Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic
The Geological Society of London Books Editorial Committee Chief Editor
BOB PANKHURST (UK) Society Books Editors
JOHN GREGORY (UK) JIM GRIFFITHS (UK) JOHN HOWE (UK) RICK LAW (USA) PHIL LEAT (UK) NICK ROBINS (UK) RANDELL STEPHENSON (UK) Society Books Advisors
MIKE BROWN (USA) ERIC BUFFETAUT (FRANCE ) JONATHAN CRAIG (ITALY ) RETO GIERE´ (GERMANY ) TOM MC CANN (GERMANY ) DOUG STEAD (CANADA ) MAARTEN DE WIT (SOUTH AFRICA )
Geological Society books refereeing procedures The Society makes every effort to ensure that the scientific and production quality of its books matches that of its journals. Since 1997, all book proposals have been refereed by specialist reviewers as well as by the Society’s Books Editorial Committee. If the referees identify weaknesses in the proposal, these must be addressed before the proposal is accepted. Once the book is accepted, the Society Book Editors ensure that the volume editors follow strict guidelines on refereeing and quality control. We insist that individual papers can only be accepted after satisfactory review by two independent referees. The questions on the review forms are similar to those for Journal of the Geological Society. The referees’ forms and comments must be available to the Society’s Book Editors on request. Although many of the books result from meetings, the editors are expected to commission papers that were not presented at the meeting to ensure that the book provides a balanced coverage of the subject. Being accepted for presentation at the meeting does not guarantee inclusion in the book. More information about submitting a proposal and producing a book for the Society can be found on its website: www.geolsoc.org.uk.
It is recommended that reference to all or part of this book should be made in one of the following ways: HOMBERG , C. & BACHMANN , M. (eds) 2010. Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341. GARDOSH , M. A., GARFUNKEL , Z., DRUCKMAN , Y. & BUCHBINDER , B. Tethyan rifting in the Levant Region and its role in Early Mesozoic crustal evolution, In: HOMBERG , C. & BACHMANN , M. (eds) 2010. Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 9–36.
GEOLOGICAL SOCIETY SPECIAL PUBLICATION NO. 341
Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic
EDITED BY
C. HOMBERG University Pierre et Marie Curie, France
and M. BACHMANN University of Bremen, Germany
2010 Published by The Geological Society London
THE GEOLOGICAL SOCIETY The Geological Society of London (GSL) was founded in 1807. It is the oldest national geological society in the world and the largest in Europe. It was incorporated under Royal Charter in 1825 and is Registered Charity 210161. The Society is the UK national learned and professional society for geology with a worldwide Fellowship (FGS) of over 10000. The Society has the power to confer Chartered status on suitably qualified Fellows, and about 2000 of the Fellowship carry the title (CGeol). Chartered Geologists may also obtain the equivalent European title, European Geologist (EurGeol). One fifth of the Society’s fellowship resides outside the UK. To find out more about the Society, log on to www.geolsoc.org.uk. The Geological Society Publishing House (Bath, UK) produces the Society’s international journals and books, and acts as European distributor for selected publications of the American Association of Petroleum Geologists (AAPG), the Indonesian Petroleum Association (IPA), the Geological Society of America (GSA), the Society for Sedimentary Geology (SEPM) and the Geologists’ Association (GA). Joint marketing agreements ensure that GSL Fellows may purchase these societies’ publications at a discount. The Society’s online bookshop (accessible from www.geolsoc.org.uk) offers secure book purchasing with your credit or debit card. To find out about joining the Society and benefiting from substantial discounts on publications of GSL and other societies worldwide, consult www.geolsoc.org.uk, or contact the Fellowship Department at: The Geological Society, Burlington House, Piccadilly, London W1J 0BG: Tel. þ44 (0)20 7434 9944; Fax þ44 (0)20 7439 8975; E-mail:
[email protected]. For information about the Society’s meetings, consult Events on www.geolsoc.org.uk. To find out more about the Society’s Corporate Affiliates Scheme, write to
[email protected]. Published by The Geological Society from: The Geological Society Publishing House, Unit 7, Brassmill Enterprise Centre, Brassmill Lane, Bath BA1 3JN, UK (Orders: Tel. þ44 (0)1225 445046, Fax þ44 (0)1225 442836) Online bookshop: www.geolsoc.org.uk/bookshop The publishers make no representation, express or implied, with regard to the accuracy of the information contained in this book and cannot accept any legal responsibility for any errors or omissions that may be made. # The Geological Society of London 2010. All rights reserved. No reproduction, copy or transmission of this publication may be made without written permission. No paragraph of this publication may be reproduced, copied or transmitted save with the provisions of The Copyright Licensing Agency Ltd, Saffron House, 6 –10 Kirby Street, London EC1N 8TS, UK. Users registered with the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, USA: the item-fee code for this publication is 0305-8719/10/$15.00. British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library. ISBN 978-1-86239-306-6 Typeset by Techset Composition Ltd, Salisbury, UK Printed by MPG Books Ltd, Cornwall Distributors North America For trade and institutional orders: The Geological Society, c/o AIDC, 82 Winter Sport Lane, Williston, VT 05495, USA Orders: Tel. þ1 800-972-9892 Fax þ1 802-864-7626 E-mail:
[email protected] For individual and corporate orders: AAPG Bookstore, PO Box 979, Tulsa, OK 74101-0979, USA Orders: Tel. þ1 918-584-2555 Fax þ1 918-560-2652 E-mail:
[email protected] Website: http://bookstore.aapg.org India Affiliated East-West Press Private Ltd, Marketing Division, G-1/16 Ansari Road, Darya Ganj, New Delhi 110 002, India Orders: Tel. þ91 11 2327-9113/2326-4180 Fax þ91 11 2326-0538 E-mail:
[email protected] Preface This volume belongs to a series of four regional Geological Society of London Special Publications presenting the results of the Middle East Basin Evolution (MEBE) Programme working groups and associated research teams. MEBE was a 4year consortium (2003– 2006) funded by several oil companies (BP, ENI, PETRONAS, SHELL, and TOTAL), the French Research Organization INSU-CNRS, as well as the French University Pierre and Marie Curie (UPMC). This programme was a multi-disciplinary study of the Middle East, spanning the Arabian –Peri-Arabian and Caucasian –Caspian areas and a detailed presentation can be found in the Preface of a companion volume (Brunet et al. 2009). The four MEBE volumes cover the Black Sea –Caucasus (Sosson et al. 2010), the South Caspian –Central Iran (Brunet et al. 2009), the Zagros –East Arabian
margin (Leturmy & Robin 2010), and the Levant (this volume). The Levant volume combines new data coming from several working groups. In 2005, MEBE workers joined colleagues from various institutes and countries leading ongoing research in the Levant. Several workshops occurred and the last one was held in the University Pierre and Marie Curie, Paris, on 14– 15 December, 2006. This volume stems from scientific presentations and active discussions arose during this last Levant Meeting. The editors would like to thank colleagues who choose to publish their data in this volume, as well as others who kindly helped us to review the papers submitted to the volume. The following companies are thanked for funding several research projects under the frame of the MEBE Programme and whose results are shown here.
References BRUNET , M.-F., WILMSEN , M. & GRANATH , J. W. (eds) 2009. South Caspian to Central Iran Basins. Geological Society, London, Special Publications, 312. LETURMY , P. & ROBIN , C. (eds) 2010. Tectonic and Stratigraphic Evolution of Zagros and Makran during the Meso– Cenozoic. Geological Society, London, Special Publications, 330.
SOSSON , M., KAYMAKCI , N., STEPHENSON , R. A., STRAROSTENKO , V. & BERGERAT , F. (eds) 2010. Sedimentary Basin Tectonics from the Black Sea and Caucasus to the Arabian Platform. Geological Society, London, Special Publications, 340.
CATHERINE HOMBERG & MARTINA BACHMANN
Contents Preface
vii
HOMBERG , C. & BACHMANN , M. Evolution of the Levant margin and western Arabia platform since the Mesozoic: introduction
1
GARDOSH , M. A., GARFUNKEL , Z., DRUCKMAN , Y. & BUCHBINDER , B. Tethyan rifting in the Levant Region and its role in Early Mesozoic crustal evolution
9
MOUSTAFA , A. R. Structural setting and tectonic evolution of North Sinai folds, Egypt
37
YOUSEF , M., MOUSTAFA , A. R. & SHANN , M. Structural setting and tectonic evolution of offshore North Sinai, Egypt
65
CAME´ RA , L., RIBODETTI , A. & MASCLE , J. Deep structures and seismic stratigraphy of the Egyptian continental margin from multichannel seismic data
85
BACHMANN , M., KUSS , J. & LEHMANN , J. Controls and evolution of facies patterns in the Upper Barremian–Albian Levant Platform in North Sinai and North Israel
99
FRANK , R., BUCHBINDER , B. & BENJAMINI , C. The mid-Cretaceous carbonate system of northern Israel: facies evolution, tectono-sedimentary configuration and global control on the central Levant margin of the Arabian Plate
133
WENDLER , J. E., LEHMANN , J. & KUSS , J. Orbital time scale, intra-platform basin correlation, carbon isotope stratigraphy and sea-level history of the Cenomanian–Turonian Eastern Levant platform, Jordan
171
MORSI , A.-M. M. & WENDLER , J. E. Biostratigraphy, palaeoecology and palaeogeography of the Middle Cenomanian– Early Turonian Levant Platform in Central Jordan based on ostracods
187
JOSEPH -HAI , N., EYAL , Y. & WEINBERGER , R. Mesoscale folds and faults along a flank of a Syrian Arc monocline, discordant to the monocline trend
211
COLLIN , P.-Y., MANCINELLI , A., CHIOCCHINI , M., MROUEH , M., HAMDAM , W. & HIGAZI , F. Middle and Upper Jurassic stratigraphy and sedimentary evolution of Lebanon (Levantine margin): palaeoenvironmental and geodynamic implications
227
HOMBERG , C., BARRIER , E., MROUEH , M., MULLER , C., HAMDAN , W. & HIGAZI , F. Tectonic evolution of the central Levant domain (Lebanon) since Mesozoic time
245
HENRY , B., HOMBERG , C., MROUEH , M., HAMDAN , W. & HIGAZI , F. Rotations in Lebanon inferred from new palaeomagnetic data and implications for the evolution of the Dead Sea Transform system
269
MU¨ LLER , C., HIGAZI , F., HAMDAN , W. & MROUEH , M. Revised stratigraphy of the Upper Cretaceous and Cenozoic series of Lebanon based on nannofossils
287
AL ABDALLA , A., BARRIER , E., MATAR , A. & MULLER , C. Late Cretaceous to Cenozoic tectonic evolution of the NW Arabian platform in NW Syria
305
Index
329
Evolution of the Levant margin and western Arabia platform since the Mesozoic: introduction CATHERINE HOMBERG1* & MARTINA BACHMANN2 1
University Pierre et Marie Curie, ISTEP, Case 129, 4 Place Jussieu, 75252 Paris Cedex 05, France
2
University of Bremen, Department of Geosciences, P.O. Box 330440, 28334 Bremen, Germany *Corresponding author (e-mail:
[email protected])
Abstract: The Levant area comprises the offshore Levant Basin (LB) (eastern corner of the Eastern Mediterranean) as well as the adjacent continental slopes and platforms of the African and Arabian plates. This area experienced major events of the geodynamical evolution of the Middle East, such as the Late Palaeozoic– Early Mesozoic Pangea break up, the Late Cretaceous – Cenozoic closure of the Neo-Tethys and individualization of the Arabian plate, as well as a set of external factors like global sea-level and climate changes. This volume combines original data from the offshore and onshore Levant in various fields, including sedimentology, palaeontology, sequence stratigraphy, geochemistry, structural geology, stress reconstitution and geophysics (seismic lines, palaeomagnetism). All together, these multidisciplinary approaches allow the review of the development of the LB and gain a better insight on the later geological history and deformation processes of the Levant provinces.
The Levant describes a composite area that includes (1) the offshore Levant Basin (LB) (eastern corner of the Eastern Mediterranean) in the west, (2) the Afro –Arabian continental slope and platform, which is today largely emerged along the eastern Mediterranean coast of Egypt, Israel, Lebanon and Syria, and in Jordan and disrupted by the Dead Sea Fault and (3) the Cyprian Arc and NW Syria collision zones that mark the southern boundary of the Eurasian plate (Fig. 1). During the last decades, acquisition of geophysical and geological data allowed imaging of the crustal structures and the sedimentary infill of the LB. Together with detailed tectonic and sedimentary field investigations in the onshore Levant, these studies discover the basin forming and filling processes. Modern interpretations indicate that the LB resulted from rifting, but much controversy exists on several aspects, such as the age and kinematics of its opening or the nature of its crust. The Late Cretaceous to Present-day period includes several discrete, more or less widespread, tectonic events in relation with the interaction of the Arabian, African and Eurasian plates that resulted in the division of the Levant into several provinces. Depending on the areas, tectonic structures and timing of deformation are more or less constrained. The sedimentation patterns observed in the Meso-Cenozoic sequences of the LB and
surroundings correlate in a high degree with the tectonic processes. However, various further factors acted on top, such as global sea-level or climate changes and the recognition of their relative contribution to the sedimentation processes is a still not finished puzzle. This Special Publication of the Geological Society of London combines new multidisciplinary datasets acquired recently onshore and offshore Levant. A special aim of the volume is to demonstrate new concepts on the tectonic, stratigraphic, sedimentologic and environmental evolution of the LB and African and Arabian platforms. Periods considered cover both the Late Palaeozoic –Early Mesozoic phase of basin development and the Late Cretaceous to Cenozoic deformation period associated with the closure of Neo-Tethys and the incipience of the Dead Sea Transform.
Middle East Basin Evolution Levant Group and Others This special volume contains new data coming from several working groups. From 2003 to 2006, the Middle East Basin Evolution (MEBE) Programme funded multidisciplinary studies in the Middle East, some of them devoted to the Levant. After field data acquisitions, a working group, the
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 1 –8. DOI: 10.1144/SP341.1 0305-8719/10/$15.00 # The Geological Society of London 2010.
2
C. HOMBERG & M. BACHMANN
Fig. 1. Situation of the Levant and surroundings. Digital topography model SRTM30 from NASA Shuttle Radar Topography Mission, 30 arc-second grids mosaic (http://www.jpl.nasa.gov.srt). Numbers indicate approximate location of areas investigated in this volume. 1, Gardosh et al.; 2, Moustafa; 3, Yousef et al.; 4, Camera et al.; 5, Bachmann et al.; 6, Frank et al.; 7, Wendler et al.; 8, Morsi & Wendler; 9, Hai et al.; 10, Collin et al.; 11, Homberg et al.; 12, Henry et al.; 13, Mu¨ller et al.; 14, Al Abdalla et al. Inset shows geodynamical frame. Ar, Eu and Nu: Arabia, Eurasia, Nubia plates. DSF and EAF: Dead Sea Fault and East Anatolian Fault. RS, CA and CZ: Red Sea rift, Cyprian Arc and Arabia– Eurasia collision zone.
‘Levant Group’ led by C. Homberg, Universite´ Pierre et Marie Curie, was created in which MEBE workers joined colleagues from various institutes and countries leading ongoing research in the Levant. Several workshops occurred and the last one was held in the University Pierre and Marie Curie, Paris, on 14 –15 December, 2006. This volume stems from scientific presentations and active discussions that arose during this last Levant meeting.
General evolution of the Levant It is generally accepted that the LB formed as a result of rifting during Early Mesozoic time, which started perhaps in the Late Palaeozoic period. This is supported by crustal thinning from 30 –35 km on the African and Arabian continents to c. 8 km below the LB overlain by a 10–14 km thick sedimentary pile of Mesozoic (and Permian?) to present age (Makris et al. 1983; Ginzburg &
EVOLUTION OF THE LEVANT MARGIN AND WESTERN ARABIA PLATFORM
Ben-Avraham 1987; Vidal et al. 2000; BenAvraham et al. 2002). The affinity of the crust below the LB is regarded either as a highly stretched continent (Woodside 1977; Hirsch et al. 1995; Vidal et al. 2000) or as oceanic (Makris et al. 1983; Ginzburg & Ben-Avraham 1987; Ben-Avraham et al. 2002). Shallow-marine conditions prevailed in the basin during Triassic and Liasic times, later followed by the formation of the deep-marine basin (Bartov & Steinitz 1977; Garfunkel 1998). These marine conditions were interrupted by a strong period of erosion during the latest Jurassic – Neocomian time, with huge and widespread volcanic activity and deposition of terrestrial, fluvio-deltaic sediments, which became very thick in some places (e.g. Said 1971; Bartov & Steinitz 1977; Hirsch 2005a). The basin and its surroundings went through intense normal faulting during this period. The most common trend recognized in many places is NE– SW (Cohen et al. 1990; Garfunkel 1998; Vidal et al. 2000; Gardosh & Druckman 2006) but others like NW –SE structures exist as well (Garfunkel & Derin 1984; Homberg et al. 2009). Kinematic models (Dercourt et al. 1986; Stampfli & Borel 2002) predict a north–south opening of the basin and some difficulties arise in adaptation of these models with the local extensional fabrics of the basin and surroundings. Another characteristic of the area is the long-lived rift activity, which likely included several rifting pulses, and various ages for the opening of the basin have been proposed, from Triassic or Late Permian to Cretaceous (Freund et al. 1975; Dercourt et al. 1986; Garfunkel 1998; Ben-Avraham et al. 2002; Stampfli & Borel 2002). General coastal onlap and the development of shallow marine facies belts is a main feature of the Levant platform starting during the late Early Cretaceous period, overprinted by second- and third-order relative sea-level changes (Said 1971; Kuss et al. 2003; Rosenfeld & Hirsch 2005). Until the Mid-Albian, second-order sea-level changes generally did not coincide with those ones observed in other parts of the Tethys, but do in parts with those ones documented from other parts of the Arabian plate, suggesting that the southern Levant regional sea-level history was triggered by the interaction of subsidence and sediment supply, with an increasing eustatic influence since the Mid-Albian (Bachmann et al. 2003). A general flooding of the Levant continued during Cenomanian to Senonian time (e.g. Bartov & Steinitz 1977; Kuss et al. 2003). Since the Mid-Cenomanian a palaeorelief developed at several localities on the southern Levant platform and the subsidence control in relation with extensional tectonics is discussed by several authors
3
(e.g. Bauer et al. 2003; Schulze et al. 2005). Since the Senonian period sedimentation was controlled by a variety of tectonic factors. In some places, large grabens developed like in the interiors of the Arabian plate (Euphrates and Azraq grabens), while other areas underwent a strong inversion. The last one is referred to as the Syrian Arc tectonism that produced broad folds and high-angle reverse faults in southern Levant and northern Africa (Israel, Egypt and offshore of the Mediterranean coast), associated with breaks in deposition, angular unconformities and lateral thickness variations within sedimentary sequences (e.g. Bartov et al. 1980; Mimran 1984; Cohen et al. 1990; Lewy 1991; Moustafa & Khalil 1994; Lu¨ning et al. 1998; Gardosh & Druckman 2006). Although the age of faulting is still debated (see references above and Shalar 1994 for a review), there is a general consensus that the Syrian Arc tectonism includes two episodes, a main one during Senonian time and a second during Palaeogene time (possibly continuing up into Early Miocene time). The curved east–west to NE –SW belt is assumed to continue northward into Lebanon and Palmyrides. Further to the north, the Tethyan ophiolites were obducted onto the Arabian platform in Late Cretaceous time and therefore now outcrop occurs in Cyprus and NW Syria. During the Neogene, individualization of the Arabia plate affected the Levant with the development of the Dead Sea (or Levant) Fault (DSF) that runs almost parallel to the eastern Mediterranean coast with a general north– south trend and connects the Red Sea Rift in the south to the collision zone in the north. The c. 1000 km transform plate boundary includes a major NNE–SSW restraining bend in its central part (Lebanon –Palmyrides), which is regarded either as primary (Butler et al. 1998) or as resulting from a later clockwise rotation (Quennel 1984). Since recognition of the 70 – 110 km left-lateral displacement along the DSF (Quennel 1958; Freund et al. 1970), continuous progress has been made in the identification of the onshore and offshore plate tectonic structures and understanding of their relationships. These tectonic features resulted in latest Cretaceous to Neogene sedimentary processes, marked by small-scale structures and a pronounced palaeorelief with basins, sub-basins, swells and local unconformities, observable in many parts of the Levant. A great range of sediments reach from outer shelf to deltaic-fluviatile deposits (e.g. Buchbinder et al. 2005; Hirsch 2005b; Rosenfeld & Hirsch 2005; Kuss & Boukhary 2008).
Main outline of the papers Papers in this volume deal with the LB, adjacent deformed margins, and platforms (Fig. 1). They
4
C. HOMBERG & M. BACHMANN
are organized in three sections according to three main study areas, namely the offshore deep basin and its margins, the southern Levant (Egypt, Israel, Jordan), and the northern Levant (Lebanon, Syria). Papers concern one or several of the three key time-periods of the Mesozoic and Cenozoic evolution of the Levant. These periods can be described as follows: (1) rifting and development of the LB since Late Palaeozoic to Early Cretaceous time; (2) later evolution of the basin; and (3) inversion and deformation of the Levant from Late Cretaceous to Present.
Rifting and tectono-stratigraphic sequences of the LB and margins The papers by Gardosh et al., Camera et al. and Yousef et al. present new subsurface data on the southern part of the LB near the Israeli and Egyptian coasts. Gardosh et al. document normal faults onshore and offshore Israel with vertical offsets of several kilometres, bordering NE –SW and NNE– SSW grabens and horsts. These structures occupy a several hundred kilometre wide deformation zone that extends from the inner part of the Levant to the marine basin offshore Israel. Gardosh et al. also show that faulting activity in this area progressed over a period of 120 Ma and took place in three main pulses: Late Palaeozoic (Carboniferous to Permian), Middle to Late Triassic and Early to Middle Jurassic. The last event was the most intense rifting phase. The authors demonstrate an extension discrepancy between the brittle deformation in the upper crust and the amount of total crustal thinning and explain the basin evolution by a depth-dependent stretching, associated with mantle upwelling and removal of lower crustal layers, by decoupling along deep detachment faults. Moustafa describes similar Early Mesozoic rifting structures onshore in northern Sinai. In addition to a gradual northward increase in thickness of the Mesozoic and Cenozoic rocks, abrupt changes in thickness of the Mesozoic rocks (especially documented from the Jurassic) are reported from seismic profiles and boreholes data. Moustafa identifies several sub-basins in North Sinai with Jurassic thicknesses up to c. 3000 m. He underlines the role of the Sinai hinge belt that extends in an ENE– WSW direction from the northern tip of the Gulf of Suez toward the DSF and represented at this period the southern boundary of the tectonically active area to the north. Camera et al. and Yousef et al. recognize a second major rifting phase during Late Jurassic – Early Cretaceous in the light of sub-subsurface data along the Egyptian margin. Yousef et al. show a detailed study of the offshore area of North Sinai in a sector crossed by c. 6 km seismic reflection sections. They document NE-trending normal faults
bounding asymmetrical half-grabens. The syn-rift package consists predominantly of clastic – dominated successions of Early Cretaceous age, unconformably overlying the shallow marine Upper Jurassic carbonates. Seismic profiles shown by Camera et al. are located several hundred kilometres farther north from the Egyptian coast in the deep basin. Camera et al. combine different processing techniques to image the deep structures. The seismic stratigraphy of the area is interpreted in terms of three main tectono-stratigraphic units; a deep unit with low frequency discontinuous reflectors, a middle on average 6 km thick unit with highly faulted reflectors, and an upper 7 km thick and well-layered unit with continuous horizons (but locally disrupted by gravity structures). They are respectively interpreted as the crustal basement, the syn-rift package, and the post-rift sedimentary cover of the basin. A strong angular unconformity marks the end of the last rifting event in this area and an Aptian age is proposed for this end by correlation of the seismic reflectors with ODP data near Eratosthenes seamount. The continental crust is estimated to be 9 to 12 km thick in this sector of the LB and a 23–25 km depth below sea floor of the Moho is proposed. Based on field data, Collin et al., Homberg et al. and Bachmann et al. characterize a similar tectonic influence on the shallow marine sedimentation and the platform architecture in the onshore northern (Lebanon) and southern Levant (Sinai, Egypt) during Late Jurassic and Early Cretaceous times. Collin et al. present sedimentologic, facies, and biostratigraphic analyses of well exposed sedimentary rocks from Lebanon and discuss the relationships between the sedimentological and tectonic evolution during Middle to Late Jurassic time. Their stratigraphical data allow for the timing of the tectonic events. They first evidence the development of a stable epicontinental shelf, with shallow marine environment extending across large parts of the Lebanon, during Bathonian–earliest Kimmeridgian times and argue for a tectonic quiescence during this period. A thick sediment package accumulated owing to intense subsidence. A regional unconformity, regression and block faulting associated with a volcanic event was recognized and dating of sediments below and above suggests a Kimmeridgian age (Middle– Late Jurassic) for the beginning of this rifting phase. Following this, a Kimmeridgian to Tithonian (Late Jurassic) marine transgression induced the development of a new shallow marine carbonate shelf. Lateral thickness variations during this period are interpreted as resulting from the platform morphology formed by the previous block faulting. These marine successions were ended by regression, resulting in erosion and the deposition of continental sandstones of the basal Cretaceous.
EVOLUTION OF THE LEVANT MARGIN AND WESTERN ARABIA PLATFORM
Homberg et al. document syn-sedimentary WNW– ESE to WSW–ENE normal faults active from Hauterivian (maybe earlier) to Albian times in Lebanon, with offsets of up to several hundreds of metres and slipping under a NNE–SSW extension. This Early Cretaceous extensional phase largely controlled the sedimentologic architecture in Lebanon, especially concerning the Neocomian sequence thickness, which reaches its maximum in Central Lebanon whereas it is strongly reduced in northern Lebanon and replaced by lava flows. The authors propose that the WNW –ESE intra-Lebanon basin is a result of an Early Cretaceous rifting phase in the Levant. This event displays the continuation and acme of the precursor Late Jurassic movements described by Collin et al. That the rifting phase was active until the Early Aptian in northern Sinai (Egypt) is discussed by Bachmann et al. on the base of facies, stratigraphical and sequence stratigraphical data from the Barremian to Albian shallow-marine succession in northern Sinai and northern Israel. They indicate the presence of small-scale sub-basins in northern Sinai until the Early Aptian characterized by significant sediment thickness changes and local normal faults. The study of time equivalent sections in northern Israel miss significant influence of extensional tectonic, which is interpreted as a local restriction of tectonic movements during the late Early Cretaceous. Furthermore, the paper investigates the complexity of factors controlling the sedimentation on the Late Barreminan to Albian southern Levant platform. A detailed stratigraphic concept based on benthic foramifers, ammonites and chemostratigraphy is used for detailed dating of the individual processes. Besides the extensional tectonics, regional and global second-order sea-level changes are determined and dated. The driving processes for the observed sedimentation patterns are interpreted. The tectonic frame defined the start-up architecture of the southern Levant Cretaceous platform, which underwent significant changes by the interaction of tectonics, sea-level changes and sedimentation rates. The sedimentation patterns and rates have a major influence on the southern Levant platform geometry. In northern Sinai, a transition from a shallow-shelf structured by sub-basins through a homoclinal ramp into a flat topped platform is recognized, while the sections in northern Israel show a transition from a homoclinal ramp into a fringing platform.
Post-rift evolution of the basin and adjacent areas The Levant platform was influenced by a set of external factors, like climate or sea-level changes, which can be a result of global changes and orbital cycles. The relatively quiet post-rift tectonic frame
5
seemed to be an optimal phase for the investigation of those factors and their influences. The post-rift evolution of the southern Levant Platform is investigated in three papers. That tectonic quietness does not match the reality is shown by Frank et al. They focus on the Cenomanian –Turonian development of northern Israel. By detailed sedimentologic and facies analyses of several sections, they described the cyclic sedimentation patterns and the hierarchical organization of those cycles. They document proximalto-distal facies and thickness changes, and the geometry of genetic –stratigraphic sequences. These datasets are interpreted and authors show that the tectono-sedimentary regime of northern Israel represents an east– NE branch-off of the depositional strike from the north –south-striking Levant margin. Frank et al. indicate several productive and demise phases of the Cenomanian – Turonian carbonate ramp and show that the ramp development was strongly influenced by eustatical and palaeoceanographical trends, similar to those observed in the Tethys. Late Cenomanian normal faults and a latest Cenomanian sequence boundary, marked by subaerial exposure, reflect tectonic activity and uplift of the Galilee during this period, which was generally regarded as tectonically quiet. The papers of Morsi & Wendler and Wendler et al. analyse shallow marine platform successions from an intra-platform basin that developed on the southern Levant platform (central Jordan). They highlight another important event characterizing the Mesozoic history: marine sediments, rich in organic matter, occurred during several Jurassic and Cretaceous intervals. They reflect a significant disturbance in the global carbon cycle related to oceanic anoxic events (OAEs). While the effects of OAEs in pelagic to hemipelagic deposits have been studied intensively, the palaeoenvironmental conditions related to these significant events in the near-shore environments still need to be investigated in more detail. Morsi & Wendler use stratigraphic and palaeontologic investigations of ostracods to complete age dating of the midCenomanian through Lower Turonian shallowmarine succession and to analyse palaeoecology and palaeoenvironmental conditions related to OAE2 at the Cenomanian –Turonian (C –T) transition interval and to study the palaeobiogeographical configurations. A detailed taxomical and detailed palaeontologic part forms a broad base for all interpretations. Wendler et al. concentrate on the influence of orbital cycles on the sedimentation during the C –T interval and on stratigraphic aspects concerning the OAE2. A high-resolution stratigraphy is presented by the use of high-resolution calcimetry and stable carbon isotope stratigraphy. They use these methods to correlate the successions to the
6
C. HOMBERG & M. BACHMANN
global record (exemplified for the well-dated C –T boundary interval) in order to eliminate ambiguities in the local stratigraphy. They refine the sequencestratigraphic model by constructing an orbital time scale and present a model of orbitally forced sequences. Finally, they indicate that the Jordan platform demise was later than the onset of the OAE2 and discuss the role of biological factors for demise of platform carbonate producers.
Late Cretaceous to Cenozoic deformation in the Levant Moustafa and Yousef et al. discuss the inversion of the southern Levant margin that began in Late Cretaceous time. New data presented by these authors provide details on the development of the Egyptian segment of the Syrian Arc. Moustafa shows results of a detailed mapping and structural analysis of North Sinai. North Sinai comprises a folded zone dominated mainly by NE– SW doubly plunging anticlines, as well as two fault belts, the ENE– WSW Sinai hinge belt and the east– west themed fault. Faults of the area show a preferred NW to WNW orientation, but other trends, like NNW and ENE ones, also exist in the hinge belt. Mechanism of deformation in this area includes reactivation of deep faults bounding Early Mesozoic basins (see previous section) associated with folding, as well as dextral shear with more or less pronounced transpression. This major NW–SE compression started in north Egypt in Late Cretaceous, reached its acme in Campanian time and ended before Early Miocene time. Moustafa also recognizes post-Early Miocene fault reactivation in the Sinai hinge belt and proposes a genetic relation between this Cenozoic deformation and the sinistral slip on the DSF. Yousef et al. observe similar inversion structures in offshore Sinai. They are the Mango, Goliath, North Sinai and Ziv structures. They correspond to NE–SW asymmetrical folds that most commonly show a steep northwestern flank bounded by a steeply dipping fold progation fault. Most of these faults show reverse offsets at their upper ends whereas the deeper layers increase in thickness towards the main bounding fault and are offset in a normal motion (see previous section). These observations indicate that the offshore Sinai structures correspond to inverted Late Jurassic –Early Cretaceous half-grabens. Yousef et al. evidence progressive onlaps on the crest of the elevated structures and data in boreholes cutting the syncompression sequence indicate a Campanian– Maastrichtian age for the inversion. Inversion continued in a middle way until Late Oligocene. A few WSW–ESE normal faults cut the younger sequences and are thought to form in response
to gravity driven extension in Miocene and post-Miocene times. New data on the tectonic an sedimentologic evolution of southern and northern Levant, in Israel, Lebanon and Syria are presented in Al Abdalla et al., Hai et al., Henry et al., Homberg et al. and Mu¨ller et al. Al Abdalla et al. combine a brittle tectonic analysis and stratigraphic study of the NW Arabia platform. They demonstrate that an extensional context prevailed in NW Syria at Senonian time with the development of meso-scale normal faults. The authors propose that this NE– SW extension is associated with the development of the Euphrate graben to the east. Investigation of the Senonian package of Lebanon by Homberg et al. suggests that little or no major compressive deformation occurred there in Late Cretaceous time. Al Abdalla et al. and Homberg et al. also recognize syn-sedimentary normal faults in the Eocene sediments of Syria and Lebanon and fault analysis indicates that they slipped under a NNE – SSW extension. According to these observations, the tectonic evolution of northern Levant implies that the Mesozoic extensional context continued here until Late Eocene, and maybe Oligocene time, with a short interruption during Maastrichtian time when the Tethyan Ophiolites were obducted to the south onto the Arabian platform. Al Abdalla et al. shows that this event is marked in the Maastrichtian sequence of NW Syria by intra-formational unconformities. Al Abdalla et al. detail the Cenozoic tectonic structures and palaeostress evolution in NW Syria. A major phase of shortening occurred here in the Early Miocene. During this NW–SE compresion event, the Baer-Bassit thrusted over the Coastal Range platform along the SE vergent Lattakia Thrust. This major thrusting induced the flexure of the Arabian platform and the formation of the Middle to Late Miocene Lattakia Basin. From the end of the Miocene, and until Present, the region experienced a NNW–SSE directed regional compression. A coeval east –west trending compression associated with the north– south folding of the Coastal Range was also recognized. The authors suggest that it probably corresponds to a stress-field deflection in relationship with the DSF activity. The left-lateral displacement along the northern segment of the DSF since latermost Miocene is estimated to 30 –40 km. Homberg et al. present a re-evaluated tectonic history of Lebanon based on new field observations of tectonic structures. Lebanon experienced two compressional tectonic events during the Cenozoic period. A first east–west compression produced moderate folding and faulting during Early Miocene, probably as a far-field effect of the ongoing collision in the north between Afro-Arabia
EVOLUTION OF THE LEVANT MARGIN AND WESTERN ARABIA PLATFORM
and Eurasia. A second, but much more severe, folding event occurred during Late Miocene time owing to a NNW–SSE compression. The authors interpret this stress reorganization during the Neogene as the signature of the initiation of the transform tectonics in Lebanon, with a pronounced early transpresssional character. A maximum Late Langhian age of the Yammouneh Fault, the main branch of the DSF in Lebanon, is given. Henry et al. present results from a palaeomagnetical analysis in Lebanon. They document two palaeomagnetical directions in Aptian– Albian formations from widespread sites, one of them corresponding to the primary magnetization and the second being acquitted after folding. Comparison of these data with previous palaeomagnetical results for the Jurassic age in Lebanon, and expected directions from African apparent polar wander path, yields evidence of three different counter-clockwise regional rotations: 338 before Aptian deposition, 118 during Late Miocene time, and 178 in postMiocene times. The regional Cenozoic 288 counterclockwise rotation observed in Lebanon rules out the possibility that the Yammouneh fault originally trended north– south similar to the other main fault branches of the DSF. According to the authors, the origin of the deflection of the fault in Lebanon is to be found in crustal rheological constrasts within the Levant, probably formed during the Mesozoic extension. A structural model of the central segment of the DSF is proposed. In this model, about 35% of the transform plate motion was first accommodated during Late Miocene time, mostly by folding in Lebanon and Palmyrides and was then later transferred onto the Yammouneh and associated faults. The first phase is associated with 118 of the counterclockwise rotation and 178 occurred during the second phase. Mu¨ller et al. studied calcareous nannofossils from predominantly marly Senonian– Maastrichtian, Paleogene, and Neogene series from the central Levant platform in Lebanon. This lithological homogenous stratigraphic interval was poorly subdivided on the existing geological maps. Palaeocene, Upper Eocene, Upper Oligocene and Lower Miocene units are identified for the first time in Lebanon. The study presented by Mu¨ller et al. contains the precise determination of hiatus and tectonic events. They document that the relatively complete marine succession was interrupted by an extended Late Cretaceous (Coniacian and Early Santonian) hiatus and several regional hiatus at the top of the Maastrichtian to lowermost Palaeocene, the Lower Oligocene, and the lower part of the Lower Miocene (Aquitanian –lowermost Burdigalian) sequences. Hai et al. combine a fold and fault analysis in Israel along the western limb of the Ramallah
7
monocline. Studied folds in Turonian and Senonian rocks have wavelengths ranging between 25 –100 m and amplitudes between 10–30 m and show two superposed axis trends. The minority of the folds are aligned NNE–SSW and are thus compatible with the WNW –ESE shortening trend of the Syrian Arc fold belt. In the contrary, the majority of measured folds are aligned ENE– WSW and are not compatible with this shortening trend. Kinematic analysis of fault attitude indicates NNW– SSE shortening and ENE –WSW extension in accordance with the shortening of the majority of folds. Based on several arguments, authors argue that folds resulted from tectonic shortening and were not formed owing to karstic activity or other processes. They propose that folds compatible with the main trend of the Ramallah monocline are parasitic small folds within the Syrian Arc fold belt whereas the other major fold trend is associated with Miocene to Recent movement along the DSF.
References Bachmann, M., Bassiouni, M. A. A. & Kuss, J. 2003. Timing of mid-Cretaceous carbonate platform depositional cycles, northern Sinai, Egypt. Palaeogeography, Palaeoclimatology, Palaeoecology, 200(1– 4), 131– 162. Bartov, Y. & Steinitz, G. 1977. The Judea and Mount Scopus Groups in the Negev and Sinai with Trend Surface Analysis of the thickness data. Israel Journal of Earth Science, 28, 119– 148. Bartov, Y., Lewy, Z., Steinitz, G. & Zak, I. 1980. Mesozoic and Tertiary stratigraphy, paleogeography and structural history of the Gebel Areif en Naqa area, eastern Sinai. Israel Journal of Earth Science, 29, 114 –139. Bauer, J., Kuss, J. & Steuber, T. 2003. Sequence architecture and carbonate platform configuration (Late Cenomanian-Santonian), Sinai, Egypt. Sedimentology, 50, 387 –414. Ben Avraham, Z., Ginzburh, A., Makris, J. & Eppelbaum, L. 2002. Crustal structure of the Levant Basin, eastern Mediterranean. Tectonophysics, 346, 23–43. Buchbinder, B., Calvo, R. & Siman-Tov, R. 2005. The Oligocene in Israel: a marine realm with intermittent denudation accompanied by mass-flow deposition. Israel Journal of Earth Sciences, 54, 63– 85. Butler, R. W. H., Spencer, S. & Griffiths, H. M. 1998. The structural response to evolving plate kinematics during transpression: evolution of the Lebanese restraning bend of the Dead Sea Transform. In: Holdsworth, R. E., Strachan, R. A. & Dewey, J. F. (eds) Continental Transpressional and Transtensional Tectonics. Geological Society, London, Special Publications, 135, 81– 106. Cohen, Z., Kapstan, V. & Flexer, A. 1990. The tectonic mosaic of the southern Levant: implications for hydrocarbon prospects. Journal of Petroleum Geology, 13, 437– 462.
8
C. HOMBERG & M. BACHMANN
Dercourt, J., Zonenshai, L. et al. 1986. Geological evolution of the Tethys belt from the Atlantic to the Pamirs since the Lias. Tectonophysics, 123, 241–315. Freund, R., Garfunkel, Z., Zak, I., Goldberg, M., Weissbrod, T. & Derin, B. 1970. The shear along the Dead Sea Rift. Philosophical Transactions of the Royal Society, London, 267, 107–130. Freund, R., Goldberg, M., Weissbrod, T., Druckman, Y. & Derin, B. 1975. The Triassic-Jurassic structure of Israel and its relation to the origin of the Eastern Mediterranean. Bulletin Geological Survey of Israel, 65, 26. Gardosh, M. A. & Druckman, Y. 2006. Seismic stratigraphy, structure and tectonic evolution of the Levantine Basin, offshore Israel. In: Robertson, A. H. F. & Mountrakis, D. (eds) Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, 201–227. Garfunkel, Z. 1998. Constrains on the origin and history of the Eastern Mediterranean basin; collision-related processes in the Mediterranean region. Tectonophysics, 298, 5– 35. Garfunkel, Z. & Derin, B. 1984. Permian-early Mesozoic tectonism and continental margin formation in Israel and its implication for the history of the Eastern Mediterranean. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 187–201. Ginzburg, A. & Ben-Avraham, Z. 1987. The deep structure of the central and southern Levant continental margin. Annales Tectonicae, 1, 105– 115. Hirsch, F. 2005a. The Jurassic of Israel. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant. Volume II: The Levantine Basin and Israel. Jerusalem: Historical Productions-Hall, Jerusalem, 361– 391. Hirsch, F. 2005b. The Oligocene-Pliocene of Israel. In: Hall, J. K., Krasheninikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant (II): The Levantine Basin and Israel. Historical Productions, Hall Publications, Jerusalem, Israel, 459– 488. Hirsch, F., Flexer, A., Rosenfeld, A. & Yellin-Dror, A. 1995. Palinspatic, crustal studies of the eastern Mediterranean. Journal of Petroleum Geology, 18, 149– 170. Homberg, C., Barrier, E., Mroueh, M., Hamdan, W. & Higazi, F. 2009. Basin tectonics during the Early Cretaceous in the Levant margin, Lebanon. Journal of Geodynamics, 47, 218– 223. Kuss, J. & Boukhary, M. A. 2008. A new upper Oligocene marine record from northern Sinai (Egypt) and its paleogeographic context. GeoArabia, 13, 59–84. Kuss, J., Bassiouni, A., Bauer, J., Bachmann, M., Marzouk, A., Scheibner, C. & Schulze, F. 2003. Cretaceous – Paleogene sequence stratigraphy of the
Levant Platform (Egypt, Sinai, Jordan). In: Gili, E., Negra, H. & Skelton, P. (eds) North African Cretaceous Carbonate Platform Systems. Nato Science Series, 171–187. Lewy, Z. 1991. Periodicity of Cretaceous epeirogenic pulses and the disappearance of the carbonate platform facies in the Late Cretaceous times (Isr). Israel Journal of Earth Science, 40, 51–58. Lu¨ning, S., Marzouk, A. M., Morsi, M. & Kuss, J. 1998. Sequence stratigraphy of the Upper Cretaceous of central-east Siani, Egypt. Cretaceous Research, 19, 153–196. Makris, J., Ben-Avraham, Z. et al. 1983. Seimic refraction profiles between Cyprus and Israel and their interpretations. Geophysical Journal of the Royal Astronomical Society, 75, 575– 591. Mimran, Y. 1984. Unconformities on the eastern flank of the Faria anticline and their implications on the structural evolution of Samaria (central Israel). Israel Journal of Earth Science, 33, 1 –11. Moustafa, A. R. & Khalil, M. H. 1994. Structural characteristics and tectonic evolution of north Sinai fold belts. In: Said, R. (ed.) The Geology of Egypt, A.A. Balkema. Rotterdam, 381– 389. Quennel, A. M. 1958. The structure and the evolution of the Dead Sea Rift. The Quarterly Journal of the Geological Society, London, 64, 1 –24. Quennel, A. M. 1984. The western Arabia rift system. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Blackwell, Oxford, 775–788. Rosenfeld, A. & Hirsch, F. 2005. The Cretaceous of Israel. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant (II): The Levantine Basin and Israel. Historical Productions, Hall Publications, Jerusalem, Israel, 393–436. Said, R. 1971. Explanatory notes to accompany the Geological Map of Egypt. Geological Survey of Egypt, Papers, 1– 123. Schulze, F., Kuss, J. & Marzouk, A. 2005. Platform configuration, microfacies and cyclicities of the upper Albian to Turonian of west-central Jordan. Facies, 50, 505– 527. Shalar, J. 1994. The Syrian arc system: an overview. Palaeogeography, Palaeoclimatology, Palaeoecology, 112, 125 –142. Stampfli, G. M. & Borel, G. D. 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrones. Earth and Planetary Science Letters, 196, 17– 33. Vidal, N., Alvarez-Marron, J. & Klaeschen, D. 2000. Internal configuration of the Levantine Basin from seismic reflection data (Eastern Mediterranean). Earth and Planetary Science Letters, 180, 77– 89. Woodside, J. M. 1977. Tectonic elements and crust of the eastern Mediterranean Sea. Marine Geophysical Researches, 3, 317– 354.
Tethyan rifting in the Levant Region and its role in Early Mesozoic crustal evolution MICHAEL A. GARDOSH1, ZVI GARFUNKEL2*, YEHEZKEL DRUCKMAN3 & BINYAMIN BUCHBINDER3 1
The Geophysical Institute of Israel, P.O. Box 182, Lod, 71100, Israel
2
Institute of Earth Sciences, Hebrew University, JerusPlealem, Israel, 91904 3
Geological Survey of Israel, 30 Malkhe Israel St. Jerusalem, 95501 *Corresponding author (e-mail:
[email protected])
Abstract: At the time of the opening of the Tethys Ocean the northern edge of Gondwana was affected by several rifting events. In this study, we used data from deep exploration wells, seismic profiles, and seismic depth maps to reconstruct the pattern of Tethyan rifting in the Levant region and to investigate its effects on the evolution of the Levant crust. The results show a several hundred kilometre wide deformation zone, comprised of graben and horst structures that extend from the inner part of the Levant to the marine basin offshore Israel. The structures are dominated by sets of NE– SW and NNE– SSW oriented normal faults with vertical offsets in the range of 1–8 km. Rifting was associated with a NW–SE direction of extension, approximately perpendicular to the present-day Mediterranean coast. Faulting activity progressed over a period of 120 Ma and took place in three main pulses: Late Palaeozoic (Carboniferous to Permian); Middle to Late Triassic; and Early to Middle Jurassic. The last, and the most intense, tectonic phase post-dates the activity in other rifted margins of northern Gondwana. Rifting was associated with the modification and stretching of the Levant crust. Our results demonstrate an extension discrepancy between the brittle deformation in the upper crust and the amount of total crustal thinning. Seismic reflection data shows that the Levant Basin lacks the characteristics of typical rifted margins, either volcanic or non-volcanic. The evolution of the basin may be explained by depth-dependant stretching, associated with the upwelling of divergent mantle flow and removal of lower crustal layers by decoupling along deep detachment faults.
The separation of Gondwana and Eurasia and the development of the Tethys Ocean during the Late Palaeozoic to Early Mesozoic period were accompanied by continental breakup, rifting and drifting of various micro-continental blocks (Sengor & Yilmaz 1981; Robertson & Dixon 1984; Robertson 1998; Garfunkel 1998, 2002; Robertson 2007). This fragmentation of northern Gondwana, referred to here as Tethyan rifting, resulted in the formation of several marine basins and extensional margins in the Eastern Mediterranean region. Some of the basins were later consumed during the Cenozoic closure of the Tethys Ocean; the Levant continental margin and basin (Fig. 1) remained however, relatively intact and the Tethyan extensional structure in this area were therefore preserved (Garfunkel 1998; Gardosh & Druckman 2006). Structures that are associated with Tethyan rifting are recognized throughout the Levant onshore, from the Palmyra area in central Syria to the Egyptian Western Desert (Fig. 1) (Freund et al. 1975; Druckman 1984; Garfunkel & Derin 1984; Druckman et al. 1995; Garfunkel 1998; Guiraud & Bosworth 1999; Brew et al. 2001;
Sawaf et al. 2001). Tethyan structures are characterized by thickness variations, faulting and magmatism spanning the Palaeozoic to Early Mesozoic succession. This sedimentary section is typically found at great depth and, therefore, provides limited amount of geological information. Indeed, the study of Tethyan structures in the Levant region was so far based on rather small number of deep boreholes and partial coverage by land seismic data. Several deep exploration wells that were recently drilled in Israel and particularly, a new set of two-dimensional (2D), marine seismic reflection lines (Fig. 2) (Gardosh & Druckman 2006; Gardosh et al. 2006, 2008) add more details on some of the known Tethyan structures and reveal new structures that were previously unknown. Tethyan tectonic activity have been studied in various regions, located along the northern margins of the ancient super-continent of Gondwana. Robertson & Mountrakis (2006) and Robertson (2006, 2007) recently discussed the confusion in geological literature regarding the Tethyan nomenclature. Two terms are commonly used: ‘Palaeo-tethys’ and ‘Neotethys’. The former refers
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 9 –36. DOI: 10.1144/SP341.2 0305-8719/10/$15.00 # The Geological Society of London 2010.
10
M. A. GARDOSH ET AL.
Fig. 1. Main tectonic elements of the Eastern Mediterranean region shown on the background of satellite imagery. The Levant Basin is located on the northeastern edge of the African plate, south of the Cyprian Arc plate boundary (marked by thick white lines). The area of study includes the central part of the basin and its margin onshore and offshore Israel. The outline of a new, seismic reflection survey offshore Israel is marked by dashed blue line. The insert shows a regional imagery of the Eastern Mediterranean and Red Sea (discussed in the text).
to oceanic basins of the Late Palaeozoic to Early Mesozoic age and the later is generally used for an Early Mesozoic, primarily Late Triassic to Early Jurassic ocean. However, these terms are often
misused by researchers that apply the same names to different marine basins (Robertson & Mountrakis 2006). Following Robertson (2007), we apply the more generalized term ‘Tethyan rifting’ for all the
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
11
Fig. 2. The Tethyan rift system of the Levant. The map in (a) depicts inferred, Triassic to Early Jurassic faults and structural highs and lows that were formed during Tethyan rifting activity; (b) is a seismic depth map of the ‘top of the crystalline basement’ taken from Gardosh & Druckman (2006); and (c) is a Bouguer gravity map of the Levant from Rybakov & Al-Zoubi (2005). The gravity map (c) shows the transition from light, continental crust inland (negative values), to heavier crust (positive values) in the Mediterranean Sea. The faults and structures in (a) are compiled from various sources: the map in (b), and data published by Roberts & Peace (2007) and by Aal et al. (2000) for the offshore area, seismic and well data (Figs 3– 6) and seismic depth maps of Gelbermann (1995) for central Israel, data published by Moustafa & Khalil (1990) for Northern Sinai and by Walley (1998) for Lebanon. The location of the Palmyra Trough in (a) is restored to its position prior to the c. 100 km of sinistral motion along the Dead Sea Transform. The outlines of magnetic anomalies that are associated with Early Mesozoic magmatic activity in (a) are adopted from Rybakov et al. (1997). A list of well names is in Table 1. Abbreviations for structural elements in (b) and (c) are: ER, Eratosthenes High; JU, Judea Graben; PL, Pleshet Basin; NS, North Sinai Basin; JN, Jonah High; LV, Leviathan High; YM, Yam High.
Late Palaeozoic to Early Mesozoic structures of the Levant, with no reference to a specific ocean system. The main goal of this paper is to present an overview of Tethyan structures and rifting activity in the Levant region based on the integration of old and new well and seismic data. The regional structural pattern is used to reconstruct the style, direction and timing of continental breakup. As one of the notable results of Tethyan rifting activity is the modification of the Levant crust, an additional goal of this study is to examine the relations between faulting and crustal thinning as compared to other continental margins and extensional basins worldwide. The relatively large amount of geological and geophysical data makes the Levant margin a unique location for testing concepts of continental margin evolution.
The main tectonic events that shaped the Levant region The crust of the Levant was stabilized in the Late Precambrian by the Pan-African orogeny c. 1000– 550 Ma (Eyal et al. 1991; Stern 1994; Alsharhan & Nairn 1997). During the Early to Middle Palaeozoic period, the Levant area formed a wide continental platform of the northern part of Gondwana. Thick sections of predominantly siliciclastic sediments of fluvio-deltaic to shallowmarine origin were deposited on this platform south of the Tethys Ocean (Klitzsch 1981; Wolfart 1981; Weissbrod 1981, 2005; Beydoun 1988; Alsharhan & Nairn 1997; Garfunkel 2002). The Palaeozoic succession of northern Gondwana were affected by regional epiorogenical
12
M. A. GARDOSH ET AL.
movements that produced broad swells and depressions, several hundred kilometres in diameter (Garfunkel & Derin 1984; Gvirtzman & Weissbrod 1984; Alsharhan & Nairn 1997; Garfunkel 2002, 2004). The swells were partly eroded, resulting in significant stratigraphic gaps. Gvirtzman & Weissbrod (1984) and Weissbrod (2005) inferred the formation of a regional swell in the Levant area, which they called the Geanticline of Helez. These authors suggested two phases of uplift: at the beginning of the Carboniferous and in the Early Permian. The Carboniferous erosion presumably produced a concentric outcrop pattern of the older Palaeozoic units, centred on the crest of the Helez Geanticline in the southern Israeli coastal plain (Gvirtzman & Weissbrod 1984; Weissbrod 2005). Garfunkel & Derin (1984) and Garfunkel (1998) found a more complex pattern of vertical motions with a structural high extending from Jordan to the Israeli coast. They further stressed the role of the mid-Palaeozoic erosion phase, which is supported by fission track dating (Kohn et al. 1992). Some authors attributed the Palaeozoic epeirogenic movements to compression of the Arabian plate that was associated with collision and subduction in the northern margin of Gondwana (Hercynian Orogeny) (Boote et al. 1998; Guiraud & Bosworth 1999; Weissbrod 2005). Others (Garfunkel 1998, 2004) think that during the Palaeozoic period most of the northern Gondwanian margin, including the Levant, was affected by extension rather than subduction, and this allowed the detachment of Gondwanian terrains that were incorporated in the Hercynian orogenic edifice in Europe. A profound change in the pattern of vertical motions and sediment distribution that took place during the latest Palaeozoic to Early Mesozoic period reflects a new tectonic regime (Garfunkel 1998). The Permian, Triassic and Lower Jurassic rock section of the Levant is dominated by shallowmarine, mostly carbonates and less common siliciclastic rocks. Considerable localized thickness and facies variations in these strata were identified in outcrop, well and seismic data from southern and central Israel (Goldberg & Friedman 1974; Druckman 1974, 1977; Freund et al. 1975; Druckman 1984; Garfunkel & Derin 1984; Gelbermann & Kemmis 1987; Bruner 1991; Druckman et al. 1995; Gelbermann 1995; Garfunkel 1998; Gardosh & Druckman 2006). Thickness variations are also documented in the Palmyrides and other parts of Syria (Al-Youssef & Ayed 1992; Brew et al. 2001; Sawaf et al. 2001) and in northern Sinai (Alsharhan & Salah 1996; Hirsch et al. 1998). These widely spread phenomena indicate repeated differential motions and extension on the edge of the Gondwanian plate during the latest
Palaeozoic to Early Mesozoic period (Freund et al. 1975; Garfunkel & Derin 1984; Druckman et al. 1995; Garfunkel 1998; Gardosh & Druckman 2006). The extensional pulses that were accompanied by magmatic activity resulted in the separation of the Tauride, Eratosthenes, and probably other small continental blocks from the AfroArabian craton (Biju-Duval & Dercourt 1980; Sengor & Yilmaz 1981; Robertson & Dixon 1984; Le Pichon et al. 1988; Robertson et al. 1996; Garfunkel 1998), and led to the opening of an ocean north and west of the Levant margin. The Early Mesozoic rifting activity was followed by subsidence and the formation of the deep marine Levant Basin in the eastern Mediterranean area (Garfunkel & Derin 1984; ten Brink 1987; Garfunkel 1988). Starting from the Middle Jurassic, a continental shelf and slope developed along the eastern margin of the basin (Fig. 1). This continental margin slope, originally identified in deep wells near the present-day Israeli coastline (Gvirtzman & Klang 1972; Bein & Gvirtzman 1977; Garfunkel 1988), formed a narrow transition zone, termed by Bein & Gvirtzman (1977) as the ‘hinge-line’. It separated the Jurassic and Cretaceous shallow-marine shelf on the east from the deep marine basin on the west. The Mesozoic shelf-edge is identified near the modern Mediterranean coastline from northern Egypt to western Lebanon (Harms & Wray 1990; Jenkins 1990; Walley 1998, 2001; Garfunkel 1998). Late Cretaceous convergence of the Eurasian and Afro-Arabian plates resulted in the formation of a northward-dipping subduction zone at the present-day areas of Cyprus and southern Turkey (Fig. 1) (Le Pichon et al. 1988; Robertson 1998). Far-field stresses related to this subduction caused contraction of the Levant margin and generated the development of anticlines and synclines throughout the Levant region known as the ‘Syrian Arc’ (Krenkel 1924) fold belt. The contractional activity continued in the Levant area through the Cenozoic period (Eyal & Reches 1983; Walley 1998; Gardosh & Druckman 2006). Well and seismic data suggest that the Syrian Arc deformation is associated in many places with inversion of Late Palaeozoic to Early Mesozoic normal faults that were reactivated in a reverse motion (Freund et al. 1975; Druckman et al. 1995), although in some structures these relations are not clearly observed (Druckman 1981). The last, major tectonic event that affected the Levant area was the breakup of the African – Arabian Craton and the formation of the Dead Sea fault zone (Fig. 1) (Freund et al. 1970; Garfunkel 1981, 1997). This new plate boundary is, in fact, a transform fault that connects the spreading centre of the Red Sea with a collision zone in the Taurus
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
Mountains of southeastern Turkey (Fig. 1). The Dead Sea transform offset some of the older Tethyan structures. The sinistral, strike –slip motion along this fault system amounts to about 100 km (Freund et al. 1970; Garfunkel 1981).
13
and lows; and (d) the Eratosthenes High. These zones include symmetric and asymmetric grabens, horsts, and down-stepping blocks, all controlled by sets of normal faults generally trending NE –SW.
The inner basins Data sources The distribution of Tethyan rift structures in the subsurface of the Levant (Fig. 2a) is compiled from the following sources: data from deep boreholes that penetrated the Palaeozoic to Lower Jurassic section in Israel (Table 1); seismic interpretation maps of the Levant offshore (Gardosh & Druckman 2006; Gardosh et al. 2006, 2008); seismic interpretation maps of central and southern Israel (Gelbermann 1995), outcrop and well studies by Freund et al. (1975), Garfunkel & Derin (1984), Druckman et al. (1995) and Garfunkel (1998). The seismic mapping of Gardosh et al. (2006, 2008) was based on the interpretation of a highquality set of 2D seismic reflection lines acquired in 2001 (Figs 1 & 2b). These regional marine lines extend from the coastal area of Israel to a distance of 150 –200 km offshore and penetrate the subsurface down to the top of the crystalline basement at depth of 10 –15 km below mean sea level (MSL). The excellent coverage allows the interpretation and mapping of the deep part of the Levant Basin that was not imaged in older seismic datasets. Two important seismic markers were used to map Tethyan rift structures: (a) top of the crystalline basement (Fig. 2b) and (b) Middle Jurassic unconformity of an assumed Bathonian age (Gardosh et al. 2006, 2008). Three regional, geological cross-sections (Fig. 3) were constructed from the seismic depth maps (Levant Basin and central Israel) and results of 18 deep exploration boreholes located onshore and offshore Israel (Table 1). The structure in the northern part of section 3 (Fig. 3c) is considered somewhat speculative because the area of northern Israel is not sufficiently covered by seismic lines. The main chronostratigraphic boundaries in the wells (Table 1) follow Derin et al. (1980), Fleischer & Varsahvsky (2002), Zion Oil and Gas Inc. (2005) and Gardosh et al. (2008).
Rifting structures in the subsurface of the Levant Basin and margin The Tethyan extensional structures of the Levant area were divided here into four zones of deformation, distributed, more or less, parallel to the Mediterranean coast (Fig. 2). These are from east to west: (a) the inner basins; (b) the highs along the Mediterranean coastline; (c) the offshore highs
The central part of Israel is occupied by the Judea Graben (Figs 2a & 3) (Freund et al. 1975). This c. 150 km long and c. 30 –50 km wide basin formed as a Triassic and Early Jurassic depocentre, bounded by elevated horst blocks (Fig. 2a). The axis of the Judea Graben is marked by the thick Triassic sections of the Ramallah-1 (1700 m, base not reached) (Derin et al. 1980) and Devora-2A wells (2600 m, base not reached) (Derin & Gerry 1979). The Palmyra Trough is a similar depocentre filled by more than 2 km of Triassic sediments that is identified by well and seismic data in southern Syria (Garfunkel 1998; Brew et al. 2001; Sawaf et al. 2001). Restoring the sinistral motion along the Dead Sea transform shifts the Palmyra Trough 100 km southward (Fig. 2a). In this position, it correlates to the northeastern part of the Judea Graben, suggesting a link between these two structures in Triassic time (Garfunkel 1998; Flexer et al. 2005). The Judea Graben changes direction and widens towards the north, but its structure north of latitude 318500 (Fig. 2a) is poorly constrained by the available subsurface data. Important evidence to the structural pattern in northern Israel is the c. 2.5 km thick, latest Triassic and earliest Jurassic series of the Asher volcanics that were penetrated by Atlit-1 Deep (Fig. 2a, Table 1) and several other wells in northern Israel (Derin et al. 1982; Gvirtzman & Steinitz 1983; Kohn et al. 1993). Stratigraphic analysis of theses wells shows that the Asher extrusive volcanics are overlain by a continuous and relatively flat, shallow-marine, Middle Liassic section, indicating that the volcanic rocks accumulated in an active, fault controlled depression (Fig. 3c) (Garfunkel & Derin 1984; Garfunkel 1989). This depression is termed here the Asher Basin. The Asher Basin is reconstructed as a several tens of kilometre long graben that extends from the Haifa Bay on the northern coast of Israel towards the SE (Fig. 2a). Owing to the limited amount of seismic and well data, this postulated graben is largely based on the shape of the magnetic field in the area. The magnetic map shows a conspicuous, positive anomaly centred NE of the Atlit-1 Deep well that is interpreted to be caused by the deeply buried, Asher basalt flows (Rybakov et al. 1997, 2000). The elongated, NW–SE oriented magnetic anomaly is assumed to reflect the shape of a fault bounded depression that was filled with the Jurassic extrusive rocks (Garfunkel & Derin 1984;
14
Table 1. Summary of well data used in this study. The depth intervals of chronostratigraphic units (in meters from KB) are taken, for most of the wells, from Fleischer & Varsahvsky (2002). Data for well 9 is taken from Derin et al. (1980); for well 11 from Zion Oil and Gas Inc. (2005); and for the Liassic to Bathonian section in wells 18 and 19 from Gardosh et al. (2008). The Cretaceous depth interval in wells 17, 18 and 19 includes the Tithonian succession Well no. Well name
Depth and thickness (in metres) of stratigraphic units Palaeozoic (Permian)
Precambrian Basement Depth Atlit-1 Deep Ashqelon-2 Betarim-1 Bessor-1 David-1 Deborah-2 Gaash-2 Gevim-1 Halal-1 Helez Deep-1 Maanit-1 Deep Meged-2 Nevatim-1 Pleshet-1 Ramallah-1 Deep Talme Yafe-4 Yam-2 Yam West-1 Yam Yafo-1 Yinnon-1
þ Base of unit was not penetrated. – Unit is missing.
Depth
Thick.
4567 –4465
102
4465 – 4230 235 5998 – 5455 543þ
4620 –4588
32
4588 – 4078
510
6093 –5978
115
5978 – 5726
252
5344 –5315
29
5315 – 4788
527
Jurassic (Liassic – Bathonian)
Jurassic (Callovian – Kimmeridgian)
Cretaceous (Neocomian – Turonian)
Depth
Thick.
Depth
Thick.
Depth
Thick.
Depth
Thick.
6531 – 5390
1141þ 65 1645 2585þ 913þ 208 685þ 926 1419þ 1599þ
4788 – 4200 6361 – 4672
588 1689þ
3525 161þ 1449þ 1762 1540 1582 1614 1564 2814 2830 937 1482 434þ 1595 3162 1493 142þ 350þ 437þ 364þ
1865 – 1446 – 1790 – 1241 2403 – 2060 2270 – 1956 1480 – 646 2981 – 2506 2306 – 1876 815 – 457 1970 – 1858 2341 – 1856 2119 – 1770 1760 – 1340 2605 – 2131 1510 – 901 – 5235 – 4765 4900 – 4400 5350 – 4410 2616 – 2210
418
4230 – 4165 5455 – 3810 5647 – 3062 5508 – 4595 4078 – 3870 4314 – 3629 5726 – 4800 4697 – 3278 5200 – 3601
5390 – 1865 4076 – 3915 3239 – 1790 4165 – 2403 3810 – 2270 3062 – 1480 4595 – 2981 3870 – 2306 3629 – 815 4800 – 1970 3278 – 2341 3601 – 2119 2194 – 1760 4200 – 2605 4672 – 1510 4204 – 2711 5377 – 5235 5250 – 4900 5787 – 5350 2980 – 2616
1447 – 60 3915 – 2249 1241 – 10 2060 – 628 1956 – 279 646 – 23 2506 – 803 1876 – 567 457 – 0 1858 – 455 1856 – 3 1770 – 10 1340 – 24 2131 – 615 901 – 4 2711 – 1199 4765 – 2730 4400 – 2874 4410 – 2380 2210 – 394
1367 666 1231 1432 1377 623 1703 1309 457 1403 1853 1760 1316 1516 897 1512 2035 1526 2030 1816
549 343 314 834 475 430 358 112 485 349 420 474 609 470 500 940 406
M. A. GARDOSH ET AL.
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20
Thick.
Triassic
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN Fig. 3. Geological cross-sections showing the main stratigraphic successions, faults and structural blocks on the Levant Basin and margin. Thickness variations in the Palaeozoic, Triassic and Lower Jurassic (Liassic to Bathonian) reflect syn-tectonic deposition, related to the development of an extensive graben and horst system during Tethyan rifting activity. The location of the sections is shown in Figure 2a. 15
16
M. A. GARDOSH ET AL.
Garfunkel 1989). The reduced thickness of the Asher volcanics in Devora-2A (Garfunkel 1989) may indicate the existence of a large, normal fault west of the well that formed the northeastern limit of this depression (Fig. 3c). The Asher Basin seems to form a north– west extension of the Tethyan, Judea Graben (Fig. 2a). It was later modified by the c. 1 km vertical offset of the Carmel Fault that is associated with the Cenozoic activity along the Dead Sea plate boundary (Fig. 3c). In Early to Middle Jurassic time, the Judea Graben may have also extended northward, as indicated by a relatively thick Bathonian section (600 m) found in outcrops at the Hermon area (Hirsch et al. 1998; Hirsch 2005). The postulated Hermon Basin (Fig. 2a) was probably limited in size and did not extend eastwards to the Palmyra Trough where the Jurassic section is only 100 m thick (Mouty 1997; Hirsch 2005). A set of NE–SW oriented faults found in the southern part of the Judea Graben (Figs 2a & 3b) formed intra-graben fault blocks. These were later inverted during the Syrian Arc contractional deformation of the northern Negev (Freund et al. 1975; Gelbermann 1995; Druckman et al. 1995). The Judea Graben is delimited in the SE by a series of small, positive structures, that is, the Massada and Heimar Highs, located on the western margin of the Dead Sea (Fig. 2a) (Goldberg & Friedman 1974; Druckman 1974; Freund et al. 1975). The Ramon High, located near the Ramon fold (Druckman 1977; Buchbinder & Le Roux 1993), delimits the graben further to the south (Fig. 2a). The Judea Graben extends from the central Negev westward into northern Sinai. A Triassic tectonic depression in this area, termed here North Sinai Basin (Fig. 2a), is indicated by a relatively thick Carnian section at the bottom of the Halal-1 well (Fig. 2a, Table 1) (Garfunkel & Derin 1984; Derin & Gerry 1986a). Subsidence of the North Sinai Basin continued during the Early to Middle Jurassic. This activity is reflected by the thick Liassic to Bathonian interval found in outcrops at Gebel Maghara (1700 m) (Alsharan & Salah 1996; Hirsch et al. 1998) and penetrated in the Halal-1 well (2600 m) (Derin et al. 1976). The shape and internal structure of the North Sinai Basin are not well delineated owing to insufficient seismic and well data. It was most likely formed by down-to-the-north sets of fault blocks, oriented to the NE–SW and east –west (Fig. 2a). This trend is inferred from the direction of Syrian Arc folds that are interpreted as inverted Early Mesozoic structures (Moustafa & Khalil 1990; Alsharhan & Salah 1996). Small intra-basin blocks and localized highs and lows, such as those found within the Judea Graben, may have also existed in this area.
The highs along the Mediterranean coastline A series of structural highs along the Mediterranean coastline forms the western boundary of the Judea Graben. The most prominent structure is the NE – SW trending Gevim High (Figs 2a, 3b, 4 & 5). This narrow structure was penetrated by the Gevim-1 and the Bessor-1 wells (Figs 3c & 4). Both wells reached Precambrian rocks in depth of 4460 m and 4293 m below sea level, respectively, thus indicating that the Gevim High is the highest basement block in central Israel. The Gevim High is bounded by two sets of faults: the Helez Fault on the NW and the Pleshet Fault on the SE (Fig. 4). The stratigraphic relations in the Gevim-1 and the Helez Deep-1 wells, located respectively on the footwall and hanging wall of the Helez Fault, indicate 1740 m of vertical offset during the Triassic and Early to Middle Jurassic (Gardosh & Druckman 2006). The stratigraphy of the Pleshet-1 well (Derin et al. 1985), located on the downthrown side of the Pleshet Fault, indicates about 700 m of vertical offset on an Upper Permian level that can be attributed to the same tectonic activity. Further evidence for the activity on the Helez Fault is found in Helez Deep-1. The well penetrated c. 300 m thick section of coarse-grained, polymictic conglomerate, within the Triassic succession [the Erez Conglomerate of Druckman (1984), or the Or Haner Conglomerate of Derin (1979) and Garfunkel & Derin 1984]. This clastic unit was interpreted as an alluvial fan that accumulated at the foot of the Helez Fault escarpment (Fig. 4) (Druckman 1984; Garfunkel & Derin 1984). In the central and northern part of Israel the Gaash-Meged and Maanit structures (Figs 2a, 3a, c & 6) delineate the western boundary of the Judea Graben and probably also of its NW extension, the Asher Basin. The two highs are inferred from the reduced thickness of the Jurassic and Triassic sections in the wells that penetrated these structures (Table 1). The Liassic to Bathonian interval reaches 1614 m in Gaash-2 (Derin et al. 1981; Gvirtzman et al. 1984), 1482 m in Meged-2 (Givot Olam Oil Ltd 1995), and 937 m in Maanit-1 Deep (Zion Oil and Gas Inc. 2005) (Fig. 3a, c, Table 1). The same stratigraphic interval is more then 3000 m thick in the Ramallah-1 well within the Judea Graben (Fig. 3a, Table 1) (Derin et al. 1980; Fleischer & Varsahvsky 2002). A similar trend is observed in the Triassic section. In Gaash-2, the Scythian to Carnian interval is only 913 m thick (Derin et al. 1981), whereas in Ramallah-1, within the Judea Graben it reaches 1700 m (Fig. 3a, Table 1) (Derin et al. 1980). Seismic data indicate that the axis of the coastal highs changes its direction from southern Israel
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
17
Fig. 4. Interpreted, seismic profile DS-512 (time migrated), across the Gevim High and the western edge of the Judea Graben. The structure is bounded by two fault zones (Helez and Pleshet) that were intermittently active from Late Palaeozoic to Middle Upper Jurassic time. The Middle Triassic Erez conglomerate (a) is an alluvial fan that accumulated at the foot of the Helez fault escar/minent. The Helez and Pleshet faults were re-activated in a reverse motion during the Late Cretaceous to Cenozoic Syrian Arc folding phase. Note the location of the Upper Jurassic to Middle Cretaceous, rimmed shelf-edge that developed on the western edge of the Gevim High. The ages of the formation tops are: Basement, Precambrian; Gevim, Permo-Triassic; Arkov, Permian; Erez, Anizian; Saharonim, Anizian to Carnian; Mohila, Carnian; Qeren, Aalenian; Shederot, Bathonian; Zohar, Callovian; Beer Sheva, Oxfordian; Yagur, Albian. Abbreviations for chronostratigraphic ages are: L., Lower; M., Middle; U., Upper. Double arrows refer to reactivation of older, normal faults in a reverse motion. The location of the profile is shown in Figure 2a.
northward. The normal faults that bound the Gevim High trend NE– SW whereas in the area of the Gaash High the main faults trend north –south (Fig. 2) (Gelbermann 1995). A north–south to NNW–SSE direction is inferred also for the axis of the Maanit High further to the north, although the seismic coverage of this structure is limited. The northward shift of the structural grain appears to be a fundamental characteristic of the central Levant area and is reflected also by the direction of two younger geological elements: the Jurassic – Cretaceous shelf –edge and the Syrian Arc fold belt.
