Academia.eduAcademia.edu
Sedimentary Geology 192 (2006) 141 – 166 www.elsevier.com/locate/sedgeo Timing of canyon-fed turbidite deposition in a rifted basin: The Early Cretaceous turbidite complex of the Cerrajón Formation (Subbetic, Southern Spain) Pedro A. Ruiz-Ortiz ⁎, Ginés A. de Gea, José M. Castro Dpto. de Geología, Universidad de Jaén, Campus Universitario, 23071-JAÉN, Spain Received 25 December 2004; received in revised form 20 March 2006; accepted 3 April 2006 Abstract The Subbetic Cerrajón Formation of latest Hauterivian–earliest Late Albian age is made up of turbidite sandstones and calcarenites intercalated in marly-limestones and marls. It can be defined as a turbidite complex with three distinct intervals where turbidites are recorded: latest Hauterivian–Barremian p.p., Late Aptian p.p. and Middle to middle-late Albian age. The distribution of four biochronostratigraphic units of these ages throughout the studied outcrops is shown. The main hiatuses, latest Barremian–early Late Aptian and latest Aptian–early Late Albian, mainly the latter, have been identified in many other outcrops of the Southern Tethys, and are related to tectonics and associated palaeogeographic and palaeoceanographic changes. Platform environments, as the source area of the clastics, the feeding submarine canyon, base-of-slope channels and related basinal lobes or sheet systems, have been identified and can be traced laterally along more than 80 km of discontinuous outcrops aligned in a downcurrent direction. The system could have extended more than 150 km in length. The mean palaeocurrent direction fits well with the present alignment of the outcrops and, after corrections for tectonics rotation, with the current palaeogeographic model of the Southern Iberian Continental Palaeomargin (SICP). This conclusion largely supports previous interpretations of a longitudinal pattern for the distribution of clastics during the Cretaceous in that part of the SICP. The ages of the main Cretaceous events deduced from the study of the Cerrajón Formation correlate quite well with the sequence sets and subsidence history of the Prebetic Zone, where the Cretaceous platform environments adjacent to the turbidite basin are recorded. Moreover, the main unconformities dated, such as the onset and the end of turbidite sedimentation, show good correlation with the sequence boundaries described for the European basins. The analysis in time and space of the turbidite distribution and its correlation with the Prebetic carbonate platforms enables us to outline some interesting conclusion that might be a suitable subject for future research. © 2006 Elsevier B.V. All rights reserved. Keywords: Subbetic; Cretaceous; Turbidites; Submarine canyon; Rift basin 1. Introduction Turbidites are common deposits in deep water basins and a wide range of styles of turbidite basins and systems ⁎ Corresponding author. Dpto. Geología, Campus Universitario, Ed. B-3-339, 23071-JAEN, Spain. Fax: +34 953 212 141. E-mail address: paruiz@ujaen.es (P.A. Ruiz-Ortiz). 0037-0738/$ - see front matter © 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.sedgeo.2006.04.003 have been identified (Mutti and Normark, 1987; Shanmugam and Moiola, 1988; Weimer and Link, 1991; Richards et al., 1998; Mutti et al., 1999; 2003; among others). Tectonics and sediment supply (Mutti and Normark, 1987; Mutti et al., 1996, 1999, 2003) are the main factors controlling the long-term stability of the locus of turbidite deposition in such a way that a turbidite complex can be formed. The terms ‘turbidite complex’ 142 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 and ‘turbidite system’ (sensu stricto) are used here as defined by Mutti and Normark (1987). The arrangement of the feeding canyon — submarine fan system in most of the basins is usually transverse to the marginal slope. However, in elongated basins, such as foreland basins, the morphology of the basin can favour a clastic longitudinal, or axial, distribution pattern. This is also the case for extensional rifted basins where subsiding troughs can result from the alignment of fault-bounded half-grabens, and sediment is distributed in a direction parallel to the fault strike (Leeder and Gawthorpe, 1987; Ravnas and Steel, 1997, 1998; Bosence, 1998; Purser et al., 1998; Gawthorpe and Leeder, 2000). From previous studies on the Cretaceous turbidites (Cerrajón Formation) in the Intermediate Domain of the Subbetic, SE Spain, Ruiz-Ortiz (1980, 1981) and Maldonado and Ruiz-Ortiz (1982) proposed a longitudinal dispersal pattern for the clastic sediments of the Cerrajón Formation. The number and location of entry points for the terrigenous clastics and their relation with the sedimentation, mainly carbonate, on the adjacent Prebetic platform, a more precise dating of both the deposits and the probable hiatuses, the size and style of the turbidite system elements and their distribution in time and space, all remain to be established. The discovery of the record of the history of a submarine canyon in the Cretaceous outcrops of the Intermediate Domain in the Huelma region (east of Jaén) (de Gea et al., 1998; Ruiz-Ortiz et al., 2001a), along with the detailed biostratigraphic framework made by de Gea (2003) from a study of calcareous nannofossils and planktonic foraminifers from Early Cretaceous hemipelagic sediments, together with the new biostratigraphic and sedimentological data supplied in this paper, provide a new setting enabling a deeper understanding of the relations in time and space of the canyon with other elements of the turbidite systems. The main goals of this paper are as follows: firstly, to describe the characteristics of the stratigraphic record of both a Cretaceous submarine canyon and the corresponding downcurrent turbidite system elements, channels and lobes, as well as their relations in time and space; then, to analyze the stratigraphic architecture of the Cerrajón turbidite complex, dating the main events recorded in it, and, finally, to identify the processes controlling turbidite sedimentation, the turbidite system and the element build-up within it, and to integrate all these into a regional and supraregional stratigraphic framework. In fact, the main phases in the evolution of this turbidite complex were related to stages of special significance in the evolution of the basin. Well known examples of turbidite systems and complexes are to be found in foreland basins and passive margin settings, whereas the Cerrajón turbidite complex was accumulated in a basin affected by extensional tectonics (Vilas, 2001; Vilas et al., 2003) in a transform continental margin setting (cf. Vera, 2001), which adds additional interest to the results presented in this paper. 2. Geological setting The External Zones of the Betic Cordillera (Fig. 1) comprise sedimentary rocks deposited on the Southern Iberian Continental Palaeomargin (SICP) (Fig. 1A) during the Alpine tectonic cycle (Mesozoic to Early Miocene ages). The palaeomargin had a WSW–ENE orientation (Fig. 1A). To the N–NW of the Cordillera, Prebetic para-autochthonous units are overthrust by southerly-derived Subbetic allochthonous units (Fig. 1B). On the basis of stratigraphic and tectonic criteria, the Subbetic is divided into Intermediate, External, Median and Internal domains or subzones, from northwest to southeast respectively (Fig. 1), based largely on the occurrence of different thrust units (GarcíaHernández et al., 1980; Ruiz-Ortiz et al., 2001b; Vera, 2001). The Intermediate subzone of the Subbetic was a subsiding trough where thick successions of Cretaceous pelagic and turbidite sediments accumulated (Fig. 1C). This trough was bounded to the north and northwest by the Prebetic shallow marine carbonate platform, or by emerged areas, and to the south and southeast by a discontinuous string of pelagic swells, the deposition site of the External Subbetic series. At present, the southwestern boundary of the Prebetic Zone in Sierra de Cazorla is a NW–SE striking fault, known as the Tiscar Fault (TF, Fig. 1B,D), probably corresponding to an ancient transfer fault which played a role in Cretaceous palaegeography similar to that of other important faults crossing the Prebetic outcrops (e.g. Vinalopó Fault, Vilas and Querol, 1999). The Tiscar Fault operated during the Miocene as a dextral strike-slip fault with a horizontal displacement of 5.8 km (Foucault, 1974; in Sanz de Galdeano, 2003) and also down throwing its southwestern block (Sanz de Galdeano, 2003). There is evidence suggesting the absence, or a drastic reduction of the extent, of any type of marine platform during the Early Cretaceous in the region currently sited west of Cazorla (Ruiz-Ortiz et al., 2001c; de Gea, 2003; Martínez del Olmo, 2003) (Fig. 1B). In this region of the central sector of the cordillera, between Sierra de Cazorla and Córdoba province, the Early Cretaceous pelagic P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 143 Fig. 1. A: Palaeogeographic reconstruction of the Southern Iberian Continental Palaeomargin (SICP) during the Late Jurassic. The numbers in A correspond to those of the upper right key. The location of section I–I′ in C is shown. B: Geologic map of the Betic Cordillera with the location of the studied area (map in D) and some of the localities cited in the text. C: Schematic section with the palaeogeography of the SICP for the Aptian–Albian. D: Geologic map of the south and east of the Jaén region with the location of the studied outcrops; Hu: Huelma outcrops; Lg: La Guardia outcrops; Vi: Los Villares outcrops; Ma: Martos outcrops. A and B, modified from García-Hernández et al. (1989). C, modified from Vera (1983). 