3D structure and tectonic evolution of the Sorbas basin

0 downloads 0 Views 13MB Size Report
Margin, Earth and Planetary Science Letters, 241(3-4), 734-749. .... Miller, K.G., Kominz, M.A., Browning, J., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman,.
1

Tectonic inversion of an asymmetric graben: insights from a

2

combined field and gravity survey in the Sorbas basin

3

Damien Do Couto1,2,3, Charles Gumiaux4,5,6, Romain Augier4,5,6, Noëmie Lebret4,5,6, Nicolas

4

Folcher7, Gwenaël Jouannic8, Laurent Jolivet4,5,6, Jean-Pierre Suc2,3, Christian Gorini2,3

5

1

6

([email protected])

7

2

8

3

9

[email protected])

Total S.A., 2 Place De La Coupole, 92078 Paris La Défense Cedex, France

UPMC Université Paris 6, UMR 7193, ISTEP, 75005, Paris, France CNRS,

UMR

7193,

ISTEP,

F-75005,

Paris,

France

10

4

11

orleans.fr,

12

[email protected])

13

5

CNRS/INSU, ISTO, UMR 7327, 45071 Orléans, France

14

6

BRGM, ISTO, UMR 7327, BP 36009, 45060 Orléans, France

15

7

16

Nouvelle-Calédonie ([email protected])

17

8

18

Haie, 54510 Tomblaine, France ([email protected])

([email protected],

Université d’Orléans, ISTO, UMR 7327, 45071, Orléans, France ([email protected]@univ-orleans.fr,

[email protected],

PPME (EA n° 3325), Université de la Nouvelle-Calédonie, BP R4 98851 Nouméa Cedex,

CETE de l’Est Laboratoire Régional des Ponts et Chaussées de Nancy, 71 Rue de la Grande

19 20

Corresponding author: Damien Do Couto, Institut des Sciences de la Terre de Paris, CNRS-

21

UMR 7193-Université Pierre et Marie Curie, 4 Place Jussieu-75005 Paris, France.

22

([email protected])

23

24

Key points

25



The Sorbas Basin shows a strong asymmetric structure

26



The basin has been initiated during an extensional regime

27



It then experienced a severe tectonic inversion from ca. 8 Ma characterized by a ca. N-S

28

shortening

29 30

Abstract

31

[1] The formation of sedimentary basins in the Alboran domain is associated with the

32

exhumation of metamorphic core complexes over a ca. 15 Ma period through a transition from

33

regional late-orogenic extension to compression. An integrated study coupling field analysis and

34

gravity data acquisition was performed in the Sorbas basin in the southeastern Betic Cordillera.

35

Detailed field observations revealed for the first time that extensional tectonics occurred before

36

shortening in this basin. Two extensional events were recorded with NW-SE to N-S and NE-SW

37

kinematics respectively; the first of which being likely responsible for the basin initiation.

38

Tectonic inversion of the basin then occurred around 8 Ma in an overall ca. N-S shortening

39

context. 2D-gravity sections reveal that the basin acted as an active depocenter as the basin floor

40

locally exceeding 2km-depth is characterized by a marked asymmetric architecture. Based on

41

this integrated study, we explore a new evolutionary scenario which can be used as a basis for

42

interpretations of the Neogene tectonic history of the southeastern Betics.

43 44

1. Introduction

45

[2] During the Neogene, the geodynamics of the Mediterranean region was marked by the

46

development of backarc basins, from the Alboran Sea in the west all the way to the Aegean Sea

47

in the east, initiated after a major change in the subduction regime, during the Oligocene

48

[Rehault et al., 1984; Faccenna et al., 1997; Jolivet and Faccenna, 2000]. The Pannonian Basin,

49

Aegean Sea, Liguro-Provençal Basin, Tyrrhenian Sea and Alboran Sea formed above retreating

50

slabs. The present-day complex geometry of the subduction zones and backarc basins results

51

from the initial geometry of the African margin [Frizon de Lamotte et al., 2011] and progressive

52

slab tearing and detachments [Wortel and Spakman, 2000] associated with a complex 3D mantle

53

convection pattern [Faccenna and Becker, 2010]. At the western end of the Mediterranean

54

region, the Alboran domain has recorded this regional episode of backarc extension and slab

55

retreat [Faccenna et al., 2004; Spakman and wortel, 2004] until ~7-8 Ma before the resumption

56

of a compressional regime due to Africa-Eurasia convergence [Jolivet et al., 2008]. Offshore and

57

onshore sedimentary basins have recorded this evolution and the respective contributions of early

58

extension and late compression in the final architecture of these basins is still debated, especially

59

in the Huércal-Overa and Sorbas basins [Weijermars et al., 1985; Ott d’Estevou and Montenat,

60

1990; Augier et al., 2005a; Meijninger and Vissers, 2006; Pedrera et al., 2010]. This study aims

61

at deciphering the respective contribution of extension and subsequent compression in the

62

shaping of the Sorbas basin.

63

[3] The Sorbas basin, located in the internal zones of the Betic Cordillera (southern Spain), is a

64

key peripheral basin where upper-Serravallian to Pliocene sediments have recorded interactions

65

of both tectonic and eustatic changes of the southwestern Mediterranean region. Several

66

scenarios of the Neogene evolution of the whole Alboran region were based upon the

67

stratigraphy and structure of this basin. However, different tectonic settings, from strike-slip to

68

pure extension, have been proposed for the initiation of the basin in the Serravallian-lower

69

Tortonian. Since the upper Tortonian, the Sorbas basin has been subjected to NW-SE shortening

70

that has shaped most of its current geometry [Weijermars et al., 1985]. The debate on the

71

tectonic origin of the Sorbas basin is mainly based on contrasting interpretations of the structural

72

relationships between sedimentary basins and metamorphic domes of the adjacent “sierras” (i.e.

73

Sierra de los Filabres and the Sierra Alhamilla) [Martínez-Martínez and Azañón, 1997; Martínez-

74

Martínez et al., 2002, 2004; Behr and Platt, 2012]. While the Huércal-Overa basin shows rather

75

well preserved shape and distribution of depocenters linked to the pre-8 Ma extension

76

[Meijninger and Vissers, 2006; Pedrera et al., 2010], late shortening had a major influence on the

77

present-day geometry of the Sorbas basin [Weijermars et al., 1985; Ott d’Estevou and Montenat,

78

1990]. The effect of the early extension has not already been discussed and a better description

79

of the deep structure of the basin is needed to access its early tectonic history. In order to better

80

constrain the geometry of the basin during its evolution, a gravity survey has been carried out

81

through the basin together with new field structural analyses. From these new structural and

82

geometrical constraints, a new tectonic evolutionary scenario is proposed for the Sorbas Basin.

83 84

2. Geological setting

85

2.1. The southeastern Betic Cordillera

86

[4] The Betic Cordillera forms the northern branch of the Betic-Rif orogenic belt that results

87

from the combination of a continental collision between the Alboran domain and the Iberian-

88

Moroccan passive margins, during the Miocene [Crespo-Blanc and Frizon de Lamotte, 2006],

89

and a strong westward movement of the Alboran domain [Lonergan and White, 1997; Faccenna

90

et al., 2004]. The former south Iberian Mesozoic-Cenozoic passive margin currently forms the

91

External Zones of the Betics and corresponds to a fold-and-thrust belt [Platt et al., 2003]. To the

92

South, the finite structure of the Internal Zones results from the stacking of metamorphic units

93

and their subsequent exhumation in the core of metamorphic domes [Martínez-Martínez et al.,

94

2002; Augier et al., 2005a, 2005b]. From bottom to top, the alpine metamorphic basement is

95

composed of the Nevado-Filabride (NF) and the Alpujarride (Alp) complexes, overlain by a third

96

one, the Malaguide (Mal) complex [Torres-Roldán, 1979]. These three complexes are presently

97

separated by crustal-scale extensional shear zones that played a major role during the

98

exhumation of the deeper complexes [García-Dueñas et al., 1992; Lonergan and Platt, 1995;

99

Crespo-Blanc, 1995; Platt et al., 2005; Augier et al., 2005a]. Exhumation of the Alpujarride

100

complex occurred during the Early Miocene (22 to 18 Ma) in a N-S to NNE-SSW extensional

101

setting [Monié et al., 1994; Crespo-Blanc et al., 1994; Crespo-Blanc, 1995; Platt et al., 2005].

102

Timing of peak metamorphic conditions and the inception of exhumation of the NF complex is

103

still debated [López-Sánchez Vizcaino et al., 2001; de Jong, 2003; Augier et al., 2005b; Platt et

104

al., 2006]. Conversely, timing of the final exhumation stages in the greenschists facies conditions

105

spans over the Middle and the Late Miocene (from 15 to 9 Ma) [de Jong, 1991; Monié and

106

Chopin, 1991; Johnson et al. 1997; Augier et al., 2005b; Platt et al., 2005; Vázquez et al., 2011]

107

through an E-W regional-scale extension [Jabaloy et al., 1992]. Final exhumation stages in the

108

brittle regime at ca. 13-12 Ma was accompanied by the development of extensional sedimentary

109

basins such as the Huércal-Overa basin [Vissers et al., 1995; Augier, 2004; Meininger and

110

Vissers, 2006; Augier et al., 2013].

111

[5] The Sorbas basin belongs to a mosaic of E-W oriented Late Neogene intramontane basins

112

with the Huércal-Overa basin and the Almanzora Corridor, to the North, the Alpujarran Corridor

113

to the West and the Vera basin to the East [Ott d’Estevou and Montenat, 1990; Sanz de Galdeano

114

and Vera, 1992; Iribarren et al., 2009; Pedrera et al., 2010]. It is bounded to the North by the

115

Sierra de los Filabres and to the South by the Sierra Alhamilla (Figure 1), two E-W elongated

116

metamorphic domes [Platt and Vissers, 1989; Vissers et al., 1995; Martínez-Martínez and

117

Azañón, 1997] formed by the exhumation of the NF complex and amplified by later folding

118

[Weijermars et al., 1985].

119 120

2.2. Tectonic and geodynamic framework

121

[6] The extensional regime which affected the Betic Cordillera and, in a more regional extent,

122

the entire Alboran domain started from the Late Oligocene – Early Miocene (see Vergés and

123

Fernàndez, 2012 for a review of tectonic evolution) and is widely documented both offshore

124

[Watts et al., 1993; Comas et al., 1999] or onshore [García-Dueñas et al., 1992; Galindo-

125

Zaldívar et al., 2003], especially in the Huércal-Overa basin [Briend et al., 1990; Augier et al.,

126

2005a; Meijninger and Vissers, 2006; Pedrera et al., 2009, 2010, 2012; Augier et al., 2013].

127

This latter displays an overall half-graben geometry accompanied by meso- to small-scale syn-

128

sedimentary normal faults. Its formation and its development have been associated with the

129

exhumation of the NF metamorphic complex in the Sierra de Los Filabres metamorphic dome

130

[Augier et al., 2005a; Meijninger and Vissers, 2006; Augier et al., 2013] (see Figure 1).

131

According to Augier et al. [2013], the half-graben geometry results from the development of

132

north dipping normal faults that are probably rooted in the detachment separating the NF from

133

the Alp complexes, called the Filabres Shear Zone. A similar tectonic contact (called the

134

Alhamilla detachment [García-Dueñas et al., 1992; Martínez-Martínez and Azañón, 1997]) also

135

crops out in the Sierra Alhamilla. Last motions on that detachment in ductile conditions occurred

136

at ca. 14 -13 Ma [Martínez-Martínez and Azañón, 1997; Augier et al., 2005b, Platt et al., 2005]

137

while exhumation tectonics under the brittle regime lasted until ca. 8 Ma.

138

[7] Unlike the Huércal-Overa basin, the Sorbas basin presents evidences for compression

139

tectonics (Figure 2). The northern boundary of the basin is characterized by an overall gentle

140

onlap of the sediments over the basement of the Sierra de Los Filabres (Fig. 2b). Conversely, in

141

the southern part of the basin, Tortonian sediments are folded, locally overturned or even

142

overthrusted by basement rocks (Figure 3) [Weijermars et al., 1985; Ott d’Estevou and

143

Montenat, 1990; Giaconia et al., 2012a,2013]. The onset of compression in the region has been

144

ascribed to the upper Tortonian into the basin at around 8 Ma [Weijermars et al., 1985; Comas et

145

al., 1999; Augier, 2004; Jolivet et al., 2006; Meijninger and Vissers, 2006; Augier et al., 2013].

146

This compressional regime resulted in the formation of the current E-W oriented open-fold

147

syncline geometry of the Messinian to Quaternary series [Weijermars et al., 1985; Ott d’Estevou

148

and Montenat, 1990] and the (Figure 2B).

149

[8] Based on field investigations, several hypotheses were proposed to explain the initiation

150

phase of both the Sorbas [Montenat et al., 1987; Weijermars et al., 1985; Ott d'Estevou et al.,

151

1990, Poisson et al., 1999] and Tabernas basins [Kleverlaan, 1989]. These mainly differ on the

152

interpretation of the regional-scale tectonic regime occurring at that time:

153

- compressional to transpressional models with an overall N-S oriented maximum shortening

154

have been put forward [Montenat et al., 1987; Weijermars et al., 1985; Ott d’Estevou et al.,

155

1990; Sanz de Galdeano and Vera, 1992; Poisson et al., 1999]. They all consider the

156

Africa/Europe convergence as the principal driving force controlling crust deformation [Sanz de

157

Galdeano, 1985; Ott d’Estevou and Montenat, 1990]. In these models, the Sorbas basin is

158

interpreted either i) as a purely flexural basin developed along a major EW trending/northward

159

verging thrust [Weijermars et al., 1985] or ii) a transpressional basin linked to a wide NE-SW

160

strike-slip shear-zone [Sanz de Galdeano, 1985; Weijermars, 1987; Montenat et al., 1987; Ott

161

d'Estevou et al., 1990; Sanz de Galdeano and Vera, 1992; Stapel et al., 1996].

162

- a strike-slip faults network is active since the Middle/Upper Miocene [Sanz de Galdeano, 1985;

163

Booth-Rea et al., 2004a; Sanz de Galdeano et al., 2010; Rutter et al., 2012] and corresponds to

164

the onshore extension of the Trans-Alboran shear zone (Figure 1) [de Larouzière et al., 1988]. In

165

that context, the deposition of Tortonian sediments would have taken place in a pull-apart trough

166

[Kleverlaan, 1989; Stapel et al., 1996; Haughton, 2001; Martínez-Martínez et al., 2006; Rutter et

167

al., 2012].

168

- extensional models in which the formation of asymmetric intramontane basins are linked to the

169

late exhumation stages within the metamorphic domes of the NF metamorphic complex during

170

the collapse of the Internal zones [García-Dueñas et al., 1992; Vissers et al., 1995; Martínez-

171

Martínez and Azañón, 1997; Meijninger and Vissers, 2006; Augier et al., 2013]. Basins would be

172

initially extensional during the upper Serravallian-lower Tortonian [Meijninger and Vissers,

173

2006; Augier et al., 2013] before they experience inversion tectonics from upper Tortonian to

174

Quaternary.

175

[9] Recent gravity surveys show an irregular and/or asymmetrical distribution of the residual

176

gravity anomalies in the Almanzora Corridor (Figure 1) [Pedrera et al., 2009], in the Sorbas

177

basin [Li et al., 2012] and in the Huércal-Overa basin [Pedrera et al., 2010]. The Huércal-Overa

178

area has been particularly well documented: modeled gravity profiles evidence that the deepest

179

parts of the sediment/basement interface, reaching 1000 to 1500 m depth are confined to the

180

southern part of the basin [Pedrera et al., 2009, 2010] and more particularly along normal faults

181

[Augier et al., 2013]. Besides, tectono-stratigraphic and palaeostress analyses from brittle fault

182

show that two main depocenters were first controlled by SW-NE trending normal faults probably

183

rooting in the Filabres Shear Zone [Augier et al., 2013]. This brittle extensional event occurred

184

during the latest Serravallian to lower Tortonian (~12-11 Ma) and is linked to the exhumation of

185

the NF metamorphic complex [Augier et al., 2013]. Sedimentation started during the late

186

denudation stages of the NF complex as evidenced by zircon and apatite fission tracks ages and

187

U-Th/He analyses on apatites (respectively at 11.9 ± 0.9 Ma, 8.9 ± 2.9 Ma and 8.7± 0.7 Ma)

188

[Johnson et al., 1997; Vázquez et al., 2011] in both the Sierra de los Filabres and Alhamilla

189

[Augier et al., 2005b; Platt et al., 2005].