The offshore highs and lows Palaeozoic to Early Mesozoic structures in the Levant offshore are revealed by the interpretation and mapping of two regional seismic markers: (a) the top of the crystalline basement and (b) the Middle Jurassic unconformity (Figs 7–10) (Gardosh & Druckman 2006; Gardosh et al. 2006, 2008). The top of the crystalline basement is
identified by significant change in seismic character from reflection-free zones below to parallel, continuous reflection series above (Fig. 8). This transition is assumed to reflect the contact of magmatic and metamorphic rocks with the overlaying Palaeozoic to Mesozoic sedimentary section. Similar characteristics of the ‘top basement’ reflector were interpreted by Vidal et al. (2000) in the northern part of the Levant Basin. In seismic profiles, the Middle Jurassic unconformity (Fig. 8) marks a change in seismic character, from high amplitude and continuous reflections below, to low amplitude and discontinuous reflections above. In the offshore Yam West-1 well, this marker is correlated to the top of the Bathonian, Shederot Formation (Fig. 8) (Gardosh et al. 2006, 2008). The change in seismic character is interpreted to reflect a transition from Lower Jurassic, shallow-marine carbonates to Middle–Upper Jurassic deepwater strata (Gardosh & Druckman 2006; Gardosh et al. 2006, 2008). Thickness changes, dipping reflections, and discontinuity of seismic
18
M. A. GARDOSH ET AL.
Fig. 5. Interpreted, seismic profile DS-3589 (time migrated), showing a Palaeozoic basin in the northeastern edge of the Gevim High. Onlapping of the Permian, Triassic and Lower Jurassic strata (marked by open arrows) is associated with subsidence and faulting. Correlation of seismic reflections to the Gevim-1 well suggests that this activity may have started prior to the deposition of the Permian Arkov Formation. The ages of the formation tops are in Figure 4. Abbreviations for chronostratigraphic ages are: L., Lower; M., Middle; U., Upper. The location of the profile is shown in Figure 2a.
events found throughout the offshore, between the top of the crystalline basement and the Middle Jurassic seismic markers, were interpreted to reflect syn-tectonic rifting activity (Figs 7 –10).
A large structure, revealed by the depth of the crystalline basement and the increased thickness of the Palaeozoic to Middle Jurassic interval, is the Pleshet Basin (Figs 7–9). This structure is a
Fig. 6. Interpreted seismic profile DS-559 (time migrated) across the Gaash-Meged High. Discontinuity of seismic reflections within the Triassic and Permian intervals indicates vertical motions during these times, depicted by the normal faults at the eastern part of the profile. The Gassh-Meged High was an elevated structure between the Pleshet Basin to the west (not shown on this profile) and the Judea Graben to the east. An apparent tilt of the Gaash area in the western part of the profile is a pull-down effect, created by the low-velocity Cenozoic cover (compare to section 1 in Figure 3a). The ages of formation tops are: Zafir, Scythian; Mohila, Carnian; Shefayim, Norian; Lower Haifa, Bajocian; Upper Haifa, Oxfordian. Abbreviations for chronostratigraphic ages are: L.-Lower, M.-Middle, U.-Upper. The location of the profile is shown in Figure 2.
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
c. 150 km long and c. 50 km wide graben that extends along the southeastern part of the Levant Basin. The Pleshet depocentre is found about 50 km west of the coastline, opposite the Gaash– Meged High (Figs 2a, b & 3a). The existence of a deep Mesozoic basin in this area was first suggested by Cohen et al. (1988) based on potential field data that indicates a magnetic basement at great depth (Domzhalski 1986). The interpretation of the new seismic data show a structural low that more or less coincides with the area of the Pleshet Basin as outlined by Cohen et al. (1988) (Fig. 2b). This basin structure is interpreted as a Tethyan Graben, similar in size and shape to the Judea Graben further to the east (Figs 2a & 3a, b). The Pleshet Basin is bounded on the SE by a 5 – 10 km wide horst, oriented in a NE–SW direction, termed here the Yam High (Figs 2, 3 & 8). A set of normal, down-to-the-west step faults displace the top of the crystalline basement marker west of the Yam High. The normal faulting is associated with a significant increase of thickness and onlapping of the Palaeozoic to Middle Jurassic section (Fig. 8). Some of the faults, such as west of the Yam West-1 well (Fig. 8), were later reactivated as high-angle thrust faults. The Yam High probably extended further to the NE converged with the
19
Gaash High near the coastline. The structural configuration of this area near the Yam Yafo-1 well is, however, highly complex and difficult to interpret owing to the younger contractional deformation (Fig. 3a). The Pleshet Basin is delimited to the NW by the Jonah High (Figs 2 & 3a). This feature, which was first identified by Folkman & Ben Gai (2004), is a conspicuous basement high, about 80 km long and 10 –20 km wide, controlled by NE–SW oriented sets of normal faults (Figs 2b, 7 & 9) (Gardosh & Druckman 2006; Gardosh et al. 2006, 2008). The Jonah High is likely underlain by a large magmatic intrusion (Fig. 10). A magnetic anomaly that is located on this narrow horst (Fig. 2a) may be associated with an extrusive volcanic body of Early Mesozoic age (Fig. 7) (Gardosh et al. 2006, 2008). A conspicuous asymmetric graben is observed NW of the Jonah High. It is characterized by southeasterly dipping seismic events and increased thickness of the Palaeozoic to Middle Jurassic interval (Figs 9 & 10). The western limit of this graben is a large basement structure, about 100 km long and 20 –40 km wide, termed here the Leviathan High (Figs 2, 3a & 9). This wide and flat horst is controlled by several sets of normal faults. It may also be associated with a magmatic intrusion in
Fig. 7. Interpreted, composite marine seismic profile (time migrated) showing Tethyan extensional structures in the Levant Basin, offshore. Thickness variations in the Palaeozoic to Middle Jurassic intervals were controlled by normal faulting and reflect syn-tectonic deposition. Most, but not all of the normal faults were reactivated in a reverse motion during the Syrian Arc contractional phase. The Jonah High is a fault controlled basement structure that bounds the Pleshet Basin on the east, and a smaller asymmetric graben on the west. The Jonah High may have been the loci of Early Mesozoic, extrusive volcanic activity (a) (Gardosh et al. 2008). Abbreviations for chronostratigraphical ages are: L., Lower; M., Middle; U., Upper. Double arrows refer to reactivation of older, normal faults in a reverse motion. The location of the profile is shown in Figure 2b.
20
M. A. GARDOSH ET AL.
Fig. 8. Interpreted marine seismic profile (time migrated), showing the southeastern edge of the Pleshet Basin. This fault-controlled, rift basin is characterized by a high-amplitude, divergent reflection package that onlaps the Yam High. The base of the syn-rift sequence is correlated to the contact between chaotic and well-layered seismic packages, interpreted as the top of the crystalline basement. The top of the sequence is a continuous reflection that is correlated to the Bathonian, Shederot Formation in the Yam West-1 well. The continuous, sub-parallel, high-amplitude seismic reflections within the Pleshet Basin are interpreted as continental to shallow-marine strata. The maximal thickness of the basin-fill in this profile is 4 –5 km (calculated with interval velocity ¼ 6000 m s21). Some, but not all of the Tethyan normal faults were reactivated as high-angle thrusts during Syrian Arc folding, observed at the edges of the profile. Abbreviations for chronostratigraphic ages are: L., Lower; M., Middle; U., Upper. Double arrows refer to reactivation of older, normal faults in a reverse motion. The location of this profile (an enlarged part of Fig. 7) is shown in Figure 2.
depth. The conspicuous, positive Bouguer gravity anomaly found in the Leviathan area (Fig. 2c), suggest the existence of such an intrusive body (Gardosh et al. 2006, 2008). Several, small basement highs, about 10 km wide and 20 km long were identified south of the Jonah and Leviathan horsts (Fig. 2a, b). The belt of deep-seated highs and lows that is identified offshore Israel extends further to the NE and SW parts of the Levant Basin. Seismic and gravity data indicate the existence of Early Mesozoic structures below the Cenozoic, Nile Delta cone (Fig. 2c) (Aal et al. 2000; Bentham et al. 2007). The Rosetta fault that extends from the Nile Delta to the eastern side of the Eratosthenes high (Fig. 2a) (Aal et al. 2000) is oriented in NE– SW direction, similar to the direction of the
Leviathan and Jonah horsts. Triassic and Jurassic highs that were recently identified on seismic reflection profiles offshore Lebanon (Fig. 2a) (Roberts & Peace 2007) fit well to the regional structural pattern.
The Eratosthenes High The Eratosthenes seamount is a large structure, located at the NW of the Levant Basin (Figs 1 & 2). It is characterized by distinct geophysical and geological properties. The seamount overlies a positive magnetic anomaly that is interpreted to be associated with a voluminous igneous body in depth (Fig. 2a) (Ben-Avraham et al. 1976; Garfunkel & Derin 1984). Wide-angle seismic refraction tests show that the Eratosthenes area is
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
21
Fig. 9. Interpreted, composite marine seismic profile (time migrated), showing Tethyan extensional structures in the Levant Basin, offshore. Three structural highs: Jonah, Leviathan and Eratosthenes, are underlain by elevated basement blocks that display a chaotic seismic character and may comprise magmatic intrusions. These fault-controlled structures bound three rift basins that are characterized by continuous, divergent seismic reflection packages. The Pleshet Basin on the right-hand side of the profile is bounded by the Gaash-Meged High, located east of the profile. Abbreviations for chronostratigraphic ages are: L., Lower; M., Middle; U., Upper. Double arrows refer to reactivation of older, normal faults in a reverse motion. The location of the profile is shown in Figure 2b.
Fig. 10. Interpreted marine seismic profile (time migrated), showing the Jonah High and the asymmetric graben on its western flank This Tethyan, fault-controlled rift basin is characterized by a high-amplitude, divergent reflection package that thickens towards the Jonah structure. The base of the syn-rift sequence is correlated to the contact between chaotic to well-layered seismic packages, interpreted as the top of the crystalline basement. Its top is a high-amplitude reflection that is correlated to the Middle Jurassic unconformity. The maximal thickness of the syn-rift sequence in this profile is 7 –8 km (calculated with interval velocity ¼ 6000 m s). Abbreviations for chronostratigraphic ages are: L., Lower; M., Middle; U., Upper. The location of this profile (an enlarged part of Fig. 8) is shown in Figure 2b.
22
M. A. GARDOSH ET AL.
underlain by a low velocity crystalline crust (Vp ¼ 6 km/s) that is interpreted as a continental or intermediate type (Makris et al. 1983; BenAvraham et al. 2002). A borehole drilled on the northern slope of the seamount penetrated Lower Cretaceous, shallow-marine limestones similar to Cretaceous rocks found in the Levant margin (Mart & Robertson 1998). Based on these lines of evidence, the Eratosthenes is interpreted by most authors as a continental block that was detached and drifted from the nearby Afro-Arabian plate during the Early Mesozoic break-up of Gondwana (Garfunkel & Derin 1984; Kempler 1998; Mart & Robertson 1998). This interpretation is generally supported by the new, offshore, seismic profiles that show a basement high and a series of fault blocks of an assumed Mesozoic age SE of the seamount (Figs 2b & 9). This structure probably represents part of the larger Eratosthenes high that is delimited on the east by the NE –SW trending Qattara – Eratosthenes (Rosetta) fault zone (Fig. 2a) (Aal et al. 2000). The Cretaceous limestones found on the seamount supports the assumption that during Late Mesozoic times this area was part of an elevated, shallowmarine shelf located at the northwestern side of the deep-marine Levant Basin (Gardosh & Druckman 2006). The Eratosthenes high was further uplifted by reactivation of the Mesozoic faults during the Cenozoic convergence of the AfroArabian and Eurasian plates (Robertson 2000). Zverev & Ilinsky (2000, 2005) recently studied the deep structure of the Eratosthenes by a new set of seismic refraction profiles. They found a complex pattern of high and low velocity layers at the upper crust, below the seamount. The different velocities were attributed to alternating layers of lavas, volcanic sediments and basic magmatic rocks. Based on these findings, Zverev & Ilinsky (2000, 2005) suggested that the Eratosthenes seamount is a long-lived volcanic edifice associated with pre-Cretaceous magmatic activity, possibly similar to Early Jurassic volcanic rocks found in northern Israel. These authors further suggested that the crust underneath the Eratosthenes is similar to the crust of the adjacent Levant Basin. Although it is not clear whether the Eratosthenes seamount overlies a drifted continental block (Garfunkel 1989) or an Early Mesozoic volcano edifice (Zverev & Ilinsky 2000, 2005), in either case its formation appears to be the result of Tethyan rifting activity.
Timing of the rifting activity The structures found in the deep subsurface of the Levant region, onshore and offshore, appear to
have been active during several pulses of extension and faulting, followed by periods of relative tectonic quiescence. Parts of these rift structures were active in varying intensity during different times. Well data indicate three main phases of activity: Late Palaeozoic, Middle to Late Triassic and Early to Middle Jurassic.
Late Palaeozoic Direct evidence for Late Palaeozoic continental breakup is found in the Palmyrides area of central Syria (Fig. 1). Well and seismic data show up to 1 km thick Permian succession in a fault-controlled trough that is interpreted as an intra-continental Permian rift (Garfunkel 1998; Brew et al. 2001; Sawaf et al. 2001). More evidence for this tectonic phase is found near the Mediterranean coastline. In the Gevim High (Fig. 2a), the Gevim-1 and Bessor-1 wells record a reduced Permian section of several tens to a few hundred metres, overlaying the Precambrian basement (Fig. 3c). These stratigraphic relations indicates at least two major tectonic events: a pre-Permian, possibly Carboniferous or earlier uplift that resulted with erosion of the entire Early to Middle Palaeozoic section, and an uplift that resulted with deposition of a thin Permian section. Palaeozoic vertical motions in the area of the Gevim High are evident also on the seismic profile in Figure 5, which shows significant thickness increase and onlapping of Palaeozoic strata on the northeastern flank of the structure. Correlation of seismic reflections to the Gevim-1 well (Fig. 5) suggests that the subsidence may have started during pre-Permian time. Permian tectonic event in this area is supported by facies change of the Saad and Arkov Formations. In the Gevim-1 well (Salhov 1996), these rock units are predominantly siliciclastic and sand rich, whereas in the David-1 well (Derin 1995), located about 40 km north of the Gevim High (Fig. 3b), they are comprised of shallow-marine carbonate. In the offshore, the Pleshet Basin, and the a-symmetric graben located between the Jonah and the Leviathan highs contain 4– 8 km thick, preMiddle Jurassic section (Fig. 3a, c) (Gardosh & Druckman 2006). The maximal thickness of the Triassic to Middle Jurassic section in the onshore Deborah-2 and Ramallah-1 wells is 4 to 5 km respectively. It is, therefore, possible that part of the excessive thickness in the offshore basins is associated with greater Triassic and Palaeozoic subsidence. This scenario may be supported by the similarity of the onlapping pattern observed on seismic profiles at the edge of the Pleshet Basin (Fig. 8) and north of the Gevim High (Fig. 5). Considering the examples described above, it is evident
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
that extension and rifting already took place in the Levant region during the latest Palaeozoic period, although the dimension and configuration of this tectonic event are not well constrained.
23
shelf deposits. It is, therefore, likely that the Triassic Tethyan rifting did not result with the development of a deep-water oceanic realm in the studied area.
Early to Middle Jurassic Middle to Late Triassic The Early part of the Triassic was a period of relative tectonic quiescence (Garfunkel 1998). Evidence for Middle Triassic faulting and vertical motions is found in several locations. An uplift of the Gevim High is interpreted from the Gevim-1 and Bessor-1 wells that penetrated a thin section of the Ladinian (Upper–Middle Triassic) Saharonim Formation overlying the Permian (Fig. 3c). Middle Triassic faulting event is interpreted from the Erez (Or Haner) fault breccias of an assumed Anisian age (Druckman 1984; Garfunkel & Derin 1984). Further evidence for this phase is inferred from thickness changes of the Anisian, Gevanim Formation in the northern and central Negev (Druckman 1974). A second, Middle to Upper Triassic faulting event is observed in the Judea Graben. The Deborah-2 and Ramalah-1 wells (Fig. 2) (Derin & Gerry 1979, 1986b) penetrated a c. 1 km of the Carnian, Mohilla Formation, which is only several tens of meters thick in other parts of the country. The evaporitic facies of the Mohilla Formation is recorded by wells and outcrops in the northern Negev (Druckman 1974). This facies is associated with restricted water circulation and the development of evaporitic conditions in tectonically active depressions (Druckman 1974) at the southeastern part of the Judea Graben. Upper Triassic subsidence also took place within the North Sinai basin as indicated by the relatively thick Carnian section at the bottom of the Halal-1 well (Garfunkel & Derin 1984; Derin & Gerry 1986a). The accumulation of a thick sedimentary section in the Pleshet Basin and in the graben NE of the Jonah High (Fig. 3a, b) may be partly attributed to Triassic tectonic activity although no well control is available to confirm this hypothesis. It is further unknown whether the Levant Basin was a deep marine basin during Triassic time. Pelagic Triassic strata is found in outcrops at Antalya (southern Turkey) and in the Mammonia blocks (Cyprus) (Druckman et al. 1982; Robertson 2000; Robertson & Xonophontos 1993), suggesting the development of deep-marine conditions north of the Levant (Garfunkel 1998; Robertson 1998, 2007). On the other hand, the Triassic section of the Levant onshore is characterized by shallow-marine to continental environments. The continuous, high-amplitude reflections that characterize the Palaeozoic to Middle Jurassic interval on offshore seismic profiles (Figs 7–10) may be interpreted as shallow-marine,
Activation of Tethyan rift structures during the Early to Middle Jurassic took place throughout the Levant region. More than 3 km of Liassic strata were deposited at the central part of the Judea Graben near the Ramallah-1 well (Figs 2a & 3a). Further to the NW the Asher Basin (Figs 2a & 3c) accumulated c. 2.5 km thick section of Liassic basalts and pyroclasts (Asher volcanics). The Deborah-2 well, located on the edge of the Asher Basin (Figs 2a & 3c), penetrated c. 200 m of the Asher volcanics and only c. 1 km thick Liassic to Bathonian section. Whereas, increased thickness of the Bajocian and Bathonian strata is found in outcrops at Mount Hermon (Goldberg et al. 1981; Hirsch et al. 1998). These thickness variations suggest segmented subsidence at the northern part of the Judea Graben (Fig. 2a) during Early to Middle Jurassic time. Reduced thickness of the Liassic to Bathonian sections in the Maanit, Gaash-Meged and Gevim highs (Fig. 3) indicates uplifting of the coastal area. The vertical offset on the Helez Fault, at the northwestern edge of the Gevim High amounts to 1.2 km (Figs 3b & 4) (Gardosh & Druckman 2006). In the southern edge of the Judea Graben, thickness changes of the Ardon and Inmar formations (Goldberg & Friedman 1974; Druckman 1977; Buchbinder & le Roux 1993) are associated with Early–Middle Jurassic tectonic activity. Further to the SW, in the North Sinai Basin the Liassic to Bathonian section is more then 2000 m thick (Halal-1 well and the Gebel Magahra outcrops; Druckman 1977; Derin & Gerry 1986a; Alsharhan & Salah 1996; Hirch et al. 1998), indicating significant subsidence during the Early to Middle Jurassic period. The thick sedimentary section in the Pleshet Basin and in the graben NE of the Jonah High (Figs 2a & 3a, b) may be partly attributed to this faulting phase. Similar to the Triassic conditions it is unknown whether the Levant Basin was a deep-marine basin during Early Jurassic time. The lithology of the Middle Jurassic Shederot Formation in the Yam West-1 well offshore indicates shallow marine conditions (Gardosh et al. 2006, 2008). Likewise, shallow-marine to fluvio-deltaic depositional environments characterize the Lower to Middle Jurassic section throughout the Levant onshore. Continuous, high-amplitude seismic reflections found below the Middle Jurassic unconformity in offshore seismic profiles are interpreted as shelf carbonate (Fig. 8). This raises the possibility that
24
M. A. GARDOSH ET AL.
the Jurassic rift structures of the Levant area developed on a continental to shallow-marine platform. Deeper-marine conditions may have locally prevailed in parts of the Pleshet Basin and in other offshore structural lows, accompanied by growth of carbonate buildups on the adjacent Jonah, Leviathan and Eratosthenes highs. The Early Jurassic rifting phase was associated with extensive magmatic activity. Extrusive magmatic rocks were penetrated by wells in the Asher Basin at northern Israel (Asher Volcanics) (Dvorkin & Kohn 1989; Kohn et al. 1993). Magnetic anomalies that are interpreted to be associated with Early Jurassic volcanism are found in the Eratosthenes and Jonah highs offshore (Fig. 2) (Zverev & Ilinsky 2000, 2005; Gardosh et al. 2006, 2008). A large magnetic anomaly found within the Judea Graben in central Israel (Fig. 2a) is also interpreted to be associated with Early Jurassic magmatic activity, probably of a more intrusive nature (Garfunkel 1989; Rybakov et al. 1995). The activity of the Levant rift system ceased during the Middle to Upper Jurassic period. The Oxfordian –Cimmerian and the younger, Cretaceous section in the inner part of the Levant show minor thickness variations indicating tectonic quiescence. At this time conspicuous, carbonate shelf-edge developed on the eastern part of the Levant Basin, along the present-day Mediterranean coast, and large volumes of deepwater siliciclastics and carbonates were deposited on the slope further to the west (Cohen 1976; Bein & Gvirtzman 1977; Gardosh 2002). This accumulation of deepwater sediments within the Levant basin is associated with post-rift subsidence, caused by cooling of newly formed crustal units (Garfunkel & Derin 1984; ten Brink 1987; Gardosh 2002).
Tectonic reconstruction The Tethyan tectonic reconstruction of the Levant region is controversial. Some important aspects, such as the timing and location of rifting, are still under debate (Robertson & Mountrakis 2006). Several models have been proposed for the rifting direction and structural style (Fig. 11). Dewey et al. (1973), Bein & Gvirtzman (1977), Robertson & Dixon (1984), Stampfli & Borel (2002), and others, suggested a NE –SW opening in the Levant Basin that was accommodated by a major north– south oriented transform fault along the eastern Mediterranean coast (eastern transform margins) (Fig. 11a). According to Robertson (1998), the Levant margin represents either an orthogonally rifted or a transform margin. Garfunkel & Derin (1984) proposed rifting and extension in a NW– SE direction that resulted in separation of the Eratosthenes and Tauride blocks from the Gondwanian
Craton and was accommodated by a transform fault along the Sinai coast (southern transform margin) (Fig. 11b). Hirsch et al. (1995) described a north – south oriented extension on a Gondwanian shallow shelf with no reference to a specific fault pattern. According to the present analysis, the dominant direction of the normal faults offshore and onshore Israel and in northern Sinai is NE– SW to ENE – WSW (Fig. 2a). This direction indicates extension that is perpendicular to the strike of the normal faults, in a NW– SE and NNW–SSE direction as proposed by Garfunkel & Derin (1984) and Garfunkel (1998) (Fig. 11d). The NW –SE direction of the extension in the Levant Basin implies an accommodating transform fault somewhere along its southern margin. It is unlikely that such a fault extended along northern Sinai, as this area appears to have been dominated by a NW–SE extension (Fig. 2a). Alternatively, a southern transform fault may have been located along the African coast west of the Nile Delta (Fig. 11d), as shown by Bentham et al. (2007). Minor, transfer fault zones may also have existed within the Levant Basin (Bentham et al. 2007). An apparent, left-stepping offset in the northern edge of the Jonah ridge (Fig. 2b) can be explained by such a fault. The existence of eastern transform margin, which requires north– south direction of extension in the Levant Basin and a strike-slip fault along the Levant coast (Fig. 11a) (Dewey et al. 1973; Stampfli & Borel 2002), is not consistent with the structures and fault directions that are identified within the Levant Basin. Furthermore, this model can not explain the formation of the coastal highs and the interior basins of the Levant that are located east of the presumed Levant transform fault. An additional important aspect of the tectonic reconstruction is the timing of activity of Tethyan structures. Previous views advocated a time span ranging from Carboniferous to Permo-Triassic (Stampfli et al. 2001), Late Permian to Late Triassic (Robertson 1998), Triassic to Liassic (Garfunkel 1998) and Triassic and Late Jurassic to Early Cretaceous (Hirsch et al. 1995). The present analysis indicates three main phases of activity. Vertical motions started in several locations during the Late Palaeozoic, either in the Permian or the Carboniferous (Fig. 11c), continued during the Middle and Late Triassic, and climaxed in the Early Jurassic (Fig. 11d) when most of the structures were active and magmatic activity was widespread. Although the intensity of the faulting significantly decreased towards the Middle Jurassic it appears that some activity continued through the Bathonian and possibly later, till the early Late Jurassic. According to the time frame described above, Tethyan rifting in the Levant was pulsated, and extended over a period of more then 120 Ma. Its initiation may
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
(a)
(b)
(c)
(d)
25
Fig. 11. Several, alternative tectonic reconstructions of Tethyan rifting in the Levant region. Two, previously proposed models are: (a) after Dewey et al. (1973) and Stampfli & Borel (2002), showing north– south extension with eastern transform margin and (b) after Garfunkel & Derin (1984) and Garfunkel (1998), showing NW–SE extension with southern transform margin. The reconstructions in (c) and (d) are based on the present study. Tethyan rifting activity on the northern edge of Gondwana was pulsed and progressed from the Late Palaeozoic (c) to Early Jurassic (d). The southeastern part of the Mediterranean region was not affected by sea-floor spreading during rifting. Oceanic lithosphere was formed further to the north and NW, between the Eratosthenes block and the Taurid micro-continent, as indicated by ophiolitic complexes found in Antalya, Mamonia and Baer-Bassit. ER, Eratosthenes block; AN, Antalya; MA, Mamonia; PA, Palmyra; BB, Baer-Bassit.
have been contemporaneous with the time of opening of the ‘Palaeotethys’ whereas its most active phase took place during the development of the ‘Neotethys ocean’ (Robertson 2007). The available well and seismic data suggest that break-up took place in an intra-continental setting. There is no evidence for the development of a deepmarine basin in the Levant region throughout the Late Palaeozoic to Middle Jurassic. Deep marine conditions were established within the Levant Basin in the Late Jurassic and Early Cretaceous, a period considered by most authors to be associated with post-rift cooling and subsidence (Garfunkel & Derin 1984; ten Brink 1987; Gardosh 2002). On the other hand, north of the study area (Cyprus, NW Syria and southern Turkey), there is a record
of deep-marine conditions already in the Triassic (Robertson 1998; Garfunkel 1998). Available data do not allow assessing the nature of the transition from these areas to the Levant. Perhaps the most problematic aspect of the Tethyan tectonic reconstruction is the effect of rifting processes on the shaping of the Levant crust. This subject is discussed in the following chapter.
Shaping of the Levant crust during rifting The deep structure of the Levant continental margin The deep structure of the Levant was investigated by several wide-angle seismic refraction tests, shot
26
M. A. GARDOSH ET AL.
onshore (Ginzburg & Folkman 1980; Weber et al. 2004) and offshore Israel (Makris et al. 1983; Ben-Avraham et al. 2002; Netzeband et al. 2006). Although differing in some details, their results provide a consistent picture of the crustal properties, summarized in Figure 12. The refraction profiles show that the depth to the Moho decreases from about 30 –35 km underneath southern Israel to about 20–22 km in the central part of the Levant Basin offshore (Fig. 12) (Makris et al. 1983; Ben-Avraham et al. 2002; Netzeband et al. 2006). A corresponding change was observed in the seismic velocities and thickness of the crust below the sedimentary cover. In southern Israel the crust is 25– 35 km thick and is separated to an upper (Vp ¼ 6.2–6.4 km/s) and a lower (Vp ¼ 6.7 km/s) layer (Fig. 10) (Makris et al. 1983; Ben-Avraham et al. 2002; Weber et al. 2004), whereas in the Levant Basin it is only 8 –10 km thick. Makris et al. (1983) and Ben-Avraham et al. (2002) interpreted one crustal unit in the Levant Basin with seismic velocity of 6.7 km/s. Netzeband et al. (2006) recently interpreted a thin (2–3 km) upper (Vp ¼ 6.0–6.3 km/s) and a lower (Vp ¼ 6.5–6.8 km/s) layers (Fig. 12). A change in crustal properties is observed further to the west. In the Eratosthenes area, the depth to the Moho increases to 26–28 km and its thickness to NW Eratosthenes Seamount 0
Levant Slope
Levant Basin
Cenozoic 6 km/s
Plz.-Tr. 6.0–6.3 km/s Depth (km)
Jur. Volcanic(?) Intruded upper crust 7.8–8 km/s
Mantle 0
40
SE
Negev
0
Jur.
Precambrian
6.2–6.4 km/s
50 km
10
Upper continental crust
6.5–6.8 km/s
20 6.7 km/s
30
coastline
Pliocene Jur.-Cret.
Messinian 10
23 km (Fig. 12) (Makris et al. 1983; Ben-Avraham et al. 2002; Netzeband et al. 2006). Makris et al. (1983) interpreted an upper (Vp ¼ 6.0 km/s) and lower (Vp ¼ 6.7 km/s) layers underneath the seamount (Fig. 12), whereas Zverev & Ilinsky (2000, 2005) suggest a more complex layering pattern of high and low velocity layers ranging from 5.3 to 7.7 km/s. In the Galilee area of northern Israel, BenAvraham & Ginzburg (1990) interpreted a 20 –25 km thick crust below the sedimentary cover. The relatively thinner crust of this area, in comparison to southern Israel, is associated with a uniform seismic velocity of 6.1–6.5 km/s (Ginzburg & Folkman 1980). The change in the crustal properties is indicated also by gravity data. The regional Bouguer gravity map (Fig. 2c) shows negative anomaly associated with a lighter, continental type crust south and east of the Mediterranean coastline and a positive anomaly associated with a denser crust in the Levant Basin (Rybakov et al. 1997; Rybakov & AlZoubi 2005). A similar, although less pronounced, change is observed from southern Israel northward (Fig. 2c). It is reasonable to assume that the variations in thickness and velocity of the Levant crust are related to the Tethyan break-up and rifting activity.
20
Lower continental crust 6.7 km/s
30
7.9 km/s 40
Fig. 12. Crustal scale section across the Levant Basin and margin, from southern Israel (SE) to the Erathosthenes seamount (NW); adapted from Garfunkel (1998). Structure and stratigraphy in the upper crust are taken from sections a and b in Figure 3 of this study; crustal thickness and seismic velocities are based on wide-angle seismic refraction tests (see details in text). The total crustal thinning along this section is considerably larger than the extension produced by the high-angle faults in the upper crust, indicating an extension discrepancy (see details in text). Depth-dependant-stretching derived by upwelling, divergent mantle flow (schematically shown by curved arrows) is associated with mobilization of the lower crust and part of the upper crust, along low-angle detachment surfaces (thick dashed line). The thinned crust in the centre of the Levant Basin is interpreted as the remaining, attenuated Upper continental crust that is heavily intruded by mantle material. In the Eratosthenes seamount, relatively thick crust with low seismic velocity and conspicuous magnetic anomaly may be explained by the existence of a large volcanic edifice of Early Mesozoic age. See location of the cross-section in Figure 2c.
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
Crustal thinning, subsidence, and accumulation of thick sedimentary wedges that are observed in the Levant area, characterize rifted, continental margin settings. It is, therefore, useful to discuss the evolution of the Levant margin with regards to other margins worldwide.
A global perspective on passive continental margin formation It is generally accepted that passive continental margins are created as a result of rifting causing extension and often also magmatism that leads to breakup of continents and separation of their fragments, followed by the formation of deep marine basins and an oceanic crust along the margins of the separating continental fragments. Rifted continental margins form transition zones between thick, relatively undisturbed continental crust and a newly formed, thin oceanic crust outboard. Based largely on the study of the central and northern Atlantic, two margin types were distinguished: volcanic and non-volcanic (Robertson 2007). Volcanic margins contain large volumes of intrusive and/or extrusive igneous rocks and are characterized by seaward-dipping seismic reflectors (SDR). The Greenland and the Norwegian margins of the north Atlantic are good examples (Planke et al. 2000). Non-volcanic margins generally lack extrusive igneous rocks, contain exhumed Upper mantle material, and are characterized by large rotated fault blocks. A typical example is the Iberian margin of the eastern Atlantic (Montadert et al. 1979). Breakup and separation of continental fragments, which move at large angles to the trend of the initial rift, provide a suitable tectonic setting for thinning of the lithosphere along newly formed continental margins of these two types. Recent studies of many rifted continental margins, both volcanic and non-volcanic, reveal an important characteristic; a so-called ‘extension discrepancy’ (Reston 2007). Studies from the Newfoundland and Iberia margins of the north Atlantic (Reston 2007), the northwestern Australian margin (Driscoll & Karner 1998), the Norwegian margin (Kusznir et al. 2004), and other areas, show that the extension and thinning of the whole crust and lithosphere greatly exceeds that of the upper crust.
Extension discrepancy in the Levant margin The variations in thickness of the Levant crust are well documented by wide-angle refraction profiles, as described above (Fig. 12). The new, offshore seismic reflection data allow imaging the entire Levant basin-fill down to the top of the crystalline
27
basement (Figs 7 & 9). With this new information, it is possible to estimate the amount of extension that was produced by brittle deformation during Tethyan rifting activity. The amount of extension is estimated from the geometry of the faults that displace the Palaeozoic and Mesozoic strata along the regional cross-section shown in Figure 12. There are 26 main fault zones in the area extending from southern Israel to the eastern edge of the Eratosthenes block (Fig. 12). These fault zones are comprised of normal faults that bound Tethyan grabens and horsts and reverse faults that are assumed to be reactivated Tethyan rift structures. By using the amount of apparent dip– slip on these faults we estimate the cumulative lateral displacement to be in the range of 6–7 km. The length of the section is 360 km (Fig. 12); therefore, the amount of extension is about 2%. On the same section the crystalline crust thins from southern Israel to the centre of the Levant Basin by 75% (Fig. 12). Clearly, the overall crustal thinning largely exceeds an extension produced by brittle deformation. It should be noted that our 2D estimation of fault related extension is highly simplified and it is reasonable to assume that more faults actually exist. However, even increasing the amount of faults and their vertical displacement by a factor of 2–3 would not fundamentally change the above relations. It is, therefore, concluded that, similar to many other margins worldwide, the Levant continental margin displays an extension discrepancy. Extension discrepancy at continental margins may be explained by partitioning of extension with depth, also termed depth-dependent-stretching (DDS) (Driscoll & Karner 1998; Reston 2007). Davis & Kusznir (2004) and Kusznir & Karner (2007) proposed that the mechanism that produces DDS is upwelling and divergence of continental lithosphere and asthenosphere. These authors applied a fluid-flow model for deformation of continental lithosphere and based on their modelling results concluded that an upwelling, divergent flow may cause thinning, followed by breakup, rifting and sea-floor spreading. Variations in the form and velocity of the divergent flow may lead to diversity of a rifted continental margin structure (Kusznir & Karner 2007). Can the DDS model be used to explain the extension discrepancy of the Levant continental margin? To further discuss this issue two regions are distinguished. The first region is the present-day coastal area and nearby slope (Fig. 12). There, the crystalline crust resembles the crust further inland in its two-fold seismic-velocity structure, although the upper and lower crustal units become notably thinner in this area compared with the normal continental crust farther inland. The second region is
28
M. A. GARDOSH ET AL.
the basin outboard the base of the continental slope (Fig. 12). There, the crystalline crust is thin and its average Vp is higher than the Vp of a typical continental crust, although in a relatively thin layer at its top the Vp may resemble that of an upper continental crust (Netzeband et al. 2006).
The coastal area and nearby slope Here the crust is thinned by a factor of 1.2–1.3. If thinning was the result of uniform stretching of the entire crust, then this should have been expressed by fault structures in its brittle upper part. Seismic reflection data reveal faulting (i.e. the coastal highs and the Pleshet Basin) but as noted previously the extension that it produces is quite inadequate to account for the observed thinning. As the uppermost crust was hardly stretched, one should look for processes at greater depth. Igneous activity is a likely process. Early Mesozoic volcanic rocks were found in several wells close to the Levant margin and offshore magnetic anomalies are probably caused by coeval volcanic activity that is widespread in the Eastern Mediterranean region. An important indication for the nature of this activity is provided by the Gevim quartzporphyry in the Helez Deep-1well (Fig. 4). This is c. 200 m thick, fine-grained volcanic unit of PermoTriassic age (Segev 2005). The significant feature of these rocks is their high initial 87Sr/86Sr. Steinitz (1980) found I ¼ 0.7100 + 0.029 (age 244 + 44 Ma); Segev and Eshet (2003) reinterpreted these data (rejecting one of the results) and proposed I ¼ 0.7075 + 0.0034 (age 275+ 47 Ma). These values show that the volcanics were formed by heating and melting of older mid-crustal rocks by a large intrusion of basic magma. Magma ascent was probably arrested below the cold lowdensity upper crust. Although the size of this presumed igneous body cannot be determined, it seems unlikely to have been an isolated feature, because magma formation is expected, by its nature, to occur in a region of substantial size. These considerations raise the possibility that hidden igneous intrusions are present in other places along the Levant margin. Their distribution can not be inferred from the known extent of volcanism. Neither can it be revealed by seismic reflection or refraction data. Deep intrusions are expected to heat and thus weaken the lower crust and underlying mantle. The existence of magma intrusions on the Levant margin supports the upwelling mantle flow model of Kusznir & Karner (2007). The weakened lower crust may have been stretched and thinned beneath the little deformed, cold uppermost crust resulting with DDS (Fig. 12). Reston (2007) noted that explaining the extension discrepancy by DDS
requires that the lower crust has somehow been displaced away. Such displacement may be produced by shear along low-angle detachment (Reston 1993; Driscoll & Karner 1998) or by necking or boudinage of the lower crust (Reston 2007). Decoupling of the crust by shear has been inferred from seismic reflection data in the North Sea (Reston 1993). The existence of deep shear zones on the Levant margin is postulated (Fig. 12), though it can not be proved or disproved owing to insufficient seismic data.
The basinal domain The main features of this domain, at the central part of the Levant Basin (Fig. 12) are the small thickness of the underlying crystalline crust and its high seismic velocity compared with normal continental crust. Shaping of this crust by uniform stretching is not compatible with either the minor brittle deformation observed in its upper part or with its high average Vp. Here also, crustal thinning may be explained by upwelling of a divergent mantle flow, as suggested by the DDS model (Fig. 12). Kusznir & Karner (2007) showed that a thinned upper crust lying directly on the mantle, as observed in the centre of the Levant Basin, is predicted by their upwelling divergent flow model, and can occur with pausing of the flow field prior to its reaching the surface. Similar to the coastal area, the complete displacement of the lower crust within the basin may have been produced by various types of decoupling mechanisms (Fig. 12) (Reston 2007). The higher velocity of the crust in this area is an expected result of the addition of basic intrusions from the postulated upwelling mantle, since their Vp is higher than that of normal lower continental crust.
The Red Sea analogue The processes that shaped the present-day northern Red Sea (Fig. 1) may provide modern analogues for the formation of the Levant basin. The Red Sea is an active rift system that formed by break-up of the continental lithosphere beginning in the Late Oligocene, leading to sea-floor spreading by about 5 Ma (Cochran 2005; Cochran & Karner 2007; and references therein). Two parts of this rift system are distinguished: (a) the southern Red Sea, where an oceanic sea-floor spreading centre that produces new crust with oceanic magnetic anomalies has developed, and (b) the northern Red Sea where organized sea-floor spreading is not observed (Gaulier et al. 1988; Martinez & Cochran 1988). Seismic refraction data show that the northern Red Sea is characterized by a thin crust (7–8 km)
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
with low, continental type seismic velocities (Vp ¼ 6.2–6.3 km/s), that overlies the mantle (Gaulier et al. 1988). Large-amplitude, linear gravity highs and lows observed in the northern Red Sea are interpreted as series of crustal-scale fault blocks that are oriented sub-parallel to the trend of the rift. Small, scattered magnetic anomalies are interpreted as axial volcanoes (Martinez & Cochran 1988; Cochran & Karner 2007). The physical characteristics of the Levant Basin are similar to those of the northern Red Sea, except for the somewhat higher velocity of the crust. It is, therefore, possible that the two basins were formed in a similar manner. Martinez & Cochran (1988) suggest that extension of the northern Red Sea is accommodated by block rotation at the upper crust and ductile flow within the lithosphere but no explanation for the complete removal of the lower crust, as observed by refraction data (Gaulier et al. 1988) is presented. Given the large separation between the two sides of the northern Red Sea rift that amounts to 130 –150 km and the small number of fault blocks (Martinez & Cochran 1988; Cochran & Karner 2007), it is most likely that ‘extension discrepancy’ occurs in this basin. An alternative model for the evolution of the northern Red Sea may be DDS, associated with some brittle deformation in the upper crust and removal of the lower crust by an upwelling mantle flow. Upwelling of magma in the basin is supported by the existence of small magnetic anomalies that are interpreted as extrusive igneous rocks, exposures of young basalts along segments of the basin axis, and by the very high-heat flow (Cochran & Karner 2007). It is worthwhile discussing the relations between the north and south parts of the Red Sea. Martinez & Cochran (1988) and Cochran (2005) proposed that the two parts of the rift represent different stages of evolution and with continued extension and magmatism, organized sea-floor spreading will eventually take place in the northern Red Sea. An alternative model suggested by Cocharan & Karner (2007) postulates that the two regions develop differently owing to differences in lithosphere rheology and it is possible that an oceanic spreading centre will not develop in the northern part. The two models reflect on the regional context of the analogous Levant Basin. Similarly, this basin may have developed differently than other rifted Tethyan domains owing to an inhomogeneous lithosphere.
Discussion and conclusions Rift structures of the Palaeozoic to Early Mesozoic age were previously recognized in well and seismic
29
data in the inland part of the Levant region. New, 2D seismic reflection profiles reveal similar structures in the Levant Basin offshore Israel. Integration of the onshore and offshore data allows reconstructing the pattern of this important rifting phase. Several fault belts that are generally oriented in a NE–SW direction, more or less parallel to the presentday coastline, are identified (Fig. 2a): (a) the inner basins, (b) the highs along the Mediterranean coastline, (c) the offshore highs and lows and (d) the Eratosthenes high. The fault belts comprise a broad zone of extension and rifting, more than 400 km wide, from southern Israel to the far offshore (Fig. 2a). Seismic and gravity data indicate that this zone extends further to the SW and north, below the Nile Delta and northern Egypt (Aal et al. 2000; Bentham et al. 2007) as well as onshore and offshore Lebanon (Walley 1998; Roberts & Peace 2007). The extensional structures of the Levant form part of a series of rifted margins that developed on the northern edge of Gondwana and are found today in a wide region extending from North Africa to northern India (Robertson 2007). The distribution of structures within the Tethyan rift belt of the Levant suggests an extension in a NW –SE direction. This extension was probably accommodated by the transform-rifted margin in western Egypt and North Africa (Garfunken & Derin 1984; Bentham et al. 2007). An alternative model of north –south oriented extension and transform-rifted margin along the Levant coast (Dewey et al. 1973; Robertson & Dixon 1984; Stampfli & Borel 2002) is not compatible with the direction and style of deformation observed throughout the Levant. Tethyan rifting in the Levant was pulsated and accentuated during three periods: Late Palaeozoic, Triassic and Early Jurassic (Garfunkel & Derin 1984; Garfunkel 1998). Evidence for Permian differential motions is found in well and seismic data from the Palmyra Trough (Sawaf et al. 2001) and the Gevim High near the Mediterranean coast (Figs 2a, 4 & 5). A small basin found NE of the Gevim structure (Fig. 5) may have formed prior to the deposition of the Upper Permian Arqov Formation, possibly implying Carboniferous to Early Permian time of activity. The great thickness of the sedimentary fill in the Pleshet Basin and other Tethyan lows offshore may be better explained by initiation of subsidence during the Late Palaeozoic period. In other parts of northern Gondwana, evidence for Permian breakup and rifting that was followed by the creation of oceanic crust is found in Oman and North India (Robertson 2007). Robertson (2006) recently identified deepwater, siliciclastic turbidites of Mid–Late Carboniferous to Early
30
M. A. GARDOSH ET AL.
Permian age in western Sicily and Crete and suggested that a deep-marine rift opened along the northern margin of Gondwana during this time. In summary, Late Palaeozoic rifting in the Levant area may have been more extensive and started earlier then previously considered. This rifting pulse produced vertical motions and some extension, but there is no evidence for significant magmatic activity and the formation of an oceanic crust as suggested by Stampfli et al. (2001). A second rifting pulse took place in the Triassic, activating the inner basins, the coastal highs, and the offshore highs and lows (Figs 2a & 3). Well data indicates two faulting episodes: during the Anizian (i.e. the syn-tectonic Erez Conglomerate in Helez Deep-1 well) and the Carnian (i.e. the Mohilla Evaporites). Although, these may reflect one pulse that gradually intensified towards the end of the Triassic. Minor magmatic activity is indicated by Triassic volcanic rocks found in wells at the Negev and Palmyrides areas (Segev 2005). In other parts of the Eastern Mediterranean, Triassic deep-water sediments and extensive volcanism, in places of mid-ocean range (MOR) type were found in the Baer– Bassit ophiolitic section of northern Syria, in the Antalya Complex of SW Turkey and in the Mamonia Complex of western Cyprus (Fig. 11) (Robertson 1998, 2007). These indicate the formation of well developed rifted continental margins and deep-marine basins. The Baer–Bassit section was located on the northern Gondwana margin whereas the Antalya and Mamonia Complexes formed the southern margin of the Tauride micro-continental block, at the opposite side of the Southern Tethys Ocean (Fig. 11) (Robertson 2007). In the Levant area, there is no indication for MOR type volcanism and the existence of a deep-marine basin during the Triassic. The lack of continental margin characteristics suggests that a deep Tethyan ocean did not extend as far south as the Levant. The most significant rifting pulse took place in the Levant during the Early Jurassic. Evidence for large-scale vertical motions and alkaline volcanism (i.e. Asher volcanics) are found throughout the Levant rift belt (Figs 2, 3 & 11d ), except for the Palmyra Trough that was not active during this time. The Early Jurassic tectonic pulse appears to be particularly significant in the Levant region. Other rifted margins of northern Gondwana that were formed in the Permo-Triassic (i.e. Baer–Bassit, Antalya, Mamonia) continued to subside, although rift-related activity and volcanism generally ceased (Robertson 2007). The time and style of rifting activity in the Levant are relatively well constrained by outcrop, well and seismic data. The effects of this activity on the shaping of the Levant crust are less well
understood and still being argued. The Levant Basin and adjacent coastal area are described by many authors as Mesozoic, passive continental margin (Bein & Gvirtzman 1977; Makris et al. 1983; Garfunkel & Derin 1984; ten Brink 1987; Ben-Avraham et al. 2002; Gardosh 2002). There are two main lines of evidence: (a) the marked westward thinning of the crust (Fig. 12) and (b) the formation of a shelf-edge and the accumulation of deepwater section within the Levant Basin during the Late Jurassic to Middle Cretaceous. Breakup and rifting that was followed by post-rift subsidence provides a tectonic framework to explain the formation of the Levant continental margins. In this framework two alternative models may be considered: (a) rifting–drifting and (b) DDS. The rifting–drifting model postulates breakup and separation between the African craton and the Eratosthenes seamount, which is assumed to be rooted in a continental crust. This separation was presumably followed by formation of a new crust with oceanic affinities in the centre of the Levant Basin. The rifting –drifting model raises several difficulties. There is no convincing geophysical evidence for the existence of an oceanic crust in the basin centre. The seismic velocity of the thin crust within the Levant Basin is not typically oceanic and in its upper part the velocity is similar to that of an upper continental crust (Fig. 12) (Netzeband et al. 2006). Linear magnetic anomalies that are characteristic of sea-floor spreading in other oceans are not observed. On the other hand, seismic reflection data show that Tethyan rift structures are preserved and the seismic characteristics of an oceanic crust are missing (Gardosh & Druckman 2006). Additionally, the coastal and slope area lack features that are found in other rifted continental margin such as: seaward-dipping-reflections (volcanic margin) or rotated continental blocks (non-volcanic margin). Finally, a comparison of the amount of brittle deformation in the upper crust to the overall crustal thinning reveals a considerable ‘extension discrepancy’ that is not explained by the rifting–drifting model. An alternative model for the evolution of the Levant margin may be DDS, derived by an upwelling, divergent mantle flow (Fig. 12) (Kuzsnir & Karner 2007). Mantle flow may explain thinning by removal of the lower crust and part of the upper crust (Hirsch et al. 1995). However, it also requires decoupling and shearing between the displaced crustal units (Reston 2007). Some brittle deformation in the remaining upper crust is compatible with the DDS model, assuming a pause of the flow field prior to its reaching the surface (Kuzsnir & Karner 2007). Intrusions from the mantle flow may explain the higher velocity at the lower part of the crust in the centre of the basin. Early
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN
Mesozoic intrusive and extrusive rocks found in wells along the Levant margin further indicate upwelling of mantle material. A modern analogue for the Mesozoic Levant Basin is the northern part of the Red Sea. In this rift basin, the DDS model may be similarly applied to explain crustal thinning in the absence of sea-floor spreading. Cochran & Karner (2007) speculated that, unlike the southern Red Sea, an oceanic spreading centre may not develop in its northern part. Similarly, in the Levant Basin, rifting activity stopped at an early magmatic stage (Gardosh & Druckman 2006). It is unclear why seafloor spreading that took place further to the north and west (Robertson 1998, 2007) did not occur in the Levant area. Possibly, it is related to variations in lithospheric properties or in the pattern of deep mantle convection. In any case, given the above considerations, the Mesozoic Levant Basin may be considered as a ‘failed breakup basin’ (Kusznir & Karner 2007) somewhat similar to the ‘aborted rift system’ proposed by Hirsch et al. (1995). Application of the DDS model implies that the amount of drifting in the Levant Basin was limited. In this framework, breakup from the Gondwanian craton and northward motion of the Eratosthenes block is not required. The seamount may constitute a Mesozoic volcanic edifice rooted within the same modified continental crust that is found in other parts of the basin (Zverev & Ilinsky 2005). The DDS model may be applied also to the areas of the Galilee and southern Lebanon. The reduced thickness (c. 14 –16 km) and high velocity of the crystalline crust and the higher Bouguer gravity anomaly in this area (Fig. 2c) bear similarity to the Levant Basin. The modification of the northern Levant crust during Tethyan rifting (Asher and Hermon basins) is an alternative model to accretion of oceanic terranes that is related to the closure of the Palaeo-Tethys, as proposed by Ben-Avraham & Ginzburg (1990). Well data from southern and central Israel show that during the Late Palaeozoic, Triassic and Early Jurassic rifting phases the Levant region was dominated by continental to shallow-marine environments. The seismic character of the marine reflection lines and the lithology of the Middle Jurassic strata in the offshore Yam West-1 well suggest that shallow marine conditions prevailed also during rifting in the area of the Levant Basin. On the other hand, the attenuation and thinning of the crust within the basin, either by rifting–drifting or DDS, should have resulted with subsidence and increase in water depth. At the present state of knowledge, this apparent disagreement is not resolved and requires further analysis of the basin’s subsidence history.
31
Finally, Robertson (2007) noted that none of the Tethyan-rifted margins that developed in northern Gondwana show the ideal margin characteristics that are observed in the modern oceans. The deep structure of the Levant area and its complex evolution further demonstrates the diversity of margin forming processes that remain to be studied in both recent and ancient examples. The authors are grateful to Y. Mimran and A. Honigstein from the Israeli Petroleum Commissioner office for allowing the use the offshore seismic data and for supporting this study. The comments and suggestions of P. Bentham, D. Frizon de Lamotte, and an ammoniums reviewer are greatly appreciated. The authors thank M. Rybakov for his contribution to this study.
References Aal, A. A., El Barkooky, A., Gerrits, M., Meyer, H., Schwander, M. & Zaki, H. 2000. Tectonic evolution of the eastern Mediterranean Basin and its significance for hydrocarbon prospectivity in the ultra-deepwater of the Nile Delta. The Leading Edge, 19, 1086– 1102. Alsharhan, A. S. & Nairn, A. E. M. 1997. Sedimentary Basins and Petroleum Geology of the Middle East. Elsevier, Amsterdam, 843. Alsharhan, A. S. & Salah, M. G. 1996. Geologic setting and hydrocarbon potential of north Sinai, Egypt. Bulletin of Canadian Petroleum Geology, 44, 615– 631. Al-Youssef, W. & Ayed, H. 1992. Evolution of Upper Palaeozoic sequences and the application of stratigraphy as a tool for hydrocarbon exploration in Syria. Egyptian General Petroleum Company 11th Exploration Seminar, Cairo, 2, 636–657. Bein, A. & Gvirtzman, G. 1977. A Mesozoic fossil edge of the Arabian plate along the Levant coastline and its bearing on the evolution of the eastern Mediterranean. In: Biju-Duval, B. & Montadert, L. (eds) Structural History of the Mediterranean Basins. Editions Technip, Paris, 95–110. Ben-Avraham, Z. & Ginzburg, A. 1990. Displaced terranes and crustal evolution of the Levant and the Eastern Mediterranean. Tectonics, 9, 613–622. Ben-Avraham, Z., Shoham, Y. & Ginzburg, A. 1976. Magnetic anomalies in the Eastern Mediterranean and the tectonic setting of the Eratosthenes Seamount. Geophysical Journal of the Royal Astronomical Society, 45, 105–123. Ben-Avraham, Z., Ginzburg, A., Makris, J. & Eppelbaum, L. 2002. Crustal structure of the Levant Basin, eastern Mediterranean. Tectonophysics, 346, 23–43. Bentham, P., Hanbal, I., Cotton, J., Longacre, M. B. & Edwards, R. 2007. Crustal structure and early opening of the eastern Mediterranean Basin – key observations from offshore northern Egypt. 3rd North African/Mediterranean Petroleum and Geosciences Conference and Exhibition, Abstract, 4. Beydoun, Z. R. 1988. The Middle East: Regional Geology and Petroleum Resources. Scientific Press Ltd., Beaconsfield.
32
M. A. GARDOSH ET AL.
Biju-Duval, B. & Dercourt, J. 1980. Les basins de la Mediterranee orientale representent-ils les restes d’un domainoceanique, la mesogee, ouvert au Mesozoique et distinct de la Tethys? Bulletin de la societe Geologique de France, 22, 43– 60. Boote, D. R. D., Clark-Lowes, D. D. & Traut, W. 1998. Palaeozoic petroleum systems of North Africa. In: MacGregor, D. S., Moody, R. T. J. & Clark-Lowes, D. D. (eds) Petroleum Geology of North Africa. Geological Society, London, Special Publications, 132, 7– 68. Brew, G., Barazangi, M., Al-Malekh, K. & Sawaf, T. 2001. Tectonic and geologic evolution of Syria. GeoArabia, 6, 573–616. Bruner, I. 1991. Investigation of the subsurface in the northern Negev, Israel, using seismic reflection techniques. PhD thesis, Tel Aviv University. Buchbinder, B. & le Roux, J. P. 1993. Inner platform cycles in the Ardon Formation: Lower Jurassic, southern Israel. Israel Journal of Earth Sciences, 42, 1– 16. Cochran, J. R. 2005. Northern Red Sea; nucleation of an oceanic spreading center within a continental rift. Geochemistry, Geophysics, Geosystems, 6, doi: 10.1029/2004GC000826. Cochran, J. R. & Karner, G. D. 2007. Constraints on the deformation and rupturing of continental lithosphere of the Red Sea: the transition from rifting to drifting. In: Karner, G. D., Manatschal, G. & Pinheiro, L. M. (eds) Imaging, Mapping and Modelling Continental Lithosphere Extension and Breakup. Geological Society, London, Special Publications, 282, 265– 289. Cohen, Z. 1976. Early Cretaceous buried canyon: influence on accumulation of hydrocarbons in the Helez oil field, Israel. AAPG Bulletin, 60, 108– 114. Cohen, Z., Flexer, A. & Kaptsan, V. 1988. The Pleshet basin, a newly discovered link in the peripheral chain of basins of the Arabian craton. Journal of Petroleum Geology, 11, 403– 414. Davis, M. & Kusznir, N. J. 2004. Depth-dependent lithospheric stretcing at rifted continental margins. In: Karner, G. D. (ed.) Proceedings of NSF Rifted Margins Theoretical Institute. Colombia University Press, New York, 92–136. Derin, B. 1979. Helez Deep 1A Stratigraphic Log. The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 1/79. Derin, B. 1995. David 1 stratigraphical log. In: Derin, B. & Gilboa, Y. (eds) Drilling and Completion Report 1: Final Geological Report. Naphtha Israel Corp. Ltd., Tel Aviv, 28. Derin, B. & Gerry, E. 1979. Devora 2A, Stratigraphic Log. The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 22/79. Derin, B. & Gerry, E. 1986a. Halal-1, Reexamination of the Triassic/EarlyJurassic (2950– 4311 m). The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 2/86, 12. Derin, B. & Gerry, E. 1986b. Microbiostratigraphy and environmental analysis of the early Jurassic (3169–4520 m) in the Ramallah-1 well and correlation with Pleshet-1 and Gaash-2. The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 8/86, 7.
Derin, B., Gerry, E., Lipson, S. & Weiller, Y. 1976. Halal-1. The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 11/76. Derin, B., Gerry, E. & Lipson, S. 1980. Ramalla-1, segment 4510– 6355 m, stratigraphic log (Early Triassic to Early Jurassic). The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 1/80. Derin, B., Gerry, E., Lipson, S. & Peri, M. 1985. Pleshet-1: Stratigraphic report. The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 2/85. Derin, B., Lipson, S. & Gerry, E. 1981. Gaash-2: Stratigraphic log (early Triassic-Quaternary). The Israel Institute of Petroleum and Energy Report 14/81, 2. Derin, B., Lipson, S. & Gerry, E. 1982. Asher Atlit 1, Stratigraphic Log. The Israel Institute of Petroleum and Energy Micropaleontology Laboratory, Tel Aviv, 7/82. Dewey, J. F., Pitman, W. C. III, Ryan, W. B. F. & Bonin, J. 1973. Plate tectonics and the evolution of the Alpine system. Geological Society of America Bulletin, 84, 3137– 3180. Domzhalski, W. 1986. Review and additional magnetic data in Israel and adjoining areas. Oil Exploration (Investment) Ltd, Tel Aviv. Driscoll, N. W. & Karner, G. D. 1998. Lower crustal extension across the Northern Carnarvon basin, Australia: evidence for an eastward dipping detachment. Journal of Geophysical Research, 103, 4975– 4991. Druckman, Y. 1974. The stratigraphy of the Triassic sequence in southern Israel. Israel Geologic Survey Bulletin, 64, 92. Druckman, Y. 1977. Differential subsidence during the deposition of the Lower Jurassic Ardon Formation in western Jordan, southern Israel and northern Sinai. Israel Journal of Earth Sciences, 26, 45– 54. Druckman, Y. 1981. Comments on the structural reversal model as a factor of the geological evolution of Israel. Israel Journal of Earth Sciences, 30, 44– 48. Druckman, Y. 1984. Evidence for Early-Middle Triassic faulting and possible rifting from the Helez Deep borehole and the coastal plain of Israel. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 203– 212. Druckman, Y., Hirsch, F. & Weissbrod, T. 1982. The Triassic of the southern margin of the Tethys. Geologische Rundschau, 71, 919– 936. Druckman, Y., Gill, D., Fleischer, L., Gelbermann, E. & Wolff, O. 1995. Subsurface geology and structural evolution of the northwestern Negev, southern Israel. Israel Journal of Earth Science, 44, 115–136. Dvorkin, A. & Kohn, B. P. 1989. The Asher Volcanics, northern Israel: petrography, mineralogy and alteration. Israel Journal of Earth Sciences, 38, 105–123. Eyal, Y. & Reches, Z. 1983. Tectonic analysis of the Dead Sea rift region since the Late Cretaceous based on mesostructures. Tectonics, 2, 167 –185. Eyal, Y., Eyal, M. & Kroner, A. 1991. Geochronology of the Elat terrain metamorphic basement and its
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN implications for crustal evolution of the NE part of the Arabian –Nubian Shield. Israel Journal of Earth Sciences, 40, 5– 16. Fleischer, L. & Varsahvsky, A. 2002. A lithostratigraphic data base of oil and gas wells drilled in Israel. The Ministry of National Infrastructures, Oil and Gas Unit Report OG/9/02, 280. Flexer, A., Hirsch, F. & Hall, K. 2005. Tectonic evolution of Israel. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant, Vol. II, the Levantine Basin and Israel. Historical Production-Hall, Jerusalem, 523 –537. Folkman, Y. & Ben-Gai, Y. 2004. The Jonah buried Seamount: intrusive structure in the southeastern Levant Basin, offshore Israel. Israel Geological Society Annual Meeting, Hagoshrim, Abstract, 29. Freund, R., Garfunkel, Z., Goldberg, M., Weissbrod, T. & Derin, B. 1970. The shear along the Dead Sea rift. Philosophical Transactions of the Royal Society, London, A267, 105– 127. Freund, R., Goldberg, M., Weissbrod, T., Druckman, Y. & Derin, B. 1975. The Triassic –Jurassic structure of Israel and its relation to the origin of the eastern Mediterranean. Geological Survey of Israel Bulletin, 65. Gardosh, M. 2002. The sequence stratigraphy and petroleum systems of the Mesozoic, southeastern Mediterranean continental margin. PhD thesis, Tel Aviv University. Gardosh, M. & Druckman, Y. 2006. Seismic stratigraphy, structure and tectonic evolution of the Levantine Basin, offshore Israel. In: Robertson, A. H. F. & Mountrakis, D. (eds) Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, 201–227. Gardosh, M., Druckman, Y., Buchbinder, B. & Rybakov, M. 2006. The Levant Basin offshore Israel: stratigraphy, structure, tectonic evolution and implications for hydrocarbon exploration. Geophysical Institute of Israel, 429/218/06, 1 –119. Gardosh, M., Druckman, Y., Buchbinder, B. & Rybakov, M. 2008. The Levant Basin offshore Israel: stratigraphy, structure, tectonic evolution and implications for hydrocarbon exploration – revised edition. Geological Survey of Israel Report GSI/4/ 2008, 1–119. Garfunkel, Z. 1981. Internal structure of the Dead Sea leaky transform (rift) in relation to plate kinematics. Tectonophysics, 80, 81– 108. Garfunkel, Z. 1988. The pre-Quaternary geology of Israel. In: Yom-Tov, Y. & Tchernov, E. (eds) The Zoogeography of Israel. Jung Publishers, Dordrecht, 7–34. Garfunkel, Z. 1989. Tectonic setting of Phanerozoic magmatism in Israel. Israel Journal of Earth Sciences, 38, 51–74. Garfunkel, Z. 1997. The history and formation of the Dead Sea basin. In: Niemi, T. M., Ben-Avraham, Z. & Gat, J. R. (eds) The Dead Sea, the Lake and its Setting. Oxford University Press, Oxford, 36–56. Garfunkel, Z. 1998. Constrains on the origin and history of the eastern Mediterranean basin. Tectonophysics, 298, 5 –35.
33
Garfunkel, Z. 2002. Early Paleozoic sediments of NE Africa and Arabia: Products of continental-scale erosion, sediments transport, and deposition. Israel Journal of Earth Sciences, 51, 135– 156. Garfunkel, Z. 2004. Origin of the Eastern Mediterranean basin: a reevaluation. Tectonophysics, 391, 11–34. Garfunkel, Z. & Derin, B. 1984. Permian–Early Mesozoic tectonism and continental margin formation in Israel and its implications to the history of the eastern Mediterranean. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 18– 201. Gaulier, J. M., Le Pichon, X. et al. 1988. Seismic study of the crust of the northern Red Sea and Gulf of Suez. Tectonophysics, 153, 55– 88. Gelbermann, E. 1995. Revised seismic structural and isopach maps, central and southern Israel. Institute for Petroleum Research and Geophysics Report 201/ 22/889. Gelbermann, E. & Kemmis, J. L. 1987. Triassic to Late Cretaceous tectonic alignments in southern coastal plain in the northwestern Negev. Israel Geologic Society Annul Meeting, Mizpe Ramon, Abstract, 39–40. Ginzburg, A. & Folkman, Y. 1980. The crustal structure between the Dead Sea rift and the Mediterranean sea. Earth and Planetary Science Letters, 51, 181– 188. GIVOT OLAM OIL LTD. 1995. Meged #2: Well completion report. Givot Olam Oil Ltd, Israel, 250. Goldberg, M. & Friedman, G. M. 1974. Paleoenvironments and paleogeographic evolution of the Jurassic system in southern Israel. Israel Geologic Survey Bulletin, 61, 44. Goldberg, M., Hirsch, F. & Mimran, Y. 1981. The Jurassic sequence of Mt. Hermon. Israel Geological Society, Annual Meeting, Katzerin, Proceedings, 14–19. Guiraud, R. & Bosworth, W. 1999. Phanerozoic geodynamic evolution of northern Africa and northwestern Arabian platform. Tectonophysics, 315, 73–108. Gvirtzman, G. & Klang, A. 1972. A structural and depositional hinge-line along the coastal plain of Israel evidenced by magneto-tellurics. Geological Survey of Israel Bulletin, 55, 18. Gvirtzman, G. & Steinitz, G. 1983. The Asher Volcanics – an Early Jurassic event in northern Israel. Israel Geological Survey Current Research, 28–33. Gvirtzman, G. & Weissbrod, T. 1984. The Hercynian geanticline of Helez and the Late Paleozoic history of the Levant. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 177– 186. Gvirtzman, G., Teleman, E. & Klang, A. 1984. Gaash-2: Geologic completion report. Oil Exploration (Investment), Tel Aviv, Report 84/33, 37. Harms, J. C. & Wray, J. L. 1990. Nile Delta. In: Said, R. (ed.) The Geology of Egypt. A. A. Balkema, Rotterdam, 329– 343. Hirsch, F. 2005. The Jurassic of Israel. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini,
34
M. A. GARDOSH ET AL.
Ch. & Flexer, A. (eds) Geological Framework of the Levant. Vol. II The Levantine Basin and Israel. Historical Production-Hall, Jerusalem, 361–391. Hirsch, F., Flexer, A., Rosenfeld, A. & Yellin-Dror, A. 1995. Palinspastic and crustal setting of the eastern Mediterranean. Journal of Petroleum Geology, 18(2), 149– 170. Hirsch, F., Bassoullet, J. P. et al. 1998. The Jurassic of the southern Levant: biostratigraphy, palaeogeography and cyclic event. In: Crasquin-Soleau, S. & Barrier, E. (eds) Peri-Tethys Memoir 4: Epicratonic Basins of Peri-Tethyan Platforms. Memoires du Museum National d’Histoire Naturelle, Paris, 179, 213– 235. Jenkins, D. A. 1990. North and central Sinai. In: Said, R. (ed.) The Geology of Egypt. A. A. Balkema, Rotterdam, 361– 380. Kempler, D. 1998. Eratosthenes seamount: The possible spearhead of incipient continental collision in the Eastern Mediterranean. In: Robertson, A. H. F., Emeis, K. C., Richter, C. & Camerlenghi, A. (eds) Proceedings of the Ocean Drilling Program, Scientific Results, 160, 709– 721. Klitzsch, E. 1981. Lower Paleozoic rocks of Libya, Egypt and Sudan. In: Holland, C. H. (ed.) Lower Paleozoic of the Middle East, Eastern and Southern Africa and Antarctica. Wiley, Manchester, 131– 163. Kohn, B. P., Eyal, M. & Feinstein, S. 1992. A major Late Devonian– Early Cretaceous (Hercynian) thrmotectonic event at the NW margin of the Arabian– Nubian shield: evidence from zircon fission track dating. Tectonics, 11, 1018– 1027. Kohn, B. P., Lang, B. & Steinitz, G. 1993. Ar40/Ar39 dating of the Atlit 1 volcanic sequence, northern Israel. Israel Journal of Earth Sciences, 42, 17–23. Krenkel, E. 1924. Der Syrische Bogen. Zentralblatt Mineralogie, 9, 274 –281; and 10, 301–313. Kusznir, N. J. & Karner, G. D. 2007. Continental lithospheric thinning and breakup in response to upwelling divergent mantle flow: application to the Woodlark, Newfoundland and Iberia margins. In: Karner, G. D., Manatschal, G. & Pinheiro, L. M. (eds) Imaging, Mapping and Modelling Continental Lithosphere Extension and Breakup. Geological Society, London, Special Publications, 282, 389– 419. Kusznir, N. J., Hunsdale, R. & Roberts, A. M. 2004. Timing of depth-dependent lithosphere stretching on the S. Lofton rifted margin offshore Norway: pre-breakup or post-breakup? Basin Research, 16, 279–296. Le Pichon, X., Bergerat, F. & Roulet, M. J. 1988. Plate kinematics and tectonics leading to the Alpine belt formation; a new analysis. Geologic Society of America Special Publication, 218, 111–131. Makris, J., Ben-Avraham, Z. et al. 1983. Seismic refraction profiles between Cyprus and Israel and their interpretation. Geophysical Journal of the Royal Astronomical Society, 75, 575–591. Mart, Y. & Robertson, A. H. F. 1998. Eratosthenes Seamount: an oceanic yardstick recording the Late Mesozoic– Tertiary geological history of the eastern Mediterranean. In: Robertson, A. H. F., Emeis,
K. C., Richter, C. & Camerlenghi, A. (eds) Proceedings of the Ocean Drilling Program, Scientific Results, 160, 701 –708. Martinez, F. & Cochran, J. R. 1988. Structure and tectonics of the Red Sea: catching a continental margin between rifting and drifting. Tectonophysics, 150, 1– 32. Montadert, L., Roberts, D. G., de Charpal, O. & Guennoc, P. 1979. Rifting and subsidence of the northern margin of Bay of Biscay. Initial Report Deep Sea Drilling Project, 48, 1025–1060. Moustafa, A. R. & Khalil, M. H. 1990. Structural characteristics and tectonic evolution of north Sinai fold belt. In: Said, R. (ed.) The Geology of Egypt. A. A. Balkema, Rotterdam, 381–389. Mouty, M. 1997. Le Jurassique de la chaine de Palmyrides (Syrie centrale). Bulletin de la societe Geologique de France, 168, 181– 186. Netzeband, G. L., Gohl, K., Hu¨bscher, C. P., Ben-Avraham, Z., Dehghani, G. A., Gajewski, D. & Liersch, P. 2006. The Levantine Basin – crustal structure and origin. Tectonophysics, 418, 167– 188. Planke, S., Symonds, P. A., Alvestad, E. & Skogseid, J. 2000. Seismic volcanostratigraphy of large-volume basaltic extrusive complexes on rifted margins. Journal of Geophysical Research, 105(B8), 19335–19351. Reston, T. J. 1993. Evidence for extensional shear zones in the mantle, offshore Britain, and their implications for the extension of the continental Lithosphere. Tectonics, 12, 492 –506. Reston, T. J. 2007. The formation of non-volcanic rifted margins by the progressive extension of the lithosphere: the example of the West Iberian margin. In: Karner, G. D., Manatschal, G. & Pinheiro, L. M. (eds) Imaging, Mapping and Modelling Continental Lithosphere Extension and Breakup. Geological Society, London, Special Publications, 282, 77–110. Roberts, G. & Peace, D. 2007. Hydrocarbon plays and prospectivity of the Levantine Basin, offshore Lebanon and Syria from modern seismic data. GeoArabia, 12, 99–124. Robertson, A. H. F. 1998. Mesozoic-Tertiary tectonic evolution of the easternmost Mediterranean area: integration of marine and land evidence. In: Robertson, A. H. F., Emeis, K. C., Richter, C. & Camerlenghi, A. (eds) Proceedings of the Ocean Drilling Program, Scientific Results, 160, 723–782. Robertson, A. H. F. 2000. Mesozoic-Tertiary tectonicsedimentary evolution of a south Tethyan oceanic basin and its margins in southern Turkey. In: Bozkurt, E., Winchester, J. A. & Piper, J. D. A. (eds) Tectonics and Magmatism in Turkey and the Surrounding Area. Geological Society, London, Special Publications, 173, 97–138. Robertson, A. H. F. 2006. Sedimentary evidence from the south Mediterranean region (Sicily, Crete, Peloponnese, Evia) used to test alternative models for the regional tectonic setting of Tethys during Late Palaeozoic–Early Mesozoic time. In: Robertson, A. H. F. & Mountrakis, D. (eds) Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, 91–154.
