Accepted Manuscript A reappraisal of the Alpine structure of the Alpujárride Complex in the Betic Cordillera: Interplay of shortening and extension in the westernmost Mediterranean J.F. Simancas PII:
S0191-8141(18)30142-1
DOI:
10.1016/j.jsg.2018.08.001
Reference:
SG 3720
To appear in:
Journal of Structural Geology
Received Date: 8 March 2018 Revised Date:
31 July 2018
Accepted Date: 1 August 2018
Please cite this article as: Simancas, J.F., A reappraisal of the Alpine structure of the Alpujárride Complex in the Betic Cordillera: Interplay of shortening and extension in the westernmost Mediterranean, Journal of Structural Geology (2018), doi: 10.1016/j.jsg.2018.08.001. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18
Abstract
19
The Alpujárride Complex has concentrated discussion on the extent and role of orogenic
20
extension in the Betic-Rif Orogen (westernmost Mediterranean). Structural analysis on Permo-
21
Triassic rocks of the Alpujárride Complex is a firm basis to assess extension and to integrate it in
22
the plate tectonic scenario. The main Alpine deformation was dominated by top-to-the-NE
23
shearing, with rare map-scale folds; as a whole, this deformation attests ductile extension during
24
exhumation of previously subducted rocks. Some authors have suggested that the Alpujárride
25
Complex continued evolving extensionally since this early exhumation, but in this paper a stage of
26
regional shortening is documented. Thus, km-scale overturned NW-vergent folds are interpreted
27
as witnessing shortening. Subsequently, post-metamorphic low-angle faults of Burdigalian-
28
Langhian age cut the train of folds in two ways: firstly, top-to-the-N thrusts generated
29
stratigraphic and metamorphic superpositions; then, top-to-the-N low-angle normal faults,
30
kinematically congruent with the thrusts, formed due to accretion at deep levels of the orogenic
31
wedge when the Nevado-Filábride Complex underthrust the Alpujárride Complex. The shortening
32
stage occurred at latest Oligocene early-middle Miocene and can be related to fast convergence
33
rates between Africa and Iberia. Later, a drop in the convergent rates prompted lithospheric
A reappraisal of the Alpine structure of the Alpujárride Complex in the Betic Cordillera: interplay of shortening and extension in the westernmost Mediterranean J.F. Simancas Departamento de Geodinámica, Universidad de Granada, Spain
RI PT
Phone number: +34 958243353 email:
[email protected] Address: Departamento de Geodinámica, Facultad de Ciencias, Universidad de Granada. 18071 Granada, Spain
AC C
EP
TE D
M AN U
SC
Keywords: Orogenic extension and shortening, crustal deformations and plate kinematics, Alpujárride Complex, western Mediterranean
ACCEPTED MANUSCRIPT 34
rollback and top-to-the-SW crustal extension since Serravallian time, coexisting with moderate
35
orthogonal shortening that gave way to E-W trending upright folds.
36 1. Introduction
38
Mountain belts, the ultimate consequence of orogenic evolution, are mainly characterized by
RI PT
37
intense crustal shortening and uplift. However, a major advance in understanding the orogenic
40
evolution has been the recognition that extensional processes may also play a significant role in
41
shaping mountain belts. The amount and timing of extensional deformation in the orogens is very
42
variable, as well as the extension scale (crustal or lithospheric) and the driving mechanisms
43
(collapse of a thickened crust, bending of a belt, upwelling of asthenospheric mantle, subduction
44
rollback, delamination, etc.). Therefore, the correct recognition and characterization of shortening
45
vs. extension, and the interpretation of all these deformations in terms of tectonic processes is a
46
main goal of structural studies in orogens.
M AN U
The Mediterranean region is a key area to study the interplay of orogenic shortening
TE D
47
SC
39
construction and extensional collapse during the Alpine orogeny (e.g., Jolivet and Faccenna, 2000;
49
Faccenna et al., 2004). This paper, by presenting a re-appraisal of the ductile-to-brittle structural
50
evolution of the Alpujárride Complex from the Betic-Rif orogen of Western Mediterranean (Fig. 1),
51
is aimed at constraining the role of shortening vs. extension in the Alpine tectonic evolution of the
52
Internal Zone of this orogen. To do so, this study focuses on the description and interpretation of
53
geometric/kinematic data of the Alpine structure that have been mistaken or overlooked in
54
previous papers. The revised structural data presented here are strong argument to decide on the
55
extensional versus contractional nature of some deformational events in this region, which, in
56
turn, is significant for unravelling the tectonic evolution of the Alpujárride Complex and the Betic
57
Internal Zone as a whole. In the light of the new interpretation of the Alpine internal structure of
58
the Alpujárride Complex, a comprehensive tectonic evolution will be finally discussed.
59
AC C
EP
48
ACCEPTED MANUSCRIPT 60 61 62
2. Geological setting 2.1. Boundary conditions and deep structure of the Betic-Rif Orogen A first main constraint to the Betic-Rif orogenic evolution is the N-S to NW-SE Africa-Eurasia plate convergence during Cenozoic time (Dewey et al., 1989; De Mets et al., 1990; Rosenbaum et
64
al., 2002), responsible for significant shortening in that direction. A second constraint is the
65
westward displacement of the Internal Zone of the orogen (Sanz de Galdeano, 1990; Galindo-
66
Zaldívar et al. 2015; González-Castillo et al., 2015), which gave way to clockwise rotation of
67
paleogeographic units of the External Zone and eastward-directed subduction. The interplay of
68
these two main constraints has originated a complex kinematics of deformations.
SC
RI PT
63
In agreement with the complex crustal kinematics, the deep structure under the Betic-Rif
70
orogen (Fig. 1A) is not simple, as revealed by geophysical studies. Thus, i) the Variscan crust of the
71
Iberian Massif penetrates at least under the Betic External Zone, as demonstrated by
72
aeromagnetic anomalies and a magnetotelluric NW-SE transect at the western Betics (Galindo-
73
Zaldívar et al., 1997; Ruiz-Constán et al., 2012); ii) seismicity of intermediate depth and seismic
74
surveys in the westernmost Mediterranean indicate continental subduction towards the E-SE
75
(Lonergan and White, 1997; Morales et al., 1999; Gutscher et al., 2002); iii) deep seismicity (640-
76
670 km) and seismic tomography suggest the existence of a lithospheric slab in the mantle
77
underlying the westernmost Mediterranean (Blanco and Spakman, 1993; Bezada et al., 2013); iv)
78
P-receiver functions suggest subduction rollback with edge delamination under the Betics and Rif
79
(Mancilla et al., 2015).
81
TE D
EP
AC C
80
M AN U
69
2.2. The Betics and The Betic Internal Zone
Five geological domains are usually distinguished in the Betics, the Iberian part of the Betic-Rif
82
orogen (Vera and Martín-Algarra, 2004, and references therein): i) the Guadalquivir foreland
83
basin; ii) the External Zone, which correspond to the paleomargin of southern Iberia and is made
84
up of Triassic to Miocene sedimentary rocks; iii) The Flysch Trough Units, constituted by detached
85
and dismembered Meso-Cenozoic rocks that represent the sedimentary cover of a subducted
ACCEPTED MANUSCRIPT oceanic or very thinned continental crust; iv) the Frontal Units or Predorsal, a narrow
87
discontinuous band of mainly Mesozoic rocks firstly imbricated towards the foreland and later
88
backfolded and backthrust over the Internal Zone, together with the Flysch Trough Units; and v)
89
the Internal Zone, dominated by metamorphic rocks of Paleozoic and Mesozoic (mostly Triassic)
90
age. Continental collision with the Maghrebian paleomargin occurred at Miocene time, after
91
subduction of the Flysch Trough basement.
RI PT
86
The Betic Internal Zone (Fig. 1A) is constituted by Mesozoic and Paleozoic rocks usually
93
metamorphosed and intensely deformed. The large-scale structure is a tectonic pile with three
94
main tectonometamorphic units, namely, from top to bottom, the Maláguide Complex, the
95
Alpujárride Complex and the Nevado-Filábride Complex. The paleogeographic location of these
96
three complexes is uncertain, but it is commonly assumed that they were located eastwards of
97
their present-day position, so that their westward migration would have contributed to shaping
98
the curvature of the Betic-Rif Orogen.
M AN U
The Maláguide Complex overthrust the Alpujárride Complex, though the original contact
TE D
99
SC
92
between them has been modified by later extensional tectonics (Aldaya et al., 1991; Lonergan and
101
Platt, 1995; González-Lodeiro et al., 1996; Booth-Rea et al., 2003). The Maláguide units have a
102
Paleozoic sole affected by pre-Alpine (Variscan) penetrative deformations and metamorphism.
103
This Paleozoic basement is unconformably overlain by Triassic-Miocene sedimentary rocks
104
affected by NW-vergent thrusts and folds (Lonergan, 1993; Booth-Rea et al., 2002).