144 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 environments sited immediately southward of the marginal platform would have formed part of a trough about 30 km wide and, at least, 150 km long with an irregular, fault-dissected basin floor (de Gea et al., 2000, 2001; Ruiz-Ortiz et al., 2001c). The trough was probably formed by the alignment of half-graben basins among which differential subsidence occurred. This was the palaeogeographic setting where the bulk of the turbidites making up the Cerrajón Formation were deposited, these representing the most important volume of Mesozoic turbidites recorded in the SICP. In this paper the outcrops of the Cerrajón Formation placed in the region comprised between the Tiscar Fault and the town of Martos, to the west of Jaén, are studied (Fig. 1B and D). They represent the better outcrops of the Cerrajón Formation all around the Intermediate Domain of the Subbetic (Betic External Zones). 3. Cerrajón formation On an outcrop scale, the Cerrajón Formation is made up of sandstones, marls and marly limestone interbeds. The marls are locally dark grey or black and the sandstones sometimes incorporate calcareous allochems, which are dominant in some calcarenite beds. The Cerrajón Formation is underlain by the Los Villares Formation and overlain by the Represa Formation, both composed of rhythmically alternating marls and marly limestones (Fig. 2). The base of the Cerrajón Formation is a sharp boundary marked by the first appearance of terrigenous turbidites in the latest Hauterivian, whereas the top is represented by a gradual transition towards the pelagic deposits of the Represa Formation at the middle-late Albian transition (Fig. 2). The thickness of the formation shows lateral Fig. 2. Chronostratigraphic chart with the ammonite and calcareous nannofossil biozonation and the lithostratigraphy of the studied outcrops. The biochronostratigraphic units differentiated in the Cerrajón Formation are also shown. ⁎: Globigerinelloides algerianus planktonic foraminifer Zone. ⁎⁎: Interval of coexistence of Cribrosphaerella ehrenbergii, Rhagodiscus achlyostaurion, Axopodorhabdus albianus and Hayesites albiensis. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 variations, reaching a maximum of about 1350 m in the outcrops of Los Villares, with a minimum of 150 m in the Huelma outcrops, where several erosive surfaces are superimposed. 145 The data and interpretations presented in this paper were obtained from outcrops of the Cerrajon Formation located to the south and southeast of the city of Jaén (Fig. 1D), specifically: 1) the Huelma outcrops, where Fig. 3. A: Microphotography of an exotic block of open platform sandy calcarenites with large orbitolinids and planktonic foraminifers of Early– Middle Albian age. B: Close-up of an exotic block of rudist-bearing limestones intercalated in the canyon-filling Cerrajón brown marls of the Huelma outcrops. C: The Cerrajón Formation with exotic blocks occurs overlying the Upper Ammonítico Rosso Formation of Middle–Late Jurassic age in the Huelma outcrops. D: Field view of the interchannel facies association of La Guardia outcrops and close-up of a typical bed modelled by ripples. E: Sandstone packet of the interchannel facies association from the Unit-I Cerrajón turbidites at the La Guardia outcrops interpreted as a lobe probably deposited in relation to crevasse splay processes; note the upward expansive character of the beds. F: Panoramic view of the channel-fill deposits of the Cerrajón Formation Unit-II in the La Guardia outcrops; note the erosive base carved on Unit-I slumped deposits of the Cerrajón Formation; encircled person (G.A. de Gea) for scale. G: Thick bed turbidite of the Unit-IV Cerrajón Formation in the Los Villares outcrops. H: Close-up of the sole marks of the thick-bed of G. I: Thin-bedded turbidites of the Cerrajón Formation Unit-III in the outcrops of the Los Villares. J: The Peña de Martos olistolith consisting in Jurassic limestones intercalated in Upper Hauterivian deposits stands up in the landscape of the Martos region. 146 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Fig. 3 (continued). the stratigraphic record of a submarine canyon (RuizOrtiz et al., 2001a) has been found (Fig. 3A to C); 2) the La Guardia outcrops, where sandstone bodies interpreted as channel-fill deposits are intercalated within marlstones (Fig. 3D to F); 3) the Los Villares–Martos outcrops, where the thickest development of turbidites is recorded (Fig. 3G to J). These outcrops are distributed along a discontinuous strip 45 km in length (Fig. 1). References to the Campillo de Arenas and Carcabuey outcrops (Fig. 1B) are also included. The outcrops of La Guardia, Los Villares and Martos all form part of the same tectonic unit, (Jabalcuz–San Cristóbal unit), whereas the P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Huelma outcrops are part of the Huelma unit. The Huelma unit crops out below Oligocene–Miocene deposits mainly made up of resedimented Triassic clays. From a palaeogeographic point of view, the Huelma unit is assigned to the Intermediate Domain of the Subbetic Zone, as the Jabalcuz–San Cristóbal unit, from the base of its Jurassic and Cretaceous stratigraphy (Ruiz-Ortiz et al., 1996). 3.1. Biostratigraphic dating The relative abundance of calcareous nannofossils, planktonic foraminifers and, locally, ammonites has enabled the Cerrajón Formation to be dated as latest Hauterivian–middle-late Albian transition (Aguado et al., 1996; de Gea, 2003) (Fig. 2). The base is slightly younger than that of the Nannoconus circularis Subzone and the top coincides with the Tranolithus phacelosus Subzone of nannofossils (Fig. 2). Within the stratigraphic succession of the Cerrajón Formation several hiatuses have been 147 recognized (de Gea, 2003), some of them being associated with clear erosive surfaces. Sedimentation is recorded in three-main time intervals separated by two main hiatuses (Fig. 2). The oldest sedimentation interval within the Cerrajon Formation has been dated as latest Hauterivian–Late Barremian, from the ammonites obtained at the base of the formation (see below) and the nannofossils (Nannoconus bucheri Zone, N. circularis Subzone and Micrantholithus hoschulzii Zone of nannofossils) from the lower part of the stratigraphic successions in different outcrops. The central interval recorded corresponds to a part of the Late Aptian dated with nannofossils (Rhagodiscus angustus Zone) and, more precisely, with planktonic foraminifers (Globigerinelloides algerianus Zone) (Fig. 2). The upper interval recorded within the Cerrajón Formation has been dated as Middle Albian (upper part of the Prediscosphaera columnata nannofossil Zone, upper part of the Cribrosphaerella ehrenbergii Subzone, characterized by the coexistence of C. ehrenbergii, Rhagodiscus achlyostaurion, Fig. 4. Interpreted cross-section of the Cerrajón Formation in the Huelma outcrops deduced from the correlation of the four shown columns and complementary field observations. Two erosion surfaces, the base of the Lower Hauterivian calcarenites and of the Cerrajón Formation respectively, occur. In this cross-section the Cerrajón Formation overlies Upper Jurassic calcareous turbidites of the Toril Formation, but locally the erosion reached the Middle Jurassic (Fig. 5). Note that the exotic blocks are more abundant in the deepest part of the erosive surface number 2. Also the change with time of the thalweg or deepest part of the canyon from the first to the second erosive surface can be observed. 148 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Axopodorhabdus albianus and Hayesites albiensis) to middle-late Albian transition (P. columnata Zone, T. phacelosuss Subzone) (Fig. 2). The lower and central intervals are identified as biochronostratigraphic Units I and II respectively. The recorded upper interval can be subdivided into two other biochronostratigraphic units, Units III and IV, of Middle Albian and middle-late Albian transition age respectively. Ammonites recollected from the top of the Los Villares Formation and at the base of the Cerrajón Formation in the outcrops of La Guardia, have enabled the base of the Cerrajón Formation to be dated as the upper part of the Pseudothurmania picteti Subzone of ammonites, Pseudothurmania ohmi Zone, whereas ammonites from the top of the biochronostratigraphic Unit-I have enabled the dating of the Imerites giraudi Zone. Field work carried out in sections of the La Guardia outcrops made it possible to identify the Taveraidiscus huggi and the Kotetishvilia nicklesi ammonite Zones of the Early Barremian in the lower part of the Cerrajón Formation (Company et al., personal communication). The four biochronostratigraphic Units (I to IV) differentiated within the Cerrajón Formation are only all present in the Los Villares section. In contrast, in the more westerly part of the Los Villares–Martos outcrops, in the vicinity of the town of Martos, only Unit-IV is present. In the outcrops of Huelma Units III and IV occur whereas in those at La Guardia, Units I, II and III are present (Figs. 1D and 2). Unit-IV shows the greatest development of turbidites, with thicknesses of about 900 m, in the most rapidly subsiding area (Los Villares outcrops). A more detailed account of the biochronostratigraphy of the Cerrajón Formation can be found in Ruiz-Ortiz et al. (2001a) and de Gea (2003). In the latter, the correlation to the Early Cretaceous ammonite biozones is also discussed. 