TETHYAN RIFTING IN THE LEVANT BASIN AND MARGIN Robertson, A. H. F. 2007. Overview of tectonic settings related to the rifting and opening of Mesozoic ocean basins in the Eastern Tethys: Oman, Himalayas and Eastern Mediterranean regions. In: Karner, G. D., Manatschal, G. & Pinheiro, L. M. (eds) Imaging, Mapping and Modelling Continental Lithosphere Extension and Breakup. Geological Society, London, Special Publications, 282, 325– 338. Robertson, A. H. F. & Dixon, J. E. 1984. Introduction: aspects of the geological evolution of the eastern Mediterranean. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 1– 74. Robertson, A. H. F. & Mountrakis, D. 2006. Tectonic development of the Eastern Mediterranean region: an introduction. In: Robertson, A. H. F. & Mountrakis, D. (eds) Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, 1– 9. Robertson, A. H. F. & Xenophontos, C. 1993. Development of concepts concerning the Troodos ophiolite and adjacent units in Cyprus. In: Prichard, H. M., Alabaster, T., Harris, N. B. W. & Neary, C. R. (eds) Magmatic Processes and Plate Tectonics. Geological Society, London, Special Publications, 76, 85–119. Robertson, A. H. F., Dixon, J. E. et al. 1996. Alternative tectonic model for the Late Paleozoic–Early Tertiary development of Tethys in the Eastern Mediterranean. In: Morris, A. & Tarling, D. H. (eds) Paleomagnetism and tectonics of the Mediterranean region. Geological Society, London, Special Publications, 105, 239– 263. Rybakov, M. & Al-Zoubi, A. 2005. Bouguer gravity map of the Levant – a new compilation. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant, Vol. II, the Levantine Basin and Israel, 539–542. Rybakov, M., Fleischer, L. & Goldshmidt, V. 1995. A new look at the Hebron magnetic anomaly. Israel Journal of Earth Science, 44, 41–49. Rybakov, M., Goldshmit, V. & Rotstein, Y. 1997. New regional gravity and magnetic maps of the Levant. Geophysical Research Letters, 24, 33–36. Rybakov, M., Goldshmit, V., Fleischer, L. & Ben-Gai, Y. 2000. 3-D gravity and magnetic interpretation for the Haifa Bay area (Israel). Journal of Applied Geophysics, 44, 353– 367. Segev, A. 2005. Phanerozoic magmatic activity associated with vertical motions in Israel and adjacent countries. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant, Vol. II, the Levantine Basin and Israel, 553– 557. Segev, A. & Eshet, Y. 2003. Significance of Rb/Sr age of Early Permian volcanics, Helez Deep 1A borehole, central Israel. Africa Geoscience Review, 10, 333–345. Sengor, A. M. C. & Yilmaz, Y. 1981. Tethyan evolution of Turkey: a plate tectonic approach. Tectonophysics, 75, 181–241.
35
Salhov, S. 1996. Gevim-1 composite log. Isramco Inc. Sawaf, T., Brew, G., Litak, R. & Barazangi, M. 2001. Geologic evolution of the intraplate Palmyride basin and Euphrates fault system, Syria. In: Ziegler, P. A., Cavazza, W., Robertson, A. H. F. & CrasquinSoleau, S. (eds) Peri-Tethys Memoir 6: Peri-Tethyan Rift/Wrench Basins and Passive Margins. Memoires du Museum National d’Histoire Naturelle, Paris, 186, 441– 467. Stampfli, G. M. & Borel, G. D. 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrones. Earth and Planetary Science Letters, 196, 17– 33. Stampfli, G. M., Mosar, J., Favre, P. A., Pillevuit, A. & Vannay, J. C. 2001. Permo-Mesozoic evolution of the western Tethyan realm: the Neotethys EastMediterranean Basin connection. In: Ziegler, P. A., Cavazza, W., Robertson, A. H. F. & CrasquinSoleau, S. (eds) Peri-Tethys Memoir 6: Peri-Tethyan Rift/Wrench Basins and Passive Margins. Memoires du Museum National d’Histoire Naturelle, Paris, 186, 51–108. Steinitz, G. 1980. Rb –Sr age determinations on basement rocks from Helez Deep 1A well. Geological Survey of Israel Report MM/1/80, 1 –10. Stern, R. J. 1994. Arc assembly and continental collision in the Neoproterozoic E African orogen: implications for the consolidation of Gondwanaland. Annual Review of Earth and Planetary Sciences, 22, 319– 351. Ten Brink, U. 1987. Quantitative basin analysis with applications to the Israeli continental margin. Oil Exploration (Investments) Report, 24. Vidal, N., Alvarez-Marron, J. & Klaeschen, D. 2000. Internal configuration of the Levantine Basin from seismic reflection data (eastern mediterranean). Earth and Planetary Science Letters, 180, 77–89. Walley, C. D. 1998. Some outstanding issues in the geology of Lebanon and their importance in the tectonic evolution of the Levantine region. Tectonophysics, 298, 37–62. Walley, C. H. 2001. The Lebanon passive margin and the evolution of the Levantine Neo-Tethys. In: Ziegler, P. A., Cavazza, W., Robertson, A. H. F. & Crasquin-Soleau, S. (eds) Peri-Tethys Memoir 6: Peri-Tethyan Rift/Wrench Basins and Passive Margins. Memoires du Museum National d’Histoire Naturelle, Paris, 186, 407 –439. Weber, M., Abu-Ayyash, K. et al. 2004. The crustal structure of the Dead Sea Transform. Geophysical Journal International, 156, 655 –681. Weissbrod, T. 1981. The Paleozoic of Israel and adjacent countries. Geological Survey of Israel, Report MP 600/81, 276 (in Hebrew with English abstract). Weissbrod, T. 2005. The Paleozoic in Israel and environs. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant, Vol. II, The Levantine Basin and Israel. Historical Production-Hall, Jerusalem, 283– 316. Wolfart, R. 1981. Lower Paleozoic rocks of the Middle East. In: Holland, C. H. (ed.) Lower Paleozoic of
36
M. A. GARDOSH ET AL.
the Middle East, Eastern and Southern Africa and Antarctica. Chichester, 5– 30. ZION OIL AND GAS INC. 2005. Ma’anit-1 (Re-Entry). Zion Oil and Gas Inc., Israel. Zverev, S. M. & Ilinsky, D. A. 2000. Deep structure and possible nature of the Eratosthenes Seamount (Eastern Mediterranean). Geotektonika, 4, 67–84 (in Russian).
Zverev, S. M. & Ilinsky, D. A. 2005. The deep structure of Eratosthenes Seamount from seismic refraction data. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant, Vol. II, the Levantine Basin and Israel. Historical ProductionHall, Jerusalem, 73– 111.
Structural setting and tectonic evolution of North Sinai folds, Egypt ADEL R. MOUSTAFA Department of Geology, Faculty of Science, Ain Shams University, Cairo 11566, Egypt (e-mail:
[email protected]) Abstract: Detailed study of outcrop and subsurface data of North Sinai indicates the existence of a NE–SW oriented region with very large asymmetric folds lying between the Nile Delta hinge zone and the Sinai hinge belt. The steep southeastern flanks of these folds are often associated with highangle reverse faults. These folds continue northeastward in the Naqb Desert toward the Dead Sea transform. The North Sinai folds represent inversion anticlines formed by inversion of Mesozoic (Jurassic and probably Cretaceous) extensional basins during Late Cretaceous to pre-Miocene times. Mesozoic extension is related to the divergence between Afro-Arabia and Eurasia and opening of the Neotethys whereas inversion is related to the convergent movement between these two plates. The acme of inversion was at Campanian time. The central Sinai hinge belt is a zone of ENE–WSW oriented, right-lateral strike–slip faults that separate the folded area to the north from the tectonically stable area of central and southern Sinai. It responded to the convergent movement between Afro-Arabia and Eurasia by dextral transpression on the faults. Later reactivation of the eastern edges of these faults by drag on the west side of the Dead Sea transform took place in post-Miocene to Recent times.
The Sinai Peninsula lies at the northeastern part of the African plate. It is bounded on the east by the Dead Sea transform, on the west by the Gulf of Suez rift and the Suez Canal, on the north by the Eastern Mediterranean, and on the south by the Red Sea (Fig. 1). Precambrian crystalline basement rocks are exposed in the southernmost part of Sinai forming the Sinai Massif, which reaches a height of 2641 m above sea level. The northern part of these Precambrian basement rocks is unconformably overlain by Phanerozoic sedimentary rocks that have gentle northward tilt and become younger in age toward the north. These Phanerozoic sediments range from Palaeozoic above the Precambrian basement, to Pliocene in the subsurface of the northern coastal area of Sinai (Figs 1, 2 & 3). Cretaceous to Cenozoic (Middle Eocene) carbonate rocks form the surface of the Tih Plateau in central Sinai. A number of large hillocks exist in the northern part of Sinai between the Tih Plateau and the Mediterranean Sea and represent prominent highs within the plains covered mostly by Quaternary alluvium and sand dunes. G. (Gebel, mountain) Yelleg, G. Halal and G. Maghara are the most prominent of these hillocks and are 1090, 890 and 735 m above sea level, respectively. The three hills represent three large anticlines among other smaller anticlines in northern Sinai. Similar folds exist farther to the NE in the Naqb Desert, which, along with those in North Sinai, represent a NE– SW oriented fold system. This fold system is a segment of Krenkel’s (1925) ‘Syrian Arc’ that extends from the Palmyra folds of Syria into the desert west of the Nile, passing through North Sinai.
The present work is intended to throw light on the structural setting and tectonic evolution of the North Sinai folds. This work is based on the results of detailed surface geological mapping at scales of 1:50 000 and 1:20 000 carried out by the author and his graduate students since 1990. In addition, subsurface [borehole and two-dimensional (2D) seismic] data have also been used. North Sinai has been lacking detailed structural studies for a long period of time. Early attempts to study the structures in this area date back to the Early 20th century (Moon & Sadek 1921; Sadek 1928) as well as Shata (1959). The first detailed structural study in Sinai is perhaps that of Bartov et al. 1980. Moustafa & Khalil (1989) carried out detailed photogeological study of the North Sinai folds, which was followed by a program of detailed geological field mapping of the exposed structures. The present work is based on these detailed structural studies. Information about the subsurface of North Sinai includes borehole and seismic data. The latter was only acquired in the late eighties and early nineties as part of the limited hydrocarbon exploration in North Sinai. The onshore area of North Sinai is covered by few 2D seismic sections compared to the northern offshore area, where more seismic data were acquired. The North Sinai folds have NE–SW orientation and include both large (several tens of kilometres long) folds; for example, Gebels Maghara, Halal and Yelleg folds; and small (several kilometres long) folds; for example, Gebels Libni, Minsherah and Falig folds. These folds have been the subjects of several studies (see Moustafa & Khalil 1994 for list).
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 37– 63. DOI: 10.1144/SP341.3 0305-8719/10/$15.00 # The Geological Society of London 2010.
38
A. R. MOUSTAFA
Fig. 1. Simplified geological map of Sinai and the Naqb Desert after Geological Survey of Egypt (1981) and Geological Survey of Israel (2000) with the main tectonic features. Major folds in the area are Maghara (M), Yellleg (Y), Halal (H), Ramon (R), Kurnub (K) and Qatan (Q).
Concerning the geometry of the North Sinai folds, most workers agree that the large folds are highly asymmetrical, with gentle northwestern flanks, dipping at about 5–208NW and steeper southeastern flanks that are vertical or overturned in places. Some workers mapped reverse faults dipping at angles exceeding 458 in some of these folds (e.g. Shata 1959; Al-Far 1966; Bartov et al. 1980; Moustafa & Khalil 1989; Moustafa & Yousif 1990). On the other hand, Abdel Aal et al.
(1992) alleged the existence of low-angle thrusts in North Sinai based on observations of few seismic reflection profiles. All the reverse faults mapped on the surface in North Sinai have a consistent northwestward direction of dip indicating southeastward vergence. Moon & Sadek (1921) and Sadek (1928) mentioned that the North Sinai folds are arranged in distinct NE-oriented parallel lines (axes). Shata (1959) identified a gently folded area in north-central Sinai and a strongly
NORTH SINAI FOLDS
39
Fig. 2. Simplified geological cross section along the Sinai Peninsula. Top Precambrian basement in the subsurface of North Sinai is after Rybakov & Segev (2004) and Abd El-Gawad & Ibraheem (2005). Rock units symbols stand for Palaeozoic (Pz), Mesozoic (Mz), Jurassic (J), Cretaceous (K), Paleocene (Tp) and Eocene (Te) rocks. A and T designate the sense of strike–slip movement on some faults (away from and toward the observer respectively). See Figure 1 for location.
folded area in North Sinai. The two areas are elongated in a NE–SW direction and are separated by a 20-km wide fractured area, which he called the Sinai hinge belt. The North Sinai folds and their associated reverse faults have been explained to be due to pureshear deformation involving tangential compressive stresses (e.g. Moon & Sadek 1921; Shukri 1954; Said 1962; Youssef 1968; Abdel Aal et al. 1992), simple-shear deformation model by a dextral shear couple (Moustafa & Khalil 1989, 1990; Moustafa & Yousif 1990; Moustafa et al. 1991), or a vertical stress model of basement uplifting (Shata 1959). Subsurface data of offshore North Sinai revealed the role of inversion tectonics in forming some structures similar to those in the onshore area (Ayyad & Darwish 1996). Inspired by this model, those authors also proposed the same mechanism for the onshore structures. Most workers agree that folding in northern Sinai took place over a long period of time. Shukri (1954) mentioned that these folds rose from the
bottom of the Cretaceous sea and have been intermittently submerged and emerged in later times up to the Late Oligocene. Shata (1959) proposed that folding acted throughout a long period of time during the Jurassic throughout most of the Early Miocene. Said (1962) stated that folding started as early as the Cenomanian and continued intermittently throughout the latest Cretaceous. He also mentioned that strong folding at the end of the Maastrichtian brought many of the folds above sea level and final uplift of North Sinai took place in post-Eocene time. Bartov et al. (1980) concluded that the Araif El Naqa fold started to develop during the Coniacian and continued developing in the Santonian and up to the Late Campanian. Faulting in this area is evident in two phases during the Senonian and in post-Early Neogene time (post 21 Ma). Sadek (1928) mentioned that the North Sinai folding took place in the Oligocene or Early Miocene time, whereas Farag & Shata (1954) mentioned that folding in G. El Minsherah was after the Early Miocene.
40
A. R. MOUSTAFA
Fig. 3. Stratigraphic sections based on borehole data and exposed stratigraphic section of G. Maghara showing the changes in thickness of Jurassic rocks across the Sinai hinge belt and of Cenozoic rocks across the Nile Delta hinge zone. Symbols designate: Palaeozoic (PZ), Triassic (Tr), Jurassic (J), Lower Cretaceous (Kl), Aptian– Albian (AA), Cenomanian (Kc), Turonian (Kt), Senonian (Ksn), Paleocene (Tp), Eocene (Te), Oligocene (Tineh, Tn; Qantara, Qt), Miocene (Sidi Salem, SS; Qawasim, QS; Abu Madi, AM) and Pliocene formations (Kafr El Sheikh, KS; El Wastani, WS; Mit Ghamr, MG).
NORTH SINAI FOLDS
Stratigraphy The oldest exposed rocks in the study area are Triassic in age and are exposed in the core of G. Araif El Naqa anticline. Jurassic rocks have wider exposures in the area and are exposed in G. Araif El Naqa and G. Minsherah in addition to the main outcrop of G. Maghara. Jurassic rocks have also been penetrated by some boreholes in the study area. Cretaceous outcrops have wider distribution in the area and are exposed mainly in the three major fold structures of Gebels Yelleg, Halal and Maghara, as well as the smaller folds. Senonian and Cenozoic rocks are exposed with gentle attitudes in the topographically low areas lying between the high fold structures, as well as some small tablelands. The youngest exposed sedimentary rocks in North Sinai are of Middle Eocene, Oligocene, Pliocene and Quaternary age, whereas Upper Eocene, most of the Oligocene, Miocene and most of the Pliocene sediments are only found in the subsurface of the northernmost onshore area of Sinai, as well as in the offshore area. Basic igneous intrusives are also exposed in a few parts of the study area, some of which have Early Miocene age (Steinitz et al. 1978) whereas others are not dated (e.g. G. Araif El Naqa area). Similar, nearby basalt outcrops in G. Ramon (Ramon Basalt) are similar to some of those of G. Araif El Naqa and were dated as Early Cretaceous (111 –116 Ma); Geological Survey of Israel (2001). Although the Mesozoic and Cenozoic rocks of northern Sinai show a gradual northward increase in thickness, abrupt changes in thickness of the Mesozoic rocks (especially the Jurassic section) occur in the study area reflecting the presence of
41
syn-depositional tectonism during the Mesozoic time. Regional changes in thickness of the different rock units are controlled by the Sinai hinge belt and the Nile Delta hinge zone (Fig. 3). The Sinai hinge belt, which extends in an ENE –WSW direction from the northern tip of the Gulf of Suez toward the Dead Sea Transform, represents the boundary between a tectonically stable (platform) area to the south and a tectonically active area to the north. The tectonically active area included some extensional basins that were actively subsiding during the Mesozoic (mainly Jurassic time). The Nile Delta hinge zone of Schlumberger (1984) or the Nile Delta flexure zone of Sestini (1989) has an arcuate outline and extends from the Bardawil Lake toward the Suez Canal and the central part of the cultivated Nile Delta (Fig. 1). It is a major zone of flexure along which listric normal faults, with a down to the north throw of 5–6 km at the Cretaceous and Middle Eocene levels exist, with abrupt changes in thickness of Cenozoic sediments across this zone (0.5– 1 km to the south versus 5 –7 km to the north), Sestini (1989). Bouguer gravity anomalies indicate a 2– 3 km thick sedimentary section to the south, v. 10 km thick section to the north of the Nile Delta hinge zone (Schlumberger 1995). Table 1 shows the maximum thicknesses of Mesozoic and Cenozoic rocks in the study area, and Figure 3a shows the effect of the Sinai hinge belt on the changes in thickness of the Jurassic rocks and the Nile Delta hinge zone on the changes in thickness of the Cenozoic rocks. Figure 3b also shows that several sub-basins existed in the area, lying north of the Sinai hinge belt as indicated by the changes in thickness of
Table 1. Maximum thicknesses of Mesozoic and Cenozoic rocks in the study area System/series/stage Oligocene Eocene Paleocene Senonian Turonian Cenomanian Lower Cretaceous Jurassic Triassic 1
Kuss & Boukhary 2008. Osman et al. 2000. 3 Farag & Shata 1954. 4 Jenkins 1990. 5 Al-Far 1966. 6 Eicher 1947. 7 Abed et al. 1996. 2
Maximum exposed thickness 1
77 m, G. Risan Aneiza 51þ m, east of G. Halal2 65 m, G. Minsherah3 295 m, G. Minsherah3 213 m, G. Halal2 624 m, G. Halal2 243 m, G. Halal4 250 m, G. Minsherah3 1980þ m, G. Maghara5 203.5þ m6 or 207þ m7, Araif El Naqa
Maximum subsurface thickness 152 m, subsurface of North Sinai 59 m, Nakhl well 326 m, Abu Hamth well 260 m, Nakhl well 370 m, Abu Hamth well 3234 m, Halal-1 well4 914 m, Halal-1 well4
42
A. R. MOUSTAFA
the Jurassic and/or Cretaceous sediments (cf. the Jurassic sections of G. Maghara and G. Halal as well as the thick Aptian –Albian section in the offshore Mango structure). The following is a brief review of the major stratigraphic units.
Triassic rocks The exposed Triassic section of G. Araif El Naqa was studied by different workers. Bartov et al. (1980) and Jenkins (1990) reported 185þ m thick section whereas Eicher (1947) and Abed et al. (1996) reported larger thicknesses (203.5þ m and 207 m, respectively). The Triassic section of this area includes two main units, a lower clastic unit and an upper carbonate unit. The lower clastic unit is 68 m thick and consists of dark and lightcoloured, fine- to medium-grained sandstone with shaly, marly and silty beds intercalated with a few fossiliferous limestone beds. It also includes two weathered dark igneous sills. The shales and sandstones contain plant imprints and petrified wood fragments. This sandstone unit was deposited in a terrestrial, shallow coastal, to shallow marine environment. The upper carbonate unit is 117 m thick and is made up of shallow marine limestone beds alternating with marls. It also includes an olivine basalt dyke. The upper part of the unit includes some gypsum intercalations.
Jurassic rocks Jurassic rocks are exposed in the core of the G. Maghara anticline (1900þ m; Al-Far 1966), G. Araif El Naqa (141 m), and G. El Minsherah (80þ m). The lowermost Jurassic sediments of G. Maghara (Mashabba Formation, Early Jurassic age) are made up of fluvial sandstones containing large wood fragments that were deposited by northerly flowing braided streams carrying detritus shed off the Arabo-Nubian massif. The Mashabba Formation is followed by interbedded shallow marine carbonates of the Rajabiah Formation and nearshore marine clastics of the Shusha Formation. Lower Jurassic sandstones and shales have also been penetrated by the Nakhl (162 m), Abu Hamth (162 m), Falig-1 (315þ m), and N. Falig-1 (383þ m) wells. A thick clastic sequence in the Ayun Musa-2 (622 m) and Hamra-1 (534 m) wells has been assigned a Middle– Early Jurassic age, while the El Khabra well has an undifferentiated Jurassic section (1430 m) of sandstones, shales and limestones. The Middle Jurassic sediments of G. Maghara are 660 m thick and include a lower carbonate unit (Bir Maghara Formation) and an upper clastic unit (Safa Formation). A Middle Jurassic clasticcarbonate sequence is also exposed at G. Minsherah
and was also penetrated in the Halal-1 (780 m), Falig-1 (310 m), N. Falig-1 (415 m), Katib El Makhazin (502þ m) and Giddi-1 (805þ m) wells. Several coal seams exist in the Lower and Middle Jurassic clastic units (Shusha and Safa Formations, respectively). The Upper Jurassic sediments (Masajid Formation) are dominated by carbonates representing a southerly marine transgression at the end of Bathonian–Callovian times. At G. Maghara, these carbonates are 575 m thick (Al-Far 1966). Shelf carbonates have also been recognized in the Halal-1 (214 m), Falig-1 (530 m), N. Falig-1 (635 m), Ayun Musa-2 (121 m), Hamra-1 (65 m) and Katib El Makhazin (388 m) boreholes. Drilling for hydrocarbon exploration in the northwestern part of G. Yelleg indicated the presence of another Upper Jurassic clastic unit above the Masajid Formation. This unit was named the Gifgafa Formation by geologists working in the Gulf of Suez Petroleum Company (see also Abdel Aal & Lelek 1994). The Gifgafa Formation is a shale section with some sandstone and a few limestone and dolomite streaks. It is 145 m thick in the Falig-1 well and 220 m thick in the N. Falig-1 well
Comments on the changes in thickness of the Jurassic rocks Although the thicknesses of Jurassic rocks penetrated by boreholes and reported in the last section represent true vertical rather than true stratigraphical thicknesses, all of them, perhaps except that of the Halal-1 well, can be considered true stratigraphical thicknesses as these wells occupy crestal positions of folded rocks. The thicknesses presented in the last section indicate that the changes in thickness of the Jurassic rocks in northern Sinai do not reflect a homogeneous northward increase in thickness as one would expect in a normal depositional setting. On a regional scale, Hirsch & Picard (1988) identified two separate ENE –WSW oriented Jurassic basins in the North Sinai–Levant area, namely the Maghara –Halal basin and the Helez – Ramallah trough with Jurassic thicknesses equal to 3000 and 4000 m, respectively. The Mahgara – Halal basin generally corresponds to the area with thick Jurassic rocks in Figure 3. These two basins are separated by a high area containing a thinner Jurassic section. Changes in thickness of the Jurassic rocks in G. Maghara and Halal-1 well indicate the existence of separate sub-basins in northern Sinai like those in the northern Western Desert. The tectonic evolution of the northern Western Desert was previously studied using numerous borehole and seismic data indicating a phase of extensional deformation in the Jurassic time, where several half graben-like basins were formed (Moustafa
NORTH SINAI FOLDS
et al. 1998; Abd El-Aziz et al. 1998). A similar situation might also have existed in northern Sinai. In this regard, G. Halal and G. Maghara would represent two separate Jurassic sub-basins. Table 2 shows a comparison of the thickness of the Jurassic rocks in the two areas indicating that basin opening in G. Maghara and G. Halal probably took place in the Early and Middle Jurassic Epochs, whereas the Late Jurassic Epoch was a period of gradual northward increase in thickness.
Cretaceous rocks Lower Cretaceous rocks. Fluvial-continental Lower Cretaceous sediments unconformably overlie the marine Upper Jurassic rocks as a result of a major eustatic fall in sea level. They belong to the Malha Formation, which has an Early Cretaceous age and have been penetrated in the Abu Hamth (370 m), Nakhl (247 m), Ayum Musa-2 (149 m), Hamra (366 m), Kabrit (103 m), Falig-1 (217þ m), N. Falig-1 (318 m), Katib El Makhazin (32 m) and Mango-2 (242þ m) wells. The exposed section of the Malha Formation is 160 –168 m at G. Yelleg (Jenkins 1990), 243 m thick at G. El Halal (Osman et al. 2000) and 250 m thick at G. El Minsherah (Farag & Shata 1954). The Malha Formation is overlain by an interbedded series of argillaceous clastics and carbonates of Aptian– Albian age. These represent fluvialparalic to shallow shelf facies deposited in a low energy system, with occasional high energy episodes, indicated by the presence of rudists and oolites. The Aptian– Albian sediments are 240 – 382 m thick in the G. Maghara area (Bachmann & Kuss 1998). Thick Aptian–Albian sediments reaching a thickness of 1749 m were encountered in the Mango-1 well in offshore North Sinai. Abu Zeid (2007), based on micropalaeontological work, considered the 300 m thick section overlying the Malha Formation in the northern part of G. Maghara area (G. El Tourkmaniya outcrop) to be of Barremian–Albian age, indicating that Cretaceous transgression in northern Sinai started in the Barremian time. Upper Cretaceous rocks. The Upper Cretaceous rocks in North Sinai are mostly carbonates and
43
form widespread exposures. The Cenomanian section in North Sinai is dominated by a thick sequence of dolomites that were deposited on a broad shallow subtidal shelf. At G. Halal, the Cenomanian Halal Formation is 624 m thick (Osman et al. 2000). According to Bartov & Steinitz (1977), the Cenomanian rocks in North Sinai and the Naqb Desert represent a continuous transgressive event of a relatively long duration without significant regressions. Transgression proceeded from the northwestern part of Sinai that was already submerged during the Aptian or even earlier (Shata 1960; Abu Zeid 2007). The Turonian section in North Sinai includes the Abu Qada and Wata Formations. The Lower Turonian Abu Qada Formation consists of interbedded shales, marls, limestones and sandstones. It is 68 –141 m thick at G. Araif El Naqa and is represented by gypsiferous shales with thick limestone beds (Bartov et al. 1980). The Upper Turonian Wata Formation is a thick carbonate sequence, which forms the prominent dip slopes of the North Sinai anticlines. Its thickness is about 180 m at G. Araif El Naqa (Bartov & Steinitz 1977), 190 m at G. Yelleg (Visser 1941), 132 m at G. El Minsherah (Farag & Shata 1954) and 213 m at G. Halal (Osman et al. 2000). A sequence of massive Turonian limestones with rare shale interbeds was penetrated in the Abu Hamth (254 m), Darag (227 m) and Nakhl (260 m) wells. Deep marine conditions persisted throughout the Cenomanian and Turonian periods in the extreme northern part of Sinai as well as in the offshore area. The Coniacian –Santonian sediments of North Sinai are represented by marine sediments of the Themed Formation (Ziko et al. 1993). The upper part of the Coniacian rocks is truncated by erosion at early Late Coniacian in NE Sinai (Lewy 1975). The Campanian–Maastrichtian sediments are represented by white, soft chalks of the Sudr Formation. Undifferentiated Senonian sediments were penetrated by the Darag (235 m), Nakhl (33 m), Abu Hamth (326 m) and El Khabra (192 m) wells. The exposed Senonian rocks are 265 m thick at G. Yelleg (Visser 1941), 295 m thick at G. El Minsherah (Farag & Shata 1954) and 206 m thick at G. Halal (Osman et al. 2000).
Cenozoic rocks Table 2. Comparison of the thickness of the Jurassic rocks in the G. Maghara outcrop and Halal-1 well Sequence Upper Jurassic Middle Jurassic Lower Jurassic
G. Maghara
Halal-1 well
575 m 660 m 665þ m
214 m 780 m 2240 m
Although the Paleocene sediments (Esna Shale) in Sinai are located in the structural lows between the major anticlinal structures, this does not rule out the possibility that they were once present at the crests of the anticlinal structures and were eroded later. The Esna Shale has a uniform lithology of greenish –grey shale. In the Nakhl and Darag boreholes it is 59 m and 38 m thick, respectively.
44
A. R. MOUSTAFA
To the north, the Esna Shale is exposed on the flanks of G. Maghara (maximum 1 m) as well as in G. El Minsherah (65 m), G. Halal (55 m) and G. Araif El Naqa (50 m). The Eocene sediments crop out in many areas throughout Sinai and have also been penetrated by several wells. The Lower Eocene rocks in Sinai are generally represented by a massive flinty limestone (Thebes limestone) and occupy the broad synclinal lowlands between the large anticlines. To the east of G. Halal, the exposed Middle Eocene rocks are 51 m thick (Osman et al. 2000). No Upper Eocene outcrops have been recognized in North Sinai. However, Upper Eocene rocks were recognized in the shallow subsurface near G. Libni to the east of G. Maghara. Few scattered outcrops of Upper Oligocene shallow marine carbonate rocks (77 m thick) have been reported by Kuss & Boukhary (2008) in G. Rizan Aneiza to the east of G. Maghara. These carbonate rocks rest unconformably above Jurassic and Lower Cretaceous rocks. Basic volcanic igneous rocks of the Late Oligocene –Early Miocene age exist in some parts of northern Sinai, as reported in G. Yelleg and its vicinity. A 40 m thick basic igneous sill is exposed in G. Ikteifa (G. Yelleg area) where it is intruded at the contact between the Maastrichtian chalk and
the Esna Shale. According to Steinitz et al. (1978), this sill is of Early Miocene age (20.5 + 0.7 Ma, K –Ar dating). A NNE oriented discontinuous dyke also exists in the area and extends from G. Ikteifa to the western side of G. El Minsherah. Several small, one-meter thick dykes were also identified at the northeastern part of G. Yelleg by Khalil (1991). A NW–SE oriented basic igneous dyke crops out to the south of G. Maghara (El-Hemma area). Also, a very long east –west oriented basic igneous dyke (Raqabat Naam dyke) exists in central Sinai and has the same age like the G. Ikteifa volcanics (Eyal et al. 1980).
Structural setting The structural pattern of North Sinai is dominated mainly by NE–SW oriented doubly plunging anticlines (Fig. 4). Locally, ENE and east –west oriented faults form boundaries to these folds. Faults of other orientations also exist, but are associated with the anticlinal folds and are thought herein to be fold-related faults. The anticlinal folds of the area are of two obvious sizes, large (tens of kilometres long) and small (several kilometres long). G. Maghara, G. Halal and G. Yelleg folds are the largest three folds in North Sinai and are
Fig. 4. Simplified structural form-line map of North Sinai and the Naqb Desert after Khalil & Moustafa (1994) showing the main structural features of the study area.
NORTH SINAI FOLDS
60, 45 and 43 km long, respectively. Other large folds also exist to the east in the Naqb Desert, for example, Ramon, Kurnub and Qatan anticlines. The small anticlines exist in several parts of the study area, but are dominant in the southern part of the area and are mostly associated with the ENE –WSW and east– west oriented faults. The ENE –WSW and east –west oriented faults exist in two main belts that extend across the study area (Fig. 4). One of these two belts is the Sinai hinge belt, which forms the southern boundary of the area affected by large anticlines. This belt also corresponds to the central Sinai –Negev shear zone of Bartov (1974). The other fault belt generally has east –west orientation and extends from the
(a)
N
45
Gulf of Suez to the northern end of the Gulf of Aqaba and was previously called the themed fault (Moustafa & Khalil 1994). Figure 5 shows the orientations of the different structures in northern Sinai. The folds of the area (both large and small) have a predominant NE– SW orientation with a mean orientation of N458E– S458W (Fig. 5a). The faults of the area (both in the Sinai hinge belt and in the folded area) show a preferred NW to WNW orientation (Fig. 5b). The faults dissecting the large folds have a N558W preferred orientation (Fig. 5c), whereas those in the hinge belt (Fig. 5d ) also include other trends like NNW and ENE. Lengthwise, the ENE –WSW oriented faults are the longest
(b)
N
All faults N = 1197
All folds N = 272
46
165
N
(c)
N
(d)
Sinai hinge belt faults N = 584
Faults dissecting the large folds N = 613
88
95
Fig. 5. Rose diagrams showing the strikes of the different field-measured structures in North Sinai. Number along each circle shows the length of the longest petal in the rose diagram.
46
A. R. MOUSTAFA
faults in the Sinai hinge belt and are the main faults that controlled the deformation of this part of North Sinai.
North Sinai folds G. Maghara folds. The structures of G. Maghara affect an area that is about 60 km long and 13– 17 km wide. They include four main asymmetric NE–SW oriented anticlines (Um Mafruth, Um Asagil, Maghara and Hamayir) separated by two main SE vergent reverse faults (Hamayir and Mizeraa faults), as shown in Figure 6. The Hamayir and Maghara asymmetric anticlines lie on the hanging walls of these two reverse faults (see cross sections in Fig. 6). The Um Asagil asymmetric anticline may lie in the hanging wall of a third reverse fault located south of the G. Maghara complex. To the NE and SW of the Maghara anticline are two narrow ENE –WSW oriented ridges (Um Mafruth and Um Latiya ridges) that extend for 13 and 23 km respectively (Fig. 6). The Hamayir reverse fault dips at an angle ranging from 44 –738NW and has throw equal to 500 m
causing the Middle Jurassic rocks to ride over the Upper Jurassic rocks. The orientation of the eastern portion of this reverse fault changes to north– south with a corresponding change in orientation of its hanging wall anticline. The Mizeraa reverse fault is at least 45 km long and is made up of several segments joined together by strike –slip (tear) faults (Fig. 6). The overall trend of this fault is NE–SW but has some segments oriented NNE and WNW. Sinistral offset characterizes the NNE faults, whereas dextral offset characterizes the WNW faults. The maximum distance of horizontal offset is equal to 2– 2.5 km. The dip angle of the Mizeraa reverse fault reaches a high value of 718. Two large anticlines lie on the hanging wall of the fault. These are the G. Maghara main anticline (where all the section of Jurassic rock is exposed) and the Um Mafruth anticline. The maximum throw of the Mizeraa fault is 1250 m at its middle part. The northwestern flank of the G. Maghara anticline dips at about 158NW whereas its southeastern (forelimb) is narrower and usually vertical to overturned. G. Manzour represents the footwall of the Mizeraa
Fig. 6. Landsat TM image (see Fig. 4 for location) and structural cross sections (from Moustafa 2005) showing the main structures of the G. Maghara area.
NORTH SINAI FOLDS
47
Fig. 7. Field photograph (looking south) showing the asymmetric syncline of G. Manzour that represents the footwall of the Miseraa reverse fault. See Figure 6 for location. Apt, Alb, Kc and Kt designate Aptian, Albian, Cenomanian and Turonian rocks, respectively.
fault and is folded by an asymmetric syncline with a very steep northwestern limb (Fig. 7). The anticlines of the G. Maghara area are pervasively dissected by transverse (NW –SE oriented) normal faults that have relatively small amounts of throw (several tens of metres). These faults are synchronous with the folds and were probably formed due to lengthening of the rocks in a direction normal to the NW–SE shortening, resulting from the folding and reverse faulting. These faults do not dissect the Paleocene– Eocene rocks exposed in the outer edges of the folds. This gives strong evidence that they are of the same age as the folds and can not be attributed to younger extensional deformation (like the Gulf of Suez rifting for instance). Similar observations are also obvious in the G. Yelleg and G. Halal folds. G. Yelleg folds. Like G. Maghara, the G. Yelleg area also has a large NE–SW oriented anticline. This anticline is 45 km long and 18 km wide and affects the exposed Lower Cretaceous to Lower Eocene rocks of the area. It is a large asymmetric fold with a gentle northwestern flank dipping at 5–138NW and a steeper southeastern flank dipping as much as 568SE. The southeastern flank has a monoclinal shape and winding outline (Fig. 8) with segments oriented ENE and NNE within the predominant NE trend. Like G. Maghara, the G. Yelleg anticline is dissected by a large number of NW to WNW oriented (transverse) faults (Fig. 8, inset) that have relatively small amounts of throw in the order of a few tens of metres.
Cretaceous and Cenozoic rocks in the northern part of G. Yelleg area are affected by a number of smaller NE–SW oriented symmetric folds, the most prominent of which are G. Falig, G. Meneidret Abu Quroun and G. Meneidret El Etheili folds (Fig. 8); Moustafa & Gibali (2005) and Moustafa & Khalil (1995). The southernmost part of the G. Yelleg area is bounded by ENE– WSW oriented faults at G. Rishat Saada, G. Rishat Lehman and G. El Minsherah (Fig. 8). These faults belong to the Sinai hinge belt. The G. Yelleg monocline gradually fades away into the ENE –WSW oriented fault of G. Rishat Saada by gradual flattening of its steep southeastern flank. Mesostructures measured in the Upper Cretaceous rocks at the northeastern side of the G. Yelleg area (G. Meneidret El Etheili area) include tectonic stylolites, calcite veins and minor normal faults. These features indicate that s1 was oriented NW–SE and s3 was oriented NE–SW (Fig. 9a). Minor faults affecting Upper Cretaceous rocks of G. Falig also indicate the same orientation of s1 (Fig. 9c). Analysis of the fault– slip data of macrofaults at G. Meneidret El Etheili and G. Falig folds (Fig. 9b, d, respectively) indicate maximum horizontal compressive stress in the NW quadrant and minimum horizontal stress in the NE–SW quadrants. G. Halal folds. The structures of the G. Halal area were mapped in detail by Abd-Allah et al. (2004). The exposed Cretaceous rocks of this area are folded by a very large asymmetric, NE–SW
48
A. R. MOUSTAFA
Fig. 8. Landsat TM image showing the structures of the G. Yelleg area (see Fig. 4 for location). Rose diagram shows the strikes of field-mapped faults of this area.
oriented, doubly plunging anticline that is 43 km long and 14 km wide. This asymmetric anticline has a gentle northwestern flank dipping at about 158NW and a steep southeastern flank that is mostly vertical to overturned (Fig. 10a, c). The steep flank of the G. Halal anticline is expected to be underlain by a NE–SW oriented reverse fault similar to the Mizeraa reverse fault of G. Maghara. Right-stepping en echelon folds in the form of three anticline –syncline pairs affect the southern flank of the G. Halal anticline and give it a characteristic winding outline (Fig. 10a). Long distances separate these three fold pairs and these folds are similar to those affecting the forelimb of the Maghara and Um Mafruth folds. The en echelon folds in these three localities include hanging wall anticlines above en echelon reverse fault segments. The intervening synclines represent transfer zones between these en echelon fault segments (Moustafa 2005). A large number of transverse (NW – SE) oriented normal faults dissect the Cretaceous rocks of G. Halal (Fig. 10b). These faults have steep dip and relatively small throws. Analysis of slip data of these faults indicates lengthening in the NE– SW direction parallel to the fold axis (Fig. 11). Lower Cretaceous sandstones are exposed in the breached core of the G. Halal anticline. The Halal-1
well was drilled in the core of the anticline in 1975 with a total depth of 4311 m subsea. The well penetrated the thickest Jurassic basin, which was inverted later in Late Cretaceous – Early Cenozoic time.
Sinai hinge belt The Sinai hinge belt is a narrow ENE– WSW oriented structural belt forming the boundary between the North Sinai folds and the platform area lying to the south. This belt is 15–20 km wide and is dominated by long segments of ENE – WSW oriented faults that show evidence of rightlateral strike–slip displacement. Detailed field mapping of different parts of the Sinai hinge belt (Moustafa & Yousif 1990; Moustafa & Gibali 2005; Moustafa & Salama 2005) indicate that the ENE –WSW oriented faults in the hinge belt form two sub-belts, northern and southern. In each subbelt, the ENE –WSW oriented faults have a rightstepping en echelon arrangement (Fig. 12). The northern sub-belt extends from the southernmost side of G. Yelleg toward G. El Riash, whereas the southern sub-belt extends from the Mitla Pass to G. Araif El Naqa (Fig. 12). Fault–slip data indicate that these ENE –WSW oriented fault segments show dextral strike–slip displacement (Fig. 13).
N
(a)
(b)
σ3
σ1
N
N 44/356 P
44/356 P
E
W
T 14/252
T 14/252 Tectonic Stylolites
σ1
(c)
S
Calcite Veins Minor Normal Faults
N
N
Meneidret El Etheili Faults (N = 11)
(d)
N
N T 10/017
77/003 T
P 09/140 G. Falig Minor Faults (N = 5)
77/004 T
P 36/280
T 10/017
NORTH SINAI FOLDS
σ3
P 36/280
P 09/140
G. Falig Faults (N = 10)
Fig. 9. (a) Lower hemisphere equal area projection of the poles to mesostructures in G. Meneidret El Etheili (NE part of G. Yelleg) after Moustafa et al. (1991). (b), (c) and (d) show lower hemisphere equal-area projections of field-measured fault planes (great circles) and slickenside striae (small filled circles with arrows) in each left-hand stereogram and analysis of fault– slip data by FaultKin software of Allmendinger (2001) in each right-hand stereogram for Meneidret El Etheili large faults (b), G. Falig minor faults (c) and G. Falig large faults (d). P and T show the orientations of the maximum and minimum principal stress axes. 49
50
A. R. MOUSTAFA
Fig. 10. (a) Landsat TM image of the G. Halal anticline (see Fig. 4 for location). Note the winding trace of the steep southern flank (dotted line). (b) Rose diagram of the strikes of field-mapped faults. (c) Structural cross section across G. Halal fold modified after Abd-Allah et al. (2004).
Some of the ENE –WSW oriented faults in the Sinai hinge belt are characterized by the presence of ENE –WSW oriented anticlines, axially dissected in their middle parts by these faults, for example, G. Rishat Saada (Fig. 14), G. El Minsherah, G. El Riash and G. Kherim folds. Each of the faults dissecting these folds is about 1.5 times the length of the respective fold and it seems that these folds were formed in the middle part of each fault at the early stages of strike– slip movement of the fault. Erosion of the crestal areas of these folds shows more deformation of the older rocks in the axial areas of the folds by tighter folding and steeper
dip of the beds, for example, G. El Minsherah fold (Fig. 15). Some of these folds are dextrally offset by the faults. Thus, in G. Rishat Saada, the southern half of the anticline is dextrally offset for about 2 km relative to the northern half (Fig. 14). Some of the faults enclose pull-apart grabens between their overlapping ends such as the area between the El Minsherah and Rishat Lehman faults (Moustafa & Yousif 1990). Horizontal and gently plunging slickenlines characterize the ENE –WSW oriented faults of the Sinai hinge belt indicating dextral slip (Fig. 13). Bartov et al. (1980) also reported rightlateral strike –slip displacement along the Araif El
NORTH SINAI FOLDS
N
51
N
T
P
N = 30 Fig. 11. Analysis of fault– slip data of the field-measured faults of G. Halal. Symbols are like those in Figure 9. N indicates the number of faults.
Naqa fault that lies in the eastern part of the southern sub-belt. The westernmost side of the southern sub-belt of the Sinai hinge belt is characterized by the existence of right-stepping en echelon doubly
plunging anticlines, rather than an ENE fault segment (Fig. 16 and Moustafa & Khalil 1989). The arrangement of these folds provides evidence for dextral simple shear deformation along the
Fig. 12. A Landsat TM image showing ENE–WSW oriented en echelon faults of the Sinai hinge belt (see Fig. 4 for location). These faults are located in two sub-belts (A & B). Main faults in the northern sub-belt (A) include G. Rishat Saada fault (1), G. Rishat Lehman fault (2), G. El Minsherah fault (3), South Talet El Badan fault (4) and G. El Burqa– G. El Riash fault (5). Main faults in the southern sub-belt (B) include G. El Bruk fault (6), G. Kherim fault (7) and G. Araif El Naqa fault (8). Rose diagram represents strikes of all field-mapped east–west to ENE–WSW oriented faults in the hinge belt.
52
A. R. MOUSTAFA
Fig. 13. Analysis of fault– slip data of the Sinai hinge belt. Symbols are as those in Figure 9.
Fig. 14. Landsat TM image (a) and vertical aerial photograph (b) of G. Rishat Saada fault and the associated fold. See Figure 8 for location.
NORTH SINAI FOLDS
53
Fig. 15. Landsat TM image of G. El Minsherah area showing the Minsherah fault that axially dissects the middle part of the Minsherah anticline (see Fig. 8 for location). Note the small tight folds in the core of the anticline, north of the fault. Structural cross section across the Minsherah fold is after Moustafa & Yousif (1990). A and T (away from and toward the observer, respectively) designate the sense of strike– slip movement on the Minsherah fault.
Sinai hinge belt. Noweir et al. (2006) also confirmed the same style of deformation for this area. The Sinai hinge belt is obvious on the Bouguer anomaly and total magnetic intensity maps of Sinai Peninsula (Egyptian General Petroleum Company 1985) where it shows up like a boundary with a change in geophysical character across it.
Themed fault The tectonically stable area lying south of the Sinai hinge belt is cut by a very long east –west oriented fault zone, the Themed fault. This fault zone
extends for about 200 km from the vicinity of the eastern margin of the Suez rift to the Dead Sea transform (Fig. 1). The themed fault disrupts the monotonously flat-lying rocks of central Sinai by the development of a narrow elongated belt of small doubly plunging tight folds. This fault was rejuvenated along a pre-existing fault marking the southernmost edge of the Early Mesozoic continental margin of the Eastern Mediterranean basin in central Sinai (Moustafa & Khalil 1994). Within its very long tract, the Themed fault truncates rocks ranging in age from Precambrian to Middle Eocene. The eastern and western parts
54
A. R. MOUSTAFA
Fig. 16. Landsat TM image showing the en echelon doubly plunging anticlines affecting the Upper Cretaceous rocks in the Mitla Pass. See Figure 4 for location.
of the fault are very clear, whereas the central part (traversing the Tih Plateau) is less obvious. A basalt dyke of Early Miocene age is intruded along a segment of the Themed fault in the vicinity of G. El Dirsa (see Fig. 4 for location). The Themed fault was mapped in detail by Moustafa & Khalil (1994). The westernmost part of the fault consists of four right-stepping en echelon segments oriented east –west and ENE – WSW. All these segments have steep dips (about 50–908) with slickenside lineations recording dominantly right-lateral strike– slip motion (Fig. 17a). East–west to NE–SW oriented, doubly plunging folds are associated with these fault segments and affect the exposed Cretaceous and Cenozoic rocks. Since a segment of the Themed fault is intruded by an Early Miocene basaltic dyke and the dyke itself does not show any fault-related deformation, Moustafa & Khalil (1994) considered the last movement on the Themed fault to be pre-Miocene in age. The easternmost one third of the Themed fault has east– west orientation where the fault has a steep northward dip that ranges between 628 and 908. Local southward dip of the fault occurs less
often. Nearly horizontal to gently plunging slickenside lineations and associated chatter marks on this segment of the fault indicate right-lateral strike–slip or oblique– slip with a predominant right-lateral strike –slip component (Fig. 17b). In addition, the Rishat El Themed doubly plunging anticline is dextrally offset by a fault for about 300 m (Fig. 17c, d). The en echelon arrangement of the segments of the Themed fault, their associated folds, and the horizontal to gently plunging slickenlines, indicate that the Themed fault is a dextral wrench zone. The Themed fault affects rocks as young as Middle Eocene indicating a post-Middle Eocene slip. The last movement on the fault must be preEarly Miocene, since an Early Miocene basalt dyke was intruded in the weak zone of the fault.
Tectonic evolution The North Sinai folds clearly show the effect of horizontal compressive stress on the deformation of the exposed Mesozoic and Cenozoic rocks. All
NORTH SINAI FOLDS
55
Fig. 17. Lower-hemisphere equal-area projections of the strikes of field-measured faults (great circles) of the western part (a) and eastern part (b) of the Themed fault and their slickenside lineations (small circles) after Moustafa & Khalil (1994). U and D refer to upthrown and downthrown sides respectively. (c) Geological map and (d) vertical aerial photograph of G. Rishat El Themed fold showing its dextral offset by the Themed fault (after Moustafa & Khalil 1994). See Figure 4 for location.
56
A. R. MOUSTAFA
structures in the three folded areas (Maghara, Yelleg and Halal) indicate shortening in the NW– SE direction and related lengthening in the NE– SW direction. Subsurface data of the study area show the effect of an earlier deformation that proceeded by crustal extension in Mesozoic time, indicating that the exposed folds were formed by basin inversion. Seismic reflection profiles and borehole data show the existence of relatively small extensional basins in the area. A 2D seismic section extending from the G. Um Mafruth anticline (G. Maghara area) to the northeastern nose of the G. Yelleg fold (Fig. 18) clearly shows the structures of this portion of North Sinai. Asymmetric folds associated with reverse faults are clear in the G. Maghara area (northwestern part of the seismic section) followed to the south by a 15-km wide synclinal box fold separating these asymmetric folds from the G. Yelleg anticline. The latter
(southeastern part of the seismic section) is a large box-shaped anticline with two steep flanks on its northwestern and southeastern sides. Both flanks are bounded by faults but the fault bounding the southeastern flank is a blind reverse fault that dissects only the lower part of the Upper Cretaceous section and older rocks. Flattening of the seismic section on the top Masajid Formation (Fig. 19) indicates that all faults had normal slip at Jurassic time and that the Jurassic section lying between the top Rajabiah Formation (intra-Lower Jurassic) and top Masajid Formation (Upper Jurassic) was deposited in two separate extensional sub-basins in G. Maghara and G. Yelleg. These two sub-basins are separated by an inter-basinal high area located at the present-day syncline separating the G. Maghara and G. Yelleg folds. The thickness of the Jurassic sediments in the G. Yelleg sub-basin is smaller than that in the
Fig. 18. Two-dimensional seismic section extending from the eastern part of G. Maghara anticline to the eastern nose of G. Yelleg anticline (see Fig. 1 for location). Symbols designate the following reflectors: top of the Rajabiah Formation (intra-Lower Jurassic) (Jl), top of the Upper Jurassic (Ju), top of the Lower Cretaceous (Kl), tops of the Intra-Upper Cretaceous units (Ku1 and Ku2), top of Santonian (Ku3) and top of Maastrichtian chalk (Ku4).
NORTH SINAI FOLDS
57
Fig. 19. Same seismic section of Figure 18 flattened on the top Jurassic reflector.
G. Maghara sub-basin. The G. Maghara sub-basin includes a southward thickening Jurassic wedge. The thickest part of the Jurassic section in the Yelleg sub-basin is located in its middle part. Small changes in thickness of the Lower and Upper Cretaceous rocks are obvious across some of the faults. Such changes in thickness could either be real or unreal and related to the seismic flattening process. The seismic resolution below the top Rajabiah reflector is not good enough to show any changes in the thickness of the Lower Jurassic rocks. The two Jurassic sub-basins of G. Maghara and G. Yelleg indicate extensional deformation by rifting in North Sinai during the Jurassic. Similar change in thickness of the Jurassic section in G. Maghara and G. Halal areas was also pointed out in a previous section (Table 2). The exposed Jurassic rocks in G. Maghara (1980þ m; Al-Far 1966) represent almost the whole Jurassic system. On the other hand, Halal-1 well that was drilled in the core of G. Halal anticline penetrated a complete Jurassic section, which is 3234 m thick (much thicker than that of G. Maghara). The location of Halal-1 well, relative to the crest of the G. Halal anticline and to its steeply dipping flank (Fig. 10c), indicates that the penetrated Jurassic thickness is close to the true thickness rather than being the vertical thickness of steeply dipping rocks. Because G. Halal lies to the south of G. Maghara, it was expected to
have a thinner Jurassic section assuming northward increase in thickness of the marine Phanerozoic rocks of Egypt. The thick Jurassic section penetrated by the Halal-1 well most probably indicates a separate extensional sub-basin that was subsiding during Jurassic time. Based on these changes in thickness of the Jurassic rocks in the Maghara, Yelleg and Halal areas, as well as the local changes in thickness of the Lower and Upper Cretaceous rocks across the Mizeraa fault, it is proposed herein that three separate extensional sub-basins existed in these areas during the Jurassic and probably during part of the Cretaceous Period. These sub-basins were inverted at a later time as the early normal faults were reactivated by reverse slip and the large anticlines of G. Maghara, Yelleg and Halal are fault-propagation folds associated with the upward propagation of these faults during basin inversion. Detailed field mapping of the G. Maghara area indicates that the NE –SW oriented Mizeraa and Hamayir reverse faults have several segments oriented north–south and WNW– ESE (Moustafa 2005). Slickenside lineations on the north–south and WNW–ESE oriented fault segments show oblique slip where the north–south fault segments show left-lateral and the WNW –ESE fault segments show right-lateral strike– slip components. The horizontal separation on these fault segments is too large to represent the amount of horizontal
58
A. R. MOUSTAFA
slip on the faults and it is proposed herein that these inverted faults (e.g. the Mizeraa fault) had zigzag shape during the early extensional phase in the Jurassic and Early Cretaceous times. Such fault patterns are common in extensional basins where different rift-parallel segments are linked together by transfer faults (e.g. Colletta et al. 1988; Morley et al. 1990; Ziegler 1992). During the phase of basin inversion, the fault segments oriented at high angle or normal to the direction of compression were reactivated by reverse slip (e.g. the NE–SW fault segments in North Sinai) whereas the transfer faults oriented at an angle to the direction of compression were reactivated by oblique slip. The nature of deformation at the Sinai hinge belt is different from that at the inversion structures of G. Maghara, Yelleg and Halal. Reliable field data support dextral simple shear deformation in the hinge belt along old ENE –WSW oriented faults that separate the inverted basins area of North Sinai from the platform area to the south. ENE – WSW oriented faults in north Egypt are known to be old, probably of Precambrian age. They dissect the exposed Precambrian basement rocks of the northern Eastern Desert, but not the nearby Phanerozoic sediments. Orwig (1982) also mapped faults of this trend in the Precambrian basement in the
subsurface in north Egypt. The Sinai hinge belt faults define the southern boundary of the area affected by Mesozoic extensional deformation. The SE-oriented compression, which led to basin inversion in North Sinai, reactivated these faults by dextral transpression (Fig. 20). The Themed fault is another fault lying further south in the platform area and was also reactivated by dextral slip. Being more oblique to compression, the Themed fault more clearly expresses the effect of strike – slip reactivation. Fault-slip data in Figures 11 and 13 perhaps indicate that faults did not slip under a single stress state. This may be attributed to the fact that all measured faults of each area are included together in one stereogram.
Time of compressional deformation Detailed field investigation indicates that compressional deformation in North Sinai took place in Late Cretaceous –Early Cenozoic times for the following reasons. (1) The en echelon folds of Mitla Pass affect the exposed Cenomanian, Turonian and Lower Senonian rocks. On the other hand, the Campanian –Maastrichtian, Paleocene, and
Fig. 20. Landsat TM image of North Sinai showing the inverted structures of Mahgara, Yelleg, Halal and the Hamayir–Amrar area. The Sinai hinge belt separates these inverted basins from the central Sinai platform area. Heavy arrow represents the direction of maximum compressive stress axis at Late Cretaceous– Cenozoic time.
NORTH SINAI FOLDS
(2)
(3)
(4)
Eocene rocks of the Sadr El Heitan Plateau, which lies exactly to the south of these folds, do not seem to be affected by folding. Reworked Lower Cretaceous sandstones deposited within the Campanian –Maastrichtian chalk section of G. Araif El Naqa were reported and explained by Luning et al. (1998) to have been eroded from the crest of the Araif El Naqa fold. This indicates that folding already started before these rocks were deposited. A similar (intra-Late-Senonian) time of folding was reported in the Abu Roash fold (west of Cairo) by Moustafa (1988). An intra-Paleocene syn-tectonic debris flow overlying the Upper Senonian chalk exists on the southern side of the G. Maghara fold (Moustafa 2005). This debris flow contains clasts derived from the Jurassic rocks indicating that intense folding and related uplift took place in pre-Paleocene time. Smaller amounts of debris flow also exist in the Lower Eocene rocks of the same area. The Paleocene and Lower Eocene rocks on the southern flank of the G. Maghara anticline show a gradual upward decrease in dip indicating progressive folding at these times. Folded Lower Eocene rocks on the NE side of G. Falig anticline (NW side of G. Yelleg area),
(5)
59
Moustafa & Gibali (2005), clearly indicate continuation of folding in post-Early Eocene time. Folded Middle Eocene rocks in the G. El Dirsa syncline (formed by the dextral slip on the Themed fault) represent the youngest folded rocks within the North Sinai outcrops (Moustafa & Khalil 1994). The lack of younger outcrops in North Sinai prohibits defining the end of the compressive deformation, but the presence of an Early Miocene basalt intrusion along the fault bordering the southern side of the G. El Dirsa syncline led Moustafa & Khalil (1994) to propose that the compressive deformation in North Sinai ended before Early Miocene time.
These observations indicate that the compressive deformation that led to basin inversion in North Sinai started in Late Cretaceous time and ended before Early Miocene time. The acme of this deformation was during the Late Cretaceous and is thought to have led to breaching and erosion along the crests of anticlines and deposition of syntectonic clastics within the Upper Cretaceous chalk of G. Araif El Naqa and in the Paleocene and Lower Eocene rocks on the southern side of the G. Maghara fold. Figure 21 shows an
Fig. 21. Detailed portion of the seismic section in Figure 18 showing thinning of the Upper Maastrichtian chalk (rock unit between the Ku3 and Ku4 reflectors) on the flanks of the G. Maghara and G. Yelleg anticlines.
60
A. R. MOUSTAFA
intra-Senonian deformation in the broad syncline separating the G. Maghara and G. Yelleg folds, where the Upper Senonian chalk is thicker in the trough of this syncline, compared with its flanks where the chalk was deposited unconformably above the tilted Santonian rocks. Such change in thickness indicates post-Santonian folding in the area. An intra-Upper Senonian angular unconformity was reported on seismic sections of the northern Western Desert of Egypt by Moustafa (2002, 2008). This angular unconformity constrains the acme of Late Cretaceous compressive deformation in north Egypt to be of Campanian age. The Late Cretaceous to pre-Miocene compressional deformation in North Sinai was followed by another deformation leading to dextral offset of the Lower Miocene basalt dyke of G. El Minsherah (120 m of dextral slip reported by Moustafa & Yousif 1990). This post-Early Miocene reactivation of the faults of the Sinai hinge belt is thought to be associated with the slip on the Dead Sea transform (Moustafa & Khalil 1994).
Synthesis The structures of North Sinai are related to the movement of the African plate relative to the nearby plates. Divergent and convergent movements between the African and Eurasian plates account for most of the tectonic deformation of this area. Extensional deformation during the Jurassic and probably part of the Cretaceous is related to the divergent movement between the Afro-Arabian and Eurasian plates that led to opening of the Neotethys in the Eastern Mediterranean area associated with northward drift of some fragments away from northern Africa (Robertson & Dixon 1984). According to Biju-Duval et al. (1979), Argyriadis et al. (1980), Garfunkel & Derin (1984), Dercourt et al. (1986), Mart (1987) and Stampfli et al. (2001), opening started in the Late Triassic –Early Jurassic time. This divergence led to thinning of the continental crust in North Sinai where crustal thickness decreases from 32 km at latitude 288N to about 27 km at the Mediterranean coast (Tealeb 1985; El-Azoni 1992). This divergent movement led to the development of several extensional basins in North Sinai (e.g. Maghara, Yelleg and Halal basins) in the present onshore area. Other extensional basins also exist in the offshore Mediterranean area (e.g. Mango basin, among others) but are not dealt with in the present study. The southern boundary of the extensional basins area is the Sinai hinge belt, which is defined by ENE –WSW oriented faults. Other faults were probably formed at the same time in the platform area lying south of the Sinai hinge belt (e.g. the Themed fault).
Convergence between the African and Eurasian plates in Late Cretaceous –Cenozoic times led to the compressional deformation in North Sinai that caused basin inversion. The direction of the maximum principal stress axis at that time was oriented NW –SE based on the study of mesostructures and/or the analysis of fault–slip data in G. Yelleg (Fig. 9), G. Maghara (Moustafa 2005) and the Sinai hinge belt (Fig. 13). Youssef (1968), Smith (1971), Eyal & Reches (1983) and Letouzey (1986) also reported similar direction of compression. Such convergence started in Late Cretaceous time, reached its acme in Campanian time and continued mildly until the pre-Miocene. Guiraud & Bosworth (1997) and Guiraud (1998) believe that inversion took place during the Santonian and latest Maastrichtian period in north Africa. The acme of the compressive deformation is almost contemporaneous with the collision of Afro-Arabia with Eurasia at the Bitlis suture (Hempton 1985) and the obduction of the Baer – Bassit ophiolites in NW Syria (Delaloye & Wagner 1984). Mild continued compressive deformation till the pre-Miocene time is related to further convergence between the two plates. Compressional deformation in northern Sinai seems to have been relieved during the OligoMiocene time when the Gulf of Suez-ancestral Red Sea rifting took place (Garfunkel & Bartov 1977). Continued deformation on the Sinai hinge belt in post-Early Miocene time is probably associated with the sinistral slip on the Dead Sea transform. The eastern parts of the faults of the Sinai hinge belt were probably dragged by the Dead Sea transform in response to the left-lateral slip in post-Miocene times leading to post-Miocene reactivation of the Sinai Hinge belt (Moustafa & Khalil 1994).
Conclusions Detailed field mapping of North Sinai structures combined with ample subsurface (seismic and borehole) data support the following statements. (1) North Sinai is a province of inverted basins that includes three main inverted structures in G. Maghara, G. Yelleg and G. Halal. The province extends northwards and includes other inverted basins in the Sinai offshore area and eastward to the Dead Sea transform, including other folds in the Naqb Desert. (2) The main asymmetrical folds of the Maghara, Yelleg and Halal areas are fault-propagation folds formed during basin inversion. (3) Neither wrenching nor thin-skinned deformation models previously proposed for the G. Maghara, Yelleg and Halal area could have
NORTH SINAI FOLDS
(4)
(5)
(6)
formed the existing structures. Although the deep geometry of faults in North Sinai is seen on a few seismic sections (e.g. Fig. 18), similarities between the North Sinai structures and those in the northern Western Desert of Egypt (where enough seismic and borehole data prove basement involvement) make it possible to postulate that the North Sinai faults affect the basement and, therefore, the deformation is thick-skinned. The inverted basins province is separated from the tectonically stable area to the south by the Sinai hinge belt. The Themed fault is one of the long faults dissecting the tectonically stable area. Dextral transpressive deformation characterizes the Sinai hinge belt, whereas pure dextral strike –slip deformation characterizes the Themed fault. Three main phases of deformation have affected North Sinai since the Mesozoic time: (a) The earliest (D1) deformation is an extensional deformation of Jurassic (and probably Cretaceous) age related to the divergence between the Afro-Arabian and Eurasian plates and opening of the Neotethys. This deformation formed extensional basins in the G. Maghara, Yelleg and Halal areas, as well as in the North Sinai offshore area and probably in the Naqb Desert. (b) The D2 deformation is compressional and related to the convergence between the African and Eurasian plates. It led to inversion of the extensional basins and dextral strike–slip movement on the Sinai hinge belt and the Themed fault. The acme of this deformation was in the Campanian associated with the collision of AfroArabia with Eurasia at the Bitlis suture and the obduction of the Bassit ophiolites in NW Syria. This deformation continued mildly until the pre-Miocene time. (c) The D3 deformation is post-Miocene and is probably associated with the sinistral strike–slip movement on the Dead Sea transform. This deformation reactivated the Sinai hinge belt and causes the present-day seismicity on faults in the eastern part of the hinge belt.
Fieldwork for this study was partly sponsored by Ain Shams University and the VW Foundation through Joint Research Project between Bremen University (Germany) and Ain Shams University (Egypt). Subsurface data were provided by the Egyptian General Petroleum Corporation. Parts of this paper are based on joint publications with some of my graduate students and colleagues. I specially
61
mention the late M. E. Salama (Mansoura University), H. G. Fouda (now with Saudi Aramco), and W. Hashem (Ain Shams University) to whom I am indebted. S. Khalil (Suez Canal University) helped me in the analysis of fault-slip data. Fruitful discussions with J. Kuss and M. Bachmann (Bremen University) were valuable for understanding the stratigraphy of Cretaceous rocks of G. Maghara area. Critical review of the manuscript by M. I. Youssef (Ain Shams University) and constructive comments by Geological Society referees P. Bentham and P. Leturmy are highly appreciated.
References Abd-Allah, M. A., Moustafa, A. R. & Hashem, W. A. 2004. Structural characteristics and analysis of the Gebel El Halal fold, Northeast Sinai, Egypt. Middle East Research Center, Ain Shams University, Earth Science Series, 18, 1 –26. Abdel Aal, A. & Lelek, J. J. 1994. Structural development of the northern Sinai, Egypt and its implications on the hydrocarbon prospectivity of the Mesozoic. Proceedings of Middle East Petroleum Geoscience conference (GEO 94), Bahrain, 1, 15–30. Abdel Aal, A., Day, R. A. & Lelek, J. J. 1992. Structural evolution and styles of the northern Sinai, Egypt. Proceedings of 11th Egyptian General Petroleum Corporation Exploration and Production Conference, Cairo, 1, 546– 562. Abd El-Aziz, M., Moustafa, A. R. & Said, S. E. 1998. Impact of basin inversion on hydrocarbon habitat in the Qarun Concession, Western Desert, Egypt. Proceedings of 14th Egyptian General Petroleum Corporation Exploration and Production Conference, Cairo, 1, 139– 155. Abd El-Gawad, A. M. & Ibraheem, I. M. 2005. Gravity-deduced faults associated with the northern Sinai Syrian arc folds. Middle East Research Center, Ain Shams University, Earth Science Series, 19, 165– 178. Abed, M. M., Ayyad, S. N. & Abu Zeid, R. H. 1996. Stratigraphic classification of Triassic-Cretaceous rocks of Gebel Arif El-Naga, Northeastern Sinai, Egypt. Newsletters on Stratigraphy, 33, 117 –131. Abu Zeid, R. H. 2007. Paleoenvironmental significance of Early Cretaceous foraminifera from northern Sinai, Egypt. Cretaceous Research, 28, 765–784. Al-Far, D. A. 1966. Geology and coal deposits of Gebel El Maghara (north Sinai). Geolocial Survey of Egypt. Paper no. 37, 59. Allmendinger, R. W. 2001. FaultKinWin version 1.1: a program for analyzing fault slip data for Windows computers. Argyriadis, I., De Graciansky, P. C., Marcoux, J. & Ricou, L. E. 1980. The opening of the Mesozoic Tethys between Eurasia and Arabia–Africa. Memoir du Bureau de Recherches Geologiques et Minieres, 155, 199–214. Ayyad, M. H. & Darwish, M. 1996. Syrian arc structure: a unifying model of inverted basins and hydrocarbon occurrences in north Egypt. Proceedings of 13th Egyptian General Petroleum Corporation Petroleum Conference, 1, 40– 59.
62
A. R. MOUSTAFA
Bachmann, M. & Kuss, J. 1998. The Middle Cretaceous carbonate ramp of the northern Sinai: sequence stratigraphy and facies distribution. In: Wright, V. P. & Burchette, T. P. (eds) Carbonate Ramps. Geological Society, London, Special Publications, 149, 253–280. Bartov, Y. 1974. A structural and paleogeographical study of the central Sinai faults and domes. PhD dissertation, Hebrew University, Israel, 134. (Quoted by Bartov, Y., Lewy, Z., Steinitz, G. & Zak, I. 1980. Mesozoic and Tertiary stratigraphy, paleogeography and structural history of the Gebel Areif en Naqa area, eastern Sinai. Israel Journal of Earth Sciences, 29, 114–139). Bartov, Y. & Steinitz, G. 1977. The Judea and Mount Scopus Groups in the Negev and Sinai with trend surface analysis of the thickness data. Israel Journal of Earth Sciences, 26, 119–148. Bartov, Y., Lewy, Z., Steinitz, G. & Zak, I. 1980. Mesozoic and Tertiary stratigraphy, Palaeogeography and structural history of the Gebel Areif en Naqa area, eastern Sinai. Israel Journal of Earth Sciences, 29, 114–139. Biju-Duval, B., Letouzey, J. & Montadert, L. 1979. Variety of margins and deep basins in the Mediterranean. American Association of Petroleum Geologists Memoir, 29, 293–317. Colletta, B., Le Quellec, P., Letouzey, J. & Moretti, I. 1988. Longitudinal evolution of the Suez rift structure (Egypt). Tectonophysics, 153, 221–233. Delaloye, M. & Wagner, J. 1984. Ophiolites and volcanic activity near the western edge of the Arabian plate. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 225. Dercourt, J., Zonenshain, L. P. et al. 1986. Geological evolution of the Tethys belt from the Atlantic to the Pamirs since the Lias. Tectonophysics, 123, 241–315. EGYPTIAN GENERAL PETROLEUM COMPANY 1985. Bouguer gravity map of Egypt, Scale 1:100 000. Eicher, D. B. 1947. Micropaleontology of the Triassic of north Sinai. Bulletin Institute de Egypte, 28, 87–92. El-Azoni, M. A. 1992. Two-dimensional gravity modeling of the basement down to Mohorovicic discontinuity, northern Western Desert, Egypt. Proceedings of 11th Egyptian General Petroleum Corporation Exploration and Production Conference, Cairo, 2, 427– 434. Eyal, Y. & Reches, Z. 1983. Tectonic analysis of the Dead Sea rift region since the Late Cretaceous based on mesostructures. Tectonics, 2, 167. Eyal, M., Bartov, Y., Shimron, A. E. & Bentor, Y. K. 1980. Sinai geological map, scale 1:500 000. Survey of Israel. Farag, I. & Shata, A. 1954. Detailed geologic survey of El Minsherah area. Bulletin Institut Desert de Egypte, 4, 1 –82. Garfunkel, Z. & Bartov, Y. 1977. The tectonics of the Suez rift. Geological Survey of Israel Bulletin, 71, 44. Garfunkel, Z. & Derin, B. 1984. Permian-early Mesozoic tectonism and continental margin formation in Israel and its implications for the history of the Eastern Mediterranean. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 178–201.