AC C
105
EP
100
The Nevado-Filábride Complex occupies the lowest structural position in the Betic Internal
106
Zone and crops out at the core of large-scale antiformal ridges. This complex is perhaps the most
107
controversial in the internal Betic Zone, being at present the subject of a strong debate regarding
108
stratigraphy, subdivision of tectonic units, internal structure and even the existence or not of an
109
Ophiolitic Unit (De Jong, 1993; Gómez-Pugnaire et al., 2000; Puga et al., 2002; Martínez-Martínez
110
et al., 2002, 2007). However, there is agreement on the existence of a pre-Alpine (Variscan) low-
111
pressure metamorphism and an Alpine metamorphic evolution consisting of an initial stage of
ACCEPTED MANUSCRIPT high-pressure occurred at 15-18 Ma (López Sánchez-Vizcaíno et al., 2001; Platt et al., 2006),
113
followed by intermediate and low-pressure conditions. The most evident deformations are late-
114
metamorphic extensional shear zones that have erased previous deformations, apparently
115
concentrated on the top of the Nevado-Filábride tectonic pile. Shear zones contain many minor
116
folds subparallel to the stretching lineation and a few major NW-vergent tight-to-isoclinal folds;
117
kinematic criteria indicate top-to-the-W sense of movement (Galindo-Zaldívar et al., 1989; Jabaloy
118
et al., 1993; De Jong, 1993; Martínez-Martínez et al., 2002). At structural levels beneath the
119
extensional shear zones, there are several shear zones of similar kinematics but interpreted as
120
contractional, since they separate stratigraphic recurrences and rocks with higher metamorphic
121
grade superposed on rocks with lower metamorphic grade (Martínez-Martínez, 2007).
SC
M AN U
122
RI PT
112
The Alpujárride Complex, at an intermediate position in the tectonic pile of the Betic Internal Zone, is the one with more widespread outcrops (Fig. 1A). The thrust contact that placed the
124
Alpujárride rocks onto the Nevado-Filábrides ones has been obliterated by the extensional shear
125
zones referred above. The lithostratigraphic sequence of the Alpujárride tectonic units is more
126
complete in the western units (Fig. 1C), with a common lithostratigraphy that consists in, from
127
bottom to top: i) dark schists and quartzschists of Paleozoic age, sometimes with gneissic rocks at
128
the base; ii) light-colored schists, quartzschists and quartzites, also of Paleozoic age; iii) Permo-
129
Triassic phyllites and quartzites; and iv) Triassic calcitic and dolomitic marbles. However, a very
130
significant difference (not shown in Fig. 1C) appears in the westernmost Alpujárride unit, which
131
includes a sole of sub-continental ultramafic rocks (the so-called Ronda peridotites) and high-
132
grade gneisses. The metamorphism of the Alpujárride units varies from low- to high-grade,
133
though its characterization in the Paleozoic schists is complicated by the existence of pre-Alpine
134
parageneses. This paper studies the Alpine structure of the Alpujárride units that crop out in a
135
central transect shown in Fig. 1B, which is then used to discuss a tectonic interpretation of the
136
Internal Betic Zone.
137
AC C
EP
TE D
123
2.3. Sequence of deformations in the Alpujárride Complex
ACCEPTED MANUSCRIPT 138
The following sequence of deformational events has been determined in Permo-Triassic
139
phyllites and marbles of the central transect of the Alpujárride Complex (Fig. 1B), being probably
140
valid for eastern and western transects too. The succession of deformations is, from older to
141
younger (the suffix A in the labels refers to Alpine):
143 144
- D1(HP)A, which corresponds to the high-pressure/low-temperature metamorphism
RI PT
142
defined by carpholite-bearing assemblages in phyllites (Azañón and Goffé, 1997).
- D2A, characterized by a non-coaxial and very penetrative deformation that gave way to a planar-linear fabric (foliation S2A, stretching lineation Ls2A) with minor folds subparallel to the
146
stretching lineation (F2A) (Avidad and García-Dueñas, 1981; Simancas and Campos, 1993; Balanyá
147
et al., 1997; Rossetti et al., 2005).
150 151
M AN U
149
- D3A, which consists in km-scale, overturned, NW-vergent folds trending NNE-SSW to ENEWSW, with a spaced axial planar foliation (S3A) (Simancas and Campos, 1993). - D4A, constituted by top-to-the-N low-angle faults (Aldaya, 1981; Avidad and GarcíaDueñas, 1981; Simancas and Campos, 1993; Crespo-Blanc et al., 1994).
TE D
148
SC
145
-Finally, there is a set of recent (Serravallian to present) but kinematically varied
153
structures, which are here grouped as DN. This set includes: i) Major low-angle normal fault zones
154
with dominant top-to-the-SW kinematics (Fig. 1B), the main one being the current Nevado-
155
Filábride/Alpujárride boundary (Jabaloy et al., 1993; Martínez-Martínez et al., 2002, 2004); ii) E-W
156
trending upright open folds, among which the Sierra Nevada antiform is the biggest one
157
(Martínez-Martínez et al., 2002, 2004; Sanz de Galdeano and Alfaro, 2004; Pedrera et al., 2007);
158
iii) NE-SW left-lateral faults with subordinate NW-SE to E-W right-lateral ones, and high-angle
159
normal faults of variable orientation (see compilation by Sanz de Galdeano and Peláez, 2011).
160
Present-day deformation in the Betics seems to be dominated by the SW-directed extension, with
161
NW-SE shortening being subordinate (Azañón et al., 2015; Galindo-Zaldívar et al., 2015; González-
162
Castillo et al., 2015). However, in the eastern Betics active tectonics is dominated by NW-SE
AC C
EP
152
ACCEPTED MANUSCRIPT 163
shortening, with folds, thrusts and strike-slip faults (Bousquet, 1979; Silva et al., 1993; Alfaro et
164
al., 2012; Echeverría et al., 2015).
165
This paper only deals with the pre-Serravallian D2A, D3A, D4A1 and D4A2 deformational events. Therefore, DN structures, which attest the recent and present-day tectonic activity in the region,
167
are beyond the scope of this work. The tectonic discussion presented at the end of this paper will
168
show that the distinction between pre- and post-Serravallian structures is not arbitrary,
169
corresponding to a change in the large-scale plate tectonic scenario. 2.4. Main issues in the Alpujárride Complex
171
2.4.1.Variscan versus Alpine imprint
Recent investigations have definitely shown the importance of the Variscan imprint in the
M AN U
172
SC
170
RI PT
166
Paleozoic rocks of the Alpujárride Complex, thus bringing into question the age of a part of
174
mineral growth and tectonic fabrics (Acosta et al. 2014; Sánchez-Navas et al. 2014). This is the
175
reason why in this paper the structural analysis has been focused on Permo-Triassic rocks, where
176
the entire tectonometamorphic imprint must be necessarily Alpine. The presumed Variscan-
177
Alpine unconformity, which should be located somewhere between the black Paleozoic schists
178
and the Permo-Triassic phyllites, has not been located yet.
180
2.4.2.Timing of the early Alpine events
EP
179
TE D
173
The metamorphic record in the Alpujárride Complex attests an Alpine early history of burial at high-pressure/low-temperature conditions followed by exhumation (Goffé et al., 1989; Azañón
182
and Goffé, 1997; Azañón et al., 1998; Booth-Rea et al., 2002). The timing of the high-pressure
183
event is uncertain, with muscovite 40Ar-39Ar dating having yielded tentative ages of c. 48 Ma (Platt
184
et al., 2005). If the high-pressure event took place at c. 48 Ma and the subsequent deformational
185
events started at earliest Miocene time (Platt et al., 2005), a time gap of unknown tectonic
186
meaning would have existed in-between. On the other hand, the age of the high-pressure
187
metamorphism in the Nevado-Filábride Complex has been found to be as young as 15-18 Ma (U-
188
Pb on zircon, López Sánchez-Vizcaíno et al., 2001; Lu-Hf on garnets, Platt et al., 2006), hence
AC C
181
ACCEPTED MANUSCRIPT 189
evidencing that the high-pressure events in the Alpujárride and Nevado-Filábride complexes are
190
unrelated.
191 192
2.4.3.Extension versus shortening The recognition of important low-angle normal faults in the Betic Internal Zone gave way to a profound rethink of the orogenic evolution. Thus, the view of the Internal Zone of the orogen as
194
the result of thrust-stacking was abandoned in favor of a collapsed tectonic pile made up of
195
sheets bounded by low-angle normal faults (Galindo Zaldívar et al., 1989; Aldaya et al., 1991;
196
García-Dueñas et al. 1992; Jabaloy et al., 1993; Crespo-Blanc et al., 1994; Crespo-Blanc, 1995;
197
Lonergan and Platt, 1995; Vissers et al., 1995; Martínez-Martínez et al., 2002). Some authors
198
consider that continuous extension has dominated the Betics since the end of the subduction-
199
related high-pressure event (Platt and Vissers, 1989; Vissers et al., 1995; Orozco et al., 1998, 2004,
200
2017; Williams and Platt, 2017), while others propose alternating contractional and extensional
201
events (Tubía et al., 1992; Simancas and Campos, 1993; Azañón et al., 1997; Balanyá et al., 1997;
202
Azañón and Crespo-Blanc, 2000; Rossetti et al., 2005). Among the latter, there is remarkable
203
confusion concerning the contractional or extensional nature of many tectonic contacts in the
204
Alpujárride Complex. In order to decide which one of these disparate interpretations is more
205
appropriate, this paper firstly provides with a precise geometric and kinematic description of the
206
structures, and then integrate these structures in a proposal of tectonic evolution for the
207
Alpujárride Complex and the Betic Internal Zone.