190 191

3. The Sorbas basin

192

3.1. Basement units bounding the basin

193

[10] While the Malaguide complex is poorly represented within the study area, the Alpujarride

194

and the Nevado-Filabride complexes are both well exposed within the metamorphic domes

195

forming the sierras [Vissers et al., 1995; Martínez-Martínez and Azañón, 1997]. The Alpujarride

196

complex is mostly represented by Permian purple to reddish phyllites overlain by Triassic

197

metacarbonates [García Monzón et al., 1973b, 1974; Azañón and Crespo-Blanc, 2000]. The

198

Nevado-Filabride complex is made of three main units [García-Dueñas et al., 1988; de Jong,

199

1991] called, from bottom to top, the Ragua Unit, the Calar-Alto Unit and the Bédar-Macael Unit

200

with respective average structural thicknesses of about 4,000, 4,500 and 600 m [Martínez-

201

Martínez et al., 2002]. The Ragua Unit is made up of Palaeozoic dark micaschists with

202

intercalated quartzites and rare dark marbles [Martínez-Martínez, 1985] while the Calar-Alto and

203

the Bédar-Macael units mainly contain Palaeozoic black micaschists, Permo-Triassic quartzites,

204

metapelites and Triassic metacarbonates. Dense eclogite and blueschist bodies or their

205

retrogressed equivalent amphibolite-facies metamorphic rocks are described as metric to

206

kilometric scattered lenses into both Calar-Alto and Bédar-Macael units [García Monzón and

207

Kampschuur, 1972; García Monzón et al., 1973a; Martínez-Martínez and Azañón, 1997].

208 209

3.2. Stratigraphy

210

[11] The oldest sedimentary record of the Sorbas basin surroundings is represented by scarce

211

Burdigalian-Langhian marly limestones badly preserved north to the Sierra Cabrera and

212

Serravallian shallowing-upward turbidites exposed along the Sierra Alhamilla and Sierra Cabrera

213

(Figure 1) [Ott d’Estevou and Montenat, 1990; Sanz de Galdeano and Vera, 1992]. The rest of

214

the series is quite similar to nearby basins [Sanz de Galdeano and Vera, 1992] and starts with

215

two different units [Ott d’Estevou and Montenant, 1990]: (i) a ~100 m thick upper Serravallian-

216

lower Tortonian continental coarse-grained deposits evolving basinward to yellowish marls and

217

turbidites (Figure 2C), which outcrop in the western (i.e. Tabernas area) and eastern (i.e.

218

Gafarillos area) parts of the basin (Figure 3A); (ii) a rather thick (up to 2000 m according to [Ott

219

d’Estevou and Montenat, 1990]) upper Tortonian marine unit, composed of grayish to yellowish

220

sandstones and turbidites (Figure 2C). It is worth mentioning that deposition of the upper

221

Serravallian-lower Tortonian sediments appear coeval with the latest exhumation of the NF

222

complex in the surrounding sierras [Ott d’Estevou and Montenat, 1990; Johnson et al., 1997;

223

Vázquez et al., 2011]. Latest Miocene sediments unconformably overly Tortonian sediments and

224

are composed of six units including from bottom to top (Figure 2C):

225

- a shallow marine fossiliferous calcarenite

226

Tortonian/Messinian boundary [Ott d’Estevou and Montenat, 1990; Sierro et al., 1993;

227

Krijgsman et al., 2001]; its thickness ranges from 10 to 80-90 m depending on its

228

paleogeographical context [Ott d’Estevou and Montenat, 1990; Puga-Bernabéu et al., 2007]

229

- an early Messinian marly unit known as the Abad marls composed of ~120 m of marls

230

alternating with diatomites. The upper part passes laterally to plurimetric fringing reefal

231

carbonates (Cantera member; Figure 2) outcropping both in the northern and in the southern part

232

of the basin (e.g. Cariatiz and Morrón de la Cantona localities; Figure 2A), bounding the basin

233

during the early Messinian [Martín et al., 1999]

(Azagador member) dated from the

234

- a 120 m thick evaporitic unit (Yesares member) consisting in gypsum alternating with clayey to

235

marly laminites [Dronkert, 1976] deposited during the Messinian Salinity Crisis [Clauzon et al.,

236

1996; CIESM, 2008] whose onset is calibrated at 5.971 Ma [Manzi et al., 2013]

237

- a latest Messinian-early Pliocene lagoonal calcarenite (Sorbas member) [Roep et al., 1998]

238

passing to a Gilbert-type fan delta (dated from 5.46 to 5.00 Ma) [Clauzon et al., in revision]; the

239

thickness of this marine unit reaches 75 m in the center of the basin at Sorbas locality

240

- finally, a mostly continental sedimentary unit made of reddish conglomerates called the

241

Zorreras member and reaching 60 m in the basin [Montenat et al., 1980; Martín-Suárez et al.,

242

2000].

243 244

4. Aims and approach

245

[12] Most of the studies on the Sorbas basin, having poor structural claims, focused on the

246

Messinian history [Rouchy and Saint-Martin, 1992; Clauzon et al., 1996; Roep et al., 1998;

247

Riding et al., 1998, 2000; Martín et al., 1999; Krijgsman et al., 2001; Braga et al, 2006;

248

Bourillot et al., 2009; Roveri et al., 2009]. Besides, while the overall structure of the post-

249

Tortonian deposits is documented as the consequence of a ~N-S compressional context (see

250

above), the tectonic control during the onset of subsidence and early development of the basin

251

remains debated and in fact mostly unexplored. This study aims at constraining the kinematics of

252

the Sorbas basin while the basin underwent its first subsidence phase and sedimentary deposition

253

during the upper Serravallian-lower Tortonian. This paper thus focuses on the structure of the

254

early deposits/basement contact which outcrops along the southern margin of the basin but also

255

on the 3D geometry of this contact, to depth. It also deals with the kinematics and stress

256

conditions that prevailed during the upper Serravallian-lower Tortonian period. Geological and

257

geophysical data have been used and integrated in this work including structural geology,

258

mapping, palaeostress computations and gravimetric survey and modelling.

259

[13] At first, a synthesis on the structure of the southern margin of the basin (i.e. the northern

260

edge of the metamorphic Sierras) is made with a particular focus on the contact between

261

Serravallian-Tortonian sedimentary units and the basement. It is based on a compilation of

262

existing data and own field observations. Then, structural and kinematical constraints are given

263

from structural field observations made in the older deposits of the basin, in specific areas. As

264

shown before, earliest deposits of the upper Serravallian–lower Tortonian are poorly exposed in

265

the study area because of the extension of post-Tortonian sedimentary series (Figure 2C, 3A) and

266

the ongoing deformation. However, systematic and accurate structural measurements have been

267

realized along the southern basin margin where outcropping conditions allow observing various

268

kinematic constraints. To first describe the initial development of the Sorbas basin and to insure

269

an accurate determination of the deformation kinematics, palaeostress orientation patterns were

270

evaluated by the computer-aided inversion method of fault-slip data [Angelier, 1984, 1990,

271

1994]. Thanks to the recognition of numerous stratigraphic time-markers including major

272

unconformities [Ott d’Estevou and Montenat, 1990; Sanz de Galdeano and Vera, 1992], the

273

succession of stress regimes is reasonably well constrained in time. As detailed in the followings

274

(see Structural Analysis), the study focused on (i) the structure of the basement/sedimentary

275

rocks interface, (ii) structural analyses and inversion of fault-slip data to obtain palaeostress

276

tensors carried out in the Serravallian-Tortonian series, and (iii) their internal structures of .

277

[14] Gravity data modeling has already proven its efficiency for imaging deep geological

278

structures at crustal-scale, especially for “basement” rock types imaging in plutonic and

279

metamorphic geological environments [Rousset et al., 1993; Lefort and Agarwal, 1999; Martelet,

280

1999; Martelet et al., 2004; Talbot et al., 2004; Joly et al., 2009; Turrillot et al., 2011]. While

281

gravity modeling is more rarely applied to sedimentary geological environments [Szafián and

282

Horváth, 2006; Pedrera et al., 2009, 2010; Li et al., 2012; Salcher et al., 2012], this method is

283

used here, in particular, to constrain the geometry of the interface between the rather dense

284

basement rocks of the metamorphic complexes and the lighter sediments of the basin. Deep

285

structures of the basin are thus constrained along a series of 2D gravity profiles overlapping the

286

basin and the surrounding basement units. Gravity models and field observations, including

287

bedding orientations measurement, were finally used to interpolate the 3D basin floor geometry.

288

These results allow improving our knowledge on the tectonic evolution during the development

289

of the Sorbas basin and precising the structural control on deposits geometry.

290 291

5. Structural analysis

292

5.1. Evidence of compressive tectonics

293

[15] The Sierra Alhamilla, Sierra Polopos and Sierra Cabrera represent three E-W elongated en

294

echelon antiforms (Figure 3). The Alhamilla and Polopos ranges form two north-verging

295

asymmetric folds with a north steep-dipping forelimb and a gentle south-dipping backlimb [Platt

296

et al., 1983]. They are together separated from the Sierra Cabrera by the dextral North Gafarillos

297

strike-slip fault (NGF, Figure 3). The northern front of the Sierra Alhamilla, Polopos and

298

Cabrera antiformal structures are composed of three major faults called, from East to West, the

299

North Cabrera reverse fault (NCRF), the North Gafarillos fault (NGF) and the North Alhamilla

300

reverse fault (NARF) [Giaconia et al., 2012a,b,2013] (Figure 3).

301

[16] The North Cabrera reverse fault affects the entire sedimentary filling and shows inverted or

302

strongly dipping Serravallian and Tortonian sediments with a N to NW transport striking [Booth-

303

Rea et al., 2004b] (cross-section I-I’, Figure 3). According to Booth-Rea et al. [2004b], the

304

activity of the NCRF initiated before the upper Tortonian as upper Tortonian conglomerates

305

unconformably overlies Langhian-Serravallian deposits. The fault activity is related to the

306

initiation of major sinistral Palomares and Carboneras strike-slip faults [Booth-Rea et al., 2004b;

307

Gràcia et al., 2006; Giaconia et al., 2012a]. The North Gafarillos and South Gafarillos faults

308

(NGF and SGF, Figure 3) represent two transpressive dextral faults accommodating the

309

deformation and the relative displacement of the Sierra Alhamilla with the Sierra Cabrera

310

[Giaconia et al., 2012a]. The E-W segment of the NGF, bounding the Sorbas basin to the south,

311

affects the Tortonian series (cross-section II-II’ in Figure 3) and most of the deformation

312

occurred during the Tortonian as Messinian reefal carbonates seal the contact toward the West

313

(Cantera Member, Figure 2B, 3) [Ott d’Estevou and Montenant, 1990; Stapel et al., 1996; Jonk

314

and Biermann, 2002]. The E-W North Alhamilla reverse fault zone, which is linked to the South

315

Gafarillos dextral strike-slip fault, can be followed over ~20 km along the southern border of the

316

Sorbas-Tabernas basin. This thrust fault zone cuts the northern limb of the asymmetric Alhamilla

317

antiform where Serravallian-Tortonian deposits display overturned bedding just below (Figure 2;

318

cross-section III-III’ in Figure 3). Its activity lasted from the upper Tortonian to the Late

319

Pleistocene with an overall top-to-the-NNW movement and controls the front of the Sierra

320

Alhamilla range by a drastic topographic gradient [Jonk and Bierman, 2002; Giaconia et al.,

321

2012a]. Several studies pointed out the syn-compression filling of upper Tortonian series

322

[Weijermars et al., 1985; Ott d’Estevou and Montenat, 1990; Haughton, 2001], reporting a

323

southward progressive gradient of deformation. Toward the west of the Sierra Alhamilla, the

324

NARF disappears and the nature of the contact separating Serravallian-Tortonian deposits from

325

the Alpujarrides metamorphic series appears unconformable (cross-section IV-IV’ in Figure 3).

326

There, the El Alfaro hill which is made of plurimetric conglomeratic and sandy interval belongs

327

to the upper Tortonian turbiditic series [Hodgson and Haughton, 2004]. According to Haughton

328

[2000] and Hodgson and Haughton [2004], the timing of deformation in the Alfaro area is also

329

estimated as Tortonian and Messinian in age from the sedimentary record (thick slump deposits

330

and internal unconformities).

331

[17] The overthrust of the Sierra Alhamilla basement range over the southern border of the basin

332

marks the strong asymmetry of both the basin and the basement range [Ott d’Estevou et al.,

333

1990; Martínez-Martínez and Azañón, 1997]. The NARF disappears toward the East of

334

Lucainena de las Torres (Figure 3), Figure 4 displays a landscape interpretation accompanied

335

with bedding measurements made across the basement/sediments contact (see Figure 2 for

336

location). The bedding attitude of the Tortonian series shows a progressive steepening over more

337

than 2 km and is even overturned in the vicinity of the Sierra Alhamilla (Figure 4); fold

338

asymmetry being illustrated by the more pronounced clustering of bedding poles in stereonet

339

projection (Figure 4). Structurally below, to the south, meta-sediments of the basement units

340

show parallel overturned bedding orientation to the basal layers of the basin marking an

341

overturned forelimb of the Sierra Alhamilla. There, upper Tortonian sediments are separated

342

from the basement by an unconformity or a stratigraphic break as the upper Serravallian-lower

343

Tortonian continental series are missing. East of the NARF, the geometry of the

344

basement/sediment interface argues in favour of a north-verging asymmetric fold in the Sierra

345

Alhamilla with an overturned steep-dipping forelimb, highlighting the recumbent syncline of

346

Tortonian deposits. Such folded structure could be interpreted as a fault-propagation fold.

347

[18] Progressive unconformities (e.g. growth-strata) in upper Tortonian sediments (8-7.24 Ma),

348

mark the growth of the Sierra Alhamilla basement range and determine a good age-control on

349

that deformation episode. This compressional episode mostly shaped the current geologic and

350

topographic features of the southeastern parts of the Internal zones of the Betic Cordillera and

351

pre-dates the Messinian to present evolution of the basin. Indeed, the overlying Messinian

352

carbonates seal most of the folding linked with the growth of the Sierra Alhamilla range (Figure

353

2, 3) [Weijermars et al., 1985; Ott d’Estevou and Montenat, 1990] but are in turn involved in the

354

present-day large-scale open fold of the Sorbas basin. Internal deformation of the upper-

355

Tortonian series is characterised by small-scale wrench faulting with only rare reverse faults as

356

shown along the road-cut near Lucainena de las Torres (Figure 4). Analysis of fault-slip data

357

yielded a strike-slip palaeostress tensor solution with a N-S maximum compression direction [1]

358

and a W-E minimum compression direction [3] in that case, consistent with the large-scale

359

folding of the northern front of the Sierra Alhamilla (Figure 2B and 4).

360 361

5.2. Evidence of extensional deformation

362

[19] Upper Serravallian-lower Tortonian deposits, which are particularly abundant in the

363

Huércal-Overa basin, are in contrast very limited within the Tabernas-Sorbas basin. They are

364

encountered along the northern limb of Sierra Alhamilla (Figure 3). These deposits are of prime

365

importance with respect to the deep structure of the basin and consequently with the buried

366

history of the Sorbas basin. Hereafter, two groups of outcrops are described as representative of

367

that discussion, the Tabernas and the Gafarillos areas (see location in Figure 2).

368

[20] In the Gafarillos area, upper Serravallian-lower Tortonian formations crop out well along

369

roads that cross the hills, particularly the ca. N020E striking road linking Gafarillos to Gacia-

370

Bajo. There, the sedimentation, characterized by conglomerates and coarse-grained sandstones

371

presents general subhorizontal bedding and is affected by a pervasive network of small-scale

372

normal faults (Figure 5A) locally showing synsedimentary character (Figure 5B). The entire

373

normal faults population, generally developed as 50-70° dipping roughly planar surfaces,

374

presents two distinctive orientations: a dominant ca. N140E and a subordinate ca. N030E

375

conjugate sets of normal faults. Associated stress tensors correspond to subhorizontal extension

376

presenting respectively N060E and N140E minimum compression direction [3] and a common

377

subvertical maximum compression direction [1] (Figure 5A). Analysis of younger rocks

378

belonging to the basal upper Tortonian in the same area (inset C in Figure 5), revealed that only

379

the ca. N140E set of normal faults was present, suggesting that the former N30E system was

380

restricted to the upper Serravallian-lower Tortonian formations.

