.

1610 Elter, G., Elter, P., Sturani, C. & Weidmann, M. (1966). Sur la prolongation du

1611 domaine ligure de l’Apennin dans le Montferrat et les Alpes et sur l’origine de

1612 la Nappe de Simme s.l. des Préalpes romandes et chablaisiennes. Archives des

1613 Sciences, 19, 279–377.

1614 Escher, A., Hunziker, J.C., Marthaler, M., Masson, H., Sartori, M. & Steck, A.

1615 (1997). Geologic framework and structural evolution of the western Swiss-

1616 Italian Alps. In: Pfiffner A.O. et al. (Eds.): Deep Structure of the :

1617 Results From NRP 20 (pp. 205-221). Birkhäuser, Basel, Switzerland.

1618 Faccenna, C., Jolivet, L., Piromallo, C. & Morelli, A. (2003). Subduction and the

1619 depth of convection in the Mediterranean mantle. Journal of Geophysical

1620 Research 108, doi:10.1029/2001JB001690.

1621 Festa, A. & Codegone, G. (2013). Geological map of the External Ligurian Units in

65 1622 western Monferrato (Tertiary Piedmont Basin, NW Italy). Journal of Maps, 9,

1623 84-97.

1624 Fassmer, K., Obermüller, G., Nagel, T. J., Kirst, F., Froitzheim, N., Sandmann, S.,

1625 Miladinova, I., Fonseca, R. O. C. & Münker, C. (2016). High-pressure

1626 metamorphic age and significance of eclogite-facies continental fragments

1627 associated with oceanic lithosphere in the Western Alps (Etirol-Levaz Slice,

1628 Valtournenche, Italy). Lithos, 252-253, 145-159.

1629 Finetti, I.R. (editor) (2005). CROP PROJECT: Deep Seismic Exploration of the

1630 Central Mediterranean and Italy (779pp & 72 plates). Amsterdam: Elsevier.

1631 Finetti, I.R., Boccaletti, M., Bonini, M., Del Ben, A., Geletti, R., Pipam, N. & Sani,

1632 F. (2001). Crustal section based on CROP seismic data across the North

1633 Tyrrhenian–Northern Apennines–Adriatic Sea. Tectonophysics, 343, 135– 163.

1634 Ford, M. & Lickorish, W.H. (2004). Foreland basin evolution around the western

1635 Alpine Arc. In: P. Joseph & S.A. Lomas (Eds.), Deep-Water Sedimentation in

1636 the Alpine Basin of SE France: New perspectives on the Grès d’Annot and

1637 related systems (pp.39-63). Geological Society London Special Publication 221.

1638 Ford, M., Duchêne, S., Gasquet, D. & Vanderhaeghe, O. (2006). Two-phase orogenic

1639 convergence in the external and internal SW Alps. Journal of the Geological

1640 Society, London, 163, 815–826.

1641 Fountain, D.M. (1976). The Ivrea-Verbano and Strona-Ceneri zones, Northern Italy:

1642 a cross-section of the continental crust: new evidence from seismic velocities of

1643 rock samples. Tectonophysics, 33, 145-165.

1644 Fox, M., Herman F., Willett, S.D. & Schmid, S.M. (2016). The exhumation history of

1645 the European Alps inferred from linear inversion of thermochronometric data.

1646 American Journal of Science, 316, 505-541.

66 1647 Froitzheim, N., Schmid, S.M. & Frey, M. (1996). Mesozoic paleogeography and the

1648 timing of eclogite-facies metamorphism in the Alps: A working hypothesis.

1649 Eclogae geologicae Helvetiae, 89, 81-110.

1650 Fry, N. (1989). Southwestward thrusting and tectonics of the western Alps. In: M.P.

1651 Coward et al. (Eds.), Alpine Tectonics (pp. 83-109). Geological Society London

1652 Special Publication 45.

1653 Fügenschuh, B. and Schmid, S.M. (2003). Late stages of deformation and

1654 exhumation of an orogen constrained by fission track data: A case study in the

1655 Western Alps. Geological Society of America Bulletin, 115, 1425-1440.

1656 Fujimoto, Y., Konob, Y., Hirajima, T., Kanagawa, K., Ishikawa, M. & Arima, M.

1657 (2010). P-wave velocity and anisotropy of lawsonite and epidote blueschists:

1658 Constraints on water transportation along subducting oceanic crust. Physics of

1659 the Earth and Planetary Interiors, 183, 219–228.

1660 Ganne, J., Marquer, D., Rosenbaum, G., Bertrand, J.M. & Fudral, S. (2006).

1661 Partitioning of deformation within a subduction channel during exhumation of

1662 high-pressure rocks: a case study from the Western Alps. Journal of Structural

1663 Geology, 28, 1193-1207.

1664 Garzanti, E. & Malusa, M.G. (2008). The Oligocene Alps: Domal unroofing and

1665 drainage development during early orogenic growth. Earth and Planetary

1666 Science Letters, 268, 487–500.

1667 Gattacceca, J., Deino, A., Rizzo, R., Jones, D.S., Henry, B., Beaudoin, B. &

1668 Vadeboin, F. (2007). Miocene rotation of Sardinia: New paleomagnetic and

1669 geochronological constraints and geodynamic implications. Earth and

1670 Planetary Science Letters, 258, 359–377.

1671 Gebauer D., Schertl, H.P., Brix M. & Schreyer W. (1997). 35 Ma old ultrahigh-

67 1672 pressure metamorphism and evidence for very rapid exhumation in the Dora

1673 Maira Massif, Western Alps. Lithos, 41, 5-24.

1674 Gidon, M. (1972). Les chaînons briançonnais et subbriançonnais de la rive gauche de

1675 la Stura entre le Val de l'Arma (province de Cuneo Italie). Géologie Alpine, 48,

1676 87-120.

1677 Giuntoli, F. & Engi, M. (2016). Internal geometry of the central Sesia Zone (Aosta

1678 Valley, Italy): HP tectonic assembly of continental slices. Swiss Journal of

1679 Geosciences DOI 10.1007/s00015-016-0225-4.

1680 Grandjacquet, C. & Haccard, D. (1977). Position structural et role paléogéographique

1681 de l’unité du Bracco au sein du context ophiolitique liguro-piémontais

1682 (Apennin – Italie). Bulletin de la Société géologique de France, 19, 901-904.

1683 Grasemann, B. & Mancktelow, N.S. (1993). Two-dimensional thermal modeling of

1684 normal faulting: The Simplon fault zone, central Alps, Switzerland.

1685 Tectonophysics, 225, 155–165.

1686 Guerrera, F., Martín-Algarra A. & Perrone, V. (1993). Late Oligocene- Miocene syn-

1687 /-late-orogenic successions in Western and Central Mediterranean Chains from

1688 the Betic Cordillera to the Southern Apennines. Terra Nova, 5, 525–544.

1689 Guillot, S. & Ménot, R.P. (2009). Paleozoic evolution of the External Crystalline

1690 Massifs of the Western Alps. Comptes Rendus Geoscience, 341, 253–265.

1691 Handy, M.R. (1987). The structure, age and kinematics of the Pogallo Fault Zone;

1692 Southern Alps, northwestern Italy. Eclogae geologicae Helvetiae, 80, 593-632.

1693 Handy, M. R. & Zingg, A. (1991). The tectonic and rheologic evolution of an

1694 attenuated cross-section of the continental crust: Ivrea crustal section, Southern

1695 Alps, northwestern Italy and southern Switzerland. Geological Society of

1696 America Bulletin, 103, 236-253.

68 1697 Handy, M.R., Franz L., Heller, F., Janott, B. & Zurbriggen, R. (1999). Multistage

1698 accretion and exhumation of the continental crust (Ivrea crustal section, Italy

1699 and Switzerland). Tectonics, 18, 1154-1177.