SC
M AN U
TE D
EP
AC C
208
RI PT
193
209
3. Description of Alpine deformations in the Alpujárride Complex
210
3.1. Early Alpine syn-metamorphic deformations (D1(HP)A; D2A)
211
Relics of an earliest foliation probably related to the high-pressure/low-temperature event
212
are exceptionally preserved at microscopic scale (e.g. Azañón et al., 1997; Rossetti et al., 2005);
213
this is apparently what Williams and Platt (2017) have noticed as cleavage arcs surrounded by the
214
main foliation in chloritoid-kyanite schists. This relic foliation that precedes the first recognizable
ACCEPTED MANUSCRIPT folds is labelled here as S1A. Moreover, in metapelites (but not in quartzites or marbles) the first
216
recognizable folds affect a planar fabric defined by preferred orientation of phyllosilicates, parallel
217
to thin grain-size layering, which could represent either the very first tectonic foliation (S1A) or a
218
recrystallized sedimentary (compaction) fabric (Hobbs et al., 1976). In any case, there are no
219
observable folds or shear-sense indicators related to that relic or presumed tectonic foliation,
220
being not possible to decipher any geometric or kinematic information.
221
RI PT
215
The main deformational fabric in the Permo-Triassic rocks, i.e. the first one associated with
recognizable folds, is frequently LS-type, characterized by foliation (S2A) and associated stretching
223
lineation (Ls2A); S2A is continuous and axial planar of tight microfolds; Ls2A orientates NE-SW (Fig.
224
1D; Simancas and Campos, 1993). Asymmetric tails in porphyroclasts indicate top-to-the-NE non-
225
coaxial flow (Fig. 2A). In the Paleozoic schists, S2A appears as an intense crenulation cleavage
226
enclosing a presumably Variscan relic foliation within microlithons. Nevertheless, in strain
227
shadows around competent Paleozoic igneous bodies intruded in the schists, the first Alpine
228
deformation may be weak, the Variscan foliation prevailing at outcrop-scale (Sánchez-Navas et al.,
229
2014). In the central and eastern Alpujárride units, D2A metamorphism is of low grade in phyllites
230
and carbonates, reaching medium grade in dark (Paleozoic) schists (Aldaya, 1981; Avidad and
231
García-Dueñas, 1981).
M AN U
TE D
EP
232
SC
222
Tight micro- and meter-scale folds of this deformational event are ubiquitous (Fig. 2B), but map-scale folds are exceptional. Indeed, inverted stratigraphy always appears in connection with
234
the short limbs of F3A folds (see next section), except in the sector east of Sierra de Lújar (Estévez
235
et al., 1985), where map-scale F2A folds appear coaxially overprinted by F3A folds (Figs. 3A and 4A).
236
In this sector, F2A axes trend NNE-SSW with gentle but variable plunge; folds are close-to-tight,
237
overturned, east-vergent (Figs. 3B and 4) and subparallel to the stretching lineation observed in
238
other sectors (compare Figs. 1D and 3B). Interestingly, the deformation associated with these F2A
239
folds lacks stretching lineation or asymmetric microstructures, so that folds probably nucleated at
240
oblique angle to the regional stretching, then undergoing only moderate passive rotation of hinge
AC C
233
ACCEPTED MANUSCRIPT 241
lines. After removing the effect of superposed F3A folding, F2A folds mapped east of the Sierra de
242
Lújar form a train of east-vergent overturned folds with inverted limbs less than 1 km-long (Fig.
243
4A). It is worth noting that F3A folds east of the Sierra de Lújar (Figs. 3A and 4) are second-order
244
folds of the km-scale fold that mostly shapes that Sierra.
246
3.2. Late-metamorphic penetrative folding phase (D3A)
RI PT
245
The existence of km-scale folds widely distributed in the Alpujárride Complex, which fold the main foliation and are cut by low-angle faults, was initially pointed out by Estévez et al. (1985)
248
and Simancas and Campos (1993). The selected cross-sections located in Fig. 1B and displayed in
249
Fig. 5 show the geometry and arrangement of these folds. F3A folds are close, overturned and NW-
250
vergent (Fig. 2C); their dominant trend is ENE-WSW, with dominant gentle to moderate plunge
251
towards WSW (Fig. 1D). The axial planar foliation of F3A folds (S3A) is a spaced crenulation cleavage
252
in the phyllites and a spaced disjunctive (pressure solution) cleavage in the carbonate rocks, more
253
penetrative in hinges and inverted limbs. At the convex side of hinges in carbonate rocks, the
254
adjacent phyllites show small triangular zones of weak strain, which are characteristic of buckling
255
with moderate competence contrast (Fig. 2D); moreover, a certain resemblance of F3A folds to the
256
cuspate (anticlines) and lobate (synclines) style (Figs. 5 and 6) also indicates some rheological
257
contrast, with the Triassic carbonate rocks being more competent than the underlying phyllites.
258
F3A folds were formed at low-grade metamorphic conditions. The timing of these folds is not
259
accurately stablished but 40Ar-39Ar data suggest an earliest Miocene age (Platt et al., 2005).
M AN U
TE D
EP
AC C
260
SC
247
The biggest F3A folds are found in the lowermost tectonic unit of the Alpujárride Complex, the
261
so-called Lújar unit, due perhaps to a thicker competent “layer” of carbonate rocks (Figs. 5E and
262
5F). In this unit, three successive km-scale folds are distinguished, from NW to SE: the Sierra de
263
Lújar fold, the Sierra de Turón fold and the Sierra Alhamedilla fold (Figs. 1B, 5E and 5F; Simancas
264
and Campos, 1993). Furthermore, F3A folds in the Lújar unit are bigger than the F2A folds
265
previously developed in the same area (compare Fig. 4A with Figs. 5E and 5F), which suggests
266
different mechanisms of formation. The major F3A folds appear on map as remarkably continuous
ACCEPTED MANUSCRIPT until being cut by the low-angle faults of the D4A event (see axial traces in Fig 1B); accordingly,
268
each one of the tectonic sheets contains different fold portions, though restoration rearrange the
269
pieces to reveal the original train of folds (Fig. 6). In this respect, there are significant differences
270
between the schematized structural map of Williams and Platt (2017) and the one in Fig. 1B. The
271
nomenclature of the tectonic sheets distinguished in this paper is shown at the bottom of Fig. 5.
272 273
3.3. Top-to-the-N low-angle faulting (D4A)
RI PT
267
Low-angle faults of two different types, thrusts (D4A1) and normal faults (D4A2), cut the F3A folds (see cross-sections in Fig. 5). The kinematics of thrusts and low-angle normal faults is identical
275
(top-to-the-NNW or -N; Fig. 1D); both types of faults seem to have developed close in time though
276
consecutive, the thrusts being systematically cut by the low-angle normal faults. Associated fault-
277
rocks are random fabric crush breccias in carbonate rocks and foliated gouges or spaced C´-type
278
shear bands in phyllites/schists (Figs. 2E and 2F); thrust fault zones contain less or no gouge and
279
more C´-type shear bands than normal fault zones. All of these low-angle faults were
280
subsequently cut off by top-to-the-SW low-angle normal faults (Fig. 1B) and finally folded by open
281
upright folds.
M AN U
TE D
282
SC
274
3.3.1.Thrusts (D4A1)
The observation of repeated stratigraphy above and below the low-angle normal faults was
284
the reason why all these faults were originally interpreted as thrusts (e.g., Navarro-Vilá and García
285
Dueñas, 1980; Aldaya, 1981; Avidad and García-Dueñas, 1981). Later on, detailed observations on
286
the contact between the Nevado-Filábride and the Alpujárride complexes proved the extensional
287
character of this main fault zone (Galindo-Zaldívar et al., 1989; Jabaloy et al., 1993; Martínez-
288
Martínez et al., 2002), and this extensional interpretation was quickly applied by many authors to
289
practically all low-angle faults in the Betic Internal Zone. However, the structural data provided
290
here show that there is a good deal of thrust contacts preserved in the Alpujárride pile, such as: i)
291
a foreland dipping duplex in the Salobreña tectonic sheet (Fig. 5A; ii) imbricate fans in various
292
tectonic sheets (Figs. 5D, 5E and 5F); and iii) ascending frontal ramps in Triassic carbonates (Figs.
AC C
EP
283
ACCEPTED MANUSCRIPT 293
5B, 5C and 5D). Accordingly, some illustrative examples of thrust superposition are as follows: the
294
thrust of the Guájares sheet onto the Salobreña one (Fig. 5B), the thrust of the Salobreña sheet
295
onto the Herradura one (Fig. 5C), the thrust of the Herradura sheet onto the Lújar one (Fig. 5D),
296
and the large-scale imbricated structure of the Lújar unit (Fig. 5E, L2 onto L1; Fig.5F, L3 onto L2). 3.3.2.Low-angle normal faults (D4A2)
RI PT
297
The extensional character of a part of the top-to-the-N low-angle faults is inferred from their
299
descending trajectories with respect to both thrusts and stratigraphy in normal limbs of F3A folds
300
(Figs. 5A, 5B, 5D and 5E). Since the low-angle normal faults are folded by late open upright folds,
301
their variable dip is not a reliable feature to stablish the extensional character (see Figs. 5B and
302
5E). The map in Fig. 1B shows the distinction made in this paper between thrusts and low-angle
303
normal faults: it is clear from this map that a large part of the extensional brittle deformation
304
affecting the Alpujárride Complex is not due to the top-to-the-N faults but to subsequent top-to-
305
the-SW or W low-angle normal faults (García-Dueñas et al. 1992; Jabaloy et al., 1993; Lonergan
306
and Platt, 1995; Martínez-Martínez et al., 2002).