GEOLOGIC SURVEY OF EGYPT 1981. Geological Map of Egypt, scale 1:2 000 000. Egyptian Geological Survey and Mining Authority, Egypt. GEOLOGICAL SURVEY OF ISRAEL 2000. Map of Lithostratigraphic Groups. Israel and environs: Current Research, 12. GEOLOGICAL SURVEY OF ISRAEL 2001. Geological Map of Israel: Har Loz Sheet, scale 1:50 000. Guiraud, R. 1998. Mesozoic rifting and basin inversion along the northern African Tethyan margin: an overview. In: Macgregor, D. S., Moody, R. T. J. & Clark-Lowes, D. D. (eds) Petroleum Geology of North Africa. Geological Society, London, Special Publications, 132, 217–229. Guiraud, R. & Bosworth, B. 1997. Senonian basin inversion and rejuvenation of rifting in Africa and Arabia: synthesis and implications to plate-scale tectonics. Tectonophysics, 282, 39– 82. Hempton, M. P. 1985. Structure and deformation history of the Bitlis suture near Lake Hazar, southeastern Turkey. Geological Society of America Bulletin, 96, 233–43. Hirsch, F. & Picard, L. 1988. The Jurassic facies in the Levant. Journal of Petroleum Geology, 11, 277– 308. Jenkins, D. A. 1990. North and central Sinai (Chapter 19). In: Said, R. (ed.) The Geology of Egypt. A. A. Balkema, Rotterdam, 361–380. Khalil, S. M. 1991. Geological Studies on the North Eastern Area of Gebel Yelleg, North Central Sinai, Egypt. MSc Thesis, Suez Canal University, 107. Khalil, M. H. & Moustafa, A. R. 1994. Tectonic framework of northern Egypt and its bearing on hydrocarbon exploration. Proceedings of 12th Egyptian General Petroleum Corporation Petroleum Exploration and Production Conference, Cario, 1, 67–86. Krenkel, E. 1925. Geologie Afrikas. Verlag von Gebruder, Berlin, 461. Kuss, J. & Boukhary, M. 2008. A new upper Oligocene marine record from northern Sinai (Egypt) and its paleogeographic context. GeoArabia, 13, 59–84. Letouzey, J. 1986. Cenozoic paleo-stress pattern in the Alpine Foreland and structural interpretation in a platform basin. Tectonophysics, 132, 215– 231. Lewy, Z. 1975. The geological history of southern Israel and Sinai during the Coniacian. Israel Journal of Earth Sciences, 24, 29– 43. Luning, S., Kuss, J., Bachmann, M., Marzouk, A. M. & Morsi, A. M. 1998. Sedimentary response to basin inversion: Mid Cretaceous–Early Tertiary pre- to syndeformational deposition at Areif El Naqa anticline (Sinai, Egypt). Facies, 38, 103–136. Mart, Y. 1987. Superpositional tectonic patterns along the continental margin of the southeastern Mediterranean: a review. Tectonophysics, 140, 213– 232. Moon, F. W. & Sadek, H. 1921. Topography and geology of north Sinai, Egypt. Petroleum Research Bulletin, Cairo, 10, 154. Morley, C. K., Nelson, R. A., Patton, T. L. & Munn, S. G. 1990. Transfer zones in the East African Rift system and their relevance to hydrocarbon exploration in rifts. American Association of Petroleum Geologists Bulletin, 74, 1234–1253. Moustafa, A. R. 1988. Wrench tectonics in the north Western Desert of Egypt (Abu Roash area, southwest
NORTH SINAI FOLDS of Cairo). Middle East Research Center, Ain Shams University, Earth Science Series, 2, 1 –16. Moustafa, A. R. 2002. Structural style and timing of syrian arc deformation in northern Egypt (abstract). American Association of Petroleum Geologists International meeting, Cairo, October 2002. Moustafa, A. R. 2005. Structural architecture of Late Cretaceous–Early Tertiary inverted structures in northern Sinai (Gebel Maghara area): abstract. 43rd Annual Meeting of the Egyptian Geological Society, Cairo, December 2005. Moustafa, A. R. 2008. Mesozoic–Cenozoic basin evolution in the northern Western Desert of Egypt. In: Salem, M., El-Arnauti, A. & Saleh, A. (eds) The Geology of East Libya, 3, 29–46. Moustafa, A. R. & Gibali, H. 2005. Structural Setting and tectonic evolution of Syrian Arc folds at Gebel Yelleg (North Sinai, Egypt), abstract. First International Conference on the Geology of the Tethys, Tethys Geological Society, Cairo University, November 2005. Moustafa, A. R. & Khalil, M. H. 1989. North Sinai structures and tectonic evolution. Middle East Research Center, Ain Shams University, Earth Science Series, 3, 215– 231. Moustafa, A. R. & Khalil, M. H. 1990. Structural characteristics and tectonic evolution of north Sinai fold belts. In: Said, R. (ed.) The Geology of Egypt. Balkema Publishers, Rotterdam, Netherlands, 381– 389. Moustafa, A. R. & Khalil, M. H. 1994. Rejuvenation of the Eastern Mediterranean passive continental margin in northern and central Sinai: new data from the Themed Fault. Geological Magazine, 131, 435– 448. Moustafa, A. R. & Khalil, S. M. 1995. Rejuvenation of the Tethyan passive continental margin of northern Sinai: deformation style and age (Gebel Yelleg area). Tectonophysics, 241, 225–238. Moustafa, A. R. & Salama, M. E. 2005. The Sinai Hinge Zone: a major crustal boundary in northern Sinai), abstract. 43rd Annual Meeting of the Egyptian Geological Society, Cairo, December 2005. Moustafa, A. R. & Yousif, M. S. 1990. Two-stage wrench deformation at Gebel El Minsherah, north Sinai. Middle East Research Center, Ain Shams University, Earth Science Series, 4, 112– 122. Moustafa, A. R., Khawasik, S. M. & Khalil, S. M. 1991. Late Cretaceous – Tertiary rotational deformation in North Sinai (Gebel Yelleg area). Neues Jahrbuch fur Geologie und Palaontologie, 11, 643–653. Moustafa, A. R., El-Badrawy, R. & Gibali, H. 1998. Pervasive E-ENE oriented faults in the northern Egypt and their relationship to Late Cretaceous petroliferous basins in the northern Western Desert. Proceedings of 14th Egyptian General Petroleum Corporation Exploration and Production Conference, Cairo, 1, 51–67. Noweir, M. A., Al Alfy, Z. & Fawwaz, E. M. 2006. Wrench deformation Syrian arc structures: El Hamraa, Umm Busal and Um Horeiba en echelon anticlines, Mitla Pass, west central Sinai, Egypt. Egyptian Journal of Geology, 50, 265– 287. Orwig, E. R. 1982. Tectonic framework of northern Egypt and the Eastern Mediterranean region. Proceedings 6th Egyptian General Petroleum Corporation Exploration Seminar, Cairo, 1, 1 –16.
63
Osman, R. A., Ahmed, S. M. & Mahmoud, N. I. 2000. Cretaceous– Lower Tertiary rocks at Gebel Halal area, northern Sinai, Egypt: a Stratigraphy and sedimentary history. Proceedings of 5th International Conference on the Geology of the Arab World, Cairo, 1309– 1332. Robertson, A. H. F. & Dixon, J. E. 1984. Introduction: aspects of the geological evolution of the Eastern Mediterranean. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 1 –74. Rybakov, M. & Segev, A. 2004. Top of the crystalline basement in the Levant. Geochemistry, Geophysics, Geosystems, 5, 1 –8. Sadek, H. 1928. The principal structural features of the Peninsula of Sinai. 14th International Geological Congress, Madrid, 1926, 3, 895–900. Said, R. 1962. The Geology of Egypt. Elsevier Pub. Co., Amsterdam, 377. Schlumberger 1984. Well Evaluation Conference, Chapter 1: Geology of Egypt, 1– 64. Schlumberger 1995. Well Evaluation Conference, Chapter 4: Western Desert, 56–71. Sestini, G. 1989. Nile Delta: a review of depositional environments and geological history. In: Whateley, M. K. G. & Pickering, K. T. (eds) Deltas: Sites and Traps for Fossil Fuels. Geological Society, London, Special Publications, 41, 99–127. Shata, A. 1959. Structural development of the Sinai Peninsula (Egypt). Proceedings of the 20th International Geological Congress, 1956, 225–249. Shata, A. 1960. The geology and geomorphology of El Qusaima area. Bulletin Society Geography d’Egypte, 33, 95– 146. Shukri, N. M. 1954. Remarks on the geological structure of Egypt. Bulletin Society Geography d’Egypte, 27, 65–82. Smith, A. G. 1971. Alpine deformation and the oceanic areas of the Tethys, Mediterranean and Atlantic. Geological Society of America Bulletin, 82, 2039–2070. Stampfli, G., Borel, G., Cavazza, W., Mosar, J. & Ziegler, P. A. 2001. The paleotectonic atlas of the PeriTethyan domain. European Geophysical Society (CD). Steinitz, G., Bartov, Y. & Hunziker, J. C. 1978. K–Ar age determinations of some Miocene– Pliocene basalts in Israel: their significance to the tectonics of the rift valley. Geological Magazine, 115, 329–340. Tealeb, A. 1985. Spectral analysis of the gravimetric bouguer anomaly of Egypt for crustal thickness studies. Helwan Institute for Applied Geophysics Bulletin, 5B, 63–82. Visser, W. A. 1941. Geological Report on Gebel Yelleq (north central Sinai, Egypt). Internal report, Standard Oil Company, Cairo, 21. Youssef, M. I. 1968. Structural pattern of Egypt and its interpretation. American Association of Petroleum Geologists Bulletin, 52, 601– 614. Ziegler, P. A. 1992. North Sea rift system. Tectonophysics, 208, 55– 75. Ziko, A., Darwish, M. & Eweda, S. 1993. Late Cretaceous–Early Tertiary stratigraphy of the Themed area, east central Sinai, Egypt. Neues Jahrbuch fu¨r Geologie und Palaontologie, 3, 135– 149.
Structural setting and tectonic evolution of offshore North Sinai, Egypt M. YOUSEF1*, A. R. MOUSTAFA1 & M. SHANN2 1
Department of Geology, Ain Shams University, Cairo 11566, Egypt
2
BP International Exploration, Chertsey Road, Sunbury on Thames, Middlesex TW16 7LN, UK *Corresponding author (e-mail:
[email protected]) Abstract: The offshore area of North Sinai represents the northern extension of the Syrian Arc inversion structures into the southeastern Mediterranean region. Integration of detailed seismic interpretation of key tectonic events in offshore North Sinai and recently acquired gravity and magnetic data reveal structural deformation represented by large buried inversion anticlines that have played an important role in the geological history and hydrocarbon potential of the area. This tectonic inversion took place in the Late Mesozoic and continued slightly during the Cenozoic, and formed NE-trending asymmetrical folds. Three different phases of deformation have been detected in offshore North Sinai: (1) A Jurassic–Early Cretaceous extensional phase, which formed NE trending normal faults bounding asymmetrical half-grabens, (2) Post-Santonian– Middle Miocene positive inversion of these faults and half-grabens and (3) Post–Middle Miocene subsidence. A set of tectonosequences related to the opening and the subsequent convergence of the Tethys was mapped. Each identified tectonosequence has its own unique drive mechanism, geometry, and location with respect to the plate boundary. Recognition of these elements allows illustration of the Tethyan basin evolution of offshore North Sinai through time as well as understanding the tectonic and stratigraphic framework and effective prediction of the petroleum system.
Although the geology of the North Sinai Peninsula has been the subject of study of several researchers, little was written about the offshore area. Being located on the Tethyan margin of the AfricanArabian plate, offshore North Sinai exhibits a sequence of shallow marine and continental sedimentary rocks, which may be up to 12 km thick (Guiraud & Bosworth 1997). Large synsedimentary normal faults, with nearly N108E and N708E average trends, were very active during Permian–Early Cretaceous times (Guiraud & Bosworth 1997). From the Early Senonian, both onshore and offshore domains of this region registered folding and faulting (Ginzburg et al. 1975; Bartov et al. 1980; Jenkins 1990). The compression caused reverse faulting along the old normal faults and the initiation of basin inversion. These events resulted in a well-exposed fold belt of NE– SW oriented, doubly plunging folds referred to as the Syrian Arc fold belt (Krenkel 1925), Figure 1. These folds are of different sizes, including folds tens of kilometres long, several kilometres long, and many others that are less than 2 km in length (Moustafa & Khalil 1989). The area selected for detailed subsurface study in the present study is geographically located in the southeastern part of the Mediterranean Sea, offshore North Sinai (Fig. 2). It covers an area
of about 6630 km2 and is bounded by latitudes 31830 and 318460 N and by longitudes 328410 and 348130 E. The aim of this paper is to address the tectonostratigraphic evolution of offshore North Sinai Basin and to document the phases of rifting and basin inversion, which affected the North Sinai margin of the Tethys from the Jurassic to Middle Miocene.
Regional setting The eastern Mediterranean passive continental margin was formed in the Early Mesozoic, when widespread rifting occurred in the entire Tethys area (Dewey et al. 1973; Garfunkel & Derin 1984; Moustafa & Khalil 1994). This margin was characterized by shallow platform carbonates in the east and by deep-water carbonates on the slope and basin in the west (Bein & Gvirtzman 1977; Druckman et al. 1995). Since the early Late Senonian time, the collision of the African-Arabian plate with the Eurasian plate resulted in the development of the Syrian Arc fold belt (Fig. 1; Neev & Ben-Avraham 1977; Moustafa & Khalil 1989; Eyal 1996; Buchbinder & Zilberman 1997; Garfunkel 1998). The main deformation in the
From: HOMBERG , C. & BACHMANN , M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 65–84. DOI: 10.1144/SP341.4 0305-8719/10/$15.00 # The Geological Society of London 2010.
66
M. YOUSEF ET AL.
Fig. 1. Simplified structural form-line map showing the main structural features of North Sinai and al-Naqab Desert after Khalil & Moustafa (1994).
Fig. 2. Location map of the study area showing the used 2D seismic sections and boreholes.
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
Syrian Arc fold belt continued into the post-Early Miocene (Moustafa & Yousif 1990; Moustafa et al. 1991; Tibor et al. 1992; Druckman et al. 1995; Eyal 1996). During the Early Miocene the Syrian Arc began to emerge on the Levant margin (Buchbinder & Zilberman 1997). The shelf area underwent localized tectonical uplift and became intermittently emergent (Buchbinder & Zilberman 1997). At the end of the Miocene, the eastern Mediterranean passive continental margin underwent extensive erosion and evaporite deposition (Gvirtzman & Buchbinder 1978; Druckman et al. 1995), a pattern common to most of the Mediterranean basin (Hsu¨ et al. 1978). Thick evaporites (.2 km) were deposited on the seafloor, while the continental margin was affected by erosion producing deeply incised valleys (Cita & Ryan 1978; Garfunkel & Almagor 1987; Abdel Aal et al. 2001). The Pliocene to recent sediments in large areas of the eastern Mediterranean continental margin and adjacent basin are affected by thin-skinned deformation (Bertoni & Cartwright 2006) owing to salt mobilization and shelf loading that caused the collapse and landward tilt of these deposits above the Messinian evaporites (Tibor et al. 1992).
Data and methodology The offshore subsurface data used for this study comprise about 6444 km (3995 miles) of twodimensional (2D) seismic reflection sections with maximum recorded two-way time of 5 to 6 s, 25 m shot point interval, 48-fold geophone coverage, and a line spacing of 400 m, in addition to bore-hole data (wire-line logs and biostratigraphic data) of 11 wells, and gravity and magnetic data. Figures 2 and 3 illustrate the database and the correlation between seismic lines and wells. Fault and horizon interpretations were carried out on a workstation using Landmark’s SeisWorks software. A seismic tectonostratigraphic approach was followed in the present study. Regionally persistent seismic horizons were mapped across the study area. After regionally consistent seismic tectonostratigraphic framework was constructed, biostratigraphic data were used to calibrate the seismic interpretation. Regional events were dated based on ties to apparent depositional hiatuses (condensed sections, erosional unconformities, periods of nondeposition, etc.) interpreted from biostratigraphic well control. Critical well ties were achieved using check shot surveys and synthetic seismograms. There was generally good correlation between the seismic tectonostratigraphy and the biostratigraphic interpretations. Integration of these interpretations
67
resulted in a consistent and robust tectonostratigraphic framework.
Tectonosequences Seismic units bounded by regional unconformities are tectonosequences that represent deposition during tectonically controlled phases of basin development. The tectonosequence boundaries reflect major changes in regional tectonic settings (Hubbard 1988) and they are referred to as TB1 to TB5 from oldest to youngest (Figs 3 & 4). The tectonosequences are referred to as TS1 to TS5 (Table 1 & Fig. 5) from oldest to youngest. The geometry of the tectonosequences is illustrated by 2D seismic sections. These sections are located in Figure 2 and are either oriented along the strike of the basin or orthogonal to the major tectonic elements. These data allied to biostratigraphical information from wells, provide important evidence for the timing of deformation in offshore North Sinai. Five tectonosequences interpreted on seismic sections are summarized below (from oldest to youngest); Figure 6.
Tectonosequence 1 (Upper Jurassic– Santonian) Most of the rocks of tectonosequence 1 represents a syn-rift package that comprises the initial clastic, basin-fill and shows significant growth into opposed, fault bounded depocentres. They were deposited during the development of Mesozoic ENE oriented rift basins in North Sinai as well as in the other parts of north Egypt (Orwig 1982; Moustafa et al. 1998), owing to opening of the Neotethys between North Africa-Arabia and Eurasia (Biju-Duval et al. 1979; Argyriadis et al. 1980). The top of tectonosequence 1 is truncated by erosion in the Mango area (Fig. 5), but the boundary becomes a disconformity in the western part of the area of study. Tectonosequence 1 consists predominantly of clastic-dominated successions of the Cretaceous age unconformably overlying the shallow marine Upper Jurassic carbonates as a result of the sharp drop in global sea level in the Late Tithonian (Guiraud 1998). The basin geometry of the tectonosequence 1, illustrated by the isochron map (Fig. 7), shows several sub-basins bounded by extensional faults.
Tectonosequence 2 (Campanian – Maastrichtian) Deposition of tectonosequence 2 locally continued away from the crests of the evolving anticlines leading to onlap of tectonosequence 2 sediments
68
M. YOUSEF ET AL.
Fig. 3. Correlation panel between the main lithostratigraphic units on well data and the interpreted seismic packages. TB1– TB5 refer to tectonosequence boundaries 1 –5.
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
69
Fig. 4. Offshore North Sinai basin cartoon, showing key surfaces mapped. TB1–TB5 refer to tectonosequence boundaries 1– 5.
on older rocks in the flanks. It is thin (,200 ms TWT) over the crests of the anticlines and gradually thickens (.450 ms TWT) in the synclines (Figs 5 & 8). This syn-compressional tectonosequence consists predominantly of shelf and slope mudstones. The onset of Tethys convergence between the Eurasian and African plates in postSantonian time was characterized by positive inversion of the previously formed NE trending extensional basins in offshore North Sinai (Yousef et al. 2006). This inversion led to the development of a series of NE–SW trending folds (Fig. 9a).
Tectonosequence 3 (Palaeocene – Upper Oligocene) Tectonosequence 3 is the second syn-compression tectonosequence identified in offshore North Sinai and is also associated with the folding event, but with a magnitude less than that in tectonosequence 2. It is characterized by onlap and truncation relationships at its margins owing to the active deformation of the antiforms, giving an indication of
the continuation of the post-Santonian basin inversion until the Late Oligocene. Folding and faulting are not as extensive as those of TB2 and TB3 but only gentle folding is obvious (Fig. 9b). Tectonosequence 3 represents deposition in shelf and slope environments. It is also thin (,200 ms TWT) over the crests of the anticlines and gradually thickens (.450 ms TWT) in the synclines (Figs 5 & 8).
Tectonosequence 4 (Miocene) Tectonosequence 4 defines a northward-thickening sedimentary wedge, which thins locally over the crests of the anticlines, delineated by the TB4 and TB5 at its base and top, respectively (Fig. 5). It is almost missing along the present-day shoreline in the south, but thicken to .800 ms in the northern part of the study area (Fig. 5). Tectonosequence 4 comprises open marine deposits, as well as an evaporite sequence deposited during the Late Miocene (Messinian) drying up of the Mediterranean (Hsu¨ et al. 1973). This evaporite sequence consists predominantly of anhydrite and gypsum with
70
M. YOUSEF ET AL.
Table 1. Summary of the principal features of the tectonostratigraphic units in offshore North Sinai Tectonosequence
Age
TWT thickness (ms)
TS5
Pliocene–Pleistocene
TS4
Miocene
400– 600
TS3
Palaeocene –Late Oligocene
700– 1000
TS2 TS1
Campanian –Maastrichtian Late Jurassic –Santonian
local halite deposits that irregularly cover marine sediments of the Middle to Late Miocene age. This is linked to the final convergence of the African and Eurasian plates at the end of the Miocene, where the Late Messinian records the closing of the Strait of Gibraltar and the subsequent evaporation of the Mediterranean Basin (Dolson et al. 2001).
Tectonosequence 5 (Pliocene – Pleistocene) Tectonosequence 5 is the youngest tectonosequence in offshore North Sinai and is represented by a northward-thickening wedge above the TB5. It increases in thickness from nearly 200 ms along the present-day shoreline in the south to nearly 2500 ms along the northern edge of the study area in the north (Figs 5 & 8). Tectonosequence 5 is characterized as a package of relatively undeformed, horizontal to gently inclined parallel reflectors showing moderate to high amplitude and remarkable lateral continuity (Fig. 5). It exhibits a wedgelike geometry and displays extensional faulting with detachment at or near its base. This Neogene extension was gravity driven with general extension directions to the NE. It comprises mainly Nile-derived sediments (marl, shale, and sandstone) prograding into the Mediterranean Sea.
Evidence for basin inversion in offshore North Sinai Tectonic structures The continental margin of offshore North Sinai is a unique example of an apparently simply structured continental margin that, when studied carefully,
0 – 2200
0 – 500 1500– 2000
Lithology Shale and siltstone with sandstone intercalations Shale, mudstone and siltstone with intercalations of sandstone and minor limestone interbeds Evaporites of restricted realm Dark grey to dark green, locally pyritic or glauconitic shale Shallow shelf limestone with few dolomite streaks Shallow shelf carbonates Dark coloured shale, occasionally highly calcareous, with interbeds of sand, sandstone and minor streaks of dolomitic limestone
displays complex structural patterns owing to resurgent and superimposed tectonic activity from the Early Mesozoic to Recent. Detailed study of seismic data interpreted with recently acquired gravity and magnetic data, revealed structural deformation represented by large buried inversion anticlines that played an important role in the geological history and the hydrocarbon potential of the area. The interpretation of different tectonics structures in the area is given in the following paragraphs. Mango structure. The Mango structure is a 24-km long, ENE-oriented (Fig. 10d), doubly plunging anticline with a relatively steep northwestern flank and breached crest. Three wells have been drilled to test the potential of the Lower Cretaceous sandstones and Upper Cretaceous to Eocene carbonates in this structure. The Lower Cretaceous succession is conformable, with the overlying and underlying sequences downdip in the southern flank and the severe erosion at the crestal part of the structure was not observed downdip in the southern flank. The stratigraphic succession, between the Top Jurassic and Top Santonian represented by tectonosequence 1, shows an increase in thickness towards the main bounding fault indicating normal slip on that fault during Jurassic– Santonian time. On the other hand, tectonosequences 2 and 3 are progressively onlapping the northward uprising tectonosequence 1. The east-northeasterly orientation of the Mango doubly plunging anticline, as with other inverted structures of offshore North Sinai, resulted from inversion-related folding over a deep-seated ENEoriented fault. Minor NW-oriented normal faults (Fig. 10b) on the northeastern side of the doubly plunging anticline indicate lengthening in the NE– SW direction, as elsewhere in offshore and
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
71
Fig. 5. 2D seismic section through the Mango structure, offshore North Sinai (see location in Fig. 2). Fault-controlled subsidence in the offshore North Sinai basin resulted in growth of the Upper Jurassic– Santonian TS1 towards the Mango structure main bounding Fault. Later inversion formed the asymmetrical anticline affecting the TS1– TS4. TB1–TB5 refer to tectonosequence boundaries 1 –5.
onshore North Sinai (Moustafa & Khalil 1989; Abd-Allah et al. 2004). Goliath structure. The Goliath structure is a 20-km long, NNE-oriented (Fig. 10d), doubly plunging anticline with a relatively steep NW flank on the northeastern side of the area of study (Fig. 9a). Goliath-1 well was drilled on the crest of the structure and indicated a major stratigraphic break between the Coniacian– Maastrichtian carbonates and the underlying Lower Albian clastics of tectonosequence 1 (Fig. 11a). The Goliath structure
is bounded on the north by a NNE-oriented fault and, like the Mango structure, it represents an example of inverted Late Jurassic– Early Cretaceous half-graben. North Sinai 21-1 structure. The North Sinai 21-1 structure is a 20-km long, ENE-oriented (Fig. 10d), doubly plunging anticline in the central part of the area of study (Fig. 9a). It has a relatively steep northwestern flank. The North Sinai 21-1 well was drilled on the crest of the structure. Updip thickening of tectonosequence 1 (Fig. 8) was
72
M. YOUSEF ET AL.
side of the area of study (Fig. 9a). Ziv-1 well was drilled on the crest of the structure and indicated an increase in thickness of tectonosequence 1 towards the main bounding fault (Fig. 8). The Ziv structure is bounded on the south by a NNE-oriented high-angle reverse fault representing an example of inverted Late Jurassic–Early Cretaceous half graben. This reverse fault has the same orientations of the normal fault affecting the Top Jurassic surface (TB1) (Fig. 10b). Similar onshore observations were reported by Shata (1959), Al-Far (1966), Bartov et al. (1980), Jenkins (1990) and Moustafa (2010).
Fig. 6. Tectonostratigraphy of offshore North Sinai basin.
observed in this structure in addition to an extensive erosional truncation giving indication of an early extensional phase during the Late Jurassic to the Early Cretaceous, followed by a compressional phase starting from the post-Santonian time. On the other hand, the overlying tectonosequences 2 and 3 show a rather progressive onlapping on the crest of the elevated structure. Ziv structure. The Ziv structure is an 8-km long, NNE-oriented (Fig. 10d), doubly plunging anticline with a relatively steep SE flank on the northeastern
Tineh structure. The Tineh structure is a 7-km long, NE-oriented (Fig. 10d), doubly plunging anticline. Three wells have been drilled to explore and evaluate the hydrocarbon potential of Oligocene clastic sequence in this structure. The Tineh-1 well, drilled on the crest of the structure, encountered significant oil indications in some sand reservoirs, interbedded with shales, within the Late Oligocene section. Numerous grabens and half grabens formed during the rifting stage and block tilting recognized in Israel and Sinai (Bartov et al. 1980), as well as in northern Egypt (Keely & Wallis 1991) characterize the Top Jurassic (TB1), Figures 5 and 8. Rose diagram of the mapped faults (Fig. 10a) indicates that they are oriented N30–60E. A few normal faults (Fig. 12) are affecting the Top Miocene (TB5), which acts as a detachment surface for such faults to the north of the study area. A rose diagram of the mapped faults at the Top Miocene surface (Fig. 10c) indicates that they are oriented N60 –70E. TB5 is a major erosional unconformity in the Late Miocene (Fig. 11a, b), which marks a change in the reflectivity pattern (Fig. 5) related to the transition from the shale of tectonosequence 5 (Pliocene–Pleistocene) to the Messinian evaporite complex of the uppermost part of tectonosequence 4 (Miocene). This change is interpreted to be associated with the presence of Messinian evaporite-bearing series (Vidal et al. 2000). This event is proposed herein to be a tectonically driven event related to late-stage convergence of the African plate against Europe causing the closure of the Strait of Gibraltar and evaporation of the Mediterranean Sea (Halbouty & El Baz 1992).
Structural evolution of offshore North Sinai The North Sinai fold belt (part of the so-called Syrian Arc system) shows examples of inversion structures represented by asymmetrical, doubly plunging anticlines that have thick syn-extension strata in the core. These inversion anticlines occur
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
73
Fig. 7. Time-thickness (isochron) map of offshore North Sinai basin for tectonosequences 1. Contour interval is of 200 ms.
onshore (Fig. 1) and also extend offshore as indicated in the present study (Figs 5, 8 & 11b). In order to simulate the structural evolution of the inverted structures in offshore North Sinai, ‘seismic flattening’ was done by Landmark’s SeisWorks/2D software. This operation allows simulating the appearance of the strata below the flattened horizon at the time that horizon was
deposited. Seismic displays before and after the flatten option have been activated for a Goliathinverted structure in the northeastern part of the area of study, as will be summarized in the following sections. Goliath-inverted structure. The Goliath structure is a large northeasterly trending anticline (Fig. 9a)
Fig. 8. 2D seismic section through Ziv and North Sinai 21-1 structures, offshore North Sinai, see Figure 2 for location. TB1–TB5 refer to tectonosequence boundaries 1 –5.
74
M. YOUSEF ET AL.
Fig. 9. Time-structure maps of offshore North Sinai basin for tectonosequence boundaries 2 and 4 (a and b). Contour interval is 200 ms.
and has been interpreted as an inversion-related asymmetrical fold. Because of the onlapping of the Campanian–Maastrichtian sediments onto the structure, the main episode of compression is believed to have occurred in post-Santonian time. By flattening the top Jurassic horizon (TB 1) (Fig. 13a), a northwestward thickening Upper Jurassic section is clear towards a southeasterly dipping fault, in the form of a half graben, indicating an extensional phase during that period. This NE– SW trending fault is bounding the Goliath structure on the NW (Fig. 9a). Another stage of extension took place during the Neocomian time. This is evident from the flattening of the Top Neocomian horizon (Fig. 13b). More
syn-depositional tilting and thickening of the Upper Jurassic and Neocomian sections are obvious towards the same main bounding fault. Seismic data show that the Early Mesozoic extensional phase, started from the Late Jurassic (and may be earlier) and continued until the Late Santonian time. This is also clear from flattening at the top Santonian horizon (Fig. 13c). Extension, followed by contraction (Basin Inversion), is evident from flattening of the top Cretaceous horizon (TB 3) (Fig. 13d). As a result of the switch in tectonic mode from extension to compression, the Goliath half graben becomes a positive structural feature, with a significant erosional truncation surface between the Albian
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
75
Fig. 10. Rose diagrams showing the trends of subsurface structural elements in the offshore North Sinai basin.
clastics and the Campanian–Maastrichtian carbonates (Fig. 5). The normal fault bounding the NW side of the Goliath basin was reactivated as a high angle reverse fault with a switch in depocentres from growth into the fault to growth away from the fault, to the SE of the inversion anticline. The syn-inversion Campanian– Maastrichtian sediments are progressively onlapping the flanks of the elevated structure, attaining a minimum thickness over the crest of the elevated structure with gradual increase in thickness over the flanks (Fig. 13d ). Figure 13e shows continuation of the compressive
event until the Oligocene time, leading to a continuous uplift of the hanging wall against the footwall along the main fault. Flattening at the top Miocene (TB 5) is shown in Figure 13f. Owing to the reduced connection of the Mediterranean basin with the world oceans during that time, an unconformity that irregularly cuts down into the underlying successions resulted due to the onset of the Messinian Salinity Crisis. Flattening such surface would only reflect the magnitude of erosion and its inflected topography on the underlying successions (Ayyad 1997). An indication of
76
M. YOUSEF ET AL.
Fig. 11. Seismic sections through the Goliath structure (a) and the south Tineh area (b), see Figure 2 for location. The Top Neocomian is not shown in the footwall of the fault of section (a) owing to bad seismic resolution. TB1– TB5 refer to tectonosequence boundaries 1 –5.
Early Miocene gentle folding is evident from southern Tineh area (Fig. 11b). Timing of deformation. Most of the mapped structures in offshore North Sinai are located within five NE-oriented inverted structures, whereas the
intervening areas are undeformed or only slightly deformed. The five inverted structures are controlled by major deep-seated reverse faults. Three phases of structural deformation affected the Mesozoic –Cenozoic succession in the study area and led to the development of the mapped
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
77
Fig. 12. 2D seismic section showing growth extensional faults affecting tectonosequences 4 and 5, see Figure 2 for location.
structures. These are Late Jurassic to Early Cretaceous, post-Santonian to Middle Miocene, and Neogene. Dating of the post-Santonian deformation is based on the presence of a stratigraphic gap and progressive onlapping between the Campanian– Maastrichtian carbonates and older rocks in the Goliath-1 well. Additional data from the Mango structure constrain the age of this phase of deformation. At that locality, the progressive onlap within tectonosequence 2 onto growing anticlines marks the climax of the basin inversion in northern Egypt. Basin inversion continued during and after the deposition of the Paleocene to Langhian (Middle Miocene) rocks, but its effect can only be seen in close vicinity of the basin-bounding faults. The third phase of deformation is contemporaneous with Neogene regional subsidence of the shelf margin offshore North Sinai. Jurassic–Early Cretaceous rifting. The earliest phase of deformation is represented by rifting in the area of the eastern Mediterranean Basin. It is mostly considered to be of Late Triassic age (Robertson 1998; Guiraud & Bosworth 1999; Garfunkel 2004). In the Western Desert and North Sinai, owing possibly to the lack of information about the Triassic, rare data suggest that significant extension and associated basin subsidence began during Jurassic time (Moustafa et al. 1998). Tectonic deformation was registered along most of northern and Central Africa basins, around the
Jurassic –Cretaceous transition times (Guiraud et al. 2005). Frequent uplift and block-tilting, sometimes accompanied by slight folding, occurred along the northern Egyptian margin and the Levant (Guiraud 1998; Le Roy et al. 1998; Coward & Ries 2003). These deformations, underlined by hiatuses in the series and unconformities, represent the distant effects of stronger tectonic activity that occurred in southeastern Europe, referred to as the ‘Cimmerian’ or ‘Berriasian’ orogenic event (Nikishin et al. 2001; Stampfli et al. 2001). It led to the development of ENE-oriented basins bounded by major normal faults of the same orientation and they have half graben geometry with a northward tilt toward the boundary faults. Detailed seismic interpretation of key tectonic events in the present study has indicated that the Jurassic –Cretaceous rifting took place in several pulses: † Late Jurassic; † Neocomian; † Aptian –Santonian. Post-Santonian–Middle Miocene basin inversion. Basin inversion in the offshore North Sinai started in the post-Santonian time and proceeded by the reverse reactivation of the NE-trending extensional faults (Yousef et al. 2006). These are the same old faults that had normal slip during the early rifting phase. The unconformity detected at the contact between the tectonosequences 1 and 2
78
M. YOUSEF ET AL.
Fig. 13. 2D seismic sections through the Goliath structure (refer to Fig. 11a), flattened on (a) top Jurassic (TB1), (b) top Neocomian and (c) top Santonian (TB2).
marks the start of basin inversion in the area of study. In most of the study area, it was concluded that basin inversion continued during and after the deposition of the Palaeocene to Oligocene rocks, but its effect can only be seen in close vicinity of the inversion anticlines. Continued folding of the Cenozoic rocks up to the top Langhian (Fig. 11b) indicates that convergence leading to positive inversion in the area continued till that time. It also indicates continuation of closure of the Neotethys till that time. In the northern Western Desert, basin inversion started in the Late Cretaceous and probably proceeded by way of oblique-slip reactivation of old NE-oriented basin-bounding faults (Sultan & Halim 1988; Bosworth et al. 1999). Neogene extension. During the Neogene time, post-Middle Miocene sediments were deposited, onlapping the northward-subsiding shelf margin as
a result of a regional subsidence of the North Sinai passive margin. Several NW-oriented normal faults with small throws cut tectonosequences 4 and 5 (Neogene) in the study area (Fig. 12). They were formed in response to gravity driven extension in Miocene and post-Miocene times.
Discussion Basin development and inversion The development of the Mesozoic ENE –NE oriented rift basins in Egypt is probably related to the rifting between North Africa-Arabia and Eurasia (Biju-Duval et al. 1979; Argyriadis et al. 1980; Garfunkel & Derin 1984; Dercourt et al. 1986; Mart 1987) when small continental blocks drifted away from North Africa (Robertson & Dixon 1984) during the opening of the Neotethys and the formation of a passive continental margin
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
Fig. 13. (Continued) (d) top Cretaceous (TB3), (e) top Oligocene (TB4) and (f) top Miocene (TB5).
79
80
M. YOUSEF ET AL.
in northern Egypt and eastern Mediterranean (Moustafa & Khalil 1988, 1989; Cohen et al. 1990; May 1991; Keeley 1994; Guiraud 1998; Moustafa et al. 1998; Ayyad et al. 1998). This phase of rifting was accompanied by alkaline volcanic activity that was reported in different parts in Egypt, for example, Gebel Arif en-Naqa and Um Bogma in Central Sinai (Weissbrod 1969; Bartov et al. 1980; Meneisy 1986), south Eastern Desert (Serencsits et al. 1979), and north Western Desert (El Shazly 1977). But in the study area of offshore North Sinai and its vicinity no volcanic activity related to such rifting event has been recorded. The inversion of the Early Mesozoic basins in North Sinai is related to the convergence between Africa-Arabia and Eurasia, during the closure of the Neotethys (Smith 1971; Moustafa & Khalil 1989; Moustafa et al. 1991, 1998; Moustafa & Khalil 1994, 1995; Guiraud & Bosworth 1997). The study of the Atlantic spreading data revealed that Africa-Arabia moved WNW relative to Eurasia during this event (Smith 1971). This was also synchronous with the obduction of the ophiolites along the northern and northeastern margins of Arabia, the onset of separation of India and Madagascar, and the development of the European Alpine Chain (Guiraud & Bosworth 1997).
Fold development and mechanism of basin inversion The type of deformation of the Syrian Arc System, including offshore North Sinai, was attributed by many authors to thin-skinned tectonic deformation (e.g. Chaimov et al. 1990, 1992, 1993; McBride et al. 1990; Al-Saad et al. 1992; Abdel Aal & Lelek 1994; Guiraud & Bosworth 1997; Bosworth et al. 1999). Although Searle (1994) does not accept this type of deformation, he relates the system to disharmonic folding in the Mesozoic sediments above Triassic evaporites with no major regional thrust. This diversity in the interpretation of the origin of the Syrian Arc folds is probably related to the different studied portions of the Syrian Arc System by different investigators, in addition to the little subsurface data on these folds. On the other hand, the origin of the folds within this system was considered by most authors as thrustrelated folds. These folds were interpreted differently from one area to another. They are considered as drape folds over reverse faults (Moustafa & Khalil 1989) or over thrust faults (Abdel Aal & Lelek 1994), fault-bend folds (Chaimov et al. 1992, 1993; Abdel Aal & Lelek 1994; Salel & Se´guert 1994), forced push-up folds (Moustafa & Yousif 1990; Abdel Aal & Lelek 1994; Moustafa et al. 1998), fault-propagation folds (Salel &
Se´guert 1994; Moustafa & Khalil 1995), and detachment folds (Abdel Aal & Lelek 1994; Salel & Se´guert 1994). The term positive inversion refers to a switch in tectonic mode from extension to contraction, whereas the term negative inversion refers to a switch from contraction to extension (Williams et al. 1989). In positive inversion extensional basins are contracted and become regions of positive structural relief (McClay 1995). In these basins, pre-existing extensional faults are reactivated by reverse slip (Cooper & Williams 1989). Reactivation might affect isolated extensional faults within the basin or all major faults. Syninversion sedimentation over reactivated faults produce growth anticlines that are characteristic of positive structural inversion (McClay 1995). Inversion-related folding, in the area of study, is represented by fault-propagation folds with steep forelimbs and gently dipping back limbs. Faultpropagation folding occurs when a propagating fault loses slip and terminates upsection by transferring its shortening to a fold developing at its tip (Mitra 1990). No intra-basin faults have been evolved and this could be due to the pre-existing structures, with localization of strain and fault lengths as inherited by updip propagation from the pre-existing fabric (Walsh et al. 2002; Paton 2006). The folding and faulting are synchronous as indicated by the studies of Jamison (1987) and Mitra (1990) in other study areas. Inversion is dependent on pre-existing basin configuration in the initial subsidence, usually extensional phase, and the resolution of compressional forces in the later shortening phase (Lowell 1995). However, deformational forces responsible for inversion are oriented from 0–908 to the preexisting basin-bounding faults. The zero direction is pure strike– slip that is not favourable in effecting inversion. Clay model experiments (Lowell 1985) show that pure strike–slip superposed on an earlier normal fault fabric is simply resolved as horizontal slip on the old normal fault surfaces with little to no evidence of the compressional component of strike –slip. Compression at 908 to existing structures should be highly effective in inversion (Letouzey et al. 1990). The offshore North Sinai is a prime example of compression-dominant inversion. During the closure of the Tethys, which resulted because of the convergence of the African-Arabian plate and the Eurasian plate (Olivet et al. 1984) in post-Santonian time, NW– SE oriented compression inverted the NE–SW trending Late Jurassic–Early Cretaceous half grabens (Yousef et al. 2006); Figure 9a. This is further evidenced by reactivation of the older normal faults and possibly creation of new thrusts, so that several kilometres of thrust overlap occurred NE of el-Arish City.
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT
Conclusions Five regional unconformities were observed and mapped on seismic sections in the study area. These tectonosequence boundaries (TB1 to TB5) reflect major changes in regional tectonic settings. Most of the mapped structures in offshore North Sinai are located within five NE-oriented inverted structures, whereas the intervening areas are undeformed or only slightly deformed. The inverted structures are controlled by major deep-seated reverse faults with doubly plunging anticlines in which there is thick syn-extensional strata in the core. Similar inversion anticlines also occur onshore (Moon & Sadek 1921; Sadek 1928; Shata 1959; Bartov et al. 1980; Moustafa & Khalil 1994). The inversion structures, in offshore North Sinai, are characterized by fault-propagation folds with steep frontal limbs and gently dipping back limbs. Five tectonosequences (TS1 to TS5) related to the main deformational events are recognized: (a) Upper Jurassic –Santonian syn-rift tectonosequence (TS1); (b) Upper Santonian–Upper Cretaceous compressional tectonosequence (TS2) deposited contemporaneously in the synforms developed between rising antiforms; (c) Paleocene –Upper Oligocene compressional tectonosequence (TS3) is also associated with the folding event, but with a magnitude less than that of tectonosequence 2; (d) northward thickening Miocene tectonosequence (TS4); (e) wedge-like Pliocene-Pleistocene tectonosequence (TS5), which displays extensional faulting with a detachment near its base. Tectonosequence boundaries (TB1 to TB5) are regional unconformities reflecting major changes in regional tectonic setting. TB1 corresponds to rifting and block tilting that started during the Jurassic or earlier in North Sinai. TB2 corresponds to the end of this rifting phase. TB3 and TB4 correspond to major regional unconformities, denoting a continuous compressional tectonic event in offshore North Sinai till Langhian (Middle Miocene) time. TB5 corresponds to an erosional unconformity in the Late Miocene, culminating with the Messinian crisis at 6.7 Ma. Three phases of structural deformation affected the Mesozoic –Cenozoic succession in offshore North Sinai. Phase 1 is a Jurassic to Early Cretaceous rifting phase, which has been indicated in this study to have taken place in several pulses during the Late Jurassic, Neocomian, and AptianSantonian times. Phase 2 is a post-Santonian to Middle Miocene inversion phase, and Phase 3 represents the Neogene extension phase. BP Egypt is thanked for providing the 2D seismic and well data, hardware and software, and for giving permission to publish this paper. We are grateful to P. Bentham and J. Cotton (BP Egypt), for fruitful discussions regarding
81
the tectonic evolution of offshore North Sinai. M. B. Longacre, M. Hakim and T. Bevan from BP Egypt are also gratefully acknowledged. We would like to thank C. Homberg (Pierre and Marie Curie University), D. Paton and an anonymous referee for constructive comments that improved the manuscript.
References A BD -A LLAH , A. M. A., M OUSTAFA , A. R. & H ASHEM , W. A. 2004. Structural Characteristics and analysis of the Gebel El Halal Fold, northeast Sinai, Egypt. Middle East Research Center, Ain Shams University, Earth Science Series, 18, 1–26. A BDEL A AL , A. & L ELEK , J. J. 1994. Structural development of the Northern Sinai, Egypt and its implications on the hydrocarbon prospectivity of the Mesozoic. GeoArabia, 1, 15– 30. A BDEL A AL , A., E L B ARKOOKY , E., G ERRITS , M., M EYER , H., S CHWANDER , M. & Z AKI , H. 2001. Tectonic evolution of the Eastern Mediterranean Basin and its significance for hydrocarbon prospectivity in the ultradeepwater of the Nile Delta. Leading Edge, 19, 1086–1102. A L -F AR , D. A. 1966. Geology and coal deposits of Gebel El Maghara (north Sinai). Geological Survey of Egypt, 37, 59. A L -S AAD , D., S AWAF , T., G EBRAN , A., B ARAZANGI , M., B EST , J. & C HIAMOV , T. 1992. Crustal structure of central Syria: The intracontinental Palmyride mountain belt. Tectonophysics, 207, 345 –358. A RGYRIADIS , I., D E G ARCIANSKY , P. C., M ARCOUX , J. & R ICOU , L. E. 1980. The opening of the Mesozoic Tethys between Eurasia and Arabia-Africa. Me´moires du Bureau de Recherches Ge´ologiques et Minie`res, 115, 199–214. A YYAD , M. H. 1997. Inverted basins in northeast Egypt; Geology and hydrocarbon prospectivities. PhD thesis, Cairo University, 299. A YYAD , M. H., D ARWISH , M. & S EHIM , A. 1998. Introducing a new structural model for north Sinai with its significance to petroleum exploration. Egyptian General Petroleum Corporation 14th Petroleum Conference, Egypt, 1, 101–117. B ARTOV , Y., L EWY , Z., S TEINITZ , G. & Z AK , I. 1980. Mesozoic and Tertiary stratigraphy, paleogeography and structural history of the Gebel Areif en Naqa area, eastern Sinai. Israel Journal of Earth Sciences, 29, 114 –139. B EIN , A. & G VIRTZMAN , G. 1977. A Mesozoic fossil edge of the Arabian plate along the Levant coastline and its bearing on the evolution of the Eastern Mediterranean. In: B IJU -D UVAL , D. & M ONTADERT , L. (eds) Structural History of the Mediterranean Basins. Editions Technip, Paris, 95–110. B ERTONI , C. & C ARTWRIGHT , J. A. 2006. Controls on the basinwide architecture of late Miocene (Messinian) evaporites on the Levant margin (Eastern Mediterranean). Sedimentary Geology, 188– 189, 93–114 B IJU -D UVAL , B., L ETOUZEY , J. & M ONTADERT , L. 1979. Variety of margins and deep basins in the Mediterranean. American Association of Petroleum Geologists Memoirs, 29, 293 –317.
82
M. YOUSEF ET AL.
B OSWORTH , W., G UIRAUD , R. & K ESSLER , L. G. 1999. Late Cretaceous (c. 84 Ma) compressive deformation of the stable platform of northeast Africa (Egypt), Far-field stress effects of the Santonian event and origin of the Syrian arc deformation belt. Geology, 27, 633–636. B UCHBINDER , B. & Z ILBERMAN , E. 1997. Sequence stratigraphy of Miocene–Pliocene carbonate–siliciclastic shelf deposits in the Eastern Mediterranean Margin (Israel): effects of eustasy and tectonics. Sedimentary Geology, 112, 7 –32. C HAIMOV , T. A., B ARAZANGI , M., A L -S AAD , D., S AWAF , T. & G EBRAN , A. 1990. Crustal shortening in the Palmyride fold belt, Syria, and implications for movement along the dead Sea fault system. Tectonics, 9, 1369–1386. C HAIMOV , T. A., B ARAZANGI , M., A L -S AAD , D., S AWAF , T. & G EBRAN , A. 1992. Mesozoic and Cenozoic deformation inferred from seismic stratigraphy in the southwestern intracontinental Palmyride foldthrust belt, Syria. Geological Society of America Bulletin, 104, 704– 715. C HAIMOV , T. A., B ARAZANGI , M., A L -S AAD , D., S AWAF , T. & K HADDOUR , M. 1993. Seismic fabric and 3-D structure of the southwestern intracontinental Palmyride fold belt, Syria. American Association of Petroleum Geologists Bulletin, 77, 2032–2047. C ITA , M. B. & R YAN , W. B. F. 1978. Messinian erosional surfaces in the Mediterranean. Marine Geology, 27(3– 4), 193–366 C OHEN , Z., K APTSAN , V. & F LEXER , A. 1990. The tectonic mosaic of the southern Levant: implications for hydrocarbon prospects. Journal of Petroleum Geology, 13, 437– 462. C OOPER , M. A. & W ILLIAMS , G. D. 1989. Inversion Tectonics. Geological Society, London, Special Publications, 44, 376. C OWARD , M. P. & R IES , A. C. 2003. Tectonic Development of North African Basins. Geological Society, London, Special Publications, 207, 61– 83. D ERCOURT , J. ET AL . 1986. Geological evolution of the Tethys belt from the Atlantic to the Pamirs since the Lias. Tectonophysics, 123, 241 –315. D EWEY , J. F., P ITMAN III, W. C., R YAN , W. B. F. & B ONIN , J. 1973. Plate tectonics and the evolution of the Alpine system. Geological Society of America Bulletin, 84, 3137– 3180. D OLSON , J. C., S HANN , M. V., M ATBOULY , S., H ARWOOD , C., R ASHED , R. & H AMMOUDA , H. 2001. The petroleum potential of Egypt. In: D OWNEY , M. W., T HREET , J. C. & M ORGAN , W. A. (eds) Petroleum Provinces of the Twenty-first Century. American Association of Petorleum Geologists Memoirs, 74, 453 –482. D RUCKMAN , Y., B UCHBINDER , B., M ARTINOTTI , G. M., S IMAN T OV , R. & A HARON , P. 1995. The buried Afiq Canyon (Eastern Mediterranean, Israel): a case study of a Tertiary submarine canyon exposed in Late Messinian times. Marine Geology, 123, 167–185. E L S HAZLY , E. M. 1977. Geology of the Egyptian region. In: N ARIN , A. E. M., K ANES , W. H. & S TEHLI , F. G. (eds) The Ocean Basins and Margins, the Eastern Mediterranean. Plenum Press, New York, 4A, 379– 344.
E YAL , Y. 1996. Stress field fluctuations along the Dead Sea Rift since the middle Miocene. Tectonics, 15, 157–170. G ARFUNKEL , Z. 1998. Constraints on the origin and history of the Eastern Mediterranean basin. Tectonophysics, 298, 5– 35. G ARFUNKEL , Z. 2004. Origin of the Eastern Mediterranean basin: a reevaluation. Tectonophysics, 391(1–4), 11–34. G ARFUNKEL , Z. & A LMAGOR , G. 1987. Active salt dome development in the Levant Basin, southeast Mediterranean. In: L ERCHE , I. & O’B RIEN , J. (eds) Dynamical Geology of Salt and Related Structures. Academic Press, London, 263– 300. G ARFUNKEL , Z. & D ERIN , B. 1984. Permian-early Mesozoic tectonism and continental margin formation in Israel and its implications for the history of the Eastern Mediterranean. In: D IXON , J. E. & R OBERTSON , A. H. F. (eds) The Geologic Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 187–202. G INZBURG , A., C OHEN , S., H AY -R OE , H. & R OSENZWEIG , A. 1975. Geology of the Mediterranean shelf of Israel. American Association of Petroleum Geologists Bulletin, 59, 2142– 2160. G UIRAUD , R. 1998. Mesozoic rifting and basin inversion along the northern African-Arabian Tethyan margin: An overview. In: M AC G REGOR , D. S., M OODY , R. T. J. & C LARK -L OWES , D. D. (eds) Petroleum Geology of North Africa. Geological Society, London, Special Publications, 133, 215– 227. G UIRAUD , R. & B OSWORTH , W. 1997. Senonian basin inversion and rejuvenation of rifting in Africa and Arabia: synthesis and implications to plate-scale tectonics. Tectonophysics, 282, 39– 82. G UIRAUD , R. & B OSWORTH , W. 1999. Phanerozoic geodynamic evolution of northeastern Africa and the northwestern Arabian platform. Tectonophysics, 315, 73–108. G UIRAUD , R., B OSWORTH , W., T HIERRY , J. & D ELPLANQUE , A. 2005. Phanerozoic geological evolution of Northern and Central Africa: An overview. Journal of African Earth Sciences, 43, 83–143. G VIRTZMAN , G. & B UCHBINDER , B. 1978. The late Tertiary of the coastal plain and continental shelf of Israel and its bearing on the history of the Eastern Mediterranean. In: Reports of Deep Sea Drilling Project, 42, U.S. Government Printing Office, Washington, DC, 1195–1222. H ALBOUTY , M. T. & E L -B AZ , F. 1992. The Mediterranean region: Its geologic history and oil and gas potentials, In: S ADEK , A. (ed.) Geology of the Arab World. Cairo University, Cairo, Egypt, 15–60. H SU¨ , K. J., R YAN , W. B. F. & C ITA , M. B. 1973. Late Miocene desiccation of the Mediterranean. Nature, 242, 240– 244. H SU¨ , K. J., M ONTADERT , L. ET AL . 1978. In: Reports of Deep Sea Drilling Project, 42. U.S. Government Printing Office, Washington, DC. 1249. H UBBARD , R. J. 1988. Age and significance of sequence boundaries on Jurassic and early Cretaceous rifted continental margins. American Association of Petroleum Geologists Bulletin, 72, 49– 72.
STRUCTURE AND TECTONICS OF OFFSHORE NORTH SINAI, EGYPT J AMISON , W. R. 1987. Geometric analysis of fold development in overthrust terranes. Journal of Structural Geology, 9, 207–219. J ENKINS , D. A. 1990. North and central Sinai. In: S AID , R. (ed.) The Geology of Egypt, Balkema Publications, Netherlands, 734. K EELEY , M. L. 1994. Phanerozoic evolution of the basins of northern Egypt and adjacent areas. Geologische Rundschau, 83, 728– 742. K EELEY , M. L. & W ALLIS , R. J. 1991. The Jurassic System in northern Egypt. Journal of Petroleum Geology, 14, 49–64. K HALIL , M. H. & M OUSTAFA , A. R. 1994. Tectonic framework of northern Egypt and its bearing on hydrocarbon exploration. Proceedings of 12th Egyptian General Petroleum Corporation Petroleum Exploration and Production Conference, Cario, 1, 67–86. K RENKEL , E. 1925. Geologie Afrikas. Verlag Von Gebruder, Berlin, 461. L E R OY , P., G UILLOCHEAU , F., P IQUE , A. & M ORABET , A. M. 1998. Subsidence of the Atlantic Moroccan margin during the Mesozoic. Canadian Journal of Earth Sciences, 35, 476 –493. L ETOUZEY , J., C OLLETTA , B., B ENARD , F. & S ASSI , W. 1990. Fault reactivation and structural inversion physical models analyzed with x-ray scanner and seismic examples (abs). American Association of Petroleum Geologists Bulletin, 74, 703. L OWELL , J. D. 1985. Structural Styles in Petroleum Exploration. Oil & Gas Consultants International Publications, Tulsa, 460. L OWELL , J. D. 1995. Mechanics of basin inversion from world-wide examples. In: B UCHANAN , J. G. & B UCHANAN , P. G. (eds) Basin Inversion. Geological Society, London, Special Publications, 88, 39– 57. M ART , Y. 1987. Superpositional tectonic patterns along the continental margin of the southeastern Mediterranean: A review. Tectonophysics, 140, 213–232. M AY , P. R. 1991. The eastern Mediterranean Mesozoic basin: evolution and oil habitat. American Association of Petroleum Geologists Bulletin, 75, 1215–1232. M C B RIDE , J. H., B ARAZANGI , M., B EST , J., A L -S AAD , D., S AWAF , T., A L -O TRI , M. & G EBRAN , A. 1990. Seismic reflection structure of intracratonic Palmyride fold-thrust belt and surrounding Arabian platform, Syria. American Association of Petroleum Geologists Bulletin, 74, 238–259. M C C LAY , K. R. 1995. The geometries and kinematics of inverted fault systems: a review of analogue model studies. In: B UCHANAN , J. G. & B UCHANAN , P. G. (eds) Basin Inversion. Geological Society, London, Special Publications, 88, 97– 118. M ENEISY , M. Y. 1986. Mesozoic igneous activity in Egypt. Qatar University Science Bulletin, 6. M ITRA , S. 1990. Fault-propagation folds: Geometry, Kinematic evolution, and hydrocarbon traps. American Association of Petroleum Geologists Bulletin, 74, 921–945. M OON , F. W. & S ADEK , H. 1921. Topography and geology of North Sinai. Petroleum Research Bulletin, Cairo, 10, 154. M OUSTAFA , A. R. 2010. Structural setting and tectonic evolution of North Sinai folds, Egypt. In: H OMBERG , C. & B ACHMANN , M. (eds) Evolution of the Levant
83
Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 37–63. M OUSTAFA , A. R. & K HALIL , M. H. 1988. Late Cretaceous-Early Tertiary dextral transpression in north Sinai: reactivation of the Tethyan continental margin (abs). American Association of Petroleum Geologists, Mediterranean Basins Conference, Nice. M OUSTAFA , A. R. & K HALIL , M. H. 1989. North Sinai structures and tectonic evolution. Middle East Research Center, Ain Shams University, Earth Science Series, 3, 215–231. M OUSTAFA , A. R. & K HALIL , M. H. 1994. Rejuvenation of the eastern Mediterranean passive continental margin in northern and central Sinai: new data from the Themed fault. Geological Magazine, 131, 435– 448. M OUSTAFA , A. R. & K HALIL , M. H. 1995. Superposed deformation in the northern Suez rift, Egypt: relevance to hydrocarbons exploration. Journal of Petroleum Geology, 18, 245–266. M OUSTAFA , A. R. & Y OUSIF , M. S. M. 1990. Two-stages wrench deformation at Gebel El Minsherah, North Sinai. Middle East Research Center, Ain Shams University, Earth Science Series, 4, 112–122. M OUSTAFA , A. R., E L B ADRAWY , R. & G IBALI , H. 1998. Pervasive E-ENE oriented faults in northern Egypt and their effect on the development and inversion of prolific sedimentary basins. Egyptian General Petroleum Corporation 14th Petroleum Conference, Egypt, 1, 51–67. M OUSTAFA , A. R., K HAWASIK , S. M. & K HALIL , S. M. 1991. Late Cretaceous-Tertiary rotational deformation in North Sinai (Gebel Yelleq Area). Neues Jahrbuch fu¨r Geologie und Palaontologie, 11, 643– 653. N EEV , D. & B EN -A VRAHAM , Z. 1977. The Levantine countries: the Israeli coastal region. In: N AIRN , E. M. & K ANES , W. H. (eds) The Ocean Basins and Margins. The Eastern Mediterranean, 4A. Plenum Press, New York, United States, 355–377. N IKISHIN , A. M., Z IEGLER , P. A. ET AL . 2001. Mesozoic and Cainozoic evolution of the Scythian PlatformBlack Sea-Caucasus domain. In: Z IEGLER , P. A., C AVAZZA , W., R OBERTSON , A. H. F. & C RASQUIN S OLEAU , S. (eds) Peri-Tethys Memoir 6: Peri-Tethyan Rift/Wrench Basins and Passive Margins. Me´moires du Muse´um National d’Histoire Naturelle de Paris, 186, 295–346. O LIVET , J. L., B ONNIN , J., B EUZART , P. & A UZENDE , J. M. 1984. Cine´matique de l’Atlantique nord et central. Rapports scientifiques et techniques du Centre national pour l’exploitation des oce´ans, 54, 108. O RWIG , E. R. 1982. Tectonic framework of northern Egypt and eastern Mediterranean region. Egyptian General Petroleum Corporation 6th Petroleum Conference, Cairo. P ATON , D. A. 2006. Influence of crustal heterogeneity on normal fault dimensions and evolution: southern South Africa extensional system. Journal of Structural Geology, 28, 868–886. R OBERTSON , A. H. F. 1998. Mesozoic-Tertiary tectonic evolution of the easternmost Mediterranean area: integration of marine and land evidence. In: R OBERTSON , A. H. F., E MEIS , K. C., R ICHTER , C. &
84
M. YOUSEF ET AL.
C AMERLENGHI , A. (eds) Proceedings of the Ocean Drilling Program Science Research, 160, 723– 782. R OBERTSON , A. H. F. & D IXON , J. E. 1984. Introduction. In: D IXON , J. E. & R OBERTSON , A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 1– 74. S ADEK , H. 1928. The principal structural features of the Peninsula of Sinai. 14th International Geological Congress Proceedings, Madrid, 3, 895– 900. S ALEL , J. F. & S EGURET , M. 1994. Late Cretaceous to Paleogene thin-skinned tectonics of the Palmyrides belt (Syria). Tectonophysics, 234, 265 –290. S EARLE , M. P. 1994. Structure of the intraplate eastern Palmyride fold belt, Syria. Geological Society of America Bulletin, 106, 1332–1350. S ERENCSITS , C., P AUL , H., R OLAND , K. A., E L R AMLY , M. F. & H USSEIN , A. A. 1979. Alkaline ring complexes in Egypt: Their ages and relationship to tectonic development of the Red Sea. Annals of the Geological Survey of Egypt, 9, 102 –116. S HATA , A. 1959. Structural development of the Sinai Peninsula (Egypt). 20th International Geological Congress, Proceedings, Mexico, 225– 249. S MITH , A. G. 1971. Alpine deformation and the oceanic area of the Tethys, Mediterranean and Atlantic. Geological Society of America Bulletin, 82, 2039–2070. S TAMPFLI , G. M., M OSAR , J., F AVRE , P., P ILLEVUIT , A. & V ANNAY , J. C. 2001. Permo-Mesozoic evolution of the western Tethys realm: the Neo-Tethys East Mediterranean Basin connection. In: Z IEGLER , P. A., C AVAZZA , W., R OBERTSON , A. H. F. & C RASQUIN S OLEAU , S. (eds) Peri-Tethyan Rift/Wrench Basins
and Passive Margins. Me´moires du Muse´um national d’Histoire naturelle de Paris, 186, 51–108. S ULTAN , N. & H ALIM , M. A. 1988. Tectonic framework of northern Western Desert, Egypt and its effect on hydrocarbon accumulations. Egyptian General Petroleum Corporation 9th Petroleum Conference, 2, 1– 22. T IBOR , G., B EN -A VRAHAM , Z., S TECKLER , M. & F LIGELMAN , H. 1992. Late Tertiary subsidence history of the Southern Levant Margin, Eastern Mediterranean Sea, and its implications to the understanding of the Messinian event. Journal of Geophysical Research, 97, 17593– 17614. V IDAL , N., A LVAREZ -M ARRO´ N , J. & K LAESCHEN , D. 2000. Internal configuration of the Levantine Basin from seismic reflection data (eastern Mediterranean). Earth and Planetary Science Letters, 180, 77– 89. W ALSH , J. J., N ICOL , A. & C HILDS , C. 2002. An alternative model for the growth of faults. Journal of Structural Geology, 24, 1669– 1675. W EISSBROD , T. 1969. The Paleozoic of Israel and adjacent countries: Part II, The Paleozoic outcrops in southwestern Sinai and their correlation with those of southern Israel. Bulletin of the Geological Survey of Israel, 48, 1– 32. W ILLIAMS , G. D., P OWELL , C. M. & C OOPER , M. A. 1989. Geometry and kinematics of inversion tectonics. In: C OOPER , M. A. & W ILLIAMS , G. D. (eds) Inversion Tectonics. Geological Society, London, Special Publications, 44, 376. Y OUSEF , M., M OUSTAFA , A. R. & S HANN , M. 2006. The Petroleum Play Systems in the Mango – Tineh area: Offshore North Sinai, Egypt (Abs). The 8th International Conference on the Geology of the Arab World, Cairo University, Cairo, Egypt.
Deep structures and seismic stratigraphy of the Egyptian continental margin from multichannel seismic data LAURENT CAME´RA*, ALESSANDRA RIBODETTI & JEAN MASCLE Geoazur UMR 6526, Observatoire Oce´anologique de Villefranche-sur-Mer, BP48 06235 Villefranche-sur-Mer cedex, France *Corresponding author (e-mail:
[email protected]) Abstract: Regional multichannel seismic reflection (MCS) profiles across the Egyptian continental slope, offshore the Nile delta, were recorded during the MEDISIS survey (conducted in 2002 on board the R/V Nadir). The results of this survey allow an interpretation of the overall structure and evolution of this passive continental margin. The MCS data were processed using an amplitude preserving pre-stack depth migration technique, which has the advantage of providing a quantitative, and geometrically correct, image of seismic horizons. Well-defined reflecting events allow the identification of three main seismic units. The upper unit (a 7 km thick) is interpreted as the post-rift sedimentary cover of the margin; it includes an undisturbed Middle Cretaceous to Upper Miocene sedimentary pile, covered by thick Messinian (latest Miocene) salt-rich layers and by Pliocene to Quaternary sediments, locally intensively deformed by gravity tectonics. The underlying intermediate acoustic unit (6 km thick on average) is interpreted as the Mesozoic syn-rift sedimentary cover of the margin; the end of the last rifting event is marked by a strong angular unconformity, tentatively of Aptian age. The lower unit may correspond to the thinned continental crust of Africa (12 km thick on average in the study area) and its pre-rift cover. Its base is identified by strong, discontinuous reflector packages about 23–25 km below sea floor, interpreted as indicative of the Moho.
The southern border of the Levantine basin is classically interpreted as a Mesozoic passive continental margin now, progressively entering a diachronous collision with the Anatolian –Aegean-European plates as a consequence of long-term convergence and subduction between Africa and Eurasia (Fig. 1). The Egyptian continental margin, presently covered by a thick sedimentary wedge including significant Messinian salt deposits, represents a 500 km long segment of this passive margin. Moreover, the Egyptian margin is an area of fairly active investigations conducted by numerous oil and gas companies as a consequence of its potential, and partly proven, giant gas reserves (Dolson et al. 2000, 2005; Abdel Aal et al. 2001; Samuel et al. 2003). For a decade, systematic surveys of the Egyptian margin have been conducted by scientists from Geosciences Azur and other academic laboratories in order to obtain multi-scale geophysical images of this margin segment and to better understand its tectonic and sedimentary structures. These investigations have focused on the deep-sea terrigenous cone built offshore of the Nile delta and have resulted in geophysical and geological datasets that have provided important results on the various active geological processes that shape this young sedimentary pile (Loncke 2003; Loncke et al. 2006).
However, little is known on the deep structures of this margin segment and only a few simplified geological cross sections are available, based on unpublished MCS data (Dolson et al. 2000, 2005; Abdel Aal et al. 2001) and gravity–magnetic modelling (Segev et al. 2006; Longacre et al. 2007). Most of the proposed models imply a 12–15 km thick sedimentary sequence beneath the Nile deep-sea fan resting above a c. 10 km thick crust, the latter being regarded either as thinned continental crust (Segev et al. 2006), and/or shortened continental crust (Abdel Aal et al. 2001), or as old oceanic crust bounded by a transform margin segment (Longacre et al. 2007). Here we report results from the MEDISIS multichannel seismic (MCS) survey, conducted in 2002 by Geosciences Azur. The processed MCS data reveal a thick sedimentary wedge (12 –15 km) beneath the Nile deep sea fan, covering, depending on the area, a 9 –12 km thick continental crust.
Tectonic setting The Egyptian margin lies in a complex geodynamical setting, where three major plates (African, Arabian, Eurasian) and several minor plates (Anatolian, Aegean and probably Sinai) interact, and where most of the so-called neo-Tethyan oceanic crust
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 85– 97. DOI: 10.1144/SP341.5 0305-8719/10/$15.00 # The Geological Society of London 2010.
L. CAME´RA ET AL.
86
Fig. 1. Shaded-relief bathymetry of the Egyptian continental margin (from Sardou & Mascle 2003) and locations of the MEDISIS MCS profiles (lines). The Egyptian margin can be divided into four different morpho-structural domains that are the Western, Central, Eastern and Levantine provinces (Loncke 2003). In the lower right corner: tectonic setting of eastern Mediterranean basin (modified from Loncke 2003).
and its margins have been consumed by subduction and partly incorporated in the Alpine and Caucasus mountain chains (Smith 1971; Sengo¨r 1979; Smith & Woodcock 1982; Sengo¨r et al. 1984; Dercourt et al. 1986; Stampfli & Borel 2002). Currently, most of the deformations occurring in the Eastern Mediterranean relate to two main plate tectonic mechanisms (Fig. 1): (1)
(2)
the northward motion of the Africa, and more recently Arabia, with respect to the Eurasian plate, probably ongoing since the Late Cretaceous (Olivet et al. 1982; Argus et al. 1989; Rosenbaum et al. 2002); the recent westward extrusion of the Aegean– Anatolian micro-plate (30 –40 Ma), which is bounded by the East and North Anatolian fault zones (McKenzie 1970, 1972; Taymaz et al. 1990; Jackson et al. 1992; Jackson 1994; Le Pichon et al. 1995; McClusky et al. 2000).
These mechanisms are expressed in: (1)
major strike –slip faulting (the East and North Anatolian and Dead Sea faults);
(2) (3) (4)
ongoing and ending subductions (Hellenic and Cyprus subduction); active extension (Aegean Sea, Red Sea); and various collision zones (Fig. 1).
In this geodynamic environment, the Egyptian continental margin cannot be considered as a simple passive margin segment. Came´ra (2006) has recently proposed an evolution of the Egyptian margin area including six major phases, following Dolson et al. (2000). In Palaeozoic to Early Triassic times, intraplate deformations, mainly documented by later reactivated structures, led to important continental erosion and sedimentation, generally associated with magmatical activity (Guiraud & Bosworth 1999). From Late Triassic to Jurassic times, the still poorly constrained break-up of Pangea resulted in several rifting events, ultimately leading to the opening of the neo-Tethyan ocean (Smith 1971; Sengo¨r 1979; Sengo¨r et al. 1984; Dercourt et al. 1986; Ben-Avraham & Ginzburg 1990; Stampfli & Borel 2002) and to the development of a passive continental margin north of Egypt (Dixon & Robertson 1984; Moustafa et al. 1998). Between Cenomanian and
DEEP STRUCTURES OF THE EGYPTIAN MARGIN
Miocene times, NW–SE-oriented compression episodes, the so-called Syrian Arc tectonism (Krenkel 1925; Moustafa et al. 1998; Dolson et al. 2000), took place in the easternmost Levantine domain and resulted in the structural inversion of former extensional neo-Tethyan rifted features (Lu¨ning et al. 1998a, b; Kuss et al. 2000; Hussein & Abd-Allah 2001; Gardosh & Druckmann 2005). In Oligocene times, renewed continental extension (early motion of Arabia away from Africa) resulted in the creation of the Gulf of Suez and, in the Middle Miocene, in the opening of the Red Sea. Geological and seismological data suggest that the Sinai block, located between the Suez and Akaba gulfs, behaved like a micro-plate trapped between the African and Arabian plates (Le Pichon & Gaulier 1988; Mascle et al. 2000). At the end of the Miocene (from 5.9 to 5.3 ma), the entire Mediterranean basin experienced the Messinian salinity crisis (Ryan et al. 1973; Rouchy 1986; Gautier et al. 1994; Clauzon et al. 1996), which resulted in thick deposits of salt and anhydrite throughout the deep Levantine basin, and erosional unconformities and terrigenous deposits on the continental shelf and upper slope. Finally, during Pliocene and to recent times, the Egyptian margin underwent rapid subsidence and the accumulation of a thick Plio-Quaternary sedimentary pile (Tibor & Ben-Avraham 2005) that constitutes the present day Nile deep-sea fan (NDSF). The NDSF is deposited above the mobile Messinian salt layers and is affected by typical thin-skinned processes, such as gravity gliding and gravity spreading, in relation with the underlying Messinian salt layers (Loncke 2003; Loncke et al. 2006).
The MEDISIS MCS survey During the 2002 MEDISIS survey, about 1000 km of MCS reflection profiles were recorded along seven regional lines crossing the Egyptian continental margin (Fig. 1). The profiles were acquired using an array of 10 GI-guns, with a total volume of 53 l every 30 s. The reflected signals were recorded via a 4700 m long, 360 channel digital streamer, with a 12.5 m group spacing, across record lengths of 15 s at a 4 ms sampling rate. The regional lines cut successively across the Levantine, Eastern, Central and Western province of the Nile deep-sea fan (Fig. 1) as distinguished by Loncke (2003). One of the challenges was to image the deep structures beneath the up to 3 km thick salt layers. The length of the streamer does not allow us, in principle, to obtain well constrained velocities greater than 4.5 km/s. A procedure of iterative adjustments of the interval velocity was thus performed during the processing of lines, described in detail below, to obtain a velocity model that was calibrated
87
with the nearby expanding spread profile experiments of De Voogd et al. (1992). Together these methods provided good estimates of the velocities and depth migration on all lines.
Data processing The data were processed at Geosciences Azur using two different techniques: (1) a standard processing using the Geovector package to produce post-stack time-migrated profiles for preliminary interpretations; (2) a preserved amplitude pre-stack depth migration method (PSDM, see Fig. 2) to generate depth sections. In addition, one line (MD06, see Fig. 3) was processed at Geomar (Kiel) using the Sirius package to obtain a reference PSDM profile for comparison (Fig. 3). Geovector software was used to prepare the data for PSDM (via pre-processing to preserve the amplitude of the acoustic signal), and reduce the signal/ noise ratio. The pre-processing included: sort data to 6.25 m common depth point (CDP), first pass velocity analysis, amplitude attenuation of noisy traces, band pass filter (2, 5, 140, 150 Hz), multiple attenuation in the frequency-wavenumber (FK) domain, normal move out (NMO) velocity analysis, loose external mute, and inverse NMO correction, and finally transform from CDPs to shot gather. Note that in order to preserve acoustic signal amplitudes, no spherical divergence and amplitude corrections were applied (Fig. 2). To obtain the final depth-migrated images we used a Ray þ Born migration technique also known as PSDM. The software (Thierry et al. 1999) is the ‘acoustic’ version of an original ‘elastic’ method proposed by Jin et al. (1992), who introduced an attractive asymptotic method for inverting seismic reflection data. Since the technique of Jin et al. (1992) was introduced, several applications to two-dimensional (2D) and three-dimensional (3D) datasets have been developed for acoustic modelling (Lambare´ et al. 1992; Thierry et al. 1999; Ribodetti et al. 2000; Operto et al. 2003). This approach is very sensitive to the initial velocity model estimate. If the model is incorrect, the reflectors are mislocated and the amplitudes of the velocity distributions are biased. To obtain a reliable image (correct geometry and correct velocity perturbations of seismic reflectors) we used a simple and efficient method, developed by Alyahya (1989) and Agudelo (2005), to perform a quantitative estimate and correction of the velocity macro-model, through a standard ‘migration –velocity –analysis’ approach. Iso-X panels (or common image gathers (CIGs)) are stored during migration and semblance panels estimated to obtain a local correction function for the
88
L. CAME´RA ET AL.
Fig. 2. Processing steps in the Ray þ Born pre-stack depth migration technique. The data are first pre-processed but amplitudes are preserved. A velocity macro-model estimated from velocity analysis, changed in depth using Dix equation, is then interpolated and smoothed. In this model, an asymptotic ray tracing (Lambare´ et al. 1996) is estimated for the computation of all the ray related parameters. Common angle migration, including velocity analysis and velocity macro-model correction, is performed iteratively until Iso-X panels appear flat and semblances are around 1 for all the main reflectors. Finally, CIGs are stacked to obtain an accurate migrated image.
velocity-macro model. Confidence in the migrated image is achieved when the Iso-X panels are flat and when the corresponding semblance panels are around 1 (Alyahya 1989). The velocity macromodel is then iteratively corrected during migration until the semblance panels remain around 1 for all the main reflectors. When such conditions are satisfied, all the CIGs are stacked to obtain the final migrated image (Fig. 2). We performed three iterative corrections of the velocity macro-model; using this technique we observe between 0 to 5 km (c. total length of the seismic streamer) very flat CIGs, suggesting a minimum error for the velocity
model; as a consequence depth location of seismic events have uncertainties of less than 30m. Between 5 and 10 km, a slight upward bend of reflectors shown on the CIGs indicates an error in depth location of about 100 m. Between 10 to 15 km, the CIG panels reveal clear downward bends of reflectors indicating over-estimated velocities and an error in seismic event location on the order of 500 m. Below 15 km, the CIG technique is no longer useful but the variability of deeper seismic event locations during the three iterations allows an estimate of the error in depth values of 1000 and 1300 m.
