22
Neotectonic evolution of the Central Betic Cordilleras (Southern Spain) Klaus R. Reicherter a, * , Gwendolyn Peters b,1 a Institut fu ¨r Geophysik und Geologie, Universita ¨t Leipzig, Talstr. 35, D-04103 Leipzig, Germany b Geophysikalisches Institut, Universita ¨t Karlsruhe, Hertzstr. 16, D-76187 Karlsruhe, Germany Received 15 July 2004; received in revised form 18 May 2005; accepted 25 May 2005 Available online 14 July 2005 Abstract Paleostress orientations were calculated from fault-slip data of 36 sites located along a traverse through the Central Betic Cordilleras (southern Spain). Heterogeneous fault sets, which are frequent in the area, have been divided into homogeneous subsets by cross-cutting relationships observed in the field and by a paleostress stratigraphy approach applied on each individual fault population. The state of stress was sorted according to main tectonic events and a new chronology is presented of the Miocene to Recent deformation in the central part of the Betic Cordilleras. The deviatoric stress tensors fall into four distinct groups that are regionally consistent and correlate with three Late Oligocene–Aquitanian to Recent major tectonic events in the Betic Cordilleras. The new chronology of the neotectonic evolution includes, from oldest to youngest, the following main tectonic phases: (1) Late Oligocene–Aquitanian to Early Tortonian: r 1 subhorizontal N–S, partly E–W directed, r 3 subvertical; compres- sional structures (thrusting of nappes, large-scale folding) and strike-slip faulting in the Albora ´n Domain and the External Zone of the Betic Cordilleras; (2) Early Tortonian to Pliocene–Pleistocene: r 1 subvertical, r 3 subhorizontal NW–SE, partly N–S directed or E–W-directed (radial extension); large-scale normal faulting in the Central Betic Cordilleras and in the oldest Neogene formations of the Granada Basin related to the gravitational collapse of the Betic Cordilleras and the exhumation of the intensely metamorphosed rock series of the Internal Zones, at the same time formation of the Albora ´n Basin and intramontane basins such as the Granada Basin; (3) Pleistocene to Recent: (3a) r 1 subvertical, r 3 subhorizontal NE–SW with prominent normal faulting, but coevally; (3b) r 1 subhorizontal NW directed, r 3 NE–SW subhorizontal with strike-slip faulting. Extensional structures and strike-slip faulting are related to the ongoing convergence of the Eurasian and African Plates and coeval uplift of the Betic 0040-1951/$ - see front matter D 2005 Elsevier B.V. All rights reserved. doi:10.1016/j.tecto.2005.05.022 * Corresponding author. Tel.: +49 341 9732911; fax: +49 341 9732809. E-mail address: [email protected] (K.R. Reicherter). 1 Present address: Tectonics Department, Faculty of Earth and Life Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1085, 1081 HV Amsterdam, The Netherlands. Tectonophysics 405 (2005) 191– 212 www.elsevier.com/locate/tecto

Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Embed Size (px)

Citation preview

Page 1: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

www.elsevier.com/locate/tecto

Tectonophysics 405 (

Neotectonic evolution of the Central Betic Cordilleras

(Southern Spain)

Klaus R. Reicherter a,*, Gwendolyn Peters b,1

aInstitut fur Geophysik und Geologie, Universitat Leipzig, Talstr. 35, D-04103 Leipzig, GermanybGeophysikalisches Institut, Universitat Karlsruhe, Hertzstr. 16, D-76187 Karlsruhe, Germany

Received 15 July 2004; received in revised form 18 May 2005; accepted 25 May 2005

Available online 14 July 2005

Abstract

Paleostress orientations were calculated from fault-slip data of 36 sites located along a traverse through the Central Betic

Cordilleras (southern Spain). Heterogeneous fault sets, which are frequent in the area, have been divided into homogeneous

subsets by cross-cutting relationships observed in the field and by a paleostress stratigraphy approach applied on each individual

fault population. The state of stress was sorted according to main tectonic events and a new chronology is presented of the

Miocene to Recent deformation in the central part of the Betic Cordilleras. The deviatoric stress tensors fall into four distinct

groups that are regionally consistent and correlate with three Late Oligocene–Aquitanian to Recent major tectonic events in the

Betic Cordilleras. The new chronology of the neotectonic evolution includes, from oldest to youngest, the following main

tectonic phases:

(1) Late Oligocene–Aquitanian to Early Tortonian: r1 subhorizontal N–S, partly E–W directed, r3 subvertical; compres-

sional structures (thrusting of nappes, large-scale folding) and strike-slip faulting in the Alboran Domain and the External

Zone of the Betic Cordilleras;

(2) Early Tortonian to Pliocene–Pleistocene: r1 subvertical, r3 subhorizontal NW–SE, partly N–S directed or E–W-directed

(radial extension); large-scale normal faulting in the Central Betic Cordilleras and in the oldest Neogene formations of the

Granada Basin related to the gravitational collapse of the Betic Cordilleras and the exhumation of the intensely

metamorphosed rock series of the Internal Zones, at the same time formation of the Alboran Basin and intramontane

basins such as the Granada Basin;

(3) Pleistocene to Recent: (3a) r1 subvertical, r3 subhorizontal NE–SW with prominent normal faulting, but coevally; (3b)

r1 subhorizontal NW directed, r3 NE–SW subhorizontal with strike-slip faulting. Extensional structures and strike-slip

faulting are related to the ongoing convergence of the Eurasian and African Plates and coeval uplift of the Betic

0040-1951/$ - s

doi:10.1016/j.tec

* Correspondi

E-mail addre1 Present addr

Amsterdam, The

2005) 191–212

ee front matter D 2005 Elsevier B.V. All rights reserved.

to.2005.05.022

ng author. Tel.: +49 341 9732911; fax: +49 341 9732809.

ss: [email protected] (K.R. Reicherter).

ess: Tectonics Department, Faculty of Earth and Life Sciences, Vrije Universiteit Amsterdam, De Boelelaan 1085, 1081 HV

Netherlands.

Page 2: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212192

Cordilleras. Reactivation of pre-existing fractures and faults was frequently observed. Phase 3 is interpreted as periodic

strike-slip and normal faulting events due to a permutation of the principal stress axes, mainly r1 and r2.

D 2005 Elsevier B.V. All rights reserved.

Keywords: Neotectonics; Paleostress; Fault-slip data; Deformation history; Betic Cordilleras

1. Introduction observed in tectono-stratigraphic units of the External

In this paper a paleostress and structural analysis is

presented of the central part of the Betic Cordilleras in

a corridor from the Mediterranean coast east of Ma-

laga to the southern border of the Guadalquivir Basin

(Fig. 1). This paper concentrates on the Miocene to

Quaternary deformation observed in the Neogene–

Quaternary sediments of the Granada Basin, and use

the deformation information (i.e. trend, character and

timing of the faulting) to calibrate the deformation

Fig. 1. A. Geological overview of the Betic Cordilleras and the working are

directed maximum in the Betics, and minor NW–SE and E–W directed faul

= Carboneras Fault Zone; GF = Gafarillos Fault; PF = Palomares Fault;

working area and geological main units of the Iberian Peninsula.

Zone and the Alboran Domain. We applied fault-slip

analysis with a paleostress stratigraphy approach,

which is complemented by joint and fractures analysis

as well lineations from satellite and aerial images.

Finally, a compilation is presented of the deformation

through time of the neotectonic history along the

studied transect.

The corridor investigated from the Mediterranean

coast to the southern border of the Guadalquivir Basin

crosses all tectono-stratigraphic units, from the inter-

a. Major faults and satellite lineations show a dominant ENE–WSW

ts occur. AF = Alhama de Murcia Fault; CF = Crevillente Fault; CFZ

TF = Tiscar Fault; VF = Ventas de Zafarraya Fault. B. Inset shows

Page 3: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 193

nal parts of the Betics to the foreland, including

Neogene intra-montane basins. Therefore a short sum-

mary of those geological units is provided.

1.1. Geological setting of the Betic Cordilleras

The Betic Cordilleras of southern Spain form

the northern branch of the Gibraltar Arc, the western-

most termination of the Alpine orogenic system. Clas-

sically, the arcuate orogen is subdivided in a

predominantly metamorphic core complex, the

Alboran Domain, comprising the Betic and Rif Inter-

nal Zone, surrounded by an external fold-and-thrust

belt (Fig. 1). From the Late Cretaceous onwards,

continuing convergence of Africa with respect to

Europe (Reicherter and Pletsch, 2000; Rosenbaum et

al., 2002) defines the tectonic framework for the

Iberian subplate. The structure of the Betic orogen is

due to the oblique convergence and collision of the

Iberian plate with the African plate, which was initi-

ated in the Santonian/Campanian (De Jong, 1990;

Reicherter and Pletsch, 2000). The collision led to

metamorphism and thrusting within the Alboran Do-

main (e.g., Sanz de Galdeano, 1997).

Whether subduction of oceanic crust was involved

in the mountain building process within the Betic

Cordilleras is still doubtful (discussion in Platt et al.,

2003). During the collision, the Alboran Domain

underwent approximately 300 km of westward move-

ment, from a more easterly position in the context of

the Alpine Cretaceous–Paleogene orogenic system

(Martınez-Martınez and Azanon, 2002). In the Early

Miocene the westward movement of the Alboran

Domain caused thinning of the thrust sheets (Bakker

et al., 1989). A widespread orogenic collapse in the

Western Mediterranean region led to extension in the

Alboran Domain and the Liguro-Provencalian Basin

during the Upper Oligocene and Burdigalian (Platt

and Vissers, 1989; Calvert et al., 2000). Convective

removal of the lithospheric root in combination with

N–S directed compression seems to have resulted in a

westward tectonic escape (Martınez-Martınez and

Azanon, 2002). During the Burdigalian–Langhian

gravitational collapse of the metamorphic core com-

plex started with extension towards the (N)NW–(S)SE

along low-angle normal faults, followed by WSW

directed orogen-parallel extension during the Serrava-

lian (Galindo-Zaldıvar et al., 1989; Martınez-Martınez

and Azanon, 2002). From the Tortonian until present

the dominating stress regime in the Betic Cordilleras

is extensional with varying extension directions in an

overall compressive plate boundary setting (Galindo-

Zaldıvar et al., 1999).