M AN U
TE D
307
SC
298
The relationships between faults and Miocene sediments suggest that the top-to-the-N lowangle faults are Burdigalian-Langhian in age, while younger top-to-the-SW low-angle normal faults
309
are Serravallian (Mayoral et al., 1994; Crespo-Blanc, 1995).
311
3.3.3.The current tectonic pile The present-day geometry of the Alpujárride Complex is dominated by a tectonic pile due to
AC C
310
EP
308
312
the interaction of the low-angle faults described above (Fig. 1B). In this respect, two main
313
considerations must be done: i) no tectonic superpositions ascribable to the D2A event (which
314
should be folded by F3A folds) has been mapped yet; ii) despite the relevance of the low-angle
315
normal faults, the tectonic pile is still dominated by stratigraphic repetitions in most sectors.
316
A complete understanding of the meaning of mapped tectonic sheets can only be attained
317
through restoration of the structure; unfortunately, this is possible only in a few areas of well
318
constrained and continuous structure. The restoration in Fig. 6 shows the position that the
ACCEPTED MANUSCRIPT 319
Herradura, Salobreña and Guájares tectonic sheets had in the D3A fold train prior to the D4A event.
320
Fig. 6 also shows that a significant N-S shortening of approximately 50 km took place in the
321
analyzed geological transect, though this is only a fraction of the total N-S orogenic shortening. Since middle Miocene to present-day, after the top-to-the-N low-angle faults in the
323
Alpujárrides, low-angle normal faults with SW-NE extensional kinematics started to operate
324
affecting the entire Internal Zone of the Betics (Fig. 1B), overprinting previous structures (Galindo-
325
Zaldívar et al., 1989; Jabaloy et al., 1993). Coevally, orthogonal shortening gave way to major E-W
326
trending upright folds (Martínez-Martínez et al., 2002, 2004; Azañón et al., 2015).
SC
RI PT
322
327
4. Discussion: tectonic interpretation of the Alpine deformations in the Alpujárride
M AN U
328 329
Complex
330
4.1. D2A event
331
PT paths in the Permo-Triassic metapelites (phyllites) attest a pressure drop of 5-6 Kbar associated with the development of the main foliation (S2A), i.e. peak pressures were of 8-10 Kbar
333
while the main foliation developed at 3- 4 Kbar (Azañón et al., 1998; Booth-Rea et al., 2005;
334
Williams and Platt, 2017). This means that the main tectonic fabric in these rocks is related to
335
exhumation, a process hard to evaluate because equating metamorphic pressures and burial
336
depth may result in significant overestimations (Yamato and Brun, 2017).
EP
TE D
332
Regarding deformation, D2A generated an LS fabric with asymmetric microstructures and
338
minor folds. Major folds have only been identified east of Sierra de Lújar (Figs. 1B and 3A), where
339
they show an eastern vergence that may be compatible with the top-to-the-NE kinematics
340
inferred from the LS fabric. To further characterize this deformation, it must be noted that: i) the
341
occasional development of the large-scale F2A folds described in this paper suggests that bedding
342
would have been rarely positioned in the shortening field during D2A shearing; ii) no tectonic
343
superposition ascribable to the D2A event (which should appear folded by the subsequent F3A
344
folds) has been described yet; and iii) the vertical proximity of isograds suggests (though attention
AC C
337
ACCEPTED MANUSCRIPT 345
has not been paid to separate the Variscan metamorphism) ductile vertical thinning of the
346
Alpujárride units (Azañón et al., 1998; Williams and Platt, 2017). Accordingly, the overall
347
extensional interpretation that is generally favored for the D2A deformation is also supported in
348
this paper. 4.2. The D3A deformational event
350
The first late- to postmetamorphic, pre-middle Miocene structures are km-scale, ENE-
RI PT
349
WSW trending, overturned NW-vergent folds (F3A), with moderate or faint hinge curvature. These
352
folds are widely and evenly distributed in the area studied (Fig. 1B).
353
SC
351
The structure of the Sierra de Lújar area deserves attention because it has been presented as an example of mega-sheath folds inside a large-scale, top-to-the-N, extensional shear zone
355
(Orozco et al., 2004, 2017). However, the data presented here do not support that interpretation.
356
Firstly, the structural pattern of this area is not due to a single structure but to fold superposition
357
(Figs. 3A, 4A and 4B); if folding superposition is not considered, the geometric appraisal is
358
incomplete and a mixture of fold styles, minor structures, axial traces and stratigraphic polarities
359
results (both F2A and F3A folds produce local inversions). Furthermore, the rocks do not exhibit the
360
intense strain and stretching lineation that characterizes sheath folds: the Sierra de Lújar is mainly
361
a major F3A syncline (Fig. 5E), with minor folds of this phase featuring moderate strain and spaced
362
foliation; on the contrary, F2A folds (Fig. 4A) show a more penetrative foliation. One more
363
geometrical constraint to be considered is the hinge curvature of the F3A Sierra de Lújar fold,
364
which is rather moderate, trending between N20E at the eastern border (Fig. 3B) and N80E (WSW
365
plunging) at the southern border of the Sierra. The cause of the hinge curvature is uncertain, but a
366
possible explanation is that the uncommon N20E trend results from adjustment to the anisotropy
367
introduced by F2A folds in this area. Interestingly, the km-scale Sierra de Turón overturned
368
syncline, located to the southeast of the Sierra de Lújar fold, can be followed along more than 20
369
km without showing significant hinge curvature (Figs. 1B and 5F); the same can be said about the
370
F3A folds mapped in the tectonic sheets west of the Sierra de Lújar (Figs. 1B and 1D). Finally, it is
AC C
EP
TE D
M AN U
354
ACCEPTED MANUSCRIPT worth noting that the top-to-the-N shear deformation is not ductile and pervasive but semibrittle
372
and discrete (see below). To conclude, this paper does not sustain the interpretation by Orozco et
373
al. (2017) and Williams and Platt (2017) on the F3A folds mapped in Fig. 1B as mega-sheath folds
374
formed in a large-scale, ductile extensional shear zone. Quite the contrary, the data presented
375
here strongly suggest that F3A folds were formed by regional NNW-SSE shortening, which, in turn,
376
implies that the continuous extension envisaged by some authors since the end of the high-
377
pressure event (Platt and Vissers, 1989; Platt et al., 1998; Orozco et al., 1998, 2017, Williams and
378
Platt, 2017) must be reconsidered.
SC
RI PT
371
4.3. The D4A deformational event
380
The superposition of tectonic sheets that characterize the Alpujárride Complex is due to a
381
deformational stage characterized by kinematically coherent (N-directed) thrusts (responsible for
382
stratigraphic and metamorphic repetitions) and low-angle normal faults (Figs. 5 and 6). Contrary
383
to this interpretation, some authors have cast doubt on the truthfulness of post-metamorphic
384
superpositions (Platt and Vissers, 1989; Orozco et al., 1998, 2004; Williams and Platt, 2017). These
385
authors claim that stratigraphic superposition is sometimes explained due to early “ghost” thrusts
386
formed during a subduction-related thickening stage and completely destroyed by transposition
387
and recrystallization later on. Furthermore, since early (pre-metamorphic) thrusts cannot explain
388
metamorphic superposition, normal faults complicating geometric relationships are invoked.
389
However, detailed structural mapping does not support these alternative interpretations, as can
390
be illustrated with the example of the area north of Motril (Fig. 7). The structural map in Fig. 7
391
shows that: i) the internal structure of the Herradura unit schists consists in F3A folds and an
392
imbricate fan of D4A1 thrusts; and ii) the Paleozoic medium-grade schists of the Herradura unit
393
overlie the Triassic low-grade carbonates of the Escalate unit, as demonstrated at the northern
394
frontal footwall-ramp and at the eastern oblique footwall-ramp, this latter re-activated as normal
395
fault. This structure is summarized in the cross-section of Fig. 5D. For the same area, Williams and
396
Platt (2017) suggest the following interpretation: i) pre-metamorphic thrusting between dark and
AC C
EP
TE D
M AN U
379
ACCEPTED MANUSCRIPT light-colored schists; and ii) east-dipping normal faults (opposite to what is observed in the field)
398
downthrowing the Triassic carbonates with respect to the Paleozoic schists. It can be hence
399
concluded that in the area north of Motril (Fig. 7), as in other areas of the Alpujárride Complex,
400
post-metamorphic superposition is well documented (e.g., Avidad and García-Dueñas, 1981;
401
Simancas and Campos, 1993; Rossetti et al., 2005).
