<<

Source rock characterization and generation modelling of the Levant Basin, onshore-offshore Lebanon: An integrated approach

Von der Fakultät für Georessourcen und Materialtechnik der

Rheinisch -Westfälischen Technischen Hochschule Aachen

zur Erlangung des akademischen Grades eines

Doktors der Naturwissenschaften

genehmigte Dissertation

vorgelegt von M.Sc.

Samer Bou Daher

aus Beirut, Libanon

Berichter: Univ.-Prof. Dr. rer. nat. Ralf Littke Univ.-Prof. Dr. Rudy Swennen

Tag der mündlichen Prüfung: 07. März 2016

Diese Dissertation ist auf den Internetseiten der Hochschulbibliothek online verfügbar

“There are no limits when you are surrounded by people who believe in you, or by people whose expectations are not set by the short-sighted attitudes of society, or by people who help

to open doors of opportunity, not close them.”

Neil deGrasse Tyson, The Sky is Not the Limit: Adventures of an Urban Astrophysicist

I dedicate this work to my family

ii

Acknowledgment

My PhD journey would not have been possible without the support, guidance, and help of many people and institutes. I would like to start by thanking the German Academic Exchange Service

(DAAD) and Maersk Oil for funding this project.

A good supervision plays a major role in the success of any project. And I have had the privilege of receiving a great supervision. I thank my supervisor Prof. Dr. Ralf Littke for his continuous support, advice, and trust. I thank my co-supervisor Dr. Fadi Nader whose positive energy, enthusiasm, and continuous support have been an invaluable asset in this project and a precious source of motivation for me. Whether on the professional or on the personal level, it has been a pleasure and an honour to be their student and to learn from them.

I thank Prof. Dr. Harald Strauss, for preparation, measurement, and discussion of organic carbon isotope data. I thank Dr. Sven Sindern for his help in producing the XRF data. I thank

Dr. Carla Müller for the biostratigraphic data and for her great company and valuable discussions in the field. I thank Dr. Nicolas Hawie for his help and fruitful discussions.

I thank the crew of the institute of geology and geochemistry of petroleum and coal at RWTH

Aachen who became like a family to me and was always welcoming, supportive, informative, and amusing. Special thanks to Olga Schefler for her administrative super powers.

I thank the IFPEN teams at the departments of geology and geochemistry who have welcomed me as a guest for 5 months and were always available for assistance and discussions, particularly Fadi Nader, who has given me the opportunity to be at IFPEN, Bernard Carpentier,

Mathieu Ducros, and Pauline Michel for their technical support and expertise.

I thank Isabelle Moretti (GDF Suez) and other anonymous reviewers for their valuable comments which improved the published manuscripts.

iii

I thank BeicipFranlab for providing a TemisFlowTM software license. I thank the Cimenterie

Nationale S.A.L. for granting access to their well cores and providing assistance in the quarry in Chekka. Special thanks go to Mr. Ahmed Hoteit. I thank Mr. Nabil Abou Dehen for his generosity and for providing the local information needed in the field in Hasbayya. I thank the

Lebanese ministry of energy and water (Petroleum Administration) for providing permissions and access.

To my friends who have supported me to pursue this challenging dream and my new friends that I made along the way, I cherish all the time we spent together camping, hiking, mountain biking, climbing, travelling, board gaming, kayaking, cooking, eating, partying, laughing, drinking, and much more….. I thank you for being part of my life.

Finally, I would like to express my deepest appreciation and gratitude to my family for their unconditional love and support. Without them I would not have been here today.

iv

Preface

The idea of this PhD thesis came as a result of new data that became available in the last decade and that gave rise to new scientific questions in the complicated east Mediterranean region.

Initiated, coordinated and co-supervised by Fadi Nader (IFPEN) three parallel PhD projects were launched in the years 2011-2012, among which was this project.

The first launched project, conducted by Nicolas Hawie at University Paris 6 UPMC, tackled the geodynamic evolution and sedimentary filling of the Levant basin using 2D seismic data, detailed outcrop sedimentologic and biostratigraphic analysis, and forward stratigraphic modelling.

The second launched project, conducted by Ramadan Ghalayini at University Paris 6 UPMC, tackled the structural evolution of the complex Cenozoic zone of the Levant Basin offshore

Lebanon using 3D and 2D seismic data, as well as structural modelling.

The third project, presented in this thesis, tackles the potential source rocks exposed along the eastern margin of the Levant basin, and analyses the thermal history of the study area and the evolution of potential petroleum systems onshore and offshore Lebanon using detailed geochemical and petrographic analysis on source rock samples, as well as 3D petroleum system modelling.

Along with these projects, several master and bachelor projects were accomplished and another set of PhD theses were launched in 2014-2015 targeting the lithospheric dynamics of the Levant basin as well as the geodynamic and structural evolution of its western margin.

All in all, this tremendous scientific and organizational effort lead to the establishment of the

“Levant research group” which includes several European and international universities and research institutes and helps to create synergy between the various research projects.

v

Abstract

This thesis consists of a combination of analytical and numerical modelling work on the east

Mediterranean Levant basin which has in the last decade become a frontier hydrocarbon province. Several gas discoveries have been recorded in Miocene reservoirs offshore Israel and seismic data suggest promising prospective plays in deeper intervals throughout the basin.

Source rock information and calibration points are hitherto scarce due to the lack of well data, especially in the offshore Lebanon part of the basin. Thus, this thesis provides a considerable contribution to the available geochemical data and thermal evolution of the study area.

The analytical work consist of geochemical data from several potential source rocks, specifically Upper Cretaceous (Campanian – lower Maastrichtian), outcropping along the eastern margin of the Levant Basin onshore Lebanon. The Upper Cretaceous rock succession has shown very good source rock properties that have a type that varies laterally and vertically throughout the study area between Type II and Type IIS kerogen. Biomarker, kinetic, petrographic, and elemental data attribute this shift in kerogen type to iron deficiency and variable organic matter content, and suggest deposition of these source rocks under a high productivity zone mostly restricted to the shelf area. This depositional model implies a decrease in organic matter (OM) richness towards deeper and distal basinal areas, which in turn would be accompanied by a shift in bulk hydrocarbon generation kinetics to higher values due to oxidation of labile OM and a decrease in organic sulphur content. This trend is observed in HI values and measured bulk hydrocarbon generation kinetics of Upper Cretaceous source rocks with various OM content onshore Lebanon. The recorded lateral and vertical variations in the

Upper Cretaceous (Campanian – lower Maastrichtian) suggest that the onset and the extent of the oil and gas window can vary considerably within the same source rock as a function of organofacies. Other source rocks that have been characterized in this thesis include the Upper

Jurassic (Kimmeridgian), Lower Cretaceous (Neocomian), Cenomanian, Albian, and upper vi

Paleocene. The Kimmeridgian, Neocomian, and Albian include Type III/IV kerogen with gas potential while the Cenomanian and upper Paleocene include type I and II/III with oil and gas potential. All analysed source rocks are thermally immature.

The numerical modelling part consists of a large scale 3D thermal history and maturity model of the Levant basin, margin, and onshore Lebanon. The constructed model has been calibrated using the publically available data in addition to the data produced in this thesis. The model suggests the presence of several potential petroleum systems including an Upper Cretaceous-

Oligo-Miocene biogenic and thermogenic system in the deep basin, a Jurassic-Cretaceous system along the margin, and a Permian-Triassic system in the onshore. Sensitivity analysis suggested an important effect of the depth of the lithospheric-asthenospheric boundary on the thermal history of the basin and showed that under any scenario, the thickness of the biogenic zone below the Messinian salt would vary between 700 and 1500 m in the deep basin offshore

Lebanon.

vii

Zusammenfassung

Die vorliegende Doktorarbeit beinhaltet eine Kombination aus analytischen Daten und numerischer Modellierung des ostmediterranen Levante Beckens, in welchem sich in der letzten Dekade eine große Explorationstätigkeit entwickelte. Vor der Küste Israels sind verschiedenste Gasfunde in miozänen Reservoiren gemacht worden; seismische Daten lassen weitere vielversprechende Lagerstätten in tieferen Gesteinsschichten des Beckens vermuten.

Informationen über das Muttergestein und damit mögliche Kalibrationsdaten für numerische

Modelle sind rar, da bisher kaum Daten aus Bohrungen existieren. Dies betrifft insbesondere die Region vor der libanesischen Küste. In diesem Sinne leistet die vorliegende Arbeit einen wesentlichen Beitrag zum Verständnis der Kohlenwasserstoff-Geochemie und zur thermischen

Entwicklung des Arbeitsgebietes.

Die geochemischen Daten beinhalten Informationen über verschiedene, potentielle

Kohlenwasserstoffmuttergesteine der Oberkreide (Campanium – unteres Maastrichtium), welche entlang des östlichen Beckenrandes im Libanon anstehen. Die lithologische Abfolge der Oberkreide zeigt sehr gute Muttergesteinseigenschaften. Der entsprechende Kerogentyp variiert innerhalb des Arbeitsgebietes in lateraler und vertikaler Richtung zwischen Typ II und

Typ IIS. Daten zu Biomarkern, Kinetiken, Petrographie und der elementaren Zusammensetzung ermöglichten die Verknüpfung der Veränderung des Kerogentypes mit einer Verarmung an

Eisen im Ablagerungsmilieu, auch gesteuert durch variable Anteile an organischem Material.

Die Daten legen eine Ablagerung des Muttergesteines in Zonen hoher Bioproduktivität, wahrscheinlich beschränkt auf die Schelfregion und den flachen Kontinentalrand, nahe. Das vorgestellte Ablagerungsmodell beinhaltet eine Abnahme des organischen Kohlenstoffanteiles in Richtung der tieferen und distaleren Beckenfazies. Dieses Modell wirkt sich, durch

Oxidation labilen organischen Materials und einer Abnahme des organischen

Schwefelgehaltes, auch auf den Kerogentyp und die Kohlenwasserstoff-bildungskinetiken aus. viii

Gemessene Kinetiken an Proben aus der Oberkreide, mit variablen Anteilen an organischem

Material bestätigen diesen Trend. Die beobachteten lateralen und vertikalen Veränderungen innerhalb der organischen Fazies der Oberkreide lassen vermuten, dass die Grenzen des Öl- und Gasfensters innerhalb des gleichen Muttergesteines auf Grund der Organofazies erheblich variieren können. Weitere, in dieser Studie charakterisierte Muttergesteine beinhalten Proben des Oberjura (Kimmeridgium), der Unterkreide (Neocomium), des Cenomanium, Albium und des oberen Paläozäns, von denen Kimmeridgium, Neocomium und Albium Typ III/IV Kerogen mit Gaspotential beinhalten und Cenomanium und oberes Paläozän Typ I und II/III Kerogene mit Öl- und Gaspotential aufweisen. Alle analysierten Muttergesteine sind thermisch unreif.

Der numerische Modellierungsteil beinhaltete ein großmaßstäbliches Modell, mit dem dreidimensional die thermische Geschichte und Reife des Levante Beckens, des Beckenrandes und der onshore Gebiete Libanons ausgearbeitet wurde. Das erstellte Modell wurde mit den hier neu vorgestellten Daten und auch zuvor publizierten Daten kalibriert. Die Ergebnisse lassen einige potentielle Kohlenwasserstoffsysteme vermuten, unter anderem ein bio- und thermogenes System in der tieferen Beckenfazies von Oberkreide-Oligozän-Miozän, ein jurassisch-kretazisches System entlang der Beckenränder und ein permo-triassisches System im Onshorebereich des Libanons.

Sensitivitätsanalysen wiesen einen erheblichen Einfluss der Tiefe der Lithosphären-

Asthenosphäre Grenze auf die thermische Geschichte des Beckens nach und zeigten, dass die

Mächtigkeit der biogenen Zone unter dem Messinischen Salz offshore Libanons zwischen 700 und 1500 m variieren.

ix

Abbreviations

CaCO3: Calcium carbonate

TOC: Total organic carbon

TIC: Total inorganic carbon

TS: Total sulphur

OM: Organic matter

HI: Hydrogen index

OI: Oxygen index

S1: Amount of free hydrocarbons in rock (mg HC/g rock); mobile organic matter in source rock.

S2: Amount of hydrocarbons produced from the breakdown of kerogen during pyrolysis (mg HC/g rock).

S3: Amount of CO2 produced from the breakdown of kerogen during pyrolysis (mg CO2/g rock).