1700 Handy, M. R., Schmid, S.M., Bousquet, R., Kissling, E. & Bernoulli, D. (2010).

1701 Reconciling plate-tectonic reconstructions of Alpine Tethys with the

1702 geological–geophysical record of spreading and subduction in the Alps. Earth-

1703 Science Reviews, 102, 121–158.

1704 Henry, C., Michard, A. & Chopin, C. (1993). Geometry and structural evolution of

1705 ultra-high-pressure and high-pressure rocks from the Dora-Maira massif,

1706 Western Alps. Journal of Structural Geology, 15/8, 965-981.

1707 Hoogerduijn Strating (1994). Extensional faulting in an intraoceanic subduction

1708 complex - working hypothesis for the Paleogene of the Alps-Apennine system.

1709 Tectonophysics, 238, 255-273.

1710 Huber, M., Ramsay, J. & Simpson, C. (1980). Deformation of the Maggia and

1711 Antigorio nappes Lepontine Alps. Eclogae Geologicae Helvetiae, 73, 593-606.

1712 Jaillard, E. (1999). The Late Cretaceous – Eocene sedimentation in the internal

1713 Briançonnais units of Vanoise (French Alps): Witnesses of early Alpine

1714 movements. Eclogae geologicae Helvetiae, 92, 211-220.

1715 Jolivet, L., Faccenna, C., Goffé, B., Mattei, M., Rossetti, F., Brunet, C., Storti, F.,

1716 Funiciello, R., Cadet, J.P., d’Agostino, N. & Parra, T. (1998). Mid-crustal

1717 shear zones in postorogenic extension: Example from the northern Tyrrhenian

1718 Sea. Journal of Geophysical Research, 103, 12123-12160.

1719 Keller, L.M. & Schmid S.M. (2001). On the kinematics of shearing near the top of

1720 the Monte Rosa nappe and the nature of the Furgg zone in Val Loranco

1721 (Antrona valley, N. Italy): tectonometamorphic and paleogeographical

69 1722 consequences. Schweizerische Mineralogische und Petrographische

1723 Mitteilungen, 81. 347-367.

1724 Keller, L., Fügenschuh, B., Hess, M., Schneider, B. & Schmid, S.M. (2006). Simplon

1725 fault zone in the western and central Alps: Mechanism of Neogene faulting and

1726 folding revisited. Geology, 34, 317–320. doi: 10.1130/G22256.1.

1727 Kerckhove, C. (1969). La “zone du Flysch” dans les nappes de l’Embrunais-Ubaye

1728 (Alpes occidentales). Géologie Alpine, 45, 5-204.

1729 Kissling, E. (1984). Three-dimensional gravity model of the Northern Ivrea- Verbano

1730 zone. In: Wagner, J.J. & Mueller, S. (Eds.): Matériaux pour la Géologie de la

1731 Suisse Geophysique No 21. Geomagnetic and Gravimetric Studies of the Ivrea

1732 Zone. Kümmerly & Frey, Geographischer Verlag, Bern, 1-61.

1733 Kramer, J. (2002). Structural evolution of the Penninic units in the Monte Rosa

1734 region (Swiss and Italian Alps) (144pp.). PhD Basel University, edoc

1735 University Library of Basel.

1736 Lacombe, O. & Jolivet, L. (2005). Structural and kinematic relationships between

1737 Corsica and the Pyrenees-Provence domain at the time of the Pyrenean

1738 orogeny. Tectonics, 24, TC1003, doi:10.1029/2004TC001673.

1739 Lahondère, D. (1996). Les schistes bleues et les éclogites à lawsonite des unites

1740 continentales et océanique de la Corse alpine: Nouvelles données

1741 pétrologiques et structurales (Corse) (240pp.) France, BRGM.

1742 Lardeaux, J.M., Schwartz, S., Tricart, P., Paul, A., Guillot, S., Béthoux N. &

1743 Masson F. (2006). A crustal-scale cross-section of the south-western Alps

1744 combining geophysical and geological imagery. Terra Nova, 18/6, 412-422.

70 1745 Laubscher, H.P. (1971). The large-scale kinematics of the Western Alps and the

1746 Northern Apennines and its palinspastic implications. American Journal of

1747 Science, 271, 193-226.

1748 Laubscher H. (1991). The arc of the Western Alps today. Eclogae Geologicae

1749 Helvetiae, 84, 631-659.

1750 Laubscher, H., Biella, G.C., Cassinis, R., Gelati, R., Lozej, A., Scarascia, S. &

1751 Tabacco, I. (1992). The collisional knot in Liguria. Geologische Rundschau,

1752 81/2, 275-289.

1753 Lemoine, M. & Tricart, P. (1986). Les Schistes lustrés piémontais des Alpes

1754 Occidentales: Approche stratigraphique, structural et sédimentolologique.

1755 Eclogae Geologicae Helvetiae, 79/2, 271-294.

1756 Leu, W. (1987). Lithostratigraphie und Tektonik der nordpenninischen Sedimente in

1757 der Region Bedretto-Baceno-Visp. Eclogae Geologicae Helvetiae, 79, 769-824.

1758 Levi, N., Ellero, A., Ottria, G. & Pandolfi, L. (2006). Polyorogenic deformation

1759 history recognized at very shallow structural levels: the case of the Antola Unit

1760 (Northern Apennine, Italy). Journal of Structural Geology, 28, 1694-1709.

1761 Lickorish, W.H. & Ford, M. (1998). Evolution of the Digne thrust system, southern

1762 Subalpine chains: kinematics and timing of deformation. In A. Mascle et al.

1763 (Eds.), Cenozoic Foreland Basins of Western Europe (pp. 189–211). Geological

1764 Society London, Special Publications 134.

1765 Lippitsch, R., Kissling, E. & Ansorge, J. (2003). Upper mantle structure beneath the

1766 Alpine orogen from high-resolution teleseismic tomography. Journal of

1767 Geophysical Research, 108, NO B8, 2376, doi:10.1029/2002JB002016.

1768 Loprieno, A., Bousquet, R., Bucher, S., Ceriani, S., Dalla Torre, F.H., Fügenschuh, B.

1769 & Schmid, S.M. (2011). The Valais units in Savoy (France): a key area for

71 1770 understanding the paleogeography and the tectonic evolution of the Western

1771 Alps. International Journal of Earth Sciences, 100, 963–992.

1772 Lustrino, M., Morra, V., Fedele, L. & Franciosi, L. (2009). Beginning of the

1773 Apennine subduction system in central western Mediterranean: Constraints from

1774 Cenozoic ‘‘orogenic’’ magmatic activity of Sardinia, Italy. Tectonics, 28,

1775 TC5016, doi:10.1029/2008TC002419.

1776 Lustrino, M., Fedele, L., Melluso, L., Morra, V., Ronga, F., Geldmacher, J., Duggen,

1777 S., Agostini, S., Cucciniello, C., Frnaciosi, L. & Meisel. T. (2013). Origin and

1778 evolution of Cenozoic magmatism of Sardinia (Italy). A combined isotopic (Sr–

1779 Nd–Pb–O–Hf–Os) and petrological view. Lithos, 180–181, 138–158.

1780 Maffione, M., Speranza, F., Faccenna, C., Cascella, A., Vignaroli, G. & Sagnotti, L.

1781 (2008). A synchronous Alpine and Corsica-Sardinia rotation. Journal of

1782 Geophysical Research, 113, B03104, doi:10.1029/2007JB005214.

1783 Malaroda, R., Carraro, F., Dal Piaz, G.V., Franceschetti, B., Sturani, C. & Zanella, E.

1784 (1970). Carta geologica del Massicio dell_Argentera alla scala 1:50.000 e Note

1785 illustrative. Memorie della Società Geologica Italiana, IX, 557–663.