1.1.1. Alboran domain or internal zone

The Alboran Domain or Internal Zone of the Betic

and Rifian Cordilleras is believed to form part of a

continental microplate with a complex tectonic and

metamorphic history (Vissers et al., 1995). The

Alboran Domain forms the core of the orogen (Fig.

1) and comprises the remains of an older collisional

mountain belt formed during Late Cretaceous and

Paleogene times (Platt et al., 2003), with a certain

evidence for a pre-Alpine (Variscan or older) meta-

morphic episode (Gomez Pugnaire et al., 2000).

Mainly Precambrian to Paleozoic and Triassic sedi-

mentary rocks were subdivided on the basis of the

different metamorphic overprint from top to base into

the Malaguide-Complex, the Alpujarride-Complex,

and the Nevado-Filabride-Complex (summary in Wei-

jermars, 1991; Azanon et al., 2002). These units

represent three thrust sheet complexes associated

with stacking and thickening during the Oligocene

and Late Miocene (Jolivet and Faccenna, 2000),

later deformed by extensional detachments (Fernan-

dez-Fernandez et al., 1992). Low- to high-grade meta-

morphosed schists and marbles form the Alpujarride-

Complex, while the Malaguide-Complex includes

Variscan basement covered by Mesozoic sediments.

The Internal-External Zone Boundary (IEZB) is a

major tectonic contact in the Central Betic Cordilleras

and separates the Internal Zones from the External

Zones (Lonergan et al., 1994). The IEZB forms a low

angle detachment (N508E strike and 258 dip to NW)

in the studied area with a top-to-the-W sense of

movement and leading to Neogene intramontane

basin formation, such as the Granada Basin

(Galindo-Zaldıvar et al., 2000). The E–W trending

mountain chains of the Sierras Tejeda and Almijara

(Fig. 2) are situated between the Granada Basin and

the Mediterranean coast and consist of Paleozoic

phyllites, quartzites and Triassic marbles of the Alpu-

jarride-Complex (e.g., Gallegos, 1975; Martın Martın,

1980). Outcrops of the weaker metamorphosed Mala-

guide-Complex with Silurian phyllites are found

along the western border of the Sierra Tejeda

Page 4: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212194

(IGME, 1979–1988). In the southern part of the study

area the Alboran Domain is locally overlain by nappes

of the Flysch Zone (Figs. 1 and 2).

1.1.2. Flysch zone

The Flysch Zone in the western Betic Cordilleras is

comprised of a stack of nappes consisting of silici-

clastic sediments of Cretaceous to Early Miocene age

(Figs. 1 and 2), mainly deposited along the passive

margin of Africa on thinned continental crust (Reich-

erter et al., 1994), but also partly on oceanic crust

(Durand-Delga et al., 2000). The nappes of the Flysch

Zone of the Campo de Gibraltar make up most of the

western part of the Cordilleras, whereas small slices

are incorporated in thrust sheets between Ronda and

Granada. The Sierra Tejeda is overlain by flysches of

the Vinuela Formation and the Colmenar–Periana

Complex in its western part (Guerrera et al., 1993,

Fig. 2. Geological map of the central Betic Cordilleras with the locality of

External Boundary Zone.

Fig. 2). Deposition of a breccia at the base of the

Vinuela Formation, with clastic components of the

metamorphic Alboran Domain, is interpreted to indi-

cate the end of the westward movement of the Internal

Zone during the Aquitanian. During the Early Burdi-

galian sedimentation of the Vinuela Formation was

terminated by overthrusting of the Flysch Zone (Bour-

gois, 1978; Sanz de Galdeano et al., 1993).

1.1.3. External zone

The External Zone of the Betics consists of non-

metamorphic sedimentary successions of Triassic to

Paleogene age, representing the Mesozoic south Ibe-

rian margin (Figs. 1 and 2). The different paleogeo-

graphic units were deformed during intracontinental

collision in the Miocene, forming now a thin-skinned

fold-and-thrust belt (e.g., Garcıa-Hernandez et al.,

1980; Lonergan and White, 1997). The External

the paleostress stations, note section line of Fig. 3. IEZB = Internal-

Page 5: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 195

Zone is made up of discrete paleogeographic domains:

(1) the Subbetic (subdivided into External, Median

and Internal Subbetic, Fig. 2); (2) the Penibetic; and

(3) the Prebetic nappes. Generally, all units are com-

prised of up to 1000 m of Lower Jurassic limestones

and dolomites, followed by approximately 500 m of

marly limestones and limestones of Middle Jurassic to

Late Cretaceous age (Garcıa-Hernandez et al., 1980).

The Triassic rocks at the base of the stratigraphic

sequences, especially the Keuper evaporites, acted

as decollement horizons during nappe emplacement

and were also subject to tectonically induced and

compressive diapirism in the Late Cretaceous (Reich-

erter and Pletsch, 2000). The individual tectono-strati-

graphic evolution of these paleogeographic units

during the Mesozoic has been outlined by Reicherter

et al. (1994).

1.1.4. Neogene basins

The most recent rocks in the study area are the

Neogene to Quaternary sediments of the Granada

Fig. 3. Simplified stratigraphic section with the timing of faulting events

Location of section see Fig. 2.

Basin (Fig. 2). The transition from sequences repre-

senting Tortonian marine sedimentation to post-Torto-

nian continental deposits, with major unconformities

related to tectonic and/or eustatic events, displays the

complex tectono-sedimentary evolution of the intra-

montane basin (Fernandez et al., 1996a). Basin-wide

sedimentation commenced with marine sediments in

the Early Tortonian, characterized by calcarenites and

marls as well as breccias (Fig. 3). A very well-pro-

nounced intra-Tortonian unconformity is interpreted

to be of tectonic origin (Fernandez et al., 1996a).

During the uppermost Tortonian–lowermost Messi-

nian the depositional environment changed to conti-

nental conditions with sedimentation of proximal

fluvial conglomerates and distal sandstones and marl-

stones (Fig. 3). The central part of the basin evapo-

rated and filled with gypsum and halite. On top of this

sequence upper were deposited Messinian to lower

Pliocene lacustrine limestones and terrigenous sand-

stones of the Turolian mammal stage. These marls

include lignitic layers and lacustrine gastropod-rich

in the Neogene Granada Basin, modified after Braga et al. (1990).

Page 6: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212196

limestones (e.g., near Arenas del Rey, Bandel et al.,

2000). The Pliocene to Pleistocene is represented by

thick alluvial conglomerates and sands of Piedmont

and Glacis sediments (Lhenaff, 1979; Fernandez et al.,

1996b), locally intercalated by clays and travertines.

The entire basin fill has a thickness of 500 m along the

basin margins (Braga et al., 1990) and occasionally

exceeds 1000 m in depocenters close to the Sierra

Elvira (Fig. 2) northwest of Granada (Morales et al.,

1990; Rodrıguez Fernandez and Sanz de Galdeano,

2001). The sedimentary infill of the Granada Basin

experienced a northward tilt of the sequences. The

present day sedimentation takes place along a NW–SE

trending fault-bounded graben W of Granada (Fig. 2).

1.2. Seismological setting

The Betic Cordilleras are located within the Afro-

European Convergence Zone in an approximately 600

km long and 200 km wide region with a dispersed

seismicity. The central and eastern Betic Cordilleras,

especially the intramontane basins of Neogene age,

form the most active seismic zones in Spain with

several moderate earthquakes reported during the

last few centuries (Morales et al., 1996, 1997; Reich-

erter, 2001). Major earthquakes occurred during his-

torical times (see catalogue in Reicherter, 2001). The

last prominent earthquake took place in 1884 and

caused more than 800 casualties close to Arenas del

Rey in the southwestern Granada Basin, (Fig. 2;

Reicherter et al., 2003). Moderate and small earth-

quakes (bM 5) take place predominantly along major

fault zones, and are concentrated along the margins of

the Granada Basin (Morales et al., 1990; Sanz de

Galdeano et al., 1995).

2. Tectonic setting

The present-day stress orientation in the studied

area has been determined from focal mechanism solu-

tions, which point to a NW–SE directed maximum

horizontal stress SHmax, whereas r1 is mainly vertical

and r3 is horizontal in NE–SW direction (Galindo-

Zaldıvar et al., 1993, 1999; Herraiz et al., 2000).

However, the stress field in the area is heterogeneous

as documented by permutations of the stress axes.

Radial extension, NE–SW extension, and NW–SE

subhorizontal compression are observed; the latter

being parallel to the regional tectonic stress field

(Buforn et al., 1988; De Mets et al., 1990; Galindo-

Zaldıvar et al., 1993).

Since Tortonian time the Betic Cordilleras were

deformed under extensional stress regimes with va-

rying extension directions in an overall compressive

plate boundary setting (Galindo-Zaldıvar et al., 1999).

The Alboran Basin and the Neogene intramontane

basins formed along major faults (Fernandez et al.,

1996a; Comas et al., 1999). Strike-slip faults (Sanz de

Galdeano, 1990; Lonergan and White, 1997) and

major normal faults (e.g., Jabaloy et al., 1992; Sanz

de Galdeano and Lopez-Garrido, 1999) were involved

in the basin formation processes.

Thrusting directions in the External Zone vary

significantly. Generally, the thrusting directions in

the Betic Cordilleras are N and NW directed (Gar-

cıa-Hernandez et al., 1980; Guezou et al., 1991;

Banks and Warburton, 1991; Allerton et al., 1993).