402
RI PT
397
The structural association of kinematically coherent thrusts and low-angle normal faults is not uncommon (e.g., Çoruh et al., 1988; Yin and Kelty, 1991; Butler, 1992; England and Molnar,
404
1993), the mechanics of thrust wedges offering a number of potential explanations (Davis et al.,
405
1983; Platt, 1986). Since there is no factual support for a change of the friction coefficient or the
406
fluid pore-pressure at the basal decollement, these factors are not invoked. Instead, accretion at
407
deep levels of the wedge is very appealing because it is well known that the Nevado-Filábride
408
Complex underthrust the Alpujárride Complex coevally (15-18 Ma, López Sánchez-Vizcaino et al.,
409
2001; Platt et al., 2006) with the top-to-the-N normal faults (Burdigalian-Langhian; Mayoral et al.,
410
1994; Crespo-Blanc, 1995). If the surface slope angle of the wedge increased due to
411
underthrusting of the Nevado-Filábride Complex, development of structures that could diminish
412
the surface slope would result, such as normal faults on top of the wedge. Additionally, the
413
deeper location of the basal decollement might have resulted in ductile deformation at the basal
414
slope, also prompting reduction of the surface slope. Thus, underthrusting of the Nevado-
415
Filábride Complex may have induced the development of top-to-the-N low-angle normal faults in
416
the Alpujárrides.
418
M AN U
TE D
EP
AC C
417
SC
403
4.4. Shortening evaluation
An accurate evaluation of the NNW-SSE total shortening due to D3A folds and D4A1 thrusts
419
is not possible for any transect of the orogen, though the following considerations may serve as
420
an approximation. The restoration of the continuous cross-section shown in Fig. 6 yields a
421
shortening of 49 km. However, this is only a part of the total shortening, due to the following
422
reasons: i) the analyzed cross-section does not completely cross the Alpujárride Complex (see
ACCEPTED MANUSCRIPT location in Fig. 1B); ii) whether or not the tectonic pile made up of Guájares + Salobreña +
424
Herradura units (Fig. 6) overstepped the Lújar unit is unknown, a superposition that should be
425
added to compute shortening; iii) the whole Alpujárride Complex thrust over the Nevado-Filábride
426
Complex, the corresponding tectonic window (Figs. 1A and 1B) providing an additional shortening
427
of 50 km; iv) the initial part of the NNW-directed shortening in the External Zone of the Betics
428
(Crespo-Blanc et al., 2007) was coeval with similar shortening in the Alpujárrides, which implies an
429
additional increase in global shortening. Summing up, a minimum figure of around 100 km of
430
NNW-SSE total shortening can be taken for granted in the Alpujárride Complex. This calculation is
431
not affected by the fact that the present-day contact between the Nevado-Filábride and
432
Alpujárride complexes is not the original thrust but a major extensional fault zone with 116 km of
433
W-SW horizontal displacement and substantial ductile thinning of the footwall rocks (Martínez-
434
Martínez et al., 2002); it simply means that the Alpujárride Complex has to be placed before
435
extension more than 100 km east of their present-day location and over a thicker pile of Nevado-
436
Filábride rocks.
TE D
M AN U
SC
RI PT
423
437
4.5. Crustal deformation and plate kinematics
438
The Betic-Rif Orogen exemplifies the interaction of complex boundary conditions constraining crustal deformation. The two most important constraints are the N-S to NW-SE
440
Africa-Eurasia plate convergence during Cenozoic time and the westward displacement of the
441
Internal Zone of the orogen. These two main constraints, together with associated processes of
442
lithospheric rollback and edge tearing (Mancilla et al. 2015) have originated a complex kinematics
443
of deformations. Despite this limitation, essaying a correlation between N-S to NW-SE orogenic
444
shortening and the Africa-Eurasia plate convergence seems still justified.
445
AC C
EP
439
There are significant differences in the Paleogene-Neogene Africa/Iberia convergence
446
rates evaluated by different authors (Dewey, 1989; Rosenbaum et al., 2002; Vissers and Meijer,
447
2012); fortunately, some common patterns can be extracted and considered to relate crustal
448
deformations in the Alpujárride rocks to plate-scale kinematics. The starting point is the
ACCEPTED MANUSCRIPT 449
computation that Africa/Eurasia motions between 80 and 40 Ma were accommodated in the
450
Pyrenean domain, i.e. NNW-SSE convergence and crustal shortening in the Betics should be
451
younger than 40 Ma (Vissers and Meijer, 2012).
452
High convergence rates exceeding 10 km/m.y with NW-SE trending vectors have been computed between ≈ 30 and 20 Ma (Rosenbaum et al., 2002). As the age tentatively attributed to
454
D2A folds is earliest Miocene (Platt et al., 2005), there is coincidence in broad terms between
455
shortening in the Alpujárride Complex and high plate-convergence rates. Actually, the overall
456
timing of the shortening period (D3A folds and D4A low-angle faults) can be estimated on a
457
different ground: i) for more than 100 km of NNW-SSE shortening, as discussed above, convergent
458
velocities ≈ 10 km/m.y need around 10 m.y to complete that figure; and ii) if the subsequent
459
extensional episode started at ≈ 15 Ma (Crespo-Blanc et al., 1994), shortening could have taken
460
place between ≈ 26 and 16 Ma. At some stage of this shortening episode, the entire Alpujárride
461
Complex thrust over the Nevado-Filábride Complex, which reached then its maximum depth (peak
462
pressure dated at 15-18 Ma; López Sánchez-Vizcaino et al., 2001; Platt et al., 2006). Moreover, the
463
new geometry of the thicker orogenic wedge would provoke NNW-directed extension at the
464
upper part of the tectonic pile.
SC
M AN U
TE D
Unlike the high convergence rates for the ≈ 30-20 Ma period, relatively slow convergence
EP
465
RI PT
453
rates suddenly started to dominate since early-middle Miocene time (Rosenbaum et al., 2002;
467
Vissers and Meijer, 2012). Slow convergence rates could have promoted lithospheric rollback and
468
extension at surface (Royden, 1993; Rosenbaum et al., 2002). Accordingly, while the NNW-SSE
469
shortening in the Alpujárride Complex fits well with fast Africa-Iberia convergence at latest
470
Oligocene to early Miocene time, the top-to-the-SW extension dominant since Langhian-
471
Serravallian time fits well with a slowdown in the convergence rate, which, in turn, would have
472
triggered west-directed lithospheric rollback (Lonergan and White, 1997; Duggen et al., 2008;
473
Mancilla et al., 2015). Thus, at Serravallian time, the new tectonic scenario of slab rollback would
474
have imposed a kinematically distinctive westward extension with exhumation of the Nevado-
AC C
466
ACCEPTED MANUSCRIPT Filábride Complex (Martínez-Martínez et al., 2002, 2004). Notwithstanding, before the dominance
476
of westward extension (or partly coeval to it), contractional shears must have operated to cause
477
the superposition within the Nevado-Filábride Complex of higher pressure metamorphic rocks
478
over lower pressure ones. Shear zones separating these contrasting Nevado-Filábride units have
479
been described by Martínez-Martínez (2007), with the same top-to-the-W kinematics as the
480
extensional shear zones affecting the upper levels of the Nevado-Filábride Complex. However, the
481
similar kinematics raises the possibility that all of the W-directed displacements in the Nevado-
482
Filábride might correspond to extensional tectonics obliterating previous contractional shears.
483
Whichever the case, if partially coeval contractional and extensional shear zones would have
484
existed, they would constitute an efficient crustal flow for the very fast Nevado-Filábride
485
exhumation (Martínez-Martínez, 2007). Since Serravallian time, after the Alpujárrides/Nevado-
486
Filábride superposition, both complexes share the same set of deformational structures.
M AN U
SC
RI PT
475
487 5. Conclusions
489
i) In accordance with previous interpretations, the main penetrative and synmetamorphic
490
Alpine deformation in the Alpujárride rocks is extensional and probably related to exhumation
491
after an initial continental subduction stage of uncertain age. The new map-scale folds described
492
in this paper are rather of local development, denoting that layering would have been rarely
493
positioned in the shortening field during D2A shearing.
AC C
EP
TE D
488
494
ii) Late-metamorphic NW-vergent buckling folds are ubiquitous in the Alpujárride Complex
495
and denote shortening at latest Oligocene-earliest Miocene time. Subsequent semi-brittle and
496
brittle low-angle faults with top-to-the-N kinematics (thrust and low-angle normal faults) cut this
497
train of folds.
498
iii) The top-to-the-N thrust system became blurred but not obliterated by the later low-angle
499
normal faults, with ascending ramps, imbricate fans and duplex structures having been preserved
500
at many places. The documented thrusts are responsible for the stratigraphic and metamorphic
ACCEPTED MANUSCRIPT 501
repetitions observed in the current tectonic pile. Separate mapping of thrusts and low-angle
502
normal faults has been achieved in the central sector of the Alpujárride Complex.
503
iv) The top-to-the-N low-angle normal faults cutting the thrust system might have formed due to accretion at deep levels of the orogenic wedge during underthrusting of the Nevado-Filábride
505
Complex at Burdigalian-Langhian time.
506
RI PT
504
v) A minimum of 100 km of NNW-SSE shortening has been evaluated from restoration of the Alpujárride tectonic pile and the NNW-SSE length of the underlying Nevado-Filábride tectonic
508
window.