Tmax: The temperature at which the rate of kerogen breakdown is at maximum.

Fe: Iron

S: Sulphur

V: Vanadium

Ni: Nickel

VRr: Vitrinite reflectance

13 δ Corg: Organic carbon isotopic ratio

HF: Heat flow

x

Table of Contents Acknowledgment ...... iii Preface ...... v Abstract ...... vi Zusammenfassung ...... viii Abbreviations ...... x Chapter I Introduction and overview ...... - 1 - 1.1 Introduction ...... - 1 - 1.2 Thesis overview ...... - 5 - Chapter II Depositional environment and source-rock characterisation of organic-matter rich upper Turonian-upper Campanian carbonates, Northern Lebanon ...... - 7 - 2.1 Abstract ...... - 7 - 2.2 Introduction ...... - 8 - 2.3 Stratigraphic framework ...... - 10 - 2.4 Materials and methods ...... - 13 - 2.4.1 Samples ...... - 13 - 2.4.2 Methods ...... - 13 - 2.5. Results ...... - 18 - 2.5.1 Lithofacies and sedimentology ...... - 18 - 2.5.2 Elemental analyses ...... - 20 - 2.5.3 Rock-Eval pyrolysis ...... - 22 - 2.5.4 Organic petrography ...... - 22 - 2.5.5 Organic geochemistry ...... - 24 - 2.5.6 Organic carbon isotopes ...... - 25 - 2.6 Discussion ...... - 25 - 2.6.1 Depositional environment and early diagenetic conditions ...... - 25 - 2.6.2 Petroleum potential and maturity ...... - 32 - 2.7 Exploration implications for the Levant basin ...... - 33 - 2.8 Conclusions ...... - 34 - Chapter III Geochemical and petrographic characterization of Campanian-Lower Maastrichtian calcareous petroleum source rocks of Hasbayya, South Lebanon ...... - 38 - 3.1 Abstract ...... - 38 - 3.2 Introduction ...... - 39 - 3.3 Geological setting ...... - 41 - 3.4. Stratigraphic framework ...... - 43 - 3.5 Samples ...... - 44 - 3.6 Methods ...... - 45 - 3.6.1 Biostratigraphy ...... - 45 - 3.6.2 Elemental analysis ...... - 45 - 3.6.3 Rock-Eval pyrolysis ...... - 46 - 3.6.4 Petrographic analysis ...... - 46 - 3.6.5 GC-MS...... - 48 - 3.6.6 CP-Py-GC-MS ...... - 49 - 3.7 Results ...... - 49 - 3.7.1 Biostratigraphy ...... - 49 - 3.7.2 Lithofacies and sedimentology ...... - 51 - 3.7.3 Elemental analysis ...... - 51 - 3.7.4 Rock-Eval pyrolysis ...... - 53 - 3.7.5 Organic petrology ...... - 55 - 3.7.6 Organic geochemistry ...... - 56 - 3.8 Discussion ...... - 58 - xi

3.8.1 Depositional environment...... - 58 - 3.8.2 Petroleum potential, maturity, and kerogen type ...... - 67 - 3.9 Conclusions ...... - 69 - Chapter IV 3D thermal history and maturity modelling of the Levant Basin and its eastern margin, offshore-onshore Lebanon ...... - 74 - 4.1 Abstract ...... - 74 - 4.2 Introduction ...... - 75 - 4.3 Geological setting ...... - 76 - 4.4 Lithostratigraphic framework ...... - 78 - 4.5 Materials and methods ...... - 80 - 4.5.1 Source rocks ...... - 80 - 4.5.2 3D data set ...... - 81 - 4.5.3 Boundary conditions ...... - 82 - 4.5.4 Vitrinite reflectance, transformation ratio, expulsion ...... - 84 - 4.6 Results and discussion ...... - 84 - 4.6.1 Calibration ...... - 84 - 4.6.2 Source rocks ...... - 85 - 4.6.3 Heat Flow ...... - 88 - 4.6.4 Burial history and maturity ...... - 89 - 4.6.5 Transformation ratio and expulsion ...... - 95 - 4.6.6 Petroleum systems ...... - 101 - 4.6.7 Sensitivity and uncertainty ...... - 105 - 4.7 Conclusions ...... - 111 - Chapter V Conclusion and outlook ...... - 115 - 5.1 Summary and conclusions ...... - 115 - 5.1.1 Upper Cretaceous (Campanian – Maastrichtian) source rock ...... - 115 - 5.1.2 Source rocks of the east Mediterranean ...... - 118 - 5.1.3 Petroleum systems of the Levant Basin ...... - 119 - 5.2 Outlook ...... - 121 - 6. References ...... - 123 - 7. Appendix ...... - 137 - Appendix 1...... - 137 - Appendix 2...... - 142 - Appendix 3...... - 142 -

xii

Chapter I Introduction

Chapter I

1.1 Introduction

Sedimentary basins are subsiding regions of earth’s crust that accumulate thousands of meters of sediments over prolonged periods of time (Allen and Allen, 2014). These basins, not only do they record earth’s history over millions of years, but also they provide 90% of our energy resources (Littke et al., 2008). Due to their enormous scientific and economic importance, many sedimentary basins have been and still are extensively studied and explored, but many others remain a poorly understood frontier due to their complexity and/or inaccessibility.

The sedimentary fill of a basin is controlled by many parameters including eustatic sea-level, climate, and tectonics (Littke et al., 2008). The interaction of these parameters results in the deposition of a large variety of sedimentary rocks among which are organic matter rich rocks.

Organic matter rich rocks are a distinctive type of rocks that form under a set of conditions that favour the production and subsequent preservation of organic matter derived from marine, lacustrine, and/or terrestrial life. As these rocks subside to deep burial depths, they experience elevated temperatures and pressures that alter many of their characteristics irreversibly and result in the generation of fluids such as petroleum and natural gas (Welte et al., 1997). Hence these types of rocks are also referred to as petroleum source rocks.

In order for petroleum and or natural gas accumulations to occur, many elements must come in place and many processes must happen in the right sequence. These elements are known as petroleum system elements. The term petroleum system was coined by Magoon (1988) and was defined as a natural system that includes all the elements and processes needed for a hydrocarbon accumulation to occur. The elements of a petroleum system are source rocks, reservoir rocks, cap rocks, and overburden. The processes involved in the formation of a

- 1 -

Chapter I Introduction petroleum system are trap formation, burial, hydrocarbon generation, migration, and eventually accumulation.

This study focuses on the source rock element as well as the burial and hydrocarbon generation processes. We also discuss the availability of the other elements and processes as part of a petroleum systems assessment of a frontier hydrocarbon province in the east Mediterranean known as the Levant region.

The Levant region (Fig. 1.1), characterized by its Mediterranean climate and beautiful mild weather, has been assumed for decades to be deprived of economic accumulations of natural resources, opposed to the rest of the Middle East region known for its warm and dry climate and containing the world’s largest oil fields. This long held assumption recently changed due to developments in sub salt seismic imaging and deep sea drilling technologies which lead to the discoveries of some of the world’s biggest gas fields of the last decade (Zohr, Tamar, Dalit,

Leviathan, Karish, Tanin, Dolphin, and Cyprus-A) (www.nobleenergyinc.com; www.eni.com) and redirected the industrial and academic interests to this promising yet challenging frontier area.

The new data that became available as well as the new discoveries that were made gave rise to many questions related to sediment sources, structural evolution, hydrocarbon source rocks, hydrocarbon type (biogenic vs. thermogenic), burial and thermal history, and many others.

The sedimentary filling of the Levant basin has been recently examined and thoroughly assessed by several authors combining detailed seismic facies interpretation with onshore field observations, source to sink approach, and forward stratigraphic modelling which elucidated many aspects of the basin’s tectono-stratigraphic evolution (Gardosh et al., 2006, 2008;

Steinberg at al., 2011; Hawie et al., 2013a&b, 2015; Gvirtzman et al., 2014). The structural styles and the tectonic evolution of the Levant basin and its eastern margin have also been

- 2 -

Chapter I Introduction documented and interpreted (Butler et al., 1998; Walley 1998; Beydoun 1999; Gomez et al.,

2007; Homberg et al., 2010; Ghalayini et al., 2014). The potential petroleum source rocks and the burial and thermal history of the Levant region, however, has been only rarely addressed in the literature. Shaaban et al. (2006) identified and evaluated potential source rocks in the north- eastern Nile delta. Feinstein et al. (2002) suggested the presence of several active source rocks in the Levant basin based on genetic characterization of gas shows offshore southwestern Israel.

Bein and Soffer (1987) studied oil samples from the Helez-Brur-Kokhav oil field in the southern coastal plain of Israel and suggested a Middle Jurassic source. Marlow et al. (2011) investigated thermal and burial history of the Levant basin using numerical 2D petroleum system modelling.

Only the Upper Cretaceous (Campanian – Maastrichtian) source rocks have been studied in detail in Israel (Bein et al., 1990; Minster et al., 1992; Almogi-Labin et al., 1993, 2012;

Ashckenazi-Polivoda et al., 2010, 2011; Schneider-Mor et al., 2012), and Jordan (Abed et al.,

2005). In Lebanon, no data on organic matter rich intervals, and neither on the thermal and burial history is publically available.

The aim of this PhD thesis is to determine and thoroughly characterize the important source rock intervals present along the eastern margin of the Levant Basin, particularly onshore

Lebanon (Fig 2.1a), in order to better predict and extrapolate their characteristics on a basin scale, and to assess the thermal history of the basin and the hydrocarbon generation potential of the identified source rocks using a 3D model in order to evaluate the potential petroleum systems. These objectives were achieved through an integrated workflow that includes sedimentologic, biostratigraphic, geochemical, petrological, and numerical modelling methods and tools.

Three sampling campaigns were accomplished during this thesis targeting several locations and several stratigraphic units where organic matter rich deposits are outcropping onshore Lebanon.

However, the source rock that was sampled in most detail was the Upper Cretaceous - 3 -

Chapter I Introduction

(Campanian – Maastrichtian), due to its high organic matter content and accessibility at several locations. Thus a detailed understanding of the depositional conditions under which this source rock was deposited and a thorough evaluation of the kerogen types and alterations were achieved.

This thesis constitutes a considerable contribution to the available petroleum source rocks database of the Levant region, provides a best fit regional thermal history and maturity model, and shows an example of organofacies variation within a single source rock and the effects that has on maturation and hydrocarbon generation.

Fig. 1.1. Simplified map showing the Eastern Mediterranean and North Arabian regions and the extent of the area where Triassic-Jurassic basins and some major oil and gas fields are found. (from May, 1991; Abou Shagar 2000; Nader, 2011)

- 4 -

Chapter I Introduction

1.2 Thesis overview

This thesis includes three independent yet complementary sections presented in chapters 2 – 4, followed by a concluding chapter 5.

Chapter II was published as Bou Daher, S., Nader, F.H., Strauss, H., Littke, R., 2014.

Depositional environment and source-rock characterisation of organic-matter rich upper

Turonian-upper Campanian carbonates, Northern Lebanon. Journal of Petroleum Geology 37,

1 - 20. This chapter presents a large geochemical data set from a systematically sampled 300 m thick Upper Cretaceous (Turonian – Campanian) rock succession penetrated in short wells in the city of Chekka, northern Lebanon. Based on the TOC and HI contents, 150 m of good oil prone source rocks were identified. Selected samples from the organic matter rich interval were used for further analyses revealing the maturity, depositional environment, and kerogen type of the investigated source rock.

Chapter III was published as Bou Daher, S., Nader, F. H., Müller, C., Littke, R. 2015.

Geochemical and petrographic characterization of Campanian-Lower Maastrichtian calcareous petroleum source rocks of Hasbayya, South Lebanon. Marine and Petroleum Geology, 64, 304

– 323. This chapter presents a set of geochemical, biostratigraphic, and sedimentologic data from an Upper Cretaceous (Campanian – Lower Maastrichtian) rock succession in the vicinity of the town of Hasbayya, southern Lebanon. An outcrop section of Santonian to Paleocene marls and fine grained carbonates has been sampled and investigated. Results proved the presence of excellent source rock deposits confined to the Campanian – lower Maastrichtian interval. These organic matter rich deposits have been analysed in detail to determine their source rock potential, maturity, depositional environment, and kerogen types. The results were then compared to the previously studied section in north Lebanon (chapter II) and to published data from equivalent sections in the region. A conceptual depositional model was constructed

- 5 -

Chapter I Introduction providing a predictive tool for the extrapolation of this source rock into the offshore Levant basin. Additionally, this chapter shows data from the historically famous Hasbayya solid bitumen found in fractures that had been mined for decades and have been assumed to be sourced from a deeper mature source rock. However, this chapter proves that the Hasbayya solid bitumen is an immature bitumen generated from the sulphur rich Campanian –

Maastrichtian source rocks and does not have a deeper origin.