In the Prebetic fold-and-thrust belt, however, fault-slip

vectors point to mainly NW directed thrusting, but

deviation varies from N to WNW trends (Platt et al.,

2003). Locally, prominent S and SE directed back

thrusts are developed (Allerton et al., 1993, 1994;

Vissers et al., 1995), which also affected the IEZB

(Lonergan et al., 1994). In the central and western

parts of the Betic Cordilleras, thrusting directions are

towards N and NW (Reicherter and Michel, 1993),

but frequently W orientations have been detected

(Platzman, 1992; Kirker and Platt, 1998; Ruano et

al., 2000).

Thrusting in the External Betics was accompanied

by differential vertical-axis rotations of individual

thrust sheets, as revealed by numerous paleomagnetic

studies (e.g. Osete et al., 1989; Platzman, 1992; Aller-

ton et al., 1993, 1994; Lonergan et al., 1994; Kirker

and Platt, 1998; summary in Platt et al., 2003). Fold

trends are often clearly related to thrusts, however

Triassic decollement horizons are not always parallel

to the major fold axes due to complex ramp-flat

geometries (De Ruig, 1990). Folding of the thrust

planes occurs during late stage folding. The major

fold orientation in the central Betic Cordilleras is

ENE–WSW, but swings locally towards N–S (Blu-

menthal, 1927; De Ruig, 1990; Reicherter and Michel,

1993). NW-vergent and overturned folds, as well as

box folds, can be observed. The Prebetic is partly

Page 7: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 197

characterized by NW trending folds (Platt et al.,

2003). In the Campo de Gibraltar fold axis orientation

varies between 358 and 858 (Kirker and Platt, 1998).

Internally, thrust nappes are often intensely deformed

with chevron-type folding.

2.1. Satellite lineations

Satellite lineations obtained from Landsat data in-

terpretation of the Iberian Peninsula show a grouping

in four major sets directed NW–SE, N–S, NE–SW and

ENE–WSW (531 lineations were evaluated). A pro-

nounced minimum is observed in E–W direction.

Taking into account the length of the lineations, the

four major directions are apparent. However, when the

number/length ratio is calculated the N–S and E–W

directions form maxima. The E–W direction is repre-

sented by only 2 lineations and, hence, is not regarded

as relevant. Directions of topographic lineations from

a digital elevation model and Landsat data of Anda-

lusia (Fig. 1) differ significantly from the overall

Iberian trend. A total of 295 lineations can be grouped

into three maxima: NW–SE, NE–SW, and a broad

range between 70 and 90 azimuth degrees. The latter

range parallels the major fold axis trend and is known

as the bBetic trendQ described already by Blumenthal

(1927) and Staub (1927). The lineation lengths and

number / length ratios are in accord with the observed

three major directions.

2.2. Fault pattern

The Central and Western Betic Cordilleras are

presently characterized by normal faulting, only

minor strike-slip faults are observed. The fault pattern

can be grouped into four main directions: NW–SE,

NE–SW, NNE–SSW and E–W (Sanz de Galdeano,

1983, 1990). The study area comprises faulted rocks

of all tectono-stratigraphic units (Fig. 2), from Paleo-

zoic schists of the Nevado-Filabride and Alpujarride

Complexes to Quaternary sediments of the Granada

Basin. The maximum displacement/uplift rate ob-

served was around 3500 m along the western low-

angle detachment of the Sierra Nevada (Sanz de Gal-

deano and Lopez-Garrido, 2000). The Ventas de

Zafarraya Fault in the western part of the Granada

depression (Fig. 2), the possible source of the 1884

earthquake, exhibits more than 1500 m of vertical

throw that has happened since the Tortonian

(Galindo-Zaldıvar et al., 2003; Reicherter et al.,

2003). Most of the related faults were reactivated

under changing stress fields and acted partly as

boundaries of the basins during the Neogene. The

Granada Basin is delimited by faults trending E–W

along the southern boundary and NW–SE at its west-

ern and eastern margins (Fernandez et al., 1996a). The

northern boundary is characterized by a low-angle

detachment (Ruano et al., 2000; Ruano, 2003).

Hence, the prolongation and evidence of the Crevil-

lente Fault or bCadiz-Alicante-FaultQ as a dextral

strike-slip fault in the study area (Figs. 1 and 2),

along the northern margin of the Granada Basin, is

doubtful (De Smet, 1984; Leblanc and Olivier, 1984;

Leblanc, 1990; Sanz de Galdeano, 1990; Sanz de

Galdeano and Lopez-Garrido, 2000; Ruano, 2003).

2.3. Previous studies

Several studies concerning the neotectonic evolu-

tion were carried out in the Central Betic mountain

range, the Subbetic Zone and the adjacent Neogene

basins (e.g., Sanz de Galdeano, 1990; Galindo-Zaldı-

var et al., 1993; Reicherter and Michel, 1993; Ruano

et al., 2000; Ruano, 2003). The results of this study

supplement previous studies by Galindo-Zaldıvar et

al. (1993, 1999), Reicherter and Michel (1993), Aller-

ton et al. (1994), Sanz de Galdeano (1997) and Munoz

et al. (2002), and will be so compared to those. The

upshot is to provide a new deformation path for the

last 15 Ma.

3. Paleostress analysis

Stress evolution was deduced from paleostress

analysis of populations of striated fault planes.

Paleostress analysis was carried out using a repre-

sentative number of striated fault planes (325 in

total) from 36 sites (with different lithology, stratig-

raphy and of different paleogeographical units) along

the traverse across the central Betic Cordilleras (Fig.

2, Table 1). The brittle mesoscale features include

striated fault planes, joints and tension gashes, dip of

bedding, as well as fold-axis directions. A set of 20

fault planes with striations was usually collected at

each site, sometimes the number of faults used is

Page 8: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Table 1

Paleostress data of the Central Betic Cordilleras: AD = Alboran Domain (Alp = Alpujarride Complex, Dors. = Dorsalian Complex); EZ = External Zone (eSub = External Subbetic, mSub = Median Subbetic,

Peni = Penibetic), GB = Granada Basin

General information Calculated results Timing of deformation

Station Domain Age Lithology Locality N r1 r2 r3 R Quality Phase Fault characterization