SC
507
vi) The internal structure of the Alpujárride Complex can be roughly correlated with the
510
Paleogene-Neogene movements between Africa and Iberia. Thus, the NNW-SSE shortening
511
episode (folds and thrusts) can be related to high convergence rates at latest Oligocene to earliest
512
Miocene time. Thrusting culminates at 16-18 Ma with superposition of the Alpujaride Complex
513
onto the Nevado-Filábride Complex, which in turn reached peak-pressure at that time. Later on,
514
relatively slow Africa/Iberia convergence rates enabled the previously subducted slab to rollback
515
and delaminate, with top-to-the-SW crustal extension becoming dominant since Serravallian time.
TE D
M AN U
509
516
518
EP
517
Acknowledgments
520
This paper has received financial support from the Spanish Ministry of Economy and
AC C
519
521
Competitiveness through Grant CGL2015-71692-P. Thanks to A. Azor for his revision of the
522
manuscript. F. Rossetti and J.M. Martínez-Martínez are also thanked for constructive reviews that
523
have improved the manuscript.
524 525 526
ACCEPTED MANUSCRIPT
TE D
M AN U
SC
RI PT
Acosta, A., Rubatto, D., Bartoli, O., Cesare. B., Meli, S., Pedrera, A., Azor, A., Tajcmanová, L., 2014. Age of anatexis in the crustal footwall of the Ronda peridotites, S Spain. Lithos 210-211, 147-167, doi:10.1016/j.lithos.2014.08.018 Aldaya, F. (1981). Albuñol, Mapa Geológico de España 1:50.000. IGME Aldaya, F., Alvarez, F., Galindo-Zaldívar. J., González-Lodeiro, F., Jabaloy. A., Navarro-Vilá, F., 1991. The Maláguide-Alpujárride contact (Betic Cordilleras, Spain): a brittle extensional detachment. C.R. Acad. Sci. Paris 313, Série II, 1447-1453 Alfaro, P., Andreu, J.M., Bartolomé. R., Borque, M.J., Estévez, A., García-Mayordomo, J., García-Tortosa, F.J., Gil, A., Gràcia, E., Lo Iacono, C., Perea, H., 2012. The Bajo Segura Fault Zone: Active blind thrusting in the Eastern Betic Cordillera (SE Sapin). Journal of Iberian Geology 38, 271-284 Avidad, J., García-Dueñas, V., 1981. Motril, Mapa Geológico de España 1:50.000. IGME Azañón, J.M., Crespo-Blanc, A., 2000. Exhumation during a continental collision inferred from the tectonometamorphic evolution of the Alpujárride Complex in the Central Betics (Alboran Domain, SE Spain) Tectonics 19, 549-565 Azañón, J.M., Goffé, B., 1997. Ferro- and magnesiocarpholite assemblages as record of high-P, low-T metamorphism in the central Alpujárrides, Betic Cordillera (SE Spain).European Journal of Mineralogy 9, 1035-1051 Azañón, J.M., Crespo-Blanc, A., García-Dueñas, V., 1997. Continental collision, crustal thinning and nappe forming during the pre-Miocene evolution of the Alpujárride Complex (Alborán Domain, Betics). Journal of Structural Geology 19, 1055-1071 Azañón, J.M., García-Dueñas, V., Goffé, B., 1998. Exhumation of high-pressure metapelites and coeval crustal extension in the Alpujarride complex (Betic Cordillera). Tectonophysics 285, 231-252 Azañón, J.M., Galve, J.P., Pérez-Peña, J.V., Giaconia, F., Carvajal, R., Booth-Rea, G., Jabaloy, A., Vázquez, M., Azor, A., Roldán, F.J., 2015. Relief and drainage evolution during the exhumation of the Sierra Nevada (Spain): Is denudation keeping pace with uplift? Tectonophysics 663, 19-32, doi: 10.1016/j.tecto.2015.06.015 Balanyá, J.C., García-Dueñas, V., Azañón, J.M., 1997. Alternating contractional and extensional events in the Alpujárride nappes of the Alboran Domain (Betics, Gibraltar Arc). Tectonics 16, 226-238 Bezada, M.J., Humphreys, E.D., Toomey, D.R., Harnafi, M., Dávila J.M., Gallart, J., 2013. Evidence for slab rollback in the westernmost Mediterranean from improved upper mantle imaging. Earth Planet. Sci. Lett. 368, 51-60 Blanco, M.J., Spakman, W., 1993. The P-wave velocity structure of the mantle below the Iberian Peninsula: evidence for subducted lithosphere below southern Spain. Tectonophysics 221, 13-34 Booth-Rea, G., Azañón, J.M., Goffé, B., Vidal, O., Martínez-Martínez, J.M., 2002. High-pressure, lowtemperature metamorphism in Alpujárride Units of southeastern Betics (Spain). C.R. Geoscience 334, 857-865 Booth-Rea, G., Azañón, J.M., García-Dueñas, V., Augier, O., Sánchez-Gómez, M., 2003. A “core-complex-like structure” formed by superimposed extension, folding and high-angle faulting. The Santi Petri dome (western Betics, Spain). C. R. Geoscience 335, 265-274 Booth-Rea, G., Azañón, J.M., Martínez-Martínez, J.M., Vidal, O., García-Dueñas, V., 2005. Contrasting structural and P-T evolution of tectonic units in the southeastern Betics: Key for understanding the exhumation of the Alboran Domain HP/LT crustal rocks (western Mediterranean). Tectonics 24, TC2009, doi: 10.1029/2004TC001640 Bousquet, J.C., 1979. Quaternary strike-slip faults in Southeastern Spain. Tectonophysics 52, 277-286 Butler, R.W.H., 1992. Thrust zone kinematics in a basement-cover imbricate stack: Eastern Pelvoux massif, French Alps. Journal of Structural Geology 14, 29-40 Çoruh, C., Bollinger, G.A., Costain, J.K., 1988. Seismogenic structures in the central Virginia seismic zone. Geology 16, 748-751 Crespo-Blanc, A., 1995. Interference pattern of extensional fault systems: a case study of the Miocene rifting of the Alboran basement (North of Sierra Nevada, Betic Chain). Journal of Structural Geology 17, 1559-1569 Crespo-Blanc, A., Balanyá, J.C., Expósito, I., Luján, M., Díaz-Azpiroz, M., 2007. Acreción miocena del Dominio Suribérico y del Complejo de Flyschs (Arco de Gibraltar): una revisión a partir de las propuestas de V. Garcia-Dueñas. Revista de la Sociedad Geológica de España 20, 135-152
EP
528 529 530 531 532 533 534 535 536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579 580 581
References
AC C
527
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
Crespo-Blanc, A., Orozco, M., García-Dueñas, V., 1994. Extension versus compression during the Miocene tectonic evolution of the Betic chain. Late folding of normal fault systems. Tectonics 13, 78-88 Davis, D., Suppe, J., Dahlen, F.A., 1983. Mechanics of fold-and-thrust belts and acretionary wedges. Journal of Geophysical Research 88, 1153-1172 De Jong, K., 1993. Large scale polyphase deformation of a coherent HP/LT metamorphic unit: the Mulhacén Complex in the eastern Sierra de los Filabres (Betic Zone, SE Spain). Geol. Mijnb. 71, 327-336 DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophys. J. Int. 101, 425-478 Dewey, J.F., Helman, M.L., Turco, E., Hutton, D.H.W., Knott, S.D., 1989. Kinematics of the Western Mediterranean. In: Coward MP, Dietrich D, Park RG (Ed) Alpine Tectonics. Geological Society, London, Special Publication 45, 265-283 Duggen, S., Hoernle, K., Klügel, A., Gelmacher, J., Thirlwall, M., Hauff, F., Lowry, D., Oates, N., 2008. Geochemical zonation of the Miocene Alborán Basin volcanism (westernmost Mediterranean): geodynamic implications. Contributions to Mineralogy and Petrology 156, 577-593 Echeverría, A., Khazaradze, G., Asensio, E., Masana, E., 2015. Geodetic evidence for continuing tectonic activity of the Carboneras fault (SE Spain) Tectonophysics 663, 302-309, doi: 10.1016/j.tecto.2015.08.009 England, Ph., Molnar, P., 1993. Cause and effect among thrust and normal faulting, anatectic melting and exhumation in the Himalaya. Geological Society London, Special Publication 74, 401-411, doi: 10.1144/GSL.SP.1993.074.01.27 Estévez, A, Delgado, F., Sanz de Galdeano, C., Martín Algarra, A., 1985. Los Alpujárrides al sur de Sierra Nevada. Una revisión de su estructura. Mediterranea 4, 5-32 Faccenna, C., Piromallo, C., Crespo-Blanc, A., Jolivet, L., Rossetti, F., 2004. Lateral slab deformation and the origin of the Western Mediterranean arcs. Tectonics 23, http://dx.doi.org/10.1029/2002TC001488 Galindo-Zaldívar, J., González Lodeiro, F., Jabaloy. A., 1989. Progressive extensional shear structures in a detachment contact in the western Sierra Nevada (Betic Cordilleras, Spain). Geodinamica Acta 3, 73-85 Galindo-Zaldívar, J., Jabaloy, A., González-Lodeiro, F., Aldaya, F., 1997. Crustal structure of the central sector of the Betic Cordillera (SE Spain). Tectonics 16, 18-37 Galindo-Zaldívar, J., Gil, A.J., Sanz de Galdeano, C., Lacy, M.C., García-Armenteros, J.A., Ruano, P., Ruiz, A.M., Martínez-Martos, M., Alfaro, P., 2015. Active shallow extension in central and eastern Betic Cordillera from CGPS data. Tectonophysics 663, 290-301, doi: 10.1016/j.tecto.2015.08.035 García-Dueñas, V., Balanyá, J.C., Martínez-Martínez, J.M., 1992. Miocene extensional detachments in the outcropping basement of the northern Alboran Basin and their tectonic interpretation. Geo-Marine Letters 12, 88-95 Goffé, B., Michard, A., García-Dueñas, V., González-Lodeiro, F., Monié, P., Campos, J., Galindo-Zaldívar, J., Jabaloy, A., Martínez-Martínez, J.M., Simancas, J.F., 1989. First evidence of high-pressure, lowtemperature metamorphism in the Alpujárride nappes, Betic Cordilleras (SE Spain). European Journal of Mineralogy 1, 139-142 Gómez-Pugnaire, M.T., Braga, J.C., Martín, J.M., Sassi, F.P., Moro, A., 2000. Regional implications of a Palaeozoic age for the Nevado-Filábride cover of the Betic Cordillera, Spain. Schw. Min. Petrog. Mitt. 80, 45-52 González-Castillo, L., Galindo-Zaldívar, J., de Lacy, M.C., Borque, M.J., Martínez-Moreno, F.J., GarcíaArmenteros, J.A., Gil, A.J., 2015. Active rollback in the Gibraltar Arc: Evidences from CGPS data in the western Betic Cordillera. Tectonophysics 663, 310-321, doi: 10.1016/j.tect.2015.03.010 González-Lodeiro, F., Aldaya, F., Galindo-Zaldívar, J., Jabaloy, A., 1996. Superposition of extensional detachments during the Neogene in the internal zones of the Betic cordillera. Geol. Rundschau 85, 350-362 Hobbs, B., Means, W.D., Williams, P.F., 1976. An Outline of Structural Geology. John Wiley & Sons, Inc., New York, 571 p. Jabaloy, A., Galindo-Zaldívar, J., González-Lodeiro, F., 1993. The Alpujárride - Nevado-Filábride extensional shear zone, Betic Cordillera, SE Spain. Journal of Structural Geology 15, 555-569 Jolivet, L., Faccenna, C., 2000. Mediterranean extension and the Africa-Eurasia collision. Tectonics 19 10951106, doi: 10.1029/2000TC900018 Lonergan, L., Platt, J.P., 1995. The Malaguide-Alpujarrid boundary: a major extensional contact in the Internal Zone of the eastern Betic Cordillera, SE Spain. Journal of Structural Geology 17, 1655-1671 Lonergan, L., White, N., 1997. Origin of the Betic-Rif mountain belt. Tectonics 16, 504-522