1786 Malavieille, J., Lacassin, R. & Mattauer, M. (1984). Signification tectonique des

1787 lineations d’allongement dans les Alpes occidentales. Bulletin de la Société

1788 Géologique de France, 26/5, 895-906.

1789 Malavieille, J., Chemenda, A. & Larroque, C. (1998). Evolutionary model for Alpine

1790 Corsica: mechanism for ophiolite emplacement and exhumation of high-

1791 pressure rocks. Terra Nova, 10, 317-322.

1792 Manatschal, G., Sauter D., Karpoff, A.M., Masini E., Mohn G & Lagabrielle,Y.

1793 (2011). The Chenaillet Ophiolite in the French/Italian Alps: An ancient

1794 analogue for an Oceanic Core Complex? Lithos, 124, 169-184.

72 1795 Mancktelow, N.S. (1992). Neogene lateral extension during convergence in the

1796 Central Alps: Evidence from interrelated faulting and backfolding around the

1797 Simplonpass (Switzerland). Tectonophysics, 215, 295–317.

1798 Manzotti, P., Ballèvre, M., Zucaki, M., Robyr, M. & Engi, M. (2014). The

1799 tectonometamorphic evolution of the Sesia–Dent Blanche nappes (internal

1800 Western Alps): review and synthesis. Swiss Journal of Geosciences, 107, 309-

1801 336.

1802 Marchant, R. (1993). The Underground of the Western Alps. Mémoires de Geologie

1803 (Lausanne), 15, 1-137.

1804 Marchant. R. & G.M. Stampfli (1997). Geodynamic evolution of the Tethyan

1805 margins of the Western Alps: geodynamic significance. In: O.A. Pfiffner, P.

1806 Lehner, P. Heitzmann, S. Mueller, A. & Steck (Eds.). Deep structure of the

1807 Swiss Alps - Results from NRP 20 (pp. 139–153). Basel, Birkhäuser.

1808 Marroni, M. & Pandolfi, L. (1996). The deformation history of an accreted ophiolite

1809 sequence: the Internal Liguride units (Northern Apennines, Italy).

1810 Geodinamica Acta, 9, 13-29.

1811 Marroni, M. & Pandolfi, L. (2003). Deformation history of the ophiolite sequence

1812 from the Balagne Nappe, northern Corsica: insights in the tectonic evolution

1813 of Alpine Corsica. Geological Journal, 38, 67-83.

1814 Marroni, M., Molli, G., Pandolfi, L. & Taini A. (1999). Foliated cataclasites at the

1815 base of the Antola unit (Italy): Structural features and geological implications.

1816 Comptes de l’Académie des Sciences Series IIA, Earth and Planetary Science,

1817 329, 135-141.

1818 Marroni, M., Molli, G., Ottria, G. & Pandolfi, L. (2001). Tectono-sedimentary

1819 evolution of the External Liguride units (Northern Apennines, Italy): insights

73 1820 in the pre-collisional history of a fossil ocean-continent transition zone.

1821 Geodinamica Acta, 14, 307-320.

1822 Marroni, M., Cerrina Feroni, A., di Biase, D., Ottria, G., Pandolfi, L. & Taini, A.

1823 (2002a). Polyphase folding at upper structural levels in the Borbera Valley

1824 (northern Apennines, Italy): implications for the tectonic evolution of the

1825 linkage area between Alps and Apennines. Comptes Rendus Geoscience, 334,

1826 565-572.

1827 Marroni, M. Molli, G., Montanini, A., Ottria, G., Pandolfi, L. & Tribuzio (2002b).

1828 The External Ligurian units (Northern Apennine, Italy): From rifting to

1829 convergence of a fossil ocean-continent transition zone. Ofioliti, 27, 119-131.

1830 Marroni, M., Meneghini, F. & Pandolfi, L. (2010). Anatomy of the Ligure-

1831 Piemontese subduction system: evidence from Late Cretaceous–middle

1832 Eocene convergent margin deposits in the Northern Apennines, Italy.

1833 International Geology Review, 52, 1160-1192.

1834 Marshak, S. (1988). Kinematics of orocline and arc formation in thin-skinned

1835 orogens. Tectonics, 7, 73-86.

1836 Marton, E., Zampieri, D., Kazmer M., Dunkl, I. & Frisch, W. (2011). New

1837 Paleocene–Eocene paleomagnetic results from the foreland of the Southern

1838 Alps confirm decoupling of stable Adria from the African plate.

1839 Tectonophysics, 504, 89–99.

1840 Mattauer, M., Faure, M., Malavieille, J. (1981). Transverse lineation and large-scale

1841 structures related to Alpine obduction in Corsica. Journal of Structural

1842 Geology, 3, 401- 409.

1843 Mauffret, A., Contrucci, I. & Brunet, C. (1999). Structural evolution of the Northern

1844 Tyrrhenian Sea from new seismic data. Marine and Petroleum Geology, 16,

74 1845 381-407.

1846 Ménard, G. & Molnar, P. (1991). Budget of crustal shortening and subduction of

1847 continental crust in the Alps. Tectonics, 10, 231-244.

1848 Menkveld. J.W. 1995: Der geologische Bau des Helvetikums der Innerschweiz

1849 (165pp & 141 attachments). Dissertation Universität Bern.

1850 Merle, O. & Brun, J.P. (1984). The curved translation path of the Parpaillon Nappe

1851 (French Alps). Journal of Structural Geology, 6/6, 7111-719.

1852 Michard, A., Avigad, D., Goffé, B. & Chopin, C. (2004). The high-pressure

1853 metamorphic front of the south Western Alps (Ubaye-Maira transect, France).

1854 Schweizerische Mineralogische und Petrographische Mitteilungen, 84, 215–

1855 235.

1856 Michard, A., Negro, F., Saddiqi, O., Bouybaouene, M.L., Chalouan, A., Montigny, R.

1857 Goffé, B. (2006). Pressure-temperature-time constraints on the Maghrebide

1858 mountain building: evidence from the Rif-Betic transect (Marocco, Spain),

1859 Algerian correlations, and geodynamic implications. Comptes Rendus

1860 Geoscience, 338, 92-114.

1861 Milnes, A.G. (1974). Structure of the Pennine Zone (Central Alps): A new working

1862 hypothesis. Geological Society of America Bulletin, 85, 1727-1732.

1863 Mohn, G., Manatschal, G., Beltrando, M., Haupert, I. (2004). The role of rift-

1864 inherited hyper-extension in Alpine-type orogens. Terra Nova, 26, 347-353.

1865 Molli, G. (1996). Pre-orogenic tectonic framework of the northern Apennine

1866 ophiolites. Eclogae Geologicae Helvetiae, 89, 163-180.

1867 Molli, G. (2008). Northern Apennine–Corsica orogenic system: an updated overview.

1868 In: S. Siegesmund et al. (Eds.) Tectonic Aspects of the Alpine-Dinaride-

1869 Carpathian System (pp. 413–442). Geological Society London Special

75 1870 Publications, 298.

1871 Molli, G. & Malavieille, J. (2011): Orogenic processes and the Corsica/Apennines

1872 geodynamic evolution: insights from Taiwan. International Journal of Earth

1873 Sciences, 100, 1207-1224.

1874 Molli, G., Tribuzio, R., Marquer, D. (2006). Deformation and metamorphism at the

1875 eastern border of the Tenda Massif (NE Corsica): a record of subduction and

1876 exhumation of continental crust. Journal of Structural Geology, 28, 1748–

1877 1766.

1878 Molli, G., Crispini, L., Mosca, P., Piana, P. & Federico, L. (2010). Geology of the

1879 Western Alps-Northern Apennine junction area: a regional review. Journal of

1880 the Virtual Explorer 36, paper 10 of the electronic edition ISSN 1441-8142.