Chapter IV is submitted to the Arabian Journal of Geosciences as Bou Daher, S., Ducros, M.,

Michel, P., Hawie, N., Nader, F. H., Littke, R. 3D thermal history and maturity modelling of the Levant Basin and its eastern margin, offshore-onshore Lebanon. This chapter presents source rock data for a number of organic matter rich intervals from Jurassic to Paleocene. Bulk hydrocarbon generation kinetics were derived for these source rocks including the Upper

Cretaceous which showed to have varying kinetic parameters controlled by the sulphur content as discussed in chapter III. A 3D numerical model was constructed using TemisFlow™, v.

2013.2 in order to assess the thermal history of the basin and to evaluate the maturity and hydrocarbon generation potential of the identified source rocks through time.

Chapter V summarizes the major findings of this thesis which can be divided into two parts.

The first is related to the understanding of source rocks depositional environment and its use as a predictive tool for extrapolation of source rock properties away from sampling locations. The second is related to the application of numerical modelling tools to elucidate important aspects of potential petroleum systems in frontier basin with little calibration data.

- 6 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Chapter II

DEPOSITIONAL ENVIRONMENT AND SOURCE-ROCK CHARACTERISATION OF ORGANIC-MATTER RICH UPPER SANTONIAN – UPPER CAMPANIAN CARBONATES, NORTHERN LEBANON

Keywords: Levant Basin, source rocks, Lebanon, Upper Cretaceous, carbonates, hydrocarbons,

Rock-Eval, kerogen type, organic geochemistry.

2.1 Abstract

Samples of Turonian – upper Campanian fine-grained carbonates (marls, mud- to wackestones; n = 212) from four boreholes near Chekka, northern Lebanon, were analysed to assess their organic matter quantity and quality, and to interpret their depositional environment. Total organic carbon (TOC), total inorganic carbon and total sulphur contents were measured in all samples. A selection of samples were then analysed in more detail using Rock-Eval pyrolysis, maceral analyses, gas chromatography – flame ionization detection (GC-FID), and gas chromatography – mass spectrometry (GC-MS) on aliphatic hydrocarbon extracts. TOC measurements and Rock- Eval pyrolysis indicated the very good source rock potential of a ca.

150 m thick interval within the upper Santonian – upper Campanian succession intercepted by the investigated boreholes, in which samples had average TOC values of 2 wt % and Hydrogen

Index values of 510 mg HC /g TOC. The dominance of alginite macerals relative to terrestrial macerals, the composition of C27–C29 regular steranes, the elevated C31 22R homohopane /

13 C30 hopane ratio (> 0.25), the low terrigenous / aquatic ratio of n-alkanes, as well as δ Corg values between -29‰ and -27‰ together suggest a marine depositional environment and a mainly algal / phytoplanktonic source of organic matter. Redox sensitive geochemical parameters indicate mainly dysoxic depositional conditions. - 7 -

Chapter II Upper Cretaceous source rocks of north Lebanon

The samples have high Hydrogen Index values (413–610 mg HC/g TOC) which indicate oilprone Type II kerogen. Tmax values (414 – 432°C) are consistent with other maturity parameters such as vitrinite reflectance (0.25–0.4% VRr) as well as sterane and hopane isomerisation ratios, and indicate that the organic matter is thermally immature and has not reached the oil window. This study contributes to the relatively scarce geochemical information for the eastern margin of the Levant Basin, but extrapolation of the data to offshore areas remains uncertain.

2.2 Introduction

Recent gas discoveries offshore Israel and Cyprus (e.g. Tamar, Leviathan, Aphrodite: Fig. 2.1a) have led to increased interest in the hydrocarbon potential of the Levant Basin (Nader, 2011).

The continental margin of the Levant Basin includes Triassic, Jurassic, Cretaceous and Oligo-

Miocene potential reservoir and source rocks (e.g. May, 1991); however little is known about source rock type and distribution. Upper Cretaceous chalks, marls and shales are prolific potential petroleum source rocks in the eastern Mediterranean region (Henson, 1951; Renouard,

1955; Ala and Moss, 1979; Beydoun, 1988; May, 1991; Sharland et al., 2001; Tannenbaum and

Lewan, 2003; Abed et al., 2005; Gardosh et al., 2006; Nader, 2011), and are proven to be a source for oil in Adyaman province, SE Turkey (Inan et al., 2010) and in the Dead Sea basin

(Tannenbaum and Aizenshtat, 1985). Previous studies have attributed these marine organic matter (OM) rich deposits to a high productivity system associated with extensive coastal upwelling developed during the Late Cretaceous along the SE Tethys margin (Almogi-Labin et al., 1993, 2012; Ashckenazi-Polivoda et al., 2010, 2011; Schneider-Mor et al., 2012). The high flux of nutrients provided by upwelling currents led to the widespread deposition of OM-rich carbonates, cherts and phosphorites (Edelman-Furstenberg, 2008; Ashckenazi-Polivoda et al.,

2010), which are interpreted to have been deposited over a folded topography resulting from the early pulses of the Afro-Arabia/Eurasia plate collision (Bein et al., 1990; Soudry, 2000; - 8 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Abed et al., 2005; Almogi-Labin et al., 2012). The influence of palaeostructure and pre-existing basin architecture is expressed in lateral variations in thickness, lithofacies and OM content which are reported in Israel (Bein et al., 1990; Edelman-Furstenberg, 2009; Almogi-Labin et al., 2012) and Lebanon (Müller et al., 2010; Hawie et al., 2013a). Quantifying these variations and characterising the various organo-facies exposed onshore in the Levant margin will help to reconstruct the palaeogeographic setting, and to assess the extent of intervals with source rock potential in the offshore Levant Basin.

Fig. 2.1a. Regional map showing the Levant Basin and the onshore Levant margin (after Hawie et al., 2013b). B. Beirut; ECB: Eratosthenes Continental Block.

The Levant Basin (Fig. 2.1a) is thought to have developed as a result of Triassic to mid-Jurassic rifting associated with the fragmentation of the northern margin of Gondwana and the opening of the Neo-Tethys, with subsequent subsidence and sediment deposition during the Late

- 9 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Jurassic to Early Cretaceous (Gardosh et al., 2010). A change in the regional tectonic regime occurred in the Late Cretaceous, and extension was replaced by compression due to the collision of the Eurasian and Afro-Arabian plates (Robertson, 1998). This inversion resulted in the development of the Syrian Arc fold belt and the initiation of structural features such as the

Palmyride belt in Syria (Krenkel, 1924; Wolfart, 1967; Walley, 1998). A compressional regime persisted in the Levant throughout the Cenozoic (Walley, 1998; Gardosh and Druckman 2006;

Gardosh et al., 2010) until the latest Oligocene when a transpressive phase was initiated, related to the opening of the Red Sea and the break-up of Arabia from Africa. This resulted in the formation of the Levant Fracture System (Fig. 2.1a) which extends from the Gulf of Aqaba to the Taurus Mountains and which includes the Dead Sea, Yammouneh and Ghab fault systems

(Walley, 1998) (Fig. 2.1b).

The aim of this study is to characterise, for the first time, the OM-rich marls and lime-mud to wackestones of Late Cretaceous age (locally referred to as the Chekka Formation) which occur in northern Lebanon in terms of their organic matter quantity, quality and petroleum generation potential, to interpret their depositional conditions, and to discuss the implications for future on and offshore Lebanon. Samples came from four shallow boreholes located near Chekka (Fig. 2.1b) which penetrate an almost 300 m thick Turonian – upper

Campanian succession.

2.3 Stratigraphic framework

The oldest exposed rocks in Lebanon are probably of Early Jurassic age and older rocks have not been penetrated by wells (Nader, 2011). The Jurassic is represented by a thick (up to 1420 m) succession of dolostones and limestones deposited in a carbonate platform setting interrupted by a brief Kimmeridgian – Oxfordian volcanic episode (Renouard, 1955; Nader,

2011). A major regression coupled with local uplift and subaerial exposure at the end of the

- 10 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Jurassic resulted in erosion of these platform carbonates followed by widespread deposition of up to 400 m of fluvio-deltaic sandstones and shallow-marine shales (Nader, 2011). The sandstones pass upward into shallow-marine and reefal Aptian – Albian carbonates with a brief return to siliciclastic deposition during the Albian (Noujaim, 1977; Walley, 2001; Nader, 2011).

The Cenomanian – Turonian is represented by up to 900 m of carbonate platform deposits

(Nader, 2011) culminating in a regional unconformity spanning the upper Turonian to upper

Santonian (Müller et al., 2010) due to regressional cycles and the early pulses of Syrian Arc deformation (Brew et al., 2001; Hawie et al., 2013a). Subsequently in the latest Cretaceous and continuing to the middle Eocene, northern Arabia underwent a renewed phase of subsidence

(Nader, 2011), which together with a global sea-level highstand (Haq et al., 1988), led to drowning of the pre-existing Cenomanian – Turonian platform carbonates (Fig. 2.1c) and the deposition of up to 900 m of outer-shelf mud-supported limestones (mainly mud- to wackestones) and marls (Walley, 1998; Hawie et al., 2013a). In Lebanon, the thickness of the upper Santonian – lower Maastrichtian succession (the Chekka Formation) varies considerably: from ca. 20 m on the eastern flank of the Tyr-Nabatiye plateau (Renouard, 1955), to ca. 300 m near Chekka in northern Lebanon (sampled for this study) (Hawie et al., 2013a), to ca. 580 m near Hasbayya in southern Bekaa (locations in Fig. 2.1b) (Carla Müller, pers. commun.). A prominent middle Eocene to lower Miocene regional unconformity (Müller et al., 2010) is overlain in northern Lebanon by Miocene reefal, conglomeratic and lacustrine deposits

(Dubertret, 1975; Hawie et al., 2013a).

- 11 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Fig. 2.1b. Simplified geological map of Lebanon showing the location of the study area (square) near the city of Chekka. Map modified after Dubertret, 1955.

Fig. 2.1c. Palaeogeographic map of the Levant region showing depositional environments in the Late Cretaceous. Isopachs indicate the thickness of the Coniacian – Campanian interval (modified after Hawie et al., 2013b). The location of Chekka near the study area in northern Lebanon is marked. - 12 -

Chapter II Upper Cretaceous source rocks of north Lebanon

2.4 Materials and methods

2.4.1 Samples

A total of 212 core samples were collected from four shallow boreholes (80 to 100 m) drilled in the Cimenterie Nationale S.A.L. quarry near Chekka city, northern Lebanon (Fig. 2.1b), and were compiled to cover the composite upper Turonian to upper Campanian interval shown in

Fig. 2.2. Boreholes coordinates are: BH #3, N 34°19' 47.658" E 35°45’49.1976"; BH #8, N

34°19’46.9056" E 35°45’21.8304"; BH #26 N 34°19’26.04" E35°45’50.4"; and BH #27 N 34°

19’54.45" E 35°44’44.80". The thickness of the studied succession is 290 m comprising 33 m of Turonian rocks (BH #27), 10 m of upper Santonian rocks (BH #27), 61 m of lower

Campanian rocks (BH #27, #8), and 186 m of upper Campanian rocks (BH #8, #3, #26).

The Turonian interval is mainly composed of grainstones. The upper Santonian – upper

Campanian succession is composed of relatively homogenous lime-mudstones to wackestones whose OM content varies considerably (Fig. 2.2). The studied succession was sampled at a 1 m spacing within dark-coloured OM-rich intervals (79.5 to 257 m), and at a 2 m spacing within light-coloured OM-poor intervals.

2.4.2 Methods

Total organic carbon (TOC) and total inorganic carbon (TIC) contents were measured for all samples using a liquiTOC II analyzer, which enables a direct determination of TOC and TIC in solids by a temperature ramp method without previous acidification. Approximately 100 mg of rock powder was heated (300°C/min) and held at 550°C for 600 seconds for TOC measurement.