AC C
582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623 624 625 626 627 628 629 630 631 632 633 634 635 636 637
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
López Sánchez-Vizcaíno, V., Rubatto, D., Gómez-Pugnaire, M.T., Trommsdorff, V., Müntener, O., 2001. Midle Miocene high-pressure metamorphism and fast exhumation of the Nevado-Filábride Complex, SE Spain. Terra Nova 13, 327-332 Mancilla, F.L., Booth-Rea, G., Stich, D., Pérez-Peña, J.V., Morales, J., Azañón, J.M., Martin, R., Giaconia, F., 2015. Slab rupture and delamination under the Betics and Rif constrained from receiver functions. Tectonophysics 663, 225-237, doi: 10.1016/j.tecto.2015.06.028 Martínez-Martínez, J.M., 2007. Coexistencia de zonas de cizalla dúctil de extensión y de acortamiento en el domo de Sierra Nevada, Béticas (SE de España). Revista de la Sociedad Geológica de España, 20, 229-246 Martínez-Martínez, J.M., Soto, J.I., Balanyá, J.C., 2002. Orthogonal folding of extensional detachments: Structure and origin of the Sierra Nevada elongated dome (Betics, SE Spain), doi: 10.1029/2001TC0012983 Martínez-Martínez, J.M., Soto, J.I., Balanyá, J.C., 2004. Elongated domes in extended orogens: A mode of mountain uplift in the Betics (southeast Spain). The Geological Society of America, Special Paper 380, 243-266 Mayoral, E., Crespo-Blanc, A., Díaz, M.G., Benot, C., Orozco, M., 1994. Rifting miocène du Domaine d’Alboran: datations de sédiments discordants sur les unités alpujarrides en extension (Sud de la Sierra Nevada, Chaîne Bétique). C. R. Acad. Sci. Paris 319(II), 581-588 Morales, J., Serrano, I., Jabaloy, A., Galindo-Zaldívar, J., Zhao, D., Torcal, F., Vidal, F., González-Lodeiro, F., 1999. Active continental subduction beneath the Betic Cordillera and Alboran Sea. Geology 27, 735-738 Navarro-Vilá, F., García Dueñas, V., 1980. La Peza, Mapa Geológico de España 1:50.000. IGME Orozco, M., Alonso-Chaves, F.M., Nieto, F., 1998. Development of large north-facing folds and their relation to crustal extension in the Alborán domain (Alpujarras region, Betic Cordilleras, Spain). Tectonophysics 298, 271-295 Orozco, M., Álvarez-Valero, A.M., Alonso-Chaves, F.F., Platt, J.P., 2004. Internal structure of a collapsed terrain. The Lújar syncline and its significance for the fold- and sheet-structure of the Alborán Domain (Betic Cordilleras, Spain). Tectonophysics 385, 85-104, doi: 10.1016/j.tecto.2004.04.025 Orozco, M., Alonso-Chaves, F.M., Platt, J.P., 2017. Late extensional shear zones and associated recumbent folds in the Alpujárride subduction complex, Betic Cordillera, southern Spain. Geologica Acta 15, 51-66, doi: 10.1344/GeologicaActa2017.15.1.5 Pedrera, A., Galindo-Zaldívar, J., Sanz de Galdeano, C., López-Garrido, A.C., 2007. Fold and fault interactions during the development of an elongated narrow basin: the Almanzora Neogene-Quaternary Corridor (SE Betic Cordillera, Spain).Tectonics 26, TC6002, doi: 10.1029/2007TC002138 Platt, J.P., 1986. Dynamic of orogenic wedges and the uplift of high-pressure metamorphic rocks. Geological Society of America Bulletin 97, 1037-1053 Platt, JP., Vissers, R.L.M., 1989. Extensional collapse of thickened continental lithosphere: a working hypothesis for the Alboran Sea and the Gibraltar arc. Geology 17, 540-543 Platt, J.P., Soto, J.I., Whitehouse, M.J., Hurford, A.J., Kelley, S.P., 1998. Thermal evolution, rate of exhumation, and tectonic significance of metamorphic rocks from the floor of the Alboran extensional basin, western Mediterranean. Tectonics 17, 671-689 Platt, J.P., Kelley, S.P., Carter, A., Orozco, M., 2005. Timing of tectonic events in the Alpujárride Complex, southern Spain. Journal of the Geological Society, London, 162, 451-462. Platt, J.P., Anczkiewicz, S.P., Soto, J.I., Kelley, S.P., Thirlwall, M., 2006. Early Miocene continental subduction and rapid exhumation in the western Mediterranean. Geology 34, 981 Puga, E., Díaz de Federico, A., Nieto, J.M., 2002. Tectonostratigraphic subdivisión and petrological characterization of the deepest complexes of the Betic Zone: a review. Geodinamica Acta 15, 23-43 Rosenbaum, G., Lister, G.S., Duboz, C., 2002. Relative motions of Africa, Iberia and Europe during Alpine orogeny. Tectonophysics 359, 117-129 Rossetti, F.C., Faccenna, C., Crespo-Blanc, A. 2005. Structural and kinematic constraints to the exhumation of the Alpujárride Complex (Central Betic Cordillera, Spain). Journal of Structural Geology 27, 199216, doi: 10.1016/j.jsg.2004.10.008 Royden, L.H., 1993. Evolution of retreating subduction boundaries formed during continental collision. Tectonics 12, 629-638 Ruiz-Constán, A., Pedrera, A., Galindo-Zaldívar, J., Pous, J., Arzate, J., Roldán-García, F.J., Marín-Lechado, J., Anahnah, F., 2012. Constraints on the frontal crustal structure of a continental collision from an
AC C
638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667 668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693
ACCEPTED MANUSCRIPT integrated geophysical research: The central-western Betic Cordillera (SW Spain). Geochemistry, Geophysics, Geosystems 13 Sánchez-Navas, A., García-Casco, A., Martín-Algarra, A. 2014. Pre-Alpine discordant granitic dikes in the metamorphic core of the Betic Cordillera: tectonic implications. Terra Nova 26, 477-486, doi: 10.1111/ter.12123 Sanz de Galdeano, C., 1990. Geologic evolution of the Betic Cordilleras in the Western Mediterranean, Miocene to the present. Tectonophysics 172, 107-119 Sanz de Galdeano, C., Alfaro, P., 2004. Tectonic significance of the present relief of the Betic Cordillera. Geomorphology 63, 175-190 Sanz de Galdeano, C., López Garrido, A.C., 2014. Structure of the Sierra de Lújar (Alpujárride Complex, Betic Cordillera). Estudios Geológicos 70, e005, doi: 10.3989/egeol.41491.290 Sanz de Galdeano, C., Peláez, J.A., 2011. Fallas activas en la Cordillera Bética. Editorial Universidad de Granada Silva, P.G., Goy, J.L., Somoza, L., Zazo, C., Bardají, T., 1993. Landscape response to strike-slip faulting linked to collisional setting: Quaternary tectonics and basin formation in the Eastern Betics, southeastern Spain. Tectonophysics 224, 289-303 Simancas, J.F., Campos, J., 1993. Compresión NNW-SSE tardi a postmetamórfica y extensión subordinada en el Complejo Alpujárride (Dominio de Alborán, Orógeno Bético). Revista de la Sociedad Geológica de España 6, 23-35 Tubía, J.M., Cuevas, J., Navarro-Vilá, F., Álvarez, F., Aldaya, F., 1992. Tectonic evolution of the Alpujárride Complex (Betic Cordillera, southern Spain). Journal of Structural Geology 14, 193-203 Vera, J., Martín-Algarra, A., 2004. Cordillera Bética y Baleares. Divisiones mayores y nomenclatura. In: Vera, J. (Ed.) Geología de España. SGE-IGME, Madrid, 347-350 Vissers, R.L.M., Meijer, P.Th., 2012. Iberian plate kinematics and Alpine collision in the Pyrenees. EarthScience Reviews 114, 61-83, doi: 10.1016/j.earscirev.2012.05.001 Vissers, R.L.M., Platt, J.P., van der Wal, D., 1995. Late orogenic extension of the Betic Cordillera and the Alboran Domain: A lithospheric view. Tectonics 14, 786-803 Williams, J.R., Platt, J.P., 2017. Superposed and refolded metamorphic isograds, and superposed directions of shear during late-orogenic extension in the Alborán Domain, southern Spain. Tectonics 36, 756786, doi: 10.1002/2016TC004358 Yamato, P., Brun, J.P., 2017. Metamorphic record of catastrophic pressure drops in subduction zones. Nature geoscience 10, doi: 10.1038/NGE02852
732 733 734 735 736 737 738 739 740 741 742 743 744
Fig. 1 A) Location and main geological domains of the Betic Orogen. B) Sketched structural map of the central sector of the Alpujárride Complex. Location of Figs. 3 and 7 is given, as well as the crosssections shown in Figs. 4 and 5. C) Lithological sequence of the Alpujárride Complex; notice tapering to the east. D) Equal-area and lower-hemisphere projections of linear structures of the successive deformational events (see text for further explanation).