1881 Mosca, P., Polino, R., Rogledi, S. & Rossi M. (2010). New data for the kinematic

1882 interpretation of the Alps–Apennines junction (Northwestern Italy). International

1883 Journal of Earth Sciences, 99, 833–849.

1884 Müller, F. (1938). Geologie der Engelhörner, der Aareschlucht und der Kalkkeile bei

1885 Innertkirchen (Berner Oberland). Beiträge zur Geologischen Karte der Schweiz,

1886 NF 74, 71pp.

1887 Mutti, E., Papani, L., Di Biase, D., Davoli, G., Mora, S., Segadelli, S., Tinterri, R.

1888 (1995). Il Bacino Terziario Epimesoalpino e le sue implicazioni sui rapporti tra

1889 Alpi ed Appennino. Memorie di Scienze Geologiche di Padova, 47, 217– 244.

1890 Oberhänsli, R. (editor) (2004). Metamorphic structure of the Alps. Commission for

1891 the Geological Map of the World; Subcommission for Magmatic and

1892 Metamorphic Maps. IUGS and IUGG, Paris. www.ccgm.org.

1893 Orti, L., Pandeli, E. & Principi, G. (2002). New geological data from Gorgona Island

1894 (Northern Tyrrhenian Sea). Ofioliti, 27, 133-144.

76 1895 Padovano, M., Elter, F.M., Pandeli, E. & Franceschelli, M. (2012). The East Variscan

1896 Shear Zone: new insights into its role in the Late Carboniferous collision in

1897 southern Europe. International Geology Review, 54, 957–970.

1898 Pertusati, P.C. & Horrenberger, J.C. (1975). Studio strutturale degli scisti di Val

1899 Lavagna (Unità del Monte Gottero, Apennino Ligure). Bolletino Società

1900 Geologica Italiana, 94, 1375-1436.

1901 Pfiffner, O.A. (2011). Cross sections of the Helvetic nappe system Plates V & IV. In:

1902 Pfiffner et al. 2011: Structural Map of the Helvetic Zone of the Swiss Alps.

1903 Geological Special Map 1:100’000. Federal Office of Topography swisstopo.

1904 Pfiffner, O.A. & Heitzmann, P. (1997). Geological interpretation of the seismic

1905 profiles of the Central Traverse (lines C1, C2 and C3-north). In: Pfiffner A.O. et

1906 al. (Eds.): Deep Structure of the Swiss Alps: Results From NRP 20. Birkhäuser,

1907 Basel, Switzerland, 115-122.

1908 Pfiffner, O.A., Lehner, P., Heitzmann, P. Müller, S. & Steck, A. (1997). Deep

1909 structure of the Alps: Results of NRP20. Birkhäuser Verlag, Basel.

1910 Pfiffner, O.A., Hänni, R., Kammer, A., Kligfield, R., Menkveld, J.W., Ramsay, J.G.,

1911 Schmid, S.M. & Zurbriggen, R. (2010). Geological Special Map 128/4 Sheet 37

1912 Brünigpass. In: Pfiffner et al. 2011: Structural Map of the Helvetic Zone of the

1913 Swiss Alps. Geological Special Map 1:100’000. Federal Office of Topography

1914 swisstopo.

1915 Philippe, Y., Deville, E. & Mascle, A. (1998). Thin-skinned inversion tectonics at

1916 oblique basin margins: example of the western Vercors and Chartreuse

1917 Subalpine massifs (SE France). Geological Society, London, Special

1918 Publications, 134, 239-262.

1919 Philippot, P. (1988). Déformation et éclogitisation progressives d’une croute

77 1920 océanique subductée: Le Monviso, Alpes occidentales. Contraintes

1921 cinématiques durant la collision alpine (269pp). PhD thesis Université de

1922 Montpellier. Documents et Traveaux Centre Géologique et Geophysique de

1923 Montpellier Numéro 19.

1924 Philippot, P. (1990). Opposite vergence of nappes and crustal extension in the

1925 French-Italian Western Alps. Tectonics, 9, 1143-1164.

1926 Piana, F. (2000). Structural Setting of Western Monferrato (Alps-Apennines Junction

1927 Zone, NW Italy). Tectonics, 19/5, 943-960.

1928 Piazza, A., Artoni, A. & Ogata, K. (2016). The Epiligurian wedge-top succession in

1929 the Enza Valley (Northern Apennines): evidence of a syn-depositional

1930 transpressive system. Swiss Journal of Geosciences, 109, 17-36.

1931 Picotti, V. & Pazzaglia F. J. (2008). A new active tectonic model for the construction

1932 of the Northern Apennines mountain front near Bologna (Italy). Journal of

1933 Geophysical Research, 113, B08412, doi:10.1029/2007JB005307.

1934 Pieri, M. & Groppi, G. (1981). Subsurface geological structure of the Po Plain, Italy.

1935 Consiglio Nazionale di Richercha, Progetto Finalizzato Geodinamica, 414,

1936 13pp.

1937 Platt, J.P., Behrmann, J.H., Cunningham, P.C., Dewey, J.F., Helman, M., Parish,

1938 M.G., Shepley, M.G., Wallis, S. & Weston, P.J. (1989). Kinematics of the

1939 Alpine arc and the motion history of Adria. Nature, 337 January 12, 158-161.

1940 Pleuger, J., Roller, S., Walter, J.M., Jansen, E. & Froitzheim, J. (2007). Structural

1941 evolution of the contact between two Penninic nappes (Zermatt-Saas zone and

1942 Combin zone, Western Alps) and implications for the exhumation mechanism

1943 and palaeogeography. International Journal of Earth Sciences, 96, 229-252.

1944 Pleuger, J., Nagel, T.J., Walter, J.M., Jansen, E. & Froitzheim, N. (2008). On the role

78 1945 and importance of orogen-parallel and –perpendicular extension, transcurrent

1946 shearing, and backthrusting in the Monte Rosa nappe and the Southern Steep

1947 Belt of the Alps (Penninic zone, Switzerland and Italy). In S. Siegesmund, B.

1948 Fügenschuh & N. Froitzheim (Eds.), Tectonic Aspects of the Alpine-Dinaride-

1949 Carpathian System (pp. 251–280). Geological Society London, Special

1950 Publications 298.

1951 Pognante, U. (1989). Lawsonite, blueschist and eclogite formation in the southern

1952 Sesia Zone (western Alps, Italy). European Journal Mineralogy, 1, 89–104.

1953 Principi, G., Treves, B. (1984). Il sistema corso-appenninico come prisma

1954 d'accrezione. Riflessi sul problema generale del limite Alpi-Appennini.

1955 Memorie della Società Geologica Italiana, 28, 549–576.

1956 Principi, G., Bortolotti, V., Chiari, M. Cortesogno, L. Gaggero, L., Marcucci, M.,

1957 Saccani, E. & Treves, B. (2004). The pre-orogenic volcano-sedimentary covers

1958 of the Western Tethys Oceanic basin: A review. Ofioliti, 29, 177-211.

1959 Ramsay, J.G. (1980). Shear zone geometry: A review. Journal of Structural Geology,

1960 2, 83-99.

1961 Regis, D., Rubato, D., Darling, J., Cenki-Tok, B., Zucali, M. & Engi, M. (2014).

1962 Multiple metamorphic stages within an eclogite-facies terrane (Sesia Zone,

1963 Western Alps) revealed by Th-U-Pb Petrochronology. Journal of Petrology, 55,

1964 1429-1456.

1965 Ricou, L.E. & Siddans A.W.B. (1986). Collision tectonics in the western Alps In:

1966 M.P. Coward et al. (Eds.), Collision Tectonics (pp.229-244). Geological

1967 Society London Special Publication, 19.

1968 Ritz, J.F. (1992). Tectonique récente et sismotectonique des Alpes du Sud.

1969 Quaternaire, 3, 111-124.

79 1970 Rosenbaum, G. & Lister, G.S. (2004). Neogene and Quaternary rollback evolution of

1971 the Tyrrhenian Sea, the Apennines and the Sicilian Maghrebides. Tectonics,

1972 23, TC1013, 10.1029/2003TC001518.