1 AD (Alp) Triassic marbles Frigiliana 11 161/14 330/74 071/02 0.4 B 3b sin, N–S; dex, NW–SE

1 AD (Alp) Triassic marbles Frigiliana 3 213/13 084/68 307/15 0.47 B 1 sin, NE–SW

2 AD (Alp) Paleozoic schists Sierra Nevada 5 024/63 159/19 255/17 0.6 B 2 nor, NW–SE, NE+SW dip

3 AD (Alp) Paleozoic marbles Sierra Nevada-2 13 008/02 222/86 098/01 0.41 B 3b

4 AD (Alp) Paleozoic schists Arroyo de la Miel 6 119/75 297/14 027/00 0.75 B 3a nor, NW–SE, SW dip

5 AD (Alp) Paleozoic micaschists La Herradura 6 154/82 337/07 247/00 0.62 B 3a nor, NW–SE, SW dip

5 AD (Alp) Paleozoic micaschists La Herradura 4 094/72 227/11 320/12 0.54 B 2 nor, NE–SW+E–W

5 AD (Alp) Paleozoic micaschists La Herradura 6 157/71 008/16 275/09 0.55 B 1 or 2? nor, N–S

6 AD (Alp) Triassic shales Vinuela 6 262/03 353/05 142/85 0.61 B 1 rev, N–S, E dip

7 AD (Alp) Paleozoic shales Vinuela 8 320/52 079/21 182/30 0.5 C 2 nor, NW–SE, NE dip

8 AD (Alp) Paleozoic shales Vinuela 6 333/33 137/55 238/07 0.45 A 3b dex, NW–SE; sin, N–S

8 AD (Alp) Paleozoic shales Vinuela 10 355/79 130/03 222/08 0.48 B 3a nor, NW–SE, NE+SW dip

8 AD (Alp) Paleozoic shales Vinuela 4 111/74 002/05 271/14 0.51 C 2 nor, N–S, E dip

9 AD (Alp) Paleozoic shales Canillas 4 209/78 300/00 030/12 0.5 B 3a nor, NW–SE, NE+SW dip

10 AD (Dors.) Jurassic limestones Alhama 18 027/53 232/33 134/12 0.49 B 2 dex, N–S; sin, NE–SW

10 AD (Dors.) Jurassic limestones Alhama 8 003/07 265/47 100/41 0.48 C 1 rev+dex, NW–SE, SW dip

11 EZ (mSub) E. Jurassic limestones Los Olivares 7 140/09 257/69 046/17 0.54 A 3b dex, E–W+NW–SE

12 EZ (eSub) E. Cretaceous marly limestones Barbahijar 16 321/10 188/74 053/11 0.69 B 3b dex, E–W; rev, NE–SW

12 EZ (eSub) E. Cretaceous marly limestones Barbahijar 6 092/56 270/33 000/01 0.9 C 2 nor, E–W, N+S dip

12 EZ (eSub) E. Cretaceous marly limestones Barbahijar 11 172/83 055/02 325/05 0.85 B 1 dex, NW–SE+N–S

13 EZ (mSub) L. Cretaceous marly limestones Rio Fardes 11 117/44 312/44 215/07 0.63 A 3a + 3b dex, E–W

13 EZ (mSub) L. Cretaceous marly limestones Rio Fardes 7 233/79 081/09 350/04 0.77 B 2 nor, E–W

13 EZ (mSub) L. Cretaceous marly limestones Rio Fardes 14 189/04 006/85 099/00 0.6 A 1 dex, NE–SW, NE+SW dip

14 EZ (mSub) E. Jurassic limestones Morron 20 100/87 340/01 250/01 0.52 A 3a nor, N–S, E+W dip

14 EZ (mSub) E. Jurassic limestones Morron 8 159/75 060/02 329/14 0.51 B 2 nor, NE–SW, NW dip

14 EZ (mSub) E. Jurassic limestones Morron 4 036/76 287/04 196/12 0.6 C 1? nor, NW–SE, NE+SW dip

15 EZ (mSub) E. Jurassic limestones Sierra Elvira 5 336/10 212/71 069/15 0.5 B 3b dex, NW–SE+E–W; sin N–S

15 EZ (mSub) E. Jurassic limestones Sierra Elvira 16 310/14 173/70 044/12 0.55 A 3b dex, E–W, S dip; sin, N–S

15 EZ (mSub) E. Jurassic limestones Sierra Elvira 12 272/71 107/16 015/04 0.51 A 3a nor, NW–SE, SW dip

15 EZ (mSub) E. Jurassic limestones Sierra Elvira 4 205/82 353/06 084/03 0.49 B 2 nor, N–S, W dip

15 EZ (mSub) E. Jurassic limestones Sierra Elvira 5 204/01 105/82 296/12 0.49 B 1 dex, N–S

16 EZ (mSub) E. Jurassic limestones Venta Nava 12 004/24 167/64 271/06 0.5 A 3b sin, NE–SW, SE dip

16 EZ (mSub) E.–M. Jurassic limestones Venta Nava 11 136/12 321/77 226/01 0.5 A 3b dex, E–W; sin, N–S

16 EZ (mSub) E.–M. Jurassic limestones Venta Nava 5 071/22 249/68 341/01 0.52 A 1 sin, NW–SE; dex, NE–SW

17 EZ (eSub) E. Jurassic limestones Cast. Locubın 9 176/39 028/45 280/16 0.5 B 3b dex, N–S; sin, NE–SW

17 EZ (eSub) E. Jurassic limestones Cast. Locubın 12 118/74 229/05 320/14 0.39 B 2 nor, NE–SW, SE dip

17 EZ (eSub) E. Jurassic limestones Cast. Locubın 9 225/23 104/51 329/30 0.46 B 1 rev, NE–SW, SE dip

18 EZ (mSub) E. Jurassic limestones Colomera 6 325/04 082/79 234/09 0.19 B 3b dex, NW–SE; sin, N–S

18 EZ (mSub) E. Jurassic limestones Colomera 4 014/08 152/78 283/07 0.45 C 3b sin, NE–SW; dex, NW–SE

18 EZ (mSub) E. Jurassic limestones Colomera 11 007/83 145/05 236/04 0.13 B 3a dip

18 EZ (mSub) E. Jurassic limestones Colomera 13 176/76 293/06 024/11 0.71 B 2 SW dip

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

198

Page 9: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

18 EZ (mSub) E. Jurassic limestones Colomera 10 088/79 199/03 290/09 0.6 B 2 nor, N–S+NE–SW, W+E dip

18 EZ (mSub) E. Jurassic limestones Colomera 10 227/21 031/67 135/15 0.41 B 1 dex, N–S; rev, NE–SW, NW dip

19 EZ (eSub) L. Jurassic limestones Casablanca 5 141/18 237/15 004/65 0.4 C 3b rev, NE–SW, SE dip

19 EZ (eSub) L. Jurassic limestones Casablanca 11 071/66 300/15 205/16 0.36 B 3a nor, NW–SE, NE+SW dip

19 EZ (eSub) L. Jurassic limestones Casablanca 10 218/04 308/02 072/84 0.39 B 1 NE–SW

20 EZ (Peni) E. Jurassic limestones Zafarraya 4 138/25 333/63 231/05 0.51 A 3b dex, NW–SE, NE dip

20 EZ (Peni) E. Jurassic limestones Zafarraya 6 179/76 285/04 016/13 0.2 A 3a nor, NW–SE, NE dip, N–S, W dip

20 EZ (Peni) E. Jurassic limestones Zafarraya 2 190/16 073/56 289/28 0.5 B 1 rev, NE–SW, SE dip

21 EZ (Peni) E. Jurassic shales Zafarraya 8 249/77 088/12 357/04 0.48 B 2 nor, E–W, N dip

22 EZ (Peni) E. Jurassic dolomites Hacho de Loja 4 346/52 204/31 102/18 0.47 A 3a nor, NW–SE, NE dip, E–W, N dip

22 EZ (Peni) E. Jurassic limestones Hacho de Loja 12 089/68 313/16 218/15 0.48 A 2 nor, N–S+E–W

22 EZ (Peni) E. Jurassic dolomites Hacho de Loja 4 107/06 014/17 214/72 0.48 A 1 rev, N–S, E dip

23 EZ (mSub) M. Jurassic limestones Algarinejo 8 288/15 082/74 196/07 0.48 A 3b dex, E–W

23 EZ (mSub) E. Jurassic limestones Algarinejo 9 208/80 321/04 052/09 0.53 A 3a nor, NW–SE, SW dip

23 EZ (mSub) M. Jurassic limestones Algarinejo 4 083/70 272/20 181/83 0.5 A 2 nor, E–W, S dip

23 EZ (mSub) E. Jurassic marly limestones Algarinejo 4 336/54 067/01 157/36 0.5 B 2 nor, NE–SW, SE dip

23 EZ (mSub) E. Jurassic limestones Algarinejo 6 020/31 269/31 143/43 0.52 B 1 rev, E–W, N dip

24 EZ (Peni) Triassic shales Loja 4 135/04 278/84 042/04 0.57 C 3b dex, E–W

24 EZ (Peni) Triassic shales Loja 6 312/71 078/11 173/14 0.54 C 2 nor, E–W, N dip

24 EZ (Peni) Triassic shales Loja 4 298/25 031/05 131/64 0.5 B 1 rev, NE–SW, NW dip

25 EZ (Peni) E. Cretaceous marly limestones Loja 6 053/80 264/09 173/05 0.49 B 2 nor, E–W, N dip

26 EZ (eSub) E. Jurassic dolomites Priego de Cordoba 4 118/08 227/68 025/21 0.5 B 3b dex, E–W

26 EZ (eSub) M. Jurassic marly limestones Priego de Cordoba 4 191/69 056/15 322/14 0.5 B 2 nor, NE–SW, NW dip

26 EZ (eSub) L. Cretaceous marly limestones Priego de Cordoba 4 091/68 300/20 207/10 0.49 A 2 nor, E–W, N dip

26 EZ (eSub) E. Jurassic dolomites Priego de Cordoba 4 308/31 046/16 160/55 0.49 A 1 rev, NE–SW, NW dip

27 EZ (Peni) E. Jurassic limestones Torcal 6 131/70 320/19 229/02 0.49 A 3a nor, NW–SE, NE dip

27 EZ (Peni) E. Jurassic limestones Torcal 5 275/02 009/53 182/36 0.5 A 1 sin, NW–SE, NE dip

28 GB Miocene conglomerates Sierra Nevada 3 282/53 161/21 058/28 0.54 A 2 nor, NW–SE, SW dip

29 GB Mio–Plioc. marls Fuensanta 15 085/85 303/03 213/02 0.37 B 3a nor, N–S; dex, E–W

30 GB Plio–Pleistocene clastics Cubillas 8 153/00 063/81 243/21 0.43 A 3b dex, E–W; sin, N–S

31 GB L. Miocene sands Otura 7 244/09 337/18 127/68 0.38 B 3b rev, NW–SE, NE dip

31 GB L. Miocene sands Otura 5 326/21 062/14 163/63 0.39 B 3b dex, E–W; sin, N–S

31 GB L. Miocene sands Otura 14 179/06 088/02 357/83 0.4 A 2 rev, NE–SW, NW+SE dip

31 GB L. Miocene sands Otura 13 029/72 258/12 165/12 0.35 A 2 nor, NE–SW+E–W

32 GB L. Miocene marlstones Arenas del Rey 3 277/83 099/07 009/00 0.46 A 3a nor, E–W, N dip

32 GB L. Miocene marlstones Arenas del Rey 6 199/78 296/01 026/12 0.48 B 3a nor, NW–SE, N+SW dip

32 GB L. Miocene marlstones Arenas del Rey 7 236/80 003/06 093/08 0.49 A/B 2 nor, N–S, E+W dip

32 GB L. Miocene marlstones Arenas del Rey 3 326/82 215/03 125/07 0.43 B 2 nor, NE–SW, NW+SE dip

33 GB L. Miocene sandstones Arenas del Rey 2 321/40 131/50 227/05 0.31 A 3b dex, NW–SE+E–W

33 GB L. Miocene sandstones Arenas del Rey 6 270/72 105/18 013/04 0.57 A 3a SE

34 GB M. Miocene calcarenites Alhama 5 030/83 120/01 210/06 0.5 A 3a nor, NW–SE, NE dip

35 GB Pliocene sands, congl. Salar 6 183/84 296/02 027/05 0.5 A 3a nor, NW–SE, SW dip

36 GB Pliocene sands Rıo Frıo 6 105/12 281/78 015/01 0.5 A 3b dex, E–W

N =number of separated faults of the individual phase; r1=main compression direction (trend and plunge), r2= intermediate direction,r3=main extension direction, axial ratio of the stress ellipsoid

R =[(r2� r3) / (r1�r3)] with r1Nr2Nr3; quality of the fault-striae data set A = very good, B = good, C = regular; deformation phase, explanation see text; fault characterization: sin = sinistral, dex =

dextral, nor = normal, rev = reverse.

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

199

Page 10: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212200

even less due to bad preservation of striae. If possi-

ble, the sense of slip (normal, reverse, sinistral or

dextral) and displacement were determined, which is

on the order of centimeters or meters (Table 1). The

sense and direction of motion along the faults were

inferred from various kinematic indicators such as

Riedel shears, lunate and tensile cracks, striation and

slickensides (calcite fibres or slickolithes; after Petit,

1987, Doblas, 1998). The quality of the kinematic

indicators was classified into very good (A), good

(B) and C (regular), and weighted during subsequent

data analysis (extend, displacement, fault plane area,

Table 1). To distinguish between the different phases

of faulting, also considered were: (1) the age of the

deformed rocks; (2) characterization of synsedimen-

tary faults; (3) use of tilted structures and including

back-rotation of strata and associated faults; (4)

cross-cutting relationships; and (5) reactivation with

the formation of superimposed striae on the fault

plane. Most of the deformation structures were

dated relatively, i.e., using overprinting criteria, and

separated in the field into different phases of fault-

ing. Correspondingly, the fault sets described in the

following section represent already separated and

homogeneous subsets.