Subsequently, the temperature was raised to 1000°C and held for 400 seconds for the TIC measurement. The CO2 released at each heating stage was measured with a nondispersive infra-

- 13 -

Chapter II Upper Cretaceous source rocks of north Lebanon red detector (NDIR) (detection limit 10 ppm; TOC error 0.6%; TIC error 1.7%). TOC and TIC data are presented in Appendix 1.

13 Fig. 2.2. Elemental analyses (TOC, TS, CaCO3), organic carbon isotopes (δ Corg) and TS/TOC ratio plotted versus depth for the Turonian – upper Campanian succession at

Chekka, north Lebanon (based on analytical data in Appendix 1). The CaCO3 values plotted for the Turonian are CaMg(CO3)2 due to the high dolomite content. (Lithological column after Hawie et al., 2013a).

The CaCO3 abundance (wt %) was calculated using the formula: CaCO3 = TIC × 8.333 (Sachse et al., 2012) for the upper Santonian – upper Campanian interval. Based on its high TIC values, resulting in calculated calcite contents greater than 100%, the Turonian appears to be dominated by dolomite; therefore CaMg(CO3)2 = TIC × 7.8333 was used for this interval (Fig. 2.2). A few lower Campanian samples also had high TIC values which may be due to the presence of dolomites. Total sulphur (TS) content was also measured for all samples using a Leco S200 analyzer (detection limit 20 ppm; error < 5%). TS data is presented in Appendix 1. - 14 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Total Fe, V and Ni contents were determined for 15 samples, selected on the basis of their TOC contents which vary from 0.03 wt % to 5.18 wt %, on pressed powder pellets using energy dispersive X-ray fluorescence (Spectro X Lab 2000) equipped with a Pd tube operated at accelerating voltage between 15 and 53kV and current between 1.5 and 12.0 mA (c.f. Sachse et al., 2011).

Rock-Eval pyrolysis (Espitalié et al., 1977) was performed on a selection of 39 samples selected from the organic-matter rich interval (79.5 to 257 m: Fig. 2.2) from the upper Santonian – upper

Campanian with TOC > 1 wt %, covering a range of TOC between 1.04 and 5.69 wt% (Table

2.1), using a Rock-Eval VI instrument. Parameters derived from this analysis were S1, S2, S3,

Tmax, Hydrogen Index (HI), Oxygen Index (OI), and Production Index (PI) (S1 error 8%; S2 error 7%; S3 error 9%).

For organic petrology, methods followed the guidelines described by Taylor et al. (1998). Ten core samples, selected on the basis of their TOC contents which vary from 1.61 to 5.18 wt %, were cut and embedded perpendicular to bedding in a 10:3 mixture of epoxy resin (Araldite®

XW396) and hardener (Araldite® XW397) and dried at 37°C for approximately 12 hours. The sample surfaces were then ground and polished following the procedures described by Sachse et al. (2012).

Vitrinite reflectance measurements were performed on the ten samples at a magnification of

500x in a dark-room using a Zeiss Axio Imager microscope for incident light equipped with a tungsten-halogen lamp (12V, 100W), a 50X/0.85 Epiplan-NEOFLUAR oil immersion objective and a 546 nm filter. Zeiss immersion oil with refraction index ne =1.518; 23°C was used. A leuco-saphire (0.592%) mineral standard was used for calibration. The standard is kept in dust free boxes at constant temperature and humidity. The vitrinite reflectance measurements were performed on randomly oriented macerals under non-polarized light. An attempt was made to

- 15 -

Chapter II Upper Cretaceous source rocks of north Lebanon measure as many points as possible, but no more than 20 vitrinite macerals, on average, were found in each sample. DISKUS Fossil software (Technisches Büro Carl H. Hilgers) was used for data processing.

Maceral analyses were also performed on the ten samples using incident light fluorescence mode, excited by ultraviolet (UV, 280-380 nm) and violet light (380-450 nm) for identification of the liptinite macerals; 500 points were counted in each sample. Vitrinite and inertinite macerals were rare (< 0.2%) in all samples and they were not counted separately in incident white light.

The mass percentage of OM was calculated from the volume percentage of liptinite following

Littke (1993):

휌 (푂푀) 푂푀 푤푡% = 푂푀 푣표푙% × (1) 휌 (푟표푐푘)

The liptinite density ρ (OM) was estimated to be 1.1 g/ cm3, and the rock density 2.5 g/cm3.

TOC wt% was calculated according to Littke (1993):

퐶% 푇푂퐶 푤푡% = 푂푀푤푡% × (2) 100 where C % is the percentage of carbon in the OM, which was estimated to be 70% based on the typical composition of algal OM (Littke, 1993). This approach was used to investigate whether the OM is dominated by visible macerals or by submicroscopic amorphous OM.

For the analysis of saturated hydrocarbons, 5 g of rock powder from 17 selected samples with

TOC > 1 wt % were extracted using ultrasonic treatment and overnight stirring with 40 ml dichloromethane and 40 ml hexane, respectively. The extracts were separated into three fractions by means of silica-gel based liquid chromatography. The aliphatic hydrocarbons were

- 16 -

Chapter II Upper Cretaceous source rocks of north Lebanon eluted with 5 ml of pentane, the aromatic hydrocarbons with 5 ml of pentane : dichloromethane

(40:60), and the polar compounds with 5 ml of methanol.

Gas chromatographic analyses were performed on the aliphatic fractions extracted from the 17 samples. A volume of 1 µl was injected in splitless mode into a Carlo Erba Instrument HRGC

5300 flame ionization gas chromatograph (GC-FID). The GC was equipped with a Zebron ZB-

1 30 m × 0.25 mm × 0.25 µm fused silica column. H2 was used as the carrier gas with a velocity of 40 ml/s. The temperature programme used started at 80°C, was held for 5 min, and was then raised by 5°C/min to 300°C and held for 20 min.

GC-MS analyses were performed on 10 samples, selected on the basis of their TOC contents which vary from 1.04 wt% to 4.77 wt%, with a Finnigan MAT 95 mass spectrometer (MS) linked to a Hewlett Packard Series II 5890 gas chromatograph, equipped with a Zebron ZB-1

30 m x 0.25 mm x 0.25 µm fused silica column. Carrier gas was He with 35 cm/s velocity. The

MS was operated in full-scan mode from m/z 35 to 700 with a scan rate of 0.5 s/decade and an interscan time of 0.1 s (Böcker et al., 2013). The oven temperature was programmed from 80°C

(3 min isothermal time) to 310°C (held 3 min) at 5°C/min. Xcalibur software from Finnigan was used for GC-MS data processing. The identification of the different compounds was based on comparison with published mass spectra (Peters et al., 2005). Molecular ratios were calculated from integrated peak areas.

13 Organic carbon isotopic composition (δ Corg) was measured via sealed-tube combustion (e.g.

Strauss et al., 1992) on 51 samples. Bulk rock powders were decarbonated with HCl, washed with deionized water, dried and subsequently combusted at 850°C for three hours in evacuated sealed quartz tubes with CuO as oxidant. The liberated carbon dioxide was cryogenically purified in a vacuum distillation line and packed into pyrex break-seal tubes. The isotopic analyses were performed using a Finnigan MAT Delta Plus gas isotope mass spectrometer. All organic carbon isotopic compositions are reported in the standard delta notation as per mil - 17 -

Chapter II Upper Cretaceous source rocks of north Lebanon deviation from the VPDB standard. Analytical performance was monitored with international reference material (USGS 24: -16.0 ± 0.2 ‰) and internal laboratory materials. Data are presented in Appendix 1.

All the analytical work was performed at the Institute of Geology and Geochemistry of

13 Petroleum and Coal, RWTH Aachen University, except the Fe and the δ Corg measurements which were carried out at the Institute of Mineralogy and Economic Geology of RWTH Aachen and the Institut für Geologie und Paläontologie, Westfälische Wilhelms-Universität Münster, respectively.

2.5. Results

2.5.1 Lithofacies and sedimentology

The Turonian interval is mainly composed of bioclastic grainstones with considerable content of pellets and peloidal grains, typical of shallow-marine and platform carbonates. Major fossil components in the Turonian limestones are rudists (Hippuritidae and Radiolitidae) and red algae (Nader, 2000). The Turonian grainstones pass up into bioturbated wackestone/mudstones indicating deeper marine conditions, and the drowning of the Cenomanian – Turonian carbonate platform.

The upper Santonian is a glauconite-rich wackestone and marks the onset of a major Late

Cretaceous transgression and drowning event (Hawie et al., 2013a).

The investigated lower Campanian succession consists of intercalated marl and limestone layers. The limestone beds are composed mainly of mudstones and wackestones showing slight variations in matrix grain-size distribution, which are rich in planktonic foraminifera. These limestones display low to moderate degrees of bioturbation. The upper Campanian is composed mainly of massive chalky mudstones that pass upward into intercalated marls and mudstones

- 18 -

Chapter II Upper Cretaceous source rocks of north Lebanon with pyritic nodules and relatively intense bioturbation in the uppermost Campanian. With the exception of bioturbation, the Santonian – Campanian succession shows massive facies with no clear sedimentary stuctures.

(a)

(b)

(c)

Fig. 2.3. (a) Graph of TS versus TOC values for the analysed samples (n = 210) from the Turonian – upper Campanian succession at Chekka; dashed lines are calculated trend lines for samples with TOC between 0.5 and 2.5%, and samples with TOC > 2.5% (“normal marine” line after Berner, 1984). (b) Plot of TS/TOC ratio versus TOC for the analysed samples (n = 210). (c) Plot of total iron (Fe) versus total sulphur (TS) for selected samples (n =15) from the upper Santonian – upper Campanian interval (after Lückge et al., 1996). - 19 -

Chapter II Upper Cretaceous source rocks of north Lebanon

(a)

(b)

Fig. 2.4. Rock-Eval data for a selection of samples (n = 39) from the upper Santonian – upper Campanian succession at Chekka, north Lebanon. (a) plot of HI versus OI; (b) plot of HI versus Rock-Eval Tmax.

2.5.2 Elemental analyses

The measured TOC contents vary around an average of 1.5 wt % with a maximum of 5.69 wt

% (Appendix 1, Fig. 2.2). An OM rich interval is confined to a ca. 150 m thick section within the upper Santonian – mid upper Campanian succession and has average TOC value of 2 wt %

- 20 -

Chapter II Upper Cretaceous source rocks of north Lebanon but includes intervals with lower and higher TOCs (Fig. 2.2). The highest TOC values were observed in the mid-lower Campanian part of the succession (Fig. 2.2). TOC contents are higher in intervals with darker colour which in general show few signs of bioturbation. The relationship between TOC variations and textural and lithological variations within the

Campanian is not clear; however, the OM content appears to decrease in chalky intervals.

Fig. 2.5. Photomicrographs in fluorescence mode incident light (a, b) and in reflected white light (c, d, e, f). (a, b, c, d) vitrinite in red circle, alginite under red arrows, inertinite in blue circle, and foraminifera (fo) (lower Campanian organic matter rich sample); (e) framboidal pyrite in organic matter rich upper Campanian sample; (f) framboidal pyrite in organic matter poor upper Campanian sample.

TS values range from 0.02 to 2.74 wt % with an average of 0.62 wt % (Fig. 2.2), and correlate with TOC (Fig. 2.3a). High TS/TOC ratios were only found for samples with low TOC values

(Fig. 2.3b).

CaCO3 abundances are generally high, typical for the carbonate lithologies, varying between

60 and 100 wt % with the exception of a few samples within the upper Campanian (possibly due to higher clay content) (Fig. 2.2). CaCO3 abundance correlates negatively with TOC and

TS (Fig. 2.2). - 21 -

Chapter II Upper Cretaceous source rocks of north Lebanon

The Fe content varies between 0.03 and 0.64 wt % and is plotted versus TS in Fig. 2.3c. Ni content varies from 21 to 88 ppm (average: 45 ppm). V content varies from < 20 (below detection limit) to 46 ppm, with only one sample (#12/147) with 201 ppm. Ni/V ratio for the samples with detectable amounts of V varies between 0.4 and 2.7 (average: 1.9).

2.5.3 Rock-Eval pyrolysis

Rock-Eval pyrolysis results are shown in Table 2.1 (see page 23). S1 values vary between 0.08 and 0.53 mg/g rock; S2 values between 5.25 and 34.70 mg/g rock; and S3 values between 0.85 and 2.66 mg/g rock. HI values range between 413 and 610 mgHC/g TOC (Fig. 2.4a), OI values between 46 and 98 mg/g TOC (Fig. 2.4a), and PI values range from 0.01 to 0.03 (Table 2.1).