381

[21] The Tabernas area provides an example of geometrical relationships between faulting and

382

folding (see location on Figure 2). There, upper Serravallian to upper Tortonian formations

383

experienced a large-scale folding episode (i.e. the so-called Tabernas fold) [Ott d’Estevou and

384

Montenat, 1990] particularly well exposed along the incised Rambla de la Sierra. A landscape

385

picture is given on Figure 6. In details, the whole area presents a pervasive network of small-

386

scale faults systems including mostly normal and wrench faults whose relative chronology and

387

relationships with folding remained so far unexplored. Close-up views of two parts of the

388

northern limb of the fold are given as insets in Figure 6; inset B documents faulting in the upper

389

Serravallian-lower Tortonian formations and inset C, the upper Tortonian. Moreover, Figure 6D

390

provides a decomposition of the total heterogeneous fault population into homogeneous,

391

cogenetic fault populations. Normal faulting appears clearly related to pre-folding extension as

392

suggested by the coexistence of both subvertical and subhorizontal normal and reverse faults (in

393

the present attitude of bedding) passively rotated with respect to bedding (Figure 6B and 6C). In

394

addition, it is to be noted that some of them reveal a syn-sedimentary character. Computed

395

principal stress regimes, in agreement with an Andersonian model after back-tilting [Anderson,

396

1942] yielded a succession of two subhorizontal extensional regimes with N50E and N140E

397

minimum compression direction [3] for an almost vertical common maximum compression

398

direction [1]. Wrench faults that often reactivate former normal fault planes present N20-80E

399

sinistral faults accompanied by N120-170E dextral faults. Analysis of fault-slip data yielded a

400

strike-slip palaeostress tensor solution characterised by a N-S maximum compression direction

401

[1] and a W-E minimum compression direction [3]. This result is consistent with the large-

402

scale W-E folding that affect passively the Sierras [Weijermars et al., 1985; Ott d’Estevou and

403

Montenat, 1990] and the basins (Figure 2 and 3) with a ca. 10-20° dispersion as exemplified by

404

the WSW-ENE Tabernas anticline (Figure 6D).

405

[22] The structural development and the sedimentation of the Tabernas-Sorbas basin thus appear

406

controlled by normal faulting from upper Serravallian-lower Tortonian onward and at least

407

during a part of the upper Tortonian. These traces of extension whose origin appear coeval and

408

may be related with the regional-scale exhumation of the Nevado-Filabride metamorphic

409

complex [Martínez-Martínez and Azañón, 1997; Augier et al., 2005a, 2013] are now mostly

410

hidden by thick younger deposits in the Sorbas basin area.

411 412

6. 2D gravity survey

413

6.1. Acquisition and treatment

414

[23] A gravity survey was conducted throughout the Sorbas basin and surrounding basement

415

areas. Ground measurements were performed using a SCINTREX CG5-M micro-gravimeter in

416

195 gravity stations emplaced along 6 ~north-south trending sections, orthogonal to regional-

417

scale structures, and one additional longitudinal (i.e. ~east-west) section for crosschecking

418

(Figure 7). Measurement stations were spaced ca. 1 km apart along the two longer sections

419

aiming at imaging regional-scale structures (see N-S trending SCS1 “Sorbas Cross-Section 1”

420

and SCS2 E-W trending section; Figure 7). Along other profiles, stations were spaced ca. 500 m

421

apart within the Sorbas basin, and ca. 250 m apart along the segments in the vicinity of the

422

basement/sediment contact. To better control the raw data quality, measurements were repeated

423

three times at each gravity station. Geographic coordinates and altitude were measured using a

424

MAGELLAN Promark3 post-D-GPS system. For each station, location was recorded and

425

averaged during more than one minute and values were adjusted through a post-treatment

426

correction process by using a secondary antenna that was fixed for the entire survey period. Such

427

acquisition method ensures ca. 10 cm and 20 cm of precision for horizontal and vertical

428

coordinate values, respectively. To ensure consistency between the different gravity profiles

429

measured and to control the temporal instrumental drift, a unique gravity reference station was

430

used during the overall field survey; it was systematically measured twice a day, before and after

431

daily surveys (Figure 7B, Los Martinez locality). To calculate the Bouguer anomaly, standard

432

free air, plateau and terrain corrections were successively performed; a 2.67 g.cm-3 Bouguer

433

reduction density was chosen as an average density value for crustal rocks. Inner terrain

434

corrections (for 50 m of distance away from station) were routinely performed using Hammer

435

charts [1939]. Following Martelet et al. [2002], far field terrain corrections (up to 167 km away

436

from the station) were computed numerically using the SRTM digital elevation model. Finally,

437

the Bouguer anomaly values were calibrated (with a constant shift) to the available regional-

438

scale gravity data [Torne et al., 2000] in order to refer to a common and absolute gravity

439

reference with other existing data sets (see Figure 8A).

440

[24] Two dimensional (2D) direct gravity modeling was performed using the Geosoft-GM-SYS

441

software. Gravity Bouguer anomalies depend on both the geometry and the density contrasts

442

between geological units. In order to avoid multi-solutions models from the gravity analysis,

443

densities of the geological units were measured at the laboratory using the hydrostatic weighing

444

method. Densities used for each lithology considered in the model are listed in Table 1 and

445

detailed below. These were also systematically compared to published density range values

446

usually considered for classical lithologies [Robinson and Çoruh, 1988; Telford et al., 1990,

447

Jacoby and Smilde, 2009]. While density values were fixed, the modeling process then consisted

448

to adjust geometries of the geological units along the sections. The Moho depth variation

449

generally induces large wavelength variations in gravity anomaly profiles and, for this study, the

450

geometry of this particular interface is fixed from the available reference dataset of Fullea et al.

451

[2006] or Ziegler and Dèzes [2006]. The geometry of each geological unit was drawn in

452

accordance to the geological maps [García Monzón et al., 1973b, 1974] (Figure 2).

453 454

6.2. Rock densities

455

[25] Rock densities were directly measured from several rock samples for each geological unit.

456

Indeed, the determination of average standard densities for a geological formation highly

457

depends on the definition of the lithological assemblage [Jacoby and Smilde, 2009]. Gravity

458

modeling based on measured densities ensures a much better control. Both the Alpujarride and

459

the Nevado-Filabride complexes metamorphic basement were sampled (Figure 2). The purple

460

micaschists of the Alpujarride complex were sampled near Lucainena de Las Torres at the

461

southern edge of the basin. The corresponding measured density (2.69 g.cm-3; see Table 1) is

462

consistent with the general bulk densities usually considered for micaschists and phyllites (i.e.

463

from 2.64 to 2.74 g.cm-3) [Jacoby and Smilde, 2009]. Due to the occurrence of dolomitic layers

464

at all scales, the density of the Triassic metacarbonate rocks from the upper part of the

465

Alpujarride complex was fixed from the literature [e.g. Jacoby and Smilde, 2009] with a bulk

466

density of 2.7 ±0.2 g.cm-3. An average density of 2.69 g.cm-3 was thus attributed to the overall

467

Alpujarride complex for modeling. Micaschists samples of all three units composing the

468

Nevado-Filabride complex (i.e. the Ragua Unit, the Calar-Alto Unit and the Bédar-Macael Unit)

469

yielded densities of 2.65, 2.67, and 2.69 g.cm-3 respectively (Table 1). These densities are in

470

good agreement with both average densities published in the literature [Jacoby and Smilde,

471

2009] and the fact that the metamorphic grade of these complexes increases in this same order.

472

Indeed, despite a relatively common lithological succession, one main distinction between those

473

units is ascribed to the presence of metabasite bodies in both Bédar-Macael and Calar-Alto units

474

(which include eclogite and amphibolite rocks) that were not directly measured [García Monzón

475

et al., 1973a; Martínez-Martínez and Azañón, 1997]. Average densities of such basic rocks are

476

respectively about 3.3 and 2.9 g.cm-3 [Jacoby and Smilde, 2009]. Hence, the adjacent Bedar-

477

Macael and Calar-Alto units have been considered as a single unit for gravity modeling,

478

hereafter called the Upper NF unit, with a slightly increased average density of 2.7 g.cm-3 (Table

479

1). The lowermost metamorphic unit has been defined as the Lower NF unit (corresponding to

480

the above mentioned Ragua unit) with a measured average density value of 2.65 g.cm-3. On the

481

regional-scale cross-section SCS1 (Figure 7A), the use of dense bodies was not necessary to

482

adjust the modeled residual Bouguer anomaly with the measured one. At that scale, far-field

483

effects of these scarce and dense bodies can be neglected. On the contrary, adjustment of the

484

modeled residual Bouguer anomaly on small-scale cross-sections required the use of such

485

bodies, in accordance with their exposures [García Monzón et al., 1973b, 1974].

486

[26] Sampling of the sedimentary cover has been performed throughout the entire basin (Figure

487

2). Apart conglomeratic layers (from the upper Serravallian to lower Tortonian), whose density

488

cannot be obtained by laboratory measurement, densities of all the sedimentary units have been

489

measured (see Table 1). Two groups have been distinguished for modeling: one including

490

Serravallian and Tortonian deposits and another one from lowermost Messinian to uppermost

491

Pliocene layers. The average density obtained for Tortonian deep marine deposits (2.54 g.cm-3;

492

Table 1) has directly been used in the 2D modeling process. Densities measured for the

493

lowermost calcarenite, messinian marls-reefs-evaporites, and the overlying latest Messinian-

494

early Pliocene deltaic deposits are summarized in Table 1. The density measured for the

495

Messinian marls, yielding a density of 1.6 presents much lower values than the density usually

496

considered for such rock type (2.4 g.cm-3; Jacoby and Smilde, 2009). Consequently, and in order

497

not to unbalance the gravity modeling, an average density of 2.4 g.cm-3 has been considered for

498

his unit. Finally, a mean 2.37 g.cm-3 density has been calculated for the entire lowermost

499

Messinian-uppermost Pliocene group. Note that the density of the topmost Pliocene

500

conglomeratic layers has not been measured because of their sample-scale lithological

501

heterogeneity and their non-cohesive structure. It has been considered equal to the Pliocene

502

marine deposits, which yielded averaged density of 2.36 g.cm-3.

503 504

6.3. Gravity profiles modeling

505

[27] A 55 km-long profile, crossing the Sierra Alhamilla to the South and the Sierra de Los

506

Filabres to the North, was acquired to image the effect of the overall crustal structure on the

507

gravity signal (Figure 8). Indeed, wavelength of the gravity anomaly primarily depends on the

508

depth (for given density contrasts) of the “source” geological bodies and structures; rather long

509

profiles are thus required to image the deepest crustal geometries. Here, the long N-S Bouguer

510

anomaly profile, hereafter named the “regional-scale” profile, shows a large northward decrease

511

from +15.15 mGal to -42.10 mGal (Figure 8A); this long wavelength anomaly corresponds to the

512

regional Gravity anomaly and, at first order, this can be related to a north-dipping Moho (Figure

513

8B). This feature is in good agreement with the Moho depth model/compilation published by

514

Ziegler and Dèzes [2006] or Diaz and Gallart [2009] showing a progressive northward

515

deepening, from 22 km to 32 km, beneath the Sierra de Los Filabres.

516

[28] The residual Bouguer anomaly values corresponding to this regional profile (Figure 8) were

517

calculated by substracting the measured Bouguer anomaly and the regional Bouguer anomaly as

518

given by Torne et al. [2000] or Fullea et al. [2010]. The obtained profile then appeared very

519

similar to the one recently published by Li et al. [2012] – who surveyed approximately the same

520

regional-scale section – with similar amplitudes and shapes of the two residual Bouguer anomaly

521

curves (compare Figure 8A and [Li et al., 2012]:figure 3). As the two surveys and studies have

522

been strictly conducted independently, such features highlight a good reproducibility of the

523

measurements and treatments and the quality of the gravity data used for this study. Here, all the

524

2D gravity models were computed from the “total” Bouguer anomaly values (instead than from

525

the residual Bouguer anomaly) by taking into account the geometry of the moho, as mentioned

526

above [Torne et al., 2000; Fullea et al., 2006; Ziegler and Dèzes, 2006]. It is noteworthy that,

527

below the lowermost NF metamorphic unit, standard densities have been used for the crust with

528

2.7 and 2.8 g.cm-3 for respectively the “mean” and lower crust levels.

529

[29] On the other hand, short-wavelengths of the measured gravity profiles highly depend on the

530

shallow crustal structure. In this study, this part of the modeling was constrained by the geometry

531

of the outcropping post-Tortonian deposits within the Sorbas basin (see Figure 2) and of the

532

metamorphic unit, into the domes, as deduced from field structural measurements. The mid-

533

wavelength anomaly (Figure 8A) constituted the only remnant part of the measured signal to be

534

fitted. If considering the density contrast in between units (see Table 1), this latter depends on the

535

depth and shape of the base of the basin but also on the geometry of internal boundaries within

536

the underlying metamorphic complexes (Figure 8B). The residual mid-wavelength anomaly was

537

then modeled by adjusting the geometry of the contact between metamorphic basement and

538

sediments (Figure 8B), as is the case to the northern boundary of the basin (Figure 8C). As a

539

whole, the modeled Bouguer anomaly can be separated in three components along this regional-

540

scale profile (see Figure 8C), displaying the combined effects of (i) the deep crustal structures

541

(red curve) [ii] the upper crust structure into the outcropping metamorphic units (blue curve) and

542

(iii) the shape and thickness of sedimentary basins (green curve in Figure 8C). Note that, even

543

though a large part of the “total” Bouguer anomaly range value is directly linked to the Moho

544

depth (see red curve in Figure 8C), a significant part of its variations is also due to the structure

545

of the upper crust of both the metamorphic and sedimentary units, with similar proportions (see

546

similar amplitudes of the blue and green curves in Figure 8C). Finally, on the regional-scale

547

SCS1 profile the mid-wavelength Bouguer anomaly decreases from 5 to 25 km, which

548

corresponds to an asymmetric shape of the basin floor as shown in the model (Figure 8D).