Where separation was difficult the data were di-

vided with graphical methods into populations with

matching orientations of the theoretical P and T axes

(pressure and tension axes). A theoretical shorting and

extension direction were constructed for each indivi-

dual fault plane based on orientation, sense of slip of

the fault assuming a frictional angle of 308, whichprovided the best results. The principal axes of the

paleostress tensors (r1Nr2Nr3) were obtained by

computer-aided numerical dynamic analysis (NDA)

based on Spang (1972) and Sperner and Ratschbacher

(1994). The applied NDA method of the TVB-pro-

gram (Ornter, 2002) calculates strain parameters rather

than stress parameters. Assuming that deformation

follows the applied stress these values equal the stress

parameters). In addition, the axial ratio of the stress

ellipsoid [R =(r2�r3) / (r1�r3)], with r1Nr2Nr3,

was calculated. The orientations of the principal stress

axes, the distribution of the P and T axes, as well as

the ratio R have all been used to infer the individual

tectonic regimes. At most outcrops, the ratio R is

approximately 0.5, indicating pure normal, reverse

or strike-slip faulting (Table 1), characterized by

small circle distribution of T-axes. Thus the deforma-

tion is mostly two-dimensional, which can also be

observed from the analysis of earthquake focal

mechanisms (see Galindo-Zaldıvar et al., 1999).

Strike-slip faulting, with an oblique transpressive

component, shows lower ratios between 0.2 to 0.5,

and is characterized by a great circle distribution of T-

axes. The paleostress orientations of each separated

fault-slip population set are presented graphically as

cumulative pseudo fault plane solutions, representing

the mean tensors of individual fault-striae data sets. A

series of maps delineates the paleostress evolution of

the central Betic Cordilleras.

The determinations of the paleostress axes have

been supplemented by the analysis of tension gashes

and joints (hybrid, extensional or shear fractures,

partly open, partly not filled or calcite-filled; Han-

cock, 1985). Asymmetric rose plots incorporate the

dip direction of fracture planes, which may allow one

to delineate small horizontal axis rotations of the

joints (Fig. 4) and bedding tilt correction. Four

major sets are directed NW–SE, N–S, NE–SW and

ENE–WSW. As already observed in satellite lineation

analysis, the NW–SE direction forms the maximum,

and may be an inherited feature of the underlying and

overridden Variscan basement of Iberia. The ENE–

WSW striking sets have also been found in the Mio-

cene sediments in the Granada Basin and the Alcala la

Real sector. These sediments post-date the major

bBeticQ trend of thrusting and folding.

3.1. Results of the paleostress analysis

The stratigraphy and the Neogene basin evolution

of the Granada Depression, which forms most of the

central part of the investigated traverse, are known

(Braga et al., 1990; Morales et al., 1990; Fernandez et

al., 1996a,b, Fig. 3). The paleostress-stratigraphy con-

cept of Kleinspehn et al. (1989) was applied to estab-

lish a chronology of the tectonic events during basin

evolution. The Neogene–Quaternary fill of the Gra-

nada Basin displays a complex tectonic evolution, as

well as a complex sedimentary and subsidence history

since its initiation in the Early Miocene. Reactivation

of Neogene faults is frequently observed in the study

area, especially in rocks of lake deposits. Sediment

ages helped to date activities along faults (e.g., syn-

sedimentary horst and graben development associated

Page 11: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 4. Map of the traverse through the Central Betic Cordilleras showing the characteristic joint patterns (N number of extensional joints),

illustrated by asymmetric rose diagrams incorporating also the fault dip direction. From numbered stations also paleostress data are available

(see Fig. 2).

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 201

with NW–SE striking normal faults) that have been

reactivated as dextral strike-slip faults, occasionally

with an oblique component (Fig. 5A). Horst and

graben systems developed (Fig. 5B, C). Synsedimen-

tary normal growth faults are frequent in lacustrine

marlstones of Late Messinian to Earliest Pliocene age

(Fig. 5D). Most of the faults in the Granada Depres-

sion within Tortonian–Pliocene sediments indicate

NW–SE to W–E extension during formation and

SW–NE extension during later reactivation.

3.1.1. Oligocene–Tortonian deformation (thrusting,

phase D1)

The oldest stress tensor group is characterized by a

maximum subhorizontal compression (r1) and direct-

ed mainly N–S and minor NW–SE, but is also partly

Page 12: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 5. Field sketches of Tortonian–Pliocene deformation features in the Granada Basin, outcrops in Fig. 2 (dip = direction and dip; ST =

striation; R = Riedel shear; j = joint). All faults indicate (N)W–S(W) extension. A: Normal faults and joints in Tortonian calcarenites (station 20,

note vertical displacement). Left normal fault has been reactivated as dextral strike-slip fault. B: Synsedimentary normal growth faults in

lacustrine marlstones with lignitic layers of Late Messinian to earliest Pliocene age (hammer for scale; station 32). C: Normal faults in Messinian

to Early Pliocene marlstones (station 32). D: Synsedimentary normal growth fault in lacustrine marlstones with lignitic layers of Late Messinian

to earliest Pliocene age (station 32).

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212202

E–W directed. The minimum compression direction

(r3) is subvertical (Fig. 6). The evolution of fault

planes is related to thrusting and folding in the main

Alpine orogenic phase. Compressional structures

(thrusts, high-angle reverse and strike-slip faults) de-

veloped during the Lower to Middle Miocene and

were found only in parts of the Internal or External

Zone. Locally, structures (oblique normal faults, thrust

faults and folds) deviate from the major NW–SE

directed deformation trend and suggest both transfer

faulting and complex ramp-flat geometries in the area.

Some of the outcrops in the External zone, which

include the Rio Fardes region (Fig. 2, station 13),

show dominant NE–SW sinistral strike-slip and

minor thrust faults, which form a compressional flow-

er structure in this area (Fig. 6B, station 13). The

Casablanca section in the External Zone is characte-

rized by shallow dipping thrust faults with a top to NE

sense of shear (Fig. 6B, station 19). Thrust faults are

associated with minor back-thrusts and sinistral strike-

slip faults. The corridor between Loja and Vinuela is

dominated by W directed thrusting (Fig. 6). Stations

10 and 20 are situated close to the IEBZ (Internal-

External Boundary Zone). Both show a compressive

sinistral strike-slip setting during Phase D1. Due to

the age, Phase D1 is not observed in the Neogene

sediments of the Granada Basin.

3.1.2. Tortonian–Early Pliocene deformation (radial

extension, phase D2)

The compressive phase is followed by a major

extensional phase D2 with varying extension direc-

Page 13: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 6. Lower-Middle Miocene deformation phase related to thrusting and the nappe emplacement in the External Betic Cordilleras. Observed is a horizontal maximum compression

direction varying from NE–NW. The cumulative pseudo fault-plane solutions (beachball diagrams) sum up several separated fault-slip data of outcrops (see also Table 1; white =

compression; black = extension). Inset: upper stereoplot shows all cumulative P-subhorizontal, B-, and T-axes of the stations ( P-axes = r1-direction = black dots; B-axes =

r2-direction = open squares; T-axes = r3-direction = filled triangles). Middle and lower stereoplots show individual fault-striae data; station 13 is dominated by NE trending strike-slip

faults, whereas station 19 shows NE and SW shallow-dipping thrust planes.

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

203

Page 14: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 7. Middle Miocene–Pliocene deformation phase related to the collapse of the thrust wedge and the initiation of the Neogene basin formation in the External Betic Cordilleras.

Observed is a vertical maximum compression direction leading to NW–SE extension (see Fig. 6 for legend). Inset: upper stereoplot shows all cumulative P-, B-, and T-axes of the

stations. Middle and lower stereoplots show individual fault-striae data; both stations are dominated by NE trending conjugate sets of normal faults.

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

204

Page 15: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 8. Pliocene–Recent deformation phase related to the collapse of the thrust wedge and the ongoing of the Neogene basin formation in the External Betic Cordilleras. Observed is a

vertical maximum compression direction leading to major NE–SW extension (see Fig. 6 for legend). Inset: upper stereoplot shows all cumulative P-, B-, and T-axes of the stations.

Middle and lower stereoplots show individual fault-striae data; both stations are dominated by NW trending conjugate sets of normal faults.

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

205

Page 16: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 9. Pliocene–Recent deformation phase related to the convergence of the Eurasian and African Plate. Observed is a horizontal maximum compression direction leading to SE–NW

compression (see Fig. 6 for legend) and coeval NE–SW extension. Inset: upper stereoplot shows all cumulative P-, B-, and T-axes of the stations. Middle and lower stereoplots show

individual fault-striae data; both stations are dominated by N–S and E–W trending conjugate sets of strike-slip faults.

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

206

Page 17: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 207

tions, but dominantly NW–SE directed, also N–S and

E–W directed (Fig. 7). High-angle normal faults in

Paleozoic to Neogene rocks, striking NW–SE and

NE–SW, are abundant in the entire study area. The

different sets of high-angle normal faults indicate E–

W, NW–SE or N–S orientated extension and vertical

maximum compression (phase D2, r1 subvertical, r3

subhorizontal NW–SE, radial, Fig. 7B). Stations 10

and 12, both in series of the External Zone, exhibit

sets of conjugate normal faults, and minor ENE–

WSW trending, steeply dipping, transtensional, sinis-

tral strike-slip faults (Fig. 7B).

NW extension affected the oldest formations of the

Neogene Granada Basin (stations 31, 32, 33), asso-

ciated with coeval deviations of the local stress field

(e.g., station 31, with N–S directed compression, also

station 23 of the External Zone with two individual

subsets). Faults of phase D2 are mostly synsedimen-

tary high-angle normal faults in Tortonian to earliest

Pliocene sediments (Fig. 5), which enabled the dating

of the onset of the NW–SE directed extension in the

Late Miocene. The high normal faults displace the

IEBZ significantly in the Zafarraya area (stations 20,

21) and, hence, is a younger tectonic feature (Reich-

erter et al., 2003).