3.2. Studied outcrops 3.2.1. The Huelma outcrops The Early Cretaceous sedimentary record at this locality is very discontinuous and is represented by (Fig. 4): 1) Upper Berriasian marls and white marly limestones with very thin-bedded sandstone intercalations (Los Villares Formation); 2) Lower Hauterivian marls and calcarenites; 3) The Middle Albian and the middle-late Albian transition, the biochronostratigraphic Units III and IV of the Cerrajón Formation, composed here of brown sandy marls with calcareous sandstone and orbitolinid-bearing calcarenite intercalations, together with exotic blocks of a variable nature (Fig. 3A to C). Detailed mapping and dating of the outcrops of this region allowed Ruiz-Ortiz et al. (2001a) to Fig. 5. Chronostratigraphic chart of the Huelma outcrops in which the estimate ages of the erosion (incision) and sedimentation (filling) processes of the submarine canyon have been differentiated. 1. Maximum time span during which the erosive surface number 1 was originated. 2. Time span of deposition of the preserved deposits of the marls and calcarenites (canyon-fill deposits). 3. Time span with no record of deposits; during the Barremian p.p. and the Late Aptian, the canyon acted as a main entry for clastics to the trough of the Intermediate Domain of the SICP, though this long-lasting hiatus is probably the result of more than one episode of cut-and-fill of the submarine canyon. 4. Maximum time interval for erosion on the adjacent platform settings and subsequent deposition of exotic blocks on the canyon floor, along with canyon incision reaching the oldest deposits found also as exotic blocks. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 identify two erosion surfaces carved in Middle–Late Jurassic and very Early Cretaceous deposits (Figs. 3C and 4). The first of these surfaces originated between the 149 Late Berriasian, the age of the youngest deposits preserved locally below the erosive surface, and the Early Hauterivian, the age of the oldest deposits overlying the erosive Fig. 6. A: Stratigraphic sections of the Lower Hauterivian marls and calcarenites of the Huelma outcrops. B: Field view of the calcarenite succession. C: Close-up of the breccia and pebbly calcarenite beds making the lowest part of the cycles. 150 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Fig. 7. Aguzadera gorge stratigraphic section and detailed logs of La Guardia outcrops. The palaeocurrents measured at different stratigraphic levels of the section are shown. For comments see text. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 surface. This surface makes up the base of the Lower Hauterivian marls and calcarenites (Figs. 2 and 4). The second erosion surface (Fig. 3C) coincides with the base of the Cerrajón Formation which directly overlies Kimmeridgian–Tithonian (Figs. 4 and 5) and, locally, Middle Jurassic limestones (Figs. 3C and 5). There is an important hiatus, from the Late Jurassic to the middle-late Albian (Fig. 4) that, locally, reaches the Middle Jurassic (Figs. 3C and 5). Lower Hauterivian marls and calcarenites make up a local stratigraphic unit organized in repetitive thinningand fining-upward cycles (Fig. 6), in such a way that the basal beds of the cycles, often pebbly calcarenites, evolve upwards to thinner and finer sandstones, and the cycle ends with a marl bed. These deposits correlate with some calcarenite beds of the same age intercalated in the marly-limestone/marls rhythmic succession of the upper part of the Los Villares Formation at the La Guardia outcrops (Fig. 2). The brown sandy marls of the Cerrajón Formation have been dated as Middle- and middle-late Albian transition, biochronostratigraphic Units III and IV of the Cerrajón Formation. The sandy marls present a chaotic appearance and the thin-bedded calcarenite intercalations are disturbed and contorted. Exotic blocks, usually having 151 a high index of roundness, occur intercalated in the UnitIII marls (Fig. 3A to C). According to their lithology they were classified by Ruiz-Ortiz et al. (2001a) into four types: 1) Inner platform limestones with rudists, corals and large bivalves (Fig. 3B), and orbitolinid-bearing calcarenites, of Late Aptian–earliest Albian age; 2) Open platform sandy calcarenites with large orbitolinids (Fig. 3A) and planktonic foraminifers, of Early-Middle Albian age; 3) Calcarenites and calcirudites of the Lower Hauterivian, resulting from the erosion of the underlying deposits; 4) Red nodular limestones of ammonitico rosso facies from Middle-Late Jurassic age. The blocks occur intercalated in the Unit-III brown marls of the Cerrajón Formation in the same stratigraphic order as described from 1, the lowest, to 4, the highest. The maximum concentration of exotic blocks coincides with the area where the second erosion surface, the base of the Cerrajón Formation in the area, is carved in the oldest rocks (Figs. 3C and 4). 3.2.2. The La Guardia outcrops A succession of thin-bedded turbidite sandstones, marls and marly-limestones with scattered massive sandstone (and/or calcarenite) bodies makes up the outcrops of La Guardia (Figs. 3D to F, 7 and 8). The section of the Fig. 8. Correlation of some of the studied sections of the La Guardia outcrops. Cerrajón Formation Units I, II and III crop out. 152 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Aguzadera gorge (Fig. 7) can be considered as a type section of these outcrops. In this section, the Cerrajón Formation is more than 200 m thick and is overlaid unconformably by Miocene bioclastic calcarenites and white marls. The biochronostratigraphic Units I, II and III of the Cerrajón Formation have been recognized in these outcrops. The Cerrajón Formation Unit-I is made up of a succession of marls with thin-bedded turbidite intercalations. The thin-bedded turbidites are often composed only of Bouma's “c” division with the top surface being modelled by ripples (Fig. 3D and Log 2 of Fig. 7). Some sandstone packets composed of beds 20–30 cm thick, locally up to 1 m, are intercalated (Figs. 3E and 7). Fig. 9. A: Palaeocurrents measured at La Guardia outcrops; the mean direction and sense deduced from the channels and interchannel deposits, respectively, are represented by arrows. B: Palaeocurrents measured from flutes and grooves in channel-fill deposits of La Guardia outcrops. C: Palaeocurrents measured from interchannel deposits of La Guardia outcrops. D: Palaeocurrents measured from turbidite beds at Los Villares–Martos outcrops; data coming from different biochronostratigraphic units are differentiated. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 In the Aguzadera gorge section, the Cerrajón Formation Unit-II is made up of a sandstone body about 30 thick composed of massive and amalgamated beds of medium to fine sand and calcarenites. It has a clearly erosive base (Fig. 3F) and a Fe-oxide stained and bioturbated top. A detailed section of this body is shown in Fig. 7 (Log 3). Two different parts of this body can be distinguished: a lower one composed of amalgamated sandstones with few 153 calcareous grains, and an upper one, made up basically of calcarenites with marly intercalations. One of the main marly intercalations lies between the lower sandstones and the upper calcarenites (Fig. 7). From the beds directly underlying the latter sandstone body, ammonites of the I. giraudi Zone, Late Barremian, have been collected. The presence of the nannofossil R. angustus in the marly intercalations Fig. 10. Some of the studied sections in the Los Villares–Martos outcrops. Note the calcareous nannofossils Zones shown on the left part of sections A and D. The stratigraphic location of logs 1 to 7, represented in Figs. 11, 12 and 13 is also shown. 154 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Fig. 11. Bed by bed section (log) of the lower part of the biochronostratigraphic Unit-I of the Cerrajón Formation in the Los Villares outcrops. The stratigraphic location of the log is shown in section A of Fig. 10. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 of the middle and upper parts of these deposits allows us to date them as Late Aptian. The age of the lower part of the body, composed of amalgamated sandstones, is therefore between Late Barremian and Late Aptian in age. The lateral continuity of this sandstone body is difficult to assess. Other sandstone bodies of the same age 155 outcropping in the area could be lateral equivalents or part of the same body tectonically detached (Fig. 8). The only valid tool for correlating at this scale, the physical correlation of the beds in the field, allows us to assume a minimum lateral extension of the Upper Aptian clastic bodies between 550 and 900 m; their thickness varies between 15 and 30 m. Fig. 12. Logs of the Cerrajón Formation biochronostratigraphic Unit-II in the Los Villares outcrops. The stratigraphic position of the logs is shown in section B of Fig. 10. The key is the same as in Fig. 11. 156 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Fig. 13. Logs of the Cerrajón Formation biochronostratigraphic Units III and IV in the Los Villares outcrops. The stratigraphic position of the logs is shown in section A of Fig. 10. The key is the same as in Fig. 11. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 The Aptian sandstone body of the Aguzadera section (biochronostratigraphic Unit-II) occurs above slumped thin bedded sandstones of biochronostratigraphic Unit-I of the Cerrajón Formation (Fig. 3F) and it is covered by about 4 m of Upper Aptian marls which are upwards replaced by marls with thin-bedded turbidites of the Middle Albian (biochronostratigraphic Unit-III) (Fig. 7). Another clastic body with an erosive base, in this case composed of amalgamated calcarenites, occurs intercalated in the Middle Albian marls of the Aguzadera section (Fig. 7). The calcarenites are massive or locally with a coarse plane horizontal lamination, and have abundant bioclasts and orbitolinids. This calcarenite body is about 8 m thick and its lateral continuity poses the same problems as the underlying Aptian sandstone body. The vertical stacking of two calcarenite bodies intercalated in the Middle Albian marls can be seen in the southern part of the La Guardia outcrops (Fig. 8). Palaeocurrents. In Fig. 9A, a total of 58 palaeocurrent measurements are represented, all of these obtained from the La Guardia outcrops. Two main directions of transport can be differentiated, one of these indicating a mean transport sense towards the NW and the other indicating a N–S or N–NE–S–SW direction of transport. The first of these main palaeocurrent directions comes from the sole marks of the sandstone bodies of Aptian and Albian age. The other one comes from flutes and tool marks from thin-bedded turbidite intercalations of, mainly, biochronostratigraphic Unit-I. In Fig. 7, in which the Aguzadera gorge section is represented, the palaeocurrents measured at each stratigraphic level of the section are shown. The two main palaeocurrent directions resulting from Fig. 9A form an angle of approximately 90°, as do the palaeocurrents deduced from ripple marks at the top of the beds, with respect to each of the mean palaeocurrents measured from the sole marks. 3.2.3. The outcrops of Los Villares–Martos The thickest sections (up to 1350 m) and the widest outcrops of the Cerrajón Formation are encountered just to the south of the city of Jaén, in the Los Villares–Martos region (Figs. 1D and 10). The Cerrajón Formation in these outcrops is made up of thin-, medium- and thick- (more than 1 m) bedded turbidites, arranged in non-channelized sequences (Fig. 3G to I). These are intercalated in background deposits composed of rhythmically alternating marly limestones and marls, with ammonites, planktonic foraminifers and nannofossils. In the Los Villares outcrops, the four Cerrajón Formation biochronostratigraphic units I to IV that have been distinguished crop out stacked one upon the other (Section A, Fig. 10), as mentioned above. Unit-I in the 157 Los Villares outcrops is about 330 m thick and its basal portion shows the onset of the turbidite sedimentation in the Cretaceous basin, which is recorded by some thinbedded sandstones intercalated with marls and marly limestones (Fig. 11). The proportion of turbidite sedimentation, sandstones and shales, increases upwards in such a way that a stepping transition with about four sandrich “steps” can be differentiated in the 133 m of sediments represented in Fig. 11. This transition is accompanied by frequent slumps which can affect bundles of beds up to 10 m thick. The turbidite beds have plane bases and tops and they show Bouma sequences that are either complete or base-cut-out (Fig. 11). The absence of the nannofossil Calcicalathina oblongata in the samples of the upper part of Unit-I in the Los Villares section (Section A, Fig. 10) allows us to identify the M. hoschulzi nannofossil Zone. The Cerrajón Formation Unit-II in the Los Villares outcrops is about 100 m thick and is represented by sandstone/shale couplets with some local orbitolinid-rich intercalations (Figs. 10 and 12). The turbidite beds have generally plane bases and tops, are massive or more often they show Bouma's sequences complete or top-cut-out (Fig. 12). The Cerrajón Formation Units III and IV in the Los Villares outcrops (Figs. 10 and 13) make up a succession about 1000 m thick that represents the time interval with the greatest volume of Cretaceous turbidite sedimentation in the whole of the SICP. Minimum values for the mean sedimentation rate could have been as high as 200 mm/ka. The mean thickness of the turbidite beds is greater in these biochronostratigraphic units (Fig. 3I and J) than in the underlying units. Base-cut-out Bouma's sequences are predominant in the turbidite sandstones (Fig. 13). Organic matter-rich black shale interbeds are common towards the upper part of Unit-IV and a single orbitolinid-rich bed occurs near the top of the Cerrajón Formation. Moreover, carbonate allochems, some of which are carbonate-coated quartz grains, are relatively frequent in the sandstones, and debris of wood are common in the sandstones and in the lutites of the Los Villares section (Figs. 10 and 13). In the more westerly Martos outcrops, the Cerrajón Formation occurs overlying some levels of breccia which contain Jurassic and Early Cretaceous clasts (section D, Fig. 10). A huge olistolith, about 0.5 km3 in volume, the “Peña de Martos”, a distinctive hill in the landscape of the region, which is made up of more than 500 m of Jurassic limestones, also occurs (Figs. 3J and 10). Unit-I Cerrajón turbidites 50 m thick crop out about 2 km to the east of the Peña de Martos olistolith (Section D in Fig. 10). This Unit-I pinches out in the surroundings of the olistolith outcrop. No Unit-II Cerrajón turbidites crop out in the 158 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 surroundings of Martos (Fig. 10). Units III and IV are represented in the Martos outcrops (N400 m in thickness; section D, Fig. 10), but always composed of fine grained and thin-bedded turbidites. Unit-IV Cerrajón turbidites overlie the Peña de Martos olistolith, while references to the presence of middle-upper Albian Cerrajón turbidites have been obtained from outcrops located as far as 45 km to the west-southwest of the Martos area (Carcabuey outcrops, Roldán-García et al., 1988), and also from more southerly Subbetic outcrops (Campillo de Arenas, Fig. 1B) (de Gea, 2003), which highlights the extensive character of the middle-upper Albian transition turbidite sedimentation (Unit-IV) of the Cerrajón Formation. The palaeoflow measures obtained from sole marks (flute and tool marks) of the beds represented in Figs. 11 to 13 and from neighbouring sections of the Los Villares outcrops are represented in Fig. 9D. A significant NW–SE direction of flow for the turbidite currents and a concentration of the measured palaeocurrents between 230° and 310° occur, which fit well in the pattern of flow of the La Guardia outcrops (Fig. 9A and B). However, a relative dispersion of the palaeocurrents also is shown. The implications of these observations are considered below. 4. Basic elements of the turbidite systems 4.1. Feeding canyon The outcrops of the Huelma region were interpreted by Ruiz-Ortiz et al. (2001a) as the record of the incision and filling of a submarine canyon. The age of the first erosive surface, Late Berriasian–Early Hauterivian, provides an estimation of the age of the beginning of the canyon incision. Lower Hauterivian calcarenites, organized in thinning and fining upward sequences, are the probable record of the filling of the canyon's axial channels, and represent the former canyon-fill deposits (Ruiz-Ortiz et al., 2001a). Lower Hauterivian turbidite calcarenites interbedded in the upper part of the Los Villares Formation in the La Guardia outcrops (Fig. 14) could be Fig. 14. Chronostratigraphic chart of the studied deposits and correlation with the sequence sets defined by Vilas et al. (2003) from the Prebetic Zone and the sequences of the European basins of Hardenbol et al. (1998). P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 the downcurrent equivalent of these deposits. The second erosive surface (Fig. 5) was originated after the deposition of the Lower Hauterivian marls and calcarenites and prior to the deposition of the former exotic blocks, which came from source rocks of Late Aptian–Early Albian age. The latest Hauterivian, age of the onset of the Cerrajón turbidite deposition in distal areas, may be considered as its most probable age. However, more than one episode of canyon incision could be involved in the genesis of this second erosive surface. Fig. 5 shows the time relations among the erosion surfaces, the lithostratigraphic units occurring in the canyon fill and the underlying deposits. A valuation of the age of the incision and filling of the submarine canyon is indicated (1 to 4 in Fig. 5). The beginning of the massive canyon fill that is currently preserved took place from the Middle Albian. No record of Upper Hauterivian to Middle Albian (≈26 Ma) sediments has been found. To elucidate what happened during this long-lasting time interval is highly speculative. The coeval sedimentary record from more distal areas (the outcrops of La Guardia and Los Villares– Martos) leads us to deduce that, at least during the Barremian and Late Aptian, the canyon acted as a main entry, funnelling clastics to basinal areas. The presence, today, of two erosive surfaces in the record of the fill of the canyon is probably just the final result of a multistorey of cut-and-fill episodes. Specifically, deep erosion associated to the second erosive surface which reaches to cut Middle Jurassic rocks (Fig. 5), might have eliminated the record of previous cut and fill episodes. 