Rock-Eval Tmax values range from 414 to 432°C (Fig. 2.4b and Table 2.1). The very low S1 values (Table 2.1) indicate small amounts of extractable organic matter (EOM).

2.5.4 Organic petrography

Petrographic analyses indicated the presence of a high proportion of liptinite macerals, specifically alginate (Fig. 2.5a, b). The volume of liptinitic macerals, estimated by point counting (assuming that the areal percentage represents the volume percentage), varies between

4 and 14 vol. %. The TOC content calculated from the volume of liptinite based on equations

(1) and (2) matches quite well with the measured TOC values which correlate positively with the liptinite vol.% (R2 = 0.63) (Fig. 2.6). This indicates that most of the OM is present as visible macerals and only a small proportion is present as submicroscopic OM.

Foraminifera and small pyrite grains were found in all samples (Fig. 2.5). The pyrite size distribution appears to vary between samples with high and low TOC values. In high TOC samples, the observed pyrite is small and has a relatively uniform size (less than 10-30 μm); in low TOC samples, the observed pyrite has a higher size variation (up to 200 μm).

- 22 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Vitrinite reflectance (VRr) values, measured on the very few vitrinites found (Fig. 2.5d), varied between 0.25 and 0.4%.

Fig. 2.6. Plot of TOC wt % versus liptinite abundance for selected upper Santonian – upper Campanian samples (n = 10) showing that the main source of OM is liptinite (after Littke, 1993).

Fig. 2.7. Typical aliphatic fraction total ion chromatogram for a representative lower Campanian sample showing the dominance of medium-range n-alkanes.

- 23 -

Chapter II Upper Cretaceous source rocks of north Lebanon

2.5.5 Organic geochemistry

Biomarker ratios were calculated from the results of organic geochemical analyses and were used to interpret the depositional environment and the thermal maturity of the samples. There is a clear dominance of medium range (nC15 – nC19) relative to long chain n-alkanes (nC27 – nC31) (Fig. 2.7). The terrigenous /aquatic ratio, TAR = (nC27 +nC29 +C31 ) / (nC15 +nC17 +C19)

(Peters et al., 2005) ranges from 0.07 to 1.24 with an average of 0.4. The carbon preference index (CPI25-31) (Bray and Evans, 1961; Leythaeuser and Welte, 1969) varies between 1.38 and

2.67 (Table 2.2). Pristane/phytane (Pr/Ph) ratios vary between 0.62 and 1.33 (average: 0.85). In most samples there is a slight dominance of nC17 and nC18 over pristine and phytane respectively (Fig. 2.8, Table 2.2).

The norhopane (C29) / hopane (C30) ratio (Peters et al., 2005) varies from 0.4 to 1.33 (average:

0.73). The C31 22R homohopane/C30 hopane ratio (Peters et al., 2005) is high for all the analysed samples, ranging between 0.42 and 1.45 (average: 0.83). The C31 homohopane 22S/(22S+22R) isomerisation ratio (Peters et al., 2005) is low for all samples, varying between 0.06 and 0.35

(average: 0.19). The only homohopane found is C31, with the exception of one sample which was found to contain C32 and C33 in very low concentration.

The higher abundance of steranes relative to hopanes in all samples is reflected by the high regular steranes/17α-hopanes ratio, ranging between 0.74 and 1.98 with an average of 1.16. The

C29 sterane 20S/ (20S+20R) isomerization ratio varies between 0.26 and 0.46, with an average of 0.36. The C27, C28 and C29 regular steranes distribution is similar in all samples (Fig. 2.9) and has an average of 42% C27, 25% C28, and 33% C29. C30 sterane is detected in some samples in small concentrations.

- 24 -

Chapter II Upper Cretaceous source rocks of north Lebanon

2.5.6 Organic carbon isotopes

13 δ Corg exhibits little variation within the Santonian – Campanian interval with values ranging between -29.37 and -27.11‰ (Fig. 2.2). The Turonian interval shows values around -25‰.

13 δ Corg results seem to show some tendency towards more negative values with increasing TOC

(R2 = 0.29) (Fig. 2.10a) and with increasing HI (R2 = 0.22) (Fig. 2.10b).

Fig. 2.8. Plot of Pr/nC17 versus Ph/nC18 for selected upper Santonian – upper Campanian samples (n = 17). The plot suggests that deposition took place under reducing conditions (after Shanmugam, 1985).

2.6 Discussion

2.6.1 Depositional environment and early diagenetic conditions

TOC contents combined with other geochemical parameters may serve as a proxy for surface palaeoproductivity and OM preservation potential (e.g. Challands et al., 2009). The relatively high TOC contents observed within the upper Santonian – upper Campanian interval (Fig. 2.2) indicate high OM input and/or relatively good preservation of OM. In carbonate systems, a - 25 -

Chapter II Upper Cretaceous source rocks of north Lebanon positive correlation between silicate and organic matter is often observed, in which organic matter productivity is limited by nutrient supply which in turn is related to input of silicate detritus from land (e.g. Sachse et al., 2011). This correlation is not very clear in the dataset presented here, where in some samples the OM content seems to vary independently of the calculated silicate content (Fig. 2.11). This may indicate an additional source of nutrients such as a marine upwelling system, or alternatively may suggest variability of OM preservation, e.g. due to changes in oxygen availability.

Fig. 2.9. Regular steranes ternary diagram for selected upper Santonian – upper Campanian samples (n = 10) suggesting that phytoplankton are the main source of the organic matter present (after Peters et al., 2005).

The petrographic analysis showed the dominance of marine liptinite (alginite) macerals relative to landderived macerals. Liptinite macerals are clearly visible and there was no dominance of the amorphous (submicroscopic) OM which is usually observed in upwelling systems (Littke and Sachsenhofer, 1994). The positive correlation between TOC content and liptinite vol. %

(Fig. 2.6), and the content of organic carbon determined from the liptinite vol. %, indicate that

- 26 -

Chapter II Upper Cretaceous source rocks of north Lebanon the OM in general is present in the form of liptinite macerals which are derived from marine algae. Marine OM is also indicated by the high HI values (Fig. 2.4a). The n-alkanes distribution

(Fig. 2.7) shows a dominance of medium range (

(a)

(b)

13 Fig. 2.10. (a) Graphs of δ Corg versus TOC wt % (n = 51) (regression line calculated 13 without the Turonian samples) and (b) δ Corg versus Rock-Eval HI (n = 19) for selected 13 upper Santonian – upper Campanian samples, showing more negative δ Corg values for better preserved OM.

- 27 -

Chapter II Upper Cretaceous source rocks of north Lebanon

The very low TAR (Table 2.2) also suggests a high contribution of marine OM relative to terrestrial-derived OM. A marine algal/phytoplanctonic source of OM is also indicated by the

C27 – C29 regular steranes distribution (Fig. 2.9), the high regular steranes / 17α- hopanes ratio

(>1) (Moldowan et al., 1985), and the high C31 22R homohopane/C30 hopane ratio (> 0.25)

(Peters et al., 2005).

13 The δ Corg record (Fig. 2.2) shows values which match quite well with the expected values for

Cretaceous marine OM (average -27‰ ) (Dean et al., 1986; Hayes et al., 1999). Some more negative values could be the result of diagenetic processes and/or a small bacterial contribution.

However a significant bacterial contribution would produce even more negative values

(Bouillon and Boschker, 2005). Also, the steranes / 17α-hopanes ratio is high (Peters et al.,

2005) and tends to increase with increasing TOC (Fig. 2.12), suggesting a very small bacterial contribution to the OM.

13 δ Corg results show a tendency to more negative values with increasing TOC (Fig. 2.10a) and

HI (Fig. 2.10b). Dean et al. (1986) reported a similar trend attributed to enhanced preservation

13 of Cretaceous OM, where labile, hydrogen-rich OM appears to have more negative δ Corg. At

13 low TOC values (<0.5 wt %) refractory organic matter with lower HI and more positive δ Corg values seems to dominate.

The positive correlation between TS and TOC (Fig. 2.3a) indicates a marine depositional environment (Berner, 1984) in which the intensity of sulphate reduction depends on the abundance of OM. During bacterial sulphate reduction, a significant percentage of the primary

OM is consumed, as indicated by the stoichiometry of the bacterial sulphate reduction equation

(Littke et al., 1991; Lückge et al., 1996; Lallier-Vergès et al., 1993). However, for low and high values of TOC, the correlation with TS is less pronounced than it is for intervening values (i.e. between 0.5 and 2.5 wt. % TOC). For TOC values > 2.5 wt.% the samples plot below the normal

- 28 -

Chapter II Upper Cretaceous source rocks of north Lebanon marine line (Fig. 2.3a). This trend is usually observed in carbonate depositional systems where there is insufficient iron to bring about appreciable pyrite formation (Berner, 1984).

Fig. 2.11. Triangular plot of original sediment composition (organic matter, silicate, carbonate) calculated after Littke 1993 (n = 210).

Fig. 2.12. Plot of steranes /17α-hopanes versus TOC for selected upper Santonian – upper Campanian samples, indicating little or no bacterial contribution to the OM.

- 29 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Thus, the negative correlation between TOC and TS/TOC (Fig. 2.3b), and the low total Fe values (Fig. 2.3c), especially for samples with TS > 0.5 wt % and high TOC, suggest that at low TOC contents the main sink of reduced sulphate (H2S) is pyrite, while at high TOC contents

OM becomes a more important sink. Thus, we assume that at low-moderate TOC values, H2S reacted with Fe to form pyrite; whereas at high TOC values, Fe became a limiting factor for pyrite formation (e.g. samples that plot below the pyrite line in Fig. 2.3c) and part of the sulphur is therefore incorporated within the OM. Due to the limited sulphur intake into OM (the experimental organic S/C saturation ratio is around 0.25: Dinur et al., 1980; Bein et al., 1990), lower TS/TOC ratios have resulted for samples with high TOC (Fig. 2.3b).

The upper Campanian OM-poor samples (< 0.5 wt % TOC) which have high TS/TOC ratios

(Fig. 2.3b) and scatter above the normal marine line (Fig. 2.3a) must have another explanation.

Possibly during a later diagenetic stage, H2S produced in TOC-rich sediments migrated

(upwards) through the sediments in which reactive iron was still present, allowing the slow growth of pyrite and causing secondary sulphidation.

An observation which supports this model is the size distribution of pyrite crystals in OM-rich as opposed to upper Campanian OM-poor intervals (Fig. 2.5 e, f). In the OM-rich samples, framboidal pyrite occurs with a uniform small size (less than 10-30 μm; Fig. 2.5e). In the upper

Campanian OM-poor interval, large pyrite crystals occur with a more variable size (Fig. 2.5f).

This pyrite size distribution reflects differences in growth patterns (Wilkin et al., 1996;

Hofmann et al., 2000). In the OM-rich intervals, the nucleation and growth of pyrite occurred rapidly within the uppermost sediments, and due to hydrodynamic effects resulted in smaller and less variable framboids compared to that in the upper Campanian OM-poor interval. In this interval pyrite formation started at a later diagenetic stage, resulting in a larger and more variable pyrite size distribution (Wilkin et al., 1996; Hofmann et al., 2000).

- 30 -

Chapter II Upper Cretaceous source rocks of north Lebanon

This indicates that the deposition of the OM-rich intervals took place under conditions that were sufficiently reducing for sulphate reduction and subsequent pyrite formation to have occurred at early diagenetic stages. The deposition of the OM-poor intervals took place under more oxygenated conditions where sulphate reduction might have occurred at later stages.

Thus, the observed TS-Fe-TOC relationship, together with the pyrite size distribution, as well as the negative correlation between bioturbation intensity and TOC content, suggests that the oxygen content in the uppermost sediments and possibly in the bottom waters fluctuated during deposition of the upper Santonian – upper Campanian succession, resulting in different source rock qualities.

The positive correlation between HI and TOC (Table 2.1) indicates enhanced preservation of

OM at high levels of TOC. This enhanced preservation could be the result of increasing productivity leading to increased oxygen depletion through microbial respiration during OM degradation. The Pr/Ph values (Table 2.2) are generally low but do not indicate completely anoxic conditions. The high Ni/V ratios (>1) (Barwise, 1990), as well as the absence of C32–

C35 homohopanes in all analyzed samples, with the exception of one sample from the most OM- rich interval (199.5 m depth) with minor concentrations of C32 and C33, may also indicate an oxygen-depleted rather than a completely anoxic depositional environment (Peters and

Moldowan, 1991; Peters et al., 2005), with possible dysoxia-anoxia at the highest organic matter contents (highest productivity). The very low norhopane C29/C30 hopane ratio (<1) is likewise consistent with an oxygen-depleted depositional setting (Peters et al., 2005).