1973 Rosenberg, C.L. & Kissling, E. (2013). Three-dimensional insight into Central-

1974 Alpine collision: Lower-plate or upper-plate indentation? Geology, 41, 1219–

1975 1222. doi:10.1130/G34584.1.

1976 Rossetti, F., Faccenna, C., Jolivet, L., Funiciello, R. Goffé, B., Tecce, F., Brunet, C.,

1977 Monié, P. & Vidal, O. (2001). Structural signature and exhumation P-T-t path

1978 of the Gorgona island blueschist sequence (Tuscan Archipelago, Italy).

1979 Ofioliti, 26, 175-186.

1980 Rossetti, F., Faccenna, C., Jolivet, L., Goffé, B. & Funiciello, R. (2002). Structural

1981 signature and exhumation P-T-t paths of the blueschist units exposed in the

1982 interior of the Northern Apennine chain, tectonic implications. Bolletino

1983 Società Geologica Italiana Volume Speciale, 1, 829-842.

1984 Roure, F., Polino, R. & Nicolich, R. (1990). Early Neogene deformation beneath the

1985 Po plain: Constraints on post-collisional Alpine evolution, in Deep Structure

1986 of the Alps. Mémoires de la Société Géologique de France, 156, 309-321.

1987 Roure, F., Bergerat, F. Damotte, Mugnier J.-L. & Polino R. (1996). The ECORS-

1988 CROP Alpine seismic traverse. Mémoires Société Géologique de France, 170,

1989 113 pp.

1990 Rubatto, D. & Hermann, J. (2001). Exhumation as fast as subduction? Geology, 29/1,

1991 3-6.

1992 Scharf, A., Handy, M.R., Favaro, S., Schmid, S.M. & Bertrand, A. (2013). Modes of

1993 orogen-parallel stretching and extensional exhumation in response to

1994 microplate indentation and roll-back subduction (Tauern Window, Eastern

80 1995 Alps). International Journal of Earth Sciences, 102, 1627-1654.

1996 Schenk, V. (1990). The exposed crustal cross section of southern Calabria, Italy:

1997 structure and evolution of a segment of Hercynian crust. In: M.H. Salisbury &

1998 D.M. Fountain (Eds.), Exposed Cross-Sections of the Continental Crust (pp.

1999 21-42). Netherlands, Kluwer Academic Publishers.

2000 Skora, S., Mahlen, N.J., Johnson, C.M., Baumgartner, L.P., Lapen, T.J., Beard, B.L.,

2001 Szilvagyi, E.T. (2015). Evidence for protracted prograde metamorphism

2002 followed by rapid exhumation of the Zermatt-Saas Fee ophiolite. Journal

2003 of Metamorphic Geology 33, 711-734.

2004 Schmid, S.M. & Kissling, E. (2000). The arc of the Western Alps in the light of

2005 geophysical data on deep crustal structure. Tectonics, 19, 62-85.

2006 Schmid, S.M., Zingg, A. & Handy, M. (1987). The kinematics of movements along

2007 the Insubric Line and the emplacement of the Ivrea Zone. Tectonophysics, 135,

2008 47-66.

2009 Schmid, S.M., Aebli, H.R., Heller, F. & Zingg, A. (1989). The role of the Periadriatic

2010 Line in the tectonic evolution of the Alps. In: M.P. Coward et al. (Eds.),

2011 Alpine Tectonics (pp.153-171). Geological Society London Special

2012 Publication 45.

2013 Schmid, S.M., Pfiffner, O.A., Froitzheim, N., Schönborn, G., and Kissling, E. (1996).

2014 Geophysical-geological transect and tectonic evolution of the Swiss-Italian Alps.

2015 Tectonics, 15, 1036-1064.

2016 Schmid, S.M., Fügenschuh, B., Kissling, E. and Schuster, R. (2004). Tectonic map

2017 and overall architecture of the Alpine orogen. Eclogae geologicae Helvetiae,

2018 97, 93-117.

2019 Schönborn, G. (1992). Alpine tectonics and kinematic models of the central Southern

81 2020 Alps. Memorie di Scienze Geologiche Padova, 44, 229-393.

2021 Schumacher, M.E. (1997). Geological interpretation of the seismic profiles through

2022 the Southern Alps (lines S1-S7 and C3-south). In: Pfiffner A.O. et al. (Eds.):

2023 Deep Structure of the Swiss Alps: Results From NRP 20. Birkhäuser, Basel,

2024 Switzerland, 101-114.

2025 Schumacher M.E. & Laubscher H. (1996). 3D crustal architecture of the Alps-

2026 Apennines join – a new view on seismic data. Tectonophysics, 260, 349-363.

2027 Schwartz, S., Lardeaux, J. M., Tricart, P., Guillot, S. & Labrin E. (2007).

2028 Diachronous exhumation of HP–LT metamorphic rocks from south-western

2029 Alps: evidence from fission-track analysis. Terra Nova, 19, 133–140.

2030 Séranne, M. (1999). The Gulf of Lion continental margin (NW Mediterranean)

2031 revisited by IBS: an overview. Geological Society, London, Special Publications,

2032 156, 15–36.

2033 Serva, L. (editor) (2005). Gravity map of Italy and surrounding seas. Agenzia per la

2034 Protezione dell’Ambiente e per i Servici Tecnici. Rome

2035 http://www.isprambiente.gov.it.

2036 Seton, M., Müller, R. D., Zahirovic, S., Gaina, C., Torsvik, T., Shephard, G., Talsma,

2037 A., Gurnis, M., Turner, M., Maus, S. & Chandler, M. (2012). Global continental

2038 and ocean basin reconstructions since 200 Ma. Earth-Science Reviews, 113, 212–

2039 270.

2040 Simon-Labric, T., Rolland, Y., Dumont, T., Heymes, T., Authemayou, C., Corsini M.

2041 & Fornari M. (2009). 40Ar / 39Ar dating of Penninic Front tectonic displacement

2042 (W Alps) during the Lower Oligocene (31–34 Ma). Terra Nova, 21, 127–136.

2043 Sommaruga A. (1997). Geology of the central Jura and the Molasse basin: new

2044 insight into an evaporite-based foreland fold and thrust belt. Mémoires Société

82 2045 Sciences Naturelles Neuchâtel, 12, 1-145.

2046 Sommaruga, A., Eichenberger, U. & Mariller, F. (2012). Seismic Atlas of the Swiss

2047 Molasse Basin. Beiträge zur Geologie der Schweiz – Geophysik (text and 24

2048 enclosures). Federal Office of Topography swisstopo.

2049 Spada, M., Bianchi, I., Kissling, E., Piana Agostinetti N. & Wiemer S. (2013).

2050 Combining controlled-source seismology and receiver function information to

2051 derive 3-D Moho topography for Italy. Geophysical Journal International, 194,

2052 1050–1068.

2053 Spakman, W., and M. J. R. Wortel (2004). Tomographic view on western

2054 Mediterranean geodynamics. In W. Cavazza et al (Eds.), The TRANSMED Atlas,

2055 The Mediterranean Region from Crust to Mantle (pp. 31–52), Springer-Verlag,

2056 Berlin - Heidelberg, Germany.

2057 Steck, A. & Hunziker, J.C. (1994). The Tertiary structural and thermal evolution of

2058 the Central Alps - Compressional and extensional structures in an orogenic belt.

2059 Tectonophysics, 238, 229–254.

2060 Steck, A., Epard, J.-L., Escher, A., Lehner, P., Marchant, R. & Masson, H. (1997).

2061 Geological interpretation of the seismic profiles through Western Switzerland:

2062 Rawil (W1), Val d’Anniviers (W2), Mattertal (W3), Zmutt-Zermatt-Findelen

2063 (W4) and Val de Bagnes (W5). In: Pfiffner A.O. et al. (Eds.): Deep Structure of

2064 the Swiss Alps: Results From NRP 20 (pp. 123–137). Birkhäuser, Basel,

2065 Switzerland.

2066 Steck, A., Della Torre, F., Keller, F., Pfeifer, H.R., Hunziker, J. & Masson, H.

2067 (2013). Tectonics of the Lepontine Alps: ductile thrusting and folding in the

2068 deepest tectonic levels of the Central Alps. Swiss Journal Geosciences, 106,

2069 427–450.