4.2. Base-of-slope channels The sandstone bodies embedded in marls with thinbedded turbidites outcropping in the La Guardia are interpreted as channel-fill deposits. These present some of the features typical of ancient depositional channels (Mutti and Normark, 1987) such as their medium to fine grain size, the absence of lag deposits, the presence of thick-bedded and amalgamated axial facies (lower part of Log 3 in Fig. 7) and alternating sandstone and mudstone beds in the upper part (upper part of Log 3 in Fig. 7). Also the morphological relations among the beds and the erosive base in the channel margin may be consistent with such an interpretation (Fig. 3F). These channels would have been active as conduits for clastics bypassing this part of the basin (Fig. 15) during the time interval represented by the hiatus associated to the erosive base of the sandstone bodies. The age of the channel-fill deposits is probably Late Aptian for the stratigraphically lowest sandstone bodies and Middle Albian for the uppermost sandstone bodies (see above). 159 Each sandstone body represents, at least, two different sedimentary episodes, an initial one in which these channels acted as transfer zones and were bypassed by the flows transporting sand and mud to more distal areas, and a second phase of channel-fill when the channel was part of the depositional zone of the system/ s. However, the presence of hemipelagic marl beds in the channel-fill deposits (Fig. 7), and also the change in composition of the sand-size clastics from sandstones to calcarenites in the Late Aptian clastic bodies, suggests that the fill of the channels was a complex polyphasic depositional process. The thick packages of mudstone facies with thinbedded turbidite intercalations making up the background deposits where the sandstone bodies are embedded in the La Guardia outcrops (Fig. 3D and E) are interpreted as having been deposited mainly through overbank processes in interchannel areas. This facies group is usually affected by sediment slumping or associated with it (e.g. Mutti and Normark, 1987). Channel margin facies can also form part of this facies association, as packages tens of metres thick with a relatively high sand/shale ratio and composed of thin-bedded turbidites usually formed only by Bouma's c division (Mutti, 1977; Pickering, 1982; Mutti and Normark, 1987; among others) (Fig. 3D). Others sandstone packets composed of a few relatively thick sandstone beds of limited lateral continuity (Figs. 3E and 7) could represent crevasse splay deposits (Mutti and Normark, 1987). The palaeocurrent measurements obtained from the turbidite beds of this facies association in the La Guardia outcrops show a transport direction that is normal or oblique to the mean transport direction (W– NW) deduced from the channel-fill deposits, as shown in Figs. 7 and 9, which is another argument supporting the interpretation given to the facies associations of the La Guardia outcrops. These outcrops, therefore, would represent the record of a base-of-slope environment throughout the late Early Cretaceous. 4.3. Basinal lobe or sheet system The Los Villares–Martos outcrops show the most distal turbidites in the whole of the study region. The abundance of slumps in the Cerrajón Formation Unit-I of the Los Villares outcrops, just coinciding with the beginning of the turbidite sedimentation, could be explained by reference to tectonic events triggering the gravity flows. Cerrajón Formation Units I to IV are recorded in the Los Villares outcrops (Figs. 10 and 15). In contrast, westward, in the westernmost part of the study area (the western part of the Martos outcrops), 160 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Fig. 15. A: Schematic and simplified map showing the currently position of the Intermediate Domain and the Prebetic Zone. B: Interpretative longitudinal cross-section along the Intermediate Domain from the platform edge (Prebetic Zone), near Cazorla, to the basinal distal areas in the surroundings of the Martos region, showing the relative palaeogeographic position of the turbidite systems elements of the Cerrajón Formation turbidite complex. only Cerrajón Formation Unit-IV turbidites outcrop (Figs. 14 and 15), overlying calcareous breccias and a huge olistolith (“Peña de Martos”, Fig. 3J) which were probably emplaced during the Late-Hauterivian. The Cerrajón Formation Unit-IV turbidites buried the Peña de Martos olistolith (Fig. 15B) and reached more distal areas located, at present, more than 45 km to the west (Carcabuey outcrops, Fig. 1 B) and south (Campillo de Arenas, Fig. 1B and D). In the Los Villares–Martos outcrops, outer fan facies, fan fringe and basin plain turbidites are present (RuizOrtiz, 1980) (Fig. 3G to I). The whole group is envisaged as a basinal lobe (Mutti et al., 1996, 1999, 2003) or “sheet system” (Pickering et al., 1989) in which it is difficult to carry out a detailed differentiation of outer fan from basin plain deposits. In other studies (e.g. Remacha and Fernández, 2003), these sheetlike lobes and basin plain deposits have been geomorphically classified as a basin plain, which comprises both lobe and basin plain facies associations. The relative dispersion of the palaeocurrents with respect to those measured at the La Guardia outcrops (base-of-slope environments, Fig. 9), supports such an interpretation. The presence of calcareous breccias in the westerly Martos section and a huge olistolith which was only buried by the Cerrajón Formation Unit-IV turbidites (Fig. 15), favours the interpretation of the Los Villares– Martos area from a palaeogeographic point of view as a distally restricted (fault-bounded?) subsiding basinal zone (Fig. 15B). This subsiding part of the basin would have remained as a major depocentre up to the middlelate Albian transition age, when a huge volume of turbidite deposits compensated the submarine relief and the turbidite flows reached the western and southernmost areas (Carcabuey and Campillo de Arenas, respectively, Fig. 15A) throughout the Early Cretaceous. A wide nonrestricted basinal area about 150 km long and at least 30 km wide was then the site of turbidite sedimentation. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 5. The Early Cretaceous turbidite complex of the Cerrajón Fm The studied outcrops show the record of Early Cretaceous turbidite sedimentation in a shelf margin– slope–basin transition setting. The present alignment of the studied outcrops (≈ 285°) (Fig. 1B) is very close to the mean palaeocurrent direction (≈ N305°E) deduced from the flute casts of the channel-fill deposits (Fig. 9). The shortening, tectonic net displacement and/or tectonic rotation of the vertical axis of the outcrops have not been considered. Although a discussion of the latter questions is beyond the scope of this article, some general results from the palaeomagnetic studies might be of interest at this point. Thrusting in the central and western Subbetic was commonly accompanied by ≈ 60° of clockwise rotation of the vertical axis (Platzman and Lowrie, 1992; Platt et al., 2003). Rotation of ≈ 60° in an anticlockwise sense of the present alignment of the studied outcrops and consequently of the main palaeocurrent direction produces an arrangement that fits very well into the palaeogeographic scheme presented in Fig. 1A. The studied outcrops are probably, at present, in a relative position very similar to their palaeogeographic location during the Early Cretaceous. Early Cretaceous continental, littoral and platform environments, now cropping out in the Prebetic Zone (Sierra del Pozo in Fig. 1D), pass successively to incised submarine canyon – base-ofslope turbidite channels – basinal turbidites (basinal lobe) sedimentary environments (Fig. 15B). Each of the turbidite system elements deduced from the studied outcrops are in a suitable position with respect to each other in a broadly east to west direction, according to classical models: platform– slope (feeding canyon)–base-of-slope–basinal lobe or sheet system (outer fan, fan fringe and basin plain) (Mutti and Ricci-Lucchi, 1972; Mutti and Normark, 1987; Pickering et al., 1989; Mutti, 1992; Normark et al., 1993; Richards et al., 1998; Mutti et al., 1999, among others) (Fig. 15). A longitudinal or axial clastic distribution pattern, that is, parallel to the margin, is therefore confirmed (RuizOrtiz, 1981; Maldonado and Ruiz-Ortiz, 1982); this is, in fact, relatively common in extensional rifted basins (Leeder and Gawthorpe, 1987; Purser and Bosence, 1998). The preservation of a relatively good alignment of the Cretaceous turbidite system elements, even though some of them crop out as part of a different tectonic unit (Huelma unit), is an important point to be taken into account in the palaeogeographic reconstructions of this part of the SICP. Fig. 15B represents an interpretative longitudinal section of the Intermediate Domain trough, showing the relative position of the turbidite system elements and the adjacent Prebetic platform. 