These results therefore indicate that the source of the OM in the samples analysed is predominantly marine phytoplankton with little contribution from bacteria (suggested by the presence of small amounts of hopanes) and with very little terrigenous contribution. Redox sensitive geochemical parameters indicate oxygen-depleted but not completely anoxic bottom-

- 31 -

Chapter II Upper Cretaceous source rocks of north Lebanon

13 water conditions during deposition. TS-Fe-TOC data, δ Corg data, and biomarker ratios suggest a fluctuation in the oxygen deficiency during the deposition of the rocks in the studied section.

Similar fluctuations have also been reported for the equivalent Mishash Formation in the Negev and central Israel, and have been attributed to variation in primary productivity (Bein et al.,

1990; Almogi-Labin et al., 1993, 2012).

Another factor which could have controlled organic matter preservation was the warm water temperature recorded globally during the Cretaceous. This led to reduced oxygen solubility and thus reduced aerobic degradation of organic matter (Geng and Duan, 2010).

2.6.2 Petroleum potential and maturity

Within the 300 m thick Turonian – upper Campanian succession studied, a 150 m thick upper

Santonian –mid-upper Campanian interval has an average TOC value of 2 wt % and a potential total petroleum yield (S1+S2) of >6 mg/g rock, making it a very good petroleum source rock

(c.f. Peters, 1986). The high HI values, 413 to 610 mg HC/g TOC, classify the kerogen as oil- prone Type II (Fig. 2.4a) (Espitalié et al., 1977; Langford and Blanc-Valleron, 1990). This kerogen classification matches well with microscope observations which show the presence of abundant alginite.

Maturity parameters indicate the thermal immaturity of the analysed samples. Tmax values show that the succession has not reached the oil window for Type II kerogen (Fig. 2.4b). The vitrinite reflectance measurements, even though conducted on a limited number of vitrinite macerals, also indicate the immaturity of the OM.

Calculated CPI25-31 ratios show values significantly >1, likewise indicating the odd-versus-even dominance of n-alkanes resulting from the immaturity of the samples. The isomerisation ratios

- 32 -

Chapter II Upper Cretaceous source rocks of north Lebanon

for C29 regular steranes (< 0.5) and C31 homohopane (< 0.6) show the dominance of R over S epimers, also indicating a low degree of thermal maturity (Peters et al., 2005).

The Pr/nC17 and Ph/nC18 ratios (Fig. 2.8) point towards an early maturity, however, this may be due to enrichment by short-chain n-alkanes (< C20) due to the dominance of algal OM.

2.7 Exploration implications for the Levant basin

This study has recorded the presence of Upper Cretaceous organic-rich carbonates in coastal northern Lebanon on the eastern margin of the Levant Basin. Seismic profiles (Hawie et al.,

2013b) show the extent of the Upper Cretaceous stratigraphic package offshore Lebanon, where it is locally buried to depths of more than 5000 m. A comparable deeply-buried Upper

Cretaceous succession has been investigated in the southern Levant basin, showing a seismic character corresponding to chalks and marls of pelagic to hemi-pelagic origin (Gardosh and

Druckman, 2006) which may have reached maturity in the deep basin (Gardosh et al., 2006).

However, the basinward extent, distribution, and quality of the source rock are still not known.

The major Late Cretaceous transgression placed central Syria, Jordan, Negev and southern Sinai in a middle- to inner-shelf position from the late Turonian to the mid Eocene (Flexer and

Honigstein, 1984; Bein et al., 1990; Hawie et al., 2013a). Lebanon, Galilee, western Israel,

Gaza and northern Sinai were located in an outer-shelf setting at that time (Fig. 2.1c) (Bein et al., 1990; Hawie et al., 2013a). More distal parts of the Levant Basin were in a pelagic to hemipelagic setting in which chalks, shales and gravity-driven deposits are expected to occur

(Hawie et al., 2013a).

The Campanian Mishash Formation in southern Israel (Fig. 2.1c) is a stratigraphic equivalent to the succession investigated in this study. The OM content of the Mishash Formation in the

Negev area is higher (up to 25% TOC) than the OM content of the analysed carbonates in the

- 33 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Chekka area boreholes (Bein et al., 1990; Minster et al., 1992). This observed decrease in source rock properties of Upper Cretaceous deposits from inner shelf locations, e.g. in the Negev area

(southern Israel), to the outer shelf, e.g. at Chekka with minimum water depths of about 200-

300 m (Hawie et al. 2013a), may indicate a restriction of source rock deposition to a productivity belt along the eastern margin of the Levant Basin. It is known from studies of present-day oceans that the highest TOC contents and best kerogen quality may be expected to occur along continental margins at water depths of a few hundred metres, at greater depths, i.e. in more distal parts of margins, TOC values and kerogen quality tend to decrease (Littke et al.,

1998).

Although productivity appears to decrease towards the offshore, another mechanism could contribute to OM accumulation in distal parts of the basin. Lückge et al. (1996) noted that turbidity currents and mass flows are just as important for OM accumulation as productivity, and could result in the transport and accumulation of OM in deep marine areas outside the zone of productivity, as is the case in the Neogene coastal upwelling system offshore Peru. A similar phenomenon has been reported in the Oligocene turbidite system of the offshore Nile Delta

(Villinski, 2013).

Another important implication of the geochemical data presented in this study is the presence of organic sulphur which could lead to kerogen cracking at relatively lower temperatures, and thus hydrocarbon generation and expulsion at earlier stages of thermal maturity.

2.8 Conclusions

Geochemical and petrographic analyses of borehole samples from the upper Turonian – upper

Campanian succession in northern Lebanon show the presence of ca.150 m thick OM-rich (oil- prone Type II kerogen) fine-grained carbonates and marls of late Santonian – mid-late

Campanian age with very good source rock potential. The OM appears to be dominated by - 34 -

Chapter II Upper Cretaceous source rocks of north Lebanon algal/phytoplanktonic material with little bacterial or terrigenous input. No important variations in organofacies were observed across this rock succession. Deposition occurred in an outer- shelf setting under dysoxic conditions with suggested fluctuations in the oxygen content of bottom waters, although there is no evidence for permanent anoxia.

Thermal maturity parameters indicate that the studied succession onshore is not sufficiently thermally mature for petroleum generation and expulsion to have occurred. However the presence of elevated organic sulphur contents may permit early petroleum generation and expulsion.

Comparison of the new data with those from more proximal areas (e.g. in the Negev) indicates enhanced source rock properties in inner shelf areas and decreasing primary productivity towards the offshore. Deep marine locations may have source rock potential as a result of OM transported from the organic-rich continental shelf by mass flows and turbidites.

- 35 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Table 2.1. Rock-Eval pyrolysis data for samples from the Turonian – upper Campanian succession at Chekka, northern Lebanon. Table

2.1 S S S HI OI Depth TOC 1 2 3 T PI Sample (mg/g (mg/g (mg/g max (mg/g (mg/g (m) (%) (°C) (S1/(S1+S2)) rock) rock) rock) TOC) TOC) 12/203 79.5 1.52 0.18 6.40 1.29 432 421 85 0.03 12/217 86.5 1.12 0.11 5.76 1.10 428 514 98 0.02 12/225 90.5 1.69 0.19 8.05 1.48 427 476 87 0.02 12/235 95.5 2.03 0.24 9.83 1.52 429 484 75 0.02 12/242 99.0 2.50 0.16 12.05 1.83 429 482 73 0.01 12/247 101.5 2.70 0.23 13.89 1.72 429 514 64 0.02 12/251 103.5 1.97 0.18 9.45 1.40 424 480 71 0.02 12/255 112.0 1.78 0.15 8.37 1.39 414 470 78 0.02 12/260 115.0 1.27 0.11 5.25 1.09 428 413 86 0.02 12/264 117.0 1.79 0.15 8.59 1.29 425 480 72 0.02 12/268 119.0 2.07 0.18 9.81 1.54 427 474 75 0.02 12/272 121.0 2.41 0.18 12.20 1.63 420 506 68 0.01 12/277 123.5 1.71 0.21 8.98 1.50 429 525 88 0.02 12/288 129.0 3.83 0.33 19.84 2.06 424 518 54 0.02 12/6 131.0 2.37 0.14 11.53 1.49 429 501 65 0.01 12/11 134.0 2.02 0.15 9.83 1.40 428 487 69 0.02 12/18 137.5 1.98 0.14 9.12 1.24 426 461 63 0.02 12/28 144.5 1.91 0.12 10.38 1.23 422 543 65 0.01 12/34 147.5 1.76 0.14 9.93 1.22 422 564 69 0.01 12/39 150.0 3.28 0.24 17.04 1.54 424 519 47 0.01 12/44 152.5 2.72 0.22 12.95 1.81 431 508 71 0.02 12/65 163.5 2.78 0.21 14.46 1.68 427 520 60 0.01 12/71 166.5 2.54 0.16 12.59 1.58 430 502 63 0.01 12/104 183.0 1.62 0.15 7.38 1.14 428 456 71 0.02 12/113 187.5 2.24 0.13 12.01 1.52 430 536 68 0.01 12/119 190.5 1.91 0.11 10.36 1.32 427 542 69 0.01 12/130 196.0 3.51 0.23 20.06 1.67 427 572 48 0.01 12/137 199.5 3.57 0.36 19.97 1.85 423 559 52 0.02 12/140 201.0 5.69 0.53 34.70 2.66 418 610 47 0.02 12/145 203.5 4.90 0.11 10.36 1.32 427 542 69 0.01 12/152 207.0 3.59 0.28 20.66 1.78 428 575 49 0.01 12/158 210.0 3.61 0.34 20.81 1.72 422 576 48 0.02 12/160 211.0 4.70 0.48 27.96 2.21 423 587 46 0.02 12/791 227.5 1.49 0.09 7.45 0.93 425 500 63 0.01 12/820 242.0 1.04 0.11 5.27 0.85 427 506 81 0.02 12/830 247.0 1.24 0.10 5.81 0.91 431 469 74 0.02 12/840 251.0 1.40 0.08 6.86 1.07 422 490 77 0.01 12/846 254.0 2.07 0.08 11.20 1.40 422 541 67 0.01 12/852 257.0 2.24 0.10 10.49 1.02 423 477 46 0.01 - 36 -

Chapter II Upper Cretaceous source rocks of north Lebanon

Table 2.2. Overview of biomarker ratios determined in this study. (Abbreviations: Pr = pristane; Ph = phytane; TAR = Terrigenous Aquatic Ratio; CPI = Carbon Preference Index; Hop = hopane; Str = sterane). Table 2.2

C C 22R Str/ C Hp C Str Pr/ Ph/ TA CPI 29 31 31 29 Sample Pr/Ph norHp/ Hp/C30 17α- 22S/ 20S/ nC17 nC18 R 25-31 C30 Hp Hp Hp (22R+22S) (20R+20S)

12/203 1.33 0.73 0.80 0.04 4.72 12/288 1.03 0.89 1.46 0.17 1.81 12/274 0.73 0.52 1.14 0.54 1.38 0.80 0.50 1.25 0.20 0.45 12/9 0.65 0.27 0.64 0.46 2.00 1.33 0.79 0.78 0.26 0.27 12/31 0.64 0.27 0.62 0.49 2.05 0.71 0.75 1.34 0.17 0.43 12/53 0.63 0.40 1.01 0.07 2.67 12/65 0.72 0.66 1.08 0.56 1.69 0.70 0.65 1.38 0.21 0.39 12/71 0.70 0.38 0.76 0.26 2.23 12/104 0.84 0.55 0.76 0.58 1.62 12/116 0.93 0.62 0.90 1.24 6.96 0.40 1.06 1.36 0.06 0.46 12/137 0.84 0.27 0.43 0.09 1.43 0.94 0.71 0.74 0.35 0.39 12/147 1.03 0.49 0.64 0.23 1.97 0.40 1.13 1.98 0.06 0.26 12/152 0.82 0.28 0.47 0.23 1.94 0.57 0.85 1.01 0.16 0.26 12/160 0.76 0.55 0.79 0.19 1.53 12/820 0.98 1.25 1.40 0.66 1.49 0.84 0.42 0.75 0.34 0.31 12/844 0.86 1.02 1.44 0.77 1.46 12/852 0.96 1.09 1.38 0.25 1.62 0.63 1.45 1.04 0.08 0.40

- 37 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Chapter III

GEOCHEMICAL AND PETROGRAPHIC CHARACTERIZATION OF CAMPANIAN-LOWER MAASTRICHTIAN CALCAREOUS PETROLEUM SOURCE ROCKS OF HASBAYYA, SOUTH LEBANON

Keywords: Levant Basin, petroleum source rocks, kerogen type, carbonates, Upper Cretaceous.