549

[30] Profiles SCS3, SCS4, SCS5, SCS6 and SCS7 (Figure 7), located in the center of the Sorbas

550

basin, are described from West to East (Figure 9). In comparison to the above described ~55 km

551

long regional-scale profile, these latter shorter profiles stand for more local-scale variations in

552

the structure of the basin. For sake of clarity, modeled geometries of the units are only displayed

553

down to 4 km on the sections but, as for the processing of SCS1 profile, the whole crust is taken

554

into account for gravity anomaly computation, down to the mantle, in order to include depth

555

variations of the Moho. Profile SCS6 extends to the North from Uleila del Campo to the west of

556

Sorbas city (Figure 7). The short wavelengths of the Bouguer anomaly are well fitted by taking

557

into account the extension of the northern boundary of the Tortonian trough beneath the

558

Messinian and Pliocene sediments (see Figure 2). On the other hand, the gravity high displayed

559

in the Bouguer profile (between 2 and 4 km) and forming a mid-wavelength anomaly cannot be

560

simply reproduced by the presence of standard metamorphic rock units below the post Tortonian

561

deposits. Such wavelength can neither be explained by variations in the Moho depth or in the

562

lower crust. Gravity modeling shows here that the occurrence of rather dense material (of 2.9

563

g.cm-3) in the upper crust is mandatory, at that local-scale, to properly model the Bouguer gravity

564

anomaly. Slices of such dense rocks are displayed in black on Figure 9. Note that such small

565

bodies are only necessary to adjust short wavelength of the anomaly in “local” gravity profiles

566

and are not mandatory for modeling the regional scale profile as described above. SCS4 profile,

567

which extends 3 km south of Sorbas, is the southern extension of the SCS6 profile. The first

568

order trend of the modeled Bouguer anomaly was adjusted with the geometry of the

569

metamorphic basement units at depth. Shortest wavelengths have been then modeled with the

570

depth of the interfaces of shallow geological units exposed at the surface. SCS3 profile is almost

571

parallel to SCS6 and extends from Sorbas to the north of El Pilar (Figure 7). The same modeling

572

protocol as described above was used here to adjust the computed Bouguer anomaly curve to the

573

measured anomaly values. Here again, this section displays the extension of the northern

574

boundary of the Tortonian trough below the younger sedimentary series. The N-S oriented SCS7

575

profile extends from Los Feos (northern part of the Níjar basin) to Los Perales (Figure 7). This

576

profile crosses a metamorphic basement high connecting the Sierra Alhamilla in the west to the

577

Sierra Cabrera in the east and where metamorphic rocks have been sampled to constrain density

578

values (Figure 2). The northward decrease of the Bouguer anomaly values, from +21 to +8 mGal

579

(Figure 9), was ascribed to the anticline shape of the metamorphic units as known laterally in

580

both Sierra Alhamilla and Sierra Cabrera. At 2.8 km distance along the profile, a trend break is

581

visible in the anomaly curve and this feature most probably coincides with the

582

basement/sediments contact outcropping there (Figure 9). This argues for a rather steep

583

downward extension of the contact. Besides, the sedimentary cover is here entirely composed of

584

Tortonian sediments and this structure may correspond to the termination of the Tortonian

585

extensional trough to the SE. The westernmost gravity section extends from La Herreria to north

586

of El Chive (SCS5 profile; Figure 7). Three main segments can be distinguished along the

587

Bouguer anomaly profile:- the first one, between kilometer 0 and 2, is characterized by a

588

northwestward strong decrease; the best fit of the model was obtained by an important deepening

589

of the sedimentary infill of the basin ;

590

- the second segment, between 2 and 7 km, is marked by a rather constant low anomaly which

591

corresponds, after adjusting the model, to the thickest sediment filling along the cross-section,

592

- the third segment, beyond 7 km, highlights a rather weak increase of the anomaly before a new

593

stabilization around +3 mGal. This “step” in the trend of the Bouguer anomaly profile may

594

correspond to the extent of the Upper NF units from the Sierra de Los Filabres southward below

595

the post Tortonian series (Figure 2, 7). Except for the extreme northern boundary of the SCS6

596

profile, the measured and modeled Bouguer anomalies show a good consistency and the resulting

597

misfit error is negligible for all profiles.

598 599

7. 3D geometry and depth constraints

600

[31] 2D gravity sections were integrated into the 3D GeoModeller software [Calcagno et al.,

601

2008; Guillen et al., 2008] in order to interpolate the 3D geometry of the basin-floor. The

602

computation principles of this software are based on a joint inversion of both the places, within

603

the 3D volume considered, where geological contacts have been observed or deduced, and

604

measured orientations of corresponding geological surfaces. Here, modeled 2D gravity sections

605

were thus used to constrain both the depth of the basin-floor and its geometry. The modeling box

606

(dashed border rectangle in Figure 7B) extends over 20 x 15 km from 37°10’12.33” N -

607

2°15’03.00” W (top-left corner) to 37°01’48.44” N – 2°02’12.75” W (bottom-right corner) and

608

from 5 km below sea-level to 3 km in altitude.

609

[32] The geometry of the basin floor is illustrated in Figure 10 by both an iso-depth curves map

610

(Figure 10A) and an oblique 3D view of the model in which the sediment infill have been

611

removed (Figure 10B). Iso-depth curves of the basin floor, as interpolated during the 3D

612

modeling process, show a strong marked asymmetry of the basin geometry (Figure 10A)

613

characterized by a steep North-dipping boundary in the southern part of the basin, passing to a

614

gently South-dipping contact to the North. This first-order geometry of the model is well

615

constrained by both (i) the gravity 2D modeling along the profiles and (ii) the very steep dip of

616

the first Tortonian layers even characterized by locally overturned bedding close to the basement

617

contact. The deepest part of the basin, a 2000-2800 m trough, is located in the South, facing the

618

North Alhamilla front zone separating the basin from the metamorphic basement of the Sierra

619

Alhamilla.

620

[33] While the asymmetrical shape of the basin evidenced by our computed 3D model is

621

compatible with the 2D image proposed by Li et al. [2012], discrepancies arise when considering

622

the modeled depth of the basement/sediment basement below the basin. Indeed, previous work

623

estimates a ~1500 m maximum depth of the Sorbas basin in its southern part whereas our 3D

624

computations result in a 2000-2800 m depth localized along the southern part of the basin. When

625

considering the SCS1 regional scale gravity section of this study, comparison of the residual

626

Bouguer anomaly values with the one obtained by Li et al. [2012] clearly shows that such

627

discrepancy could not be attributed to measurement deviations. Thus, differences in depth

628

computations must be assigned to the choice of the density values used for gravity modeling. On

629

the one hand, previous gravity studies conducted in the Sorbas basin and surrounding areas

630

where considering geological units in, basically, two groups: (i) sedimentary rocks with a

631

constant density fixed at 2.35 g.cm-3 and (ii) basement rocks with a constant density fixed at 2.67

632

g.cm-3 [Pedrera et al., 2009, 2010; Li et al., 2012]. We have tested this same approach on the N-

633

S regional cross-section to model the basal contact separating the sediments from the

634

metamorphic basement (dashed line on Figure 8D). The overall asymmetric shape of the basin

635

fill can be reproduced, however the main difference between these approaches is the resulting

636

depth of the basin-floor estimated between 1.6 to 2.8 km (Figure 8D). However, the Bouguer

637

anomaly primarily depends on density contrasts [Martelet, 1999; Martelet et al., 2004; Jacoby

638

and Smilde, 2009; Joly et al., 2009]. Accurate determination of the density is thus essential to

639

model a more reliable geometry to depth. The Sorbas basin, as well as its basement, is composed

640

of a large spectrum of rocks and average densities assigned to each geological layer sufficiently

641

contrasts with both adjacent units to be considered as a representative density interval (Table 1).

642

Our study strives to consider more specific density values for geological units or groups as

643

recognized in the area as well as measuring densities from field samples rather than using

644

standard average density values. When comparing densities of rock units measured in this study

645

to standard ones, we suggest that the distinction of as many geological units as possible to

646

separate “geophysical” units is a more accurate approach to estimate both the geometries and

647

depths of the geological boundaries.

648 649

8. DISCUSSION

650

[34] Contrasted tectonic settings have previously been proposed to explain the initiation and the

651

development of sedimentation within the Sorbas basin: (i) N-S compression to transpression

652

[Montenat et al., 1987; Weijermars et al., 1985; Ott d’Estevou et al., 1990; Sanz de Galdeano

653

and Vera, 1992; Poisson et al., 1999], including strike-slip tectonics [Kleverlaan, 1989;

654

Haughton, 2001] or (ii) extension [García-Dueñas et al., 1992; Vissers et al., 1995; Martínez-

655

Martínez and Azañón, 1997; Augier et al., 2005a; Augier et al., 2013]. Based on field structural

656

investigations and the gravity survey, the 3D geometrical model shows a marked asymmetric

657

shape with the deepest parts localized to the South, along the contact with the Sierra Alhamilla.

658

The southern basin/basement rocks contact is steeply dipping, so as the bedding within the

659

Tortonian basal deposits. The southern deepest parts of the basin correspond to the Tortonian

660

trough, as previously defined. The strong asymmetry of the basin might result from the

661

compressive event, when the Sierra Alhamilla “thrusted” over the southern edge of the Sorbas

662

basin. Indeed, the North Alhamilla Reverse Fault has been described as a prominent structure

663

and its development could have been accompanied by a significant subsidence in its footwall.

664

However, structural analysis highlighted that while north-verging thrusting can be evidenced

665

along this margin, some segments of the north Alhamilla front display a preserved structure of

666

unconformable stratigraphic contact of the Tortonian series on the basement rock units. Such

667

interface was subsequently tilted or even overturned in front of the Sierra Alhamilla but no trace

668

of thrusting can be evidenced then. In these areas (as presented in Figure 4) the deformation

669

marks the growth of a fault-propagated fold at the tip of a blind thrust and such quick change in

670

tectonic style along the north Alhamilla front argues for a rather limited dip slip movements

671

along the NARF. On the other hand, while previous works estimate ~1500 m of maximum depth

672

for the Sorbas basin, our new gravimetric models and 3D computations result in a 2000-2800 m

673

maximum depth localized along the southern part of the basin. This result reinforced the

674

asymmetric character of this basin. It also highlights a much larger cumulated subsidence in this

675

area compared to the Huércal-Overa basin for instance.

676

[35] Moreover, several syn-sedimentary normal faults have been locally described from the

677

upper Serravallian to the upper Tortonian series, in close surroundings to the Sorbas basin

678

(Figure 5, 6). Two extensional paleostresses, whose minimum compression direction [3] are

679

N140E and N50-60E oriented, controlled the deposition of upper Serravallian to upper Tortonian

680

sediments (Figure 5, 6). The analysis of upper Tortonian series, in particular in the Gafarillos

681

area, suggests that the N140E oriented system was mainly restricted to the upper Serravallian-

682

Early Tortonian. In the Tabernas area, both normal faults sets appear clearly as pre-folding

683

extensional faulting as suggested by tilted conjugate sets of normal faults (Figure 6B and 6C).

684

The asymmetry of the basin floor most probably results from the successive extensional and

685

compressive deformations evidenced in the Sorbas basin. Indeed, the limited dip-slip movements

686

along the northern “thrust” front of the Sierra Alhamilla could hardly explain the ~2500 m of

687

subsidence in the south Sorbas trough. In addition, the amount of shortening due to the folding of

688

the Alhamilla range remains hardly assessable because of the dextral movements of the Polopos

689

and North Alhamilla Fault Zone [Jonk and Biermann, 2002; Giaconia et al., 2012a] and the

690

involvement of the strike-slip Cantona Fault Zone in the deformation [Haughton, 2001].

691

Therefore, we assume that a significant imprint of the early extensional event contributed on the

692

finite global architecture and structure of the Sorbas basin.

693 694

8.1. The initial extension history of the Sorbas basin and surroundings

695

[36] The onset of sedimentation roughly occurred at the same time in all intramontane basins of

696

SE Betics, during upper Serravallian to lower Tortonian [Cloething et al., 1992; Sanz de

697

Galdeano and Vera, 1992; Iribarren et al., 2009]. The Huércal-Overa and Sorbas basins

698

underwent a common subsidence history during their early development stage (Figure 11): this

699

first phase is characterized, in both basins, by the deposition of a thick coarse-grained breccia

700

containing Nevado-Filabride clasts [Sanz de Galdeano and Vera, 1992]. Both basins display a

701

synchronous onset and a similar evolution of sedimentation [Sanz de Galdeano and Vera, 1992]

702

while their basement were exhumed by normal faulting. Evidences of extensional tectonics

703

during the lower Tortonian have also been extensively described into the Huércal-Overa basin, to

704

the north (Figure 12A) [Briend et al., 1990; Augier et al., 2005b; Meijninger and Vissers, 2006;

705

Pedrera et al., 2010; Augier et al., 2013]. In that area, the latest exhumation stages of the

706

metamorphic Nevado-Filabride complex (green arrows in Figure 12A) [Johnson et al., 1997;

707

Augier et al., 2005b] coincides with the kinematics of normal faults which controlled the

708

sedimentation into the basin [Augier et al., 2013]. Both basins may have experienced common

709

extensional tectonics during the initiation of sedimentation (Figure 12A). The effect of extension

710

is obviously hardly detectable on the southern edge of the Sorbas basin. However, similar

711

extensional regime boarding metamorphic core-complexes in other surrounding areas led to the

712

formation of half-graben basin such as the Granada basin [Ruano et al., 2004], Fortuna-

713

Guadalentin basins [Amores et al., 2001, 2002], and in the Alboran Sea [Mauffret et al., 1992;

714

Watts et al., 1993; Comas et al., 1999]. We assume that the Sorbas basin which initiated under

715

similar tectonic conditions (Figure 12) may have formed an asymmetric half-graben tilted toward

716

the south. The fact that extensional paleostresses calculated in this study, and especially the

717

N140E oriented system, is consistent with the latest ductile deformation highlighted in the Sierra

718

Alhamilla (green arrows in Figure 12A) reinforces the comparison with the Huércal-Overa basin

719

and the interpretation of half-graben initiation. In the Sorbas basin, gravity modeling evidences

720

that the rather dense rocks of the Alpujarride complex with respect to the sedimentary cover,

721

cannot be prolonged northward, below the basin (Figure 9, 13E), without inducing a large misfit

722

on the measured Bouguer anomaly. Hence, the southern fault wich probably controlled the

723

sediments deposition during the initiation of the subsidence may cross-cut and post-date the

724

Alhamilla detachment zone itself (Figure 13E). This is fairly consistent with the tectonic history

725

of the Huércal-Overa basin and the ca. 12-11 Ma ages for the initiation of sedimentation during

726

latest exhumation stages of the sierras under brittle conditions [Platt et al., 2005].

727 728

8.2. Evolutive scenario of the Sorbas basin

729

[37] A N-S evolutive cross-section of the Sorbas basin is given on Figure 13. Situations are

730

restored for key main steps from the upper Serravallian-lower Tortonian stage to the present

731

view by deconvoluting the tectonic history of the Sorbas basin. As discussed before, we argue

732

here for an onset of sedimentation occurring in an extensional regime in the Sorbas basin. The

733

strong asymmetry of the basin is compatible with a half-graben geometry controlled by a

734

supposed north dipping fault zone, located along the southern margin during Serravallian and

735

part of the Tortonian (Figure 5 and 6). Geometry and major bounding faults of this initial basin

736

can obviously not be retrieved accurately as most of this basin is now hidden by younger

737

deposits; a significant part of the early basin even being eroded in particular during Messinian

738

evolution. Extensional kinematics, as studied within the preserved and outcropping parts of the

739

basins, is compatible with a NW-SE (~N140E) and NE-SW (~N50-60E) directed subhorizontal

740

extension accompanied by a subvertical maximum compression direction (Figure 6). This event

741

may have favoured an initial subsidence pulse and the deposition of coarse-grained sediments

742

grading to marine deposits. Such pulse thus seems recorded over most the SE Betic basins [Sanz

743

de Galdeano and Vera, 1992; Meninger and Vissers, 2006; Augier et al., 2013]. Inception of the

744

main sedimentation occurred shortly after the crossing of the ductile-brittle transition at ca. 14-

745

13Ma during the final exhumation and the denudation of the Nevado-Filabride-cored domes

746

[Johnson et al., 1997; Augier et al., 2005b; Vasquez et al., 2011]. In the Huércal-Overa basin,

747

where the corresponding oldest deposits are exposed and even very well preserved, this time

748

interval corresponds to sedimentation of about 1 km thick series along the normal-faulted

749

southern boundary [Pedrera et al., 2009, 2010; Augier et al., 2013]. As the two basins show a

750

similar tectonic subsidence evolution during upper Serravallian-lower Tortonian, one can assume

751

that, in the Sorbas basin, the subsidence pulse led reasonably to the deposition of the same

752

amount sediments.

753

[38] Later, during the upper Tortonian, the Sorbas basin experienced a drastic inversion tectonics

754

stage (Figure 4 and 6). This formation shows a progressive steepening over more than 2 km and

755

is even overturned in the vicinity of the Sierra Alhamilla in the footwall of the North Alhamilla

756

reverse fault zone (Figure 4). Besides, the development of upper Tortonian growth strata

757

highlights that deformation occurred during sedimentation (Figure 4). Second order folding (e.g.

758

the Tabernas fold) and small-scale faults are all consistent with a N-S to NW-SE shortening

759

(Figure 4 and 6). Progressive steepening on the southern margin of the basin (i.e. the mobile

760

border) and the significant uplift of the Sierra Alhamilla [Braga et al., 2003] probably resulted in

761

a continuous tectonic subsidence within the basin and the persistence of marine deposits

762

[Weijermars et al., 1985; Kleverlaan, 1989]. Flexural type subsidence then superimposed on the

763

former extensional depocenter and migrated northward (Figure 13D). This marked inversion of

764

the basin appears mechanically in agreement with the development of dextral strike-slip faults

765

recognised at the western and eastern edges of the basin [Giaconia et al., 2013].