DEEP STRUCTURES OF THE EGYPTIAN MARGIN
km
SSE
(a)
10
89
30
50
70
90
110
NNW
130
150
170
190
210
230
s twt
0 5 10 15
Depth km
(b) 0 5 10 15 20
(c)
0
Depth km
5 10 15 20 25 30
Fig. 3. Examples of three different processing techniques applied to profile MD06. (a) post-stack time migration using Geovector. (b) PSDM performed at Geomar using Sirius package. (c) PSDM using the Ray þ Born technique.
Results and interpretation The different processing techniques (post-stack time migration versus PSDM using the Sirius package or Ray þ Born technique) result in comparable seismic successions on most of the MEDISIS profiles. As an example, the three different techniques applied to profile MD06 (see location on Fig. 1) clearly yield similar images (Fig. 3). † On Figure 3a, which illustrates the result of a post-stack time migration using Geovector, we observe a series of well-layered seismic sequences (7 –8 s-twt thick) resting on an acoustic basement that includes a few discontinuous internal reflectors (up to 11 –12 s-twt ). † On Figure 3b, a PSDM of the same line made at Geomar using the Sirius package (only processed down to 10 s), a similar succession is observed: well layered and continuous reflecting units can be seen down to 15 km, resting on a pattern of strong and discontinuous horizons detected down to 18–20 km (assuming adjusted velocities between 5.5 to 6 km/s). † Finally on Figure 3c, PSDM of the same MD06 profile processed using the Ray þ Born technique shows a similar succession, above
discontinuous and low frequency reflectors at around 25 km depth. Using the seismic characteristics of line MD06, we have distinguished five main seismic sequences on the Egyptian margin (Figs 4 & 5). From top to bottom, these are referred as sequences E to A. The uppermost, sequence E, is a well-layered unit whose thickness varies between 1 to 3 km (Figs 4– 6). It corresponds to the Pliocene and Quaternary sedimentary cover of the margin. Below this recent cover, an almost reflection-free sequence D shows numerous diapir-like deformations and is inferred to correspond to the Messinian evaporite deposits (Figs 4–6). Its thickness varies between a few hundred meters (upslope) and 3 km (at the base of the slope). It is bounded at its base by a series of strong and continuous reflectors that underline an important velocity inversion (average interval velocities are in the order of 4000 m/s within the salt and 2500 m/s just below). Below the Messinian horizons, sequence C comprises welllayered and undisturbed reflectors (Figs 4–6). It is separated by a strong angular unconformity from underlying discontinuous and strongly fractured horizons of sequence B. Finally, below sequence B, discontinuous, low frequency, and locally
L. CAME´RA ET AL.
90
0 0
km
10
S
N
Sequence E
5 Sequence D Sequence C
Break-up unco
nformity
10
km
Sequence B
15
20 Sequence A
25
30 Fig. 4. Detail from PSDM profile MD08 showing five main acoustic sequences (see location on Fig. 1).
faulted reflectors define sequence A (Fig. 4), interpreted as evidence of a weakly layered continental crust. At its base, sequence A is bounded by a series of discontinuous and very low frequency reflector packages (see Figs 4 & 7) situated at an average 23 –25 km depth, which we interpret as indicative of the Moho. From these data, the seismic stratigraphy of the Egyptian margin can be interpreted in terms of three main tectono-stratigraphic units (Fig. 8). The two uppermost units 1 and 2 are well-layered and are interpreted as the sedimentary cover of the margin. Unit 1 includes sequences E to C, which are all characterized by continuous horizons (locally disrupted by gravity structures). In contrast Unit 2, which includes only sequence B, shows medium to high-amplitude reflectors and is highly faulted. A strong unconformity separates Unit 1 and 2. Below, Unit 3, which corresponds to sequence A, shows low-frequency discontinuous reflectors bounded at their base by a few, very thick, reflector packages. It thus corresponds to the crustal basement. Detailed description of these units is presented below in dedicated sections.
The sedimentary cover From top to bottom, the sedimentary cover of the margin includes sequences E to B.
As already indicated, sequence E is referred to as the recent Pliocene and Quaternary sediment blanket constituting most of the present Nile deepsea fan. This Pliocene –Quaternary cover, which locally reaches thickness in excess of 4500 m, is intensively cut by a dense set of extensional faults resulting from salt tectonics, creating small polygonal grabens on the upper slope, and elongated crestal grabens on the middle slope (Diegel et al. 1995; Loncke 2003; Loncke et al. 2006). At the base of the slope, this Pliocene–Quaternary pile is strongly deformed by folds, reverse faulting and thrusts that also result from salt movements downslope (Cramez & Jackson 2000; Loncke et al. 2006) (Figs 6 & 7). Sequence D corresponds to the Messinian evaporite section. The lower boundary comprises highamplitude reflectors, while its internal seismic character is a transparent seismic facies that includes a few high-amplitude and low-frequency reflectors; chaotic reflectors are also locally observed, particularly near the base of the sequence and in areas of intense deformation. Since its deposition, the Messinian evaporites have been moving downslope and this motion has introduced important salt thickness variations. Such variations are illustrated in Figure 6 along the MD06 profile that extends from the continental slope up to the deep basin. Along the upper slope/shelf area, the Messinian salt layers are almost absent, either because they were not deposited or/and because part of the salt has moved downslope inducing growth fault systems [see (1) on Fig. 6]. On the middle slope, the salt layers are almost in concordance with the upper cover and their thickness reaches up to 1500 m, except in several crestal grabens [see (2) on Fig. 6]. At the base of the slope, the Messinian layers reach thickness of up to 3000 m as a consequence of their stacking, which leads to various and intensively folded features [see (3) on Fig. 6]. Directly below the salt, bounded at its base by strong reflective horizons, the variably thick sequence C is made of several sub-sequences, including a poorly layered upper section and a lower section made of more continuous horizons. On most profiles sequence C is separated from sequence B, either by a strong acoustic contrast or by an angular unconformity along which internal reflectors are truncated below and onlapping above (Fig. 5). No specific evidence of tectonic activity can be recognized within this sequence. Sequence B includes concordant reflectors, of variable continuity, amplitude and frequency content, characterized by sub-parallel configurations and often dislocated and offset by small faults. Its thickness is highly variable, but is estimated to be 6000 m on average (see Figs 4 & 5).
DEEP STRUCTURES OF THE EGYPTIAN MARGIN SSE
1000
91 NNW
10 000 m
3000 5000
metres
7000 9000 11000 13000 15000 17000 19000
SSE
1000 3000
NNW
10 000 m
Sequence E Sequence D
5000
Sequence C metres
7000 9000 11000
formity Uncon
Sequence B
13000 15000 17000
Sequence A
19000
Fig. 5. Pre-stack depth-migrated MD06 profile without and with interpretation (see Fig. 1 for location). Note the strong regional unconformity between sequences B and C.
The identification of an strong unconformity between sequences B and C is an important result of this study. This surface is clearly observed on profile MD06 (Fig. 5) and has been detected on most of the profiles across the margin, even if locally it is less distinct. These observations indicate that this surface has a regional significance. Using results from ocean drilling program (ODP) site 967 (Emeis et al. 1996) on the Eratosthenes seamount, near profile MD06 (Fig. 1), we correlate this unconformity with the top of indurated limestones of Aptian age (119 –115 Ma). According to Whiting (1998), these sediments, which include interbedded breccias, were deposited in a lower bathyal environment (a few hundred meters water depth). We therefore are inclined to interpret the fractured sequence B (see Fig. 8) as a syn-rift sequence, dominantly made of carbonate and deposited in a rather shallow environment. The unconformity, seen at its top, may mark the end of a rifting phase and is interpreted as a break-up unconformity
occurring here in Aptian time. If correct, this hypothesis implies that sequences C to E correspond to the post-rift evolution of the Egyptian margin, initiated in Middle Cretaceous time (Fig. 8).
The acoustic basement Sequence A is interpreted as the basement of the Egyptian continental margin (see Fig. 8). Its seismic facies is characterized by packages of intermittent, sub-horizontal, low-frequency reflectors. Such deep discontinuous reflectors can be identified on most of the profiles. Moreover one of the post-stack time-migrated profiles (MD08, Fig. 7) displays more continuous seismic events around 11 s-twt slightly deepening toward the south. Despite the presence of acoustic noise generated during MCS data acquisition and possibly during processing, we interpret these few scattered strong reflectors, detected at the base of this unit, as potential indications of the Moho. If our
92
10 000 metres
SSE
10 000 metres
NNW
SSE
1000
1000
2000
2000
3000
3000
4000
4000
5000
5000
6000
6000
NNW
Sequence E Sequence D Sequence C
7000 metres
(1)
7000
10 000 metres
10 000 metres
1000
2000
2000
3000
3000
4000
4000
5000
5000
6000
6000
(2)
Sequence D Sequence C
7000
metres
10 000 metres
ity
nform
Unco
Sequence B
(2)
10 000 metres
1000
1000
2000
2000
3000
3000
4000
4000
5000
5000
6000
6000
7000 metres
Sequence E
(3)
7000
metres
Sequence E Sequence D Sequence C
(3)
Fig. 6. Three extracts of profile MD06 (without and with interpretations, at right). Main faults are shown by heavy black lines. The upper sedimentary sequence (sequence E), from Messinian to Present, is strongly deformed by thin-skinned gravity tectonics.
L. CAME´RA ET AL.
1000
7000 metres
(1)
metres
DEEP STRUCTURES OF THE EGYPTIAN MARGIN
S 0
10
20
30
40
50
60
km
93
N 70
80
90
100
110
120
130
0
s twt
5 10 15 Fig. 7. Post-stack time migration of profile MD08 illustrating deep seismic reflections around 11– 12 s-twt.
interpretation is correct, this indicates a Moho depth averaging 23–25 km over most of the Egyptian continental slope and a basement thickness between 10 to 12 km (Fig. 8), assuming velocities between 5.6 to 6.5 km/s as provided by MCS processing. In absence of refraction data to better constrain crustal velocities, we have used for the basement the average velocity model (between 5.6 to 6.5 km/s) deduced from MCS processing. This indicates that the thickness of the basement decreases from south to north; although beneath the upper part of the slope the thickness is around 14 km, it reaches only 10 to 12 km in most of the study area, whereas below Eratosthenes seamount it reaches at least 18 to 20 km (see Fig. 9). We
0 0
km
interpret this basement as the thinned continental crust on which the Egyptian continental margin cover has been later deposited (Fig. 9). Our interpretation is in good agreement with the refraction data and interpretations of Netzeband et al. (2006) for the nearby southern-eastern Levantine basin, where crustal thickness is found to be in the order of 10 km; they are also in agreement with gravity modelling of Segev et al. (2006) who show a thinned continental crust (10 km thick) beneath the eastern Nile cone. Our hypothesis differs, however, from that of Longacre et al. (2007) who used gravity, magnetic, and seismic data modelling to infer a potential oceanic crust below part of the western Nile continental margin.
10
S
N
UNIT 1
post-rift
UNIT 2
syn-rift
UNIT 3
pre-rift
Jurassic to Lower Cretaceous
2000 to 6000 m
UNIT 3
200 to 4500 m
Plio-Quaternary
UNIT 2
Sequence D Sequence C
Thickness in average
UNIT 1
Sequence E
5
Litho-acoustic interpretation
Continental Crust (Upper and Lower)
10 to 12 km except near Eratosthenes sea mount
Moho
23 to 25 km depth
Break-up unco
nformity
0 to 3000 m Salt Messinian Aptian to Upper Miocene 1500 to 4500 m
10
km
Sequence B
15
20 Sequence A
25
30
Fig. 8. Detail of PSDM profile MD08 illustrating the different seismic sequences and their geological interpretation.
94
L. CAME´RA ET AL.
Fig. 9. Geological cross sections of the Egyptian continental margin based on the interpretation of MEDISIS MCS data. The Egyptian continental margin is made of a succession of three main seismic units: (1) a thinned continental crust (and its pre-rift cover) (Unit 1); (2) a syn-rift sedimentary section (Unit 2) and (3) a post-rift sedimentary cover of the margin (Unit 3).
DEEP STRUCTURES OF THE EGYPTIAN MARGIN
Conclusions Processing and analyses of MCS reflection data recorded during the 2002 MEDISIS survey allow the recognition of several seismic sequences along the Egyptian continental margin offshore the Nile delta. The tectonic and sedimentary structures imaged in these seismic profiles lead us to interpret these sequences in terms of three main tectonostratigraphic units, in the overall geodynamical context of the Eastern Mediterranean (Fig. 9): † The upper Unit 1 is on the order of 7 km thick and corresponds to the post-rift sedimentary cover of the Egyptian margin. This post-rift unit comprises three main sedimentary sequences that are, from top to bottom: (a) a 2 to 4 km thick Pliocene–Quaternary cover, intensively deformed by gravity tectonics; (b) a variably thick (few hundred meters to 3 km) latest Miocene (Messinian) blanket in which evaporites dominate along the middle to lower slope, whereas clastic deposits prevail along the upper continental slope – the inter-bedded thick Messinian evaporites has induced a strong decoupling (Loncke et al. 2006) between the upper cover and the remaining sedimentary pile and therefore the superficial tectonic features do not reflect any basement structural trends; (c) a Middle Cretaceous to Upper Miocene sedimentary pile (1 to 4 km thick depending on the area). † The middle Unit 2 averages 6 km in thickness and is interpreted as the syn-rift sedimentary cover of the margin. This unit is characterized by strong and low frequency reflectors that are frequently disrupted by faults, in response to either successive rifting events or to a very slow and long rift evolution. Correlation of the seismic reflectors with ODP data near Eratosthenes seamount suggests that Unit 2 may correspond to relatively shallow water carbonates of Jurassic to Aptian age. If this hypothesis is correct, it would imply the last rifting event to have ended sometime in the Early Cretaceous. This event is marked either by a major angular unconformity or by strong seismic contrasts observed at the boundary between Unit 1 and Unit 2. † The lower Unit 3 is interpreted as a stretched continental crust extending over most of the studied area, with the exception of the northwesternmost region where evidence of oceanic crust has been found from expanding spread profile seismic experiments (De Voogd et al. 1992). This continental basement is bounded by scattered, low-frequency reflector packages, which are good candidates for a Moho lying at about 23– 25 km.
95
The authors gratefully acknowledge the MEDISIS crew and scientific staff aboard the R/V Nadir. We thank F. Sage and L. Schenini for help in seismic data processing. We thank D. Klaeschen for help in MD06 profile processing performed using Seismic and Sirius at Geomar centre (Kiel). This work was partly supported by GDR Marges. Detailed reviews by G. Netzeband, D. J. Shillington and D. Praeg have greatly improved preliminary versions of this paper.
References Agudelo, W. 2005. Imagerie sismique quantitative de la marge convergente d’Equateur- Colombie: Application des me´thodes tomographiques aux donne´es de sismique re´flexion multitrace et re´fraction-re´flexion grand angle des campagnes SISTEUR et SALIERI. The`se de doctorat de l’Universite´ Paris VI, Spe´cialite´ Sciences de la Terre, PhD thesis, 203. Abdel Aal, A., El Barkooky, A., Gerrits, M., Meyer, H. J., Schwander, M. & Zaki, H. 2001. Tectonic evolution of the eastern Mediterranean basin and its significance for the hydrocarbon prospectivity of the Nile Delta deepwater area. GeoArabia, 6, 363– 384. Alyahya, K. M. 1989. Velocity analysis by iterative profile migration. Geophysics, 54, 718–729. Argus, D. F., Gordon, R. G., De Mets, C. & Stein, S. 1989. Closure of the Africa– North America plate motion circuit and tectonics of the Gloria fault. Journal of Geophysical Research, 94, 5585– 5602. Ben-Avraham, Z. & Ginzburg, 1990. Displaced terranes and crustal evolution of the Levant and the eastern Mediterranean. Tectonics, 9, 613– 622. Came´ra, L. 2006. Structure profonde de la marge passive egyptienne au large du Nil: contribution des donne´es de sismique multitrace et de gravime´trie. The`se de doctorat de l’Universite´ Paris VI, Spe´cialite´ Sciences de la Terre, PhD thesis, 157. Clauzon, G., Suc, J.-P., Gautier, F., Berger, A. & Loutre, M.-F. 1996. Alternate interpretation of the Messinian salinity crisis; controversy resolved. Geology, 24, 363–366. Cramez, C. & Jackson, M. P. A. 2000. Superposed deformation straddling the continental-oceanic transition in deep-water Angola. Marine and Petroleum Geology, 17, 1095–1109. Dercourt, J., Zonenshain, L. P. et al. 1986. Geological evolution of the Tethys belt from the Atlantic to the Pamirs since the Lais. Tectonophysics, 123, 241– 315. De Voogd, B., Truffert, C., Chamot-Rooke, N., Huchon, P., Lallemant, S. & Le Pichon, X. 1992. Two-ship seismic soundings in the basins of the Eastern Mediterranean Sea (Pasiphae Cruise). Geophysical Journal International, 109, 536–552. Diegel, F. A., Karlo, J. F., Schuster, D. C., Shoup, R. C. & Tauvers, P. R. 1995. Cenozoic structural evolution and tectono-startigraphic framework of the northern gulf coast continental margin. In: Jackson, M. P. A., Roberts, D. G. & Snelson, S. (eds) Salt Tectonics: A Global Perspective. American Association of Petroleum Geologists Memoir, 65, 109– 151.
96
L. CAME´RA ET AL.
Dixon, J. E. & Robertson, A. H. F. 1984. The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 844. Dolson, J. C., Shann, M. V., Matbouly, S. I., Hammouda, H. & Rashed, M. R. 2000. Egypt in the twenty-first century: petroleum potential in offshore trends. GeoArabia, 6, 211–230. Dolson, J. C., Boucher, P. J., Siok, J. & Heppard, P. D. 2005. Key challenges to realizing full potential in an emerging giant gas province: Nile Delta/ Mediterranean offshore, deep water, Egypt. In: Dore´, A. G. & Vining, B. A. (eds) Petroleum Geology: North–West Europe and Global PerspectivesProceedings of the 6th Petroleum Geology Conference. Petroleum Geology Conferences Ltd. Published by the Geological Society, London, 607– 624. Emeis, K.-C., Robertson, A. H. F. & Richter, C. Shipboard Scientific Party 1996. Proceedings of the Ocean Drilling Program, Initial Reports, 60, 215–287. Gardosh, M. & Druckmann, Y. 2006. Stratigraphy and tectonic evolution of the Levantine Basin, offshore Israel. In: Robertson, A. H. F. & Mountrakis, D. (eds) Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, 201– 227. Gautier, F., Clauzon, G., Suc, J.-P., Cravatte, J. & Violanti, D. 1994. Age et dure´e de la crise de salinite´ messinienne. Comptes Rendus de l’Acade´mie des Sciences de Paris, 318, II, 1103– 1109. Guiraud, R. & Bosworth, W. 1999. Phanerozoic geodynamic evolution of northeastern Africa and the northwestern Arabian platform. Tectonophysics, 315, 73–108. Hussein, I. M. & Abd-Allah, A. M. A. 2001. Tectonic evolution of the northeastern part of the African continental margin, Egypt. Journal of African Earth Sciences, 33, 49–68. Jackson, J. A. 1994. Active tectonics of the Aegean region. Annual Review of Earth and Planetary Sciences, 239 –271. Jackson, J. A., Haines, A. J. & Holt, W. E. 1992. The horizontal velocity field in the deforming Aegean Sea region determined from the moment tensors of earthquakes. Journal of Geophysical Research, 97, 17657– 17684. Jin, S., Madariaga, R., Virieux, J. & Lambare´, G. 1992. Two-dimensional asymptotic iterative elastic inversion. Geophysical Journal International, 108, 1– 14. Krenkel, E. 1925. Geologie Afrikas, Verlag Von Gebruder, Berlin, 1925, 461. Kuss, J., Westerhold, T., Groß, U., Bauer, J. & Lu¨ning, S. 2000. Mapping of Late Cretaceous stratigraphic sequences along a Syrian Arc Uplift – Examples from the Areif el Naqa/Eastern Sinai. Middle East Research Center, Ain Shams University, Earth Science Series, 14, 171 –191. Lambare´, G., Lucio, P. S. & Hanyga, A. 1996. Twodimensional multivalued traveltime and amplitude maps by uniform sampling of ray field. Geophysical Journal International, 125, 584–598. Lambare´, G., Virieux, J., Madariaga, R. & Jin, S. 1992. Iterative asymptotic inversion in the acoustic approximation. Geophysics, 57, 1138– 1154.
Le Pichon, X. & Gaulier, J.-M. 1988. The rotation of Arabia and the Levant fault system. Tectonophysics, 153, 271 –294. Le Pichon, X., Chamot-Roocke, N. & Lallemant, S. 1995. Geodetic determination of the kinematics of central Greece with respect to Europe: implications for Eastern Mediterranean Tectonics. Journal of Geophysical Research, 100, 12 675– 12 690. Loncke, L. 2003. Le delta profond du Nil: structure et e´volution depuis le messinien (Mioce`ne terminal). The`se de doctorat de l’Universite´ Paris VI, Spe´cialite´ Sciences de la Terre, PhD thesis, 184. Loncke, L., Gaullier, V., Mascle, J., Vendeville, B. & Came´ra, L. 2006. The Nile deep-sea fan: an example of interacting sedimentation, salt tectonics, and inherited subsalt paleotopographic features. Marine and Petroleum Geology, 23, 297–315. Longacre, M., Bentham, P., Hanbal, I., Cotton, J. & Edwards, R. 2007. New crustal structure of the Eastern Mediterranean basin: detailed integration and modeling of gravity, magnetic, seismic refraction, and seismic reflection data. EGM 2007 International Workshop Innovation in EM, Grav and Mag Methods: a new Perspective for Exploration Capri, Italy, April 15–18, 2007, 4. Lu¨ning, S., Kuss, J. & Bachmann, M. 1998a. Sedimentary response to basin inversion: Mid Cretaceous– Early Tertiary Pre- to syndeformational deposition at the Areif El Naqa anticline (Sinai, Egypt). Facies, 38, 103– 136. Lu¨ning, S., Marzouk, A. M., Morsi, A. M. & Kuss, J. 1998b. Sequence stratigraphy of the Upper Cretaceous of central-east Sinai, Egypt. Cretaceous Research, 19, 153–196. Mascle, J., Benkhelil, J., Bellaiche, G., Zitter, T., Woodside, J. & Loncke, L. 2000. Marine geologic evidence for a Levantine-Sinai plate, a missing piece of the Mediterranean puzzle. Geology, 228, 779– 782. McClusky, S., Balassanian, S. et al. 2000. GPS constraints on plate kinematics and dynamics in the eastern Mediterranean and Caucasus. Journal of Geophysical Research, 105, 5695– 5719. McKenzie, D. P. 1970. Plate tectonics of the Mediterranean region. Nature, 226, 239–243. McKenzie, D. P. 1972. Active tectonics of the Mediterranean region. Geophysical Journal of the Royal Astronimical Society, 30, 109– 185. Moustafa, A. R., El-Badrawy, R. & Gibali, H. 1998. Pervasive E-ENE oriented faults in northern Egypt and their effect on the development and inversion of prolific sedimentary basins. EGPC 14th Exploration and Production Conference, Cairo, Egypt, 51– 67. Netzeband, G. L., Gohl, K., Hubscher, C. P., BenAvraham, Z., Dehghani, G. A., Gajewski, D. & Liersch, P. 2006. The Levantine Basin – crustal structure and origin. Tectonophysics, 418(3–4), 167– 188. Olivet, J.-L., Bonnin, J., Beuzart, P. & Auzende, J.-M. 1982. Cine´matique des plaques et pale´oge´ographie, une revue. Bulletin de la Socie´te´ Ge´ologique France, 7, 875–892. Operto, S., Lambare´, G., Podvin, P., Thierry, P. & Noble, M. 2003. 3-d rayþborn migration/inversion – part 2: application to the seg/eage overthrust experiment. Geophysics, 68, 1357– 1370.
DEEP STRUCTURES OF THE EGYPTIAN MARGIN Ribodetti, A., Thierry, P., Lambare´, G. & Operto, S. 2000. Improved multiparameter rayþborn migration/ inversion. Society of Exploration Geophysicists, 70, 1032–1035. Rosenbaum, G., Lister, G. S. & Duboz, C. 2002. Relative motions of Africa, Iberia and Europe during Alpine orogeny. Tectonophysics, 359, 117 –129. Rouchy, J.-M. 1986. Les e´vaporites mioce`nes de la Me´diterrane´e et de la mer rouge et leurs enseignements pour l’interpre´tation des grandes accumulations e´vaporitiques d’origine marine. Bulletin de la Socie´te´ Ge´ologique de France, 8, II, n. 3, 511– 520. Ryan, W. B. F., Hsu¨, K. J. et al. 1973. Initial Reports of the Deep Sea Drilling Project, 13, 1447. Samuel, A., Kneller, B., Raslan, S., Sharp, A. & Parsons, C. 2003. Prolific deep-marine slope channels of the Nile delta, Egypt. American Association of Petroleum Geologists Bulletin, 87, 541–560. Sardou, O. & Mascle, J. 2003. Cartographie par sondeur multifaisceaux du delta sous marin profond du Nil et des domaines voisins. Deux cartes (Morphobathyme´trie et mosaiques d’images acoustiques). Special Publication, CIESM, Monaco. Segev, A., Rybakov, M., Lyakhovsky, V., Hofstetter, A., Tibor, G., Goldshmidt, V. & Ben Avraham, Z. 2006. The structure, isostasy and gravity field of the Levant continental margin and the southeast Mediterranean area. Tectonophysics, 425, 137–157. Sengo¨r, A. M. C. 1979. The North Anatolian transform fault: its age, offset and tectonic significance. Journal of the Geological Society of London, 136, 269– 282. Sengo¨r, A. M. C., Yilmaz, Y. & Sungurlu, O. 1984. Tectonics of the Mediterranean Cimmerides: nature
97
and evolution of the western termination of Palaeo-Tethys. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean, Geological Society, London, Special Publications, 17, 77–113. Smith, A. G. 1971. Alpine deformation and the oceanic areas of the Tethys, Mediterranean and Atlantic. Bulletin of the Geological Society of America, 82, 2039– 2070. Smith, A. G. & Woodcock, N. H. 1982. Tectonic synthesis of the Alpine-Mediterranean region: a review. In: Berckhemer, H. & Hsu, K. (eds) AlpineMediterranean Geodynamics. AGU, Geodynamics Series, 7, 15–38. Stampfli, G. M. & Borel, G. D. 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrons. Earth and Planetary Science Letters, 196, 17– 33. Taymaz, T., Jackson, J. A. & Westaway, R. 1990. Earthquake mechanisms in the Hellenic Trench near Crete. Geophysical Journal International, 102, 695–731. Tibor, G. & Ben-Avraham, Z. 2005. Late Tertiary paleodepth reconstruction of the Levant margin off Israel. Marine Geology, 221, 331– 347. Thierry, P., Operto, S. & Lambare´, G. 1999. Fast 2-d rayþborn migration/inversion in complex media. Geophysics, 64, 162– 181. Whiting, B. M. 1998. Subsidence record of early-stage continental collision, Eratosthenes platform (sites 966 and 967). In: Robertson, A. H. F., Emeis, K.-C., Richter, C. & Camerlenghi, A. (eds) Proceedings of the Ocean Drilling Program, Scientific Results, 60, 509– 515.
Controls and evolution of facies patterns in the Upper Barremian– Albian Levant Platform in North Sinai and North Israel MARTINA BACHMANN*, JOCHEN KUSS & JENS LEHMANN Department of Geosciences, Bremen University, P.O. Box 330440, 28334 Bremen, Germany *Corresponding author (e-mail:
[email protected]) Abstract: The Upper Barremian–Albian Levant Platform was studied in North Sinai and Israel (Galilee and Golan Heights) by bio- and lithostratigraphy, facies analyses, and sequence stratigraphy. Integrating shallow-marine benthic foraminifera (mainly orbitolines), ammonite, and stable isotope data resulted in a detailed stratigraphic chart. Transects across the shallow shelf in both regions are based on facies analysis and form the basis for depositional models. In both transects five platform stages (PS I–V) were identified, which differ significantly in their stratigraphic architecture, mainly controlled by local tectonics, climate and second-order sea-level changes. In North Sinai, a transition from a shallow-shelf that is structured by sub-basins through a homoclinal ramp into a flat toped platform is recognized, while the sections in North Israel show a transition from a homoclinal ramp into a fringing platform. Local normal faults influenced the depositional architecture of the Upper Barremian–Lower Aptian strata in North Sinai and were attributed to syn-rift extensional tectonics. Four second-order sequence boundaries were identified, bounding MidCretaceous Levant depositional sequences. These well-dated second-order sequence boundaries are MCL-1 (Late Barremian), MCL-2 (earliest Late Aptian), MCL-3 (Lower Albian), and MCL-4 (Late Albian). The sea-level history of the Levant Platform reflects the Late Aptian–Albian global long-term transgression, while the second-order sea-level changes show good correlation with those described from the Arabian plate.
The Late Barremian– Albian Levant Platform is an ideal case study for studying platform development in a setting of global long-term sea-level change, climate variation, and geodynamics. The platform was located at the northern rim of the North African/Arabian plate and the southern border of the Tethyan Ocean, respectively (Fig. 1a). Extending from southern Lebanon to northern Egypt the Levant Platform strikes out in a narrow stripe parallel to the recent coastline (Saint-Marc 1974; Braun & Hirsch 1994; Kuss & Bachmann 1996; Rosenfeld et al. 1998). Its central part was studied in Israel (Galilee and Golan Heights) and its southwestern part in North Sinai, both characterized by good exposures (Fig. 1b, c). During the Late Barremian–Albian about 500 m of shallow marine deposits accumulated on the Levant Platform. During the Mesozoic, the region was characterized by extensional and compressive tectonical processes related to the opening and closure of the Neotethys (e.g. Keeley 1994; Hirsch et al. 1995; Stampfli & Borel 2002; and other papers in this volume) forming the main tectonical structures such as the Syrian Arc fold belts and the Dead Sea Transform (DST). However, the analysed interval represents a consistent platform succession deposited since the Late Barremian. This interval has been interpreted to be deposited in a post-extensional setting during mid-Cretaceous
transgression and indicating relative tectonical quiescence before the onset of Late Cretaceous compression (Moustafa & Khalil 1990; Hirsch et al. 1995; Kuss & Bachmann 1996). Most former studies concentrate on small areas and/or selected stratigraphical or palaeontological parameters. Litho- and biostratigraphical concepts of the Lower Cretaceous of Galilee and the Golan Heights are summarized in Rosenfeld & Hirsch (2005) and Bachmann & Hirsch (2006); tectonic concepts are summarized in Flexer et al. (2005) and Gilat (2005). Stratigraphical subdivisions of the Lower Cretaceous of the Sinai, are mainly based on Said (1971) and Bartov & Steinitz (1977). Younger interpretations include facies, stratigraphic and sedimentological data (Aboul Ela et al. 1991; Askalany & Abu-Zeid 1994; Bachmann & Kuss 1998; El-Araby 1999; Steuber & Bachmann 2002; Bachmann et al. 2003). Only few studies document marine Barremian– Lower Aptian sediments in North Sinai (Arkin et al. 1975; Morsi 2006; Abu-Zied 2007, 2008). In the present paper, we summarize various stratigraphical data with respect to a consistent correlation of various lithological units in the area and add new data from the Barremian–Lower Aptian succession of North Sinai. The methodology includes bio- and isotope stratigraphy, the interpretation of sedimentological structures in the field,
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 99– 131. DOI: 10.1144/SP341.6 0305-8719/10/$15.00 # The Geological Society of London 2010.
100
M. BACHMANN ET AL.
Fig. 1. Location of the study area. (a) Tectonic map, simplified after Garfunkel (1998) with the location, the two study areas in North Sinai and northern Israel (Galilee and Golan Heights) straddling the DST fault. (b) Aptian–Albian palaeoenvironmental map of the Levant Platform, modified from Rosenfeld et al. (1998) and supplemented by results from North Sinai indicating the extension of the shallow platform facies belts. (c) Gebel (mountains) examined in North Sinai indicated on a satellite map. ‘B’ indicates the tectonically restored position of the Golan Heights section B, when assumed the lateral movement of 100 km as indicated, for example by Garfunkel & Ben-Avraham (1996).
small-scaled geological mapping, and microfacies analysis. Our interpretations are focused on facies reconstructions and long-term, second-order, sealevel changes within a high-resolution stratigraphical frame. Moreover, we estimated the climatic and tectonic influences on the sedimentation. The data allow us to reconstruct and interpret the varying Late Barremian–Albian shelf geometry, as well as the timing and interpretation of the platform development. In this respect syn-depositional tectonical processes were determined, which were previously unknown. These data are interpreted, allowing the detailed timing of the Late Barremian–Albian succession. This results in an evaluation of the facies, long-term local and regional second-order sea-level changes, and climate framework, characterizing the complex sedimentary system of the Levant Platform. Furthermore, the data allow us to the reconstruct and interpret the varying Late Barremian–Albian shelf geometry. In this respect,
the paper is a review. Together with new data, especially from the Barremian–Lower Aptian succession of North Sinai, the dataset allows the interpretation of the platform development. Most former studies concentrate on small areas and/or selected stratigraphical or palaeontological parameters. Litho- and biostratigraphic concepts of the Lower Cretaceous period of Galilee and the Golan Heights are summarized in Rosenfeld & Hirsch (2005) and Bachmann & Hirsch (2006), tectonic concepts are summarized in Flexer et al. (2005) and Gilat (2005). Stratigraphic subdivisions of the Lower Cretaceous of the Sinai are mainly based on Said (1971) and Bartov & Steinitz (1977). Younger interpretations include (Aboul Ela et al. 1991; Askalany & Abu-Zeid 1994; Bachmann & Kuss 1998; El-Araby 1999; Steuber & Bachmann 2002; Bachmann et al. 2003). Only few studies document marine Barremian– Lower Aptian sediments in North Sinai (Arkin et al. 1975; Morsi 2006; Abu-Zied 2007).
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
Geological framework The Levant Platform developed at the northern passive continental margin of the African– Arabian plate close to the plate boundary with the Anatolian plate. The tectonical pattern in the Eastern Mediterranean Levantine region has important influence on the Cretaceous platform development characterized by this position at the triple junction of the African, Arabian, and Anatolian plates (Flexer et al. 2005). The Early Mesozoic opening of the Neotethys generated an east –west striking rift system with subsiding areas and thus created the Levant Basin at the northeastern edge of the African– Arabian plate (Keeley 1994; Hirsch et al. 1995; Garfunkel 1998; Flexer et al. 2005). The shallow-marine Levant Platform formed at the southeastern passive margin of the Levant Basin trending in a narrow strip parallel to the Mediterranean coastline from Lebanon and Syria through Israel to North Sinai, Egypt (Garfunkel 1998). The breakup of the Levant Platform commenced with the Late Coniacian –Palaeogene inversion of extensional faults (Flexer et al. 2005). In northern Israel, the initial ramp geometry of the Early Cretaceous Levant Platform developed into a flat-topped platform in the Early Aptian (Bachmann & Hirsch 2006). In the Albian it is saddled by the formation of fringing rudist-reefs at the shelf break (Sass & Bein 1982; Ross 1992). In North Sinai, the initial Upper Barremian–Lower Aptian platform is not yet analysed. The Upper Aptian– Lower Albian platform geometry is described as a distally steepened ramp (Bachmann & Kuss 1998). The region in between North Sinai and northern Israel is described as a gradual shallow –deep transition (Rosenfeld et al. 1998) known only from the subsurface (Rosenfeld & Hirsch 2005).
The investigated sections and lithostratigraphical concepts In northern Israel/Galilee and the Golan Heights, most outcrops are located along strike–slip faults orientated parallel to the DST fault and in asymmetric Syrian Arc anticlinal structures (Flexer et al. 2005; Gilat 2005). In North Sinai, several anticlinal Syrian Arc structures form good outcrop conditions (Moustafa & Khalil 1990). Owing to the regional differentiated nomenclature, with a high number of units, members and formations, we compile standard sections for the two platform edges composed from several sections in both regions (Fig. 1). Galilee/Golan. We follow the lithostratigraphical subdivisions of the Cretaceous succession in the Golan Heights and Galilee established by Eliezri
101
(1965) and modified by Rosenfeld et al. (1995), while additionally distinguishing a new member (Fig. 2). The terrestrial sandstones at the base of the succession are known as the Hatira Formation and occur in the subsurface of northern Israel (Rosenfeld et al. 1998). The overlying marine succession is subdivided into five formations: Nabi Said (ooid-rich limestones), Ein el Assad (limestones and subordinated marlstones), Hidra (sandstones, marlstones, and few limestone beds), Rama (limestones and marlstones), and Yagur (limestones and dolomites) (Fig. 2). The studied succession is 440 m thick; the standard section is based on three sections studied in Galilee (Har Ramin, Rama and Ein Netofa) and one on the Golan Heights (Ein Quniya) (Bachmann & Hirsch 2006), (Fig. 1). Hence outcrops of marine Lower Cretaceous sediments are generally rare in Galilee and the Golan Heights, two sections (Har Ramin/Galilee and Ein Quniya/Golan Heights) comprise the entire 440 m thick Upper Barremian –Albian succession. Weathering profiles, bedding surfaces, sedimentary structures, stratigraphical, and facies interpretation are described in Bachmann & Hirsch (2006). North Sinai. We use the lithostratigraphical subdivision of the Upper Barremian–Albian succession of North Sinai of Said (1971), applied in current geological maps (Fig. 2). The platform succession starts above terrestrial sandstones (Malha Formation) with a limestone, dolomite, and marlstone alternation characterized by detrital influence (lower Rizan Aneiza Formation). Sandstone, marlstone, limestone, and dolomite alternations with deltaic influence above represent the upper Rizan Aneiza Formation; the subsequent succession of limestones and dolomitic marlstones without considerable siliclastic input was related to the Halal Formation. The thickness of individual sections varies greatly, depending on their palaeogeographical position. From south to north, the thickness of the lower part of the Rizan Aneiza Formation varies from 0 to 170 m and from 120 to 250 m for the upper part of the Rizan Aneiza Formation. The lower Halal Formation studied at Gebel Mansour reaches 220 m. The Upper Barremian–Upper Albian succession of North Sinai was studied in five anticlinal structures, comprising the following Gebels: Amrar, Rizan Aneiza, Raghawi, Mansoura, and southeastern Maghara (Fig. 1c). Within this frame, the Raghawi section (A) describes a proximal setting of the lower Rizan Aneiza Formation (Upper Barremian –lower Upper Aptian) (Fig. 2), while the Amrar section (GA) shows a more distal expression of the same interval. One Rizan Aneiza section (D) includes the Lower/Upper Aptian boundary, while the eastern Maghara section
C. pavonia
oa3
FO O. (M.) s.
deshayesi
Upper
waagenoides sarasini giraudi feraudianus satousi vandenheckii
P. cormyi
tuarkyricus
P. infracretacea
weissi
E. charollaisi
Lower
furcata
O. (M.) texana O. (M.) subconcava O. (M.) parva
P. wienandsi O. (M.) lotzei
P. lenticularis
subnodosocostatum
C. decipiens
Upper
melchioris
Ot-Op O.(M.) texana O. (M.) parva
FO O. (M.) t.
Op O. (M.) parva
FO O. (M.) p.
Pc-Ol P. cormyi O.(M.) lotzei
FO P. c.
local LO C. d. local FO P. i.
Pl P. lenticularis
Pl-Cd P. lenticularis C. decipiens FO P. l.
Yagur Fm
Unit A Halal Fm.
Rama Fm Rama Fm Limestine Mb
oa2
Hidra Fm oa1
Unit A Rizan Aneiza Fm.
Hidra Fm P. wienandsi, H. dinarica, P infracretacea P infracretacea, C. sugdeni, A. podolica, H. dinarica T. marsicana E. charollaisi
Unit 0
Ein El Assad Fm
Nabi Said Fm
Malha Fm Rizan Aneiza Fm.
Hatira Fm
Fig. 2. Biostratigraphic subdivision of the Upper Barremian –Albian ranges of the LFBs, timing of the lithostratigraphical formations and members and ranges of the sections. The ranges of the most important benthic foraminifers and the ammonites are compiled from several authors (see text) and arranged in the chronostratigraphic framework of Ogg et al. (2004). Accompanying organisms are taken from Bachmann & Hirsch (2006) and Bachmann et al. (2003). ‘oa’ refers on ostracod assemblages observed in North Sinai (Bachmann et al. 2003).
M. BACHMANN ET AL.
nodoso costatum
S
Raghawi
O. (M.) subconcava
N
Amrar Raghawi Mansour
Rizan Aneiza
mammillatum
Range of sections
Formations Sinai
Amrar
Os
nolani
Aptian
E. plicatius periadriatic A. dardei E. murgendsis
dentatus
W
Golan
Galilee Golan / Ein Quinya section
loricatus
Range of sections
Gebel Masoura
oa4
jacobi
Barremian
E
Formations Galilee/ Golan Heights
offshore Galilee
lautus
tardefurcate
Lower Cretaceous
Ammonites Accompanying Northern Sinai organisms
Talme Yafe Fm (basin) Carmel reefs
Thierry et al. 1998
LFB (Larger benthic foraminifer biozones) FO / LO Name
Kutatissites bifurcatus Costidiscus recticostatus Heteroceras coulleti Cheloniceras (C.) cornuelianum C. (C.) cf. quadrarium Paradeshayesites cf. grandis Aconeceras nisus C. (Epicheloniceras) tschernyschewi
Substage Middle Lower
Stage Albian
Time (Ma) Series
Larger benthic foraminifers
102
European ammonite zones
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
(MgE) completes the Lower–Middle Albian succession (upper Rizan Aneiza Formation) to the SE. These sections are summarized and span a 40 km wide transect from the distal Amrar anticline in the NW to the proximal Maghara structure in the SE (Fig. 1c). Eight sections from the upper Rizan Aneiza and lower Halal formations (Upper Aptian –Albian) were presented in former publications (Bachmann & Kuss 1998; Bachmann et al. 2003); three of them are added in this work (Fig. 1c) to illustrate facies and sequence stratigraphy.
Methods For the North Sinai sections, weathering profiles, bedding surfaces, sedimentary structures, and facies characteristics are documented. Sample distances of the limestone intervals are usually less than 1 m and may be higher in the marlstone and sandstonedominated intervals. Stratigraphy and facies were analysed by using field data, thin sections, ammonite findings and washed samples, and stable isotopes analyses required powdered samples. Biostratigraphy. Biostratigraphy is based on benthic foraminifers (mainly orbitolinids) and ammonites. The orbitolinid distribution (taxonomy according to Schroeder 1975) allows the definition of six larger benthic foraminifera biozones (LFBs), of which five are originally defined on the first occurrence and/or last occurrence (FO/LO) of the eponymous species in the Galilee/Golan area (Fig. 2, Bachmann & Hirsch 2006). We compare our biostratigraphic data with those of the northern Tethys (Schroeder & Neumann 1985; Arnaud-Vanneau 1998; Arnaud et al. 1998; Bernaus et al. 2002; Schroeder et al. 2002; Conrad et al. 2004) and the Middle East (Saint-Marc 1974; Simmons & Hart 1987; Simmons 1994; Simmons et al. 2000), and integrate the LFBs with the chronostratigraphy of Hardenbol et al. (1998a, b) and Ogg et al. (2004) also considering calcareous algae following Bachmann & Hirsch (2006). Ammonites from nearly all sections in North Sinai are determined and allow comparison with the Tethyan ammonite zonation (e.g. Rawson et al. 1999; Garcı´a-Monde´jar 2009) and to correlate the LFBs with the ammonite range charts. The stratigraphical framework of the Upper Aptian –Albian succession additionally comprises the ranges of ostracods and rudists and is supported by graphical correlation (Bachmann et al. 2003). Stable isotopes. A total of 69 bulk rock samples were analysed for carbon isotopes from the Upper Barremian–Lower Aptian of section A (Fig. 1c), North Sinai. To avoid diagenetic alteration, all samples were selected from the micritic parts of
103
the limestones. The stable isotope composition was measured using a Finigan mass spectrometer at Research Center Ocean Margins (RCOM) Bremen. The results are expressed in the common d-notation in per mille relative to PDB (PeeDee belemnite) standard. Facies analysis. The thin-section analysis includes the determination of the relative abundance of skeletal and non-skeletal components, matrix composition as well as grain size. We distinguish between abundant, common, and rare occurrence of the components. Microfacies types (MFT) were distinguished according to their texture, matrix, and components, added to data observed already in the field such as lithology, bedding patterns terrigenous input, and abundance of ferruginous ooids and macrofossils. The microfacies types are summarized in facies zones (FZ), which include microfacies types occurring in similar environments. Palaeoenvironment reconstruction and sequence stratigraphical interpretation. Within the stratigraphic framework, the combination of small-scale mapping, log-correlation, and microfacies data (including a review of further data) results in palaeoenvironmental maps of North Sinai to northern Israel for five time slices. Despite the limited number of sections, the general platform geometries could be reconstructed. The stratigraphical correlation of all sections within the two analysed depositional areas allows the interpretation of the platform architecture as a function of accommodation, sediment supply, and production. Lateral and vertical facies changes reflect water-depth and palaeoenvironment variations, which were compared between the Galilee/Golan Heights and Sinai and are interpreted in considering sea-level history and tectonic development. Second-order sequences are reflected by lateral and vertical facies distribution patterns. Sequence stratigraphical terminology as originally developed for thirdorder sequences (Vail et al. 1991) is used herein for the second-order cycles. Our tectonic interpretations benefit from the models given by Moustafa (2010).
Stratigraphy The stratigraphical subdivision is focused on biostratigraphy supplemented by stratigraphical interpretations of d13-C curves. Biostratigraphy of the Upper Barremian–Lower Aptian from northern Sinai on ammonites and orbitolinids is integrated with carbon isotope measurements, summarized in Figure 2. This chart allows the comparison of the studied shoal-water environments with stratigraphical subdivisions of basinal settings and platform settings. Subdivision of the Upper
104
M. BACHMANN ET AL.
Aptian– Albian is based on orbitolinids, ostracods, rudists and ammonites, data based mainly on Bachmann et al. (2003). The stratigraphical interpretation of the sections is shown in Figure 3a, b.
Benthic organisms The abundant orbitolinids in all sections and their wide range of habitats (lagoonal to deeper platform areas, e.g. Pittet et al. 2002) allow the subdivision of the Upper Barremian–Middle Albian succession into five larger benthic foraminifer biozones (LFBs). From base to top these are: Pl–Cd Palorbitolina lenticularis–Choffatella decipiens, Pl P. lenticularis, Pc–Ol Praeorbitolina cormyi – Orbitolina (Mesorbitolina) lotzei, Op O. (M.) parva, Ot –Op O. (M.) texana– O. (M.) parva in the Golan Heights/Galilee area (Bachmann & Hirsch 2006). The occurrence of Eopalorbitolina charollaisi, which is not younger than Late Barremian (Clavel et al. 1995; Arnaud et al. 1998; Conrad et al. 2004), in North Sinai (Amrar section) and Galilee (Ein Netofa area), confirms the Late Barremian age for the LFB Pl–Cd. This stratigraphical frame is extended by a sixth LFB (Os) occurring in North Sinai, which is characterized by the first occurrence (FO) of O. (M.) subconcava (Fig. 2). LFB (Os) ranges from uppermost Aptian to Albian and coincides with the FO of O. (M.) subconcava in the latest Aptian of the Tethyan realm (Castro et al. 2001; Schroeder & Neumann 1985) and in the Middle East (Simmons et al. 2000). The upper boundary of LFB (Os) is placed at the FO of Orbitolina (C.) corbarica in North Sinai (Bachmann et al. 2003), coinciding with the Albian/Cenomanian boundary. The orbitolinid-distribution of the Upper Barremian–Lower Aptian sections in Sinai is presented in Figure 3. Ranges of the LFBs for Golan Heights –Galilee and the younger sediments of Sinai are shown in Figs 3 and 4.
Ammonites Ammonites were sampled in North Sinai, at Gebel Raghawi (section A) and three sections 3 km east and west of section A (R1, K and JJ, Fig. 1c), which can be directly correlated by characteristic marker beds. Additionally, a few ammonites were sampled in the Rizan Aneiza area. The Upper Barremian –Lower Aptian of these sections contains rare heteromorph ammonites, including heteroceratids. One of these, Kutatissites bifurcatus (section R), is an index species for the Upper Barremian, but it ranges into the Lower Aptian (Aguado & Company 1997). Heteroceras coulleti, occurring at Gebel Raghawi, is known from the Upper Barremian of southern France (Delanoy 1997; Delanoy & Ebbo 2000). This
species occurs first in the uppermost Hemihoplites feraudianus Zone and ranges up to the lower Imerites giraudi Zone (middle Late Barremian, Fig. 2) (Delanoy 1997). Cheloniceras (Cheloniceras) and deshayesitids are the most common ammonites of the Lower Aptian in North Sinai. Cheloniceras (C.) cornuelianum and Cheloniceras (C.) cf. quadrarium (section JJ, Raghawi,) are widely distributed taxa. C. (C.) cornuelianum is known from the Boreal of southern England (Casey 1962) and from Spain, Colombia, Texas, Georgia and Turkmenistan. C. (C.) quadrarium is known from southern England and Colombia (Kotetishvili 1970; Young 1974; Lillo Bevia 1975; Sharikadze et al. 2004). These cheloniceratids indicate the Deshayesites deshayesi and Tropaeum bowerbanki zones of the Boreal zonation [sensu Casey et al. (1998), the T. bowerbanki Zone, corresponding to the D. furcata zone in the present paper, late Early Aptian, Fig. 2]. In the same interval Pseudohaploceras matheroni occurs, that is of limited use for biostratigraphy. The base of the Upper Aptian is characterized by an Aconeceras nisus found at northern Raghawi (sections R and A), a species that is also common in Russia and northern Germany (Kemper 1995; Mikhailova & Baraboshkin 2002). A. nisus is accompanied by Cheloniceras (Epicheloniceras) tschernyschewi, which is described from the Boreal in the equivalent of the Cheloniceras (Epicheloniceras) subnodosocostatum zone (early Late Aptian, Fig. 2).
Stable isotopes Carbon isotope values vary between – 1.0‰ and 3.5‰ (Fig. 3). While the Upper Barremian – lowermost Aptian values display a gradually increase from 0.5 to 3.0‰ (interrupted by a small drop only), the upper part of the Lower Aptian succession is characterized by a distinct drop (to – 1‰), a subsequent increase (to 3.5‰), and a drop (to 1‰) of the d13C values until the Lower/Upper Aptian boundary (Fig. 3). Although, the measured data show general lighter values than those from pelagic sections (e.g. Menegatti et al. 1998; Luciani et al. 2001; Weissert & Erba 2004) or shallow platform sections (e.g. Strasser et al. 2001; Wissler et al. 2003) in the northern Tethys, the general trend of our data is similar. The absolute values fit well to measurements from the southern Tethys (e.g. Oman: Immenhauser et al. 2005; Tunisia: Heldt et al. 2008) or the Pacific (Jenkyns & Wilson 1999; Price 2003). Menegatti et al. (1998) divided the Cismon carbon isotope stratigraphy curve for the Upper Barremian and the Aptian in eight segments (C1 –C8), widely used for subdividing Barremian –Aptian carbon isotope curves (e.g. de Gea et al. 2003;
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
105
Fig. 3. Sections A (Gebel Raghawi), D (Rizan Aneiza), and GA (Gebel Amrar) reflect a transect across the Upper Barremian– Lower Aptian shallow shelf in North Sinai. The sections include lithology, d-13C isotope bulk rock values, the occurrence of biostratigraphic marker fossils, and the facies interpretation.
106
Fig. 3. (Continued)
M. BACHMANN ET AL.
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
107
Fig. 4. Correlation of the North Sinai and Galilee–Golan Heights succession. A North Sinai standard section and the Raghawi standard section was chosen, for Galilee–Golan Heights a composite section was drawn from the sections Har Ramin and Ein Quniya. Indicated are the biostratigraphic interpretation (based on the LFBs), the depositional environment, the interpretation of the second-order sequences, the main changes in detrital input, and ferruginous ooid content, as well as for the Galilee–Golan Heights area the main environmental interpretation and the position of the palaeocoastline.
Heimhofer et al. 2004; Immenhauser et al. 2005; Renard et al. 2005; Heldt et al. 2008). The segments are dated by planktic foraminifera and calcareous nanoplankton zones (Menegatti et al. 1998; de Gea et al. 2003; Fo¨llmi et al. 2006). The sections studied can be correlated with the segments as follows: the Upper Barremian interval, characterized by bulk carbon isotope values between 20.5‰ and 3.0‰, may correlate with segments C1 and C2. The strong decrease to values, around –0.1‰, and
a subsequent increase to 1.0‰, reflect the intervals C3–C4. A relative stable segment may correlate with C5, widely interpreted as reflecting the Oceanic Anoxic Event (OAE) 1a (Menegatti et al. 1998; Luciani et al. 2006). The following increasing carbon isotope values in Cismon (C6) are reflected by an increase to a value around 3.5‰. C7 is indicated again by stable carbon isotope values. The subsequent strong decrease in carbon isotope values reflects the change from segment C7 to C8.
108
M. BACHMANN ET AL.
The Barremian/Aptian boundary in North Sinai coincides with a negative carbon isotope peak between C1 and C2 at 115 m (Raghawi section A, Fig. 3a), which correlates with an interval around the Barremian/Aptian boundary of Moullade et al. (1998) and Godet et al. (2006). Fo¨llmi et al. (2006) interpreted this peak as marking the Barremian/Aptian boundary. The equivalent of an OAE 1a is located at 160 to 175 m in the Raghawi section, which is in the Middle L. cabri/deshayesi zone according to Ogg (2004). The Lower/Upper Aptian boundary is suggested in a position just below the C7 –C8 interval boundary around 200 m in the same section (Fig. 3a).
Stratigraphical interpretation and correlation of the sections All data on stratigraphy are summarized in the regional correlation for the southern Levant Platform (Fig. 4). The first marine sediments are dated as Late Barremian by Kutatissites bifurcatus and Heteroceras coulleti at the lower part of Raghawi sections A, K and R in North Sinai and by Eopalorbitolina charollaizi in the Amrar section (GA, North Sinai) and Ein Netofa section (Galilee). The occurrence of P. lenticularis and C. decipiens in all sections indicate LFB Pl –Cd. The Barremian/Aptian boundary is characterized by the FO of the Lower Aptian index species Triploporella marsicana (Masse 1998) in Galilee/Golan Heights and is correlated with an interval above the last occurrence of H. couleti and K. bifurcatus in North Sinai. The boundary is possibly located at the negative carbon isotope peak at 115 m (Raghawi section A). This interpretation coincides widely with that of Abu-Zied (2007), based on small benthic and planktic foraminifers for the Raghawi area. The occurrence of P. lenticularis, C. decipiens, P. cormyi, P. wienandsi and O. (M.) lotzei allows the subdivision of the Lower Aptian into the LFBs Pl and Pc–Ol in both areas (Figs 3 & 4). While LFB Pl comprises the basal Lower Aptian, LFB Pc–Ol widely coincides with the negative isotope interval marking the OAE 1a in the upper Lower Aptian. In North Sinai, the Lower/Upper Aptian boundary is indicated by the FO of A. nisus [C. (E.) subnodosocostatum ammonite zone] and a decrease of d13C values (Raghawi, section A, Fig. 3). This is 35 m below the boundary defined by Abu-Zied (2007) on the base of planktic foraminifers and suggested by Morsi (2006) on the base of ostracods. The upper part of the basal Upper Aptian C. (E.) subnodosocostatum ammonite zone furthermore comprises the FO of O. (M.) parva (e.g. Rizan Aneiza sections D and RN) and C. (E.) tschernyschewi (Fig. 3). In the Raghawi area, LFB Op–Ot is indicated
only a few metres above the FO of O. (M.) parva (Fig. 3) indicating a late FO of O. (M.) parva in comparison to other studies (Lower/Upper Aptian boundary according to Arnaud-Vanneau 1998). In Galilee/Golan Heights, the Lower/Upper Aptian boundary is suggested to be 40 m below the FO of O. (M.) parva, according sequence stratigraphical and lithological correlations with North Sinai. The correlation of the Upper Aptian –Albian succession between both areas is more difficult because of the rare biostratigraphical markers in the Galilee– Golan Heights region, while the North Sinai succession is quite well dated. The Aptian/ Albian boundary in North Sinai is marked by two distinct ostracod assemblages and confirmed by graphic correlation (Bachmann et al. 2003). This boundary was interpreted to be within the Limestone Member (lower Rama formation) in the Golan Heights (Bachmann & Hirsch 2006). Based on the correlation of second-order sequences between Sinai and the Golan Heights (presented in this paper, Fig. 4) the boundary is expected to be located lower in the section, around the boundary Hidra/Rama Formation, which is conformed to ostracod data from Rosenfeld et al. (1995). In North Sinai, the Early/Middle Albian boundary is placed above the FO of Eoradiolites liratus (occurring since the latest Early Albian, Steuber & Bachmann 2002), by graphic correlation and the FO of Desmoceras (D.) latidorsatum (Bachmann et al. 2003), which first appears in the Late Albian (Gale et al. 1996). The Middle/Late Albian boundary is characterized by the LO of O. (M.) parva and the FO of N. simplex (Bachmann et al. 2003). Biostratigraphical data on the upper part of the succession in Golan–Galilee are generally rare and is based on ostracod assemblages or rare findings of the ammonite Knemiceras in Galilee (Rosenberg 1960; Rosenfeld et al. 1995).
Platform development Facies zones: platform environments The classification of facies zones (FZs) and the interpretation of the depositional environments are summarized in Figure 5, following the classification given by Bachmann & Hirsch (2006, Galilee – Golan Heights) and Bachmann & Kuss (1998, Sinai). The FZs range from deeper open platform to terrestrial and siliciclastic-influenced restricted platform. Nearly every FZ contains more than one facies type reflecting similar environments of deposition. Facies interpretations involve the comparison with other Barremian– Albian Tethyan platforms (Arnaud-Vanneau & Arnaud 1990; Masse 1993; Masse et al. 1995; Husinec 2001; Bernaus
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
Fig. 5. Facies classification and interpretation on the base of components, composition and sedimentary structures.
109
110
M. BACHMANN ET AL.
Fig. 5. (Continued)
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
Fig. 5. (Continued)
111
112
M. BACHMANN ET AL.
et al. 2003; Masse et al. 2003), and particularly with those in the southern Tethys (Pittet et al. 2002; Van Buchem et al. 2002; Immenhauser et al. 2005). Most FZs occur in both study areas and at different stratigraphical levels. Their distribution is indicated for three Upper Barremian– Aptian sections: section A (Fig. 6e & f) (supplemented by data from the sections K and R), sections GA and D (Figs 1c, 3a, b & 6d). To investigate the lateral facies variations in North Sinai during the Upper Aptian– Lower Albian (Fig. 7) sections D and MgE are added to data by Bachmann & Kuss (1998) and Bachmann et al. (2003). For the Galilee –Golan Heights the facies description summarizes data presented in Bachmann & Hirsch (2006). Two standard sections illustrate the facies variations between the platform settings at Sinai and Galilee– Golan Heights (Fig. 4). General trends for terrigenous input, formation of ferruginous ooids, main carbonatic components, and grain diversity are added. Lateral variations of FZs are recognized when correlating the different sections and allow an interpretation of the platform geometry for both areas studied during several time slices. Vertical variations indicate the changes of the platform environment associated with varying external and internal parameters (e.g. climate, sea-level change, sediment accumulation). The distribution of FZs characterizes major lithological, facies, and environment changes, leading to a subdivision of five PS (PS I to V) within the evolution of the Levant Platform. These PS are decoupled from the definition of formations. Facies development is described separately for both study areas, to focus on the regional geometric patterns.
The North Sinai platform stages PS I. PS I comprises the first Lower Cretaceous marine sediments and is characterized by limestones, marlstones, and siliciclastics deposited in varying depositional environments reaching from terrestrial to deeper marine. PS I comprises the lowermost Rizan Aneiza Formation (Upper Barremian– basal Lower Aptian, LBFs Pl –Cd, and lower Pl). An alternation of marl, siltstone, dolomite, and limestone belonging to various FZs (FZ 1 to 8, Fig. 5) in the northern sections A (Raghawi, 85 m thickness) and GA (Amrar, 95 m thickness, Figs 3 & 7) is interfingering with terrestrial sandstones (Malha Formation, FZ 9) 5 km south of Raghawi (Mansoura area – Fig. 1c). At section A (Figs 1c & 3a) PS I is dominated by protected lagoonal environments, characterized by bioclastic wacke and packstones, with benthic foraminifers, shell and echinoderm debris and
oncoids (FZ 6), partly rich in orbitolinids and cyanophyceans. These limestones alternate with marlstones, silty marlstones, and siltstones containing few autochthonous benthic foraminifers, ostracods, small solitary coral colonies, echinoderm debris, and further bioclasts (FZ 6-2) that are indicating a varying terrestrial input. Several levels contain ferruginous ooids. In the lower part, intercalated peritidal to coastal sandstones and siltstones (FZ 8a) indicate two regressive events. In the same interval, ammonite-bearing marlstones indicate a sporadic influence of open-marine conditions. Among those ammonites are several phylloceratids (e.g. Euphylloceras aff. inflatum), which are interpreted as reflecting deeper-marine conditions (Westermann 1996). Intercalations of oolithic, intraclastic and bioclastic grainstones (FZ 2) mark the repeated influence of shallow subtidal, high-energy conditions. Upward increasing, intercalated limestones rich in oncoids (FZ 3, mainly MFT 3a) reflect increasing open platform shallow subtidal conditions. In the uppermost part of PS I frequent intercalated dolomites (early diagenetic dolomite formation), are suggested as representing short events of emergence. The upper boundary of PS I is settled at the point of increasing carbonate content and marked by the last occurrence of the ferruginous ooids and the first occurrence of dasycladacean-rich open lagoons (FZ 4) above. At section GA (Figs 1c & 3a) the succession is dominated by oncoid-rich limestones of FZ 3, showing a trend to more open lagoonal/platform conditions. Siliciclastic input is less frequent and of finer grain size, while ferruginous ooids occur only rarely. Deeper platform environments (FZ 1), comparable to the ammonite bearing beds at section A, occur only in the upper part of PS I at section GA. About 5 km towards the south, terrestrial sediments characterize PS I and indicate that the coastline was located north of the Gebel Mansour (Fig. 8a, f). The proximal position of section A is reflected by intercalated coastal sediments and high contents of clay and ferruginous ooids. Further north, at the distal section GA, the openmarine facies lacks coastal influence. However, depositional environments are partly shallower than in the proximal SW, suggesting deposition on a submarine high, separating the protected area at section A from open-marine environment. The open-marine deeper environments are expected to be in a close position to the study area because of the occurrence of phylloceratids. PS II. During PS II open platform environments in the north are interfingering with terrestrial sediments in the south.
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
The main part of the Lower Aptian sediments and the first metres of the lowermost Upper Aptian (lower LFB Op-zone/lower C. (E.) subnodosocostatum zone), belong to PS II. Marine limestones, marlstones, siltstones, and sandstones of PS II crop out at section GA (90 m thickness, Fig. 3b) and at section A (110 m thickness, Figs 3a & 6e, f), while at Gebel Mansour further to the south, only terrestrial siliciclastics occur. PS II starts with shallow-subtidal openmarine environments containing diverse bioclasts and dasycladaceans (section A, FZs 3 and 4). Intercalated near-shore sandstones (FZ 8) and a remarkable red-brownish dolomite horizon (early diagenetic dolomite formation, Fig. 6e) are suggested as representing events of regression and emergence. Above, marlstones containing some ammonites (FZ 1) are overlain by limestones rich in calcareous algae (FZ 4), limestones with diverse biota and planktic foraminifers (FZ 1a) and limestones characterized by high amounts of orbitolinids (FZ 1b). About 40 m of marlstones with corals at the base and ammonites and planktic foraminifers (FZ 1) are marked by upward decreasing carbonate content. About 4 m of intercalated grainstones with ferruginous ooids (FZ 2), indicate a short event of shoaling. Maximal 10 m of marly claystones of FZ 1 form the top of this interval. The upper boundary of PS II is formed by an erosional unconformity, which cuts into the upper marly claystone at section A and into the ferruginous-ooid interval below E of the section A. At Gebel Amrar limestones deposited in high-energy shoal environments dominate the lower part of PS II, while silty marls of open-marine and lagoonal environments (FZ 1 and 6-2) occur in the upper part. Generally PS II is characterized by subtidal open-platform environments with upward increasing influence of open-marine conditions. The coarse grained terrestrial input is strongly reduced in the upper part and ferruginous ooids are missing until a short shoaling interval, allowing the deposition of near coast sediments with ferruginous ooids, at Gebel Raghawi. Its southward proximal regional extension is very similar to PS I. Terrestrial sediments crop out in the entire southern region and indicate a widely unchanged position of the coastline between the northern Raghawi and Mansour area (Fig. 1). The only exception is one marly bed with corals, which was observed in a single locality at southern Gebel Maghara close to section A (Fig. 7). Because of the erosive nature of PS III above, a now eroded uppermost part of PS II, ranging further to the south, cannot be excluded. The deposition of high-energy shoals at section GA with deeper sediments behind makes the existence of a barrier in that area highly probable.
113
PS III. Sandstones deposited in terrestrial and coastal environments characterize PS III. This interval is attributed to the lowermost Upper Aptian (subnodocostatum zone; Fig. 2). At section A (most proximal section) the interval consists of only 10 m of caolinitic quartzose sandstones (Fig. 6e), which are interpreted as terrestrial deposits (FZ 9). The erosional base of the succession cuts into different levels of the underlying succession. At section D (Fig. 1c) 40 m of coastal siliciclastic sediments (FZ 8a) mark that interval, which are two times interrupted by lagoonal and shoal originated carbonates (FZ 4 and 2). At section GA the interval is formed of 10 m of tidal flat dolomites (Fig. 3b). Compared to PS I and II, the coastline moved to the north and was now located between sections A and GA, close to sections RN and D (Fig. 1). Section A comprises the terrestrial realm, while sections RN and D indicate a near shore position. The more distal section GA was also partly emerged, but not affected by siliciclastic deposition, speaking for a local high/swell (Figs 7 & 8a). PS IV. Shallow subtidal marine limestones, deltaic siliciclastics, and a southward transgression mark this platform stage. PS IV comprises the upper Rizan Aneiza Formation, which was deposited during the Late Aptian until the end of the Middle Albian (Bachmann et al. 2003). Marine sediments of this interval were documented from sections D (Fig. 3a) and RN (240 m, Fig. 7), section R (225 m, Fig. 4), sections ME and M (190 m, Figs 7 & 6h), and section MgE (140 m, Fig. 7). In the northern and central distal sections, the marine sedimentation starts in the upper C. (E.) subnodocostatum zone (sections C, RN and R), while in the proximal southeastern area the first marine sediments are dated as Lower Albian (sections M, ME; Bachmann et al. 2003; and section MgE, Fig. 7). At the northern section RN marlstone –limestone alternations of the shallowsubtidal FZ 3 to protected lagoon FZ 6 with some intercalated small-scaled rudist biostromes (FZ 2-2), bioclastic to oolithic shoals (FZ 2), and some meters of open-marine sediments (FZ 1) prevail. The Raghawi– Mansour area was characterized by a delta system with sandstones, siltstones, and marlstones of FZ 8, prograding repeatedly into the area studied and alternating with shallow subtidal, openmarine (FZ 3), and lagoonal sediments (FZs 4, 6 and 7; Bachmann & Kuss, 1998) and furthermore high energy shoals (FZ 2, Fig. 5). The amount and frequency of fine and coarse grained siliciclastic intercalations increase to the SE until they dominate the succession at section MgE (Fig. 7). Up to 4 m thick rudist biostromes (FZ 2-2) with
114
M. BACHMANN ET AL.
Fig. 6. Facies and bedding-patterns of the Upper Barremian– Albian strata. (a) Nabi Said Fm. The Har Ramin section, Galilee, comprises the entire Lower Aptian– Middle Albian succession. (b) Ein El Assad Fm. In the Ein Netofa section (Galilee), the cross-bedded limestones of the Upper Barremian Nabi Said Formation are well developed. (c) Hidra Fm. Well-bedded limestones mark the Ein El Assad Formation in the Har Ramin section. (d) Rama Fm. The Rizan Aneiza
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
115
Fig. 7. Correlation of the sections in North Sinai summarizing new and known sections (Bachmann et al. 2003). The sections are correlated along marker horizons, sequence boundaries, and the biostratigraphic frame. Platform stages are indicated by different colours. The location of the sections is shown on the small map.
lateral extensions of tens of metres are intercalated in the succession at sections RN, D, A and ME (Figs 7 & 6h). At section A, an up to 4 m-thick coral biostrome (FZ 2-2a) is intercalated in the lower part of PS IV, while the entire succession is marked by high amounts of ferruginous ooids, which decrease in frequency and amounts distally (section RN, Fig. 7). Intercalated emergence
horizons are common and marked by rhizolithes or ferruginous crusts that can be correlated over large areas in the region (Bachmann & Kuss 1998). The facies evolution of PS IV reflects third-order sealevel changes (Bachmann & Kuss 1998). All the described sections of PS IV follow a proximal–distal transect across the shallow shelf (Fig. 8). A retrogradation of facies belts and
Fig. 6. (Continued) Formation at its type-locality is characterized by marly and sandy limestones (section D). (e) Lower part of the Raghawi Section A: coastal sandstones marlstones, and upward increasing dolomites and limestone mark the Upper Barremian and Lower Aptian platform stages PS I and PS II (Rizan Aneiza Formation, Gebel Raghawi). The arrow indicates a prominent dolomitic horizon 150 m above the base of the section, which is interpreted as emergence surface. (f) At northern Gebel Raghawi the entire Barremian– Middle Albian succession crops out and is bounded in its upper part by a major reverse fault. The arrow indicates the terrestrial sandstones of PS II shortly above the Lower– Upper Aptian boundary. The succession above the sandstone shows the Upper–Middle Albian PS IV marked by increasing limestone content (Upper Rizan Aneiza Fm). (g) The figure comprises nodular limestones, marlstone, and the sandstone (arrow) of the uppermost Lower Aptian to lowermost Upper Apian. (h) The Upper Aptian to Lower Cenomanian succession at Gebel Mansoura is characterized by deltaic sediments in its lower part and pure limestones with huge rudist biostromes (arrow) in the upper part.
116
M. BACHMANN ET AL.
Fig. 8. Schematic transect through the North Sinai shallow shelf displays the Late Barremian– Aptian evolution and involves the factors extension along normal faults and sea level change (a –e). Palaeoenvironmental maps indicate the lateral extension of facies zones (f–h). (a, b) PS I and PS II are characterized by increasing sea level. (f) The southern coastline was located at the Raghawi–Mansoura normal fault, with a possible short-termed southward transgression
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
transgression onto former terrestrial areas is obvious (Fig. 8). The coastline was located between Gebels Raghawi and Mansour during Late Aptian times, between Mansour and Maghara East around the Aptian/Albian boundary, south of Maghara area during Early and early Middle Albian times, and more than 50 km further south during the late Middle Albian (Bachmann & Kuss 1998). This stepwise retrogradation is also indicated by a southward decreasing thickness, which is mainly owing to a later onset of sedimentation (Fig. 8). PS V. PS V is characterized by relatively uniform deposition of limestones in the entire region. PS V comprises the lower part of the Halal Formation (Late Albian –Early Cenomanian, Bachmann et al. 2003). At Rizan Aneiza and Raghawi only the lower part of PS V crop out (section RN: 40 m, section R: 145 m, Fig. 4), but at Mansour (section M, Fig. 6h) and Maghara SE (section MgE) 200 m of sediment occur (Fig. 7). The base of PS V is characterized by an up to 20 m thick rudist biostrome (FZ 2-2, sections RN, R, M, Figs 4 & 7), which is interfingering southward (section MgE) with lagoonal limestones (FZ 6a). In all localities, a sudden stop of the siliciclastic input occurs below this bed. Overlying this, limestone– marlstone alternations of FZs 5, 6 and 7 prevail. Marly lagoonal environments characterize PS V with high amounts of gastropods and orbitolinids (FZ 6-2) and the repeated formation of small and larger rudist biostromes (FZ 2-2) of protected lagoons, as well as stromatolithic tidal flats (FZ 7) above. Intercalated oolithic and bioclastic shoals (FZ 2) indicate high-energy facies. Open-marine subtidal facies are rare, while deeper platform facies types are missing. Repeated emergence horizons characterized by rhizolithes, tepee structures, dolomitic, and in the lowermost part by carstification are intercalated and reflect sequence boundaries of the third-order sequences (Bachmann & Kuss 1998). The increasing abundance of dolomitized tidal flat deposits (FZ 7) in the upper part of PS V indicates a general shoaling of the depositional area. During PS V, the main position of the coastline was located clearly south of the investigated area (Bachmann & Kuss 1998). Decrease in siliciclastic input and short-termed sea-level lowstands produced emergence in contrast to the prograding delta wedges of former periods. Thickness of the
117
sediments is clearly higher in the respective distal sections. The lateral distribution of FZs indicates a very similar, shallow, protected marine environment covering large parts of the North Sinai platform (Fig. 7).