83 2070 Steck, A., Masson, H. & Robyr, M. (2015). Tectonics of the Monte Rosa and

2071 surrounding nappes (Switzerland and Italy): Tertiary phases of subduction,

2072 thrusting and folding in the . Swiss Journal of Geosciences, 108, 3-

2073 34.

2074 Stipp, M., Stünitz, H., Heilbronner, R. & Schmid, S.M. (2002). Dynamic

2075 recrystallization of quartz: correlation between natural and experimental

2076 conditions. In: De Meer et al. (Eds), Deformation Mechanisms, Rheology and

2077 Tectonics: Current Status and Future Perspectives (pp. 171-190). Geological

2078 Society London Special Publication, 200.

2079 Thouvenot, F., Senechal, G., Truffert, C. & Guellec S. (1996). Comparison between

2080 two techniques of line-drawing migration (ray tracing and common tangent

2081 method). Mémoires de la Société Géologique de France, 170, 53-59.

2082 Tricart, P. (1986). Le chevauchement de la zone briançonnaise au Sud-Est du

2083 Pelvoux: clé des rapports zone externe – zones internes dans les Alpes

2084 occidentales. Bulletin de la Société géologique de France, 11, 233-244.

2085 Trullenque, G. (2005). Tectonic and microfabric studies along the Penninic Front

2086 between Pelvoux and Argentera massifs (Western Alps) (293pp). University of

2087 Basel, PhD thesis. Pdf available at http://edoc.unibas.ch/353/

2088 Trullenque, G., Kunze, K., Heilbronner, R. Stünitz, H. Schmid, S.M. (2006).

2089 Microfabrics of calcite ultramylonites as records of coaxial and non-coaxial

2090 deformation kinematics: Examples from the Rocher de l'Yret shear zone

2091 (Western Alps). Tectonophysics, 424, 69–97.

2092 Van der Voo, R. (2004). Presidential Address: Paleomagnetism, oroclines, and

2093 growth of the continental crust. Geological Society of America Today, 14/12, 4-9.

2094 van Hinsbergen, D.J.J. & Schmid, S.M. (2012). Map view restoration of Aegean–

84 2095 West Anatolian accretion and extension since the Eocene. Tectonics, 31,

2096 TC5005, doi:10.1029/2012TC003132.

2097 van Hinsbergen, D.J.J., Vissers, R.L.M. & Spakman W. (2014a). Origin and

2098 consequences of western Mediterranean subduction, rollback, and slab

2099 segmentation. Tectonics, 33, 393-419.

2100 Van Hinsbergen, D.J.J., Nebsink, M., Langereis, C.G., Maffione, M., Spalluto, L.,

2101 Tropeano, M. & Sabato, L. (2014b). Did Adria rotate relative to Africa? Solid

2102 Earth, 5, 611-629.

2103 Van Wamel, W.A. (1987). On the tectonics of the Ligurian Apennines (northern

2104 Italy). Tectonophysics, 142, 87-98.

2105 Vialon, P. Rochette, P. & Ménard, G. (1989). Indentation and rotation in the western

2106 Alpine arc. In: M.P. Coward , D. Dietrich & R.G. Park (Eds.), Alpine tectonics

2107 (pp. 329-338). Geological Society of London Special Publications, 45.

2108 Villa, I.M., Bucher, S., Bousquet, R., Kleinhanns, I.C. & Schmid, S.M. (2014).

2109 Dating polygenic metamorphic assemblages along a transect across the Western

2110 Alps. Journal of Petrology, 55, 803-830.

2111 Vitale, S. & Ciarcia, S. (2013). Tectono-stratigraphic and kinematic evolution of the

2112 southern Apennines/Calabria–Peloritani Terrane system (Italy).

2113 Tectonophysics, 583, 164–182.

2114 Vitale Brovarone, A. & Herwartz, D. (2013). Timing of HP metamorphism in the

2115 Schistes Lustrés of Alpine Corsica: New Lu–Hf garnet and lawsonite ages.

2116 Lithos, 172–173, 175–191.

2117 von Blanckenburg, F. (1992). Combined high-precision chronometry and

2118 geochemical tracing using accessory minerals: Applied to the Central-Alpine

2119 Bergell intrusion (central Europe). Chemical Geology, 100, 19-40.

85 2120 Wagner, M., Kissling, E. & Husen, S. (2012). Combining controlled-source

2121 seismology and local earthquake tomography to derive a 3-D crustal model of

2122 the western Alpine region. Geophysical Journal International, 191, 789-802.

2123 Walcott, C.R. & White, S.H. (1998). Constraints on the kinematics of post-orogenic

2124 extension imposed by stretching lineations in the Aegean region:

2125 Tectonophysics, 298, 155–175.

2126 Weber, S., Sandmann, S., Miladinova, I., Fonseca, R.O.C., Froitzheim, N., Münker, C.

2127 & Bucher, K. (2015). Dating the initiation of Piemont-Liguria Ocean

2128 subduction: Lu-Hf garnet chronometry of eclogites from the Theodul

2129 Glacier Unit (Zermatt-Saas Zone, Switzerland). Swiss Journal of Geosciences,

2130 108, 183-199.

2131 Weissert, H.J. & Bemoulli, D. (1985). A transform margin in the Mesozoic Tethys:

2132 Evidence from the Swiss Alps. Geologische Rundschau, 71, 665-679.

2133 Worthington, J.R., Hacker, B.R. & Zandt, G. (2013). Distinguishing eclogite from

2134 peridotite: EBSD-based calculations of seismic velocities. Geophysical Journal

2135 International, 193, 489-505. doi: 10.1093/gji/ggt004.

2136 Ye, S., Ansorge, J. Kissling, E. & Mueller, S. (1995). Crustal structure beneath the

2137 eastern Swiss Alps derived from seismic refraction data. Tectonophysics, 242,

2138 199-221.

2139 Zhao, L., Paul, A., Guillot, S., Solarino, S., Malusà, M.G., Zheng, T. , Aubert, C.,

2140 Salimbeni, S., Dumont, T., Schwartz, S., Zhu, R. & Wang, Q. (2015). First

2141 seismic evidence for continental subduction beneath the Western Alps.

2142 Geology, 43, 815-818.

2143 Zingg, A., Handy, M.R., Hunziker, J.C. & Schmid, S.M. (1990).

2144 Tectonometamorphic history of the Ivrea Zone and its relationship to the

86 2145 evolution of the Southern Alps. Tectonophysics, 182, 169-192.

2146

2147 Figure legends

2148

2149 Fig. 1:

2150 Tectonic map of the arc of the Western Alps after Schmid et al. (2008) and Bousquet

2151 et al. (2012b) with indication of the traces of the profiles depicted in Fig. 2. The

2152 selected faults indicated in red are: RSL: Rhone-Simplon Line; PL: Periadriatic Line;

2153 PF: Penninic Front; ACL: Argentera-Cuneo Line; AFSV: front of the metamorphic

2154 core of the Alps and Sestri-Voltaggio line; VVL: Villalvernia-Varzi line.

2155 Abbreviations in black are: LA: Ligurian Alps; TH: Torino hills; MH: Monferrato

2156 hills; GB: Giudicarie transpressive belt.

2157

2158 Fig. 2:

2159 Five selected geological-tectonic transects across the Western Alps. The sections are

2160 superimposed with and partly based on vertical cross sections across the final P-wave

2161 velocity model of Diehl et al. (2009) indicated by red dashed lines. The Moho (blue

2162 line) is taken from a combination of controlled source seismology, local earthquake

2163 tomography and receiver function analysis (Wagner et al. 2012; Spada et al. 2013).