766

[38] Many studies pointed out an important strike-slip component of the northern and southern

767

Alhamilla-Cabrera faulted boundaries during the Tortonian [Martínez-Martínez et al., 2006;

768

Rutter et al., 2012] leading to drainage captures [Mather, 2000 ; Giaconia et al., 2012b, 2013].

769

In addition, several authors suggested that strike-slip tectonics parallel to the E-W depocenter

770

have been important in the evolution of the Sorbas basin in particular during the Tortonian

771

turbiditic filling [Ott d´Estevou and Montenat, 1990; Haughton, 2001]. This assumption is

772

mainly based on the occurrence of contained turbidites and a reversal location of its ponding

773

depocenter in the late Tortonian [Haughton, 2001]. However, the initial sinistral strike-slip

774

movement advanced by the authors cannot be observed in the basin due to its intense late

775

deformation. The associated primary east-flowing sedimentation could result from the first-step

776

extension highlighted in this study. Moreover, the tilted-block geometry inferred by Haughton

777

[2001] seems unrealistic with respect to the 2D gravity inversion.

778

[39] Compared to the Huércal-Overa basin, inversion tectonics appears much stronger in the

779

Sorbas basin [Weijermars et al., 1985; Augier et al., 2013]. Onset of regional inversion has been

780

proposed to approximately occur before the Tortonian/Messinian boundary (e.g. ~7.7 Ma)

781

[Weijermars et al., 1985]. Besides, inversion tectonics affecting the northern flank of the Sierra

782

de Los Filabres occurred roughly at the same time, around 8.2 Ma [Augier, 2004]. This period

783

coincides in a broad sense with a period of geodynamic reorganisations of the southeastern

784

Betics (Figure 12B) as described throughout the Alboran region, around 8 Ma [Comas et al.,

785

1999; Jolivet et al., 2006].

786

[40] At the Tortonian-Messinian transition, the basin experienced an overall uplift and erosion as

787

attested by the deposition of the Azagador shallow marine calcarenite over the upper Tortonian

788

(Figure 4 and 13C). The large open-fold syncline of the lowermost Messinian calcarenite shows

789

the relatively limited shortening deformation affecting the basin after its deposition (Figure 13B).

790

From that period, and despite a global sea-level fall of 35 m [Haq et al., 1987], deep marine

791

marls were deposited into the Sorbas basin between 7.3 and 6.5 Ma (Figure 13B). The

792

subsequent deposition of Messinian clays and diatomites, known as the Upper Abad Member,

793

and of the coeval fringing reef, in a restricted marine environment, is once more coeval with a

794

global sea-level fall estimated at 60 to 70 m (Figure 10) [Haq et al., 1987; Miller et al., 2005].

795

These features highly suggest that the global eustatic variations have locally weaker effects on

796

sedimentation than the tectonic control during the Sorbas basin tectonic inversion.

797 798

9. Conclusions

799

[41] While most of the previous studies on Sorbas basin focused on the Messinian salinity crisis

800

event or on compressive tectonics acting from middle- to upper-Tortonian times, this work aims

801

at constraining the early tectonic context during the initiation phase of the basin. Two principal

802

targets were identified as key points to address this problem: the geometry/structure of the early

803

deposits with the basement units and the internal deformation/stress fields taking place during

804

the onset and first stages of sedimentation into the basin. At first, structural analysis and

805

compilation point out that the well-known northward thrusting movement observed along the

806

northern edge of the Sierra Alhamilla may be limited; deformation being characterized by

807

pervasive shearing and folding in some places. Besides, structural analysis of the earlier deposits

808

along the southern margin of the basin, including computation of palaeostress tensors from fault

809

kinematics analysis, highlights the prime effect of an initial extensional event during the onset of

810

basin development. This phase of deformation observed in sedimentary series is coeval with the

811

late exhumation stages of the Nevado-Filabride complex, further south, into the Sierra Alhamilla,

812

which is characterized by a top-to-the NW normal sense of shear [Augier et al., 2005a]. This

813

extensional context must probably controlled the inception of sedimentation in the continental

814

basin.

815

[42] On the other hand, gravity modeling revealed the deep and asymmetric structure of the

816

Sorbas basin. This asymmetry is highlighted both by a gently south-dipping basal unconformity

817

in the north and a steep south-dipping southern contact. Gravimetric survey and modelling

818

results in the determination of a maximum ~2500 m of sediments accumulated and preserved

819

along the southern margin of the basin. The Sorbas basin thus displays as a particularly thick

820

basin in comparison to the other neighbouring continental basins of the area and its well-marked

821

asymmetry could hardly be explained by the limited thrust movements as observed along the

822

northern Sierra Alhamilla. All these results thus argue for a two stages evolution of the Sorbas

823

basin development with (i) extension driving at first the initiation and first stages of subsidence

824

and (ii) inversion of the basin and flexural control of the sedimentation during a subsequent

825

compression.

826

[43] Our integrated study allows estimating the evolution through time of the basin behaviour.

827

We propose that during the early extensional stage, a normal to normal-strike-slip fault system

828

initiated the subsidence along the northern edge of the metamorphic range, south of the basin,

829

during the last stages of exhumation of the metamorphic units. Later on, the asymmetry of the

830

basin was reinforced during the basin inversion by the N-S compressional event occuring from

831

ca. 8 Ma. Since the deposition of the Azagador Member, the Sorbas basin evolved as a flexural

832

basin, resulting from the ongoing overall N-S shortening (acting since ca.~8 Ma). This tectonic

833

setting is marked by the strengthening of the subsidence and leads to the deposition of deep

834

marine sediments (Abad marls). Then, Messinian and post-Messinian deposits are only slightly

835

folded (Figure 13).

836 837

[43] Acknowledgments

838

We warmly thank Georges Clauzon without whom this study would not have been achieved. He

839

provided us a new vision of the evolution of the Sorbas basin based on thorough

840

geomorphological and biostratigraphic researches. He could not see the achievement of this

841

study but his memory will be forever associated to the Sorbas basin.

842

Field investigations have been supported by the CIFRE PhD grant N° 584/2010

843

(TOTAL/UPMC). This paper is a contribution to the Projects AMEDITER (“Actions Marges”

844

CNRS/INSU Programme), “Bassins néogènes et manteau en Méditerranée” (TerMEx

845

CNRS/INSU Programme). We are grateful to Michel Diament (INSU/IPGP) for providing us a

846

SCINTREX CG5-M micro-gravimeter. Geophysical interpretations were drawn using Geosoft

847

software and the 3D modeling process was performed using the 3D Geomodeller (Intrepid

848

Geophysics) developed by the BRGM (Orléans, France). We are grateful to Guillaume Martelet

849

(BRGM) for his help and assistance on the gravity modeling process and also to Gabriel

850

Courrioux (BRGM-GEO) for providing us a 3D Geomodeller licensing and for his technical

851

support. We thank two anonymous referees for their constructive contribution to improving the

852

paper.

853 854

References

855

Anderson, E.M. (1942), The dynamics of faulting and dyke formation with application to Britain.

856

Oliver and Boyd (Eds.), Edinburgh.

857

Amores, L.R., Hernandez-Enrile J.L. and Martínez-Díaz, J.J. (2001), Sobre los factores

858

relacionados con la evaluación de la peligrosidad sísmica en la región de Murcia. In:

859

Secundo Congreso Iberoamericano de Ingeniería Sísmica, Madrid, Spain, Asociación

860

Española de Ingeniería Sísmica.

861

Amores, L.R., Hernandez-Enrile J.L. and Martínez-Díaz, J.J. (2002), Estudio gravimétrico previo

862

aplicado a la identificación de fallas ocultas como fuentes sismogenéticas en la depresión del

863

Guadalentín (región de Murcia), Geogaceta, 32, 307-310.

864

Angelier, J. (1984), Tectonic analysis of fault slip data sets, J. Geophys. Res., 89, 5835–5848.

865

Angelier, J. (1990), Inversion of field data in fault tectonics to obtain the regional stress: a new

866 867 868 869 870

rapid direct inversion method by analytical means, Geoph. J. Int., 103, 363–376. Angelier, J. (1994), Fault slip analysis and palaeostress reconstruction, in Continental Deformation, 53–100, Hancock, L., Pergamon Press (Eds.), Oxford. Augier, R. (2004), Evolution tardi-orogénique des Cordilères Bétiques (Espagne): Apports d'une étude intégrée, Ph.D. thesis, Université Pierre et Marie Curie, Paris.

871

Augier, R., Jolivet, L., and Robin, C. (2005a), Late Orogenic doming in the eastern Betic

872

Cordilleras: Final exhumation of the Nevado-Filabride complex and its relation to basin

873

genesis, Tectonics, 24, TC4003.

874

Augier, R., Agard, , Monié, , Jolivet, L., Robin, C. and Booth-Rea, G. (2005b), Exhumation,

875

doming and slab retreat in the Betic Cordillera (SE Spain): in situ 40Ar/39Ar ages and P–T–

876

d–t paths for the Nevado-Filabride complex, J. Metam. Geol., 23, 357–381.

877

Augier, R., Jolivet, L., Do Couto, D., and Negro, F. (2013), From ductile to brittle, late- to post-

878

orogenic evolution of the Betic Cordillera: Structural insights form the North-eastern

879

Internal Zones, Bull. Soc. Geol. Fr., 184(4-5), 405-425.

880

Azañón, J.M., and Crespo-Blanc, A. (2000), Exhumation during a continental collision inferred

881

from the tectonometamorphic evolution of the Alpujarride Complex in the central Betics

882

(Alboran Domain, SE Spain), Tectonics, 19, 549–565.

883 884

Behr, W. M., and Platt, J.P. (2012), Kinematic and thermal evolution during two-stage exhumation of a Mediterranean subduction complex, Tectonics, 31(4), TC4025.

885

Booth-Rea, G., Azañón, J.M., and García-Dueñas, V. (2004a), Extensional tectonics in the

886

northeastern Betics (SE Spain): case study of extension in a multilayered upper crust with

887

contrasting rheologies, J. Struct. Geol., 26, 2039–2058.

888

Booth-Rea, G., Azañón, J.M., Azor, A., and García-Dueñas, V. (2004b), Influence of strike-slip

889

fault segmentation on drainage evolution and topography. A case study: the Palomares Fault

890

Zone (southeastern Betics, Spain), J. Struct. Geol., 26, 1615–1632.

891

Bourillot, R., Vennin, E., Rouchy, J.M., Durlet, C., Rommevaux, V., Kolodka, C., and Knap, F.

892

(2009), Structure and evolution of a Messinian mixed carbonatesiliciclastic platform: the

893

role of evaporites (Sorbas Basin, South-east Spain), Sedimentology, 57, 477–512.

894

Braga, J.C., Bassi, D., Martín, J.M., Riding, R., Aguirre, J., Sánchez-Almazo, I.M., and Dinarès-

895

Turell J. (2006), Testing models for the Messinian salinity crisis: The Messinian record in

896

Almería, SE Spain, Sed. Geol., 188-189, 131–154.

897 898

Braga, J.C., Martín, J.M. and Quesada, C. (2003), Patterns and average rates of late Neogene– Recent uplift of the Betic Cordillera, SE Spain, Geomorphology, 50, 3–26.

899

Briend, M., Montenat, C., and Ott d'Estevou, P. (1990), Le bassin de Huercal-Overa, in Les

900

bassins Néogènes du domaine Bétique Oriental (Espagne), Documents et Travaux IGAL, C.

901

Montenat (Ed.), 12-13, 239–260.

902

Calcagno, P., Chilès, J.P., Courrioux, G., and Guillen, A. (2008), Geological modelling from

903

field data and geological knowledge: Part I. Modelling method coupling 3D potential-field

904

interpolation and geological rules, Phys. Earth Planet. Inter., 171(1–4), 147-157.

905 906 907 908

CIESM, 2008, The Messinian Salinity Crisis from mega-deposits to microbiology - A consensus report, no. 33, in CIESM Workshop Monographs, F.Briand (Ed.), pp. 168. Clauzon, G., Suc, J., Gautier, F., Berger, A. and Loutre, M.F. (1996), Alternate interpretation of the Messinian salinity crisis: Controversy resolved?, Geology, 24, 363–366.

909

Clauzon, G., Suc, J.-, Do Couto, D., Jouannic, G., Melinte-Dobrinescu, M.C., Jolivet, L., Lebret,

910

N., Mocochain, L., Popescu, S.-M., Martinell, J., Doménech, R., Rubino, J.-L., Gumiaux, C.,

911

Warny, S., Bellas, S.M., Gorini, C., Bache, F., Rabineau, M., and Estrada, F. (revised), New

912

insights on the Sorbas Basin (SE Spain): the onshore reference of the Messinian Salinity

913

Crisis, Bas. Res.

914

Cloetingh, S., van der Beek, A., van Rees, D., Roep, T.B., Biermann, C., and Stephenson R.A.

915

(1992), Flexural interaction and the dynamics of neogene extensional Basin formation in the

916

Alboran-Betic region, Geo-Marine Letters, 12, 66–75.

917 918

Comas, M.C., Platt, J., Soto, J.I. and Watts, A.B. (1999), The origin and tectonic history of the Alboran basin: insights from leg 161 results, Ocean Drilling Program, 161, 555–580.

919

Crespo-Blanc, A. (1995), Interference pattern of extensional fault systems: a case study of the

920

Miocene rifting of the Alboran basement (North of Sierra Nevada, Betic Chain), J. Struct.

921

Geol., 17, 1559–1569.

922

Crespo-Blanc, A., and Frizon de Lamotte, D. (2006), Structural evolution of the external zones

923

derived from the Flysch trough and the South Iberian and Maghrebian paleomargins around

924

the Gibraltar arc: a comparative study, Bull. Soc. Geol. Fr., 177, 267–282.

925

Crespo-Blanc, A., Orozco, M., and García-Dueñas, V. (1994), Extension versus compression

926

during the Miocene tectonic evolution of the Betic chain. Late folding of normal fault

927

systems, Tectonics, 13, 78–88.

928

de Jong, K. (1991), Tectono-metamorphic studies and radiometric dating in the Betic Cordilleras

929

(SE Spain) - with implications for the dynamics of extension and compression in the western

930

Mediterranean Area, Ph.D. thesis, Vrije University, Amsterdam.

931 932

de Jong, K. (2003), Very fast exhumation of high-pressure metamorphic rocks with excess 40Ar and inherited 87Sr, Betic Cordilleras, southern Spain, Lithos, 70, 91-110.

933

de Larouzière, F.D., Bolze, J., Bordet, , Hernandez, J., Montenat, C. and Ott d'Estevou, P.

934

(1988), The Betic segment of the lithospheric Trans-Alboran shear zone during the Late

935

Miocene, Tectonophysics, 152, 41–52.

936

Díaz, J., and Gallart, J. (2009), Crustal structure beneath the Iberian Peninsula and surrounding

937

waters: A new compilation of deep seismic sounding results, Phys. Earth Planet. Inter.,

938

173(1-2), 181-190.

939 940 941 942 943 944

Dronkert, H. (1976), Late Miocene evaporites in the Sorbas Basin and adjoining aeras, Mem. Soc. Geol. Italiana, 16, 341–361. Faccenna, C., and Becker, T. W. (2010), Shaping mobile belts by small-scale convection, Nature, 465(7298), 602-605. Faccenna, C., Mattei, M., Funiciello, R., and Jolivet, L. (1997), Styles of back-arc extension in the Central Mediterranean, Terra Nova, 9, 126-130.

945 946

Faccenna, C., Piromallo, C., Crespo-Blanc, A., Jolivet, L., and Rossetti, F. (2004), Lateral slab deformation and the origin of the western Mediterranean arcs, Tectonics, 23, TC1012.

947

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

948

Ringenbach, J.C. (2011), The southernmost margin of the Tethys realm during the Mesozoic

949

and Cenozoic: Initial geometry and timing of the inversion processes, Tectonics, 30(3),

950

TC3002.

951

Fullea, J., Fernandez, M., and Zeyen, H. (2006), Lithospheric structure in the Atlantic–

952

Mediterranean transition zone (southern Spain, northern Morocco): a simple approach from

953

regional elevation and geoid data, C.R. Geosciences, 338, 140–151.

954

Fullea, J., Fernàndez, M., Afonso, J.C., Vergés, J., and Zeyen, H. (2010), The structure and

955

evolution of the lithosphere–asthenosphere boundary beneath the Atlantic–Mediterranean

956

Transition Region, Lithos, 120(1–2), 74–95.