The Galilee –Golan Heights platform stages The platform stages PS I –V are well documented at Galilee –Golan Heights and allow the correlation of four sections (Figs 1b & 9) with North Sinai. These sections represent a transect across the shallow shelf of the northeastern Levant platform located on both sides of the DST fault. The marine succession starts above terrestrial sandstones of the Hatira Formation. PS I. In northern Israel/the Golan Heights PS I comprises the first marine Lower Cretaceous sediments similar to North Sinai. Limestones deposited in shallow subtidal environments predominate. PS I correspond with the Nabi Said Formation, which is of Late Barremian to basal Early Aptian age (LBFs Pl– Cd and lower Pl) (Fig. 4). In central Galilee, they are well exposed in the Ein Netofa section, but at the Golan Heights only a few metres of basal Lower Aptian occur (Bachmann & Hirsch 2006). The Ein Netofa section is characterized by about 32 m of cross-bedded oolithic and bioclastic grainstones, partly enriched in ferruginous ooids and quartz (FZ 2). They increasingly alternate upward with bioclastic packstones (FZ 3) and wackestones reflecting protected environments (FZ 4, 5 and 6). At the Golan Heights 10 m of sediment were deposited in protected lagoons with tidal flats (FZ 4 and 7). During the Barremian, an open-platform highenergy facies belt (Ein Netofa section) was interfingering to the east and SE with terrestrial sediments (Hatira Formation) in the region between Galilee and Golan (Hirsch 1996). The abundance of ferruginous ooids and coarse grained quartz indicates the vicinity of the proximity to the coastal area, which retrograded during the Late Barremian– earliest Aptian to an area east of the Golan Heights. A slightly inclining ramp geometry is documented by a gradient from high- and low-energy open subtidal (FZ 2 and 3) facies occurring in the Ein Netofa section to protected subtidal environments (FZ 3 and 4) at Har Ramin, and intertidal facies-types (FZ 7) at the Golan Heights.
Fig. 8. (Continued) at the end of PS II. (c) A significant drop in sea level caused emergence of wide areas and erosion of the former relief. (d) During PS IV, a homoclinal ramp characterizes the depositional architecture. A delta system developed in the proximal areas, interfingering with lagoonal and shallow ramp sediments. (g) The strike of the facies belts changes from SW– NE to WSW–ENE. (h) During Late Albian drown by rising sea-level, a shallow platform without major changes developed.
118
M. BACHMANN ET AL.
Fig. 9. Six models explaining the development of the depositional architecture of the Levant Platform in North Sinai and Galilee–Golan Heights area.
PS II. Limestones upward alternating with marlstones characterize the marine sediments of PS II. PS II comprises the Ein El Assad Formation and the lower part of the Hidra Formation, corresponding to the Lower Aptian (upper LBFs Pl and Pc–Ol, Fig. 2) – lowermost Upper Aptian. At the boundary between PS I and PS II terrigenous input was interrupted and the lower part of PS II (Ein El Assad Formation) forms a prominent landmark of 55 m-thick pure limestones in Galilee and the Golan Heights. The upper part (lower Hidra Formation) crops out in the Golan Heights and is known from the subsurface only in the Galilee (Rosenfeld et al. 1998). Bioclastic packstones and wackestones of the Lower Ein El Assad Formation belong mainly to the low-energy open and protected lagoon FZs 3, 5 and 6 with high amounts of calcareous green algae in some layers (FZ 4). Intercalated are a few beds deposited under high-energy (FZ 2), restricted or tidal flat conditions (FZ 7). In the upper part of the Ein El Assad Formation increasing
open marine conditions are indicated by the occurrence orbitolinids-rich and planktic foraminiferscontaining sediments of FZ 1 in all sections. Relatively uniform depositional conditions prevailed in the entire area with an only small gradient to more protected environments from west to east characterizing a homogenous shallow-platform environment reaching to an area west of the Golan area. Decreasing siliciclastic input of quartz and ferruginous ooids indicate greater distance of the shoreline. The upper part of the Ein El Assad Formation (LFB Pc –Ol) is characterized by a significant synchronous deepening event establishing deeper marine environments over the entire Galilee – Golan Heights region (Fig. 4). The onset of the Hidra Formation marks increasing terrigenous input. The siltstones, marlstone and limestone alternations were deposited in open-marine subtidal (FZ 3, 4) to protected environments (FZ 6 and 8). In the higher part of PS II shoaling and progradation of the coast line occur (Fig. 4).
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
PS III. Siliciclastic sediments characterize PS III in northern Israel. PS III comprises 5 m of the middle part of the Hidra Formation, marked by fluvial sandstones of FZ 9 that can be not dated exactly. The underlying PS II extends at least into the late Early Aptian and the overlaying PS VI is clearly related to the lower Upper Aptian. Thus, PS III is not older than late Early Aptian and not younger than early Late Aptian. PS III represents an important emergence horizon. A terrestrial facies can be proved from the Golan Heights, documenting the progradation of the coast line. Rosenfeld et al. (1998) described similar siliciclastic sediments of the Hidra Formation from the subsurface of the Galilee area, however without emergence. They describe a brackish facies from the Middle Hidra Formation indicating at least a near coast position for that region. PS IV. PS IV comprises about an 150 m thick succession of marine marlstones and limestones. It corresponds to the Upper Aptian to Middle Albian (LBFs Op and Ot, possibly Os) Rama Formation and is documented from the Golan Heights (Ein Quniya section) and from Galilee (Har Ramin section, Fig. 6a). The 45 m-thick limestone member forming the base of the lower Rama Formation (35 m at the Golan Heights) is characterized by often nodular bedded limestone deposited in the various FZs, reaching from openmarine deeper platform (FZ 1; only at Har Ramin) to tidal flat sediments (FZ 7) with a maximum occurrence of the open-marine shallow platform environments (FZ 3). Terrestrial input is low and rarely ferruginous ooids occur. The upper part of the Rama Formation is composed of limestone – marlstone alternations, mainly deposited in protected environments (FZs 5, 6 and 7) in Galilee and the Golan. Open lagoonal environments (FZ 3) occur only at Har Ramin. Slightly higher siliciclastic input of sand and marlstones occurs in the Lower Albian. During deposition of the Rama Formation, the marine facies belt reaches again to an area significantly west of the Golan area. However, siliciclastic input of quartz and ferruginous ooids is still obvious. The facies indicates that a low-energy lagoon dominated the entire Formation, reaching from Golan Heights to Galilee, and limited to the west by fringing rudist bioherms at the Carmel area in the western Galilee (described by Sass & Bein 1982), and to the east by terrestrial deposits in Jordan. PS V. PS V includes dolomites of the Yagur Formation and was analysed only at Har Ramin (Galilee), where pure dolomites reflect protected lagoon (FZ 6) and tidal flat deposition (FZ 7). According to Rosenfeld et al. (1998) a late Middle
119
Albian can be assumed. Similar sediments are described from cores in Galilee, from the Golan (Hirsch 1996), and from southern Israel (Rosenfeld et al. 1998; Rosenfeld & Hirsch 2005) indicating an enlarged shallow platform covering Israel.
Structural control on the Late Barremian – Albian deposits of the Levant Platform The North Sinai record. Deposition of Early Cretaceous strata took place between Late Jurassic extensional and Late Cretaceous compressional tectonics. While the Late Jurassic extension resulted in predominantly NW dipping extensional normal faults (Moustafa & Khalil 1990; Garfunkel 1998), the Late Cretaceous compressional stage resulted in three fold ranges in North Sinai, owing to the inversion of the former extensional faults. The interpretation of the Late Barremian –Albian sedimentary record and the arrangement of facies belts in the palaeogeographical maps allow the reconstruction of extensional movements along these faults during Early Cretaceous times. Facies zones are generally orientated in WSW– ENE trending belts parallel to these major extensional faults (Figs 8 & 9). The prominent anticlinal structure of Gebel Maghara lies at the northernmost fold range, in between east –NE elongated belts of right-stepping en-echelon folds, interpreted as representing the strike –slip rejuvenation of deep-seated earlier extensional faults (Moustafa & Khalil 1990). Those reverse faults were observed at Gebel Amrar and Gebel Raghawi (Fig. 8g; compare Moustafa 2010), where abrupt facies and thickness changes of Early Cretaceous sediments across these faults indicate their importance during earlier extensional stages and might have been active until the early Late Aptian. The southern SW –NE trending fault (Raghawi –Mansour fault) separates the Raghawi and the Mansour areas with its lateral elongation further NE, at Rizan Aneiza. The Amrar section (Fig. 8f ) is located at a second extensional SW –NE trending fault, again rejuvenated during the Upper Cretaceous compression (Amrar fold belt sensu, Moustafa & Khalil 1990). Two similar sets of faults with vertical movement were also described by Moustafa (2010). The following sedimentological and stratigraphical observations suggest a structural control on Early Cretaceous deposition along normal faults of North Sinai. (1) The Upper Barremian to lower Upper Aptian (PS I –II) sections at the neighbouring Gebels Raghawi and Mansour (located c. 5 km apart) show significantly different sedimentary environments (Fig. 7). In the Mansour area, sedimentation took place under terrestrial
120
(2)
M. BACHMANN ET AL.
conditions, while more than 240 m of marly and limy sediments were deposited in marine environments at Gebel Raghawi. To interpret this record we either have to assume a primary, very steep, declining coast (with vertical differences of more than 200 m along a 5 km horizontal distance), or more probable, an active normal fault that resulted in higher subsidence rates in the northern areas. At the end of PS II (latest Early Aptian and earliest Late Aptian) water-depth at the Gebel Raghawi area increased significantly in contrast to still terrestrial environments at Gebel Mansour, and may thus reflect different subsidence owing to syndepositional tectonic activities along a normal fault. (Alternative interpretations of deposition and later erosion of marine sediments at Gebel Mansour seem unlikely because of missing reworking horizons). Our facies data hint on a SE – NW striking normal fault active during the Late Barremian–earliest Late Aptian (and reactivated in Late Cretaceous times), which caused the different depositional environments (Fig. 8f ). At Gebel Amrar, further to the NW, shallow marine Upper Barremian– lower Upper Aptian deposits indicate a swell, which may be correlated with another active normal fault belonging to the Amrar fold belt (Moustafa & Khalil 1990).
During PS III (early Late Aptian) terrestrial sedimentation occurred in the entire northern Sinai, which was accompanied by erosion in the southern sections (Fig. 8c). During PS IV (latest Aptian – Middle Albian) stepwise encroaching of the marine sediments from north–NW to south–SE indicates a slightly dipping ramp (Fig. 8d ). Uniform sediments at Raghawi and Mansour characterize the next stage (PS V), with some beds traceable for long distances (Fig. 8e). This means that the Raghawi –Mansour fault was active only until the end of the Early Aptian, possibly until the earliest Late Aptian. During the younger stages (PS III) erosion and deposition reduced the former elevation gradient along the normal faults, until more homogenous depositional environments became obvious during PS IV (Fig. 8). The sedimentological and stratigraphical data suggest extensional tectonical activity during the Barremian–Early Aptian, which is younger than described before (Garfunkel 1998; Moustafa & Khalil 1990; Moustafa et al. 1990). Only since the Late Aptian, subsequent to tectonical activity, a stepwise retrogradation took place and the facies development was controlled only by global sealevel change and supraregional tectonics.
The Galilee– Golan Heights structural development. The northern Israel Barremian –Albian structural development is much simpler as in North Sinai. During the Late Barremian, flooding of the terrestrial area created a slightly inclining ramp structure. Owing to different depositional rates, a landward-thinning wedge of sediments (described by Rosenfeld et al. 1998) filled the available accommodation space until a shallow, uniform platform developed during the Early Aptian (Bachmann & Hirsch 2006). Distally steepening of the ramp and a shelf break is suggested for the Carmel area, where rudist bioherms fringe the shelf-break and are closely allied to deeper marine sediments during the Albian period and deeper marine sediments were deposited nearby (Bein 1971; Sass & Bein 1982). This is confirmed by seismic profiles (Garfunkel 1998). In summary, we suggest a depositional regime, which was mainly controlled by sedimentation rates, sediment production, and sealevel changes without evident tectonic influence. Facies models for the Galilee –Golan Heights area include a post-depositional, Oligocene–recent sinistral strike–slip movement along the Jordan– Gulf of Aqaba line (Fig. 1b, e.g. Garfunkel & Ben-Avraham 1996; Flexer et al. 2005; Mart et al. 2005). The rate of sinistral displacement is controversially discussed, reaching from 10 km (Mart et al. 2005) to 110 km (Garfunkel & Ben-Avraham 1996; Gilat 2005). Our sections show no evidence for strong lateral displacement since displacement was parallel to facies belt boundaries.
Second-order sequences Second-order sequences were already discussed for the Galilee –Golan area and for the upper part of the North Sinai succession (Bachmann et al. 2003; Bachmann & Hirsch 2006). Comparing both platform settings allow us to recognize second-order sequences that have controlled a large part of the Levant Platform. We adopt three second-order sequences, formerly described from the Galilee –Golan Heights area and prefixed MC (mid-Cretaceous) EL (eastern Levant) and numbered (e.g. MCEL-1). For the larger extension of the analysed region, we transfer the prefix into MC (mid-Cretaceous) L (Levant) and add another sequence (Fig. 10).
MCL-1 In both areas, terrestrial sediments were flooded in the Late Barremian. For both areas, we document initial deepening and increasing accommodation during the Late Barremian –earliest Aptian, and at Galilee –Golan landward retrogradational patterns characterize the transgressive systems tract (TST).
Levant platform
AlSin8
K100 (101) SB MCL 4
SB 103
M. inflatum
H. dentatus
L. tardefurcata
Al 3 Al 2 Al 1
H. jacobi
Ap6
N. nolani
Late
Aptian Early
D. deshayesi D. weissi
AlSin1
TST
TST
SB 109.5
SB MCL 3
HST
HST
K90 (111) Late Aptian unconformity
ApSin3
SB 113
Op ApSin2
TST
TST K80 (117)
Op O.(.M.) parva
Pc-Ol
P. waagenoides C. sarasini I. giraudi H. feraudianus G. sartousi A. vandenheckii
ApSin1 Ap4
P. cormyi O.(.M.) lotzei
Pl
D. oglanlensis
Late
AlSin2
SB 107
P. melchioris
D. furcata
Barremian
K100 (106)
HST
SB
O.(.M.) parva O.(.M.) texana
E. subnodosocostatum
125
Al 4
D. mammillatum
115
120
Os O.(.M.) subconcava
Al 6 Al 5
AlSin7 AlSin6 AlSin5 AlSin4
Pl-Cd P. lenticularis C. decipiens
ApEl4 Ap3
ApEl1
Ap2 Ap1
BaEl2
Barr6
LST
SB MCL 2
HST
LST
(120.1) K70 (122.5)
HST
SB (125)
BaEl1
TST
Emergence
SB
III
E. loricatus
HST
Al 7
TST
SB (127)
OAE1a
K60 (126)
I
Middle Early
Albian
Early Cretaceous
110
Al 8
E. lautus
IV
105
Levant platform
arid
arid
humid
Europe Price, 1999
humid
Arabian Platform
AlSin9
Tectonics
fault movement/extensional tectonics (this work) Moustafa and Khalil, 1990
Late
Al 9
Israel
terrigenous input and ferruginous ooids orbitolinid beds
Al 10
Sinai
karstification at 3rd-order SBs tepees, stromatolithes
Al 11
Sinai Israel
Sharland Bachmann Bachmann modified from et al. 2004 et al. 2003 et al. 2003 Bachmann & Hirsch 2006 recalibrated by Bachmann & this work Haq & AlHirsch 2006 Qahtani 2005
V
Hardenbol et al.1998
Climate
Sinai: ramp-platform Israel: fringed platform
Global or Tethyal
121
Onlap curve, crisis & platform-types
Sinai: deltaic ramp Israel: carbonate platform
Bachmann & Hirsch 2006, 2010
nd
2 -order sequences
II
Gradstein et al. 2004
100
Sea level rd
3 -order sequences
Sinai: sub-basin Sinai: sub-basin Israel: siliciclastic ramp Israel: platform
Ammonites
Stage
Substage
Epoch
Chronostratigraphy
larger benthic foraminifer zones
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
Fig. 10. Chronostratigraphic chart summarizing the stratigraphic and sequence stratigraphic interpretation, as well as the climate and tectonic data, in comparison with data on the Arabian Plate and the regions around the Tethys.
These retrogradation patterns were retarded in North Sinai, when the coastline reached the Raghawi area by the activity of the Raghawi – Mansour fault (Figs 7 & 8), while westward retrogradation crossed the Golan Heights without barrier (Fig. 4). The TST, characterized by high siliciclastic input and ferruginous ooid, was deposited under more or less open marine locally high-energy conditions (Fig. 4). A major change to low-energy inner platform conditions in the basal Lower Aptian marks the maximum-flooding surface (MFS) at the Galilee –Golan Heights. The early highstand systems tract (HST) comprises limestones deposited under lagoonal to open-marine conditions and the late HST more protected environments with increasing siliciclastic input (Fig. 4). No former southward transgression is documented and the creation of accommodation space is clearly reduced. In North Sinai, a steady coast line and a decrease in terrigenous input characterize the HST with a depositional environment clearly changing to open marine, deeper-shelf environments. However, no reduction of accommodation is documented, before the late HST, because of subsidence owing to local fault activity. In both areas an exceptional
third-order deepening event is observed superimposing the second-order HST, which largely correlates with the platform drowning observed at the Oceanic Anoxic Event 1a (Weissert et al. 1998; Bachmann & Hirsch 2006).
MCL-2 The sequence boundary (SB) of the MCL-2 is the most prominent emergence horizon in the entire succession, indicated by terrestrial sandstones which eroded into the underlying sediments (Sinai only, Fig. 4). For North Sinai the age of the SB can be dated as being earliest Late Aptian (Figs 3 & 4). In Israel it is recognized only in the Golan Heights, where its dating is unsure. Similarly pronounced is the transgressive surface (TS), when retrogradation of the facies belts is marked by the renewed marine sedimentation in both areas. During the TST, both areas are marked by increasing accommodation, retrograde submergence of the platform, and siliciclastic input (reduced in the late TST) in dominantly protected environments. Slightly higher accumulation at the Golan Heights, compared to North Sinai, may result from earlier
122
M. BACHMANN ET AL.
onset of the marine deposition. While at North Sinai the TS is dated as early Late Aptian [C. (E.) subnodosocostatum zone], the age is not confirmed yet for the Golan Heights. The basal Upper Aptian MFS is documented in the Golan area by the deepest environment in the study area, indicated by a sudden reduction in terrigenous input and followed by HST carbonates. The mfs in the North Sinai is interpreted as lying at a surface of re-establishment of carbonate dominated sediments (Raghawi and Rizan Aneiza sections), which coincides with a decrease of the southward transgression, indicated by interfingering of high-energy shoals at Raghawi with sediments of the proximal delta at Gebel Mansour. The depositional rate of MCL-2 is slightly reduced compared to MCL-1, while the observed transgression to the south, especially at North Sinai, is much greater (Fig. 8g). However, observed depositional environments, which are generally characterized by shallower water-depth during MCL-2 indicates a general lower accommodation increase than before.
MCL-3 The Lower Albian SB MCL-3 is clearly defined at the Golan Heights (Ein Quniya section) and at Galilee (Har Ramin section) by a facies change from open marine to restricted environments (Fig. 4). At North Sinai, it is interpreted as lying below a distinct zone of delta progradation. The LST is indicated by terrigenous delta-influenced sediments in the northern distal sections and by coarse-grained sand dominated delta sediments in the proximal section M. The TSs are marked in both areas by deepening of the environment and by a clear trend to high-energy setting in the individual sections indicating increasing accommodation (Fig. 4). In North Sinai, prominent transgression during the TST results in submergence of the southern and eastern part of the Maghara area (section MgE, Fig. 8h). At Galilee the MFS is indicated by a significant change from open marine to protected environments under constant accommodation. At North Sinai, a change from consequent transgression to a succession with higher-frequency facies changes may result from reduced accommodation and mark the late Early Albian MFS. The HST in both areas is characterized by repeated siliciclastic input, but also by the development of rudist biostromes (North Sinai) or rudist debris fearing sediments (Galilee –Golan Heights).
MCL-4 The SB MCL-4 is marked by emergence in North Sinai and by facies change at Galilee (Fig. 4). The
Late Albian age confirmed at North Sinai may correlate with Galilee, where an accurate dating of the SB is not possible. In both areas, the LST contains few beds of intertidal dolomite. While only few intercalations of protected and open platform sediments mark the TST at Galilee, a succession with shoals, protected lagoons, and rudist biostromes dominate the TST at North Sinai. In the latter area the maximum flooding is characterized by the establishment of tidal flat sediments on large parts of the platform in the uppermost Upper Aptian. The uppermost part of the sequence is present in the Mansour north sections only; here the subsequent SB is of Early Cenomanian age.
Discussion Platform development – geometry and facies development as a consequence of tectonics, siliciclastic input, production, and sea-level change. The five platform-stages (I –V) reflect different depositional environments, accumulation patterns, and platform geometries (Fig. 9). We observe significant variations in one or more of the following parameters: tectonical activity, siliciclastic input (driven by climate and tectonics), carbonate production, and second-order sea-level. To chart the depositional models for the Levant Platform stages (Fig. 9) published subsurface data were incorporated, spanning the region between Galilee – Golan Heights and North Sinai. PS I: Late Barremian–earliest Early Aptian. In both areas siliciclastic influenced carbonate sediments characterize the near shore environments, rich in ferruginous ooids and quartz (Fig. 9a). Marly sediments, with upward increasing carbonate content, are described from the central Israel hinge belt. A shallow water depositional environment was defined on the base of ostracod assemblages, with intercalated fresh-water signals, observed in central and northern Israel (Rosenfeld et al. 1998). A strongly increasing sea level (TST MCL-1) resulted in transgression. Different subsidence, basic geometric regimes, and tectonical influence resulted in different depositional regimes in Sinai and Israel. At North Sinai, low-energy sub-basins developed between active normal faults, which caused enhanced subsidence rates, a stable coast line, and phases of higher siliciclastic input owing to higher-frequency sea-level changes. Sedimentological and tectonical data additionally indicate the existence of a swell in the northernmost part of the Sinai (Gebel Ambra region, Fig. 8) causing a lowenergy depositional regime south of the swell with shallow and deeper lagoonal environments interrupted by sporadical coastal progradation. Those
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
features result in small-scaled variation of the sedimentation rates, with high sediment accumulation in front of the normal fault at Gebel Raghawi and low in the area south. We suggest a shelf break situated only a few kilometres north of the studied area, possibly linked with normal faults further north. This fits with deeper marine marls characterizing the offshore area only 20 km north of the studied sections at Rizan Aneiza (Martinotti 1993) and the occurrence of phylloceratid ammonites in the Lower Aptian sediments probably suggesting that deeper marine conditions were present in an area close to the studied ramp sediments (Lehmann et al. 2007). However, an odd taphonomic history has been suggested for this group of ammonites, with surfacing and floating of shells first that is followed by resinking by cameral puncture, and therefore a substantiation of an autochthonous origin can be complex (Maeda & Seilacher 1996). In northern Israel– Galilee, sediments form an eastward thinning wedge (Rosenfeld et al. 1998; Rosenfeld & Hirsch 2005), with more restricted environments in the eastern part (Fig. 9a). There is no sign of active faulting and sea-level rise result in gradual flooding of the former terrestrial sediments, with a coastline moving from the Galilee to east of the Golan Heights (Bachmann & Hirsch 2006). At Galilee a thick succession of high-energy grainstones indicate increasing accommodation. Gradual transgression and the landward increasing protection of the environments suggest a shallow ramp geometry with a huge high-energy facies belt. During the entire succession a continuous input of detrital components and ferruginous ooids indicates a humid weathering regime, confirmed by the intercalations of freshwater ostracod assemblages (Rosenfeld et al. 1998). PS II: Early Aptian –earliest Late Albian. During the Early Aptian (HST MCL-1) both platform areas pass through significant changes; coarsegrained siliciclastic input is reduced, and thus carbonate dominates the submerged part of the platform (Fig. 9b). In the Galilee –Golan Heights and North Sinai, low-energy environments rich in calcareous algae and bioclasts occur. In the Galilee –Golan Heights, only slight lateral facies variations in large parts of the proximal platform indicate a large extension of the low-energy openplatform facies belt with similar water depths. Thus, flat-topped platform geometry was suggested to form homogenous shallow platform environments at the beginning of PS II. The seaward margin of this platform is not exposed in the studied area and rudist fringing reefs found in the Carmel area (30 km west of Galilee sections) were documented from younger successions only (Bein 1971; Sass & Bein 1982).
123
In North Sinai no change of the depositional geometry occurs compared to PS I. The Mansour– Raghawi fault and the Gebel Amrar swell controlled the depositional processes within the small subbasin. In between Sinai and Galilee the subsurface descriptions indicate a transitional zone, without a major shelf break (Rosenfeld et al. 1998). Deepening in the late Early Aptian starts with a succession of orbitolinid beds, followed by openmarine, deeper platform, marly sediments in both areas (Fig. 9c). This deepening causes a shorttermed southward transgression exceeding the Mansour –Raghawi fault until the southwestern Maghara area and may reach into southern Negev (southern Israel), where short-termed marine transgressions is known from the Lower Aptian (Deshayesites deshayesi and D. forbesi ammonite zones) dated by fossils and radiometric ages of an overlying basalt (Gvirtzman et al. 1996). PS III: Earliest Late Aptian. This short-termed platform stage marks a major break in the North Sinai depositional geometry, but is characterized by shorttermed emergence in the Galilee –Golan Heights only (Fig. 9d ). At North Sinai, the fault controlled differential subsidence patterns terminated and the sub-basins were filled with near shore and erosional deposits. Moreover, terrestrial (Raghawi), marine sandstones (Rizan Aneiza), and dolomites (Amrar) correlate with a major second-order lowstand. Regression is also marked by a northward movement of the coastline to an area in between Raghawi and Amrar (Fig. 9d ). Erosional unconformities observed at Gebel Raghawi are possibly owing to denudation of the relief around the normal fault (Fig. 8) and may have shaped the initial stage of a low dipping ramp geometry (observed in the subsequent platform stages). Fluvial sandstones deposited at the Golan Heights indicate a major regression in the northeastern part of the study area. However, erosion was not observed and the geometry persisted homogenous. Altogether PS III presents a stage characterized by a widely emerged platform with distinct regression. PS IV: Late Aptian–Middle Albian. The Late Aptian –Middle Albian PS IV comprises secondorder sequence MCL-2 and the LST and TST of MCL-3, in both areas characterized by temporarily detrital influenced carbonate production. The platform was widely submerged and a gradual transgression marked the entire area (Fig. 9e). At North Sinai, the transgression submerged the Rizan Aneiza, Amrar, Raghawi, and Mansour regions during Late Aptian and the southeastern part of the Maghara area during the Lower Albian (Fig. 7). First marine sediments occur 20 km south of Maghara at the end of PS IV (Bachmann et al.
124
M. BACHMANN ET AL.
2003). A broad delta system developed, which strongly influenced the depositional processes at Mansour and Maghara SE area interfingering with marine marl and limestones around Gebel Raghawi, while Rizan Aneiza was only slightly influenced by siliciclastic deltaics (Fig. 8g). A north dipping ramp geometry is indicated by increasing thickness of open marine sediments with rudist patch reef, oolithic shoals, and lagoonal sediments towards the north. This gradient was considerably developed at the end of PS IV, when mid-ramp environments at Rizan Aneiza gradually passed into lagoons at Raghawi. The dipping direction changed from NW during PS I– III to N (NNW) during PS IV, triggered by the deltaic input from the southern direction. The ramp can be subdivided in three main facies belts: a southern deltaic depositional facies (Mansour and Maghara SE), a shallow protected subtidal facies (Raghawi), and an open subtidal facies with rudist patch reefs in the north (Rizan Aneiza, Fig. 8g). Those facies belts prograded and retrograded following higher frequency sea-level changes. The extensive siliciclastic input in the delta realm caused higher sedimentation rates in the proximal parts of ramp. Only in the upper PS IV did the depocentre move towards north– NW, when the proximal platform was emerged during the third-order LSTs of the early second-order HST of MCL-3 and sedimentation rates increased in the more distal parts of the platform. At Galilee –Golan Heights the entire platform was submerged since the Late Aptian. Marine sedimentation in the adjacent Jordan is not older than latest Albian (e.g. Schulze et al. 2004), which indicates that the coastline was close to the Golan Heights. Rudist fringing reefs were established at the shelf hinge (Bein 1971; Sass & Bein 1982). Ross (1992) pointed out that the back reef gravel may have formed a major hydrologic barrier, while Bein & Weiler (1976) documented the lower slope nature of muddy sediments deposited in close vicinity to the fringing reefs at the Carmel area. Homogenous flat-topped platform geometry and a strong decrease in hydrologic energy in the proximal platform areas resulted in a shallow lagoonal inner platform environment that varied little over a large area. Sedimentation rates were very similar in the inner platform area, comprising Galilee and the Golan Heights. In the area in between Sinai and northern Israel, a facies established similar to the northern region (Rosenfeld & Hirsch 2005). PS V: Late Albian. PS V is characterized by gradual change of the platform geometry at Sinai and constant filling of the accommodation at Galilee and Golan Heights (Fig. 9f ). PS V comprises the last
herein studied second-order sequence. In both areas sediments of the protected inner platform area with strong reduced siliciclastic amounts are prevailing. Crucial for the development of North Sinai was the dislocation of the rising sea-level. The sedimentation rates increased in the more distal sections resulting in filling of distal accommodation. A very homogenous flat-topped platform developed. During the Late Albian this platform extended southwardly to an area 100 km south of Maghara (Bachmann & Kuss 1998). Within this large extended platform, a succession of rudist biostromes, alternating with protected lagoonal and algal-laminate dominated tidal flats, were deposited resulting from higher frequency sea-level changes. Deposition of grainstones and oolithic facies was restricted to short intervals at Gebel Raghawi, indicating the low-energy dominance during that interval and suggesting a shelf break in an area further north. In Galilee, dolomites representing tidal flats and lagoon prevail, which are also described from the Golan Heights and from Judea in between Galilee and North Sinai (Rosenfeld & Hirsch 2005). The Upper Albian distribution of facies belts indicates a large extended homogenous, very shallow, platform without major changes in facies reaching from northern Israel and the Golan Heights to North Sinai. A continuously increasing sea-level resulted in high accumulation rates and southward transgression. However, platform growth kept up with the sea-level rise and shallow-water, and intertidal deposition continued.
External controlling factors As external factors controlling the sedimentation patterns on the carbonate platform we consider climate and global sea-level changes. Imprints of those changes should occur at both platform areas studied, independent from different tectonical regimes, and are comparable to similar changes observed on other Tethyan platforms (Fig. 10). Second-order sea-level change and its influence on the formation of the platform stages. Second-order sea-level changes highly influenced the Levant platform development by controlling the accommodation and the centres of deposition. During the Late Barremian–Early Aptian a second-order TST created the accommodation to start up the marine sedimentation. Reduced accommodation during second-order HST resulted in distal dislocation of the depocentre and thus controlled the replacement of the ramp by flat topped platform architecture in Galilee –Golan Heights. Erosion during the early Late Aptian LST (SB MCL-2) influenced the ramp
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
geometry in North Sinai, while reduced accommodation during MCL-3 HST and MCL-4 LST initially induced the distal dislocation of the depocentre and the change from ramp to platform architecture in North Sinai. A comparison of global sequences with that of the Arabian plate origin is shown in Figure 10. Data are constrained by the stratigraphical frame slightly clouded by the use of different time scales in the several studies. The southern Arabian plate exhibit similar features concerning the widespread continental siliciclastic sediments flooded during the Late Barremian, and marine sedimentation during Aptian –Albian times (Haq & Al-Qahtani 2005; Sharland et al. 2001, Fig. 10). The early Late Aptian sequence boundary (MCL-2) coincides with prominent sequence boundaries observed at various localities of the Arabian plate (Van Buchem et al. 2002; Gre´selle & Pittet 2005; Haq & Al-Qahtani 2005). A second match with the Arabian plate feature is observed for the Late Aptian SB MCL-4, while SB MCL-3 is younger than that observed on the southern Arabian plate (Gre´selle & Pittet 2005; Haq & Al-Qahtani 2005). Some further second-order SBs observed on the southern Arabian plate by Gre´selle & Pittet (2005) or Haq & Al-Qahtani (2005) coincides with SBs regarded as a higher frequency in origin in North Sinai, because they are not clearly indicated at Galilee –Golan Heights. Important differences to the southern Arabian plate were observed in the Upper Aptian succession. While a Late Aptian unconformity is documented from several platforms (Van Buchem et al. 2002; Haq & Al-Qahtani 2005), continuous marine sedimentation took place in the Levant Platform at the northern edge of the Arabian plate. Comparing the sea-level changes with patterns supposed to be global, similarities are much less obvious. However, the transgression characterizing the Upper Aptian– Albian Levant succession coincides with the long-term global transgression pointed out by Hardenbol et al. (1998), while the lower part of the succession falls into a global sealevel minimum. This agrees with transgressions of smaller amplitude during that time interval on the Levant Platform. Additionally SB MCL-2 possibly correlates with Ap 4 defined by Hardenbol et al. (1998) and with a few time-equivalent exceptional SBs in the northern Tethyan realm, like the Alpine mountains (e.g. Roter Sattel, Strasser et al. 2001), while other areas are characterized by transgression or high stand during the same time. Altogether, the sea-level history of the Levant Platform reflects the Late Aptian– Albian global long-term transgression and the second-order sealevel changes are highly correlated with that one observed on the Arabian plate.
125
Climate. The composition of the siliciclastic input and the formation of ferruginous ooids allow the interpretation of general climate conditions in the study area (Fig. 10). The Upper Barremian– Albian succession comprises two intervals with coarse grained siliciclastic input and occurrence of ferruginous ooids, one reaches from the Late Barremian to the earliest Aptian (PS I) and the second comprises the Late Aptian until the mid-Albian (PS IV). Within these parts of the succession marly facies dominates and carbonatic samples often show a marly matrix. These intervals were interpreted as indicating humid conditions (Bachmann & Hirsch 2006; Bachmann & Kuss 1998) in concordance with Mu¨cke (2000), who demonstrated that lateritic weathering in the hinterland was producing the protoliths for Upper Cretaceous oolithic ironstones. Those interpretations are underlined by the common occurrence of plant remains in the succession of PS I at Gebel Raghawi and by the formation of the pronounced delta system and carstification horizons in North Sinai during PS IV. Orbitolinid-rich facies types, associated with abundant calcareous algae and echinoderms and argillaceous muddy limestones, are common in PS IV. Those facies types are typically interpreted as reflecting mesotrophic conditions (Van Buchem et al. 2001; Pittet et al. 2002), which may reflect nutrient input owing to weathering and humid conditions. Two intervals are characterized by reduced siliciclastic input and the absence of the ferruginous ooids (Early Aptian/PS II and the Late Albian/PS V). Such a reduced siliciclastic input can result from reduced sediment supply owing to less intensive weathering or from an increasing distance of the delivery area. For PS II, the steady coast line between Mansour and Raghawi during lower PS II argues for a climate origin of the reduction of detrital input. The Late Albian (PS V) of reduced siliciclastic input is accompanied by an increase of rudist and miliolid dominated pure limestones in North Sinai. Those facies types are commonly interpreted as reflecting oligotrophic conditions (Van Buchem et al. 2002), which fit well with the observed transgressional, less humid system. Changing third-order low stand features accompany humid and less humid phases. LSTs are characterized by detrital input during the humid phases and by emergence and dolomitization during the less humid periods (Bachmann & Kuss 1998). Our data fit with the general ideas of Price (1999) and Hillga¨rtner et al. (2003) for the Oman, who indicated a small trend to humidity in the Late Barremian –Early Aptian. However, the position of the Levant Platform near the equator may have triggered humidity in northeastern Africa, while more
126
M. BACHMANN ET AL.
arid conditions influence the northern Tethyan margin during the Barremian–earliest Aptian. Tectonics. Tectonical activity played an important role, especially during the initial phase of the platform development in North Sinai (Fig. 10). Our data suggest that during the Late Barremian–Early Aptian, extensional normal faults influenced the North Sinai environment and result in the formation of sub-basins, while at the same time a slightly dipping ramp, without fault segmentation, is recorded in northern Israel. The Late Aptian – Albian distribution of the facies belts in North Sinai indicates a uniform deepening in a NW direction and confirms that sedimentation geometry was orientated parallel to SW –NE striking normal faults pointed out by Moustafa & Khalil (1990). Pre-deposition extensional tectonics creating normal faults are described from the Levant margin offshore Israel as well as from northern Sinai until the mid-Jurassic (Moustafa & Khalil 1990; Garfunkel 2004; Gardosh & Druckman 2006). In North Sinai, a set of extensional basins was created from which the Maghara area represents the northernmost point that was located onshore (Moustafa 2010). Extensional tectonics at the North African –Arabian continental margin are interpreted as reflecting divergent movement between the Afro-Arabian and the Eurasian plate, which led to the Mesozoic to Middle Jurassic opening of the Neotethys (Stampfli et al. 2001). Our data indicate that extensional faults were active until the late Early Aptian in North Sinai and may represent a Lower Cretaceous syn-rift stage according Guiraud et al. (2005), who observed continental rifts along the African –Arabian Tethyan margin in the Late Berriasian –earliest Aptian.
Conclusions During the Late Barremian–Albian, the southern Levant Platform was studied in the two regions: northern Israel (Galilee –Golan Heights) and North Sinai. Facies, stratigraphy and stable isotopes of several sections were studied, to reconstruct transects across the shallow shelf. In conclusion we establish depositional models and discuss the controlling factors for the shallow water deposition. We combined biostratigraphy on the base of benthic foraminifers, mainly orbitolinids, and ammonites with stable isotope data, which allow us to date the shallow water strata and all controlling factors in detail. The depositional architecture was controlled by local tectonics, climate, and second-order sea-level changes affecting the sedimentation patterns. Four secondorder sequence boundaries were identified in the Levant area (MCL-1– MCL-4). They partly
correlate with those observed on the Arabian plate (Haq & Al-Qahtani 2005; Sharland et al. 2001), suggesting a regional control. Until the late Early Aptian, the North Sinai was influenced by normal fault development, while the Galilee area– Golan Heights exhibit continuous sedimentation without tectonical influence. Both regional transects reveal five platform stages (PS I –V) that differs with respect to platform architecture, siliciclastic input, and response to sea-level changes. PS I (Late Barremian–earliest Aptian). The Upper Barremian marine sediments transgraded on terrestrial deposits. In northern Israel a homogenous ramp existed, while North Sinai was subdivided in SW –NE striking sub-basins, marginally bounded by active normal faults. Open marine high-energy sedimentation and continuous transgression characterized the Galilee– Golan Heights area during the TST of MCL-1, while near coast siliciclastic and protected lagoonal carbonates alternate with deeper marine sediments in North Sinai. PS II (Early Aptian –earliest Late Aptian). During the HST of MCL-1, continuous filling of the accommodation resulted in shallow marine protected facies belts marking flat-topped platform architecture at northern Israel. Owing to higher subsidence around normal faults, sub-basin development kept on in North Sinai. The sedimentation was characterized by protected– partly deeper– subtidal environment within these sub-basins, separated from eachother by shallower swells. Significant reduction of detrital input in both areas may result from changing weathering regimes. PS III (early Late Aptian). During the LST of MCL-2 the tectonical activity terminated resulting in reorganization of the North Sinai platform. The former fault-controlled sub-basins became inactive and were covered by a shallow ramp architecture that controlled the depositional processes of the western Levant Platform. Emergence was evidenced from North Sinai to northern Israel indicating an extended platform system. PS IV (Late Aptian– Middle Albian). A stable platform architecture, marked by a homoclinal ramp in North Sinai and a flat-topped platform in Galilee –Golan Heights, was sandwiched between transgressive surface MCL-2 and the end of TST of MCL-3. In the North Sinai a broad delta system with high siliciclastic input was deposited, interfingering in the north with carbonate ramp deposits. High accumulation rates on parts of the ramp resulted in slight changes of the dipping direction. Simultaneously,
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION
a less siliciclastic input and inner platform sediments are characterized in the Galilee –Golan Heights region. PS V (Late Albian and younger). During HST of MCL-3 and TST of MCL-4 pure limestones and dolomites were accommodated in North Sinai, during the ongoing transgression a flat topped platform was established here as well as in northern Israel. Our data indicate that extensional faults controlled sedimentation until the late Early Aptian, which is significantly younger than observed before (Moustafa & Khalil 1990; Garfunkel 2004; Gardosh & Druckman 2006). They may represent a syn-rift stage according to Guiraud et al. (2005). Concerning the climate history, we point out that the larger input of detrital sediment, including ferruginous ooids, indicates an increased humidity owing to an accelerated continental weathering, particularly in the Late Barremian –Early Aptian. We are indebted to F. Hirsch (Jerusalem) and R. Stein (Bremerhaven) for discussions on the Lower Cretaceous of Israel, respectively geochemical data. We are especially thankful for J. Thielemann (Bremen) and M. Heldt (Hannover) for assistance in the field, many successful discussions, and providing valuable data of their theses. A. R. Moustafa (Cairo) helped with discussing the tectonical background of the working-areas. Technical support by M. Krogmann (Bremen) is appreciated for preparing a part of the ammonite fauna. We thank the two reviewers M. Simmons and R. Scott for helpful and instructive discussions and comments. We appreciate financial support by the German Science Foundation and the VW foundation.
References Aboul Ela, N. M., Abdel Gawad, G. L. & Aly, M. F. 1991. Albian fauna of Gabal Manzour, Maghara area, north Sinai, Egypt. Journal of African Earth Sciences, 13, 201–220. Abu-Zied, R. H. 2007. Palaeoenvironmental significance of Early Cretaceous foraminifera from northern Sinai, Egypt. Cretaceous Research, 28, 765–784. Abu-Zied, R. H. 2008. Lithostratigraphy and biostratigraphy of some Lower Cretaceous outcrops from Northern Sinai, Egypt. Cretaceous Research, 29, 603–624. Aguado, R. & Company, M. 1997. Biostratigraphic events at the Barremian/Aptian boundary in the Betic Cordillera, southern Spain. Cretaceous Research, 18, 309– 329. Arkin, Y., Bartov, Y. & Goldberg, M. 1975. The Geology of Rizan Aneiza, Northern Sinai. Internal Report, No. MM 2/75, 1 –19. Arnaud-Vanneau, A. 1998. (Coord. L. Cret.). Larger benthic foraminifera. In: Hardenbol, J., Jacquin, T., Farley, M. B., de Graciansky, P.-C. & Vail, P. (eds) Cretaceous Biochronostratigraphy, chart 5.
127
Arnaud-Vanneau, A. & Arnaud, H. 1990. Hauterivian to Lower Aptian carbonate shelf sedimentation and sequence stratigraphy in the Jura and northern Subalpine chains (southeastern France and Swiss Jura). Special Publication, International Association of Sedimentologists, 9, 203–233. Arnaud, H., Arnaud-Vanneau, A. et al. 1998. Re´partition stratigraphique des orbitolinide´s de la plateforme urgonienne subalpine et jurassienne (SE de la France). Ge´ologie Alpine, 74, 3–89. Askalany, M. M. & Abu-Zeid, K. A. 1994. Contribution to the geology of El-Amrar and Rizan Aneiza areas, Northern Sinai (Egypt). Bulletin of the Faculty of Sciences Assiut University, 23, 81– 99. Bachmann, M., Bassiouni, M. A. A. & Kuss, J. 2003. Timing of mid-Cretaceous carbonate platform depositional cycles, northern Sinai, Egypt. Palaeogeography, Palaeoclimatology, Palaeoecology, 200, 131– 162. Bachmann, M. & Hirsch, F. 2006. Lower Cretaceous carbonate platform of the eastern Levant (Galilee and the Golan Heights) – stratigraphy and secondorder sea-level change. Cretaceous Research, 27, 487– 512. Bachmann, M. & Kuss, J. 1998. The Middle Cretaceous carbonate ramp of the northern Sinai: sequence stratigraphy and facies distribution. In: Wright, V. P. & Burchette, T. P. (eds) Carbonate Ramps. Geological Society, London, Special Publications, 149, 253–280. Bartov, Y. & Steinitz, G. 1977. The Judea and Mount Scopus Groups in the Negev and Sinai with Trend Surface Analysis of the thickness data. Israel Journal of Earth Science, 28, 119– 148. Bein, A. 1971. Rudist reef complexes (Albian to Cenomanian) in the Carmel and the coastal plain, Israel. Geological Survey of Israel Report. 86. Bein, A. & Weiler, Y. 1976. The Cretaceous Talme Yafe formation: a contour current shaped sedimentary prism of calcareous detritus at the continental margin of the Arabian Craton. Sedimentology, 23, 511– 532. Bernaus, J. M., Arnaud-Vanneau, A. & Caus, E. 2002. Stratigraphic distribution of Valanginian-Early Aptian shallow-water benthic foraminifera and algae, and depositional sequences of a carbonate platform in a tectonically-controlled basin: the Organya` Basin, Pyrenees, Spain. Cretaceous Research, 23, 25– 36. Bernaus, J. M., Josep, M., Arnaud-Vanneau, A. & Caus, E. 2003. Carbonate platform sequence stratigraphy in a rapidly subsiding area; the late Barremianearly Aptian of the Organya Basin, Spanish Pyrenees. Sedimentary Geology, 159, 177– 201. Braun, M. & Hirsch, F. 1994. Mid Cretaceous (Albian– Cenomanian) carbonate platforms in Israel. Cuadernos de Geologı´a Iberı´a, 18, 59– 81. Casey, R. 1962. A monograph of the Ammonoidea of the Lower Greensand. Part IV. Palaeontographical Society Monographs, 116, 217–288. Casey, R., Bayliss, H. M. & Simpson, M. 1998. Observations on the lithostratigraphy and ammonite succession of the Aptian (Lower Cretaceous) Lower Greensand of Chale Bay, Isle of Wight, UK. Cretaceous Research, 19, 511– 535.
128
M. BACHMANN ET AL.
Castro, J. M., Company, M., de Gea, G. A. & Aguado, R. 2001. Biostratigraphy of the Aptian-Middle Cenomanian platform to basin domain in the Prebetic Zone of Alicante, SE Spain: calibration between shallow water nectonic and pelagic scales. Cretaceous Research, 22, 145–156. Clavel, B., Charollais, J., Schroeder, R. & Busnardo, R. 1995. Re´flexions sur la biostratigraphie du Cre´tace´ infe´rieur et sur sa comple´mentarite´ avec l’analyse se´quentielle: exemple de l’Urgonien jurassien et subalpin. Bulletin de la Societe´ Ge´ologique de France, Huitieme Series, 166, 663–680. Conrad, M. A., Schroeder, R., Clavel, B., Charollais, J., Busnardo, R., Cherchi, A. & Decrouez, D. 2004. Dating the Lower Cretaceous in the Organya` section (Catalan Pyrenees, NE Spain): a reinterpretation. Cretaceous Research, 25, 35–41. De Gea, G. A., Castro, J. M., Aguado, R., Ruiz, P. A. & Company, M. 2003. Lower Aptian carbon isotope stratigraphy from distal carbonate shelf setting: the Cau section, Prebetic zone, SE Spain. Palaeogeography, Palaeoclimatology, Palaeoecology, 200, 207–219. Delanoy, G. 1997. Biostratigraphie des faunes d’Ammonites a` la limite Barre´mien-Aptien dans la re´gion d’Angles-Barreˆme-Castellane. Etude particulie`re de la famille des Heteroceratina Spath, 1922 (Ancyloceratina, Ammonoidea). Annales du Muse´um d’Histoire Naturelle de Nice, 12, 1 –270. Delanoy, G. & Ebbo, L. 2000. Une nouvelle espe`ce d’Heteroceras: H. mascarelli sp. nov. (Ancyloceratina, Ammonoidea) dans le Barre´mien du Sud-Est de la France. Annales du Muse´um d’histoire Naturelle de Nice, 15, 1– 17. El-Araby, A.-M. 1999. Facies analysis and sequence stratigraphy of the Late Aptian–Albian Risan Aneiza Foramtion in northern Sinai, Egypt. Egyptian Journal of Geology, 43, 151– 180. Eliezri, I. Z. 1965. The Geology of the Beit-Jann region (Galilee, Israel). Israel Journal of Earth Science, 14, 51–66. Flexer, A., Hirsch, F. & Hall, S. H. 2005. Tectonic Evolution of Israel. In: Hall, J. K. et al. (eds) Geological Framework of the Levant. Volume II: The Levantine Basin and Israel. Jerusalem: Historical Productions–Hall, 523– 538. Fo¨llmi, K. B., Godet, A., Bodin, S. & Linder, P. 2006. Interactions between environmental change and shallow water carbonate buildup along the northern Tethyan margin and their impact on the Early Cretaceous carbon isotope record. Paleoceanography, 21, 1 –16. Gale, A. S., Kennedy, W. J., Burnett, J. A., Caron, M. & Kidd, B. E. 1996. The Late Albian to Early Cenomanian succession at Mont Risou near Rosans (Droˆme, SE France): an integrated study (ammonites, inoceramids, planktonic foraminifera, nannofossils, oxygen and carbon isotopes). Cretaceous Research, 17, 515– 606. Gardosh, M. A. & Druckman, Y. 2006. Seismic stratigraphy, structure and tectonic evolution of the Levantine Basin, offshore Israel. In: Robertson, A. H. F. & Mountrakis, D. (eds) Tectonic Development of the Eastern Mediterranean Region. Geological Society, London, Special Publications, 260, 201–227.
Garcı´a-Monde´jar, J., Owen, H. G., Raisossadat, S. N., Milla´n, M. I. & Ferna´ndez-Mendiola, M. A. 2009. The Early Aptian of Aralar (northern Spain): stratigraphy, sedimentology, ammonite biozonation, and OAE1. Cretaceous Research, 30, 434– 464. Garfunkel, Z. 1998. Constraints on the origin and history of the Eastern Mediterranean basin. Tectonophysics, 298, 5 –35. Garfunkel, Z. 2004. Origin of the Eastern Mediterranean basin: a reevaluation. Tectonophysics, 391, 11– 34. Garfunkel, Z. & Ben-Avraham, Z. 1996. The structure of the Dead Sea basin. Tectonophysics, 266, 155–176. Gilat, A. 2005. Strike-slip faulting west of the Dead Sea Escarpment. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant. Volume II: The Levantine Basin and Israel. Jerusalem: Historical Productions-Hall, 515– 522. Godet, A., Bodin, S. et al. 2006. Evolution of the marine stable carbon-isotope record during the early Cretaceous: a focus on the late Hauterivian and Barremian in the Tethyan realm. Earth and Planetary Science Letters, 242, 254– 271. Gre´selle, B. & Pittet, B. 2005. Fringing carbonate platforms at the Arabian Plate margin in northern Oman during the Late Aptian–Middle Albian: evidence for high-amplitude sea-level changes. Sedimentary Geology, 175, 367 –390. Guiraud, R., Bosworth, W., Thierry, J. & Delplanque, A. 2005. Phanerozoic geological evolution of Northern and Central Africa: an overview. Journal of African Earth Sciences, 43, 83– 143. Gvirtzman, G., Weissbrod, T., Baer, G. & Brenner, G. J. 1996. The age of the Aptian stage and its magnetic events: new Ar– Ar ages and palaeomagnetic data from the Negev, Israel. Cretaceous Research, 17, 293– 310. Haq, B. U. & Al-Qahtani, A. M. 2005. Phanerozoic cycles of sea-level change on the Arabian Platform. GeoArabia, 10, 127 –160. Hardenbol, J., Jacquin, T., Farley, M. B., de Graciansky, P. C. & Vail, P. R. 1998a. Mesozoic and Cenozoic sequence chronostratigraphic framework of European basins. In: de Graciansky, P. C., Hardenbol, J., Jaquin, T. & Vail, P. R. (eds) Mesozoic and Cenozoic Sequence Stratigraphy of European Basins. Society for Sedimentary Geology (SEPM), Tulsa, 3– 13. Hardenbol, J., Thiery, J., Farley, M. B., Thierry, J., de Graciansky, P.-C. & Vail, P. R. 1998b. Cretaceous sequence chronostratigraphy – Mesozoic and Cenozoic sequence chronostratigraphical framework of European basins. In: de Graciansky, P.-C., Hardenbol, J., Jaquin, T. & Vail, P. R. (eds) Mesozoic and Cenozoic Sequence Stratigraphy of European Basins. Society for Sedimentary Geology (SEPM), Tulsa, Chart 5. Heimhofer, U., Hochuli, P. A., Herrle, J. O., Andersen, N. & Weissert, H. 2004. Absence of major vegetation and palaeoatmospheric pCO2 changes associated with oceanic anoxic event 1a (early Aptian, SE France). Earth and Planetary Science Letters, 223, 303– 318. Heldt, M., Bachmann, M. & Lehmann, J. 2008. Microfacies, biostratigraphy, and geochemistry of the
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION hemipelagic Barremian–Aptian in north-central Tunisia: Influence of the OAE1a on the southern Tethys margin. Palaeogeography, Palaeoclimatology, Palaeoecology, 261, 246 –260. Hillga¨rtner, H., Van Buchem, F. S., Gaumet, F., Razin, P., Pittet, B., Gro¨tsch, J. & Droste, H. 2003. The Barremian– Aptian evolution of the eastern Arabian carbonate platform margin (northern Oman). Journal of Sedimentary Research, 73, 756 –773. Hirsch, F. 1996. Geology of the southeastern slopes of Mount Hermon. Current Research – Geological Survey of Israel, 10, 22– 27. Hirsch, F., Flexer, A., Rosenfeld, A. & Yellin-Dror, A. 1995. Palinspastic and crustal setting of the eastern Mediterranean. Journal of Petroleum Geology, 18, 149–170. Husinec, A. 2001. Palorbitolina lenticularis from the Northern Adriatic Region: paleogeographical and evolutionary implications. Journal of Foraminiferal Research, 31, 287– 293. Immenhauser, A., Hillga¨rtner, H. & Van Bentum, E. 2005. Microbial-foraminiferal episodes in the Early Aptian of the southern Tethyan margin: ecological significance and possible relation to oceanic anoxic event 1a. Sedimentology, 52, 77–99. Jenkyns, H. C. & Wilson, P. A. 1999. Stratigraphy, paleoceanography, and evolution of Cretaceous Pacific Guyots: relicts from a greenhouse earth. American Journal of Science, 299, 341– 392. Keeley, M. L. 1994. Phanerozoic evolution of the basins of northern Egypt and adjacent areas. Geologische Rundschau, 83, 728– 742. Kemper, E. 1995. Die Entfaltung der Ammoniten und die Meeresverbindungen im borealen Unter- und MittelApt. In: Kemper, E. (ed.) Geologisches Jahrbuch. Hannover: Schweizerbart’sche Verlagsbuchhandlung, Na¨gele & Obermiller, 171– 199. Kotetishvili, Z. V. 1970. Stratigraphy and fauna of Colchidites and adjacent horizons of western Georgia [in Russian]. Akademija Nauk Grusinskoj SSR Trudy, 25, 5– 116. Kuss, J. & Bachmann, M. 1996. Cretaceous paleogeography of the Sinai Peninsula and neighbouring areas. Comptes Rendus de l’Academie des Sciences, Serie II. Sciences de la Terre et des Planetes, 322, 915–933. Lehmann, J., Heldt, M., Bachmann, M. & Negra, M. E. H. 2007. Cephalopodenfaunen aus dem Apt ¨ gypten: Vergleich (Unterkreide) von Tunesien und A von Faunen, Fazies und Pala¨obiogeographie. In: Elicki, O. & Schneider, J. W. (eds) Wissenschaftliche Mitteilungen. Technische Universita¨t Bergakademie, Freiberg, 78– 79. Lillo Bevia, J. 1975. Sobre algunos Hoplitidos del Creta´cico inferior del sur de Alicante. Boletin de la Real Sociedad Espan˜ola de Historia Natural, Seccion Geologica, 73, 81– 101. Luciani, V., Cobianchi, M. & Jenkyns, H. C. 2001. Biotic and geochemical response to anoxic events: the Aptian pelagic succession of the Gargano Promontory (southern Italy). Geological Magazine, 138, 277–298. Luciani, V., Cobianchi, M. & Lupi, C. 2006. Regional record of a global oceanic anoxic event: OAE1a on
129
the Apulia Platform margin, Gargano Promontory, southern Italy. Cretaceous Research, 27, 754 –772. Maeda, H. & Seilacher, A. 1996. Ammonoid taphonomy. In: Landman, N. H., Tanabe, K. & Davis, R. A. (eds) Ammonoid Paleobiology. Topics in Geobiology, 13, Plenum Press, New York, 543– 578. Mart, Y., Ryan, W. B. F. & Lunina, O. V. 2005. Review of the tectonics of the Levant Rift system: the structural significance of oblique continental breakup. Tectonophysics, 395, 209– 232. Martinotti, G. M. 1993. Foraminiferal evidence of a hiatus between the Aptian and Albian, offshore Northern Sinai. Journal of Foraminiferal Research, 23, 66–75. Masse, J.-P. 1993. Valanginian-Early Aptian carbonate platforms from Provence, southeastern France. In: Simo, J. A. T., Scott, R. W. & Masse, J.-P. (eds) Cretaceous Carbonate Platforms. American Association of Petroleum Geologists, Tulsa, 363–374. Masse, J.-P. 1998. (Coord. U. Cret.). Calcareous algae. In: Hardenbol, J., Jacquin, T., Farley, M. B., de Graciansky, P.-C. & Vail, P. (eds) Cretaceous Biochronostratigraphy. SEPM Special Publication, Tulsa, chart 5. Masse, J.-P., Philip, J. & Camoin, G. 1995. The Cretaceous Tethys. In: Nairn, A. E. M., Ricou, L.-E., Vrielynck, B. & Dercourt, J. (eds) The Tethys Ocean. Plenum Press, New York, 215–236. Masse, J.-P., Fenerci, M. & Pernarcic, E. 2003. Palaeobathymetric reconstruction of peritidal carbonates, Late Barremian, Urgonian, sequences of Provence (SE France). Palaeogeography, Palaeoclimatology, Palaeoecology, 200, 65–81. Menegatti, A. P., Weissert, H., Brown, R. S., Tyson, R. V., Farrimond, P., Strasser, A. & Caron, M. 1998. High-resolution d13C stratigraphy through the early Aptian “Livello Selli” of the Alpine Tethys. Paleoceanography, 13, 530– 545. Mikhailova, I. A. & Baraboshkin, E. J. 2002. Volgoceratoides and Koeneniceras – new small-size Lower Aptian heteromorphs from the Ulijanovsk region (Russian Platform). In: Summesberger, H., Hilston, K. & Daurer, A. (eds) Abhandlungen der Geologischen Bundes-Anstalt. Vienna, 539– 553. Morsi, A.-M. M. 2006. Aptian ostracodes from Gebel Raghawi (Maghara area) in northern Sinai, Egypt: taxonomic, biostratigraphic and paleobiogeographic contributions. Revue de Pale´obiologie, 25, 537– 565. Moullade, M., Kuhnt, W., Bergen, J. A., Masse, J.-P. & Tronchetti, G. 1998. Correlation of biostratigraphic and stable isotope events in the Aptian historical stratotype of La Be´doule (southeast France). Comptes Rendus de l’Acade´mie des Sciences – Series IIA – Earth and Planetary Science, 327, 693– 698. Moustafa, A. R. 2010. Structural setting and tectonic evolution of North Sinai folds, Egypt. In: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 37–63. Moustafa, A. R. & Khalil, M. H. 1990. Structural characteristics and tectonic evolution of north Sinai
130
M. BACHMANN ET AL.
fold belts. In: Said, R. & Kenawy, A. (eds) Geology of Egypt. Balkema, Rotterdam, 381–389. Moustafa, A. R., Khalil, M. H., Patton, T. L. & Thompson, T. L. 1990. Structural styles and tectonics of the Sinai Peninsula, Egypt: a field trip guidebook. Supplementary Material for the 1987 Field Trip. Egyptian Petroleum Exploration Society. Mu¨cke, A. 2000. Environmental conditions in the Late Cretaceous African Tethys: conclusions from a microscopic-microchemical study of ooidal ironstones from Egypt, Sudan and Nigeria. Journal of African Earth Sciences, 30, 25–46. Ogg, J. G., Agterberg, F. P. & Gradstein, F. M. 2004. The Cretaceous Period. In: Gradstein, F. M., Ogg, J. G. & Smith, A. G. (eds) A Geologic Time Scale 2004. Cambridge University Press, Cambridge, 344– 383. Pittet, B., Van Buchem, F. S. P., Hillga¨rtner, H., Razin, P., Grotsch, J. & Droste, H. 2002. Ecological succession, palaeoenvironmental change, and depositional sequences of Barremian– Aptian shallow-water carbonates in northern Oman. Sedimentology, 49, 555– 581. Price, G. D. 1999. The evidence and implications of polar ice during the Mesozoic. Earth Science Reviews, 48, 183– 210. Price, G. D. 2003. New constraints upon isotope variation during the early Cretaceous (Barremian–Cenomanian) from the Pacific Ocean. Geological Magazine, 140, 513– 522. Rawson, P. F., Hoedemaeker, P. J. et al. 1999. Report on the 4th International Workshop of the Lower Cretaceous Cephalopod Team (IGCP-Project 362). In: Rawson, P. F. & Hoedemaeker, P. J. (eds) Proceedings 4th International Workshop of the Lower Cretaceous Cephalopod Team (IGCP-Project 362). Scripta Geologica, Special Issue, 3, 3– 13. Renard, M., Rafe´lis, M. D. et al. 2005. Early Aptian d13C and manganese anomalies from the historical Cassis-La Be´doule stratotype sections (S.E. France): relationship with a methane hydrate dissociation event and stratigraphic implications. Carnets de Ge´ologie/Notebooks on Geology, Article 2005/04 (CG2005_A04), 1 –18. Rosenberg, E. 1960. Geologische Untersuchungen in den Naftalibergen (am Rande des no¨rdlichsten Teiles der Jordansenke, Israel). Mitteilungen aus dem Geologischen Institut der Eidgeno¨ssischen Technischen Hochschule und der Universita¨t Zu¨rich, C80, 1– 180. Rosenfeld, A. & Hirsch, F. 2005. The Cretaceous of Israel. In: Hall, J. K., Krasheninnikov, V. A., Hirsch, F., Benjamini, Ch. & Flexer, A. (eds) Geological Framework of the Levant. Volume II: The Levantine Basin and Israel. Historical ProductionsHall, Jerusalem, 393– 436. Rosenfeld, A., Hirsch, F. & Honigstein, A. 1995. Early Cretaceous ostracodes from the Levant. In: Riha, J. (ed.) International Symposium on Ostracoda, 12. Chapman and Hall, London, 111– 121. Rosenfeld, A., Hirsch, F., Honigstein, A. & Raab, M. 1998. The palaeoenvironment of Early Cretaceous ostracodes in Israel. In: Crasquin-Soleau, S., Braccini, F. & Lethiers, F. (eds) What about Ostracoda!
Bulletin de Centre Recherche et ExplorationProduction de Elf-Aquitaine, Me´moire, 20, 179– 195. Ross, D. J. 1992. Sedimentology and depositional profile of a Mid-Cretaceous shelf edge rudist reef complex, Nahal Ha’mearot, northwestern Israel. Sedimentary Geology, 79, 161 –172. Said, R. 1971. Explanatory notes to accompany the Geological Map of Egypt. Geological Survey of Egypt, Papers, 1– 123. Saint-Marc, P. 1974. Etude stratigraphique et micropaleontologique de l’Albien, du Cenomanien et du Turonien du Liban. Museum National d’Histoire Naturelle, Paris, 342. Sass, E. & Bein, A. 1982. The Cretaceous carbonate platform in Israel. Cretaceous Research, 3, 135 –144. Schroeder, R. 1975. General evolutionary trends in orbitolinas. Revista Espan˜ola de Micropaleontologı´a, Nu´m. especial, 117 –128. Schroeder, R. & Neumann, M. 1985. Les grands Foraminife`res du Cre´tace´ moyen de la Re´gion Me´diterrane´enne. Geobios, me´moire spe´ciale, 7, 1 –160. Schroeder, R., Clavel, B., Cherchi, A., Busnardo, R., Charollais, J. & Decrouez, D. 2002. Ligne´es phyle´tiques d’Orbitolinide´s de l’intervalle Hauterivien supe´rieur – Aptien infe´rieur; leur importance stratigraphique. Revue de Pale´obiologie, Gene`ve, 21, 853–863. Schulze, F., Marzouk, A. M., Bassiouni, M. A. A. & Kuss, J. 2004. The late Albian–Turonian carbonate platform succession of west-central Jordan: stratigraphy and crises. Cretaceous Research, 25, 709–737. Sharikadze, M. Z., Kakabadze, M. V. & Hoedemaeker, P. J. 2004. Aptian and Early Albian Douvilleiceratidae, Acanthohoplitidae and Parahoplitidae of Colombia. In: Donovan, S. K. (ed.) Early Cretaceous Ammonites from Colombia. Scripta Geologica, 128, 313–514. Sharland, P. R., Archer, R. et al. 2001. Arabian Plate sequence stratigraphy. Gulf PetroLink, Bahrain, 369. Simmons, M. D. 1994. Micropalaeontological biozonation of the Kahmah Group (Early Cretaceous), central Oman Mountains. In: Simmons, M. D. (ed.) Micropalaeontology and Hydrocarbon Exploration in the Middle East. Chapman and Hall, London, 177–219. Simmons, M. D. & Hart, M. B. 1987. The biostratigraphy and microfacies of the Early to midCretaceous carbonates of Wadi Mi’aidin, Central Oman Mountains. In: Hart, M. B. (ed.) Micropaleontology of Carbonate Environments. Ellis Harwood, Chichester, 176– 207. Simmons, M. D., Whittaker, J. E. & Jones, R. W. 2000. Orbitolinids from Cretaceous sediments of the Middle East – a revision of the F.R.S. Henson and Associated Collection. In: Hart, M. B., Kaminski, M. & Smart, C. (eds) Proceedings of the Fifth International Workshop on Agglutinated Foraminifera, 411–437. Stampfli, G. M. & Borel, G. D. 2002. A plate tectonic model for the Paleozoic and Mesozoic constrained by dynamic plate boundaries and restored synthetic oceanic isochrons. Earth and Planetary Science Letters, 196, 17– 33.
BARREMIAN –ALBIAN LEVANT PLATFORM EVOLUTION Stampfli, G. M., Borel, G., Cavazza, W., Mosar, J. & Ziegler, P. 2001. The Paleotectonic Atlas of the Peritethyan Domain. European Geophysical Society. CD-ROM, realized by Electronic Publishing & Consulting, Berlin. Steuber, T. & Bachmann, M. 2002. Upper Aptian – Albian rudist bivalves from Northern Sinai, Egypt. Palaeontology, 45, 725– 749. Strasser, A., Caron, M. & Gjermeni, M. 2001. The Aptian, Albian and Cenomanian of Roter Sattel, Romandes Prealps, Switzerland: a high-resolution record of oceanographic changes. Cretaceous Research, 22, 173– 199. Vail, P. R., Audemard, F., Bowman, S. A., Eisner, P. N. & Perez-Cruz, C. 1991. The stratigraphic signatures of tectonics, eustasy and sedimentology – an overview. In: Einsele, G., Ricken, W. & Seilacher, A. (eds) Cycles and Events in Stratigraphy. Springer, Berlin, Heidelberg, 617– 659. Van Buchem, F., Casanova, J. et al. 2001. Southern Tethys platform evolution and chemostratigraphy during the Middle Cretaceous (Southeastern Arabian Plate, N. Oman and U.A.E.). In: Wortmann, U. & Funk, H.-P. (eds) IAS-2001, 21st Meeting, Davos, Abstracts, 167.
131
Van Buchem, F. S. P., Pittet, B. et al. 2002. High-resolution sequence stratigraphic architecture of Barremian/Aptian carbonate systems in Northern Oman and the United Arab Emirates (Kharaib and Shuaiba Formations). GeoArabia, 7, 461–500. Weissert, H. & Erba, E. 2004. Volcanism, CO2 and palaeoclimate: a Late Jurassic-Early Cretaceous carbon and oxygen isotope record. Journal of the Geological Society, London, 161, 695–702. Weissert, H., Lini, A., Fo¨llmi, K. B. & Kuhn, O. 1998. Correlation of Early Cretaceous carbon isotope stratigraphy and platform drowning events: a possible link? Palaeogeography, Palaeoclimatology, Palaeoecology, 137, 189– 203. Westermann, G. E. G. 1996. Ammonoid life and habit. In: Landman, N. H., Tanabe, K. & Davis, R. A. (eds) Ammonoid Paleobiology. Topics in Geobiology. New York: Plenum Press, 607 –707. Wissler, L., Funk, H. & Weissert, H. 2003. Response of Early Cretaceous carbonate platforms to changes in atmospheric carbon dioxide levels. Palaeogeography, Palaeoclimatology, Palaeoecology, 200, 187 –205. Young, K. 1974. Lower Albian and Aptian (Cretaceous) ammonites of Texas. Geoscience and Man, 3, 175– 228.
The mid-Cretaceous carbonate system of northern Israel: facies evolution, tectono-sedimentary configuration and global control on the central Levant margin of the Arabian Plate RAN FRANK1*, BINYAMIN BUCHBINDER2 & CHAIM BENJAMINI1 1
Department of Geological and Environmental Sciences, Ben – Gurion University of the Negev, P.O. Box 653, Beer-Sheva 84105, Israel
2
Geological Survey of Israel, 30 Malkhe Yisrael Street, Jerusalem 95501, Israel *Corresponding author (e-mail:
[email protected]) Abstract: This study deals with the sedimentary evolution, tectonic configuration and global imprints of a Cenomanian –Turonian carbonate system located in northern Israel, on the central part of the Levant margin of the north Arabian Plate. Detailed sampling of field sections, mesoscopic features, petrography and microfacies form the database for this study. Facies units are integrated into high- and low-order cycles that comprise a sequence stratigraphic model. Two palaeo-highs, separated by a subsiding trough, all striking east– NE, govern the pattern of carbonate deposition in northern Israel. An additional subsiding region extended northward into Lebanon. Eustatic and palaeoenvironmental imprints are represented by earliest Cenomanian subaerial exposure; Early Cenomanian maximum flooding and oxygenation of hypoxic sea-floor; Middle Cenomanian high-stand progradation followed by forced regression and mass transport; Middle Cenomanian subaerial exposure; Late Cenomanian eutrophication during sea-level rise; Late Cenomanian subaerial exposure; latest Cenomanian–Turonian eutrophication and gradual development of the OAE-2 (oceanic anoxic event). A Late Cenomanian eustatic rise was locally masked by uplift and subaerial exposure. We conclude that the tectono-sedimentary regime of northern Israel represents an east– NE branch-off of the depositional strike from the north– south striking Levant margin, and that the carbonate system of this region was strongly influenced by eustasy and palaeoceanographic trends of the Tethys.