3.1 Abstract

Santonian – Paleocene marls and fine grained carbonates have been sampled in Hasbayya locality, south Lebanon, in order to evaluate their organic matter (OM) content, petroleum source rock potential, and assess their depositional environment. Methods included total organic carbon (TOC), total inorganic carbon (TIC), total sulphur (TS), Rock-Eval pyrolysis, organic and inorganic petrography, X-ray fluorescence (XRF), gas chromatography-mass spectrometry (GC-MS) and Curie-point-pyrolysis-gas chromatography-mass spectrometry

(CP-Py-GC-MS). TOC, Rock-Eval, and vitrinite reflectance (VRr) results reveal excellent immature petroleum source rocks within the Campanian – lower Maastrichtian interval with

TOC up to 11.6 wt%, hydrogen index (HI) up to 872 mg/gTOC, Tmax up to 433°C and VRr average of 0.36%. Biomarker ratios and maceral analysis suggest a marine depositional environment with a dominance of algal as well as submicroscopic OM. Original sediment composition and redox sensitive geochemical parameters suggest deposition of the OM rich intervals under an oxygen minimum zone (OMZ) that was emplaced and controlled by primary productivity and nutrient supply. Pyrolysate composition shows an important content of organic sulphur compounds (thiophenes) increasing with TOC and thus indicating the presence of Type

II and Type IIS kerogen in the analysed sample set, which is consistent with the presence of immature solid bitumen in the Hasbayya region.

- 38 -

Chapter III Upper Cretaceous source rocks of south Lebanon

The data produced in this study, coupled with regional correlations, allow to construct a conceptual depositional model for the Upper Cretaceous OM rich rocks of the eastern

Mediterranean suggesting deposition under a productivity belt localized along the inner and outer shelf leading to a decrease in source rock quality and a shift in kerogen type toward the deeper parts of the Levant Basin.

3.2 Introduction

The distribution of ancient marine organic matter (OM) rich sediments is mostly controlled by paleoceanographic, paleogeographic, and paleoclimatic factors (Kruijs and Barron, 1990;

Littke 1993). During the Late Cretaceous, a relatively high sea level (Haq, 2014) coupled with oceanic upwelling (Parrish and Curtis, 1982; Ashckenazi-Polivoda et al., 2010) and higher sea- water temperatures (Berger, 1979; Arthur and Jenkyns 1981) led to the formation of several productivity belts along the southern Tethys margin and a widespread deposition of OM rich sediments (Parrish and Curtis, 1982; Almogi-Labin et al., 1993, 2012). In the East

Mediterranean region, OM rich carbonates within the Santonian to Maastrichtian interval have been described in Lebanon (Bou Daher et al., 2014), Israel (Almogi-Labin et al., 1993, 2012;

Ashckenazi-Polivoda et al., 2010, 2011; Schneider-Mor et al., 2012), and Jordan (Abed et al.,

2005).

The total organic carbon (TOC) content of these Upper Cretaceous OM rich carbonates reaches values up to 25 wt.% in the Zin Valley (southeastern Israel) (Bein et al., 1990), 20 wt.% in the

Shefela Basin (western and northwestern Israel) and northwestern Jordan (Abed et al., 2005;

Almogi-Labin et al., 2012), and 5 wt.% in northern Lebanon (Bou Daher et al., 2014) (Fig. 3.1).

The OM consists mostly of preserved marine planktonic algae (Abed et al., 2005; Schneider-

Mor et al., 2012; Bou Daher et al., 2014). The hydrogen index varies between 400 and 700 mg

HC/g TOC and may exceed 800 mg HC/g TOC for some samples (Bein et al., 1990; Alsharhan

- 39 -

Chapter III Upper Cretaceous source rocks of south Lebanon and Salah, 1997; Bou Daher et al., 2014). The kerogen in these Upper Cretaceous source rocks is classified as Type II and Type IIS, due to its high organic sulphur content (Bein et al., 1990;

Minster et al., 1992; Abed et al., 2005; Bou Daher et al., 2014). This high organic sulphur content lead to formation of immature solid bitumen across the Levant due to cracking of C-S bonds, e.g. the Dead Sea asphalt in Israel and Jordan, and the Hasbayya asphalt in Southern

Lebanon (Tannenbaum and Aizenshtat, 1985; Connan and Nissenbaum, 2004). With the exception of the Dead Sea Basin, and presumably the Levant Basin, such Upper Cretaceous source rocks have been reported to be thermally immature throughout the Middle East (Bein and Amit, 1982; Gardosh et al., 2008; Bou Daher et al., 2014). The thick Neogene sedimentary cover in the Levant Basin (Fig. 3.1) (Hawie et al., 2013a) would bring these potential source rocks to burial depth and temperature in many parts of the basin where they might be part of a viable thermogenic petroleum system.

Source rock parameters are generally among the least constraint parameters in basin analysis studies. Due to the lack of data, the source rock information is often taken from the nearest sampling location. However, the OM quantity and quality varies considerably as a function of depositional environment, resulting also in different kinetic parameters leading to erroneous estimation of onset, quality, and quantity of petroleum generation in basin modelling studies.

In the Levant basin, the type, quality, and quantity of Upper Cretaceous source rocks are yet unknown due to the lack of well data. The aim of this study is to characterize, for the first time, the OM rich Upper Cretaceous carbonates in Hasbayya, south Lebanon, to assess its depositional environment, and to suggest a conceptual depositional model based on lithofacies and organofacies variations observed in regional equivalent rock successions in order to provide a better estimation of the Upper Cretaceous source rocks distribution and quality in the Levant

Basin.

- 40 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Fig. 3.1. Simplified geological map of Lebanon showing the location of the study area and the sampled section (AA’) near the town of Hasbayya. Map modified after Dubertret (1955). AF: Akkar Fault, BF: Batroun Fault, BQF: Beit ed Dine-Qab Elias Fault, DS: Dead Sea, ES: Eratosthenes seamount, N: Negev, RAF: Rachaya Fault, RF: Roum Fault, SB: Shefela Basin, SF: Serghaya Fault, YF: Yammouneh Fault, ZB: Zin Basin.

3.3 Geological setting

Located at the present day northwestern margin of the Arabian plate, Lebanon and the Levant region have been affected by several tectonic events resulting from the interaction with the

Eurasian and the African plates. Rifting along the northern margin of Gondwana, related to the opening of the Neo-Tethys Ocean, started in the mid Permian and lead to the formation of the

Levant Basin and the Palmyra Basin to the west and northeast of Lebanon, respectively (Nader,

2014). Rifting in the Levant Basin continued until the Late Jurassic (Garfunkel, 1989; Hawie et al., 2013a; Montadert et al., 2014). Subsequent cooling and subsidence occurred during the

- 41 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Cretaceous, followed by a change in the tectonic regime in the Late Cretaceous and the onset of a compressional system related to the closure of the Neo-Tethys and the collision of the Afro-

Arabian and Eurasian plates (Robertson, 1998a). This compressional regime persisted until the late Miocene and was associated with the deposition of a thick Cenozoic succession in the

Levant Basin (Nader, 2011; Hawie et al., 2013a). In the Palmyra basin, subsidence started during the Late Permian and was later interrupted by Late Cretaceous crustal shortening and tectonic inversion marking the beginning of Syrian Arc deformation (Ponikarov, 1966;

Chaimov et al., 1992). The latest event in the tectonic history of the Levant region was the propagation of the Levant Fracture System from the Gulf of Aqaba to the Taurus Mountains in the Miocene (Beydoun, 1999).

Fig. 3.2. Simplified paleogeographic map (after Dercourt et al., 2000) of the Levant region in the Late Cretaceous showing the position of the investigated rock succession in this study (Hasbayya) and in Chekka (Bou Daher et al., 2014). - 42 -

Chapter III Upper Cretaceous source rocks of south Lebanon

3.4. Stratigraphic framework

The stratigraphic column of Lebanon is mostly dominated by carbonate rocks (Nader, 2014).

The Jurassic rock succession is the oldest surface exposed unit in Lebanon and it consists of ca.

1500 m of dolostones and limestones deposited in shallow warm waters on a carbonate platform

(Nader et al., 2008). Deposition was briefly interrupted by a phase of Kimmeridgian –

Oxfordian volcanism (Renouard, 1955). At the end of the Jurassic, local uplift and subareal exposure resulted in an erosion event that was followed by deposition of up to 400 m of Early

Cretaceous fluvio-deltaic sandstones and shallow marine shales (Nader, 2014). The return to carbonate deposition occurred gradually during the Aptian – Albian (Noujaim, 1977; Walley,

2001; Nader, 2011). The Cenomanian – Turonian witnessed the deposition of ca. 900 m of platform carbonates (Nader, 2011). The upper Turonian – upper Santonian is marked by a regional unconformity (Müller et al., 2010) resulting from the early pulses of Syrian Arc deformation coupled with regressional cycles (Brew et al., 2001; Hawie et al., 2013b). In the latest Cretaceous, starting at the upper Santonian, drowning of the pre-existing carbonate platform was caused by a global sea-level highstand (Haq et al., 1988; Haq, 2014) and a renewed phase of subsidence of the northern margin of the Arabian plate (Nader, 2011), leading to the deposition of up to 900 m of upper Santonian – middle Eocene outer-shelf (Fig. 3.2) mud- supported limestones and marls (Hawie et al., 2013b; Bou Daher et al., 2014). These deposits include several OM rich intervals, specifically within the Campanian – Maastrichtian succession which varies considerably in thickness (Renouard, 1955; Hawie et al., 2013b; Bou

Daher et al., 2014), OM content (Minster et al., 1992; Almogi-Labin et al., 2012; Bou Daher et al., 2014) and lithofacies across the region (Edelman-Furstenberg, 2009; Bou Daher et al.,

2014). The lower Maastrichtian is truncated by a regional unconformity, and is overlain by upper Paleocene deposits (Hawie et al., 2013b). The Oligocene is almost entirely missing

(Müller et al., 2010), and the Miocene rocks rest unconformably on Eocene chalks, limestones,

- 43 -

Chapter III Upper Cretaceous source rocks of south Lebanon and marls (Nader, 2014) consisting of open marine, reefal, and lacustrine deposits (Dubertret,

1975; Hawie et al., 2013b).

Fig. 3.3. Cross section of the “Jourit Fares” rock succession investigated in this study. Arrows and stars mark the limits of traverse 1 and 2, respectively (see Table 3.1).

3.5 Samples

Samples were collected from Santonian – Paleocene outcrops in the area of Hasbayya, south

Lebanon (Fig. 3.1). A section locally referred to as “Jourit Fares” (Fig. 3.3) was sampled in two parallel traverses (Table 3.1). Samples were collected to represent the organic rich Campanian and Maastrichtian intervals and covering the whole spectrum of TOC content. All beds are dipping 60°NW. Additionally several other samples were collected from various outcrops found around the Hasbayya area. For all samples GPS coordinates are provided in Table 3.1.

Parallel samples were collected for dating purposes by means of calcareous nannofossils. The samples are composed mostly of marly limestones and silicified marly limestones with varying

- 44 -

Chapter III Upper Cretaceous source rocks of south Lebanon

OM content, with the exception of samples 67 and 68 which are solid bitumen found near the old “asphalt” mines in the vicinity of the Hasbayya town. OM rich beds observed in this section have a total thickness of approximately 100 m.

3.6 Methods

3.6.1 Biostratigraphy

Calcareous nannofossils were studied in smear slides under a polarizing microscope at 1250 x magnification. The biozonation used follows Thierstein (1976), Sissingh (1977) and Perch-

Nielsen (1985).