957

Galindo-Zaldívar, J., González-Lodeiro, F., and Jabaloy, A. (1989), Progressive extensional

958

shear structures in a detachment contact in the Western Sierra Nevada (Betic Cordilleras,

959

Spain), Geodinamica Acta, 3, 73–85.

960

Galindo-Zaldívar, J., Gil, J.A., Borque, M.J., González-Lodeiro, F., Jabaloy, A., Marín-Lechado,

961

C., Ruano, , and Sanz de Galdeano, C. (2003), Active faulting in the internal zones of the

962

central Betic Cordilleras (SE, Spain), J. Geodynamics, 36, 239–250.

963

García-Dueñas, V., Balanyá, J.C., and Martínez-Martínez J.M. (1992), Miocene extensional

964

detachments in the outcropping basement of the northern Alboran Basin (Betics) and their

965

tectonic implications, Geo-Marine Letters, 12, 88–95.

966

García-Dueñas, V., Martínez-Martínez, J.M., Orozco, M. and Soto, J.I. (1988), Plis-nappes,

967

cisaillements syn- à post-métamorphiques et cisaillements ductiles-fragiles en distension

968

dans les Nevado-Filabrides (Cordillères bétiques, Espagne), C. R. Acad. Sci., 307, 1389–

969

1395.

970

García Monzón, G., and Kampschuur, W. (1972), Mapa geológico de España, Hoja 1014, Vera,

971

Instituto Geológico y Minero de España (IGME), Ministerio de Industria y Energía, Madrid,

972

scale 1:50000.

973

García Monzón, G., G., Kampschuur, W., and Verburg, J. (1974), Mapa geológico de España,

974

Hoja 1031, Sorbas, Instituto Geológico y Minero de España (IGME), Ministerio de Industria

975

y Energía, Madrid, scale 1:50000.

976

García Monzón, G., G., Kampschuur, W., and Vissers, R.L.M. (1973a), Mapa geológico de

977

España, Hoja 1013, Macael, Instituto Geológico y Minero de España (IGME), Ministerio de

978

Industria y Energía, Madrid, scale 1:50000.

979

García Monzón, G., G., Kampschuur, W., Vissers, M., R.L., Verburg, J., and Wolff, R. (1973b),

980

Mapa geológico de España, Hoja 1030, Tabernas, Instituto Geológico y Minero de España

981

(IGME), Ministerio de Industria y Energía, Madrid, scale 1:50000.

982

Giaconia, F., Booth-Rea, G., Martínez-Martínez, J.M., Azanon, J.M., Pérez-Peña, J.V., Pérez-

983

Romero, J., and Villegas, I. (2012a), Geomorphic evidence of active tectonics in the Sierra

984

Alhamilla (eastern Betics, SE Spain), Geomorphology, 145, 90-106.

985

Giaconia, F., Booth-Rea, G., Martínez-Martínez, J. M., Azañón, J. M., and Pérez-Peña, J.

986

(2012b), Geomorphic analysis of the Sierra Cabrera, an active pop-up in the constrictional

987

domain of conjugate strike-slip faults: The Palomares and Polopos fault zones (eastern

988

Betics, SE Spain), Tectonophysics, 580, 27–42.

989

Giaconia, F., Booth-Rea, G., Martínez-Martínez, M.J., Azañón, M.J., Pérez-Romero, J., and

990

Villegas, I. (2013), Mountain front migration and drainage captures related to fault segment

991

linkage and growth: The Polopos transpressive fault zone (southeastern Betics, SE Spain),

992

Journal of Structural Geology, 46, 76-91.

993

Gracia, E., Pallàs, R., Soto, J.I., Comas, M., Moreno, X., Masana, E., Santanach, P., Diez, S.,

994

García, M., Dañobeitia, J. and HITS scientific party (2006), Active faulting offshore SE

995

Spain (Alboran Sea): Implications for earthquake hazard assessment in the Southern Iberian

996

Margin, Earth and Planetary Science Letters, 241(3-4), 734-749.

997

Guillen, A., Calcagno, P., Courrioux, G., Joly, A., and Ledru, P. (2008), Geological modelling

998

from field data and geological knowledge: Part II. Modelling validation using gravity and

999

magnetic data inversion, Phys. Earth Planet. Inter., 171(1–4), 158-169.

1000

Hammer, S. (1939), Terrain corrections for gravimeter stations, Geophysics, 4, 184–194.

1001

Haq, B.U., Hardenbol, J., and Vail, R. (1987), Chronology of Fluctuating Sea Levels Since the

1002 1003 1004 1005 1006

Triassic, Science, 235, 1156–1167. Haughton, P. (2000), Evolving turbidite systems on a deforming basin floor, Tabernas, SE Spain, Sedimentology, 47, 497-518. Haughton, P. (2001), Contained turbidites used to track sea bed deformation and basin migration, Sorbas Basin, south-east Spain, Bas. Res., 13, 117–139.

1007

Hodgson, D.M., and Haughton, P. (2004), Impact of syndepositional faulting on gravity current

1008

behaviour and deep-water stratigraphy: Tabernas-Sorbas Basin, SE Spain, in Confined

1009

Turbidite Systems, edited by Lomas, S. A. Joseph and P. (eds), pp. 135-158, Geological

1010

Society, London.

1011 1012

Iribarren, L., Vergés, J., Fernàndez, M. (2009), Sediment supply from the Betic-Rif orogen to basins through Neogene, Tectonophysics, 475, 68–84.

1013

Jabaloy, A., Galindo-Zaldívar, J., and González-Lodeiro, F. (1992), The Mecina Extensional

1014

System: Its relation with the post-Aquitanian piggy-back Basins and the paleostresses

1015

evolution (Betic Cordilleras, Spain), Geo-Marine Letters, 12, 96–103.

1016 1017

Jacoby, W., and Smilde, L. (2009), Gravity Interpretation - Fundamentals and Application of Gravity Inversion and Geological Interpretation, Springer, Heidelberg.

1018

Johnson, C., Harbury, N., and Hurford, A.J. (1997), The role of extension in the Miocene

1019

denudation of the Nevado-Filabride Complex, Betic Cordillera (SE Spain), Tectonics, 16,

1020

189–204.

1021 1022

Jolivet, L., and Faccenna, C. (2000), Mediterranean extension and the Africa-Eurasia collision, Tectonics, 19(6), 1095-1106.

1023

Jolivet, L., Augier, R., Faccenna, C., Negro, F., Rimmele, G., Agard, P., Robin, C., Rossetti, F.,

1024

and Crespo-Blanc, A. (2008), Subduction, convergence and the mode of backarc extension

1025

in the Mediterranean region, Bull. Soc. Geol. Fr., 179, 525–550.

1026 1027

Jolivet, L., Augier, R., Robin, C., Suc, J. and Rouchy, J.M. (2006), Lithospheric-scale geodynamic context of the Messinian salinity crisis, Sed. Geol., 188–189, 9–33.

1028

Joly, A., Faure, M., Martelet, G., and Chen Yan (2009), Gravity inversion, AMS and

1029

geochronological investigations of syntectonic granitic plutons in the southern part of the

1030

Variscan French Massif Central, J. Struct. Geol., 31, 421–443.

1031

Jonk, R., and Biermann, C. (2002), Deformation in Neogene sediments of the Sorbas and Vera

1032

Basins (SE Spain): constraints on simple-shear deformation and rigid body rotation along

1033

major strike-slip faults, J. Struct. Geol., 24, 963–977.

1034 1035

Kleverlaan, K. (1989), Neogene history of the Tabernas basin (SE Spain) and its Tortonian submarine fan development, Geol. Mijnbouw, 68, 421–432.

1036

Krijgsman, W., Fortuin, A.R., Hilgen, F.J., and Sierro, F.J. (2001), Astrochronology for the

1037

Messinian Sorbas basin (SE Spain) and orbital (precessional) forcing for evaporite cyclicity,

1038

Sed. Geol., 140, 43–60.

1039 1040 1041 1042

Leblanc, D., and Olivier, P. (1984), Role of strike-slip faults in the Betic-Rifian Orogeny, Tectonophysics, 101, 345–355. Lefort, J.P., and Agarwal B.N. (1999), Of what is the center of the Ibero-Armorican arc composed?, Tectonophysics, 302, 71–81.

1043

Li Qiao, Ruano, P., Pedrera, A., and Galindo- Zaldívar, J. (2012), Estructura de la cuenca

1044

sedimentaria de Tabernas-Sorbas mediante prospección gravimétrica y magnética (Zonas

1045

Internas, Cordillera Bética Oriental), Geogaceta, 52, 117–120.

1046

Lonergan, L., and Platt, J. (1995), The Malaguide-Alpujarride boundary: a major extensional

1047

contact in the Internal Zone of the eastern Betic Cordillera, SE Spain, J. Struct. Geol., 17,

1048

1655–1671.

1049 1050

Lonergan, L., and White, N. (1997), Origin of the Betic-Rif mountain belt, Tectonics, 16(3), 504522.

1051

López Sánchez-Vizcaíno, V., Rubatto, D., Gómez-Pugnaire, M.T., Trommsdorff, V., and

1052

Müntener, O. (2001), Middle Miocene high-pressure metamorphism and fast exhumation of

1053

the Nevado-Filábride complex, SE Spain, Terra Nova, 13, 327-332.

1054

Manzi, V., Gennari, R., Hilgen, F., Krijgsman, W., Lugli, S., Roveri, M., and Sierro F.J. (2013),

1055

Age refinement of the Messinian salinity crisis onset in the Mediterranean, Terra Nova,

1056

25(4), 315-322.

1057

Martelet, G. (1999). Modélisation de la structure crustale et du comportement mécanique de la

1058

lithosphère à partir des anomalies gravimétriques. Applications à l’Himalaya et au massif

1059

granitique du Mont Lozère, Cévennes, Ph.D. thesis, Institut de Physique du Globe, Paris.

1060

Martelet, G., Calcagno, P., Gumiaux, C., Truffert, C., Bitri, A., Gapais, D. and Brun, J.P. (2004),

1061

Integrated 3D geophysical and geological modeling of the Hercynian Suture Zone in the

1062

Champtoceaux area (south Brittany, France), Tectonophysics, 382, 117–128.

1063 1064

Martelet, G., Debeglia, N., Truffert, and C. (2002), Updating and validating the French gravity terrain corrections out to a distance of 167 km, C. R. Geosci., 334(7), 449-454.

1065

Martín, J.M., Braga, J.C., and Sánchez-Almazo, I.M. (1999), The Messinian record of the

1066

outcropping marginal Alboran basin deposits: significance and implications, in Proceedings

1067

of the Ocean Drilling Program, Scientific Results, 161, 543–551.

1068 1069

Martínez-Martínez, J.M. (1985), Las Sucesiones Nevado-Filábrides en la Sierra de los Filabres y Sierra Nevada, Cuadernos de Geología Ibérica, 12, 127–144.

1070

Martínez-Martínez, J.M., and Azañón, J.M. (1997), Mode of extensional tectonics in the

1071

southeastern Betics (SE Spain): Implications for the tectonic evolution of the peri-Alboran

1072

orogenic system, Tectonics, 16(2), 205–225.

1073

Martínez-Martínez, J.M., Soto, J.I., and Balanyá, J.C. (2002), Orthogonal folding of extensional

1074

detachments: Structure and origin of the Sierra Nevada elongated dome (Betics, SE Spain),

1075

Tectonics, 21, TC001283.

1076

Martínez-Martínez, J. M., Soto, J. I., and Balanyá, J. C. (2004), Elongated domes in extended

1077

orogens: A mode of mountain uplift in the Betics (southeast Spain), in Gneiss Domes in

1078

Orogeny: Boulder, Geological Society of America Special Paper 380, edited by Whitney,

1079

D.L, Teyssier, C., and Siddoway, C.S., pp. 243-266.

1080

Martínez-Martínez, J.M., Booth-Rea, G., Azañon, J.M., and Torcal, F. (2006), Active transfer

1081

fault zone linking a segmented extensional system (Betics, southern Spain): Insight into

1082

heterogeneous extension driven by edge delamination, Tectonophysics, 422, 159-173.

1083

Martín-Suárez, E., Freudenthal, M., Krijgsman, W., and Rutger Fortuin, A. (2000), On the age of

1084

the continental deposits of the Zorreras Member (Sorbas Basin, SE Spain), Geobios, 33, 505–

1085

512.

1086 1087

Mather, A.E. (2000), Adjustment of a drainage network to capture induced base-level change: an example from the Sorbas Basin, SE Spain, Geomorphology, 34, 271-289.

1088

Mauffret, A., Maldonado, A. and Campillo, A.C. (1992), Tectonic framework of the eastern

1089

Alboran and western Algerian Basins, western Mediterranean, Geo-Marine Letters, 12, 104–

1090

110.

1091

Meijninger, B.M.L. and Vissers, R.L.M. (2006), Miocene extensional basin development in the

1092

Betic Cordillera, SE Spain revealed through analysis of the Alhama de Murcia and

1093

Crevillente Faults, Bas. Res., 18, 547–571.

1094

Miller, K.G., Kominz, M.A., Browning, J., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman,

1095

J., Cramer, B.S., Christie-Blick, N. and Pekar, S.F. (2005), The Phanerozoic Record of Global

1096

Sea-Level Change, Science, 310, 1293–1298.

1097

Monié, P., Torres-Roldán, R.L., and García-Casco, A. (1994), Cooling and exhumation of the

1098

Western Betic Cordilleras, 40Ar/39Ar thermochronological constraints on a collapsed terrane,

1099

Tectonophysics, 238, 353–379.

1100 1101

Monié, P. and Chopin, C. (1991),

40

Ar-39Ar ages through high-pressure low-temperature

metamorphism in the Western Alps, European Journal of Mineralogy, 2, 343–361.

1102

Montenat, C., Ott d'Estevou, , Plaziat, J.C., and Chapel, J. (1980), La signification des faunes

1103

marines contemporaines des évaporites messiniennes dans le Sud-Est de l'Espagne.

1104

Conséquences pour l'interprétation des conditions d'isolement de la Méditerranée

1105

occidentale, Géologie Méditerranéenne, 7, 81–90.

1106

Montenat, C., Ott d´Estevou, P., and Masse, P. (1987), Tectonic-Sedimentary characters of the

1107

Betic Neogene Basins evolving in a crustal transcurrent shear zone (SE Spain), Bull. C. R.

1108

Explo. Elf-Aquitaine, 11, 1–22.

1109

Ott d'Estevou, P., and Montenat, C. (1990), Le bassin de Sorbas-Tabernas, in Les bassins

1110

Néogènes du domaine Bétique Oriental (Espagne), Documents et Travaux IGAL, C.

1111

Montenat (Ed.), 12-13, 101–128.

1112

Ott d'Estevou, P., Montenat, C., and Alvado J.-C. (1990), Le bassin de Vera-Garrucha, in Les

1113

bassins Néogènes du domaine Bétique Oriental (Espagne), Documents et Travaux IGAL, C.

1114

Montenat (Ed.), 12-13, 165–187.

1115

Pedrera, A., Galindo-Zaldívar, J., Ruíz-Constán, A., Duque, C., Marín-Lechado, C., and Serrano,

1116

I. (2009), Recent large fold nucleation in the upper crust: Insight from gravity, magnetic,

1117

magnetotelluric and seismicity data (Sierra de Los Filabres–Sierra de Las Estancias, Internal

1118

Zones, Betic Cordillera), Tectonophysics, 463, 145–160.

1119

Pedrera, A., Galindo-Zaldívar, J., Tello, A. and Marín-Lechado, C. (2010), Intramontane basin

1120

development related to contractional and extensional structure interaction at the termination

1121

of a major sinistral fault: The Huércal-Overa Basin (Eastern Betic Cordillera), J.

1122

Geodynamics, 49, 271–286.

1123

Pedrera, A., Galindo-Zaldívar, J., Lamas, F., and Ruiz-Constán, A. (2012), Evolution of near-

1124

surface ramp-flat-ramp normal faults and implication during intramontane basin formation

1125

in the eastern Betic Cordillera (the Huércal-Overa Basin, SE Spain), Tectonics, 31, TC4024.