3.1.3. Pliocene-Recent deformation (SW directed

extension and SE compression, phase D3)

The youngest stress tensor group D3 has been

divided into two dominant sub-sets after P–T-sepa-

ration: (3a) with r1 subvertical, r3 subhorizontal

NE–SW directed, associated with normal faulting

(Fig. 8); and (3b) with r1 subhorizontal NW directed

and with r3 subhorizontal NE–SW directed, associ-

ated with strike-slip faulting (Fig. 9). The extension

direction is constant during phase 3 (Figs. 8 and

9B), and parallels the orogen. The P–T axis distri-

bution is characterized by sharp maxima (Fig. 8B) in

sub-set 3a, the extension dominated faulting. The

sub-set of phase 3b shows a NW–SE trending gir-

dle-like distribution of P- and B-axes. The P-axes

plot close to the circle border (shallow dips, r1-

direction is subhorizontal), whereas the B-axes (r2-

direction) are subvertical and scatter around the cen-

ter of the circle.

Older NW–SE striking high-angle normal faults

were reactivated during the NW–SE directed subho-

rizontal maximum compression in the Betic Cordil-

leras (Figs. 8 and 9). NW–SE striking faults are

mostly normal faults or dextral strike-slip faults, oc-

casionally oblique with a dextral sense of movement,

which form conjugate sets (Fig. 8B, stations 8 and

19). New faults developed in Pliocene to Quaternary

sediments, mainly striking E–W with a dextral sense

of shear and N–S with a sinistral sense of shear (Fig.

9B, stations 15 and 30). The high normal Ventas de

Zafarraya Fault (stations 20, 21) is reactivated dex-

trally during this younger phase (Reicherter, 2001;

Munoz et al., 2002).

4. Discussion and conclusions

Paleostress and structural analysis on small-scale

faults in the central part of the Betic Cordilleras,

from the Mediterranean coast east of Malaga to the

Guadalquivir Basin, allows one to establish the tim-

ing and orientation of stress regimes since the Early

Miocene. Brittle fault populations, folds, joints, and

satellite lineations in different lithologies and strati-

graphic units were studied and analyzed. Four tensor

groups were distinguished and correlate with three

distinct phases in the Late Oligocene–Aquitanian to

Recent tectonic evolution of the Betic Cordilleras

(Fig. 10):

Phase D1: Thrusting-related N–S to E–W com-

pressional structures in the Internal and External

Zones of the Betic Cordilleras correlated to a top-

to-N to NW and, partly, top-to-W directed thrusting

of the nappes, including large scale folding. Strike-

slip faults are also observed. The data are consider-

ably variable, but are consistent with NW thrusting

and NE trending fold axes. This tectonic phase could

represent continent–continent collision of the Iberian

(Eurasian) plate and the Alboran Domain, which is

considered to constitute a microplate during the Me-

sozoic and early Cenozoic (e.g., Sanz de Galdeano,

1990; Reicherter et al., 1994; Platt et al., 2003). In

addition, Galindo-Zaldıvar et al. (1993) reported E–

W compression for the Burdigalian with a switch to

approximately N–S compression during the Langhian

to Serravalian.

Phase D2: NW dominated extension is characte-

rized by large-scale normal faulting. Both orogen

parallel extension and orogen perpendicular exten-

sion are observed, and could represent the collapse

Page 18: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

Fig. 10. The evolution of the local stresses (main compression and extension directions) and related tectonic phases in the Central Betic Cordilleras since the Aquitanian to Recent (EZ

= External Zone; AD = Alboran Domain; NB = Neogene basins; GB = Granada Basin). Main fault directions and kinematics of the central Betic Cordilleras, based on fault-slip data,

joints and satellite lineations. The general Betic ENE-trend (in gray) is not frequently observed in the younger sedimentary successions. Reactivation of pre-existing faults is

frequently common. Barbs indicate dip direction of faults. Note that Phases 3a and 3b are coeval.

K.R.Reich

erter,G.Peters

/Tecto

nophysics

405(2005)191–212

208

Page 19: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 209

of the thickened thrust wedge. Faulting due to this

phase is observed along the entire transect and

appears not to be localized at faults along the

basin margins. Timing of the extensional event is

well constrained by synsedimentary deformed Tor-

tonian and Messinian sediments in the Granada

Basin (Fig. 5). Phase 2 is probably related to the

rapid exhumation of the intensely metamorphosed

rock series of the Internal Zones, i.e. the Nevado-

Filabride and Alpujarride-Complexes, and hence re-

lated to the ongoing gravitational collapse of the

Betic Cordilleras and the formation of the Alboran

Basin and intramontane basins (Azanon and Crespo-

Blanc, 2000). The time calibration of the regional

stress field D2 was further constrained considering

data of adjacent areas from Galindo-Zaldıvar et al.

(1993, 1999), Reicherter and Michel (1993), Reich-

erter (1999), Ruano et al. (2000) and Ruano (2003).

NW–SE directed extension probably occurred from

the Tortonian on and was contemporaneous with

intramontane basin formation (e.g., Sanz de Gal-

deano, 1985, 1990). NW–SE extension has also

been reported from the Eastern Betic Cordilleras

during the Tortonian (Stapel et al., 1996; Huibregtse

et al., 1998; Jonk and Biermann, 2002). Fault-slip

data from the Rif and the Atlas of Morocco show

also NW directed extension during the Tortonian

(Ait Brahim et al., 2002).

Neogene basin formation in the central Betic

Cordilleras is not related to large strike-slip move-

ments along the Cadiz–Alicante Fault that, on some

maps, is shown to extend to Cadiz (e.g., De Smet,

1984; Leblanc and Olivier, 1984; Sanz de Galdeano,

1990). In the Rio Fardes area, along the northern

margin of the Granada Basin and further west (sta-

tions 13, 14, 18, 22, 24, 27; Fig. 6), fault slip data

do not reveal any dextral strike-slip movements; on

the contrary, kinematic indicators suggest sinistral

strike-slip. The Cadiz–Alicante Fault seems to die

out west of the Tiscar Fault (Fig. 1) and cannot be

traced as a discrete structure through the central

Betic Cordilleras, as already evidenced by Platt et

al. (2003). They noted also the absence of any

significant geological change nor the structural

style. Hence, the displacement on the Cadiz–Alicante

Fault was not large and is confined to the Mesozoic

and Tertiary cover nappes in the eastern part of the

Betics (Allerton et al., 1993).

Phase D3 : Strike-slip faults and extensional struc-

tures indicate NW–SE compression and NE–SW ex-

tension since the latest Messinian to Recent due to the

ongoing convergence of the Eurasian and African

Plates and coeval uplift in the Betic Cordilleras. The

reactivation of pre-existing fractures and faults is

frequently observed. The permutation of the principal

stress axes, mainly r1 and r2, reflects periodic strike-

slip and normal faulting in the Betic Cordilleras. The

recent stress field remained stable during the Pliocene

to Quaternary (D3). It is characterized by homoge-

neous NE–SW directed extension in the central Betic

Cordilleras. Such is not the case in the Eastern Betics,

where intense strike-slip deformation is observed

(Huibregtse et al., 1998; Jonk and Biermann, 2002;

Faulkner et al., 2003). The fault pattern, including

reactivation of inherited structures, fits the observa-

tions from seismic focal mechanisms (Herraiz et al.,

2000), which largely point to a vertical maximum

compression direction. In contrast, NW–SE directed

subhorizontal compression is coevally observed (Fig.

10; Galindo-Zaldıvar et al., 1999).

The permutation of r1 and r2, as shown in the P–

B–T-axis distribution (Fig. 9) suggests an interaction

between the convergence of Eurasia and Africa and

the ongoing uplift in the mountain ranges. Conver-

gence rates over the European African Convergence

Zone in the western Mediterranean are on the order of

4–5.6 mm/a (Jolivet et al., 1999). However they are

distributed between the High Atlas in Morocco and

the Variscan Iberian Meseta. The deformation history

of the Rif Cordillera and the Atlas in Morocco is

characterized by an comparable stress history (Ait

Brahim et al., 2002). Furthermore, the permutation

of the principal stress axes from vertical to subho-

rizontal and vice versa points to almost comparable

magnitudes of r1 and r2. The isostatic equilibrium of

the thrust wedge in the central Betic Cordilleras has

yet not been reached.

The interaction of endogenic forces (plate conver-

gence, uplift) with changing exogenic factors (e.g.

climatic changes, precipitation, erosional and deposi-

tional rates) seems to play a major role in the reacti-

vation of pre-existing faults and the generation of new

faults. Also, the decoupling along the mid-crustal

shear zone (Galindo-Zaldıvar et al., 1999), with

major detachments and block tilting, influences the

basinward shift of depocentres.

Page 20: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212210

Acknowledgements

Financial support from the Deutsche Forschungs-

gemeinschaft is gratefully acknowledged (project Re

1361/3). U. Dyrssen is thanked for his help during

field work and data analysis. Thanks are extended to

J. Galindo-Zaldıvar, A. Jabaloy, P. Ruano and C. Sanz

de Galdeano (Univ. of Granada) for fruitful discus-

sions and joint field excursions. We are grateful to

Prof. I. Lerche (Univ. of Leipzig), who helped to

improve the English.

References

Ait Brahim, L., Chotin, P., Hinaj, S., Abdelouafi, A., El Adraoui,

A., Nakcha, C., Dhont, D., Charroud, M., Sossey Alaoui, F.,

Amrhar, M., Bouaza, A., Tabyaoui, H., Chaouni, A., 2002.

Paleostress evolution in the Moroccan African margin from

Triassic to Present. Tectonophysics 357, 187–205.

Allerton, S., Lonergan, L., Platt, J.P., Platzman, E.S., McClelland,

E., 1993. Palaeomagnetic rotations associated in the Eastern

Betic Cordillera, Southern Spain. Earth Planet. Sci. Lett. 119,

225–241.