161 In this setting, erosion and deposition related to gravity flows and en mass deposition lasted from at least the Early Hauterivian to the Late Albian, that is, about 30 Ma (Gradstein et al., 2004). Between latest Hauterivian and early Late Albian, age of the Cerrajón Formation, turbidites are recorded only in the latest Hauterivian–Barremian p.p. (≈ 4 Ma), Late Aptian (≈ 2.8 Ma; G. algerianus planktonic foraminifer Zone), and Middle-middle-late Albian transition (≈ 5 Ma) ages (Figs. 2 and 14). The numbers in brackets indicate the maximum time span according to the time scale of Gradstein et al. (2004). Large hiatuses, latest Barremian–early Late Aptian (≈ 7.5 Ma) and Late Aptian–Middle Albian p.p. (≈ 8 Ma) ages, only one hiatus embracing about 26 Ma in the Huelma outcrops, separate the turbidite record into three different pieces or units and allow us to conclude that the Cerrajón Formation is a turbidite complex in the sense of Mutti and Normark (1987) (Figs. 2 and 14). 6. Discussion: sedimentary evolution, controlling factors and regional correlation The existence of large hiatuses in the stratigraphic record of the Aptian–Albian of the Tethys realm is a common feature. Regional and local tectonics shaping basin morphology and modifying the paths of oceanic currents have been the main causes argued to explain large and widespread hiatuses such as that of the Late Aptian– Middle Albian in the Tethys. In Sicily, Bellanca et al. (2002) describe a hiatus embracing a significant portion of the Late Aptian, the entire Early Albian and part of the Middle Albian, which exactly coincides with that described in this paper. Also, Erba et al. (1999) in the Southern Alps, in Cismon (Italy), describe a hiatus that is Late Aptian p.p.–Middle Albian in age. Bersezio et al. (2002) discuss different sections of the Lombardian Basin, Southern Alps, in the north of Italy, with hiatuses embracing the Early Aptian and even all the Early Cretaceous up to the Middle Albian. In the External Zones of the Betics, there are many Early Cretaceous examples, which allows us to test the model originally stated by Bernoulli and Jenkyns (1974). Intense synsedimentary blockfaulting created many hiatuses and extreme variations in sediment thicknesses (e.g. de Gea, 2003). A dynamic palaeogeographic model with extensional tectonics favouring differential subsidence and changing sedimentary depocentres from one place of the basin to another over time, has been proposed by Nieto et al. (2001), who also stressed the importance of halokinetic processes, and other authors (Vera, 2001; Vilas et al., 2003; de Gea, 2003, among others). P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 162 The age of the Cerrajón Formation exactly coincides with that of the third subsidence interval (latest Hauterivian to early Late Albian) identified by Vilas et al. (2003) in the Prebetic Zone. These authors subdivide the latest Hauterivian to early Late Albian subsidence interval into three sequence sets, K3 to K5, bounded by regional unconformities that were induced by successive tectonic events. The ages of these three sequence sets, latest Hauterivian–Late Barremian (K3), latest Barremian–Late Aptian (K4), and Early–Middle Albian (K5), show that each of the regional unconformities bounding them coincides with the lower part of the two main hiatuses (latest Barremian–early Late Aptian and latest Aptian–early Middle Albian respectively), detected in the Cerrajón Formation (Fig. 14). Moreover, Vilas et al. (2003) show how this interval of subsidence and tectonic activity has been recognised in all the sedimentary domains of Iberia, in the Iberian basin by Vilas et al. (1983) and by Salas and Casas (1993), in the Table 1 Size and age of some ancient turbidite systems Setting and example Length Width Thickness Age and duration (Ma) (m) (km) (km) Blanca (transform margin) Brae (transtensional margin) Butano (active margin) Campos (divergent passive margin) Cengio (active margin) Chugach (active margin) Ferrelo (active margin) Gottero (active margin) Hecho (active margin) Kongsfjord (transtensional margin) Marnoso– Arenacea (active margin) Cerrajón (transtensional margin) 215 30 1000 mid. Miocene (5.2) 15–20 b10 600 Late Jurassic (10) 80 40 3000 100 130 3600 earl.–mid. Eocene (17.5) Eoc.–Oli.–Mio. (50) 6.4 4.8 170 2000 100 5000 400 1800 75 40– 100 50 175 40–50 3500 80– 160 40– 100 100– 150 20–30 1000 mid.–l ate Miocene (11.1) ≈ 150 ≈ 30 latest Hauterivian– Late Albian (30) 1500 3200 ≈1350 lat. Olig.–earl. Mioc. (17.3) Late Cretaceous (6.3) Eocene (11.7) Late Cretac.–earl. Paleoc. (38) earl.–mid. Eocene (17.5) late Precambrian Pyrenees by Berástegui et al. (1990) and in the Basque– Cantabrian basin by García-Mondéjar (1990) and by Rosales (1999). All these data and correlations suggest that the deposition of the turbidites of the Cerrajón Formation were very probably controlled by allocyclic processes, mainly tectonic processes, of regional or supra-regional scale. The onset of the deposition of the Cerrajón Formation was preceded by the deposition of the calcareous breccia and the huge “Peña de Martos” olistolith (outcrops of Martos) (Fig. 10D). On the other hand, the lowest Unit-I of the Cerrajón Formation in the La Guardia and the Los Villares outcrops shows abundant slid beds and slumps. All these features support an interpretation in which tectonics appears to be the main factor triggering turbidite deposition. Tectonic subsidence favoured a net accumulation of turbidite sediments during the latest Hauterivian–Late Barremian, Late Aptian, and Middle–middle-late Albian transition. In the same way, the genesis of the main hiatuses, latest Barremian–early Late Aptian and latest Aptian–early Middle Albian respectively, was not linked to the gravity flows erosion as the main cause; these hiatuses were probably the consequence of both the absence of tectonic subsidence and of palaeographic and palaeoceanographic changes related to regional tectonics. Such a genesis of the hiatuses would explain the absence of clear erosive surfaces, apart from the base of the canyon and the channels, as is the case of the superposition of Middle Albian deposits that overlie Upper Aptian deposits in both the La Guardia and the Los Villares–Martos outcrops. Considering the parts of the basin reached by the gravity flows throughout the Early Cretaceous, a cyclicity of long period can be deduced. Five steps can be distinguished: 1) Early Hauterivian deposition of calcarenites on the canyon floor (Huelma outcrops) and in proximal parts of the basin (La Guardia outcrops); 2) Latest Hauterivian and Barremian deposition of basically thin bedded turbidites by overbank processes, often slumped, in slope and base-of-slope environments (La Guardia outcrops) and a single, or composite, basinal turbidite lobe in the Los Villares outcrops that thins and pinchs out towards more westerly basinal settings (Martos outcrops) (Fig. 15B); 3) Late Aptian deposition of channel-fill deposits in base-of-slope settings (La Guardia outcrops) and a basinal lobe that only reached the more proximal basinal areas (Los Villares outcrops) (Fig 15B); 4) Middle Albian deposition of marls with exotic blocks intercalated filling in the canyon (Huelma outcrops), deposition of thick calcarenites as channel-fill deposits in base-of-slope settings and a P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 single, or composite, basinal lobe that reached the western part of the basin in the surroundings of a huge calcareous olistolith, the “Peña de Martos”, where it pinches out (Figs. 14 and 15B); 5) Deposition of marls with thin bedded turbidites that complete the fill of the canyon, fill all the accommodation space generated in basinal areas homogenising the sea floor relief and cover the Peña de Martos olistolith (Fig. 15B); this step represents a drastic forestepping of the depositional zone of the clastics then reaching basinal areas (Carcabuey and Campillo de Arenas, Fig. 1) placed up to 45 km more distal than the areas reached by the turbidites during previous stages. At the middle-late Albian transition the turbidite system reached is maximum extension with about 150 km from the platform edge to distal basinal areas. This size is similar to that of others ancient turbidite systems (Table 1). A correlation between the volume of the turbidite flows, the distance reached by them in basinal settings and the relative strength of tectonics and related relative sea-level fall and lowstands can be assumed (cf. Mutti, 1992). In this sense, Early Hauterivian and Late Aptian would have been times of a lower rate of tectonics/relative sea level fluctuations than Barremian and middle-late Albian in the SICP. Canyon and channel incision could be related to periods of sea-level falls and lowstand. The presence of carbonate-coated quartz grains and other allochems, as well as that of wood debris in the turbidite sandstones, would indicate that the erosion of mixed siliciclasticcarbonate platform environments supplied the sediment to the gravity flows, probably during times of sea-level falls and lowstands. The dating of the two stages of canyon incision is not precise enough to correlate them with sea-level fluctuation curves or with sequence chronostratigraphy charts. However, Upper Aptian turbidites deposited during the G. algerianus planctonic foram Zone, about 2.8 Ma (Gradstein et al., 2004), could correspond to a third depositional sequence, that is, to a single turbidite system (Mutti and Normark, 1987). The erosive base of the Late Aptian channels (outcrops of La Guardia, Figs. 8 and 14) would correspond in that case to a sea level fall (Mutti, 1985, 1992) located in the G. algerianus Zone. In the sequence chronostratigraphy charts proposed by Hardenbol et al. (1998), the sequence boundary Ap 5 is located in the G. algerianus Zone, Parahoplites