Mid-Cretaceous carbonate rocks are spectacularly exposed in the Galilee and Carmel regions of northern Israel, part of the central Levantine passive margin of the Arabian Plate (Fig. 1). Previously, sequence stratigraphic and palaeoenvironmental evolution of mid-Cretaceous carbonate systems were described for parts of the southern and eastern Levant margin (Bachmann & Kuss 1998; Bauer et al. 2003; Schulze et al. 2003, 2004, 2005). Sedimentary configuration for the Galilee and Carmel of northern Israel was previously studied by Freund (1965), Kafri (1972, 1991), Bein (1974) and Sass & Bein (1982), but a comprehensive sequence stratigraphic evolution for northern Israel was not described. Therefore, the integration of this region into the tectono-sedimentary framework of the northern Arabian plate is lacking. Previous work on the carbonate successions of northern Israel were related separately to the Carmel and Galilee regions (Fig. 1). Bein (1976, 1977), Bein & Weiler (1976), Sass & Bein (1982) and Segev & Sass (2006) considered rudist accumulations in the Carmel as the key to understanding
the configuration of the carbonate system in this region. The Carmel (Fig. 1b) was considered as a transitional zone, where shallow-marine carbonate platform sediments passed westwards via a rim of rudist barrier-reefs into deeper water laminites, turbidites and contourites. NE of the Carmel, in the Galilee, this type of east to west (proximal-distal) facies transition was not described. Freund (1965) and Kafri (1972, 1991) considered the Galilee to be a shallow carbonate platform transected by intrashelf basins fringed in places by rudist-reefs. Our comprehensive study of the Cenomanian– Turonian succession of northern Israel is based on bed-by-bed sampling, mesoscopic observations in the field, and petrography and microfacies of these rocks. Using the sequence stratigraphic paradigm, stage-by-stage evolution of this region was reconstructed, emphasizing the genesis of the carbonate system, tectonic processes, and the roles of eustasy and global palaeoenvironments. Our results fill a knowledge gap regarding the evolution and configuration of the Arabian plate margin in this area by supporting a markedly different model of
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 133–169. DOI: 10.1144/SP341.7 0305-8719/10/$15.00 # The Geological Society of London 2010.
134
R. FRANK ET AL.
Fig. 1. (a) Location of the study area within the Arabian plate, with distribution of mid-Cretaceous sedimentary units and major tectono-sedimentary features in this region (modified after Ziegler 2001). The ‘Levantine hinge-belt’ along the present coastline of Israel and Lebanon (thick bold line) is a narrow north–south striking zone across which east-to-west facies changes occur, from shallow water carbonates (Judea, Ajlun groups) to the east to deeper water carbonates (Talme Yafe Group) in the west. (b) Details of the study area in northern Israel with location of columnar sections and other relevant outcrops. (c) Graphic legend for columnar sections presented in Figures 6–10, 17 and 18.
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
tectono-sedimentary evolution for northern Israel than presented in previous works, emphasizing global palaeoceanographic control including sealevel change, nutrient levels, oxygenation of the water mass, and shelf edge configuration. Specific goals of this study were (1) to establish facies types describing the depositional environments in the mid-Cretaceous of northern Israel, (2) to integrate facies types into genetic units and to determine cyclic patterns and their hierarchical organization, (3) to document proximal-to-distal facies-thickness changes and the geometries of genetic-stratigraphic units at system-tract level, (4) to establish a local tectonic framework for these trends, (5) to uncover effects of Tethyan and global palaeoceanographic/palaeoecological events and eustatic fluctuations on the succession and (6) to present the mid-Cretaceous depositional and structural configuration of the Arabian plate margin as expressed in northern Israel.
Geological setting Tectonic setting of the Levant margin in the Cretaceous The Levant margin of the Neotethys (Fig. 1a) was shaped by Late Permian, Middle to Late Triassic and Early Jurassic extensional tectonics. Rifting led to separation of the Tauride and Eratosthenes blocks from the Arabo-Nubian Platform, forming the Levantine Basin of the eastern Mediterranean (Bein & Gvirtzman 1977; Garfunkel & Derin 1984; Garfunkel 1998; Robertson 1998). Normal faulting associated with this rifting determined the
135
depositional strike along the Levant margin, constituting the eastern margin of the Levantine Basin. The north–south trend exercised control on facies transitions along a narrow depositional belt, termed the ‘Levantine hinge-line’ or ‘belt’ (Gvirtzman & Klang 1972; Bein & Gvirtzman 1977). Cessation of faulting in the Middle Jurassic, lengthy tectonic quiescence, and slow passivemargin type subsidence prevailed along the Levant margin from the Mid-Jurassic to the mid-Cretaceous. During this period, east –west proximal-to-distal facies transitions were recorded in the hinge-belt, normal to the north–south striking shelf-margin. The mid-Cretaceous ‘Levantine hinge-belt’ is described from the southern coastal plain of Israel extending to the Carmel region (Bein 1971, 1974; Gvirtzman & Klang 1972; Bein & Weiler 1976; Bein & Gvirtzman 1977; Sass & Bein 1982). Northwards in the Galilee and southern Lebanon, the directional trend and associated facies transitions of the ‘Levantine hinge-belt’ were not reported. However, Walley (1998) described a similar hingebelt trend (Fig. 1) along the western Lebanon flexure in northern Lebanon, based on interpretation of data from Saint-Marc (1974) and others. He considered Galilee and southern Lebanon to be underlain by a southwestern extension of the Palmyride basin of Syria, representing an interruption of the north– south trend.
Cenomanian – Turonian lithostratigraphy in northern Israel Figure 2 represents the lithostratigraphy currently used for the Upper Judea Group carbonates
Fig. 2. Lithostratigraphic scheme for Cenomanian and Turonian mapped units in the Galilee and Carmel. Based mainly on the 1:200 000 geological map of Sneh et al. (1998) with some new observations. Not to scale.
136
R. FRANK ET AL.
(Cenomanian–Turonian) of northern Israel. It incorporates schemes used by Picard & Kashai (1958), Freund (1959), Kashai (1966), Kafri (1972, 1991), Sneh et al. (1998), Sneh (2002) and Segev & Sass (2006) with some local modifications. There has been a tendency to use different lithostratigraphic concepts for Galilee and Carmel, and the schemes are presented here separately. In Galilee, limestones of the Deir Hanna Formation occur at the base of the Upper Judea Group, overlying Upper Albian Yagur or Kamon dolomites (Fig. 2). The Deir Hanna Formation consists of well bedded or laminar chalks and limestones with chert nodules or bands, sometimes dolomitized near the top. Sakhnin Formation dolomites directly overlie the Deir Hanna Formation in most of Galilee. The Sakhnin dolomites become wellbedded upwards, or are sometimes highly brecciated (e.g. in the regions of Adamit, Peqi’in, Deir El-Assad, Fig. 1). In some localities, breccias of the Sakhnin Formation pass laterally into dolomitized laminites of the Upper Deir Hanna Formation (e.g. Adamit, Betzet, Fig. 1). The Sakhnin Formation was not mapped in the western Galilee (e.g. Picard & Kashai 1958; Freund 1965), where bioturbated limestones of the Yanuch Formation (Freund 1959) unconformably overlie chalky laminites of the Deir Hanna Formation. Kafri (1972) showed that the bioturbated limestones of the Yanuch Formation in western Galilee pass into chalks and bedded limestones toward the NW Galilee (e.g. Hila and Hurfesh regions, Fig. 1). Typical lithologies, microfacies and cyclic patterns of the Yanuch Formation also occur in NE Galilee (e.g. Dishon, Manara sections, Fig. 1). From western to central Galilee, the Yanuch Formation is overlain by ammonite-bearing marls of the Lower Turonian Yirka Formation. In other parts of Galilee, a limestone complex mapped as the Bina Formation (Shadmon 1959) directly overlies either the Yanuch or Sakhnin Formations. In the Carmel, the Yagur Formation dolomites of the Late Albian age (Fig. 2) form the lower part of the Judea Group. The Yagur Formation is overlain by the Isfiyye Formation, consisting of well-bedded chalks with chert nodules and bands, and in some places with pyroclastics interbedded at the base (e.g. Sass 1980; Segev & Sass 2006). The Isfiyye Formation is overlain by bioclastic limestones of the Beit-Oren Formation. The Beit-Oren limestones are overlain by chalks with some pyroclastics of the Arkan Formation (Segev et al. 2002; mapped as Khureibe and Junediyye Formations in older studies). The Arkan Formation is overlain by dolomites and limestones of the Muhraqa Formation. This transition is unconformable (Bein 1974; Lipson-Benitah et al. 1997), in places accompanied by pyroclastics (Segev & Sass 2006). Above the
Muhraqa Formation are Lower Turonian ammonitebearing marls, chalks and limestones of the Daliyya Formation in the central-northern Carmel (Freund & Raab 1969). Limestones of the Bina Formation form the top of the succession. In the region of southern Carmel, the Cenomanian succession is only partly exposed and is mostly dolomitic. At the base a dolomitized chert-bearing equivalent of the Isfiyye Formation is overlain by bioturbated dolomites of the Zikhron Formation, becoming well bedded toward the top. Dolomites of the Zikhron Formation are considered by Segev & Sass (2006) as the lateral continuation of the Arkan chalk complex to the north. The Zikhron dolomites are overlain by a pyroclastic bed (Segev & Sass 2006) followed by massive dolomites and limestones mapped as Sakhnin and Bina Formations respectively. These lithostratigraphic units, for the most part, are useful for depiction on geological maps. In order to determine the sedimentary evolution of this region, however, it was necessary to deconstruct them into their basic genetic components, and to reconstruct their cyclic patterns, systems tracts and sequences, using biostratigraphic and other chronostratigraphic controls.
Material and methods The database for this study was acquired by bed-by-bed sampling and analysis at outcrop, mesoscopic (hand sample) and petrographic scales. Thin sections were examined for non-skeletal and skeletal grains, cements and early diagenetic history. Limestone classification follows Dunham (1962) and Embry & Klovan (1971). Key micro- and macro-fossils were determined by experts. Beds in each columnar section were classified by their sedimentological features into facies units representing palaeoenvironments, broadly following the subdivisions of Read (1985) and Burchette & Wright (1992). In each section vertical cyclic pattern and cyclic hierarchies were analysed. The sequence-stratigraphic approach implemented here is based on the ‘four system-tract model’ (Hunt & Tucker 1992; Helland-Hansen & Gjelberg 1994) modified for carbonate systems (Hunt & Tucker 1993). Lowstand systems tract (LST), transgressive systems tract (TST), highstand systems tract (HST) and forced regressive (FRST) systems tract are recognized, with bounding surfaces that include (a) a basal surface of forced regression (BSFR) at the base of the FRST; (b) a sequenceboundary (SB) unconformity at the top of the proximal part of the HST, and basinwards, overlying the FRST formed during relative sea-level fall. Proximally, the SB unconformity may coalesce with the younger BSFR, while distally the SB may be expressed by a correlative conformity; (c) a
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
transgressive surface at the base of the TST; (d) a maximum-flooding surface [in fact, maximum flooding interval, (MFI)] at the top of the TST. The concept of a type-3 SB (drowning unconformity) introduced by Schlager (1999) was adopted for short-lived subaerial exposures topped by transgressive deposits, without intervening LST deposits. A time-boundary definition of some of the system tracts employed the transgressive– regressive sequence stratigraphic method of Embry (2002), with SB and maximum-flooding surfaces as main bounding surfaces defining a regressive system tract (RST). Time-correlative bounding surfaces were established using published biostratigraphy (mostly Freund & Raab 1969; Lewy & Raab 1978; LipsonBenitah et al. 1997), radiometric age determinations (Segev et al. 2002; Segev & Sass 2006), and some personal communication with experts, as explained where presented.
Results: facies types and cyclic patterns Sedimentary facies types and their environments of deposition Types of carbonate facies occurring in the Cenomanian–Turonian of northern Israel are summarized in Table 1. In total, 20 facies types are described (marked FT1– FT20 in Table 1 and Fig. 3). Column 2 in Table 1 provides details of sedimentology, bedding and faunal components. These sedimentary facies reflect a variety of depositional environments ranging from autochthonous basinal deposits and condensed surfaces, to supratidal and subaerially exposed sea-floor, as well as environments characterized by submarine transport. Column 3 in Table 1 and Figure 3 shows the palaeoenvironments indicated by these features, following the carbonate ramp subdivision into basin, outer ramp, mid-ramp and inner ramp of Burchette & Wright (1992). Column 4 lists the lithostratigraphic units within which these facies types were found, while column 5 presents their position within the sequence stratigraphic scheme. Figures 4 and 5 depict some of these features in the field and under the microscope.
Cycles and cyclic hierarchy in the Cenomanian –Turonian succession of northern Israel Sedimentary cycles in northern Israel were classified into two orders according to the following criteria: high-order cycles that cannot be further divided into smaller cycles, and therefore termed undividable cycles (UC); and low-order cycles that
137
are composed of high-order cycles, therefore termed composite cycles (CC). High-order cycles are arbitrarily numbered UC-1 –12. They are graphically presented and explained in Figures 6 and 7. UC are mostly decimetres to a few metres in thickness, but may in some cases be more than ten metres thick. UC are combinations of different facies types (Table 1, Fig. 3). UC-1 –6 (Fig. 6), are shallowing-upward cycles and UC-7 –12 are deepening-upward cycles (Fig. 7). UC-1 –UC-6, and UC-7, 11 and 12 are combined into low-order CC, but deepening-upward cycles UC-8 –10 (Fig. 7) are independent. The low-order composite cycles CC-1–7, are graphically simplified in Figure 8. Their thicknesses ranges between a few metres to a few tens of metres. † Type-1 low-order CC-1 (Figs 8a & 9) begins with basinal laminites (in some parts dolomitized), passes upwards into bioturbated fossiliferous mid-ramp deposits, and is topped by stacked UC-1 peritidal cycles. This is a shallowing-upward, progradational cycle, with the mid-ramp facies commonly rich in macrofossils and skeletal debris, reflecting high skeletal production on the mid-ramp (facies type-12 in Table 1). This high production results in filling the accommodation space to near sea level, as reflected in development of UC-1 peritidal cycles at the top. † Type-2 low-order CC-2 (Figs 8b & 9) is composed of stacked UC-3 high-order cycles (Fig. 6). Each UC-3 cycle may begin with a basal hiatal shell-concentration, then passes upwards into pelagic basinal or outer-ramp facies, rich in pithonella calcisphaeres, phosphatic grains and bioeroded bioclasts, and is topped by cross-bedded or bioturbated shoreface grainstones. The resulting CC-2 cycle indicates pulsed progradation/aggradation and filling of the accommodation space to fair-weather wave base only. In comparison with the CC-1 described above, CC-2 reflects a lower rate of skeletal carbonate production. † Type-3 low-order CC-3 (Figs 8c & 9) is composed of lower shoreface UC-4 cycles, passing upwards into upper shoreface UC-5 cycles, and ending with peri-tidal UC-1 cycles. In-situ skeletal growth in CC-3 is limited, but accommodation space was filled to near sea level, to a great extent by transported peloids. The composition and fabric of the peloids indicate origination in a predominantly shallow inner-ramp setting. Thus, progradation and accommodation filling were maintained by grain supply from the inner-ramp to the shoreface. † Type-4 low-order CC-4 (Figs 8d & 9) begins with stacked peritidal UC-1 cycles with rare
138
Table 1. Facies types of the Cenomanian –Turonian succession of northern Israel 1. Facies type FT-1
FT-2 FT-3
FT-5 FT-6 FT-7
FT-8
Platy or laminated chalk, indurated chalk, or marl, bioclastic/ peloidal calcisiltite, or micrite mudstone. Chert as nodules or bands. Fish debris, echinoids, pelagic crinoids. Planktonic, esp. heterohelicid foraminifera, pithonellid calcispheres. Pyrite, glauconite. Phosphatic grains, especially in Yanuch/ Yirka Fm. Decimeter-thick calcarenite, calcisiltite beds. Cross-laminated, graded to massive, rippled, planar. Occasionally dolomitized. Cm- to m-scale breccias, embedded in laminated or homogeneous dolomite, or chalk matrix. Internally deformed (fragmented, brecciated, folded, faulted). Dolomitized matrix. Shear features at base: boudins, folds, shear foliation, d-microstructures. Thick (.100 m) well-stratified, thin-bedded/laminated calcarenitic clinoforms dipping 228 – 358. Beds to 30 cm thick. Well sorted, mud-free, mostly non-graded. Bioturbated or thin-bedded fine-grained packstones. Pelagic microfossils: pithonellid calcispheres, planktonic foraminifera, pelagic crinoids. Phosphatic grains; mm-sized Fe concretions. One or more limestone beds within pelagic beds. Coarse bioclastic or poorly washed peloidal grainstone. Pelagic microfossils, crinoids, echinoderm debris in matrix. Bioerosion. Quartz, phosphatic grains. Chalcedony cement in borings. Densely packed bivalve shell-beds embedded in basinal chalks. Gryphaeids, Pycnodonte vesiculosa in Carmel. Shells bioeroded, some fragmented. Matrix calcisiltite and pelagic microfossils. NE Galilee: flat elongated orbitolinids; serpulids; some large bivalves. Rudists shell beds in the Bina (Kishk Fm).
3. Environment
4. Lithostratigraphic units
5. Sequence stratigraphic position
Hypoxic basinal or outer-ramp deposits. Yirka Fm – hypoxia, nutrient rich.
Deir Hanna, Yanuch, Yirka, Isfiyye, Arkan
Ce TST-1; Low Ce RST-1; Ce TST-2; Tu TST
High – low density Turbidites.
Sakhnin; mid Muhraqa. Sakhnin, Bina, Muhraqa Sakhnin
Ce FRST-1; Ce FRST-2 Ce FRST-1; Ce FRST-2 Ce FRST-1
Grain flow (Beds ,5 cm); Sandy debrites (Beds .5 cm).
Yanuch (W Galilee)
Ce FRST-2
Pelagic basinal (pelagites).
Muhraqa, Yanuch, Bina
Ce HST-2; Ce TST-2
Fragmental lag concentration. Winnowing of basinal hypoxic sea-floor.
Deir Hanna, Yirka
Top Ce TST-1; Tu TST (Max.-flooding).
Lag concentration. Colonization of hypoxic sea-floor. Low sed. rate.
Within upper Deir Hanna; part of Beit Oren (Carmel).
Top Ce TST-1; Tu TST ;
Bina Fm – mid-ramp rudist conc. during storm decline.
Within Bina (Kishk) Fm of western Galilee.
Tu HST.
Debrites. Translational sliding.
R. FRANK ET AL.
FT-4
2. Description
Thin reddish-black crusts, mm thickness, paving irregular bedding planes and exposure surfaces. Lag conglomerate.
FT-10
Bioturbated coarse-grained pack- or wackestone; chert nodules. Bioclasts: rudists, oysters, gastropods, echinoderms, fish. Pithonellid calcispheres, planktonic, rare benthonic foraminifera; pelagic crinoids. Phosphatic grains. Bioeroded bioclasts, occasional chalcedony cement. Laminated, thin-bedded or well-stratified, very fine-grained mudstones or wackestones, sterile to slightly bioclastic. Rotaliid foraminifera (especially gavelinellids); fecal pellets; rare planktonic foraminifera and fragments of echinoderms (Bina Fm). In South Galilee: dolomitized with some shear folds and synsedimentary breccias. Dolomitized bioturbated pack-, float- or wackestones. Complete and disarticulated rudists, oysters, gastropods, echinoderm debris; poriferan spicules; pithonellid calcispheres. Shallow-water benthic foraminifera (Nezazzatids, Cuneolina, Dicyclina) upwards. Hummocky Grainstones with peloids, quartz cross-stratification. silt. Algal, molluscan, echinoid SE-dipping bioclasts. Shallow-water benthic foraminifera (especially clinoforms. miliolids, Cuneolina). Massive, cross-bedded or planar grain to rudstones. Rudist and oyster bioclasts, calc. algae. Abundant mud-peloids and lithoclastic grains; some ooids (Yanuch, Bina). Isopachous rim cement; blocky spar; poikilotopic calcite. Bioturbated mudstones and wackestones with shallow-water benthic foraminifera: Nezazzatids, alveolinids, Cuneolina, Dicyclina. Rarely, small gastropods, rudist and echinoderm debris.
FT-11
FT-12
FT-13 FT-14 FT-15
FT-16
Can be found in the basin and on the platform. ferruginous mineralization of submarine omission surfaces; some reworking forming lag cgl. Outer ramp. Nutrient-rich surface water. Reduced sedimentation rate.
Sakhnin, NE Galilee; Yanuch, West Galilee; Muhraqa, Carmel.
Base UC-1; base Ce TST-2; base Pelech seq., base Tu TST
Muhraqa; Yanuch; Bina.
Ce HST-2; Ce TST-2
Hypoxic mid- or outer ramp. Hypoxic lagoonal (e.g. Betzet) Mass-transport features in South Galilee (Mt. Kedimim).
Sakhnin (Mt. Kedumim); Bina
Ce RST-1; Tu TST
Mid-ramp. Increased skeletal production.
Lower Sakhnin (Galilee), Zikhron (Carmel), Bina (NE Galilee)
Mid- e HST-1
Lower shoreface.
Bina; Muhraqa
Tu HST
Offshore bars. Upper shoreface.
Bina
Tu HST
Upper shoreface.
Yanuch, Bina
Ce HST-2
Low energy, subtidal lagoon. Intensive dolomitization beneath supratidal facies (FT-19).
Sakhnin, Bina, Daliyya
Ce HST-1; Tu TST-2; Ce HST-2
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
FT-9
(Continued)
139
140
Table 1. Continued 1. Facies type FT-17
FT-19
FT-20
Bioturbated limestone, decimeter to 2 m beds. Acteonella sp., chondrodonts, radiolitid, hippuritid rudists, poriferan, echinoderm debris; shallow-water benthic foraminifera. Platy or massive foraminiferal packstones to poorly washed grainstones. Oncoids, microbial lumps and films, micritized bioclasts. Dolomite, laminated auto-micrite, peloids. Laminae irregular to wavy; coating free surfaces and bioclasts. Fenestrae, birds-eyes, fragmentation, flat-pebble conglomerates; dissolution cracks; Fe stained. a) Thin horizon with spheroidal calcitic, silicic or partly-silicified vadose pisoids, pore-filling speleothems (flowstones), empty or infilled dissolution vugs, concentric pendant calcites. b) Irregular rugged pitted surface covered by highly-weathered marl layer. c) Thin reddish crust (millimetre-thick) with pedogenic pisoliths, circumgranular cracks and alveolar septal fabrics.
3. Environment
4. Lithostratigraphic units
5. Sequence stratigraphic position
Open-marine shallow subtidal.
Top Bina
Tu HST
Semi-restricted lagoon. Slow carbonate sed. rate. Condensed.
Bina
Base Ce TST-2
Supra- and inter-tidal flats with microbial laminites.
Sakhnin and Bina
Ce HST-1; Ce HST-2; Tu HST
Sub-aerial exposure facies: a- Calcrete; b- Karst (Lapies); c- Sub-aerially exposed surface.
a) Top Sakhnin (E Galilee). b, c) Top Yanuch
SBs:Ce SB-2. Ce SB-3. Ce SB-4.
R. FRANK ET AL.
FT-18
2. Description
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
141
Fig. 3. Schematic model for position of the major Cenomanian –Turonian facies types (facies types 1 –20, Table 1) on two carbonate ramp profiles. Note that facies type-4 includes the range from incipient slides on the inner ramp to disconnected blocks downslope. Transgressive surfaces (facies type-9) are expressed both in the basin and on the inner ramp. Dashed line represents different types of basinal condensed sections (facies types- 7, 8 and 9). Temporal distribution is presented in Table 1; spatial distribution is discussed at length in the text.
†
†
†
†
macrobenthos and passes upwards into openmarine inner-ramp UC-2 cycles in which the sub-tidal mud substrate is colonized by rudists, large chondrodont bivalves, opisthobranch gastropods (Acteonella) and sponges. CC-4 is a peritidal shallowing-upward progradational cycle. Type-5 low-order CC-5 (Figs 8e & 9) is constructed of stacked UC-6 cycles, each beginning with well-laminated mudstones and topped by massive shallow subtidal wackestones. The laminated mudstones of CC-5 were deposited on a stagnant hypoxic sea-floor and the massive beds reflects oxic bioturbated mid-ramp facies. CC-5 represents fluctuations in the oxygen level on the mid-ramp and not necessarily expansion or contraction of accommodation space. Type-6 low-order CC-6 (Figs 8f & 9) is composed of stacked UC-7 cycles with platy or laminated marl or chalk at the base (turbidites, ammonites) and skeletal hiatal concentrations at the top. CC-6 is a deepening-upwards basinal cycle. Type-7 low-order CC-7 (Figs 8g & 9) is composed of stacked UC-11 and/or UC-12 deepening-upward cycles. CC-7 reflects retrogradation of hypoxic laminites. Type-8 low-order CC-8 (Fig. 8h) is composed of UC-8 cycles, each commencing with
shallow-lagoonal facies, deepens upwards into shoreface grainstones, and ends with pelagic outer ramp facies. CC-8 is a low-order deepening-upward cycle.
Sequence stratigraphic subdivision and interpretation of systems tracts The correlations shown in Figures 9 and 10 are based on common cyclic trends found in the columnar sections, identification and extension of bounding surfaces, and biostratigraphic and radiometric age controls (mostly from Freund & Raab 1969; Lewy & Raab 1978; Lipson-Benitah et al. 1997; Segev et al. 2002). These correlations show that the Upper Judea Group consists of three Cenomanian sequences and one Turonian sequence, summarized in Figure 11. Systems tracts and bounding surfaces are presented with discussion of their facies composition and reconstruction of palaeoenvironmental configuration.
Sequence 1: Early – Middle Cenomanian Albian/Cenomanian sequence boundary (Alb/Ce SB-1). The boundary between the Yagur dolomites, considered as Albian in age (Lewy & Raab 1978), and the overlying indurated chalks of the Deir
142
R. FRANK ET AL.
Fig. 4. (a) Laminites (facies type 1) of the Ce TST-1. Lower Deir Hanna Formation, Kziv east section, eastern Galilee. (b) Bored and bioeroded bioclasts in poorly washed packstones (facies type-7) of the Early Cenomanian maximum-flooding bed (Ce MFI-1). Borings filled by chalcedonic cement. Upper Deir Hanna Formation, Betzet section, NW Galilee. (c) Poorly washed peloidal grainstones (facies type-7) of the Early Cenomanian
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
Hanna/Isfiyye Formations (Galilee/Carmel respectively; Figs 2 & 9) is a discontinuity surface, described previously by Karcz (1959), Folkman (1969), Bein (1974), Kafri (1986) and LipsonBenitah et al. (1997). Folkman (1969) and Kafri (1986) described pisolitic calcretes, monomictic conglomerates, oxidized Fe crusts, quartz grains and silicification and dedolomitization phenomena, and concluded that this discontinuity surface represents subaerial exposure. Gardosh et al. (2006) associated this subaerial exposure event with sea-level fall and a submarine canyon incision offshore. Ammonites and planktonic foraminifera (Avnimelech 1965; Lewy & Raab 1978; LipsonBenitah et al. 1997) from chalks somewhat above the Albian/Cenomanian SB in the Carmel region give an Early Cenomanian age; thus this surface is close to the Albian –Cenomanian transition. The Early Cenomanian transgressive system tract (Ce TST-1). The Ce TST-1 and the overlying Early Cenomanian maximum-flooding interval (Ce MFI-1) are in the lower –middle part of the Deir-Hanna Formation of Galilee, and in the Isfiyye and Beit-Oren Formations of Carmel (Figs 9 & 11). Ammonites and planktonic foraminifera suggest that the age of the TST is Early Cenomanian (Fig. 11). The Cenomanian TST-1 is composed of wellbedded to laminated, fine-grained, mostly calcisiltitic bioclastic debris (mudstone), with some pelagic microfossils, pyrite, and glauconite grains (facies type-1; Fig. 4a). The fine-grained bioclastic debris of this facies type, originated from the remote proximal environment, and the pelagic microfossils from the overlying water column. The laminated texture with pyrite and glauconite suggests that deposition took place in proximity to the hypoxic basin floor, presumably where the oxygen minimum zone (OMZ) impinged on the slope. This system is composed of one high-order (UC-7) deepening-upward cycle, bounded at the base by the Albian/Cenomanian SB, and at the top by the Early Cenomanian maximum-flooding interval. Lateral thickness variations of Ce TST-1 across northern Israel, from the Manara section in the NE to the Isfiyye section in the SW, are
143
simplified and scaled in Figure 12a. The Ce TST-1 is relatively thin and condensed in NE Galilee (14 m at Manara), thickens toward NW Galilee (50 m at Betzet), and reaches a maximum in the Carmel region (up to 120 m; Bein 1974). Thicker sections of more basinal facies occur towards Carmel in the SW. Early Cenomanian maximum-flooding interval (Ce MFI-1). The second correlated surface, the Ce MFI-1 (Figs 9 & 11), consists of a limestone bed intercalated within basinal chalks of the Deir Hanna Formation in Galilee, and within the Isfiyye-Arkan chalk complex in Carmel. This bed is a few centimetres to a few metres in thickness, overlying basinal thin-beds or laminated beds of the Ce TST-1. Orbitolina sefini occurs in this bed in NE Galilee (identified by M. Simmons, 2007, pers. comm.), and Early Cenomanian ammonites and planktonic foraminifera occur below and above this bed in Carmel, suggesting that it lies within the Early Cenomanian M. mantelli Zone (Fig. 11). In NE Galilee (Manara section) the Ce MFI-1 occurs as a limestone bed rich in large and flat orbitolinids, serpulids and large bivalves (facies type-8; Fig. 4e). In NW and central Galilee (e.g. Kziv-east, Betzet sections; Fig. 9) this bed appears as skeletal or peloidal hiatal concentrations with bioeroded bioclasts and pelagic microfossils (Fig. 4b, c; facies type-7) or as a shell-concentration (Kziv-east section; Fig. 4d; facies type-8). In Carmel and at Mt. Kedumim near Nazerath, the Ce MFI-1 is identified as a highly bioturbated bed or cemented basinal shell bed (‘Beit-Oren’ limestones in Carmel) capped by glauconites (Fig. 9) (Weiler 1968; LipsonBenitah et al. 1997). The matrix of the Ce MFI-1 contains pelagic micro-elements. Bioturbation, colonization by macrobenthos, and formation of lag concentrations by winnowing suggest that the Ce MFI-1 represents oxygenation of the Ce TST-1 submarine hypoxic bottom, accompanied by decreased sedimentation. Facies configuration of the Ce MFI-1 (Fig. 12) is based on the subdivision of this interval into three zones: the horizon of flat orbitolinid of NE Galilee represents the proximal facies zone of the Ce MFI-1
Fig. 4. (Continued) maximum-flooding bed (Ce MFI-1). Lower Deir Hanna Formation, Kziv east section, eastern Galilee. (d) Bivalve shell-bed (facies type-8) of the Early Cenomanian maximum-flooding bed (Ce MFI-1). Lower Deir Hanna Formation, Kziv east section, eastern Galilee. (e) Flat orbitolinids, serpulids and bioclastic debris (facies type-8) in the Early Cenomanian maximum-flooding bed of NE Galilee. Deir Hanna Formation, Manara section, NE Galilee. (f) Distal debrites (facies type-3) between basinal Deir Hanna laminites and Sakhnin Formation dolomites at Sheikh-Danun, western Galilee. Clasts are dolomitized. Matrix is composed of dolomitized Deir Hanna laminites. Breccias of this type appear locally at the base of the Sakhnin Formation in western Galilee. Lower Sakhnin Formation, Sheikh-Danun section, western Galilee. (g) Debrite breccias (facies type-3). Clasts are of peritidal origin, matrix is a textureless dolomite. Sakhnin Formation, Betzet section, NW Galilee. Hammer for scale (circled).
144
R. FRANK ET AL.
Fig. 5. (a) Northern slope of Betzet Valley, NW Galilee, near Betzet section. Albian Yagur Formation at base. Above are basinal laminites (facies type-1, indurated chalks) of the Deir-Hanna Formation passing upwards into laminites and bioturbated limestones (facies types-1, 10) of the Yanuch Formation. Two lenticular dolomite bodies (Channels 1 and 2) composed of dolomitized graded grainstones (a.1, a.2), interpreted as turbitidic channel fill
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
145
Fig. 6. Shallowing-upward high-order UC units found in the Cenomanian –Turonian succession of northern Israel. UC-6 is a cyclic unit controlled by repeated bottom stagnation on the mid-ramp zone (sterile laminites) and does not necessarily reflect expansion or contraction of accommodation space.
Fig. 5. (Continued) (facies type-2). They are incised into Deir-Hanna laminites. (b) Well-laminated graded, planar, and ripple cross-stratified dolomitized calcisilts (facies type-2) of the Sakhnin Formation, interpreted as the turbiditic Ta– Td Bouma (1962) divisions. Pen for scale. Sakhnin Formation, Gat section, western Galilee. (c) Synsedimentary foliated (yellowish) shear-zone lithology (facies type-4). Sakhnin Formation, Betzet section region, NW Galilee. (d) Photomicrograph of the shear-zone lithology (facies type-4): d– texture – mudstone fragment in foliated fine-grained dolomite. Shape indicates elongation and rotation owing to sub-horizontal shear. Sakhnin Formation, Betzet section region, NW Galilee. (e) Bedding-parallel shear planes. Shear-zone lithology such as that seen in (c) is found on these bedding planes. Sakhnin Formation, Betzet section region, NW Galilee. (f) Boudins and planar foliation in shear-zone lithology, Sakhnin Formation, Betzet section region, NW Galilee. (g) Steeply inclined slump fold in upper slope laminites (facies type-4), Sakhnin Formation, Mt. Kedumin section near Nazerath, southern Galilee. (h) Pendant calcitic/silicic cement of the Middle Cenomanian sequence boundary (Ce SB-2). Sakhnin Formation, Manara section, NE Galilee. (i) Deformed pisoid at the Ce SB-2. Interpreted as a vadose/calcrete pisoid (facies type-20). Sakhnin Formation, Manara section, NE Galilee. (j) Disolved vadose pisoids (facies type-20) at the Ce SB-2. Sakhnin Formation, Manara section, NE Galilee.
146
R. FRANK ET AL.
Fig. 7. Deepening-upward high-order UC found in the Cenomanian– Turonian succession of northern Israel. Early Turonian cycles overlying latest Cenomanian sequence boundary are the UC-7 of the Yirka Formation, the UC-8 of the Daliyya Formation, the UC-9 of the Bina Formation, the UC-11 of the Bina Formation and the UC-12 of the Daliyya and Bina Formations. They are assigned here to OAE-2.
(Fig. 12a). According to Vilas et al. (1995) and Hottinger (1997), flat orbitolinids are more common than conical forms under conditions of water cloudiness, low illumination, and increased water depth. Simmons et al. (2000) and Di Lucia et al. (2007) showed that flat orbitolinids characterize transgressive systems tracts, with the flattest forms found around maximum-flooding surfaces. Thus, flat orbitolinids at MFI-1 at Manara section indicate increased nutrients and phytoplankton at the
surface, causing turbidity and low light levels on the sea floor. Other authors also attributed flat orbitolinids to increased water cloudiness, either owing to terrigenous influx or eutrophic conditions in transgressive contexts (e.g. Immenhauser et al. 1999; Pittet et al. 2002; Al Juboury et al. 2006). The intermediate facies belt of shell-beds and bioclastic fragmental concentrations in NW and central Galilee (Fig. 12b) represents weak bottom-currents and winnowing of fine laminated
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
Fig. 8. Low-order CC-1-8 occurring in northern Israel. They are composed of high-order UC cycles. Schematic and simplified presentation: the exact number of cycles and their thicknesses are shown in Figure 9. Inferred ramp profiles are indicated for CC-1 and CC-2.
147
148 R. FRANK ET AL. Fig. 9. Sequence stratigraphic correlation of cyclic units in the Cenomanian–Turonian succession of northern Israel, from NE Galilee to the Carmel region. See Figure 1 for locations of columnar sections. Detailed explanation in text. The Pelech sequence (Seq. 3) is locally exposed in western Galilee (Pelech and Hamra Valley sections). The Middle Cenomanian lowstand system tract (Ce LST), represented in the Sheikh-Danun section, is not expressed in this correlation scheme, but is clearly expressed in the correlation scheme of Figure 10. Large parts of the sequences and system tracts are composed of two cyclic ranks, high-order UC and low-order CC.
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
149
Fig. 10. SW to NE sequence-stratigraphic correlation across the Galilee. Ce RST-1 (HST-1 and FRST-1) at Dishon is proximal, composed of CC-1 peritidal cycles; towards the SW it is much more distal and composed of basinal laminites and mass-transport deposits. CC-2 cycles build the LST in Sheikh-Danun and Kziv-west sections.
sediments of the TST-1, leaving fragmentalbioclastic or peloidal lag concentrations. In the Kziv-east section, the oxic sea-floor was subsequently colonized by bivalves and bioeroding organisms (e.g. Fig. 4d). The lower slope, or basinal facies zone, of the Ce MFI-1 occurs in the Nazareth Hills and Carmel. It is characterized by bioturbated and biodegraded macrobenthic shell material indicating oxygenation of the sea-floor, decelerated sedimentation, and probably also increased flux of nutrients (cf. Kidwell 1986; Hallock 1988). Glauconites at the top of the Ce MFI-1 were formed in this distal zone as hypoxia resumed (cf. Porrenga 1967; Carson & Crowley 1993). The Lower –Middle Cenomanian regressive system tract (Ce RST-1). The Early–Middle Cenomanian system tract corresponds in Galilee to the Upper Deir Hanna limestones and the mid-ramp/peritidal facies of the Sakhnin dolomites, and in the Carmel region to the Zikhron and Arkan Formations. The Ce RST-1 is composed of two parts, the midCenomanian HST (Ce HST-1) and the midCenomanian forced-regressive system tract (Ce FRST-1) (Figs 9, 10 & 11). These systems tracts are combined into a ‘regressive systems tract’ sensu Embry (2002) (Ce RST-1) for the purpose of time-boundary definition and palaeogeographic interpretation, but each part will be also addressed separately. The Ce RST-1 is bounded below by the Ce MFI-1 and above by the mid-Cenomanian
SB (Ce SB-2; Fig. 11), a termination most probably placed within the mid-Cenomanian A. rhotomagense zone. Its age is bracketed by Early Cenomanian Orbitolina sefini in the underlying Ce MFI-1, and by mid-Cenomanian radiometric and ammonite data from overlying beds (Fig. 11). Radiometric dating of pyroclastic rocks above the Ce SB-2 in southern Carmel gives midCenomanian date of 95.4 + 0.5 Ma (Segev et al. 2002; top Zikhron Formation). Ammonites of the A. jukesbrownei zone (Kennedy & Jolkicˇev 2004; Fig. 11) occur in chalks from above the Ce SB-2 in Galilee (Calycoceras sp. identified by Z. Lewy, pers. comm. 2007, and Protacanthoceras sp. in Freund 1958). The mid-Cenomanian HST (Ce HST-1) forms the lower part of the Ce RST-1 (Fig. 11). In most of Galilee and southern Carmel, Ce HST-1 is a progradational type-1 low-order shallowing-upward cycle (CC-1), 60– 200 m in thickness. The vertical architecture of this cycle and the arrangement of its facies components on the ramp are shown in Figure 8a. The mid-ramp zone of this cycle (facies type-12 in Table 1) is highly productive, providing lime mud downslope to the outer ramp/basin by wave and current action and by suspension, and also upslope to the inner ramp. A sharp boundary within CC-1 separating lower basinal laminites (facies type-1) from overlying bioturbated midramp facies (Fig. 8a) reflects the transition from the zone of impingement of the OMZ to the mixing zone. This boundary represents a marked
150 R. FRANK ET AL. Fig. 11. Correlation of sequences and systems tracts with chronostratigraphic controls (cf. Gradstein et al. 2004) and event chronology across northern Israel and adjacent regions. T1 to T7 are local ammonite zones of the Upper Cenomanian– Turonian succession in Israel (Freund & Raab 1969). Lithostratigraphy follows Sneh et al. (1998) for Galilee and Segev & Sass (2006) for the Carmel region. Biostratigraphic control based on planktonic foraminifera biostratigraphy in Carmel (Lipson-Benitah et al. 1997), ammonite biostratigraphy in Carmel (Kashai 1966; Lewy & Raab 1978) and Galilee (Freund 1958; Glikson 1966; Freund & Raab 1969), rudist biostratigrapy in Carmel and Galilee (Freund 1965; Buchbinder et al. 2000), identification of orbitolinids (M. Simmons 2007, pers. comm.), and other benthic foraminifera. Radiometric dating of volcanic rocks in Carmel are by Segev et al. 2002 [Dating of pyroclastics overlying Yagur Formation (97.1 + 1.7 Ma) omitted owing to wide error-bar]. Comparisons of bounding surfaces are to the northern Negev (Lewy 1990; Buchbinder et al. 2000), west-central Jordan (Schulze et al. 2003), Sinai (Bauer et al. 2003), Oman Platform (Philip et al. 1995; Van Buchem et al. 1996, 2002) and parts of the Arabian plate (Sharland et al. 2001).
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
151
Fig. 12. (a) SW to NE facies-thickness changes in the Ce TST-1 across northern Israel. Note system-tract thickening and MFI-1 deepening toward Carmel. (b) Palaeogeographic reconstruction of Early Cenomanian maximum-flooding interval.
ecological transition between non-productive sea-floor in the OMZ to highly productive mid-ramp in the mixing zone, a transition that could generate distally steepened ramp morphology. The mid-Cenomanian forced-regression system tract (Ce FRST-1) is the upper part of Ce RST-1 (Figs 10 & 11). Its base corresponds to the BSFR (sensu Hunt & Tucker 1992, 1993), upon which mass-transport deposits were emplaced (mapped as Sakhnin Formation dolomites). The upper boundary corresponds to the mid-Cenomanian SB (Ce SB-2) at the transition to overlying lowstand or transgressive deposits (Fig. 10). Mass-transport features of the Ce FRST-1 include debrite breccias (Fig. 4f, g; facies type-3), channelized and nonchannelized turbidites (Fig. 5a, b; facies type-2), and shear-planes and translational slides (Fig. 5c, g; facies type-4). Facies, geometries and sedimentary configuration of the Ce RST-1 (HST-1 and FRST-1) are presented in Figure 13. The type-1 low-order cycle of the Ce HST-1 (CC-1 in Fig. 8a) is present in most of Galilee and in the southern Carmel region (peritidal zone in Fig. 13a). Blocks and skeletal grains derived from the ramp were redeposited as forced-regressive debris-flow breccias, deformed and sheared slabs, and channelized and nonchannelized turbidites in western Galilee (facies types 2, 3 and 4). Toward NE Galilee the peritidal ramp passed distally into a productive mid-ramp zone (Fig. 13a, b). In southern Carmel, the equivalent mid-Cenomanian peritidal ramp of the Zikhron Formation passed toward the north into
mid-Cenomanian basinal deposits of the Arkan Formation (Figs 9 & 13c; see also Segev & Sass 2006). This transition is associated with ramp disintegration and formation of synsedimentary breccias. These south to north facies changes in Carmel are reflected in the transition from the Hotem –Carmel section to the Rakit–Beit Oren section (Figs 9 & 13c). A NE–SW transverse in the Ce RST-1 across Galilee is presented in Figure 13d. The peritidal system of Galilee passes gradually toward the Mt. Kedumim section at the south into thicker and deeper mid-outer-ramp deposits with some slumps and debrites (facies type-3). Farther to the south, this facies shallows once more, as represented by the peritidal facies of the southern Carmel region (Zikhron Formation of the Hotem –Carmel section). A more distal expression of this south – north transition is shown in Figure 13c. In summary, facies distribution in the Ce RST-1 indicates peritidal palaeo-highs prevailing in most of Galilee and in southern Carmel. The slope of the shallow ramp of Galilee faced north, west and south. The slope of the shallow ramp of southern Carmel faced north (and west; Bein 1974). These two palaeo-highs were separated by a deeper zone of increased subsidence that extended roughly east –west from central Carmel to southern Galilee. The mid-Cenomanian sequence boundary (Ce SB-2). The mid-Cenomanian SB separates the Ce RST-1 from the overlying TST or LST (Figs 9, 10 & 11). Pyroclastic rocks above this SB in southern Carmel (top Zikhron Formation; HC section in
152
R. FRANK ET AL.
Fig. 13. Ce RST-1. (a) Platform-basin configuration of Ce RST-1 with locations of cross sections shown in (b), (c) and, (d). Ce HST-1 forms autochthonous mid/peritidal ramp zones in Galilee and southern Carmel. The peritidal zone of Galilee passed toward northern Galilee into a deeper mid-ramp zone. Mass-transport deposits of the Ce FRST-1 were deposited on the slope. (b) Facies-thickness changes across the A–A0 –A00 line. Note ramp deepening toward the north (A0 –A) and mass-transport complex emplaced toward the west (A0 – A00 ). (c) Facies-thickness changes across line C–D. Progradational ramp system of the southern Carmel (CC1 cycle) deepens toward the north. (d) Facies-thickness changes across line A– B. The peritidal shallow-water ramp system of Galilee passes into thicker and deeper facies in southern Galilee, and further to the south into shallow-water ramp system of southern Carmel.
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
Fig. 9) were dated to 95.4 + 0.5 Ma (Segev et al. 2002), corresponding to the Middle Cenomanian A. rhotomagense ammonite zone. The ammonite Calycoceras sp., recovered from above the Ce SB-2 at the Manara section, supports this age also in Galilee (Figs 9 & 11). Diagenetic events connected with the SB affected the upper surface of the Ce HST-1. On this surface at the Manara section is a thin reddishblack ferruginous pavement, and some special diagenetic micro-features penetrate a few centimetres beneath the ferruginous crust. These features include sphaeroidal calcitic/silicic or partly silicified vadose pisoids, empty or infilled dissolution vugs, pore-filling karstic cements and concentric pendant calcites (Fig. 5h –j; facies type-20). These are vadose diagenetic features indicating subaerial exposure of the Ce HST-1 ramp. Teepee structures reported by Bogoch et al. (1994) from the Sakhnin Formation/Bina Formation transition in eastern Galilee (Kadarim region, Fig. 1) are additional evidence for subaerial exposure of the proximal ramp. The thin reddish-black mineralized surface that paves this zone of vadose vugs and cements, is interpreted as a transgressive feature (facies type-9) of the following TST. Rapid sea-level rise at the TST phase, and flooding of the SB led to sediment starvation, condensation, and mineralization. The mineralized crust sealed open cavities and inhibited sea-water percolation. This interpretation is supported by Longman (1980) who proposed that preservation of vadose cements requires rapid transgression or subsidence. Ce HST-1 (CC-1 cycle) is absent in western Galilee and in central-northern Carmel (e.g. Yanuch, Rakit/Beit-Oren sections; Fig. 9). The Ce SB-2 is represented in these regions by truncation of bedded/laminated chalks of the Deir Hanna and Arkan Formations, but with no evidence for subaerial exposure. Bein (1974) showed this truncation at the top of the chalk complex of the Carmel region (top Arkan Formation) and a coeval truncation surface is found in western Galilee at the transition from the Deir Hanna chalk laminites to the bioturbated limestones of the Yanuch Formation (Yanuch section, Fig. 9). These truncations represent submarine erosion of basinal deposits on the distal outer slopes of the western Galilee and Carmel regions and represent the distal part of the Ce SB-2 (Fig. 13a, c).
Sequence 2: late –Middle to Late Cenomanian The Middle Cenomanian lowstand system tract (Ce LST). The Middle Cenomanian lowstand system tract is shown in Figure 10. Breccias of the Ce FRST-1 (facies type-3 in Table 1) and Ce SB-2 features in
153
western Galilee (Sakhnin Formation of the SheikhDanun and Kziv-west sections, Fig. 10) are overlain by a number of subtidal UC-3 shallowing-upward cycles, forming a single CC-2 low-order cycle (Fig. 10). Such shallowing-upward cycles, directly overlying a SB, reflect progradation at the initial stage of relative sea-level rise, characterizing a lowstand systems tract (cf. Hunt & Tucker 1993). The late –Middle to Late Cenomanian transgressive system tract (Ce TST-2) and the LateCenomanian maximum-flooding interval (Ce MFI-2). The Ce TST-2 is composed of limestones and chalks bounded at the base by the Ce SB-2 or transgressive surface and at the top by a maximumflooding interval (Ce MFI-2). It corresponds to bioturbated limestones of the lower Yanuch Formation of western Galilee (Yanuch section), to limestones of the lower Bina Formation of NE Galilee (Dishon section), and to bedded chalks at Manara (part of the Yanuch Formation, but considered Deir Hanna Formation by Kafri 1991; Sneh & Weinberger 2003) (Figs 9, 10 & 11). Ammonites of the late Middle Cenomanian Acanthoceras jukesbrownei zone were recovered from the lower part of this systems tract in northern Galilee (Fig. 11) but proximal sections are more poorly constrained: benthic foraminifera from the proximal part of the Ce TST-2 at the Dishon section (Fig. 9) belong to the Pseudorhapidionina dubia total range zone (Fig. 11), spanning the Middle to Late Cenomanian (Aguilera-Franco 2001, 2005). Thus, the correlation of proximal to distal sections is based on similar facies patterns and the comparison of bounding surfaces, and supported by low-resolution biostratigraphic correlation. In both proximal sections (e.g. Dishon) and distal sections (e.g. Manara), the Ce TST-2 begins with relatively low rates of carbonate deposition, or submarine omission. Slow deposition in proximal settings is expressed by the growth of heavily micritized grains, microbial lumps, coated grains and oncoids (base Ce TST-2 at Dishon section; facies type-18). Submarine omission in distal settings is expressed by the transgressive ferruginous pavement of the Ce SB-2 at Manara (facies type-9; see above). The facies configuration of the Ce TST-2 is shown in Figure 14. The Ce TST-2 is thin and proximal at Betzet and Dishon sections. At Betzet it is composed of mid-ramp UC-6 redox cycles and at Dishon it is composed of lagoonal to outer-ramp UC-8 deepening upward cycle. The Ce TST-2 becomes much more distal toward the Carmel region in the SW, and also toward northern Galilee. In the Carmel region, it is represented by relatively condensed fine grained basinal pelagites (facies type-6 at Oren-Valley section). In northern
154
R. FRANK ET AL.
Fig. 14. (a) SW–NE cross-section along line A –B from central Carmel to the Manara section, NE Galilee, showing facies-thickness changes in the Ce RST-1 and overlying Ce TST-2. Note thickness variations (in metres) of the two superimposed system tracts. Ce TST-2 is relatively condensed and proximal in Galilee and more distal and deeper in Carmel and NE Galilee. Also note rapid thickening of the Ce TST-2 toward the north that could be related to downfaulting. (b) Sedimentary/structural configuration of northern Israel in the Ce TST-2. The Galilee palaeo-high is bounded by a basinal chalky facies belt in the north and by outer-ramp to basinal facies in the south; boundary to the basin in the north may correspond to a normal fault.
Galilee (Kedesh and Manara sections) it is relatively thick (65 –70 m) and composed of laminated calcisiltite chalks (facies type-1). The rapid thickening of the Ce TST-2 toward northern Galilee, from the Dishon section towards the Kedesh and Manara sections (Fig. 9), may reflect downfaulting to the north (Fig. 14). A deep basinal facies belt extended from the Kedesh and Manara sections in the east toward the Sarach Valley and Adamit in the west (Fig. 14b). The maximum-flooding interval terminating the Ce TST-2 (Ce MFI-2) was recognized on the Galilee palaeo-high, where it forms well-laminated pelagites (facies type-6) at the termination of the deepening-upward cycles UC-8 and UC-10 (Fig. 7) of the Dishon and Yanuch sections.
Late Cenomanian regressive system-tract (Ce RST-2). Included in the composite Ce RST-2 are the Late Cenomanian highstand and the Late Cenomanian forced-regressive systems tracts (Ce HST-2 and Ce FRST-2; Fig. 9, 11). The Ce RST-2 is placed in the Late Cenomanian (Fig. 11), as it is bounded at the base by the Late Cenomanian maximumflooding interval (Ce MFI-2) and at the top by the Late Cenomanian SB (Ce SB-3 of western Galilee; Fig. 11). The Late Cenomanian highstand systems tract (Ce HST-2) forms the lower part of the Ce RST-2. It consists of parts of the Yanuch and Bina Formations in the Galilee, and the lower part of the Muhraqa Formation at Oren Valley section in the western Carmel region (Figs 9 & 11). Latest
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
Middle Cenomanian ammonites of the previous TST, and the occurrence of the benthic foraminifera Cisalveolina fallax in the upper part of the Ce HST-2 at the Kishor section, place this system tract in Late to latest Cenomanian. In western Galilee the Ce HST-2 is topped by a basal surface of forced regression (Ce BSFR-2; Fig. 9). The Ce HST-2 is composed of a type-2 loworder cycle (CC-2) forming an aggradational or poorly progradational homoclinal ramp (Fig. 8b). In central and NW Galilee the Ce HST-2 is thin and shallow, composed of peritidal UC-1 cycles (e.g. at the Betzet section). This peritidal zone passes into thicker and deeper UC-3 subtidal cycles (outer ramp to shoreface) toward the south (Yanuch and Beit Ha’Emek sections) and NE (Dishon section). These thickness-facies changes show that central Galilee was peritidal and elevated, while to both the south and north, the ramp was deeper and more section accumulated. Further toward the south, in Carmel (Oren Valley section; Fig. 9), this system is composed of deep basinal fine-grained bioclastic pelagic ooze in the Lower Muhraqa Formation (facies type-6). The Late Cenomanian forced-regressive system tract (Ce FRST-2) terminating the Ce RST-2 in the regions of western Galilee and Carmel (Figs 9 & 11) is corresponding to parts of the Yanuch and Muhraqa Formations. In western Galilee, spectacular calcarenitic clinoform units (facies type-5) occurs from the Kishor section to the Shagor canyon, 5 km to the south (Figs 1, 15a & 16a). The clino-beds composing this unit dip 228 –358 to the west –SW, and the clinoform unit thickens rapidly toward the south, from 0 m at Beit-Ha’Emek section, 35 m at Kishor section, to more than 100 m in Hamra section (Figs 9 & 15a). The base of the clinoform body is a sharp, erosive contact, considered a basal surface of forced regression (BSFR-2 in Figs 9 & 15a). The top of the clinoform unit is truncated by a subaerial unconformity (Fig. 16b, c) onlapped by latest Cenomanian limestones (Fig. 17). In the Carmel region c. 40 km to the south –SW, the Ce FRST-2 is composed of 25– 30 m of thin-bedded graded and rippled calcarenites and calcisiltites in the middle part of the Muhraqa Formation (Figs 15b, 16d, e). The unconformity-bounded clinoform unit of western Galilee, with subaerial erosion at its top, is a forced-regression feature (cf. Plint 1988; Hunt & Tucker 1993; Posamentier & Morris 2000). The rapid thickening of the clinoform unit toward the south, from the Beit-Ha’Emek section to the Hamra section, is explained by downfaulting. Accommodation space created on the hanging walls was filled by steeply dipping calcarenites (Fig. 15a). The calcarenitic grains originated from shoreface erosion of HST-2 deposits (UC-3 cycles)
155
on the nearby shelf to the north, in the region of the Beit-Ha’Emek and Yanuch sections. The calcarenites were swept off-shelf toward the south –SW across a fault scarps developed on the south –SW facing Galilean slope (Fig. 15a, b). The toe of slope or basin can be found c. 40 km to the south – SW in distal Carmel, where thin-bedded calcarenites and calcisiltites were deposited as a turbidite sheet composed of partial Bouma (1962) sequences within the Muhraqa Formation (Figs 15b, 16d, e).
Sequence 3: Late Cenomanian SB (Ce SB-3) and Pelech sequence In the Hamra Valley of western Galilee, near the village of Pelech (PL section in Fig. 1), a 35 m section of well-bedded limestones overlies the Ce RST-2. This succession is bounded at the base and top by SBs, thus forming an independent sequence, termed the Pelech sequence (Fig. 17). The basal discontinuity is the Late Cenomanian SB (Ce SB-3), a ferruginous marl bed penetrated by decimetre-scale iron concentrations interpreted as root casts. The upper discontinuity is the latest Cenomanian SB (Ce SB-4), an irregular surface with karst features (Fig. 16c). The Pelech sequence is composed of fine-grained pelagic packstones, skeletal wackestones, or sterile mudstones. The detailed lateral relations between the Pelech sequence and adjacent facies units are not fully presented here, but this locally exposed sequence onlaps to the west over the forced-regressive clinoform unit of Ce FRST-2 (Fig. 17). The benthic foraminifera Cisalveolina fallax from the underlying HST, and caprinid rudist fragments from the Pelech sequence, suggest a latest Cenomanian age, probably corresponding to the M. geslinianum ammonite zone (cf. biostratigraphic scheme of Aguilera-Franco et al. 2001). The Pelech sequence is composed of finegrained pelagic outer-ramp packstones and midramp wackestones (facies types 10 and 17; Table 1). The SB at its base reflects sea-level fall at the end of the Ce RST-2, and the Pelech sequence represents the subsequent rise. Most of the Pelech sequence was removed by erosion during the latest Cenomanian exposure of the latest Cenomanian SB (Ce SB-4). The latest Cenomanian sequence boundary in northern Israel (Ce SB-4). The latest Cenomanian SB occurs between Cenomanian and Lower Turonian strata in northern Israel. In western Galilee, the clinoform unit (Ce RST-2), and the Pelech sequence are truncated by a reddish, encrusted surface with pedogenic pisoliths, circumgranular cracks, alveolar septal fabrics and karst features (facies type-9; Figs 16b, e & 17). In the Carmel
156
R. FRANK ET AL.
Fig. 15. Ce RST-2 and Tu TST. (a) facies-thickness changes in the Ce FRST-2 (Ce HST-2 and Ce FRST-2) and Tu TST showing a SW– NE cross-section along line A– B from the Oren Valley section of western Carmel to the Dishon section of NE Galilee. Note thickness variations (in metres) of the superposed system tracts. All systems tracts show that part of Galilee is proximal and condensed with respect to a thicker and deeper facies to the NE and SW. The Ce FRST-2 of western Galilee consists of thick (35–100 m) and steep (22–358) calcarenitic clinoform unit rapidly thickening to the south. Thickness variations suggest that this calcarenitic clinoform body reflects a syndepositional response to block downfaulting toward the south (see text for details). (b, c) Sedimentary/structural configurations in the Ce RST-2 and Tu TST.
region, the Cenomanian–Turonian transition lies within the Muhraqa Formation complex, which has individual caprinid rudists in its lower part and individual hippuritids in the upper part (Buchbinder et al. 2000). In the Muhraqa Formation of the Isfiyye
section of the Carmel (Fig. 1) a ferruginous crust (facies type-9) overlies Upper Cenomanian forcedregressive debrite breccias (Fig. 18). In western Galilee, subaerial exposure accompanied by pedogenesis and karstification (facies
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
157
Fig. 16. (a) Well-stratified clinoform unit at Kishor dipping 20–258 to the SW. Clinobeds downlap on erosion surface (BSFR). Underlying beds are tilted 10– 128 to the NE. Yanuch Formation, Kishor section, western Galilee. (b) Palaeo-calcrete fabric from top Mt. Gamal clinoform unit. Left: Dense micrite with a complex network of cracks forming irregular micrite nodules. Right: heterogeneous fabric with sub-spherical concentric pisoids (rectangle). Top Yanuch Formation, Mt. Gamal section, western Galilee. (c) Highly irregular karst surface at the top of the Yanuch Formation interpreted as ancient lapie´s. Above is highly weathered soft horizon, either basal Yirka Formation or a palaeosol. Top Yanuch Formation, Hamra section, western Galilee. (d) Well-stratified or thin-bedded limestones of the Muhraqa Formation. These beds are composed of peloidal-bioclastic grainstones that are commonly graded. Muhraqa Formation, Oren Valley section, western Carmel. (e) Turbidite from the middle Muhraqa Formation. Sample is graded at the base (Ta), passed upwards to sub-planar lamination (Tb) and is rippled at top (Tc). Muhraqa Formation, Oren Valley section, western Carmel.
158
R. FRANK ET AL.
Fig. 17. The Pelech sequence of the Hamra Valley, western Galilee (region of Hamra Valley section in Fig. 1) as seen by correlation between two adjacent sections, the Pelech (PL) section in the east and the Hamra (HM) section in the west. For location of these sections see Figure 1. The Pelech sequence, recognized only in western Galilee, is composed of well-stratified wackestones and packstones of outer- to mid-ramp origin. It is bounded by discontinuities and onlaps the forced-regressive clinoform unit of the Hamra Valley from the east. Note faulted clinoform unit and amalgamation of the Late Cenomanian SB (Ce SB-3) with the younger latest Cenomanian SB (Ce SB-4).
type-20, Table 1) affected Upper Cenomanian strata prior to Early Turonian transgression (Fig. 17). On the other hand, Upper Cenomanian strata in the Isfiyye section of the Carmel region lack subaerial exposure features and the thin ferruginous crust is interpreted as submarine omission surface at the beginning of the subsequent Early Turonian transgression. Therefore, at the end of the Cenomanian, relative sea-level fall resulted in exposure in Galilee while the Carmel region remained submerged. Buchbinder et al. (2000) placed the latest Cenomanian SB in northern Israel between the Upper Cenomanian ‘Yanuch and Lower-Muhraqa’ Formations and the Turonian ‘Yirka and UpperMuhraqa’ Formations (Fig. 2). A prolonged hiatus spanning 1.25 ma was indicated, extending from the upper part of the Late Cenomanian N. vibrayeanus zone, until the middle part of the Early Turonian W. coloradoense zone. In the present study, the benthic foraminifera Cisalveolina fallax was recovered from below this SB in western Galilee (Fig. 11) and ammonites of the Turonian C. securiforme Zone (T4 of Freund & Raab 1969) occur c. 20 m above it. A C. fallax zone was defined by Saint-Marc (1974) as spanning the latest Cenomanian to Early Turonian. Therefore, the hiatus in this region may not be as extensive as the previous estimation.
Sequence 4: Early and Middle Turonian Turonian transgressive system tract (Tu TST). The Tu TST (Figs 9 & 11) is bounded at the base by the latest Cenomanian SB (Ce SB-4) and at the top by a maximum-flooding interval. In the Carmel region it corresponds to part of the Upper Muhraqa Formation limestones and the Daliyya Formation, and in Galilee to the uppermost Yanuch Formation, Yirka Formation and part of the Bina Formation. The Tu TST mostly consists of deep basinal facies, turbidites, hiatal concentration (sensu Kidwell 1986), and hypoxic laminites, but shoreface deposits are present as well (facies types 1, 2, 7, 11 and 15 in Table 1). In the Carmel region, the Tu TST begins with deepening upwards cycles from the lower shoreface to the basin (UC-12 cycles forming a single CC-7 cycle, Fig. 8). Towards southern Carmel these cycles pass into deepening-upward cycles with lagoonal aspects (UC-8 cycles forming one CC-8 cycle, Fig. 18). This facies transition indicates that in the Lower Turonian shallow ramp conditions prevailed in the southern Carmel region and deeper outerramp to basinal conditions prevailed toward the north. In western Galilee, the Tu TST similarly begins with basinal marls (facies type-1) of the Yirka Formation (e.g. at Hamra section, Fig. 9)
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
159
Fig. 18. Correlation between the Rakefet section (RKF) and the Isfiyye section (ISF) located in the Carmel region (for location of these sections see Fig. 1). Datum is the latest Cenomanian SB (Ce SB-4). Deepening-upward CC-8 cycles are noted in the Rakefet section of southern Carmel and CC-7 cycles are noted in the Isfiyye section at the north. Note that abundant shallow-water microfauna is present at the CC-8 cycle at Rakefet, compared to deep water pelagic microfauna (especially planktonic foraminifera and calcisphaeres) in the CC-7 cycle at Isfiyye.
and with retrogradational shoreface grainstones passing upward to turbidite/pelagic marl farther to the north (facies types 2 and 15). In other parts of Galilee the Tu TST corresponds to hypoxic midto outer-ramp deepening-upward cycles in the Bina Formation (UC-9, UC-11 and UC-12). Geometry and facies of the Tu TST are shown in Figure 15a, c. In central and NW Galilee the Tu TST is condensed (mid- to outer-ramp UC-9 cycle). At the Dishon section in the NE, it is a much thicker deepening-upward cycle (55 m) composed mainly
of mid-outer-ramp laminites (CC-7; facies type-11). At the Yanuch section in the SW the Tu TST is composed of a 30 m thick deepening-upward cycle (30 m) (CC-6), and at the Hamra section to the SW it is composed of 80 m of basinal marls of the Yirka Formation (facies type11). Composite facies configuration of the Tu TST is shown in Figure 15c. The thin condensed cycle of central and NW Galilee reflects condensation on a palaeohigh. This zone was bounded to the north and south by deeper and more rapidly subsiding
160
R. FRANK ET AL.
regions. Outer-ramp laminites of NE Galilee onlapped the Galilean palaeo-high from the north, and basinal marls onlapped the steep clinoformic slope of western Galilee from the south.
Discussion Tectono-sedimentary framework of the central Levant margin, northern Israel The mid-Cretaceous ‘Levantine hinge-belt’ (see above; Fig. 1) extended from the southern coastalplain of Israel northwards to the Carmel region (Gvirtzman & Klang 1972; Bein & Gvirtzman 1977), and also formed a coast-parallel, north– south striking facies belt in northern Lebanon (Walley 1998). Galilee and southern Lebanon were considered by Walley (1998) as an interruption of this trend, caused by the extension of the Palmyride basin to the SW. The data and sequence stratigraphic interpretations presented here support the overall palinspastic model presented by Walley (1998). We have shown that a large part of Galilee was structurally elevated and bounded to the south and north by deeper, more rapidly subsiding regions. This tectono-sedimentary configuration is consistent throughout the studied succession, expressed in system tracts Ce TST-1 (Fig. 12), Ce RST-1 (Figs 13 & 14), Ce TST-2 (Fig. 14), Ce RST-2 (Fig. 15) and Tu TST (Fig. 15). There is also some evidence that the southern Carmel region was elevated with respect to a deeper, more subsident zone in central-northern Carmel, a trend expressed in system tracts Ce RST-1 and Tu TST. Data from the Ce RST-1 at Mt. Kedumin and southern Carmel (Figs 9 & 13) and data from the Tu TST in southern Carmel (Figs 9, 15c & 18) suggest that a subsiding trough, also trending east– NE, separated the shallow and more elevated southern Carmel region from a likewise shallow and more elevated Galilee. This trough extended from central-northern Carmel toward the southern part of the Galilee. Thus, the Cenomanian –Turonian sedimentary configuration in northern Israel is characterized by facies transitions indicating that the north– south striking hinge-belt (depositional strike) of the Levant margin (Fig. 1) shifted to the east –NE, northwards of the southern Carmel region (Hotem Carmel section, Fig. 1). Structural control on facies and systems-tract geometry extends beyond Galilee. Data of Saint-Marc (1972, 1974) from Mt. Sannine in central Lebanon show that the Cenomanian succession further thickens towards central Lebanon (Fig. 19). This thickening represents the northward continuation of the subsidence trend recorded here for northernmost Galilee.
Fig. 19. SSW–NNE cross-section from central Carmel, via Galilee, towards Mt. Sannine, central Lebanon, showing thickness variations of the Cenomanian succession. Note significant thickening from Galilee toward Mt. Saninne at the north. Data from Mt. Saninne are from Saint-Marc (1972, 1974).
The east –NE structural trend characterizing northern Israel in the Cenomanian and Turonian stages corresponds to other large-scale trends in the Levant. The SW –NE Palmyride trend in Syria was active, at least since the Triassic and possibly even Late Palaeozoic period (Ponikarov et al. 1967; Brew et al. 2001). Similarly, Ferry et al. (2007) showed that depositional strikes represented in the Lower Aptian Jezzine Formation and Upper Albian Niha Formation of Lebanon trend east –NE. This trend is also sub-parallel to depositional strikes described from mid-Cretaceous carbonate platforms in the Levant such as those studied by Carter & Gillcrist (1994) from the northern margin of the Arabian Plate (SE Turkey), by Schulze et al. (2005) from west-central Jordan, and by Bachmann & Kuss (1998) and Bauer et al. (2003) from northern Sinai. These depositional trends have not been described from the mid-Cretaceous of Israel. The integration of all these trends into a complete framework of the northern Arabian margin
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
may be premature. Clearly, a comprehensive compilation of all thickness and facies data is necessary to complete the tectono-sedimentary configuration of the Levant margin in the Mesozoic.