1126

Platt, J., Van Den Eeckhout, B., Janzen, E., Konert, G., Simon, O. J., and Weijermars, R. (1983),

1127

The structure and tectonic evolution of the Aguilón fold-nappe, Sierra Alhamilla, Betic

1128

Cordilleras, SE Spain, Journal of Structural Geology, 5(5), 519–538.

1129 1130

Platt, J., and Vissers, R.L.M. (1989), Extensional collapse of thickened continental lithosphere: A working hypothesis for the Alboran Sea and Gibraltar Arc, Geology, 17, 540–543.

1131

Platt, J., Allerton, S., Kirker, A.I., Mandeville, C., Mayfield, A., Platzman, E., and Rimi, A.

1132

(2003), The ultimate arc: Differential displacement, oroclinal bending, and vertical axis

1133

rotation in the External Betic-Rif arc, Tectonics, 22, 1017.

1134

Platt, J.P., Anczkiewicz, R., Soto, J.I., Kelley, S.P., and Thirlwall, M.F. (2006), Early Miocene

1135

continental subduction and rapid exhumation in the western Mediterranean, Geology,

1136

34(11), 981-984.

1137

Platt, J., Kelley, S., Carter, A. and Orozco, M. (2005), Timing of tectonic events in the

1138

Alpujarride Complex, Betic Cordillera, southern Spain, J. Geol. Soc. London, 162, 1–12.

1139

Poisson, A.M., Morel, J.L., Andrieux, J., Coulon, M., Wernli, R., and Guernet, C. (1999), The

1140

origin and development of Neogene Basins in the SE Betic Cordillera (SE Spain): A case

1141

study of the Tabernas-Sorbas and Huércal Overa Basins, Journal of Petroleum Geology,

1142

22(1), 97-114.

1143 1144 1145 1146

Puga-Bernabéu, A., Martin, J.M., and Braga, J.C. (2007), Tsunami-related deposits in temperate carbonate ramps, Sorbas Basin, southern Spain, Sedimentary Geology, 199, 107-127. Rehault, J-, Boillot, G., and Mauffret, A. (1984), The Western Mediterranean Basin geological evolution, Marine Geology, 55, 447–477.

1147

Riding, R., Braga, J.C., and Martín, J.M. (2000), Late Miocene Mediterranean desiccation:

1148

topography and significance of the ‘Salinity Crisis’ erosion surface on-land in southeast

1149

Spain: Reply, Sed. Geol., 133, 175–184.

1150

Riding, R., Braga, J.C., Martin, J.M., and Sánchez-Almazo, I.M. (1998), Mediterranean

1151

Messinian Salinity Crisis: constraints from a coeval marginal basin, Sorbas, southeastern

1152

Spain, Marine Geology, 146, 1–20.

1153 1154

Robinson, E. and Çoruh, C. (1988), Basic Exploration Geophysics, Wiley, J. and Sons, (Eds.), New York.

1155

Roep, T.B., Dabrio, C.J., Fortuin, A.R., and Polo, M.D. (1998), Late highstand patterns of

1156

shifting and stepping coastal barriers and washover-fans (late Messinian, Sorbas Basin, SE

1157

Spain), Sed. Geol., 116, 27–56.

1158 1159

Rouchy, J.M. and Saint-Martin, J. (1992), Late Miocene events in the Mediterranean as recorded by carbonate-evaporite relations, Geology, 20, 629–632.

1160

Rousset, D., Bayer, R., Guillon, D., and Edel, B., J. (1993), Structure of the southern Rhine

1161

Graben from gravity and reflection seismic data (ECORS-DEKORP Program),

1162

Tectonophysics, 221, 135–153.

1163 1164 1165 1166

Roveri, M., Gennari, R., Lugli, S., and Manzi, V. (2009), The Terminal Carbonate Complex: the record of sea-level changes during the Messinian salinity crisis, GeoActa, 8, 63–78. Ruano, P., Galindo-Zaldívar, J., and Jabaloy, A. (2004), Recent Tectonic Structures in a Transect of the Central Betic Cordillera, Pure and Applied Geophysics, 161, 541–563.

1167

Rutter, E.H., Faulkner, D.R., and Burgess R. (2012), Structure and geological history of the

1168

Carboneras Fault Zone, SE Spain: Part of a stretching transform fault system, J. Struct.

1169

Geol., 42, 227–245.

1170

Salcher, B.C., Meurers, B., Smit, J., Decker, K., Hoelzel, M., and Wagreich, M. (2012), Strike-

1171

slip tectonics and Quaternary basin formation along the Vienna Basin fault system inferred

1172

from Bouguer gravity derivatives, Tectonics, 31, TC3004.

1173 1174 1175 1176

Sanz de Galdeano, C. (1985), Estructura del borde oriental de la Sierra de Gádor (zona Alpujárride, Cordilleras Béticas), Acta Geológica Hispánica, 20, 145–154. Sanz de Galdeano, C. and Vera, J.A. (1992), Stratigraphic record and palaeogeographical context of the Neogene basins in the Betic Cordillera, Spain, Bas. Res., 4, 21–36.

1177

Sanz de Galdeano, C., Shanov, S., Galindo-Zaldívar, J., Radulov, A., and Nikolov, G. (2010), A

1178

new tectonic discontinuity in the Betic Cordillera deduced from active tectonics and

1179

seismicity in the Tabernas Basin, J. Geodynamics, 50, 57–66.

1180

Sierro, F.J., Flores, J.A., Civis, J., González Delgado, J.A., and Francés, G. (1993), Late Miocene

1181

globorotaliid event-stratigraphy and biogeography in the NE-Atlantic and Mediterranean,

1182

Marine Micropal., 21, 143–167.

1183

Spakman, W., and R. Wortel (2004), A tomographic view on Western Mediterranean

1184

Geodynamics, in The TRANSMED Atlas, The Mediterranean Region from Crust to Mantle,

1185

edited by W. Cavazza, F. Roure, W. Spakman, G. M. Stampfli and P. Ziegler, pp. 31-52.

1186

Stapel, G., Moeys, R., and Biermann, C. (1996), Neogene evolution of the Sorbas basin (SE

1187

Spain) determined by paleostress analysis, Tectonophysics, 255, 291–305.

1188

Szafián, P., and Horváth, F. (2006), Crustal structure in the Carpatho-Pannonian region: insights

1189

from three-dimensional gravity modeling and their geodynamic significance, Int. J. Earth

1190

Sci., 95, 50–67.

1191

Talbot, J.Y., Martelet, G., Courrioux, G., Chen Yan, and Faure, M. (2004), Emplacement in an

1192

extensional setting of the Mont Lozère–Borne granitic complex (SE France) inferred from

1193

comprehensive AMS, structural and gravity studies, J. Struct. Geol., 26, 11–28.

1194 1195

Telford, W.M., Geldart, L. and Sheriff, R.E. (1990), Applied Geophysics, Cambridge University Press.

1196

Torne, M., Fernandez, M., Comas, M. C., and Soto, J. I. (2000), Lithospheric structure beneath

1197

the Alboran Basin: Results from 3D gravity modeling and tectonic relevance, J. Geophys.

1198

Res., 105, 3209–3228.

1199

Torres-Roldán, R.L. (1979), The tectonic subdivision of the Betic Zone (Betic Cordilleras,

1200

southern Spain); its significance and one possible geotectonic scenario for the westernmost

1201

Alpine Belt, American Journal of Science, 279, 19–51.

1202

Turrillot, P., Faure, M., Martelet, G., Chen, Y., and Augier, R. (2011), Pluton-dyke relationships

1203

in a Variscan granitic complex from AMS and gravity modeling. Inception of the

1204

extensional tectonics in the South Armorican Domain (France), J. Struct. Geol., 33, 1681–

1205

1698.

1206

Vázquez, M., Jabaloy, A., Barbero, L. and Stuart, F.M. (2011), Deciphering tectonic- and

1207

erosion-driven exhumation of the Nevado-Filabride Complex (Betic Cordillera, Southern

1208

Spain) by low temperature thermochronology, Terra Nova, 23, 257–263.

1209 1210 1211 1212 1213 1214

Vergés, J., and Fernàndez, M. (2012), Tethys-Atlantic Interaction along the Iberia-Africa Plate Boundary: The Betic-Rif Orogenic System, Tectonophysics, 579, 144–172. Vissers, R.L.M., Platt, J., and van der Wal, D. (1995), Late orogenic extension of the Betic Cordillera and the Alboran Domain: A lithospheric view, Tectonics, 14, 786–803. Watts, A.B., Platt, J. and Buhl, P. (1993), Tectonic evolution of the Alboran Sea basin, Bas. Res., 5, 153–177.

1215

Weijermars, R., Roep, T.B., van den Eeckhout, B., Postma, G., and Kleverlaan, K. (1985), Uplift

1216

history of a Betic fold nappe inferred from Neogene-Quaternary sedimentation and tectonics

1217

(in the Sierra Alhamilla and Almeria, Sorbas and Tabernas Basins of the Betic Cordilleras,

1218

SE Spain), Geol. Mijnbouw, 64, 397–411.

1219 1220 1221 1222

Wortel, M.J.R., and W. Spakman (2000), Subduction and slab detachment in the MediterraneanCarpathian region, Science, 290, 1910-1917. Ziegler, A., and Dèzes, P. (2006), Crustal evolution of Western and Central Europe, Mem. Geol. Soc. London, 32, 43–56.

1223 1224

1225

FIGURE CAPTIONS

1226 1227

Figure 1: Simplified structural map of Southeastern Spain. Located are the main metamorphic

1228

domes of the Sierra de los Filabres/Nevada and the Sierra Alhamilla/Cabrera and the main

1229

sedimentary basins modified after [Augier et al., 2005b]. Also indicated are the main tectonic

1230

contacts such as the Filabres and Alhamilla shear zones and the prominent sinistral fault zones

1231

pertaining to the Trans-Alboran transcurrent zone (CF : Carboneras Fault, PF : Palomares Fault

1232

et AMF : Alhama de Murcia FaultF) and smaller-scale structures.

1233 1234

Figure 2: Structure and stratigraphy of the Sorbas Basin.

1235

(A) Geological map of the Sorbas Basin, modified from [Ott d’Estevou and Montenat, 1990;

1236

García Monzón et al., 1974]. Projection UTM Zone 30 (WGS84 ellipsoid). (B) Synthetic N-S

1237

cross-section X-Y representing the main architecture of the Sorbas basin (location in Figure 2A).

1238

(C) Synthethic stratigraphic log of the Sorbas basin. Modified after [Montenat and Ott

1239

d’Estevou, 1977; Ott d’Estevou and Montenat, 1990; Krijgsman et al., 1999, 2001; Manzi et al.,

1240

2013; Clauzon et al., revised].

1241 1242

Figure 3: Simplified geological map showing the four main faults accommodating the

1243

deformation along the Sierra Alhamilla, Sierra Polopos and Sierra Cabrera: from East to West,

1244

the North Cabrera reverse fault (NCRF), the North Gafarillos fault (NGF), the South Gafarillos

1245

Fault (SGF) and the North Alhamilla reverse fault (NARF). Four cross-sections, orthogonal to

1246

the sediments/basement contact, present the geometry and architecture of the sedimentary units

1247

with respect to the metamorphic ridges (location of cross-sections in the geological map). See

1248

text for more details.

1249 1250

Figure 4: Evidences of upper Tortonian compressive tectonic.

1251

Panoramic view of the southern folded boundary of the Sorbas Basin East of Lucainena de las

1252

Torres (see location in Figure 2). Here, the metamorphic basement describes a north-verging

1253

asymmetric fold with an overturned steep-dipping forelimb. Tortonian series, separated from the

1254

basement by an unconformity, display a rather continuous growth-strata structure forming a

1255

recumbent syncline. Fault-slip data inversion indicates a N-S maximum compression direction

1256

[1].

1257 1258

Figure 5: Field evidences of extensional deformation affecting upper Serravallian to upper

1259

Tortonian sediments, (location on Figure 2). (A) Uninterpreted and interpreted panorama of

1260

lower Tortonian reddish conglomerates affected by two sets of small-scale normal faults: a

1261

dominant ca. N140E and a subordinate ca. N30E conjugate sets of normal faults. Inset (B) show

1262

a close-up view of syn-sedimentary normal faults affecting sandy to microconglomerate intervals

1263

and interrupted by a coarse-grained conglomerate. (C) Upper Tortonian marine sediments

1264

affected by a N140E oriented set of small-scale normal faults.

1265 1266

Figure 6: Geometrical relationships existing between the upper Serravallian-Tortonian

1267

extensional deformation and the upper Tortonian inversion.

1268

(A) Uninterpreted and interpreted panoramic view of the northern flank of the so-called Taberns

1269

fold [Ott d’Estevou and Montenat, 1990] showing (B) the presence of tilted normal faults

1270

affecting the lower Tortonian deposits and also (C) the upper Tortonian marine sediments. (D)

1271

The decomposition of the total heterogeneous fault population into homogeneous, cogenetic fault

1272

populations shows the presence of two subhorizontal extensional regimes with N50E and N140E

1273

minimum compression direction (3) that pre-date inversion tectonics visible at all scales form

1274

microfaults to large-scale folding.

1275 1276

Figure 7: (A) Simplified geological map showing gravimetric cross-sections realized across the

1277

basin. Red circle represents the reference station. (B) Location of the gravimetric stations and

1278

gravity anomaly plotted over the geological map of the Sorbas Basin. All measurement stations

1279

are represented by small circles colored in function of their relative Bouguer anomaly (see the

1280

color bar in the upper right corner). The dashed border rectangle represents the extension of the

1281

modeling box used to model in 3D the sediments/basement interface at depth.

1282 1283

Figure 8: 2D gravity profile treatment (A) Comparison between the measured total Bouguer

1284

anomaly (yellow diamond shape) and the regional gravity deviation exposed in Torne et al.

1285

[2000] of the N-S regional cross-section (SCS1). White diamond-shape points represent the

1286

calculated residual gravity anomaly profile (see text for details). (B) Superposition of the 2D

1287

gravity model anomaly (green curve), calculated from the density summarized in Table 1, with

1288

the measured total Bouguer anomaly (Black dots). The red curves correspond to the difference

1289

between the modeled gravity anomaly and the measured one. (C) Deconvolution of the measured

1290

gravity signal (black dashed line) showing the combined effects of the deep structures of the

1291

crust and more particularly the Moho geometry (red curve); the upper crust structure of the

1292

outcropping metamorphic units (blue curve) and the shape and thickness of the sedimentary

1293

basins (green curve). (D) Shallow field based cross-section and interpretation of the modeled 2D

1294

gravity anomaly across the Sorbas Basin. Note the strong asymmetry of the sedimentary

1295

depocenter.

1296 1297

Figure 9: Total gravity anomaly, shallow field-based cross-section and corresponding 2D model

1298

constructed for the 5 cross-sections called SCS3 to SCS7. The gravity modeling of these cross-

1299

sections is equal to that detailed in the text for the basin-scale SCS1 cross-section. Vertical

1300

exaggeration = 2, see Figure 7 for location.

1301 1302

Figure 10: Basin-floor 3D geometry. (A) Iso-depth curves of the base of the Tortonian deposits

1303

are plotted over the geological map of the modeled area. The depth is indicated in meters below

1304

sea-floor. (B) 3D oblique view of the base of the Sorbas Basin. Vertical exaggeration = 3.

1305 1306

Figure 11: Tectonic subsidence and uplift curves for the Sorbas and the Huércal-Overa basins

1307

during the Neogene (thick black and red curves, respectively). The subsidence curve of the

1308

Sorbas Basin is drawn from the average estimated paleobathymetries compiled in Table 2.

1309

Absolute eustatic curves are from Haq et al. [1987] and Miller et al. [2005]. Exhumation and

1310

cooling path of the Nevado-Filabrides complex is estimated from ZFT, AFT and U-Th/He

1311

methodologies [Johnson et al., 1997; Augier et al., 2005b; Platt et al., 2005; Vázquez et al.,

1312

2011].Grey rectangles noted A to E refer to steps of the evolutive cross-section of the Sorbas

1313

basin in Figure 13.