2164 See text for data sources and further discussion and Fig. 1 for the traces of the

2165 profiles. a: Argentera transect; b: Pelvoux transect; c: ECORS-CROP transect; d:

2166 NFP 20-West transect; e: Ticino transect.

2167

2168 Fig. 3:

2169 Horizontal section across 3-D P-wave tomographic model of Diehl et al. (2009) at a

87 2170 depth of 20km. Yellow triangles indicate seismic stations used for tomographic

2171 inversion, crosses indicate inversion grid nodes. Note that the velocity structure is

2172 shown as percentage change relative to the 1-D initial reference model rather than in

2173 terms the P-wave velocities shown in Fig. 2. Bold black contours include well-

2174 resolved regions covering most of the area depicted. Compare Fig. 1 for major

2175 tectonic units outlined: RSL: Rhone-Simplon Line; PL: Periadriatic Line; PF:

2176 Penninic Front; ACL: Argentera-Cuneo Line. Comparison with Fig. 2 shows that the

2177 entire area with P-wave anomaly >10%, as well as a part of the area with P-wave

2178 anomaly >5% between north of Locarno and Cuneo is within the Ivrea mantle wedge.

2179

2180 Fig. 4:

2181 Three generations of stretching lineations and associated senses of shear around the

2182 arc of the Western Alps. Red arrow: stage 1 top N transport; blue arrows: stage 2 top

2183 WNW transport; green: stage 3 top SW transport; arrow pairs reflect the variation of

2184 the measured lineations; see text for further details. Numbered areas of measurements

2185 and literature sources are as follows: area 1, Valaisan and Penninic Front (Loprieno et

2186 al. 2011); area 2, Zone Houillère of the Briançonnais Unit (Bucher et al. 2004); area

2187 3, Subbriançonnais and Penninic Front (Ceriani et al. 2001; Ceriani & Schmid 2004);

2188 area 4, Penninic Front (top NW) and basal decollement of Cenozoic flysch of the

2189 Dauphinois (top SW) (Bürgisser & Ford 1998; Trullenque 2005; Trullenque et al.

2190 2006); area 5, eclogite facies lineations (top N) and greenschist facies lineations (top

2191 SW) in M. Viso ophiolites (Philippot 1988, 1990); area 6, Latest Eocene “schistes è

2192 blocs” at the base of the Embrunais-Ubaye nappes (Trullenque 2005); area 7,

2193 Penninic Front N of the Argentera massif (Trullenque 2005); area 7, thrust transport

2194 constraints from the Dauphinois of the Digne thrust system (Lickorish & Ford 1998,

88 2195 Ford et al. 2006).

2196

2197 Fig. 5:

2198 Kinematic restoration of the Alps-Apennines orogenic system using Gplates

2199 reconstruction software (e.g. van Hinsbergen et al. 2014a) holding Europe fixed; for

2200 easier orientation all time slices also show the present day position of the external

2201 massifs and Tauern window in the Alps, as well as the present-day position of

2202 northern Adria. Arrowheads indicate displacement from a given time slice to the next

2203 one, they are anchored at the position of a chosen a set of key locations at the chosen

2204 time slice. Red lines: coast lines and outlines of external massifs and Tauern Window

2205 in the Alps; blue lines: outlines of the Argentera, Pelvoux and Tambo-Suretta nappes;

2206 green area: horizontal section across the Ivrea mantle slice taken from Fig. 3. The

2207 present day position of numbered locations shown is as follows: 1: Sardinia; 2:

2208 Corsica; 3: Dora Maira massif (Briançonnais); 4: Gran Paradiso massif

2209 (Briançonnais); 5: Tambo-Suretta nappes (Briançonnais); 6: Oberammergau (front

2210 Austroalpine nappes); 7: Dimaro (eastern end Tonale Line); 8: Mauls (western end

2211 Pustertal Line; 9: Slovenske Konjice (eastern end Gailtal line); 10: Cuneo (S end of

2212 Ivrea wedge); 11: Brione (N end of Ivrea wedge); 12: Pavia (Adria microplate); 13:

2213 Monti Lessini (Adria microplate); 14: Lignano (Adria microplate); 15: W of Ancona

2214 (frontal thrust of Umbria-Marche Apennines); 16: S of Parma (frontal thrust of

2215 external Ligurides, Apennines); 17: offshore Baleares at 25 Ma ago (N front of

2216 oceanic Adria microplate = Piedmont-Liguria lower plate).

2217

2218 Fig. 6:

2219 Interpretative map sketch of the Alps-Apennines transition; see text for further

89 2220 discussion.

90 Cenozoic basins Tectonic map of the Western Alps Cenozoic in general 6° 7° 8° 9° 10° 11° Cenozoic of Piedmontese basin

Basel Lower Plate Units of the Alps European continental margin Zürich St. Gallen Dijon Cenozoic flysch Besançon (Dauphinois and Helvetic Nappes) 4 NFP 20 West 5 Ticino Innsbruck European Mesozoic undeformed 47° European Mesozoic deformed 47° Bern Allochthonous Mesozoic & Cenozoic (Helvetic Nappes)

Chur European basement (External Massifs)

European upper crust of the Subpenninic Units

High-pressure Subpenninic Units 3 Ecors CROP Bolzano Valaisan partly oceanic domain Sion RSL Valaisan Genève PL 46° Locarno Briançonnais micro-continent GB 46° Sub-Briançonnais

PF Briançonnais Aosta High-pressure Briançonnais

Pre-Piemontais Acceglio

Moûtiers Ivrea Sesia-Dent Blanche Extensional Allochthons Milano Verona Sesia, Dent Blanche, Margna, Sella Units Grenoble Piedmont-Liguria Ocean 45° 2 Pelvoux Torino TH Lower Plate Piedmont-Liguria MH 45° VVL Upper Plate Units of the Alps Briançon AFSV Upper Plate Piedmont-Liguria and distal continental margin of Adria

Adria Permo-Mesozoic Gap PF Bologna Adria upper crust Cuneo Argentera Genova Adria lower crust ACL LA Austroalpine Permo-Mesozoic 44° Digne Austroalpine basement 44° 0 20 100 km Cenozoic plutons 1 Argentera Apennines Nice Apenninic Units in general 6° 7° 8° 9° 10° 11° a. a. Argentera transect Piedmont-Liguria Upper Plate Oligocene & Miocene & Sub-Briançonnais Acceglio Piedmont-Liguria Lower Plate Miocene Pliocene Black line = Late Jurassic = Pre-Piemontais Canavese & Piedmont- Eclogites km European Mesozoic Jura Liguria Upper Plate Pliocene & 10 Dauphinois & Helvetic n. Pleistocene 10

0 Argentera 0 High-pressure Adria Mesozoic 5.5 Briançonnais Apenninic 10 Units 10 Adria upper crust 5.5 Sub-Briançonnais 6.0 European upper crust Briançonnais 20 6.0 Adria lower crust 20 6.5 6.5 6.5 7.0 30 7.0 30 7.0 European lower crust 7.5 Adria mantle 40 7.5 40

European mantle 50 8.0 50 8.0 km / s 60 60

8.0

b. b. Pelvoux transect European Mesozoic Jura Dauphinois & Helvetic n. Briançonnais Acceglio Piedmont-Liguria Sub-Briançonnais Adria Mesozoic km = Pre-Piemontais Piedmont-Liguria Upper Plate km Black band = Lower Plate Eclogites 10 Late Jurassic Miocene & Oligocene & 10 Pliocene Miocene