Controlling mechanisms and correlation across the Arabian Plate and beyond Regional significance of the Albian–Cenomanian SB of northern Israel. The Alb/Ce SB-1 of northern Israel was described as a subaerial unconformity surface (Folkman 1969; Kafri 1986). The Alb/Ce SB-1 in central and southern Israel corresponds to a discontinuity characterized by a shell-bed of Pycnodonte vesiculosa (Lewy & Weissbrod 1993; Braun & Hirsch 1994) but evidence for subaerial exposure was not reported. The Albian– Cenomanian boundary in central Jordan (Abed 1984) was described as a subaerial erosion surface paved by conglomerates, and in Lebanon (Ferry et al. 2007), as an emergent surface. Approximately coeval hiatuses and relative sea-level fall have been recorded globally, for example, in SW England (Simmons et al. 1991), the Anglo-Paris basin, and as far as Crimea, Kazakhstan, Turkmenistan and Iran (Gale et al. 1996), as well as from the North American western interior basin (Gro¨ke et al. 1998). Lower part of Cenomanian sequence-1: eustatic and environmental controls. A plate-wide earliest Cenomanian maximum-flooding surface K120 was described by Sharland et al. (2001) at 98 Ma. It was recognized in SE Turkey, Lebanon, eastern Syria, Iraq, Iran, Kuwait, Saudi Arabia, Qatar, UAE, Oman and Yemen. In the Arabian Plate, the K120 interval largely corresponds to organic-rich bituminous shales, or carbonate mudstones and wackestones. A section consisting of organic-rich shales from the base of the Natih-E member in Oman was chosen as a ‘reference section’ of this maximum-flooding event. The Early Cenomanian maximum-flooding episode of northern Israel (Ce MFI-1) correlates with K120 (Fig. 11). This is indicated by the presence of Orbitolina sefini in this interval at the Manara section and occurrences of Early Cenomanian ammonites below and above it in the Carmel region (Fig. 11). However, the Ce MFI-1 has a different sedimentological expression than the K120. It corresponds to a sharp facies transition from stagnant basinal laminites (Ce TST-1) into bioturbated –bioeroded shell-beds, bioclasticpeloidal hiatal-concentrations, glauconite-rich horizons and flat-orbitolinid limestones. This facies transition represents weak current-winnowing and submarine bottom reworking, oxygenation of hypoxic sea-floor, and possible elevation of nutrient levels in the water body. A similar timeequivalent shell bed occurs within the well-stratified
161
basinal succession of the Ein-Yorkeam Formation in northern Negev, southern Israel (Figs 1 & 11). In north-central Jordan, Schulze et al. (2003) correlated the K120 to the surface separating Lower Cenomanian marls of the Naur-A member from shallower rudist buildups and shallowing upward cycles of the Naur-B member. Other Early Cenomanian maximum-flooding episodes from Europe (Caus et al. 1997; Wilmsen 2008) are somewhat younger than the K120 of the Arabian plate but may provide some explanations for the sedimentological differences between the shelly bioclastic Ce MFI-1 of northern Israel and the bituminous K120 described from the Arabian plate. In the Sopeira Formation of SE Spain, Caus et al. (1997) showed basinal deposits capped by a maximum-flooding bed near the transition from the Early Cenomanian to the Middle Cenomanian. This is also an Early Cenomanian maximumflooding event with nearly identical facies to Galilee, for example, a proximal facies with abundant orbitolinids. Another maximum-flooding event was described by Wilmsen (2008) in the Early Cenomanian mid-M. dixoni zone of northern Germany. It corresponds to inoceramid shell-beds (Schloenbachia/Inoceramus virgatus bioevent) widely distributed in NW Europe. Despite uncertain timing or diachrony, Lower Cenomanian shell-beds of Israel and Europe were similarly affected by mechanisms of bottom reworking near or below storm wave base. Therefore, while Sharland et al. (2001) attributed the K120 event in the Arabian plate to eustatic rise and local subsidence, the Cenomanian maximum-flooding intervals in Israel and Europe record an additional palaeo-environmental input. End of Cenomanian sequence-1: eustatic and palaeoenvironmental controls. The Lower to Middle Cenomanian highstand system tract of northern Israel (Ce HST-1) represents a progradational ramp system. In most of Galilee and southern Carmel it is a type-1 low-order cycle (CC-1; Fig. 8a) characterized by complete filling of the accommodation space to peritidal depths as a consequence of two synchronously operating factors: (a) the development of an effective, mollusc-dominated carbonate-production zone on the mid-ramp that supplied carbonate mud both upslope to the innerramp zone and downslope to the basinal areas (Fig. 8a, facies type-12 in Table 1), and (b) reduction of the accommodation space owing to eustatic fall. The development of a mollusc-dominated carbonate-production zone on the mid-ramp indicates mesotrophic conditions favoured by rudists and other molluscs. Termination of this highly productive carbonate ramp was first by gravity collapse represented by the Middle Cenomanian FRST, and
162
R. FRANK ET AL.
subsequently by ramp emergence forming the Middle Cenomanian SB. The ramp system of this stage represents an Early to Middle Cenomanian regressive event in northern Israel. This Early to Middle Cenomanian regressive event of northern Israel can be identified in the northern Negev of southern Israel (Fig. 11). Lewy (1990) described an Early to Middle Cenomanian regressive event (Zafit Formation), bounded at the top by subaerial exposure surface (Lewy & Avni 1988) and overlain by transgressive deposits of the Avnon Formation. The age and facies stacking pattern of the Zafit Formation are similar to those of the CC-1 cycle of northern Israel (Fig. 8a). An Early to Middle Cenomanian regressive event was also recorded from other locations in the Arabian Platform (Fig. 11) and Europe. In west central Jordan, Schulze et al. (2003, 2004) considered the rudist-bearing shallowing-upward cycles of the Naur-D member as a regressive HST. The Naur-D member is capped by a bored ferruginous hardground (SB CeJo2) overlain by the drowning succession of the Fuheis Formation. Philip et al. (1995) recognized an Lower to Middle Cenomanian shallowing-upward HST pattern in the Natih-E member of the Natih Formation in the Oman Platform. Van Buchem et al. (2002) assigned this interval in the Natih Formation to sequence-1 of the Cenomanian, terminated by a Middle Cenomanian exposure surface (type-1 SB) 10 m below top Natih-E member, and Gre´laud et al. (2006) associated this boundary to eustatic fall, incisions, and deposition of forced-regressive wedges. Caus et al. (1997) considered the Middle Cenomanian part of the Santa-Fe limestones of NE Spain as a progradational system redeposited as breccias in the basin and topped by a subaerial type-1 SB at the Middle/Late Cenomanian boundary. In the deeper marine environments of the Anglo –Paris basin, Robaszynski et al. (1998) described a major fall in sea-level in the late Early Cenomanian, in the later part of the M. dixoni zone, followed by a strong transgression in the earliest A. rhotomagense zone, somewhat earlier than the late Middle Cenomanian transgressive episode recorded in northern Israel. Middle Cenomanian hiati, omission, condensation, and deep erosion were reported across western Europe, England, and Ireland and the Caucasus, as well as from the western interior seaway of America (e.g. Baraboshkin et al. 1998; Hancock 2003; Tro¨ger 2003). These studies suggest that the pattern of sedimentation in the late Early to Middle Cenomanian in marginal marine environments of the Arabian plate and parts of Europe is regressive –progradational, terminating by subaerial or subaquatic omission. In deeper environments sedimentation is associated with condensation and hiati, and redeposition of shallow-water deposits
(northern Israel, Spain and England). Hancock & Kaufman (1979), Van Buchen et al. (2002) and Hancock (2003) associated this Middle Cenomanian regressive event with an eustatic sea-level fall. The beginning of sequence-2: eustatic and palaeoenvironmental controls. Sequence-2 of northern Israel begins with transgression in the late Middle Cenomanian (Ce TST-2). This interval is represented by deposition of pelagites in basinal regions and by starvation, non-deposition, and development of a proximal deepening-upward cycle on the palaeo-high of Galilee. This transgressive event can be correlated across Israel and the Arabian plate (Fig. 11). In southern Israel (Fig. 1), transgressive outer-ramp nodular limestones of the Avnon Formation appear above the Middle Cenomanian SB of the top Zafit Formation, indicating platform drowning (Wald 2004; Fig. 11). As in Galilee, the transgressive succession of the northern Negev terminates with a maximum-flooding bed composed of laminated pelagites (facies type-6) with abundant small calcisphaeres, planktonic foraminifera, and other pelagic fossils. The last occurrence of orbitolinid foraminifera is slightly above this maximum-flooding bed, suggesting that the Ce MFI-2 of the northern Negev is Middle Cenomanian in age as in Galilee. In west-central Jordan, the equivalent drowning succession above the CeJo2 SB starts with dark bituminous marls and limestones of the Middle Cenomanian Fuheis Formation (Schulze et al. 2003). This transgressive event may correspond with the Ce TST-2 of northern Israel but the overlying progradational phase (Karak limestone) was not identified in northern Israel. In the Oman Platform, the Middle Cenomanian transgression is recognized above shallowwater lowstand deposits at the Upper Natih-E member with a maximum-flooding within the Lower Natih-D Member (Sharland et al. 2001; Van Buchem et al. 2002). This maximum-flooding in Oman corresponds to the calcareous shale unit of the K130 MFS of the Arabian plate and approximately corresponds to the laminated pelagites of the Ce MFI-2 of northern Israel. The Middle Cenomanian drowning succession in northern Israel represents the beginning of a lengthy episode of eutrophication associated to low skeletal productivity, spanning the entire later Cenomanian and Early Turonian (cf. Buchbinder et al. 2000). Eutrophication is especially expressed by blooms of pithonellid calcisphaeres. Cretaceous pithonellids are planktonic, unicellular, thermophilic and opportunistic dinoflagellates typical of surface waters of the continental margins in the neritic –pelagic transitional zone (c. 80–300 m water depth). They thrive under stressed conditions of elevated salinity, temperature and saturation with
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
respect to CaCO3 (Dias-Brito 2000). Other studies connect mid-Cretaceous pithonellid-rich deposits to nutrient-rich surface waters above coastal upwelling (Banner 1972; Jarvis et al. 1988; Wendler et al. 2002). In central Jordan eutrophication controlled coeval facies of sequence-2, and Schulze et al. (2004) invoked both sea-level rise and deposition of dysoxic bituminous successions in local basins to explain the Jordanian succession. Van Buchem et al. (2002) preferred increased fresh-water, nutrient and clay injection to explain coeval water cloudiness in the Oman platform. Latter part of sequence-2: eustatic, palaeoenvironmental, and tectonic controls. Late Cenomanian changes in northern Israel are expressed in the Ce RST-2 and the overlying Pelech sequence (Figs 9 & 11). Depositional processes were affected both by sea-level fluctuations and localized tectonic movements. Late Cenomanian aggradation began with the establishment of a CC-2 cycle reflecting the establishment of a homoclinal ramp (Ce HST-2; Figs 8b & 11). Sea-level fall and normal faulting in Galilee led to the transformation of this ramp into an open, non-rimmed shelf, with a steep clinoform-bearing slope facing south– SW, and with toe-of-slope turbidites distally in the Carmel region (Ce FRST-2; Fig 15). This shelf-slope system was exposed for the first time during the Late Cenomanian as evidenced by the Late Cenomanian SB (Ce SB-3; Fig. 11). Shortly afterwards, the exposed shelf-slope system of western Galilee was drowned below the storm-wave base, forming the drowning succession of the Pelech sequence (Figs 11 & 17), partly removed by erosion in a second phase of emergence in the latest Cenomanian (Ce SB-4). The Late Cenomanian SB (Ce SB-3) is coeval with subaerial exposure recorded by Voigt et al. (2006) in the M. geslinianum zone of NW Europe, correlated by them to other surfaces in the Anglo– Paris basin, NE Germany, Crimea, Ukraine, Mangyshlak (Kazahstan) and SE India. As in Galilee, this surface was subsequently submerged, but a second exposure surface, as found in northern Israel (Ce SB-4; Figs 11 & 17) was not reported. Equivalents of the Late Cenomanian SB, the Pelech sequence, and the latest Cenomanian SB of northern Israel are recorded from Wadi Feiran of Sinai. In this region, Kassab & Obaidalla (2001) show a basal Late Cenomanian erosional surface truncating part of the M. geslinianum ammonite zone, overlain by a 7.5 m thick succession, and then with a second hiatus at the Cenomanian– Turonian boundary. The Late Cenomanian SB may correspond to the Ce Sin-6 reported by Bauer et al. (2003) from Sinai, while their Ce Sin-7 may
163
correspond to the overlying latest Cenomanian SB. The Late Cenomanian SB may also correspond to the slightly younger Ce Jo4 SB reported by Schulze et al. (2003) from Jordan and to the Ce Up SB at the top of the Tamar Formation of the Negev (Buchbinder et al. 2000), both within the Late Cenomanian N. vibrayeanus zone (Fig. 11). The termination of the Upper Cenomanian homoclinal carbonate ramp system of Galilee (Ce HST-2) is unique, as drowning (by the Pelech sequence) was preceded by forced-regression and subaerial exposure, and followed by a second subaerial exposure event (latest Cenomanian SB). Regression and platform exposure at the end of the Cenomanian in Israel and Sinai were suggested by Flexer et al. (1986), Lewy & Avni (1988) and Bauer et al. (2003) on the basis of Fe-crusts and borings, but robust indications for subaerial exposure such as karst, calcrete, palaeosols, or fluvio-deltaic deposits were not reported. For example, the top of the uppermost Cenomanian in the Negev and central Israel is characterized by ferruginous burrowed or brecciated horizons that are more likely to be indicative of subaquatic omission (see recent example of Heck et al. 2007, ancient example of Immenhauser et al. 2000). Furthermore, in many marginal carbonate systems around the Tethys subaerial exposure of Upper Cenomanian platforms was not recognized and platform termination resulted from drowning owing to sealevel rise, or eutrophication and anoxia (Jenkyns 1991; Philip & Airaud-Crumiere 1991; Gusˇic & Jelaska 1993; Philip et al. 1995; Caus et al. 1997; Drzewiecki & Simo 1997; Schlager 1999; Buchbinder et al. 2000; Scott et al. 2000; Bauer et al. 2001, 2003; Schulze et al. 2004). The first Late Cenomanian subaerial exposure event (Ce SB-3, Fig. 11) was influenced by a eustatic fall at the M. geslinianum zone (Flexer et al. 1986; Voigt et al. 2006), and also by local faulting in Galilee (Fig. 15). On the other hand, the younger subaerial latest Cenomanian SB (Ce SB-4, Fig. 11) seems to result from a local tectonic uplift of Galilee and does not have equivalents in Europe. Late Cenomanian tectonic movements associated to local uplifts were also reported from North Sinai by Bauer et al. (2003), and localized subaerial unconformities attributed to local tectonic events were reported from the Cenomanian on the eastern Arabian plate and the northern margins of the Tethys (Van Buchem et al. 1996 and Scott et al. 2000 – Oman; Borgomano 2000 – Apulian platform; Wilmsen 2000 – Cantabrian platform; Mouty et al. 2003 – NE Palmyrides). Manifestation of oceanic anoxic event-2 (OAE-2; ‘Bonarelli’ Event) across the latest Cenomanian sequence boundary: palaeoenvironmental and eustatic control. The latest Cenomanian SB of
164
R. FRANK ET AL.
Galilee represents both a subaerial unconformity and transgressive surface onlapped by Lower Turonian deepening-upward cycles UC-7, UC-8, UC-9, UC-11 and UC-12 (Figs 7 & 8). These Early Turonian cycles represent initiation of drowning caused by sea-level rise and onlap of deep-water facies onto the Galilee palaeo-high (Tu TST in Fig. 15). Above the latest Cenomanian SB, the Lower Turonian Daliyya marls of Carmel contain heterohelicid foraminifera, dinoflagellates, evidence for high vegetative productivity, high concentrations of total organic carbon (1.06–2.02%), and pyrite (Honigstein et al. 1989). The Daliyya marls were considered by Honigstein et al. (1989) to have formed under ecological stress and hypoxia, and were related to the second oceanic anoxic event of the Tethys (OAE-2; Arthur et al. 1987). The correlative marls in the CC-6 cycles of the Yirka Formation, onlapping Galilee from the south (Fig. 15), contain similar features indicating increased nutrients, and hypoxia (facies type-1). The well-laminated and thin-bedded CC-7 and UC-9 cycles of the Bina Formation (facies type-11), onlapping Galilee palaeo-high from the north (Fig. 15) represent hypoxia and stagnation of Lower Turonian sea-floor deposits as well. These
well-stratified, well-laminated mudstones and wackestones are characterized by sparse but consistent occurrences of rotaliid foraminifera, especially gavelinellids (facies type-11 in Table 1, Fig. 3). Rotaliid foraminifera began to dominate oxygenpoor bottoms from the Albian, and small benthic foraminifera including gavelinellids, were used as deep-water hypoxic markers of the Cenomanian – Turonian OAE in the Tarfaya basin, Morocco (Gebhardt et al. 2004). This well-laminated rotaliidbearing facies in the Bina Formation (facies type-11) therefore also reflects hypoxia related to the OAE-2. The succession of events above the Middle Cenomanian SB (Fig. 20) shows that Lower Turonian hypoxia related to the OAE-2 of the Tethys was at the culmination of a lengthy trajectory of eutrophication (Fig. 20). In the Ce TST-2, mass occurrences of pithonellid calcisphaeres, phosphatic grains, and condensation indicate increased nutrients in the water body. During the subsequent Ce HST-2 the carbonate system failed to keep up with the rising sea level as a result of low auto-production of skeletal carbonate (as reflected in the UC-3 cycles), hence development of a homoclinal ramp (Fig. 8b). The Late Cenomanian SB reflects final
Fig. 20. Palaeoenvironmental evolution toward the Early Turonian OAE-2 climax, as reflected in the Upper Cenomanian– Lower Turonian sections of Galilee. The Upper Cenomanian carbonate system preceding the Middle Cenomanian SB shows gradual decrease in skeletal production, increase in eutrophication and deterioration to significant hypoxia in the Early Turonian.
THE CARBONATE SYSTEM OF NORTHERN ISRAEL
deterioration of this non-productive homoclinal ramp at a time of Late Cenomanian emergence. The subsequent Pelech sequence, depleted in skeletal grainstones but rich with pithonellid calcisphaeres, reflects further eutrophication-induced reduction of skeletal production leading to drowning of the system. This sequence was uplifted and subaerially exposed within the latest Cenomanian, but the overlying Lower Turonian ammonite marls and laminites reflect culmination of eutrophication and hypoxia of the OAE-2. This chain of events (Fig. 20) tracks the gradual reduction in production of skeletal carbonate through the Late Cenomanian owing to increased eutrophication.
(4)
(5)
Summary (1)
(2)
(3)
Faciestypesofthe Cenomanian– Turoniansuccession of northern Israel are highly variable, and indicate environments ranging from basinal, outer, mid- and inner ramp, to peritidal and subaerially exposed. Facies units are organized into two orders of sedimentary cycles. High-order cycles may be shallowingor deepening-upward, and are usually up to a few metres in thickness. Most high-order cycles are stacked into low-order cycles. Low-order cycles reflect progradational, aggradational, or retrogradational trends and homoclinal and distally steepened ramp profiles. The cycles construct three Cenomanian and one Turonian sequence. The origin of major Cenomanian–Turonian depositional events that can be correlated across the Arabian plate and Europe is eustatic. The Late Cenomanian eustatic rise was, to some extent, decoupled and masked by local tectonism and uplift in northern Israel. Sedimentary overprints of bottom winnowing, oxygenation of hypoxic sediments, eutrophication, or hypoxia reflect palaeoceanographic influence on some of the depositional events. Eustatic and/or palaeoenvironmental imprints were revealed for the first Cenomanian subaerial exposure; Early Cenomanian maximum flooding and oxygenation of hypoxic sea-floor; Middle Cenomanian highstand progradation followed by forced-regression and masstransport toward the basin; Middle Cenomanian subaerial exposure; increased Late Cenomanian eutrophication throughout sea-level rise; Late Cenomanian subaerial exposure; latest Cenomanian –Turonian eutrophication, and anoxia of the OAE-2 (‘Bonarelli’) event. Two episodes of carbonate ramp demise were recorded: (a) an abrupt deterioration of productive prograding ramp in the
(6)
165
Middle Cenomanian by sea-level fall, gravity collapse, subaerial exposure, and transgression and (b) deterioration of a non-productive aggradational homoclinal ramp by sea-level fall, faulting, subaerial exposure and drowning in the Late Cenomanian. Throughout the Late Cenomanian, the carbonate system of Galilee experienced a continuous decrease in skeletal grains production owing to increasing eutrophication. These conditions culminated in the Early Turonian by extreme hypoxia related to the OAE-2. Facies-thickness trends of Cenomanian – Turonian systems-tracts show that southern Carmel and much of Galilee were elevated during the Cenomanian and Turonian stages. Galilee and southern Carmel palaeo-highs were separated by a subsiding trough extending fromcentral–northernCarmeltowardsouthern Galilee. The Galilee palaeo-high was bounded at the north by a deeper zone of rapid subsidence, extending across northernmost Galilee and into Lebanon. This sedimentary configuration represents a shift of the north– south depositional strike of the mid-Cretaceous Levantine Hinge Belt toward the east –NE. This trend corresponds to other large-scale tectono-depositional features in the Mesozoic Levant. Late Cenomanian normal faults and a latest Cenomanian SB reflect tectonic activity and uplift of Galilee to above sea level near the end of the Cenomanian. Late Cenomanian faulting was associated with transformation of a Late Cenomanian homoclinal ramp into a steep margin. Locally, calcarenitic clinoform unit-formed steep slope beyond this margin, facing to the south–SW. A concomitant, turbiditic toe-of-slope developed distally in Carmel to the south.
This research was done in the framework of a PhD thesis of the senior author at Ben-Gurion University of the Negev, funded by the Geological Survey of Israel and the Earth Sciences Administration, Israel Ministry of National Infrastructures. The authors thank Z. Lewy (Geological Survey of Israel) for identifying macrofossils and M. Simmons (Neftex) for identifying orbitolinid foraminifers. The authors especially acknowledge the insightful and constructive contribution of the reviewers and editors. Y. Raphael (GSI) and D. Kusashvili (BGU) provided logistic assistance in the field and laboratory.
References Abed, A. M. 1984. Emergence of Wadi Mujib (Central Jordan) during lower Cenomanian time and its regional tectonic implications. In: Dixon, J. E. & Robertson,
166
R. FRANK ET AL.
A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 213–216. Aguilera-Franco, N. & Allison, P. 2005. Events of the Cenomanian–Turonian succession, Southern Mexico. Journal of Iberian Geology, 31, 25–50. Aguilera-Franco, N., Herna´ndez-Romano, U. & Allison, P. 2001. Biostratigraphy and environmental changes across the Cenomanian– Turonian boundary, Southern Mexico. Journal of South American Earth Sciences, 14, 237–255. Al-Juboury, A., Al-Zoobay, B. & Al-Juwainy, Q. 2006. Foraminifera as Palaeoenvironmental Indicators in Albian– Cenomanian Carbonates, NE Iraq. International Symposium on Foraminifera, Natal, RN, Brazil, 29, 317–318. Arthur, M. A., Schlanger, S. O. & Jenkyns, H. C. 1987. The Cenomanian– Turonian Oceanic Anoxic Event, II. Palaeoceanographic controls on organicmatter production and preservation. In: Brooks, J. & Fleet, A. (eds) Marine Petroleum Source Rocks. Geological Society, London, Special Publications, 26, 401–420. Avnimelech, M. A. 1965. Sur la pre´sence de Hyphoplites falcatus (Mantell) (Ammonoidea: Hoplitidae) dans le Ce´nomanien infe´rieur du Carmel (Israe¨l). Comptes rendus des se´ances de la Socie´te´ Ge´ologique de France, 5, 160– 161. Bachmann, M. & Kuss, J. 1998. The middle Cretaceous carbonate ramp of the northern Sinai: sequence stratigraphy and facies distribution. In: Wright, V. P. & Burchette, T. P. (eds) Carbonate Ramps. Geological Society, London, Special Publications, 149, 253– 280. Banner, F. T. 1972. Pithonella ovalis from the Early Cenomanian of England. Micropaleontology, 18, 278– 284. Baraboshkin, E. J., Kopaevich, L. F. & Olferiev, A. G. 1998. The Mid-Cretaceous events in Eastern Europe: development and palaeogeographical significance. In: Crasquin-Soleau, S. & Barrier, E. (eds) PeriTethys Memoir 4: Epicratonic Basins of Peri-Tethyan Platforms. Me´moires du Muse´um National d’Histoire Naturelle, Paris, 179, 93– 110. Bauer, J., Marzouk, A. M., Steuber, T. & Kuss, J. 2001. Lithostratigraphy and biostratigraphy of the Cenomanian–Santonian strata of Sinai, Egypt. Cretaceous Research, 22, 497– 526. Bauer, J., Kuss, J. & Steuber, T. 2003. Sequence architecture and carbonate platform configuration (Late Cenomanian–Santonian, Sinai, Egypt). Sedimentology, 50, 387–414. Bein, A. 1971. Rudistid reef complexes (Albian to Cenomanian) in the Carmel and the coastal plain, Israel. Institute for Petroleum Research and Geophysics Report, 1051. Bein, A. 1974. Reef developments in the Judea Group from the Coastal Plain and the Carmel, Israel. Geological Survey of Israel Report, OD/74/1, 154 pp. (in Hebrew, English abstract). Bein, A. 1976. Rudistid fringing reefs of Cretaceous shallow carbonate platform of Israel. American Association of Petroleum Geologists Bulletin, 60, 258– 272.
Bein, A. 1977. Shelf– basin sedimentation: mixing and diagenesis of pelagic and clastic Turonian carbonates, Israel. Journal of Sedimentary Petrology, 47, 382–391. Bein, A. & Gvirtzman, G. 1977. A Mesozoic fossil edge of the Arabian plate along the Levant coastline and its bearing on the evolution of the eastern Mediterranean. In: Biju-Duval, B. & Montadert, I. (eds) Structural History of the Mediterranean Basins. Editions Technip, Paris, 95– 110. Bein, A. & Weiler, Y. 1976. The Cretaceous Talme-Yafe Formation: a contour current shaped sedimentary prism of calcareous detritus at the continental margin of the Arabian Craton. Sedimentology, 23, 511– 532. Bogoch, R., Buchbinder, B. & Magaritz, M. 1994. Sedimentology and geochemistry of lowstand peritidal lithofacies at the Cenomanian –Turonian boundary in the Cretaceous carbonate platform of Israel. Journal of Sedimentary Research, 64, 733– 740. Borgomano, J. R. F. 2000. The Upper Cretaceous carbonates of the Gargano-Murge region, southern Italy: a model of platform-to-basin transition. American Association of Petroleum Geologists Bulletin, 84, 1561– 1588. Bouma, A. H. 1962. Sedimentology of some Flysch Deposits: A Graphic Approach to Facies Interpretation. Elsevier, Amsterdam. Brew, G., Barazangi, M., Al-Maleh, A. K. & Sawaf, T. 2001. Tectonic and geologic evolution of Syria. GeoArabia, 4, 573– 616. Buchbinder, B., Benjamini, C. & Lipson-Benitah, S. 2000. Sequence development of Late Cenomanian– Turonian carbonate ramps, platforms and basins in Israel. Cretaceous Research, 21, 813–843. Burchette, T. P. & Wright, V. P. 1992. Carbonate ramp depositional systems. Sedimentary Geology, 79, 1 –41. Carson, G. A. & Crowley, S. F. 1993. The glauconite– phosphate association in hardgrounds: examples from the Cenomanian of Devon, SW England. Cretaceous Research, 14, 69– 89. Carter, J. M. L. & Gillcrist, J. R. 1994. Karstic reservoirs of the Mid-Cetaceous Mardin Group, SE Turkey: tectonic and eustatic controls on their genesis, distribution and preservation. Journal of Petroleum Geology, 17, 253–278. Caus, E., Teixel, A. & Bernaus, J. M. 1997. Depositional model of a Cenomanian– Turonian extensional basin (Sopiera Basin, NE Spain): Interplay between tectonic, eustasy and biological productivity. Palaeogeography, Palaeoclimatology, Palaeoecology, 129, 23–36. Di Lucia, M., Parente, M. & Frijia, G. 2007. The Orbitolina level of southern Apennines: a tale of nutrient fluctuations and stratigraphic condensation. European Geosciences Union, Geophysical Research Abstracts, 9, 1607– 7962. Dias-Brito, D. 2000. Global stratigraphy, palaeobiogeography and palaeoecology of Albian-Maastrichtian pithonellid calcispheres: impact on Tethys configuration. Cretaceous Research, 21, 315– 349. Drzewiecki, P. A. & Simo, J. A. 1997. Carbonate platform drowning and oceanic anoxic events on a Mid-Cretaceous carbonate platform, South-Central Pyrenees, Spain. Journal of Sedimentary Research, 67, 698– 714.
THE CARBONATE SYSTEM OF NORTHERN ISRAEL Dunham, R. J. 1962. Classification of carbonate rocks according to depositional texture. In: Ham, W. E. (ed.) Classification of Carbonate Rocks. American Association of Petroleum Geologists Memoir, 1, 108–121. Embry, A. F. 2002. Transgressive– Regressive (T –R) Sequence Stratigraphy. Proceedings of the 22nd Annual Gulf Coast Section SEPM Foundation Bob F. Perkins Research Conference, 151– 172. Embry, A. F. & Klovan, J. E. 1971. A Late Devonian reef tract on northeastern Banks Island, NW Territories. Bulletin of Canadian Petroleum Geology, 19, 730–781. Ferry, S., Merran, Y., Grosheny, D. & Mroueh, M. 2007. The Cretaceous of Lebanon in the Middle East (Levant) context. Notebooks on Geology, Memoir, 2007/02 (Abstract 8), 38– 42. Flexer, A., Rosenfeld, A., Lipson-Benitah, S. & Honigstein, A. 1986. Relative sea-level changes during the Cretaceous in Israel. American Association of Petroleum Geologists Bulletin, 70, 1685–1699. Folkman, Y. 1969. Diagenetic dedolomitization in the Albian–Cenomanian Yagur dolomite on Mount Carmel (northern Israel). Journal of Sedimentary Petrology, 39, 380– 385. Freund, R. 1958. The geology of the Yirka-Peqi’in region. MSc thesis, Hebrew University, Jerusalem (in Hebrew). Freund, R. 1959. On the stratigraphy and tectonics of the Upper Cretaceous in Western Galilee. Bulletin of the Research Council of Israel, 8G, 43–49. Freund, R. 1965. Upper Cretaceous reefs in Northern Israel. Israel Journal of Earth Sciences, 14, 108– 121. Freund, R. & Raab, M. 1969. Lower Turonian Ammonites from Israel. Palaeontological Association, London, Special Papers in Palaeontology, 4, 83. Gale, A. S., Kennedy, W. J., Burnett, J. A., Caron, M. & Kidd, B. E. 1996. The late Albian to early Cenomanian succession at Mont Risou near Rosans (Droˆme, SE France): an integrated study (ammonites, inoceramids, planktonic foraminifera, nannofossils, oxygen and carbon isotopes). Cretaceous Research, 17, 515–606. Gardosh, M., Druckman, Y., Buchbinder, B. & Rybarkov, M. 2006. The Levant Basin Offshore Israel: Stratigraphy, Structure, Tectonic Evolution and Implications for Hydrocarbon Exploration. Geological Survey of Israel Report, 14/2006, 119. Garfunkel, Z. 1998. Constraints on the origin and history of the Eastern Mediterranean basin. Tectonophysics, 298, 5 –35. Garfunkel, Z. & Derin, B. 1984. Permian-early Mesozoic tectonism and continental margin formation in Israel and its implications for the history of the Eastern Mediterranean. In: Dixon, J. E. & Robertson, A. H. F. (eds) The Geological Evolution of the Eastern Mediterranean. Geological Society, London, Special Publications, 17, 187 –201. Gebhardt, H., Kuhnt, W. & Holbourn, A. 2004. Foraminiferal response to sea-level change, organic flux and oxygen deficiency in the Cenomanian of the Tarfaya basin, southern Morocco. Marine Micropalaeontology, 53, 133–157.
167
Glikson, Y. A. 1966. Geology of southern Naftali Mountains (northeastern Galilee, Israel). Israel Journal of Earth Sciences, 15, 135– 154. Gradstein, F. M., Ogg, J. G. et al. 2004. A Geologic Timescale 2004. Cambridge University Press, Cambridge. Gre´laud, C., Razin, P., Homewood, P. W. & Schwab, A. N. 2006. Development of incisions on a periodically emergent carbonate platform (Natih Formation, Late Cretaceous, Oman). Journal of Sedimentary Research, 76, 647 –669. Gro¨ke, D. R., Ludvigson, G. A., Witzke, B. L., Robinson, S. A., Joeckel, R. M., Ufnar, D. F. & Ravn, R. L. 1988. Recognizing the Albian– Cenomanian (OAE1d) sequence boundary using plant carbon isotopes: Dakota Formation, Western Interior Basin, USA. Geology, 34, 193–196. Gusˇic, I. & Jelaska, V. 1993. Upper Cenomanian–Lower Turonian sea-level rise and its consequences on the Adriatic– Dinaric carbonate platform. Geologische Rundschau, 82, 676–686. Gvirtzman, G. & Klang, A. 1972. A structural and depositional hinge-line along the Coastal Plain of Israel, evidenced by magnetotellurics. Geological Survey of Israel Bulletin, 55, 18. Hallock, P. 1988. The role of nutrient availability in bioerosion: consequences to carbonate buildups. Palaeogeography, Palaeoclimatology, Palaeoecology, 63, 275 –291. Hancock, J. M. 2003. Lower sea levels in the Middle Cenomanian. Notebooks on Geology, Letter 2003/02, 1– 6. Hancock, J. M. & Kaufman, E. G. 1979. The great transgressions of the Late Cretaceous. Journal of the Geological Society, 136, 175 –186. Heck, P. R., Frank, M., Anselmetti, F. S. & Kubik, P. W. 2007. Origin and age of submarine ferromanganese hardgrounds from the Marion Plateau, offshore NE Australia. In: Anselmetti, F. S., Isern, A. R., Blum, P. & Betzler, C. (eds) Proceedings of the Ocean Drilling Program, Scientific Results, 194, 1–22. Helland-Hansen, W. & Gjelberg, J. 1994. Conceptual basis and variability in sequence stratigraphy: a different perspective. Sedimentary Geology, 92, 1 –52. Honigstein, A., Lipson-Benitah, S., Conway, B., Flexer, A. & Rosenfeld, A. 1989. Mid-Turonian anoxic event in Israel – a multidisciplinary approach. Palaeogeography, Palaeoclimatology, Palaeoecology, 69, 103 –112. Hottinger, L. 1997. Shallow benthic foraminiferal assemblages as signals for depth of their deposition and their limitations. Bulletin de la Societe´ Ge´ologique de France, 168, 491–505. Hunt, D. & Tucker, M. E. 1992. Stranded parasequences and the forced regressive wedge systems tract: deposition during base level fall. Sedimentary Geology, 81, 1 –9. Hunt, D. & Tucker, M. E. 1993. Sequence stratigraphy of carbonate shelves with an example from the Mid-Cretaceous (Urgonian) of SE France. In: Posamentier, H. W., Summerhayes, C. P., Haq, B. U. & Allen, G. P (eds) Sequence Stratigraphy and Facies. International Association of Sedimentologists, Special Publications, 18, 307–341.
168
R. FRANK ET AL.
Immenhauser, A., Schlager, W. et al. 1999. Late Aptian to Late Albian sea-level fluctuations constrained by geochemical and biological evidence (Nahr Umr Formation, Oman). Journal of Sedimentary Research, 69, 434–466. Immenhauser, A., Creusen, A., Esteban, M. & Vonhof, H. B. 2000. Recognition and interpretation of polygenic discontinuity surfaces in the Middle Cretaceous Shu’aiba, Nahr Umr, and Natih Formations of Northern Oman. GeoArabia, 5, 299–322. Jarvis, I., Carson, G. A., Cooper, M. K. E., Hart, M. B., Leary, P. N., Tocher, B. A., Horne, D. & Rosenfeld, A. 1988. Microfossil assemblages and the Cenomanian–Turonian (late Cretaceous) oceanic anoxic event. Cretaceous Research, 9, 3– 103. Jenkyns, H. C. 1991. Impact of Cretaceous sea-level rise and anoxic events on the Mesozoic carbonate platform of Yugoslavia. American Association of Petroleum Geologists Bulletin, 75, 1007–1017. Kafri, U. 1972. Lithostratigraphy and environments of deposition, Judea Group, western and central Galilee, Israel. Geological Survey of Israel Bulletin, 54, 56. Kafri, U. 1986. The Late Albian to Early Cenomanian sedimentary break in northern Israel. Israel Journal of Earth Sciences, 35, 211–219. Kafri, U. 1991. Lithostratigraphy of the Judea Group in eastern Galilee, emphasizing the Naftali Mountains. Geological Survey of Israel Report, 24/91. Karcz, Y. 1959. The structure of the northern Carmel. Bulletin of the Research Council of Israel, 8G, 119– 130. Kashai, A. 1966. The geology of the eastern and southwestern Carmel. PhD thesis, Hebrew University, Jerusalem, Israel (in Hebrew, English abstract). Kassab, A. S. & Obaidalla, N. A. 2001. Integrated biostratigraphy and inter-regional correlation of the Cenomanian–Turonian deposits of Wadi Feiran, Sinai, Egypt. Cretaceous Reseach, 22, 105 –104. Kennedy, W. J. & Jolkicˇev, N. 2004. Middle Cenomanian ammonites from the type section of the Sanandinovo Formation of northern Bulgaria. Acta Geologica Polonica, 54, 369– 380. Kidwell, S. M. 1986. Models for fossil concentrations: palaeobiologic implications. Palaeobiology, 12, 6– 24. Lewy, Z. 1990. Transgressions, regressions and relative sea-level changes on the Cretaceous shelf of Israel and adjacent countries. A critical evaluation of Cretaceous sea-level correlations. Paleoceanography, 5, 619– 637. Lewy, Z. & Avni, Y. 1988. Omission surfaces in the Judea Group, Makhtesh Ramon region, southern Israel, and their palaeogeographic significance. Israel Journal of Earth Sciences, 37, 105– 113. Lewy, Z. & Raab, M. 1978. Mid-Cretaceous stratigraphy of the Middle-East. Annales Muse´e Histoire Naturelle, Nice, 4, 1 –20. Lewy, Z. & Weissbrod, T. 1993. Stratigraphy of the Cretaceous in ‘Makhtesh Hatira’. Israel Geological Society Annual Meeting, Arad, Field Trip Guidebook, 1 –13 (in Hebrew). Lipson-Benitah, S., Almogi-Labin, A. & Sass, E. 1997. Cenomanian biostratigraphy and palaeoenvironments in the NW Carmel region, northern Israel. Cretaceous Research, 18, 469–491.
Longman, M. W. 1980. Carbonate diagenetic textures from nearsurface diagenetic environments. American Association of Petroleum Geologists Bulletin, 64, 461–487. Mouty, M., Al Maleh, A. K. & Abou Laban, H. 2003. Le Cre´tace´ moyen de la chaıˆn des Palmyrides (Syrie centrale). Geodiversitas, 25, 429– 443. Philip, J. M. & Airaud-Crumiere, C. 1991. The demise of the rudist-bearing carbonate platforms at the Cenomanian/Turonian boundary: a global control. Coral Reefs, 10, 115–125. Philip, J., Borgomano, J. & Al-Maskiry, S. 1995. Cenomanian– Early Turonian carbonate platform of Northern Oman: Stratigraphy and palaeoenvironments. Palaeogeography, Palaeoclimatology, Palaeoecology, 119, 77–92. Picard, L. & Kashai, E. 1958. On the lithostratigraphy and tectonics of the Carmel. Bulletin of the Research Council of Israel, 7G, 1 –19. Pittet, B., Van Buchem, F. S. P., Hillga¨rtner, H., Gro¨tsch, J., Razin, P. & Droste, H. 2002. Ecological succession, palaeoenvironmental change, and depositional sequences of Barremian–Aptian shallow-water carbonates in northern Oman. Sedimentology, 49, 555–581. Plint, A. G. 1988. Sharp-based shoreface sequences and “offshore bars” in the Cardium formation of Alberta; their relationship to relative changes in sea-level. In: Wilgus, C. K., Hasting, B. S., Ross, C. A., Posamentier, H., Van Wagoner, J. & Kendall, C. G. St. C. (eds) Sea-level Changes: An Integrated Approach. Society of Economic Paleontologists and Mineralogists, Special Publications, 42, 357–370. Ponikarov, V. P., Kazmin, V. G. et al. 1967. The Geological Map of Syria, Scale 1:1,000,000, Explanatory Notes. Technoexport USSR, 111. Porrenga, D. H. 1967. Glauconite and chamosite as depth indicators in the marine environment. Marine Geology, 5, 495– 501. Posamentier, H. W. & Morris, W. R. 2000. Aspects of the stratal architecture of forced regressive deposits. In: Hunt, D. & Gawthorpe, R. L. (eds) Sedimentary Responses to Forced Regressions. Geological Society, London, Special Publications, 172, 19–46. Read, J. F. 1985. Carbonate platform facies models. American Association of Petroleum Geologist Bulletin, 69, 1 –21. Robaszynski, F., Gale, A., Juignet, P., Amedro, F. & Hardendol, J. 1998. Sequence stratigraphy in the Upper Cretaceous series of the Anglo-Paris basin: exemplified by the Cenomanian stage. In: De Graciansky, P. C., Hardenbol, J., Jacquin, T. & Vail, P. R. (eds) Mesozoic and Cenozoic Sequence Stratigraphy of European Basins. Society for Sedimentary Geology, Special Publications, 60, 363– 386. Robertson, A. H. F. 1998. Mesozoic–Tertiary tectonic evolution of the eastern Mediterranean area: integration of marine and land evidence. In: Robertson, A. H. F., Emeis, K. C., Richter, C. & Camerlenghi, A. (eds) Proceedings of the Ocean Drilling Program, Scientific Results, 160, 723–782. Saint-Marc, P. 1972. Le Cre´tace´ infe´rieur et moyen du bord occidental du Jabal Sannine (Liban). Notes et Me´moires sur le Moyen-Orient, 12, 217–226.
THE CARBONATE SYSTEM OF NORTHERN ISRAEL Saint-Marc, P. 1974. E´tude stratigraphique et micropale´ontologique de l’Albien, du Ce´nomanien et du Turonien du Liban. Notes et Me´moires sur le MoyenOrient, 13, 342. Sass, E. 1980. Late Cretaceous volcanism of Mount Carmel, Israel. Israel Journal of Earth Sciences, 29, 8–24. Sass, E. & Bein, A. 1982. The Cretaceous carbonate platform in Israel. Cretaceous Research, 3, 135– 144. Schlager, W. 1999. Type 3 sequence boundaries. In: Harris, P., Saller, A. & Simo, A. (eds) Carbonate Sequence Stratigraphy: Application to Reservoirs, Outcrops and Models. Society for Sedimentary Geology, Special Publications, 63, 35– 46. Schulze, F., Lewy, Z., Kuss, J. & Gharaibeh, A. 2003. Cenomanian– Turonian carbonate platform deposits in west central Jordan. International Journal of Earth Sciences, 92, 641– 660. Schulze, F., Marzouk, M., Bassiouni, M. A. A. & Kuss, J. 2004. The Late Albian–Turonian carbonate platform succession of west-central Jordan: stratigraphy and crises. Cretaceous Research, 25, 709–737. Schulze, F., Kuss, J. & Marzouk, M. 2005. Platform configuration, microfacies and cyclicities of the Upper Albian to Turonian of west-central Jordan. Facies, 50, 505– 527. Scott, R. W., Schlager, W., Fouke, B. & Nederbragt, S. A. 2000. Are Mid-Cretaceous eustatic events recorded in Middle East Carbonate platforms? In: Alsharan, A. S. & Scott, R. W. (eds) Middle East models of Jurassic/Cretaceous Carbonate Systems. Society for Sedimentary Geology, Special Publications, 69, 77– 88. Segev, A. & Sass, E. 2006. The Geology of the Central and Southern Carmel Region. Israel Geological Society Annual Meeting, Beit-Shean, Field Trip Guidebook, 69–88 (in Hebrew). Segev, A., Sass, E., Ron, H., Lang, B., Kolodny, Y. & McWilliams, M. 2002. Stratigraphic, geochronologic and palaeomagnetic constraints on Late Cretaceous volcanism in northern Israel. Israel Journal of Earth Sciences, 51, 297– 309. Shadmon, A. 1959. The Bina limestone. Geological Society of Israel Bulletin, 24, 4. Sharland, P. R., Archer, R. et al. 2001. Arabian Plate sequence stratigraphy. GeoArabia Special Publications, Bahrain, 2, 371. Simmons, M. D., Williams, C. L. & Hart, M. B. 1991. Sea-level changes across the Albian-Cenomanian boundary in SW England. Proceedings of the Ussher Society, 7, 408– 412. Simmons, M. D., Whittaker, J. E. & Jones, R. W. 2000. Orbitolinids from the Cretaceous sediments of the Middle East – a revision of the F. R. S. Henson and Associates Collection. In: Hart, M. B., Kaminsky, M. A. & Smart, C. W. (eds) Proceedings of the 5th International Workshop on Agglutinated Foraminifera. Grzybowski Foundation Special Publications, 7, 411–437. Sneh, A. 2002. Lithostratigraphic relationships between the Cenomanian Deir-Hanna and Sakhnin formations
169
in the Galilee in the light of recent geological mapping. Israel Journal of Earth Sciences, 51, 103– 116. Sneh, A. & Weinberger, R. 2003. Geology of the Metulla Quadrangle, northern Israel: Implications for the offset along the Dead Sea Rift. Israel Journal of Earth Sciences, 52, 123–138. Sneh, A., Bartov, Y. & Rosensaft, M. 1998. Geological Map of Israel 1:200000. Geological Survey of Israel, Jerusalem, Sheet 1-North. Tro¨ger, K. A. 2003. The Cretaceous of the Elbe valley in Saxony (Germany) – a review. Notebooks on Geology, Article 2003/03, 1 –14. Van Buchem, F. S. P., Razin, P. et al. 1996. High resolution sequence stratigraphy of the Natih Formation (Cenomanian/Turonian) in northern Oman: distribution of source rocks and reservoir facies. GeoArabia, 1, 65–91. Van Buchem, F. S. P., Razin, P., Homewood, P. W., Oterdoom, W. H. & Philip, J. 2002. Stratigraphic organization of carbonate ramps and organic-rich intrashelf basins: Natih Formation (Middle Cretaceous) of northern Oman. American Association of Petroleum Geologists Bulletin, 86, 21–53. Vilas, L., Masse, J. P. & Arias, C. 1995. Orbitolina episodes in carbonate platform evolution: the early Aptian model from SE Spain. Palaeogeography, Palaeoclimatology, Palaeoecology, 119, 35– 45. Voigt, S., Gale, A. S. & Voigt, T. 2006. Sea level change, carbon cycling and palaeoclimate during the Late Cenomanian of NW Europe; an integrated palaeoenvironmental analysis. Cretaceous Research, 27, 836 –858. Wald, R. 2004. Stratigraphy and paleoecology of the Avnon/Tamar cycle, northern Negev, Israel. MSc thesis, Ben-Gurion University of the Negev, BeerSheva (in Hebrew, English abstract). Walley, D. W. 1998. Some outstanding issues in the geology of Lebanon and their importance in the tectonic evolution of the Levantine region. Tectonophysics, 298, 37–62. Weiler, Y. 1968. Geology of Nazareth Hills and Mount Tabor (Southern Galilee, Israel). Israel Journal of Earth Sciences, 17, 63–82. Wendler, J., Grafe, K. U. & Willems, H. 2002. Palaeoecology of calcareous dinoflagellate cysts in the MidCenomanian Boreal Realm: Implications for the reconstruction of palaeoceanography of the NW European shelf sea. Cretaceous Research, 23, 213– 229. Wilmsen, M. 2000. Evolution and demise of a MidCretaceous carbonate shelf: the Altamira limestones (Cenomanian) of northern Cantabria (Spain). Sedimentary Geology, 133, 195– 226. Wilmsen, M. 2008. An Early Cenomanian (Late Cretaceous) maximum flooding bioevent in NW Europe: Correlation, sedimentology and biofacies. Palaeogeography, Palaeoclimatology, Palaeoecology, 258, 317– 333. Ziegler, M. A. 2001. Late Permian to Holocene paleofacies evolution of the Arabian plate and its Hydrocarbon occurrences. GeoArabia, 6, 445– 504.
Orbital time scale, intra-platform basin correlation, carbon isotope stratigraphy and sea-level history of the Cenomanian – Turonian Eastern Levant platform, Jordan JENS E. WENDLER1,2*, JENS LEHMANN1 & JOCHEN KUSS1 1
Department of Geosciences, Bremen University, P.O. Box 330440, 28334 Bremen, Germany 2
Present address: Smithsonian Institution National Museum of Natural History, 10th & Constitution NW, Washington, DC 20560-0121, USA *Corresponding author (e-mail:
[email protected]) Abstract: Two Cenomanian– Turonian boundary (CTBE) sections (KB3 and GM3) of the Karak– Silla intra-platform basin of the Eastern Levant carbonate platform, Jordan, are correlated based on high-resolution calcimetry. KB3 contains black shales with over 7 wt% total organic carbon (TOC). GM3 was deposited at shallower water depth and reveals four conspicuous gypsum beds used for sea-level reconstruction. Spectral analysis of carbonate content and TOC reveals forcing, mainly by the 100 ka cycle of Earth’s orbit eccentricity. Whole rock stable carbon isotope data show a conspicuous positive d13C excursion representing the Oceanic Anoxic Event 2 (OAE2). The carbon isotope records of KB3 and GM3 correspond well with the cycles in the d13C record of the global stratotype (GSSP) at Pueblo (USA). The GSSP orbital timescale, thus, can be applied to the Jordan record. Furthermore, all stable isotope events defined in the English chalk reference record are recognized in Jordan. Our orbital model for the Jordan sequence-stratigraphical framework reveals approximately 1.2 (þ0.2) Ma duration of a third-order sequence, proposed to represent one cycle of the long obliquity (1.2 Ma). This longterm period is superimposed on three fourth-order fluctuations of 400 ka length (long eccentricity; fourth-order sea-level fluctuations), each of which comprises four carbonate cycles (100 ka eccentricity; fifth-order sea-level fluctuations). Demise of the Levant platform occurred during the phase of decreasing d13C values after OAE2 in the interval between the Cenomanian –Turonian (C– T) boundary and the end of the Early Turonian.
The Levant carbonate platform deposits of central Jordan represent a textbook-like shallow-marine platform setting subdivided into intra-platform basins during Cenomanian–Turonian (C– T) times (Kuss et al. 2003). The morphological structuring by these basins induced lithologically highly variable successions, especially during black shale deposition that preferably occurred in deeper subbasins. A correlation of these different successions is possible by means of high-resolution calcimetry and stable carbon isotope stratigraphy. In this paper we present such a correlation that enables the study of three research aspects: a) linking different palaeoenvironments over the C –T boundary interval; b) refining the sequence-stratigraphic model by constructing an orbital time scale; c) correlating the successions to the global record (exemplified for the well-dated C– T boundary interval) in order to eliminate ambiguities in the local stratigraphy. Furthermore, our data support the global picture and estimates of duration of the C –T boundary interval that includes the global Oceanic Anoxic Event 2 (OAE2).
The duration of OAE2 has been the matter of many studies in recent years resulting in a range of values from 320 to 960 ka (Obradovitch 1993; Sageman et al. 2006) partly caused by considering different intervals. An orbital timescale for the Pueblo stratotype section was given by Sageman et al. (2006) providing a precise time measure for the period investigated in our study. In addition, a comprehensive study of stable carbon isotope records in Europe was given by Jarvis et al. (2006) providing a detailed set of isotope events suitable for global correlation. Recently, an orbital model was also presented for the Wunstorf (Germany) section (Voigt et al. 2008). This record correlates well with the present results and can be used to support the new hypothesis put forward in the present paper regarding the orbital trigger of third-order sequences.
Geological setting The Eastern Levant carbonate platform in Jordan is characterized by 300–400 m thick successions of
From: Homberg, C. & Bachmann, M. (eds) Evolution of the Levant Margin and Western Arabia Platform since the Mesozoic. Geological Society, London, Special Publications, 341, 171–186. DOI: 10.1144/SP341.8 0305-8719/10/$15.00 # The Geological Society of London 2010.
172
J. E. WENDLER ET AL.
nodular limestone, massive limestone and laminated limestone with intercalated clay, marl and gypsum beds that were deposited during the Cenomanian and Turonian. According to Schulze et al. (2004), carbonate platform demise occurred during the Late Cenomanian and Early Turonian when the deposition of marl and clay became dominant. The sections GM3 (Ghawr Al Mazar) and KB3 (Kuthrubbah) are outcrop sections, about 30 km apart, in the Wadi system cutting east– west into an extended plateau area east of the Dead Sea (Fig. 1). Palaeogeographically the NW-deepening Levant carbonate platform extended over the passive margin of the Arabo– Nubian shield during C –T times. The sections studied comprise the Upper Cenomanian Hummar Formation and the Upper Cenomanian to Middle Turonian Shueib Formation, which are part of the Ajlun Group (Powell 1989). They represent deposits of an intra-platform basin (Karak-Silla basin) at c. 100 km distance from the palaeocoastline of the Arabian Shield. Schulze et al. (2003) report a range of facies representing supratidal to shallow subtidal deposits, based on the analysis of a diverse shallow-water benthic association including calcareous algae, rudists, larger benthic foraminifera, oysters and ostracodes of brackish to hypersaline environments (see Morsi & Wendler 2010). The intra-platform basin was connected to the open marine environment and only temporarily experienced restricted conditions (formation of evaporites) during regressions. GM3 had a marginal position, while KB3 represents deeper parts of the Karak –Silla intra-platform basin. The described facies indicate prolonged phases of low activity of the platform carbonate factory from the Late Cenomanian to Early Turonian. Despite the general carbonate platform setting, the profiles investigated here show only short periods of normal carbonate production, while most parts of the section consist of marls and clays. So, the material represents a depositional system with relatively high siliciclastic input.
Material and methods For this paper we focus on the interval section metre 47–83 of section GM3 (Ghawr Al Mazar: 318150 3400 N; 358350 4100 E) representing the OAE2. It is part of a mid-Cenomanian to Lower Turonian section published in separate publications (see Morsi & Wendler 2010). The section develops from green clays and marls into a unit of platy, bituminuous limestone beds, followed by brown marly clays, grey marls and limestone at the top. A total of 155 samples were collected from the GM3 section interval 47– 83 m at sample spacing of 10 –25 cm.
Section KB3 (Kuthrubbah: 318090 1300 N; 358360 0600 E) is 25 m thick and comprises an alternation of black shales and platy, bituminuous limestone beds, and shows a massive limestone at the top. An interval of about 2.5 m below this topmost limestone could not be sampled owing to poor outcrop conditions. 117 samples were taken with 10 –20 cm sample spacing (section metre 0 –16 m) and 30 –50 cm (above 16 m). Bulk samples were crushed with an agate mortar. The measurements of the carbonate content and the total organic carbon (TOC) were performed at the Alfred Wegener Institute Bremerhaven, Germany, using a LECO CS-125 carbon –sulphur determinator. For total carbon a LECO CNS-2000 was used. Stable carbon isotopes were measured on bulk carbonate at the isotope laboratory of Bremen University, Germany, using a Finigan MAT 251 mass spectrometer. The results are reported relative to the V-PDB standard. Thin sections were prepared from limestone samples for microfacies analysis. Standard smear slides were used for determination of coccolith assemblages.
Results Biostratigraphy The integrated biostratigraphic framework of Schulze et al. (2003) using nannofossils and ammonites, supported partly by larger benthic foraminifera, forms the stratigraphic basis (Fig. 2). The sections comprise nannoplankton zones CC10 and CC11. For details on index species see Schulze et al. (2003). A revision of the nannofossil content of all correlated sections previously studied and new analyses resulted in a repositioning of the zone boundary (Fig. 2) based on earlier appearances of the Turonian index fossil Quadrum gartneri already in the section part formerly placed in the Late Cenomanian by Schulze et al. (2003), taken parallel to the new GM3 section. The ammonite occurrences in the C –T of Jordan (Schulze et al. 2004) provide a more detailed biostratigraphy that can be correlated with ammonite zone schemes of Southern Europe (Hardenbol et al. 1998), Israel (Lewy 1989, 1990), and the Middle East in general (Lewy & Raab 1976). In conjunction with the isotope record presented here it enables a good time control. The Hummar Formation (section metre 47– 56 m, Fig. 2) contains abundant Neolobites vibrayeanus, encountered also in the present material at section metre 47.5 m. Wiese & Schulze (2005) stated that the stratigraphic range of N. vibrayeanus in Jordan is not yet clear, but as far as biostratigraphical control is given, the species range correlates approximately with the early Late
CENOMANIAN– TURONIAN ORBITAL TIMESCALE FOR THE LEVANT
(a)
36°00'
37°00'
173
(b) gypsum beds
SAUDI ARABIA
d Re
Amman
500
32°00'
a Se
0 Km
Dead
Sea
Madaba
Wala Limestone Member
Wadi Al Karak Al Karak KB 3 31°00' section
Wadi Musa
36°00'
0
25
u. Cenomanian
city major Wadi -300-0 m 0-300 m 300-600 m 600-900 m 900-1200 m
l. Turonian
Wadi Mujib
GM 3
Wadi Musa
middle Turonian
JORDAN
50 km
platy limestones
green clay unit
10 m
37°00'
(c)
West
East
Fig. 1. (a) Location of sections GM3 and KB3 in Jordan. (b) Investigated section part GM3 showing lithological markers. (c) Overview of the northern slope of Wadi Al Karak; frame marks the section enlarged in (b).
174 J. E. WENDLER ET AL. Fig. 2. Biostratigraphy, stable carbon isotope, carbonate and TOC content data of sections GM3 and KB3. Carbonate data connected by correlation lines; lines are extended toward the isotope curves at key points to show correspondence of d13C records. TOC, main correlated TOC maxima. Black shales in GM3 ¼ grey shaded intervals. Timescale in the carbonate panel of GM3 assumes 100 ka (eccentricity) duration of the numbered cycles, zero point according to Pueblo (see Fig. 3). Dolo., dolomite unit; TS/FS, transgressive/ flooding surface; mfs, maximum flooding surface (according to Schulze et al. 2003). Vertical bar marks OAE2 as indicated by Sageman et al. (2006). Stratigraphy panel includes the referred lithological units of GM3. m1, m3 – Ammonite marker beds. Positive d13C excursions within OAE2: a, b, c; negative d13C excursions: N1, N2.
CENOMANIAN– TURONIAN ORBITAL TIMESCALE FOR THE LEVANT
Cenomanian Calycoceras guerangeri zone (Lehmann in Wiese & Schulze 2005). Concerning the first occurrence of N. vibrayeanus, this is contrary to Schulze et al. (2003, fig. 7; 2004, fig. 3) who correlated it in Jordan with the middle of the Acanthoceras rhotomagense zone of the northern European zonation. N. vibrayeanus is accompanied by further ammonites in Jordan, Proeucalycoceras haugi, Pseudocalycoceras harpax and Turrilites acutus. This assemblage can be assumed to represent Middle to Late Cenomanian (pers. comm. Lewy in Schulze et al. 2004). P. haugi is a species of the early Late Cenomanian (Kennedy & Juignet 1994). P. harpax is also early Late Cenomanian (Kennedy & Juignet 1994), though the delineation between the type material from India and that from the Near East still needs clarification. T. acutus mainly occurs in the Middle Cenomanian, and can range into the Late Cenomanian (Juignet & Kennedy 1976). To summarize, the first appearance of N. vibrayeanus in Jordan occurs already in the Middle Cenomanian and the last occurrence is correlated with the top of the European Calycoceras guerangeri zone. Section part 56 –64 m contains the platy limestone beds described as ‘ammonite marker bed 1’ by Schulze et al. (2003, 2004) who reported Vascoceras cauvini, Metoicoceras geslinianum and Burroceras transitorium, which indicate Near East ammonite zone T1 of Lewy & Raab (1976) correlating with the European M. geslinianum and N. juddii zones. The record of B. transitorium needs further confirmation, since this species is hitherto recorded from New Mexico, Arizona and Brazil only (Kirkland 1996; Gale et al. 2005). In the middle part of this interval ammonites are abundant at section metre 60.5 m in section GM3 and at 3.2 m in section KB3. Although most of the specimens are poorly preserved, besides Puzosia sp. the largest part can be possibly attributed to very feebly ribbed Watinoceras spp., namely W. guentheri, W. hesslandi, W. inerme and W. semicostatum. All of these are previously known from western Morocco and western Afrika (Reyment 1955, 1957; Collignon 1967). This group of species might be conspecific following Wright & Kennedy (1981), if so W. hesslandi Reyment (1955) has priority. The total range of these smooth Watinoceras spp. is poorly known, but the genus is widely accepted as appearing not before the Lower Turonian of the modern substage definition following the discussion of Collignon’s (1967) data by (Lehmann & Herbig 2009). The first and possibly most common occurrence of these forms can be placed into the basal Turonian W. moremani zone (Wiedmann & Kuhnt 1996), just above the latest Cenomanian Neocardioceras juddii zone.
175
Ammonite zones T2-4 correlate to the basal CC11 and can be only roughly determined by scarce findings of Choffaticeras pavillieri and Ch. quaasi in section part 67–74 m. The limestone in the top of the section represents the Wala Limestone Member, which was described as ‘ammonite marker bed 3’ by Schulze et al. (2003, 2004). It is characterized at the base by an Early Turonian ammonite assemblage with Vascoceras durandi, Thomasites rollandi, Choffaticeras luciae, Ch. quaasi and Fagesia lenticularis, indicative of zones T5 –6a. The first three species are typical for the Pseudaspidoceras flexuosum to Thomasites rollandi zones in the middle part of the Lower Turonian in Tunisia, the detailed range of Ch. quaasi and F. lenticularis, the latter is a somewhat obscure species following Chancellor et al. (1994) in the Early Turonian. Transition into zone 6b (M. nodosoides/C. woollgari zone boundary, base Middle Turonian) was placed by Schulze et al. (2003, 2004) in the Middle Wala Limestone.
Lithological evolution at the different intra-platform basin sections The GM3 section can be subdivided into six lithological units (Fig. 2). Overlying the platform limestone called Karak Limestone, the analysed profile starts at profile metre 47 with a 12 m succession of greenish marls and clays (lower 4.5 m not exposed) with interbedded nodular limestone beds (green clay unit). Occasionally, gypsum beds and crosscutting diagenetic gypsum veins are intercalated at section metre 49 –50 m. At 52.30 m a thin (c. 15 cm), iron-rich black clay layer marks a first event of increased TOC accumulation (Fig. 2). This bed is very conspicuous in the field and contains a 3 cm thick, very dense, limestone bed. Above this black shale, the succession of greenish clays continues to 54.8 m. A 2.5 m thick unit of dolomite and ankeritized platform carbonate (dolomite unit) follows with a layer of strongly distorted gypsum-carbonate breccia with signs of reworking at the base, possibly connected to a minor hiatus (TS in Fig. 2). It marks an abrupt change in lithoand bio-facies. The section continues into a 7.5 m thick alternation of brownish marl and bituminuous, platy, partly laminated limestone beds (platy limestone unit). The third bed of this unit is an oysterlimestone typical of the platform carbonate facies of the Cenomanian investigated here. The two basal and the upper two bituminuous limestone beds are calcisphere (calcareous dinoflagellate cysts Pithonella) packstones showing bioturbation (Fig. 2). The other beds, in contrast, are laminated. Thin sections reveal a strong alteration of the limestone. The dark brown marl at the base of the platy limestone unit forms a second black shale (TOC 1).
176
J. E. WENDLER ET AL.
This interval is followed by 10 m of monotonous brown clays/marls (brown marl unit). The lower half of this unit can be considered a third black shale. The next unit is a 6 m thick interval of greenish-grey marls (grey marl unit) that contain TOC peak 2 representing the fourth black shale of the section. At section metre 80, a 2 metre thick green to grey, gypsum-rich marl is present. It is followed by a 1 m thick interval of red marls at the base of the following limestone sequence (socalled Wala Limestone). This red marl represents a marine red bed, which potentially is a marker horizon throughout the Levant platform (Wendler et al. 2009a). The grey, bioclastic Wala Limestone is approximately 14 m thick and represents platform-type carbonate deposits rich in oysters. Section KB3 is essentially a succession of black shale with intercalated limestones. Section interval 0–5 m displays six equally-spaced, dm-thick, limestone beds, light-beige in colour, alternating with dark-grey marl. Section interval 5 –16 m is dominated by dark grey to greenish marl with regular intercalations of dm-thick harder beds. At section metre 16.1 a limestone bed, comparable to the beds in the lower part of the section, is present. From section metre 16.2– 24 of the section greenishgrey marl were deposited. The upper part of this unit is barely exposed and strongly weathered and distorted. Thus sampling ends at section metre 20. The Wala Limestone member in the top of the section is dislocated by some metres owing to rock sliding.
Isotope record Whole rock stable carbon isotopes are in the range of –3 to þ4‰ (Fig. 2). High-frequency fluctuations (metre-scale) are evident and a superimposed trend (2–3 m bundles) forms conspicuous negative and positive excursion phases. A broad positive excursion above mean value is present in both sections. The carbon isotope excursion (CIE) of GM3 and KB3 (Fig. 2) shows the following shape (section metres are given for GM3): † pre-excursion background around 1‰ (with a decreasing trend) in section GM3 part 47– 51 m; † pre-excursion strong negative excursion to – 4‰ (section GM3 part 51 –54.5 m, double peak: negative peak 1: N1; negative peak 2: N2); † first build up (peak a) – slow increase to 1.7‰; † a significant decrease to 0‰ at 57.5 m (the so-called ‘trough’); † second build up (peak b) – maximum around 3‰, † after peak b, a period of significant fluctuation down to 0‰ and below occurs (around 60 m); † a third build up (peak c) starts around section metre 63 and is part of a plateau of d13C values c. 2.5‰ from 62 to c. 70.5m;
† the post excursion interval 70.5– 83 m (slow decrease to mean value), which contains one significant positive excursion of above 3‰ at 74.5 m, and a strong negative excursion to –2.8‰ at 83 m. This negative excursion represents the lowest value since the pre-excursion negative spike.
Calcimetry and gypsum marker beds In section GM3 carbonate content is between 10 and 90%, and the record shows a distinctive cyclic pattern consisting of high-frequency fluctuations bundled into sets (Fig. 2). These bundles are numbered starting at the base of the isotope excursion. Section part 47–55 m below the isotope excursion can be subdivided into three bundles (– 1 to –3) of about 3 m thickness. Abundant disperse gypsum (19%) is present in the clay and marl in the lower and upper parts of bundle –2 (gypsum bed 1). Gypsum content otherwise is generally below 1%. The platy limestone facies (section part 57 –64.5 m) shows high-frequency cyclicity in the carbonate content values that range between 50 –90%. A conspicuous minimum of 11% carbonate related to elevated gypsum content occurs at bundle boundary 2/3 (gypsum bed 2). From 64.5 m to 80 m the cyclicity continues to show bundling, but bundle thickness increases. At the base of bundle 7 a reddish bed reflects increased contents of iron oxides and gypsum (gypsum bed 3). A bed with a substantial minimum of ,10% carbonate content at 80 m exhibits the highest gypsum content of the section and marks the boundary between bundles 10 and 11 (gypsum bed 4, 69% gypsum). The four so-called gypsum beds are clays and marls with disperse gypsum rather than massive sulphate beds. The most conspicuous one, gypsum bed 4, potentially is connected to a minor hiatus. TOC is mostly around 0.5 wt% throughout the section (Fig. 2). With the onset of the platy limestone facies related to isotope peak b a 2 m thick peak in TOC (1.5–3%) can be observed (TOC1). Throughout the platy limestone and brown marl units TOC remains around 0.5% with short peaks between 1– 2%. Another major peak in TOC straddling about 4 m in thickness (1.5–3% TOC) occurs in the brown marl unit in the carbonate bundles 9 and basal 10 (TOC2). In the KB3 section carbonate values are between 30 and 95% (Fig. 2). Carbonate content also shows high-frequency cyclicity with superimposed longer-period bundles. The carbonate-minima of these bundles are correlated with the GM3 record. The carbonate cycles of both sections are positively correlated with the cyclic stable carbon isotope excursion.
CENOMANIAN– TURONIAN ORBITAL TIMESCALE FOR THE LEVANT
TOC values abruptly rise from a background level of ,1% at the base of the section to a first peak of .7% at 2.5 m (TOC1). Above that level the mean value is constantly decreasing from 4 to 2% and sharply increases to a second major peak of 7% at section metre 18 m to form a 3.5 m thick double peak with maxima of 7% (TOC2). Between these two main peaks TOC values remain high from 5.5 m to 12.5 m with peaks .5%. Similar to the carbonate content, high-frequency cyclicity characterizes the TOC record.
Isotope and calcimetry correlation: spectral analysis The correlation of carbonate cycles (bundles) between GM3 and KB3 is corroborated by the comparable shape of the d13C records and the position of major TOC peaks. Thus, section KB3 comprises cycles 1–10 of GM3. Owing to the high sampling resolution of KB3 the carbonate and TOC records of this section are suitable for spectral analysis in order to determine the type of cyclicity represented. We estimate mean accumulation rates using possible ages given for the stratigraphic interval studied (based on Ogg et al. 2004; Sageman et al. 2006) assuming an equal length of the carbon isotope excursion all over the globe. These theoretical accumulation rates are applied to the spectra of carbonate and TOC in order to search for the best fit to the Milankovitch frequency band. According to the Ogg et al. (2004) timescale the duration from the base of the Middle Cenomanian to the basal Middle Turonian is about 95.7 to 92.15 ¼ 3.55 Ma. This interval is represented by c. 80 m in the GM section (see Schulze et al. 2004 for entire section) resulting in 22.5 m/Ma mean accumulation rate. With respect to the isotope excursion the time span between the first increase (peak a) and the end of the plateau is in both time scales about 870– 1000 ka (vertical bar in Figs 2, 3 & 4) resulting in mean accumulation rates of 16– 19 m/Ma (KB3) and 20–23 m/Ma (GM3). This interval contains eight cycles in both sections. Since cycle thicknesses in GM3 are highly variable (Fig. 2) the range of estimated accumulation rates is larger: cycles 4 –6 are 2 m thick while all other cycles are 3–3.8 m thick. Thus, accumulation rates around 30 m/Ma dominate. Power spectra were generated using the Lomb – Scargle algorithm. The TOC spectrum of KB3 (Fig. 5a) shows a strongest power signal at 7.2 m, followed by peaks in power at a 16 m long period (close to section thickness of 20 m: low significance), 2.83 m, 1.65 m, 0.34–0.41 m (only two samples per cycle: low significance). A less significant signal occurs at 0.88 m. Assuming 18 m/Ma mean accumulation rate the 7.2 m cycle would be
177
400 ka long eccentricity, 2.83 m represents a period of c. 200 ka not known from the Milankovitch frequency band. The cycle at 1.65–1.84 m is in the range of short eccentricity, and 0.34 –0.41 m would represent precession. The peaks at 0.66 m and 0.88 m could reflect the obliquity cycle. The spectrum of carbonate content (Fig. 5b) of the most symmetrical, highest resolution part of KB3 (3–11 m) shows two prominent peaks of 1.6 –2.65 m cycles. Power is also recorded at a cycle of c. 7 m thickness which is, however, about the thickness of the measured section. Two small signals occur at 0.34 m and 0.36 m. Cycle and bed thickness is lower in this part of the sections (Fig. 2). Thus, using 17 m/Ma proposed accumulation rate for this particular interval (derived from high-resolution correlation of Fig. 4) the two strongest peaks are related to 156 and 97 ka, the 7 m cycle spans 400 ka, and the 0.34 m cycle would comprise 20 ka. The cycle ratio 1.6:0.34 m is 4.7:1. Figure 5c shows the power spectrum of carbonate content of section GM3. Three peaks occur in the frequency band of short eccentricity (2.65 m, 3.3 m, 4.1 m), and strong peaks are present in both the obliquity (1.29 m) and precession (0.71 m and 0.59 m) frequency bands. In Figure 5d the section part 67 –79.8 m is analysed. We find the strong 3.42 m cycle (cycles 8–11 in Fig. 2) corresponding to a duration of 107 ka when 32 m/Ma accumulation rate is assumed. This latter analysis suggests that accumulation rates of GM3 are higher than those of KB3, especially in the Lower Turonian part of GM3.
Discussion High-resolution correlation with Pueblo: an orbital timescale for the Levant In Figure 3 we attempt a correlation of the Jordan GM3 isotope record with the Pueblo GSSP section record of Sageman et al. (2006). We can assume a reasonably complete stratigraphic record based on the lithological properties of the GM3 section, which indicates only two potential positions of minor hiati. Those are, first, the base of the dolomite unit where reworking is indicated, and second, the gypsum bed 4 where a substantial sulphate content suggests enhanced evaporitic conditions during deposition linked to temporary emersion (Wendler et al. 2009a). The 100 ka cycles in the carbonate record positively correlate with the d13C curve but are more detailed. Thus, we use the clearly defined cycles of the carbonate record and correlate excursions of decreased carbonate content (negative d13C excursions) with negative excursions in the Pueblo
178 J. E. WENDLER ET AL. Fig. 3. Correlation of the GM3 section and Pueblo (Colorado) GSSP; Pueblo orbital time scale from Sageman et al. (2006); correlation lines use minima in the isotope/carbonate curves. 100 ka-cycle numbering (carbonate curve) indicates Jordan fifth-order cycles; 100 ka-cycle numbers have been suggested accordingly for the Pueblo d13Ccarb record between the base of the CIE and the C–T boundary in concert with the Pueblo orbital model. GM3 isotope record: Isotope events from Jarvis et al. (2006) are labelled according to Figure 4: M, Monument; h, Holywell; c1; c2; L, Lulworth; lpm, late plateau minimum (new event introduced here); the respective numbering according to Voigt et al. (2006) is TU1 to TU5. Jordan orbital model: stippled curve, 1.2 Ma long-obliquity cycle (third-order sea-level cycle); solid curve, 400 ka long-eccentricity (fourth-order sea-level cycle). In the stratigraphy column, sequence S5 is grey-shaded; 5a, 5b and 5c are the fourth-order sub-sequences according to orbital model; SB ¼ sequence boundary.
brown marl
93.2
86
94.0 94.2
trough
3 1
OAE 2
a
77 max δ Ccarb 13
67 63
Monument
TS
CENOMANIAN
dolo.
CC10
93.8
79
b
94.4
δ Ccarb 13
VPDB –28 –27 –26 –25 –24 –23 –22
CeJo4
δ Corg 13
VPDB Stage
green clay
N. vibrayeanus
CENOMANIAN
Hummar HST LST
8
platy limestone
T1
c Plenus Marl beds
93.6
TST
Q. gartneri
93.4
A. albianus
Ammonite biozone ages (orbital model) Q. gartneri
Holywell
93.0
92.77 Ma
93.19 93.28 93.38
93.55
93.85
94.09
Nannofossils
c1 92.6
Ammonites
c2
Bridge Creek Limestone H. helvetica W. archeaocretacea P. V. N. juddii W. devon. flex. birc. M. nodosoides C. woollgari Rhagodiscus asper E. floralis E. eximius M. chiastius
Lulworth
92.2
Hartland Shale Rotalipora cushmani M. mosbyense S. gracile
6
Formation
marker beds
5
92.0 Ma
92.4
grey marl
T5-6a CC11
2 3 4
4
92.8
T2 - T4
HST Shueib
VPDB
3
Foraminifera
TuJo1
0 1
Pueblo
English Chalk 2
TURONIAN
δ13Ccarb
–3 –2 –1
Age (Geol. timescale)
GM3 Jordan
179
Fig. 4. Correlation of d13C records Jordan and Pueblo (Sageman et al. 2006) with the English chalk reference curve (Jarvis et al. 2006). SB ¼ sequence boundary. Positions of marker beds of Plenus Marl (England) and Bridge Creek Limestone (Pueblo) are given to aid lithological comparison with Jordan (platy limestone unit grey shaded). Left panel: qualitative representation of platform extent redrawn from Philip & Airaud Crumiere (1991).
CENOMANIAN– TURONIAN ORBITAL TIMESCALE FOR THE LEVANT
TURONIAN
Near East Ammonite zones Wala Lithological units Limestone
Nannofossils
Formation
Systems tracts
TST S4/S5 (3rd order)
broad extent
LST
reduced extent
Stage
Carbonate platforms
180
(a)
J. E. WENDLER ET AL.
(b)
1.2
1.2
KB3 TOC record 1.0
1.1
7.2 m (400 ka)
Power
2.83 m (157 ka)
0.6 0.5
0.35 m (19.4 ka)
1.84 m (102 ka) 1.65 m (91.6 ka)
0.4
0.88 m (49 ka)
0.3
0.4 m (22.5 ka)
2.65 m (156 ka) 1.6 m (95 ka)
thickness: 8 m est. acc. rate: 17 m/Ma
0.8 0.7 0.6 0.5
7m (400 ka)
0.4 0.66 m (37 ka)
0.1
0.0
0.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.2
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8
Frequency (1/m)
Frequency (1/m)
(d)
1.2 3.3 m 16 m (500 ka) (102 ka)
GM3 carbonate 55–80 m, cycles1–10
4.1 m (128 ka)
2.65 m (83 ka)
1.8 m (56 ka) 1.29 m (40 ka)
0.9
7 cycles = >30 m/Ma 3 cycles =