1314 1315

Figure 12: Simplified tectonic map showing average upper Serravallian-lower Tortonian

1316

extension directions and average upper Tortonian compressive directions of the southeastern part

1317

of the Internal Zones of the Betic Cordillera. Stretching direction and associated sense of shear

1318

for the Nevado-Filabride complex are synthetized from [Galindo-Zaldívar et al.,1989; Jabaloy et

1319

al.,1992; Augier et al., 2005a,b]. Extension and compression directions shown in the Huércal-

1320

Overa basin are derived from Meijninger [2006] and Augier et al. [2013].

1321 1322

Figure 13: Interpretative N-S cross-sections of the Sorbas basin representing its evolution from

1323

Present (A) to the deposition of Messinian evaporites (B), the Late Tortonian syntectonic

1324

sedimentation of turbidites (C) and the period of tectonic inversion (D) post-dating the extensive

1325

deformation during the upper Serravallian-lower Tortonian (E).

1326 1327

Table 1: Sampling and density measurements.

1328 1329 1330 1331 1332

Table 2: Palaeobathymetry and age of the sedimentary formations of the Sorbas Basin.

IBERIA

5°E

sin ir Ba

Lorca Lorca

Vélez Rubio

aF ur ci

Al ha m a

SIERRA DE LAS ESTANCIAS

Alger

Alboran Sea

de M

Malaga

TASZ

35°N Gharb Basin

2°00'W

2°30'W

e Bal

uiv

dalq

Gua

s

land

Is aric

TELL

Oran

RI F

37°30'

200 km

MOROCCO

Huercal Overa

Baza

s Fau lt

Almanzora corridor

Flysch nappes InternalGuadix Zones

mare

Foreland Basins External Zones

Vera

GRANADA

Palo

38°N



au lt

5°W

N

Fig. 2A

SIERRA DE LOS FILABRES

Alfaix Sorbas

SIERRA NEVADA

SIERRA CABRERA Tabernas

Trevenque idor

an corr

Alpujarr

SIERRA

37°00'N

LA

ALHAMIL

Nijar

SIERRA DE GADOR

r Ca

Murtas

LAS ALPUJARRAS

Polopos

bo

ne

r

as

Fa

ul

t

20 km

ALMERIA

Lujar

Motril Adra

Salobreña

Cabo de Gata

Campo de Dalias

ALBORAN SEA 3°30'

Major depocenter Sedimentary basins Volcanics

3°00'

Alpine metamorphic complexes Alpujarride Filabres & Alhamilla shear zones

Nevado-Filabride

2°30'

Main structures

2°00' W

Secondary structures

Strike-slip faults

Normal fault

Thrust Antiforms

Strike-slip fault Undifferenciated fault

La Mela

X

* *

*

* *

*

* * * * * * * * *

*

*

* * * * * ** * Cariatiz * * * * * * Los Martínez Martinez 4 * ** * * * * ** * * * Moras *

* * * *

Gochar Góchar

* * ** * * * * * * * *

Cortijo del Hoyo

Sorbas 2

2

*

3 5

Los Yesos *

*

Fig. 6

*

*

*

Lucainena de las Torres Turrillas

B

* *

Gacia Bajo

* * *

*

*

CFZ

Morrón de la *Cantona

*

8

Fig. 4

Cerrón de Hueli

Y

* * * * * * * *

* ** * * * *

Penas Peñas Negras 7

CFZ

13°

25°

Cariatiz reef

Moras (projected) 27°



Rio Aguas

16°

25°

85°

45° *

*

*

* *

*

400 200 0

0

1000

?

2000 m

Pliocene-Pleistocene conglomerates (Zorreras)

* *

Messinian reef (Cantera)

Turbidites (upper Tortonian)

Main structures

Messinian diatomites & marls (Abad)

Messinian gypsum (Yesares)

Tortonian-Messinian rich fossiliferous calcarenites (Azagador)

Upper Serravallian-lower Tortonian Breccias and conglomerates

CFZ: Cantona Fault Zone

Alpujarride nappe

Syncline axis

Névado-Filabride nappe

Anticline axis

Undifferenciated fault

?

Other Cities

Thrust

Pliocene marine sediments (Sorbas)

7.24

Serr. Tort. Basement

10

Morrón de la Cantona

800

5.97

Gafarillos

10

* *

5.60 5,60

Fig. 5

X

600

* *

*

Tabernas

1000

* 6

*

* ** *

4 100 000

* * *

*

5.46 5,46

1

*

*

*

Plio.

Tortonian

Casa Blanca

Q.

C

4 105 000

Sierra de los Filabres

N

El Chive

Messinian

9

Uleila Del Campo

Projection: UTM 30/WGS84 (coordinates in meters)

A

580 000 4 115 000

570 000

4 110 000

560 000

10

Sampling sites for density measurement Location of the presented outcrops

Y

Sierra Alhamilla

2°20’ W Pliocene

Alpujarrides detachment Nevado-Filabrides

Messinian gypsum Messinian

2° W

II

Thrust

Sorbas

Tabernas

Lucainena de las Torres

NARF

III’

Sierra Alhamilla

IV’

Pliocene continental sediments

Tortonian

Pliocene marine sediments

Upper Serravallian Lower Tortonian

Messinian gypsum

Basal messinian calcarenite

II

N

Sierra Polopos II’

SGF

37° N

Sierra Cabrera

Anticlinal Thrust

Nevado-Filabrides

Rio Aguas

20°

40°

I’

60°

NCRF

1000 m

Cariatiz

27°

Gafarillos



Rio Aguas

25°

15°

30°

21°

500 m

28°

NGF

I

Alpujarrides detachment

Messinian reef and marls

Mojàcar

Sierra Cabrera

500 m

IV

I’

NCRF

Strike-slip

El Alfaro

37°10’ N

III

Anticlinal

Tortonian Serravallian Tortonian

10°

I

II’

Sierra Polopos

NGF

18°

SGF

1000 m Sierra Los Filabres

Moralla

16°

15-20°

26°

20° 47°

53°

Sierra Alhamilla

CFZ 80°

III’

NARF

500 m

III

1000 m IV

IV’

500 m

El Alfaro 37°

1000 m

S

N

Sierra Alhamilla

S

N

Basement

Overturned Tortonian

Basement/Cover contact

Calar Alto

80° 75°

80°

Main stress axes

σ1 σ2 σ3 Strain axes Stretching N:27 faults 09 t-gashes

N:23 planes Fold axis: N086 06NE

Shortening

Messinian

86°

Fault nature Normal fault Sinistral fault Dextral fault Reverse fault pole of t-gashes pole of bedding Fold axis (inversion of bedding data)

90°

Cerro Los Llanos 60°

25°

20°

68°

S

N

B

S

A

N

0

3m

N:38 faults

N:06 faults

S

B

N:27 faults

C

N

0

1m

SE

NW

SE

NW

Tilted normal fault sets Tabernas fold NW-dipping fold flank B C 42° 24°

59°

A

Rambla de la Sierra Present geometry (Classified by fault nature)

Post-tilting faults

Present geometry Original geometry (restored) (Homogenous sets of fauts)

Folding

SE

NW

Early Tortonian

N:27 faults

N:23 planes Fold axis: N245 04SW

B S

Total population Present geometry

0

5m

N Quaternary terrace

N:09 faults

Pre-tilting faults

D

Late Tortonian N:34 faults

C

0

2m

±

2°20’W

2°W

±

Huercal-Overa

Gravity anomaly (mGal) +25

El Pilar

Uleila Del Campo

El Chive

SCS 1

S6

Filabres Sierra de Los

SC

Casa Blanca

La Mela

Vera

Los Martinez

SCS 3

37°20’N

SC

S

Gochar

5

Sorbas S

SCS 4

2

Sie

Campico

SCS

1

Tabernas Tabernas 37°N

La Herreria

SCS 7

SCS

ra

abre C a rr

lhamilla Sierra A

-17

Penas Negras

Lucainena de las Torres Turrillas 0

A

5

10

Km

B

0

5

10

Km

Sedimentary basins

Alpujarride nappe

Messinian and Pliocene

Alpujarride nappe

Miocene volcanic rocks

Nevado-Filabrides nappe

Tortonian

Nevado-Filabrides nappe

Measurement station: (1) reference station (2) basin scale cross-section (3) 500m spaced station

SCS

Name of the cross-section

Gravimetric station

10 0 -10 -20 Total Bouguer anomaly and its linear regression

-30 -40

B

+20

Gravity (mGals)

Gravity (mGals)

A

0

0

Regional gravity variation Torne et al. (2000) Residual Bouguer anomaly

10

20

30

40

50

-20

Measured

-40

Calculated

Error = 0.831

Gravity (mGals)

C 0 -20

Deep crust

-40

D

Metamorphic units

Sedimentary units

S

N

Depth (km)

0 5 10

0

V.E.=1

10

20

Post-Tortonian deposits

Alpujarrides

Serravallian? to Tortonian

Upper NF units Lower NF units

Distance (km)

40

Geometry of the basement/sedimentary cover contact: Sediments: 2,35 g.cm

50

N

-5 Mesuread

Calculated

2 4

0

2

4

6

+10 0 20°

Depth (km)

0

2

0

2

4

Section length (km) NW

SCS6

Error = 0.205

-10 -15

2

0

2

4

8

6

N

+20

Error = 0.138

+10 25° 45°

Depth (km)

Depth (km)

0

SCS7

S

17°



12°

4

6

10

Section length (km)

RBa (mgals)

RBa (mgals)

SE

12°





0

4

8

NW Error = 0.235

16°





Depth (km)

Error = 0.953

SCS5

SE

Section length (km)

20°

2

2

4

Section length (km)

Post-Tortonian deposits

Alpujarrides

Upper NF units

Serravallian? to Tortonian

Dense units

Lower NF units



25°

2

4

N

Error = 0.338

-10

0

0

SCS4

0

0

4

8

55° 35°

S RBa (mgals)

-10

RBa (mgals)

RBa (mgals)

SCS3

S

0

2 Section length (km)

A

* *

*

*

*

*

* * * * * * *

* 0* m *

*

±

B

*

0m

-2000

m 1000 *

0m

200 *

*

0m

±

*

20

*

* * ** * * * * * * * *

0

km

* *

*

-10 0

*

*

*

0 1

*

*

* * *

0m

*

2

1000 m

*

*

3

**

* * * *B* * * * * * * * *

0

*

1

*

2 Km

4 5 km

15 km

Gypsum

Huércal-Overa Basin

E

Serr. 12,0

11,5

11,0

D

10,5

10,0

9,5

Tortonian 9,0

8,5

Global sea-level curves Haq et al., 1987

Sorbas Basin 200

-200

Messinian Salinity Crisis

-200

600 m

C

8,0

Messinian 7,5

7,0

6,5

6,0

-600

B

Pliocene 5,5

5,0

4,5

4,0

3,5

Huércal-Overa basin (Augier, 2004)

A

Quaternary 3,0

2,5

Subsidence curves Miller et al., 2005

Paleobathymetry (meters)

?

ZFT: 250-290 °C

Regional uplift Gilbert delta Sorbas limestone - Zorreras

200

Regional Compression

Basin filling

Reef

AFT: 60-110 °C

Azagador Member

U-Th/He: 60-50°C

Tortonian turbidites

Tortonian turbidites

Subsidence

Sea-level (meters)

Local subsidence

Uplift

Abad marls and diatomites

Regional Extention

Sorbas basin (this study)

2,0

1,5

1,0

0,5

0 Ma

Upper Serravallian 2°30'W Lower Tortonian N (~12 - 8 Ma) de Sierra

A Mu lham rci a a F de au lt

2°00'W

N

s

tancia

las Es

2°00'W

2°30'W Late Tortonian (from 8 Ma)

de Sierra

37°30'

s

tancia

las Es

37°30'

H-O

SIERRA DE LOS FILABRES

SIERRA DE LOS FILABRES

S

bo

ne

Sierra Cabrera

37°00'N

Sierra Alhamilla r Ca

S

Sierra Cabrera

?

Palo

Palo

mare

mare

s Fau lt

s Fau lt

H-O

r

as

F

l au

Sierra Alhamilla

t

r Ca

bo

ne

r

as

Fa

ul

t

ALMERIA

ALMERIA Volcanics 20 km

2°30'

20 km

Alpujarride

Cabo de Gata

Shear zone

2°00' W

Nevado-Filabride

Amphibolite lineation in the Nevado-Filabrides [Galindo Zaldivar et al.,1989; Jabaloy et al.,1992; Augier et al., 2005a] Greenschist lineation and flow lines in the Nevado-Filabrides [Augier et al., 2005a]

Cabo de Gata

2°30'

Extensional shear zones

2°00' W

Mean extension direction

Thrust Normal fault Strike-slip fault

Mean compression direction

37°00'N

A

1000

Basin-scale cross-section at present

800 600

Uplift of the basin since the Pliocene and the deposition of marine deltaic sediments

400 200 0

Local overthrusting of the Sierra Alhamilla over the sedimentary basin

0

0

B

400 200

2

1000

4 km

2000 m

Messinian: deposition of the gypsum unit Deposition of an alternance of gypsum/marls layers relative to the first sea-level drop of the Messinian Salinity Crisis

0

C

0

1000

2000 m

0

1000

2000 m

Depocenter axis

Upper Tortonian: syn-kinematic sedimentation of late Tortonian marine sediments 200

CFZ

0

Tilting of preceding normal faults sets

0

1000

North Alhamilla blind thrust

2000 m

Vertical exaggeration: 2

D

200

Lower Tortonian: syn-kinematic sedimentation of late Tortonian marine sediments Tectonic inversion of the area: once the metamorphic complexes forming the Sierras are finally exhumed, the Africa/Europe convergence implies a N-S compression in the basin

0

0

E

1000

2000 m

Depocenter axis

Lower Tortonian: syn-kinematic sedimentation of late Tortonian marine sediments

Depocenter axis Vertical exaggeration: 2

200 0

Normal faulting 0

1000

2000 m

Depocenter axis Vertical exaggeration: 2

Syn-tectonic sedimentation following the denudation of the Nevado-Filabride core complex in the Sierra Alhamilla antiform.

Sedimentary units Latest Messinian early Pliocene Gilbert-delta

Lithologies

Messinian evaporites Messinian reef and marls Lowermost Messinian Tortonian turbiditic Alpujarride nappe Bedar-Macael Unit Calar-Alto Unit Ragua Unit

Gyps um

N° Sample

Fluviatile conglomerates Coarse-grained conglomerates Marls and clays Reef Wi ti s h di a tomi te a nd l a mi ni te Ri ch fos s i l i ferous ca l ca reni te Sa nds tone Purpl e s chi s ts

NevadoFilabride complex

Lubri n gra ni te, ma rbl e Gra phi ti c a nd bl a ck s chi s ts Bl a ck s chi s ts

1 2 3 4 5 6 7 8 9 10

Density Measured Reference 2.36 2.4 2.36 2.32 2.31 2.31 1.82 2.4 1.57 2.46 2.46 2.54 2.54 2.69 2.69 2.67 2.9 2.67 2.67 2.65 2.65

Literature Limestone: 2.49 ± 0.12 Marl: 2.4 ± 0.1; clay marl: 2.06 ± 0.3 Gypsum: 2.35 -/+ 1.5,2.5 Marl: 2.4 ± 0.1 Conglomeratic limestone: 2.48 Schist: 2.64 ± 0.26 Gneiss, schist: 2.65 -/+ 0.25,0.35 Phyllite: 2.74 ± 0.6 Amphibolite: 2.96 ± 0.8 ; Eclogite: 3.37 ± 0.17

Forma on Zorreras coquina Sorbas calcarenite Gypsum Messinian reef

Age Palaeobathymetry (m) ≈ 5.2 Ma 20 m 5.67 - 5,60 Ma 20 - 40 m 5.96 - 5.67 Ma 0 - 40 m 6.04 - 5.89 Ma 0 - 80 m

Abad marls

7.24 - 5.96 Ma

200 - 600 m

Tort/Mess calcarenites Upper Tortonian

≈ 7.24 Ma > 7,24 Ma

20 - 40 m 350 - 600 m

References Montenat et O d'Estevou, 1977 Montenat et al., 1980 Dronkert, 1976; Goubert et al., 2001 Braga et al. , 2003; Sanchez-Almazo et al. , 2007 Dronkert, 1976; Troelstra et al., 1980; O d'Estevou et Montenat, 1990; Poisson et al., 1999; Krijgsman et al. ,2001 O d'Estevou et Montenat, 1990. Montenat et Seilacher, 1978; O d'Estevou et al., 1981