0 0 Belledonne Grandes Pelvoux 5.0 Rousses Apenninic Units 10 European upper crust High-pressure 5.5 10 East Variscan Briançonnais Adria upper crust 6.0 Shear Zone Valaisan 6.0 Adria lower crust 20 20 6.5 6.5 7.0 30 7.0 30 European lower crust 7.5 7.5 40 40 Adria mantle European mantle 50 50 8.0 km / s

60 60

c. c. ECORS-CROP transect Pliocene & Pleistocene European Mesozoic Jura Briançonnais Piedmont-Liguria Sesia Miocene Dauphinois & Helvetic n. Lower Plate D. Blanche km Black line = Zone Eclogites Paleogene km Valaisan 10 Oligocene & Late Jurassic 10 Miocene Canavese & Piedmont- Liguria Upper Plate

0 High-pressure 0 Briançonnais 5.0 Adria Mesozoic Belledonne 5.5 10 6.0 M. Blanc Adria upper crust 10 East Variscan 7.0 6.0 20 6.5 shear Zone Adria lower crust 20 6.5 European upper crust 7.0 6.5 30 30 7.5 7.5 7.5 40 Adria mantle 40 8.0 km / s European lower crust 7.25 50 European mantle 50

60 60 8.0 d. d. NFP 20-West transect

European Mesozoic Jura Canavese & Piedmont- Sesia Miocene Dauphinois & Helvetic n. Black line = Briançonnais Liguria Upper Plate D. Blanche Piedmont-Liguria Late Jurassic High-pressure Sesia Adria Mesozoic Black line = Oligocene & Zone Briançonnais Lower Plate D. Blanche km Valaisan Pliocene km Late Jurassic Miocene Zone 10 Valaisan & Pleistocene 10

0 0 Eclogites 5.5 Adria upper crust 6.0 M. Blanc 6.0 10 European upper crust 10 6.0 6.5 Adria lower crust 20 20 7.0 30 30 6.5 European lower crust 6.5 7.5 7.0 Adria mantle 7.0 40 European lower crust 40 7.5 European mantle 7.5 50 8.0 km / s 50 8.5 8.0 European mantle 60 60 8.0 km / s 7.5

e. e. Ticino transect High-pressure Sesia European Piedmont-Liguria D. Blanche Black band = Valaisan basement Zone km European Mesozoic Jura Lower Plate Pliocene & km Late Jurassic Briançonnais 10 Dauphinois & Helvetic n. Pleistocene 10 Eclogites Paleogene

Gotthard 0 5.5 Aar 0 Adria Mesozoic 6.0 10 10

6.0 20 Adria upper crust 6.5 20 6.5 European upper crust 30 Adria lower crust 30 7.0 European lower crust 7.25 7.0

40 7.5 7.5 40 Adria mantle 8.0 8.0 50 European mantle 50

60 8.5 km / s 60

Europe Moho documented P-wave velocity model by CSS, LET, and RF of Diehl et al. (2009) Cenozoic Cenozoic flysch (Dauphinois and Helvetic nappes) Piedmont-Liguria Ocean Adria Pliocene & Valaisan Piedmont-Liguria Paleogene European Mesozoic Jura and Miocene Pleistocene Dauphinois & Helvetic n. Upper Plate Sub-Briançonnais Canavese & Montgenevre Adria Mesozoic European upper crust nappe superieure High-pressure European Briançonnais Piedmont-Liguria Adria upper crust Apenninic Units Adula, Cima Lunga Lower Plate High-pressure Adria lower crust European lower crust Briançonnais Extensional Allochthons Acceglio European mantle = Pre-Piemontais Sesia, D. Blanche unit Adria mantle 6˚E 7˚E 8˚E 9˚E 10˚E

z = 0020 km, ref vp: 6.27 km/s

NFP-20RSL WEST Locarno

Genève Ticino PL 46˚N ECORS-CROP

PF Milano

Torino Pelvoux 45˚N Briançon

Cuneo Genova ACL

Argentera 44˚N

−20 −15 −10 −5 0510 15 20 P-wave velocity change relative to 1D-reference model (%) a. 6 b. Today 25 Ma arrow head 0 Ma 5 8 7 9 11 14 4 13 12 3 Piedmont - Liguria upper plate 16 0 100 200 km 10 15

Ligurides

2

remaining oceanic lithosphere of Adria

1 0 100 200 km 17

c. 35 Ma d. 45 Ma arrow head 25 Ma arrow head 35 Ma

0 100 200 km 0 100 200 km

Piedmont- Liguria upper plate

oceanic lithosphere of Adria ALKAPECAoceanic lithosphere of Adria

e. f. 6 55 Ma Eastern Alps and 1: Sardinia Periadriatic Line arrow head 45 Ma 2: Corsica 8 5 9 3: Dora Maira 11 7 4: Gran Paradiso 5: Tambo-Suretta Briançonnais 13 6: Oberammergau 4 14 6 12 fixed Europe 7: E end of Tonale Line 6 3 8: W end Pustertal Line 11 0 100 200 km Australpine nappes Mauls 16 10 9 9: E end Gailtal Line 10 8 5 Slov. Konjice 5 15 10: SW Ivrea Valais ocean 9 8 Continental Adria 11: NE Ivrea 7 4 12: Adria Plate Pavia 11 4 13: Adria Plate Ivrea mantle wedge 3 Monti Lessini 3 Briançonnais 14: Adria Plate 2 Ligniano 12 13 15: front Umbria- 2 10 13 Iberia 7 14 Marche Apennine 12 1 Corsica- 14 W of Ancona 2 1 Sardinia 16: external Ligurides N. Apennine Adria Plate S of Parma 16 17: N front Adria Plate 16 17 near Baleares 1 oceanic litho- 15 0 100 200 km 15 sphere of Adria 17 Piedmont - Liguria lower plate 17 Alps Innsbruck Lower Plate units of the Alps Luzern Bern undeformed or little deformed European foreland eroded part of Austroalpine Lower Plate units of the Alps nappes Chur Lower Plate units of the Alps strongly affected by Miocene-age oroclinal bending and Apennines orogeny Bolzano Sion Upper Plate units of the Alps Genève Ligurian ocean - Adria continent transition units (Canavese, “Nappes Supérieures” of Préalpes and Locarno Embrunais-Ubaye) and Chenaillet ophiolite Upper Plate units of the ocean-continent transition substantially affected by Miocene oroclinal bending and Apennines orogeny (Torino hills, Helminthoid flysch of 36 Antola nappe and Alpes Maritimes, “Nappes supérieures” 20° of Corsica) Milano Austroalpine nappes (Adria-derived continental upper Verona crustal nappe stack formed during Cretaceous orogeny 40 and displaced over the Alpine Lower Plate units in Torino the Eocene) 44 Briançon Southern Alps retro-belt derived from and thrust onto continental Adria 48 48 Undeformed continental Adria 52 Apennines 47° Bologna Oceanic upper Plate units of the Alps intensely deformed during Alpine orogeny and reworked Genova during Apennines orogeny 44 117° Internal Ligurides (Ligurian Ocean) 40 Oceanic and ocean-continent transition units little or not affected by Alpine orogeny 36 Nice Firenze Tuscan Ligurides (Ligurian Ocean)

Gorgona External Ligurides and Subligurian units island (Ligurian Ocean - Adria continent transition) Units derived from the Adria continent and the Apennines foredeep orogenic fronts of Alps and Apennines Tuscan nappe (distal Adria continental margin) topped by Chattian-Aquitanian Macigno turbidites upper plate of Apennines Cervarola thrust sheets made up of Aquitanian to Penninic Front Early Langhian turbidites contours of European and Adriatic Moho after Spada et al. (2013) Umbria-Marche thrust sheets made up of Langhian to Tortonian Marnoso Arenacia foredeep turbidites outlines of Ivrea mantle wedge at a 36 depth of 20 km (see Fig. 3) 50° post-Messinian frontal imbricates approximate former western limit of the 100 km eroded part of the Austroalpine nappes 10° E 42° N Tertiary Piedmont basin and Epiligurian basins