melchoris ammonite Zone, which can then be correlated with the base of biochronostratigraphic Unit-II of the Cerrajón Formation. The comparison of the events recorded in the Cerrajón Formation and the sequences of Hardenbol et al. (1998) shows also other interesting correlations (Fig. 14). The onset of the deposition of the 163 Cerrajón Formation correlates with the sequence boundary Ha 7 of the quoted authors. Biochronostratigraphic Unit-I of the Cerrajón Formation extends from sequence boundary Ha 7 to sequence boundary Barr 6, which is located in the I. giraudi ammonite Zone. Six sequences are differentiated by Hardenbol et al. (1998) between the Ha 7 and Barr 6 sequence boundaries. The boundary between T. hugii and K. nicklesi ammonite Zones in the outcrops of La Guardia (see above, Section 3.1) is represented by marly limestones, which would correspond to the correlative conformity of a major sequence boundary, Barr 1, of Hardenbol et al. (1998). The lowermost part of biochronostratigraphic Unit-I of the Cerrajón Formation at the Los Villares section (Fig. 11) shows four sand-rich intervals (the “steps” under Section 3.2.3; Fig. 11) or “sandy levels” that could correspond to four lowstand system tracts of the depositional sequences Barr 1 to Barr 4 of Hardenbol et al. (1998). Finally, the base of biochronostratigraphic Unit-IV of the Cerrajón Formation could correspond to the major sequence boundary Al 7 placed at the Middle– Upper Albian boundary by the latter authors (Fig. 14). Some interesting conclusions also arise on comparing the ages of the three intervals of turbidite deposition recorded in the Cerrajón Formation and the age of the carbonate platform development described by Castro (1998) in the Prebetic Zone: Llopis Formation (Late Barremian p.p.–Early Aptian p.p.), Seguili Formation (Late Aptian p.p.–earliest Albian p.p.), and Jumilla Formation (Late Albian). Each of these three carbonate platforms developed after the deposition of the turbidites making up each of the three intervals of turbidite sedimentation recorded in the Cerrajón Formation. This could be the record of relative sea level fluctuations of low frequency (5–9 Ma), with the turbidites representing the sea level falls and lowstands and the carbonate platforms the highstands of each cycle, as is common in mixed clastic-carbonate systems (Bosence, 1998). This tentative correlation would form a basis for future research, as the apparent fitting of the age, type and nature of the deposits with theoretical models is very unlikely to be coincidental. 7. Conclusions 1. The Cerrajón Formation of the Intermediate Domain of the Subbetic was deposited between the latest Hauterivian and the middle-late Albian transition and can be defined as a turbidite complex, a first order cycle, made up of several turbidite systems. 2. The feeding canyon, base-of-slope channels and basinal turbidites (or sheet system), are found to be 164 3. 4. 5. 6. 7. 8. 9. P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 aligned in a broadly east to west direction in the studied outcrops along a strip about 45 km long. Cretaceous carbonate platform deposits occur in the Prebetic Zone to the east of the studied outcrops, which complete a fossil example with the whole spectrum of environments involved in turbidite sedimentation. The palaeocurrents measured in base-of-slope environments reveal that the turbidite currents flowed in a mean N305°E sense. After correcting for the vertical axis rotation generated during Cenozoic thrusting and folding, the resulting main palaeoflow fits well with the current palaeogeographic model for the Southern Iberian Continental Palaeomargin. The platform edge and the distal basin environments are now about 150 km away from one another in a mainly east to west direction, which means the scale of the system is similar to other ancient turbidite systems. Some events in the Cretaceous sedimentation have been biostratigraphically well characterized: these include the onset of turbidite sedimentation (P. picteti ammonite Zone, latest Hauterivian), the deposition of the Upper Aptian turbidites, which probably made up a single turbidite system (G. algerianus planktonic foraminifera Zone), and the end of the turbidite sedimentation (T. phacelosus Subzone of nannofossils, middle-late Albian transition). These main Cretaceous events deduced from the study of the Cerrajón Formation show quite a good correlation with the sequence boundaries of the European basins identified by Hardenbol et al. (1998). Turbidite sedimentation is recorded in latest Hauterivian–Barremian p.p., Late Aptian and Middle– middle-late Albian transition, and these three periods with turbidite sedimentation are separated by two main hiatuses, latest Barremian–early Late Aptian and latest Aptian–early Middle Albian in age. A very good correlation is also obtained for the age of the Cerrajón Formation, the onset and the end of the turbidite sedimentation, and even the age of its main internal hiatuses, with the sequence sets, bounded unconformities and subsidence history described by Vilas et al. (2003) in the Prebetic Zone, which from a palaeogeographic point of view represents the record of the Cretaceous platform adjacent to the turbidite sedimentation. Tectonics is envisaged as the main control of relative sea-level fluctuations, basinal subsidence and hence turbidite sedimentation in a transtensional geodynamic setting. The development of the carbonate platform in the Prebetic zone after the main stages of deposition of the Cerrajón turbidites could constitute the record of relative sea-level fluctuations of low frequency (5– 9 Ma); thus, it presents a new and interesting line of study for future research. Acknowledgements This paper has been financed by Research Projects BTE2000-1151 and CGL2005-06636-C02-01 of the Spanish administration and Working Group RNM-200 from the Junta de Andalucía. Dr. M. Company (Granada, Spain) has determined the ammonite faunas and Drs. W. P. Wright (Cardiff, U.K.) and L. Vilas (Madrid, Spain) read a former version of this paper; their comments and suggestions are much appreciated. We thank A. Carrillo and A. Piedra at the Universidad de Jaén for the sample preparation. We are grateful to Dr. A. D. Miall, editor of the journal, G. Ghibaudo and an anonymous referee for their comments and suggestions. References Aguado, R., de Gea, G.A., Ruiz-Ortiz, P.A., 1996. Datos Bioestratigráficos sobre las formaciones cretácicas del Dominio Intermedio en el corte tipo (sur de Jaén). Zonas Externas de la Cordilleras Béticas. Geogaceta 20 (1), 197–200. Bellanca, A., Erba, E., Neri, R., Premoli Silva, I., Sprovieri, M., Tremolada, F., Verga, D., 2002. Palaeoceanographic significance of the Tethyan ‘Livello Selli’ (Early Aptian) from the Hybla Formation, northwestern Sicily: biostratigraphy and high-resolution chemostratigraphic records. Palaeogeogr. Palaeoclimatol. Palaeoecol. 185, 175–196. Berástegui, X., García-Senz, J.M., Losantos, M., 1990. Tectonosedimentary evolution of the Organyà extensional basin (Central South Pyrenean Unit, Spain) during the Lower Cretaceous. Bull. Soc. Geol. Fr. 8 (VI, 2), 251–264. Bernoulli, D., Jenkyns, H.C., 1974. Alpine, Mediterranean and central Atlantic Mesozoic facies in relation to the early evolution of the Tethys. In: Dott, J.R.H., Shaver, R.H. (Eds.), Modern and Ancient Geosynclinal Sedimentation. Special Publication, vol. 19. Society of Economic Paleontologists and Mineralogists, pp. 129–160. Bersezio, R., Erba, E., Gorza, M., Riva, A., 2002. Berriasian–Aptian black shales of the Maiolica Formation (Lombardian Basin, Southern Alps, Northern Italy): local to global events. Palaeogeogr. Palaeoclimatol. Palaeoecol. 180, 253–275. Bosence, D.W.J., 1998. Stratigraphic and sedimentological models of rift basins. In: Purser, B.H., Bosence, D.W.J. (Eds.), Sedimentation and Tectonic in rift basins. Red Sea–Gulf of Aden. Chapman and Hall, London, pp. 9–25. Castro, J.M., 1998. Las plataformas carbonatadas del Valanginiense superior–Albiense superior en el Prebético de Alicante. Tesis Doctoral, Univ. de Granada. 464 pp. de Gea, G.A., 2003. Bioestratigrafía de eventos en el Cretácico Inferior de las Zonas Externas de la Cordillera Bética. Tesis Doctoral. Univ. de Jaén. 687 pp. de Gea, G.A., Aguado, R., Ruiz-Ortiz, P.A., 1998. Canyon-like erosive feature in the Lower Cretaceous of the External Zones of the Betic Cordillera. Huelma Unit (Jaén, SE Spain). In: Cañaveras, J.C., P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 García del Cura, M.A., Soria, J. (Eds.), 15th International Sedimentological Congress, Alicante, p. 286. Abstracts. de Gea, G.A., Aguado, R., Ruiz-Ortiz, P.A., 2000. Diferencias de facies y espesores en secciones vecinas del Cretácico Inferior del Dominio Intermedio Subbético (Ermita de Cuadros, provincia de Jaén). Geotemas 1 (2), 177–181. de Gea, G.A., Buscalioni, A.D., Aguado, R., Ruiz-Ortiz, P.A., 2001. Restos fósiles de un cocodrilo marino en el Berriasiense inferior de la Unidad Intermedia del Cárceles-Carluco. Cordilleras Béticas, Sur de Bedmar (Jaén). Geotemas 3 (2), 201–203. Erba, E., Channell, J.E.T., Claps, M., Jones, C., Larson, R., Opdyke, B., Premoli-Silva, I., Riva, A., Salvani, G., Torricelli, S., 1999. Integrated stratigraphy of the Cismon Apticore (Southern Alps, Italy): “Reference section” for the Lower Cretaceous at low latitudes. J. Foraminiferal Res. 29, 371–391. García-Hernández, M., López-Garrido, A.C., Rivas, P., Sanz de Galdeano, C., Vera, J.A., 1980. Mesozoic paleogeographic evolution of the External Zones of the Betic Cordillera. Geol. Mijnb. 