Allerton, S., Reicherter, K., Platt, J.P., 1994. A structural and

palaeomagnetic study of a section through the eastern Subbetic,

Southern Spain. J. Geol. Soc. London 151, 659–668.

Azanon, J.M., Crespo-Blanc, A., 2000. Exhumation during a con-

tinental collision inferred from the tectonometamorphic evolu-

tion of the Alpujarride Complex in the central Betics (Alboran

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

Azanon, J.M., Galindo-Zaldıvar, J., Garcıa-Duenas, V., Jabaloy, A.,

2002. Alpine tectonics: II. Betic Cordillera and Balearic Islands.

In: Gibbons, W., Moreno, T. (Eds.), The Geology of Spain, vol.

16, pp. 401–416.

Bakker, H.E., De Jong, K., Helmers, H., Biermann, C., 1989. The

geodynamic evolution of the Internal Zone of the Betic Cordil-

leras (south-east Spain): a model based on structural analysis

and geothermobarometry. J. Metamorph. Geol. 7, 359–381.

Bandel, K., Reicherter, K., Dyrssen, U., Reiss, S., 2000. Late

Neogene lacustrine sedimentary facies and gastropod assem-

blages (Granada Basin, southern Spain). Mitteilungen Geol.-

Palaont, vol. 84. Institut Universitat Hamburg, pp. 111–130.

Banks, C.J., Warburton, J., 1991. Mid-crustal detachment in

the Betic system of Southeast Spain. Tectonophysics 191,

275–289.

Blumenthal, M., 1927. Zum Bauplan betischer und penibetischer

Decken imNorden der ProvinzMalaga. Geol. Rdsch. 17, 37–45.

Bourgois, J., 1978. La transversale de Ronda. Cordilleres Betiques,

Espagne. Donnees geologiques pour un modele dTevolution de

l’Arc de Gibraltar. Ann. Scient. Univ. Besancon 30 (445 pp.).

Braga, J.C., Martın, J.M., Alcala, B., 1990. Coral reefs in coarse-

terrigenous sedimentary environments (upper Tortonian, south-

ern Spain). Sediment. Geol. 66, 135–150.

Buforn, E., Udias, A., Colombas, M.A., 1988. Seismicity, source

mechanisms and tectonics of the Azores–Gibraltar plate boun-

dary. Tectonophysics 152, 89–118.

Calvert, A., Sandvol, E., Seber, D., Barazangi, M., Roecker, S.,

Mourabit, T., Vidal, F., Alguacil, G., Jabour, N., 2000. Geody-

namic evolution of the lithosphere and upper mantle beneath the

Alboran Region of the Western Mediterranean: constraints from

travel time constraints from travel time tomography. J. Geophys.

Res. 105, 10871–10898.

Comas, M.C., Platt, J.P., Soto, J.I., Watts, A.B., 1999. The origin

and tectonic history of the Alboran Basin: insights from leg 161

results. In: Zahn, R., Comas, M.C., Klaus, A. (Eds.), Proc. ODP,

Scientific Results, vol. 161, pp. 555–579.

De Jong, K., 1990. Alpine tectonics and rotation pole evolution of

Iberia. Tectonophysics 184, 279–296.

De Mets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current

plate motions. Geophys. J. Int. 101, 425–478.

De Ruig, M.J., 1990. Fold trends and stress deviation in the Alicante

fold belt, southeastern Spain. Tectonophysics 184, 393–403.

De Smet, M.E.M., 1984. Wrenching in the external zone of

the Betic Cordilleras, Southern Spain. Tectonophysics 107,

57–79.

Doblas, M., 1998. Slickenside kinematic indicators. Tectonophysics

295, 187–197.

Durand-Delga, M., Rossi, P., Olivier, P., Puglisi, D., 2000. Situation

structurale et nature ophiolitique de roches basiques jurassiques

asocies aux Flyschs maghrebiens du Rif (maroc) et de Sicile

(Italie). C. R. Acad. Sci. Paris 331, 29–38.

Faulkner, D.R., Lewis, A.C., Rutter, E.H., 2003. On the internal

structure and mechanics of large strike-slip fault zones: field

observations of the Carboneras Fault in southeastern Spain.

Tectonophysics 367, 235–251.

Fernandez, J., Soria, J., Viseras, C., 1996a. Stratigraphic architec-

ture of the Neogene basins in the central sector of the Betic

Cordillera (Spain): tectonic control and base-level changes. In:

Friend, P.F., Dabrio, C.J. (Eds.), Tertiary Basins of Spain. Cam-

bridge University Press, Cambridge, pp. 353–365.

Fernandez, J., Viseras, C., Soria, J., 1996b. Pliocene–Pleistocene

infilling of the Granada and Guadix Basins (Betic Cordillera,

Spain): the influence of allocyclic and autocyclic processes on

the resulting stratigraphic organization. In: Friend, P.F., Dabrio,

C.J. (Eds.), Tertiary Basins of Spain. Cambridge University

Press, Cambridge, pp. 366–371.

Fernandez-Fernandez, E., Campos, J., Gonzalez-Lodeiro, F.,

1992. Estructuras extensionales en los materiales Alpujarrides

al E de Malaga (Sierra Tejeda Cordilleras Beticas). Geogaceta

12, 13–16.

Galindo-Zaldıvar, J., Gonzalez-Lodeiro, F., Jabaloy, A., 1989. Prog-

ressive extensional shear structures in a detachment contact in

the western Sierra Nevada (Betic Cordilleras, Spain). Geodin.

Acta 3, 73–85.

Galindo-Zaldıvar, J., Gonzalez-Lodeiro, F., Jabaloy, A., 1993. Stress

and palaeostress in the Betic–Rif cordilleras (Miocene to the

present). Tectonophysics 227, 105–126.

Galindo-Zaldıvar, J., Jabaloy, A., Serrano, I., Morales, J., Gonzalez-

Lodeiro, F., Torcal, F., 1999. Recent and present-day stresses in

the Granada Basin (Betic Cordilleras): example of a late Mio-

Page 21: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212 211

cene–present-day extensional basin in a convergent plate boun-

dary. Tectonics 18, 686–702.

Galindo-Zaldıvar, J., Ruano, P., Jabaloy, A., Lopez-Chicano, M.,

2000. Kinematics of faults between Subbetic Units during the

Miocene (central sector of the Betic Cordillera). C. R. Acad. Sci.

Paris 331, 811–816.

Galindo-Zaldıvar, J., Gil, A.J., Borque, M.J., Gonzalez-Lodeiro, F.,

Jabaloy, A., Marın-Lechado, C., Ruano, P., Sanz de Galdeano,

C., 2003. Active faulting in the internal zones of the Betic

Cordilleras (SE Spain). J. Geodyn. 36, 239–250.

Gallegos, J.A., 1975. Los Alpujarrides al W de Sierra Nevada.

Unpubl. Ph.D. Thesis, Univ. of Granada, 494 pp.

Garcıa-Hernandez, M., Lopez-Garrido, A.C., Rivas, P., Sanz de

Galdeano, C., Vera, J.A., 1980. Mesozoic palaeogeographic

evolution of the external zones of the Betic Cordillera. Geol.

Mijnb. 59, 155–168.

Gomez Pugnaire, M.T., Braga, J.C., Martın, J.M., Sassi, F.P., Del

Moro, A., 2000. Regional implications of a Palaeozoic age for

the Nevado-Filabride Cover of the Betic Cordillera, Spain.

Schweizer. Min. Petrogr. Mitt. 80, 45–52.

Guerrera, F., Martin-Algarra, A., Perrone, V., 1993. Late Oligo-

cene–Miocene syn-/late-orogenic successions in the Western

and Central Mediterranean chains from the Betic Cordillera to

the Southern Apennines. Terra Nova 5, 525–544.

Guezou, J.C., Frizon de Lamotte, D., Coulon, M., Morel, J.L., 1991.

Structure and kinematics of the Prebetic nappe complex (south-

ern Spain): definition of a bBetic Floor ThrustQ and implications

in the Betic–Rif orocline. Ann. Tecton. 5, 32–48.

Hancock, P.L., 1985. Brittle microtectonics, principles and practice.

J. Struct. Geol. 7, 437–457.

Herraiz, M., De Vicente, G., Lindo-Naupari, R., Giner, J.,

Simon, J.L., Gonzalez-Casado, J.M., Vadillo, O., Rodrı-

guez-Pascua, M.A., Cicuendez, J.I., Casas, A., Cabanas, L.,

Rincon, P., Cortes, A.L., Ramırez, M., Lucini, M., 2000.

The recent (upper Miocene to Quaternary) and present tec-

tonic stress distributions in the Iberian Peninsula. Tectonics

19, 762–786.

Huibregtse, P., van Alebeek, H., Zaal, M., Biermann, C., 1998.

Palaeostress analysis of the northern Nijar and southern Vera

basins: constraints for Neogene displacement history of major

strike-slip faults in the Betic Cordilleras, SE Spain. Tectono-

physics 300, 79–101.

IGME— Instituto Geologico y Minero de Espana: Geological maps

1 :50,000, sheet 1040, Zafarraya (1979), sheet 1041, Durcal

(1981), sheet 1025, Loja (1988).

Jabaloy, A., Galindo-Zaldıvar, J., Gonzalez-Lodeiro, F., 1992. The

Mecina Extensional System: their relation with the post-Aqui-

tanian piggy-back basins and the paleostress evolution (Betic

Cordilleras, Spain). Geo Mar. Lett. 12, 96–103.

Jolivet, L., Faccenna, C., 2000. Mediterranean extension and the

Africa–Eurasia collision. Tectonics 19, 1095–1106.

Jolivet, L., Frizon de Lamotte, D., Horvath, F., Mascle, A., Serrane,

M., 1999. The Mediterranean Basins: tertiary extension within

the Alpine Orogen — an introduction. In: Durand, B., Jolivet,

L., Horvath, F., Serrane, M. (Eds.), The Mediterranean Basins:

Tertiary Extension within the Alpine Orogen, Geological Soci-

ety of London, Special Publication, vol. 156, pp. 1–14.