EP
TE D
M AN U
SC
RI PT
694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711 712 713 714 715 716 717 718 719 720 721 722 723 724 725 726 727 728 729 730 731
AC C
Figure captions
Fig. 2 A) Main D2A deformation in marbles, with stretching lineation (Ls2A) and asymmetric tails indicating top-to-the-NE sense of shear. B) Two generations of microfolds in marbles: F2A, pointed by the arrow, and F3A. C) Hinge of an F3A syncline in marbles, with axial-planar foliation at right angle to layering; the fold is overturned to the NW. D) Second-order M-type folds at the hinge strain shadow of a first-order F3A fold, developed in quartzites and phyllites underlying marbles. E) C’-
ACCEPTED MANUSCRIPT type shear bands in low-grade schists, indicating top-to-the-N semibrittle shearing. F) Crush breccia at the thrust contact between two carbonate slabs.
RI PT
Fig. 3 A) Structural map of the area east of the Sierra de Lújar, showing the nearly coaxial but opposite vergent superposition of the D2A and D3A folding phases; the location of the cross-sections in Fig. 4 is indicated. B) Equal-area and lower-hemisphere projections of foliations (S2A and S3A) and fold axes (F2A and F3A) of the two superposed folds. Fig. 4 A) Geological cross-section of the structure east of the Sierra de Lújar (see location in Fig. 3A). B) Detailed cross-section showing local controls of vergence and polarity.
M AN U
SC
Fig. 5 Geological cross-sections of the Alpujárride Complex (location in Fig. 1B). The structure shown is made up mainly of NW-vergent folds (D3A), top-to-the-N thrusts (D4A1) and top-to-the-N low-angle normal faults (D4A2). Additionally, some late structures (top-to-the-SW normal faults and upright open folds) are also represented in the cross-sections. See full explanation in the text. The nomenclature and relative position of the tectonic sheets distinguished in the central sector of the Alpujárride Complex is given in the lower part of the figure.
TE D
Fig. 6 Restoration of the cross-section A in Fig. 5. The original position of the Guájares, Salobreña and Herradura tectonic sheets is shown, and the NNW-SSE shortening is evaluated.
EP
Fig. 7 Structural map of the area north of Motril, to document the internal structure of the Herradura Paleozoic schists and their position onto the Escalate Triassic marbles (see text for further explanation). See location in Fig. 1B.
AC C
745 746 747 748 749 750 751 752 753 754 755 756 757 758 759 760 761 762 763 764 765 766 767 768 769 770 771 772 773
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT
Structural map of the area east of the Sierra de Lújar
A) 80
RI PT
75
N
30
Quaternary
B)
1000 m
743
Sierr a
de Ju b
Structural data
iley
(equal-area projection, Kamb contours)
aJ oy a
865
35
45
ra d Sie r
1076
80
875
30
poles to S0/S1A and S2A
Road to Torvizcón
el
Road to Órgiva
M AN U
SC
792
n=117
poles to S3A
TE D
1184
Sierra de Lújar
35
n=19
35 F2A fold axes
10
40
Cross-section Fig. 4A
AC C
20
EP
45
40 n=76
Marbles
Syncline
F3A fold axes
Phyllites and quartzites
Anticline
35
50
75
65
Axial trace of DA2 folds
Cross-section Fig. 4B
n=56
Axial trace of DA3 folds
Thrust
High-angle normal fault
ACCEPTED MANUSCRIPT
Sierra de la Joya
A)
WNW - ESE
Top-to-the north thrust
1250 m
RI PT
Sierra de Lújar 1000 m
M M: marbles Ph: phyllites and quartzites
750 m
Ph
topography
500 m
1000 m
SC
W-E
B)
east local F1A vergence
road to TV antene
inverted
inverted
AC C
EP
TE D
east local F1A vergence
west local F1A vergence
S1A
g
normal
normal
bedding polarity (inverted)
din
east local F1A vergence
M AN U
topography
d be
200 m
west local F1A vergence
east local F1A vergence
inverted
ACCEPTED MANUSCRIPT
NNW-SSE
N-S
N-S
Granja Escuela
Northern klippes
A)
Almuñécar
Otívar
Poor control of the structure
SW extension
RI PT
N extension
G
G
Sa
H
km (V=H)
B)
1
4
3
2
5
SC
0
N-S
N-S Sa
H
0m
Sa
2
3
4
0
5
E) Sierra de Lújar
N-S
L1 Ad
1000 m 500 m
L2
L1
Pile of tectonic sheets (W to E):
1
2
AC C
km (V=H)
0
3
4
2
3
4
D)
Motril
N-S
topography
H
H 0m
L2
H km (V=H)
5
0
1
2
4
3
5
F) N-S
EP
M
0m
1
TE D
1
H
km (V=H)
km (V=H)
0
0m
0m
Sa
topography
Sa
Al (G)
Sa
G
M AN U
C)
Al
Rambla de Huarea
L3
M
El Pozuelo
M topography
Ad
Ad
M
0m
L3
L2
5
G: Guájares tectonic sheet (Al: Alberquilla extensional unit) Sa: Salobreña tectonic sheet ...................................... Ad: Adra tectonic sheet (similar position to the one of Salobreña) H: Herradura tectonic sheet ..................................... M: Murtas tectonic sheet (similar position to the one of Herradura) L1: first imbricate of the L2: second imbricate of the L3: third imbricate of the Lújar tectonic sheet Lújar tectonic sheet Lújar tectonic sheet (Sierra de Lújar) (Sierra de Turón) (Sierra Alhamedilla)
km (V=H)
0
1
Top-to-the-N thrust Top-to-the-N low-angle normal fault Top-to-the-SW low-angle normal fault
2
3
4
5
ACCEPTED MANUSCRIPT
RI PT
NNW - SSE Marbles
G
3
2
4
5
G: Guájares tectonic sheet (A: Alberquilla extensional unit) S: Salobreña tectonic sheet: foreland dipping duplex structure H: Herradura tectonic sheet
H
5 km (V=H) 0
1
3
2
4
A
SC
1
Sa
H
M AN U
5 km (V=H) 0
Metapelites and metasandstones
G
5
Extensional faults removed
0
1
2
3
4
5
AC C
5 km (V=H)
EP
Herradura
TE D
11 km
Thrusts removed
Guájares Salobreña
50 km
Shortening due to top-to-the-N thrusts: 50-11=39 km Shortening due to folding: 60-50=10 km Total shortening: 49 km
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
ACCEPTED MANUSCRIPT Highlights Thrust orogenic wedge obliterated by low-angle normal faults Top-to-the-N extension caused by crustal underplating Top-to-the-SW extension caused by lithospheric rollback Lithospheric rollback prompted by slow Africa-Iberia convergence rates
AC C
EP
TE D
M AN U
SC
RI PT
• • • •