59, 155–168. García-Hernández, M., López-Garrido, A.C., Martín-Algarra, A., Molina, J.M., Ruiz-Ortiz, P.A., Vera, J.A., 1989. Las discontinuidades mayores del Jurásico de las Zonas Externas de las Cordilleras Béticas: análisis e interpretación de los ciclos sedimentarios. Cuad. Geol. Iber. 13, 35–52. García-Mondéjar, J., 1990. The Aptian–Albian carbonate episode of the Basque–Cantabrian Basin (northern Spain): general characteristics, controls and evolution. In: Tucker, M.E., Wilson, J.L., Crevello, P.D., Sarg, R.J., Read, J.F. (Eds.), Carbonate Platforms: Facies, Sequences and Evolution. International Association of Sedimentologists Special Publication, vol. 9, pp. 257–290. Gawthorpe, R.L., Leeder, M.R., 2000. Tectono-sedimentary evolution of active extensional basins. Basin Res. 12, 195–218. Gradstein, F.M., Ogg, J., Smith, A., 2004. A geologic Time Scale. Cambridge University Press, Cambridge. 589 pp. Hardenbol, J., Thierry, J., Farley, M.B., Jacquin, Th., de Graciansky, P.C., Vail, P.R., 1998. Mesozoic and Cenozoic sequence chronostratigraphic chart. In: de Graciansky, P.C., Hardenbol, J., Jacquin, Th., Vail, P.R. (Eds.), Mesozoic and Cenozoic Sequence Stratigraphy of European Basins, Special Publication, vol. 60. Society for Sedimentary Geology. Chart n° 1. Leeder, M.R., Gawthorpe, R.L., 1987. Sedimentary models for extensional tilt-block/half-graben basins. In: Coward, M.P., Dewey, J.F., Hancock, P.L. (Eds.), Continental Extensional Tectonics. Geol. Soc. London Spec. Publ., vol. 28, pp. 139–152. Maldonado, A., Ruiz-Ortiz, P.A., 1982. Modelos de sedimentación turbidítica antiguos y modernos: La Fm. Cerrajón (Cretácico inferior, Cordilleras Béticas) y los Abanicos Submarinos del Mediterráneo noroccidental. Cuad. Geol. Iber. 8, 499–526. Martínez del Olmo, W., 2003. La plataforma Cretácica del Prebético y su falta de continuidad por el Margen Sudibérico. J. Iber. Geol. 29, 111–133 (Spec. issue). Mutti, E., 1977. Distinctive thin-bedded turbidite facies and related depositional environments in the Eocene Hecho Group (Southcentral Pyrenees, Spain). Sedimentology 24, 107–131. Mutti, E., 1985. Turbidite systems and their relations to depositional sequences. In: Zuffa, G.G. (Ed.), Provenance of Arenites. NATOASI Series. Reidel Publishing Company, Dordrecht, pp. 65–93. Mutti, E., 1992. Turbidites Sandstones. AGIP — Ist. Geol. University Parma, San Donato Milanese. 275 pp. Mutti, E., Normark, W.R., 1987. Comparing examples of modern and ancient turbidite systems: problem and concept. In: Legget, J.K., Zuffa, G.G. (Eds.), Marine Clastic Sedimentology. Concepts and Case Studies. Graham and Trotman, London, pp. 1–38. 165 Mutti, E., Ricci-Lucchi, F., 1972. Le torbiditi dell' Appennino settentrionale: introduzione all'analisi di facies. Mem. Soc. Geogr. Ital. 11, 161–199. Mutti, E., Davoli, G., Tinterri, R., Zavala, C., 1996. The importance of fluvio-deltaic systems dominated by catastrophic flooding in tectonically active basins. Mem. Sci. Geol. 48, 233–291. Mutti, E., Tinterri, R., Remacha, E., Mavilla, N., Angella, S., Fava, L., 1999. An introduction to the analysis of ancient turbidite basins from an outcrop perspective. Am. Assoc. Pet. Geol., (Cont. Educ. Course Note Series) 39 97 pp. Mutti, E., Tinterri, R., Benevelli, G., di Biase, D., Cavanna, G., 2003. Deltaic, mixed and turbidite sedimentation of ancient foreland basins. Mar. Pet. Geol. 20, 733–755. Nieto, L.M., de Gea, G.A., Aguado, R., Molina, J.M., Ruiz-Ortiz, P.A., 2001. Procesos sedimentarios y tectónicos en el tránsito Jurásico/ Cretácico: precisions bioestratigráficas (Unidad del Ventisquero, Zona Subbética). Rev. Soc. Geol. Esp. 14, 35–46. Normark, W.R., Posamentier, H., Mutti, E., 1993. Turbidite systems: state of the art and future directions. Rev. Geophys. 31, 91–116. Pickering, K.T., 1982. Middle-fan deposits from the late precambrian kongsfjord formation submarine fan, Northeast Finnmark, Northern Norway. Sediment. Geol. 33, 79–110. Pickering, K.T., Hiscott, R.M., Hein, F.J., 1989. Deep Marine Environments. Clastic Sedimentation and Tectonics. Unwin Hyman, London. 416 pp. Platt, J.P., Allerton, S., Kirker, A., Mandeville, C., Mayfield, A., Platzman, E.S., Rimi, A., 2003. The ultimate arc: differential displacement, oroclinal bending and vertical axis rotation in the external betic-rif arc. Tectonics 22 (3), 1017, doi:10.1029/2001TC001321. Platzman, E., Lowrie, W., 1992. Paleomagnetic evidence for rotation of the Iberian Peninsula and the external Betic Cordillera, Southern Spain. Earth Planet. Sci. Lett. 108, 45–60. Purser, B.H., Bosence, D.W.J., 1998. Sedimentation and Tectonic in Rift Basins. Red Sea-Gulf of Aden. Chapman and Hall, London. 663 pp. Purser, B.H., Barrier, P., Montenat, C., Orszag-Sperber, F., Ott d'Estevou, P., Plaziat, J.C., Philobbos, E., 1998. Carbonate and siliciclastic sedimentation in an active tectonic setting: Miocene of the northwestern Red Sea rift, Egypt. In: Purser, B.H., Bosence, D.W.J. (Eds.), Sedimentation and Tectonic in rift basins. Red Sea-Gulf of Aden. Chapman and Hall, London, pp. 239–270. Ravnas, R., Steel, R.J., 1997. Contrasting styles of Late Jurassic synrift turbidite sedimentation: a comparative study of the Magnus and Oseberg areas, northern North Sea. Mar. Pet. Geol. 14, 417–449. Ravnas, R., Steel, R.J., 1998. Architecture of marine rift-basin successions. Am. Assoc. Pet. Geol. Bull. 82, 110–146. Remacha, E., Fernández, L.P., 2003. High-resolution correlation patterns in the turbidite systems of the Hecho Group (South-Central Pyrenees, Spain). Mar. Pet. Geol. 20, 711–726. Richards, M., Bowman, M., Reading, H., 1998. Submarine-fan systems I: characterization and stratigraphic prediction. Mar. Pet. Geol. 15, 689–717. Roldán-García, F.J., Ruiz-Ortiz, P.A., Molina Cámara, J.M., 1988. Mapa geológico y memoria de la Hoja 967 (Baena), escala 1:50.000. ITGE, Madrid, 56 pp. Rosales, I., 1999. Controls on carbonate platform evolution on active fault blocks: the Lower Cretaceous Castro Urdiales platform (Aptian-Albian, northern Spain). J. Sediment. Res. 69, 447–465. Ruiz-Ortiz, P.A., 1980. Análisis de facies del Mesozoico de las Unidades Intermedias. (Entre Castril, prov. de Granada y Jaén). Tesis Doctoral, Univ. de Granada. 274 pp. 166 P.A. Ruiz-Ortiz et al. / Sedimentary Geology 192 (2006) 141–166 Ruiz-Ortiz, P.A., 1981. Sedimentación turbidítica del Cretácico de las Unidades Intermedias. Programa Internacional de Correlación Geológica (P.I.C.G.) Real Acad. Cienc. Exactas, Fis. Nat. 2, 261–280. Ruiz-Ortiz, P.A., Molina, J.M., Nieto, L.M., 1996. Turbiditas calcáreas y otros fenómenos de resedimentación/erosión en el Jurásico superior-Cretácico inferior de la unidad de Huelma (Jaén). Zonas Externas de las Cordilleras Béticas. Geogaceta 20 (2), 323–326. Ruiz-Ortiz, P.A., de Gea, G.A., Aguado, R., 2001a. Cañón submarino del Cretácico Inferior en un área de pendiente del Paleomargen Sudibérico. Unidad de Huelma (Jaén). Rev. Soc. Geol. Esp. 14 (3–4), 175–188. Ruiz-Ortiz, P.A., Molina, J.M., Nieto, L.M., Castro, J.M., de Gea, G. A., 2001b. Introducción al Mesozoico de la parte externa del paleomargen sudibérico. Cordillera Bética. In: Ruiz-Ortiz, P.A., Molina, J.M., Nieto, L.M., Castro, J.M., de Gea, G.A. (Eds.), Itinerarios Geológicos por el Mesozoico de la Provincia de Jaén. Departamento de Geología, Univ. de Jaén, pp. 13–23. Ruiz-Ortiz, P.A., de Gea, G.A., Aguado, R., 2001c. Sedimentación clástica submarina asociada a escarpes de falla en bloques basculados. Barremiense-Aptiense basal. Cordilleras Béticas. Sur de Bedmar (Jaén). Geotemas 3 (2), 65–69. Salas, R., Casas, A., 1993. Mesozoic extensional tectonics, stratigraphy and crustal evolution during the alpine cycle of the eastern Iberian Basin. Tectonophysics 228, 33–55. Shanmugam, G., Moiola, R.J., 1988. Submarine fans: Characteristics, models, classification, and reservoir potential. Earth Sci. Rev. 24, 383–428. Sanz de Galdeano, C., 2003. Presencia de estructuras oblicuas en el sector central del Subbético y significado de la falla de Tíscar (Cordillera Bética). Rev. Soc. Geol. Esp. 16 (1–2), 103–110. Vera, J.A., 1983. Las Zonas Externas de las Cordilleras Béticas. Geología de España. Libro Homenaje a J.M. Ríos, vol. 2. IGME, Madrid, pp. 218–251. Vera, J.A., 2001. Evolution of the Southern Iberian Continental Margin. In: Ziegler, P.A., Cavazza, W., Robertson, A.H.F., Crasquin-Soleau, S. (Eds.), Peri-Tethyan Rift/Wrench Basins and Passive Margins. Mem. Mus. Natl. Hist. Nat. Paris, vol. 186, pp. 109–143. Vilas, L., 2001. El episodio extensional del Cretácico Inferior en el Prebético del altiplano de Jumilla–Yecla (Murcia). Geotemas 3 (1), 25–30. Vilas, L., Querol, R., 1999. El límite septentrional de la extensión prebética en el sector de Murcia. Libro Homenaje a José Ramírez del Pozo. A.G.G.E.P., Madrid, pp. 219–226. Vilas, L., Alonso, A., Arias, C., Garcia, A., Mas, J.R., Rincon, R., Melendez, M.N., 1983. The Cretaceous of the southwestern Iberian Ranges (Spain). Zitteliana 10, 245–254. Vilas, L., Martín-Chivelet, J., Arias, C., 2003. Integration of subsidence and sequence stratigraphic analyses in the Cretaceous carbonate platforms of the Prebetic (Jumilla–Yecla Region), Spain. Palaeogeogr. Palaeoclimatol. Palaeoecol. 200, 107–129. Weimer, P., Link, M.H., 1991. Seismic Facies and Sedimentary Processes of Submarine Fans and Turbidite Systems. SpringerVerlag, New York. 447 pp.