Jonk, R., Biermann, C., 2002. Deformation in Neogene sediments

of the Sorbas and Vera Basins (SE Spain): constraints on simple-

shear deformation and rigid body rotation along major strike-

slip faults. J. Struct. Geol. 24, 963–977.

Kirker, A.I., Platt, J.P., 1998. Unidirectional slip vectors in the

western Betic Cordillera: implication for the formation of the

Gibraltar arc. J. Geol. Soc. London 155, 193–208.

Kleinspehn, K., Pershing, J., Teyssier, C., 1989. Paleostress stratig-

raphy: a new technique for analyzing tectonic control on sedi-

mentary basin subsidence. Geology 17, 253–256.

Leblanc, D., 1990. Tectonic adaptation of the external zones around

the curved core of an orogen: the Gibraltar Arc. J. Struct. Geol.

12, 1013–1018.

Leblanc, D., Olivier, P., 1984. Role of strike-slip fault in the Betic-

Rifian orogeny. Tectonophysics 101, 345–355.

Lhenaff, R., 1979. Le bassin de Grenade (Andalousie, Espagne):

morphogenese d’une cuvette intra-montagneuse. Mediterranee

3, 37–44.

Lonergan, L., White, N., 1997. Origin of the Betic-Rif mountain

belt. Tectonics 16, 504–522.

Lonergan, L., Platt, J.P., Gallagher, L., 1994. The internal/external

zone boundary in the Eastern Betic Cordillera. J. Struct. Geol.

16, 175–188.

Martınez-Martınez, J.M., Azanon, J.M., 2002. Orthogonal extension

in the hinterland of the Gibraltar Arc (Betics, SE Spain). In:

Rosenbaum, G., Lister, G.S. (Eds.), Reconstruction of the evo-

lution of the Alpine–Himalayan orogeny, Journal of the Virtual

Explorer, vol. 8, pp. 3–22.

Martın Martın, J., 1980. Las Dolomıas de las Cordilleras Beticas.

Unpubl. Ph.D. Thesis, Univ. of Granada, 210 pp.

Morales, J., Vidal, F., De Miguel, F., Alguacil, G., Posadas, A.M.,

Ibanez, J.M., Guzman, A., Guirao, J.M., 1990. Basement struc-

ture of the Granada Basin, Betic Cordilleras, Southern Spain.

Tectonophysics 177, 337–348.

Morales, J., Singh, S.K., Ordaz, M., 1996. Analysis of the Gra-

nada (Spain) earthquake of 24 June, 1984 (M =5) with em-

phasis on seismic hazard in the Granada Basin. Tectonophysics

257, 253–262.

Morales, J., Serrano, I., Vidal, F., Torcal, F., 1997. The depth of the

earthquake activity in the Central Betics (Southern Spain).

Geophys Res. Lett. 24, 3289–3292.

Munoz,D.,Cisternas,A.,Udıas,A.,Mezcua, J., SanzdeGaldeano,C.,

Morales, J., Sanchez-Venero,M., Haessler, H., Ibanez, J., Buforn,

E., Pascual, G., Rivera, L., 2002.Microseismicity and tectonics in

the Granada Basin (Spain). Tectonophysics 356, 233–252.

Ornter, H., 2002. TectonicVB, 1.5a release, a program for analysis

of orientation data, Univ. of Innsbruck, Austria.

Osete, M.L., Freeman, R., Vegas, R., 1989. Preliminary palaeomag-

netic results from the Subbetic Zone (Betic Cordillera, southern

Spain). Phys. Earth Planet. Inter. 52, 283–300.

Petit, J.P., 1987. Criteria for the sense of movement on fault surfaces

in brittle faults. J. Struct. Geol. 9, 597–608.

Platt, J.P., 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.

Platt, J.P., Allerton, S., Kirker, A., Mandeville, C., Mayfield, A.,

Platzman, E.S., Rimi, A., 2003. The ultimate arc: differential

Page 22: Neotectonic evolution of the Central Betic Cordilleras (Southern Spain)

K.R. Reicherter, G. Peters / Tectonophysics 405 (2005) 191–212212

displacement, oroclinal bending, and vertical axis rotation in the

External Betic–Rif arc. Tectonics 22 (3), 1017. doi:10.1029/

2001TC001321.

Platzman, E.S., 1992. Paleomagnetic rotations and the kinematics of

the Gibraltar arc. Geology 20, 311–314.

Reicherter, K.R., 1999. Results from Paleostress Analysis of the

Granada Basin (southern Spain). Mitt. Geol.-Palaont. Inst. Univ.

Hamburg 83, 59–74.

Reicherter, K., 2001. Paleoseismological advances in the Granada

Basin (Betic Cordilleras, southern Spain). Acta Geol. Hisp. 36,

267–281.

Reicherter, K., Michel, G.W., 1993. Incremental deformation from

fault slip data — an example form the Subbetic Zone, Betic

Cordillera, Spain. Z. Dtsch. Geol. Ges. 144, 45–56.

Reicherter, K., Pletsch, T., 2000. Evidence for a synchronous cir-

cum-Iberian subsidence event and its relation to the African–

Iberian convergence in the Late Cretaceous. Terra Nova 12,

141–147.

Reicherter, K., Pletsch, T., Kuhnt, W., Manthey, J., Homeier, G.,

Wiedmann, J., Thurow, J., 1994. Mid-Cretaceous paleogeogra-

phy and paleoceanography of the Betic Seaway (Betic Cordillera,

Spain). Palaeogeogr., Palaeoclimatol., Palaeoecol. 107, 1–33.

Reicherter, K., Jabaloy, A., Morales, J., Galindo-Zaldıvar, J., Ruano,

P., Becker-Heidmann, P., Reiss, S., Gonzalez Lodeiro, F., 2003.

Repeated palaeoseismic activity of the Ventas de Zafarraya fault

(S-Spain) and its relation with the 1884 Andalusian earthquake.

Int. J. Earth Sci. 92, 912–922.

Rodrıguez Fernandez, J., Sanz de Galdeano, C., 2001. La Cuenca de

Granada. Interpretacion de su estructura a partir de perfiles

sısmicos. In: Sanz de Galdeano, C., Pelaez Montilla, J.A.,

Lopez Garrido, A.C. (Eds.), La Cuenca de Granada-Estructura,

tectonica activa, sismicidad, geomorfologıa y dataciones exis-

tentes. Univ. de Granada, Espana, pp. 109–116.

Rosenbaum, G., Lister, G.S., Duboz, C., 2002. Relative motions of

Africa, Iberia and Europe during Alpine orogeny. Tectonophy-

sics 359, 117–129.

Ruano, P., 2003. Estructuras tectonicas recientes en la transversal

central de las Cordilleras Beticas. Unpubl. Ph.D. Thesis, Univ.

of Granada, 322 pp.

Ruano, P., Galindo-Zaldıvar, J., Jabaloy, A., 2000. Evolucion geo-

logica desde el Mioceno del sector noroccidental de la Depre-

sion de Granada (Cordilleras Beticas). Rev. Soc. Geol. Espana

13, 143–155.

Sanz de Galdeano, C., 1983. Los accidentes y fracturas principales

de las Cordilleras Beticas. Estud. Geol. 39, 157–165.

Sanz de Galdeano, C., 1985. La fracturacion del borde sur de la

Depresion de Granada (discusion acerca del ecenario del terre-

moto del 25-XII-1884). Estud. Geol. 41, 59–68.

Sanz de Galdeano, C., 1990. Geologic evolution of the Betic

Cordilleras in the Western Mediterranean, Miocene to the pre-

sent. Tectonophysics 172, 107–119.

Sanz de Galdeano, C., 1997. La zona interna Betico-Rifena (Ante-

cedentes, unidades tectonicas, correlaciones y bosquejo de

reconstruccion paleogeografica). Univ. of Granada, Granada.

316 pp.

Sanz de Galdeano, C., Lopez-Garrido, A.C., 1999. Nature and

impact of the Neotectonic deformation in the western Sierra

Nevada (Spain). Geomorphology 30, 259–272.

Sanz de Galdeano, C., Lopez-Garrido, A.C., 2000. Las fallas Tor-

tonienses a Cuaternarias entre Granada y la costa: el lımite

occidental del Nevado-Filabride y de las unidades Alpujarrides

inferiores. Rev. Soc. Geol. Espana 13, 519–528.

Sanz de Galdeano, C., Serrano, F., Lopez-Garrido, A.C., Martın

Perez, J.A., 1993. Palaeogeography of the late Aquitanian–early

Burdigalian Basin in the Western Betic internal zone. GEOBIOS

26, 43–44.

Sanz de Galdeano, C., Lopez-Casado, C., Delgado, J., Peinado,

M.A., 1995. Shallow seismicity and active faults in the

Betic Cordillera. A preliminary approach to seismic sources

associated with specific faults. Tectonophysics 248, 293–302.

Spang, J.H., 1972. Numerical method for dynamic analysis of

calcite twin lamellae. Geol. Soc. Amer. Bull. 83, 467–472.

Sperner, B., Ratschbacher, L., 1994. A Turbo-Pascal program pack-

age for graphical presentation and stress analysis of calcite

deformation. Z. Dtsch. Geol. Ges. 145, 414–423.

Stapel, G., Moeys, R., Biermann, C., 1996. Neogene evolution of

the Sorbas basin (SE Spain) determined by paleostress analysis.

Tectonophysics 255, 291–305.

Staub, R., 1927. Uber Gliederung und Deutung der Gebirge Ma-

rokkos. Eclog. Geol. Helv. 20, 275–288.

Vissers, R.L.M., Platt, J.P., van der Wal, D., 1995. Late orogenic

extension of the Betic Cordillera and Alboran Domain: a litho-

spheric view. Tectonics 14, 786–803.

Weijermars, R., 1991. Geology and tectonics of the Betic zone, SE

Spain. Earth Sci. Rev. 31, 153–236.