3.6.2 Elemental analysis

Total organic carbon (TOC) and total inorganic carbon (TIC) contents were measured on all samples (Table 3.1) using a temperature ramp method, without previous acidification, in a liquiTOC II analyzer which enables a direct determination of TOC and TIC. Aliquots of ca. 100 mg of rock powder were heated (300°C/min) and held at 550°C for 600 seconds during which the TOC peak was recorded. The temperature was then raised to 1000°C and held for 400 seconds during which the TIC was recorded. The released CO2 at every heating stage was measured with a non-dispersive infra-red detector (NDIR) (detection limit 10 ppm; TOC error

0.6%; TIC error 1.7%) (Bou Daher et al., 2014). The CaCO3 content was calculated, assuming that most of the inorganic carbon (TIC) is locked in calcite, using the formula CaCO3 = TIC x

8.333. Total sulphur (TS) content was measured in all samples using a Leco S200 analyzer

(detection limit 20 ppm; error < 5%).

Total Fe, V, and Ni contents were determined for 17 samples, selected according to their TOC contents, varying from 0.3 wt % to 9 wt %, by means of an X-ray fluorescence device (spectro

- 45 -

Chapter III Upper Cretaceous source rocks of south Lebanon

X Lab 2000) equipped with a lead tube operated at accelerating voltage between 15 and 53 kV and current between 1.5 and 12 mA (c.f. Sachse et al., 2011).

Fig. 3.4. Field photographs of (a) the Santonian-Eocene rock succession exposed in the vicinity of the town of Hasbayya at the sampled “Jourit Fares” section. Dashed line marks traverse 1, arrow marks the valley where traverse 2 was taken (see Table 3.1). (b) The outcropping organic rich Campanian carbonates at “Jourit Fares” section. (c) Solid bitumen found within the Upper Cretaceous rock succession in the vicinity of the town of Hasbayya.

3.6.3 Rock-Eval pyrolysis

Rock-Eval pyrolysis (Espitalié et al., 1977) was conducted on 35 samples selected according to their TOC content, varying from 0.5 to 11.6 wt % (Table 3.2), using a Rock-Eval VI instrument

(Lafargue et al., 1998). Derived parameters were S1, S2, S3 and Tmax. Hydrogen Index (HI),

Oxygen Index (OI), and Production Index (PI) were calculated as shown in table 3.2.

3.6.4 Petrographic analysis

Petrographic studies were undertaken for 16 thin sections representing the major sedimentary facies and varying degree of OM content. Conventional microscopic techniques (optical

- 46 -

Chapter III Upper Cretaceous source rocks of south Lebanon microscope; Nikon ECLIPSE LV 100 POL) including transmitted, incident and fluorescence light viewing have been applied. The examination of thin sections under transmitted light aimed to classify the carbonate rocks according to Folk (1959) and Dunham (1962). Incident light viewing helped in distinguishing pyrite from OM (both of which show opaque patterns under transmitted light). Standard fluorescent light was used in order to distinguish OM in matrix and moldic/vuggy porosity.

Vitrinite reflectance measurements were conducted on nine samples with varying TOC contents between 1.9 wt % and 9.4 wt % (Table 3.3). Sample preparation followed the guidelines described by Taylor et al. (1998). Cuttings were embedded in 10:3 mixture of epoxy resin

(Araldite® XW396) and hardener (Araldite® XW397) and dried at 37°C for approximately 12 hours. This was followed by surface grinding and polishing according to procedures described by Sachse et al. (2012). Random vitrinite reflectance was measured at a magnification of 500x in a dark room with a Zeiss Axio Imager microscope for incident light equipped with a tungsten- halogen lamp (12V, 100W), a 50X/0.85 Epiplan-NEOFLUAR oil immersion objective and a

546 nm filter, and using Zeiss immersion oil with refraction index ne =1.518; 23°C. For calibration, a leuco-saphire mineral standard (0.592%) was used. For data processing, DISKUS

Fossil software (Technisches Büro Carl H. Hilgers) was employed. Maceral analysis was performed on five samples using incident light fluorescence mode, excited by ultraviolet (UV,

280-380 nm) and violet light (380-450 nm) for identification of liptinite macerals and white light for identification of vitrinite and inertinite. Approximately 1200 points were counted in each sample.

Solid bitumen reflectance was measured on the two solid bitumen samples (67; 68) using the same methodology as for measuring vitrinite reflectance.

- 47 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Fig. 3.5. Photomicrographs in transmitted light for samples (a) 28 (TOC = 0.2 wt %) (b) 5 (TOC = 2 wt %) (c) 24 (TOC = 4.2 wt %) and (d) 37 (TOC = 6.2 wt %) showing a gradual change from mud-supported (a) to grain-supported (d) rock matrix with increasing TOC

3.6.5 GC-MS

The organic geochemical analysis of the saturated hydrocarbons was conducted on 14 samples

(Table 3.4). 5 g of bulk rock powder were extracted using ultrasonic treatment and overnight stirring with 40 ml dichloromethane and 40 ml hexane. The extracts were fractionated by means of silica gel liquid chromatography. The aliphatic hydrocarbons were eluted by 5 ml of pentane, the aromatic hydrocarbons with 5 ml of pentane : dichloromethane (2:3), and the polar compounds with 5 ml of methanol. The aliphatic fractions were then injected in a Hewlett

Packard Series II 5890 gas chromatograph, equipped with a Zebron ZB-1 30 m x 0.25 mm x

0.25 μm fused silica column, linked to a Finnigan MAT 95 mass spectrometer. Helium was used as a carrier gas with 35 cm/s velocity. The mass spectrometer operated in full-scan mode from m/z 35 to 700 with a scan rate of 0.5 s/decade and an interscan time of 0.1 s (Böcker et

- 48 -

Chapter III Upper Cretaceous source rocks of south Lebanon al., 2013). The oven temperature was set to 80°C for 3 minutes, then it was raised at a rate of

5°C/min to reach 310°C and held there for 3 minutes. Xcalibur software from Finnigan was used for GC-MS data processing. The different compounds identified were based on comparison with published mass spectra (Peters et al., 2005). Molecular ratios were calculated from integrated peak areas.

3.6.6 CP-Py-GC-MS

Curie-point pyrolysis-GC-MS was performed on 11 samples selected to cover a wide range of

TOC values. Approximately 1 mg of bulk rock powder was pyrolysed at 590°C for 10 seconds in a Fischer GSG CPP 1040 PSC Curie-point pyrolyser coupled to a Fisons 8065 GC instrument connected to a Thermoquest MD 800 MS instrument. A 30 m Zebron ZB-5, 0.25 mm i.d., 0.25 µm film thickness non-polar column was used. The GC oven temperature was set to

40°C and held for 3 min, then raised at a rate of 3°C/min to 310°C and held for 20 min. Helium was used as a carrier gas. The MS operated in a scan range of m/z 35-550, source temperature

200°C, and 70 eV electron impact (EI) (al Sandouk-Lincke et al., 2013). Identification of peaks was based on comparison of MS data with mass spectral libraries.

3.7 Results

3.7.1 Biostratigraphy

The Turonian was identified by the presence of the nannofossils Micula staurophora and

Watznaueria barnesae. Around 30 m of latest Santonian were identified by the presence of the nannofossils Marthasterites furcatus, Reinhardtites anthophorus, Eiffellithus eximius. The lower Campanian was identified by the presence of the nannofossils Broinsonia parca, E. eximius, R. anthophorus and has a thickness of ca. 160 m. The upper Campanian was identified by the presence of the nannofossils B. parca, E. eximius, R. anthophorus, Quadrum trifidum,

- 49 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Quadrum Gothicum and has a thickness of ca. 200 m. The lower Maastrichtian was identified by the presence of the nannofossils Q. trifidum, Q. gothicum and has a thickness of ca. 160 m.

The upper Paleocene was identified by the presence of the nannofossils Heliolithus riedeli,

Fasciculithus tynpaniformis (NP8) and Discoaster multiradiatus, F. Tynpaniformis,

Cyclolithus robustus (NP9). The Eocene was not dated in this study because of barren samples, thus the Paleocene – Eocene boundary was tentatively placed at the boundary between the marly limestone and the limestone (Fig. 3.3).

(a)

(b)

Fig. 3.6. Geochemical log showing TOC, CaCO3, TS, and TS/TOC values for traverse 1 (a) and traverse 2 (b) (see Table 3.1 and Fig. 3.3).

- 50 -

Chapter III Upper Cretaceous source rocks of south Lebanon

3.7.2 Lithofacies and sedimentology

The upper Santonian to upper Paleocene rock succession in Hasbayya consists of chalky and marly limestones with beds varying in thickness between 0.2 and 2 m. The outcropping rock succession is mostly massive, lacking any clear sedimentary structures. The major OM rich interval, which is the main interest of this study, is confined to the Campanian and Maastrichtian succession, which also includes ca. 20 m chert bed with phosphate nodules. The marly intervals are intensely weathered and mostly covered by vegetation (Fig. 3.3). The rest of the section shows an interbedding of OM rich and OM lean layers. The OM rich beds have a total thickness of ca. 150 m, and have a dark grey to black color in fresh cut and white to light grey weathered color (Fig. 3.4).

In terms of inorganic petrography, the investigated 16 samples vary between lime mudstone to foraminiferal packstone rock textures (Fig. 3.5). The texture seems to vary with TOC content, from OM lean mudstones to OM rich grainstone (Fig. 3.5). The rock matrix also changes from typical mud supported into micrograiny recrystallized matrix with increasing TOC.

3.7.3 Elemental analysis

TOC contents vary between 0.2 and 11.6 wt %, and have an average of 3.8 wt % (Table 3.1)

(Fig. 3.6a&b) excluding samples 67 and 68 which have a much higher TOC since they consist of solid bitumen (Table 3.1). The highest TOC values were recorded in the upper Campanian

(Table 3.1). The OM content seems to vary significantly even between beds of similar lithology.

TIC values vary between 8.5 and 12 wt %. Assuming that the inorganic carbon is mostly in calcite, CaCO3 values range from 71 to 100 wt %. TS values vary between 0 and 1.7 wt %, with an average of 0.9 wt % for samples with TOC > 1.5 wt%, showing a positive correlation with

TOC (Fig. 3.7a). Most values plot below the normal marine line (Fig. 3.7a) of Berner (1984).

TS/TOC ratios vary between 0 and 0.32 with the exception of samples 4 and 62 with 0.5 and

- 51 -

Chapter III Upper Cretaceous source rocks of south Lebanon

1.1 ratio, respectively. Total Fe contents are very low (<0.7 wt %), and most of the analysed samples plot below the pyrite line (Fig. 3.7b). Ni and V contents vary between 50 and 150 ppm and show a positive correlation with TOC (Fig. 3.8a). Most samples have a higher content of

Ni than V (Fig. 3.8b). V/(V+Ni) ratio stays relatively uniform around an average of 0.3 with

(a)

(b)

varying TOC (Fig. 3.8b).

Fig. 3.7. (a) Graph of TS vs. TOC for the analysed samples from the Santonian-lower

Maastrichtian succession in Hasbayya. (b) Plot of total iron (Fe) vs. total sulphur (TS) for selected samples from the lower Campanian-lower Maastrichtian interval.

The calculated original sediment composition (OM, silicate, carbonate) shows two different trends (Fig. 3.9). The first trend, which consists of most samples, shows an increase in original - 52 -

Chapter III Upper Cretaceous source rocks of south Lebanon

OM content with increasing silicate content. The second trend, which consists of mostly

Maastrichtian samples, shows very low original OM content independently of silicate content variation.

(a)

(b)

Fig. 3.8. (a) Ni + V vs. TOC (b) V/(Ni + V) vs. TOC showing values <1 indicating that the depositional environment were not anoxic.

3.7.4 Rock-Eval pyrolysis

Rock-Eval results are presented in Table 3.2 and Figures 3.10a&b. S1 values vary between 0.04 and 7.67 mg HC/g rock and have an average of 1.23 mg HC/g rock, with the exception of sample 67 and 68 which have 5.8 mg HC/g rock and 23.90 mg HC/g rock, respectively. S2 - 53 -

Chapter III Upper Cretaceous source rocks of south Lebanon values vary between 1.37 to 82.98 mg HC/g rock and an average of 40.9 mg HC/g rock, except for samples 67 and 68 which have 155.16 and 463.19 mg HC/g rock, respectively. S3 values vary between 1.22 and 4.92 mg CO2/g rock and have an average of 2.13 mg HC/g rock. HI values vary between 257 and 872 mg HC/g TOC and have an average of 620 mg HC/g TOC.

HI values seem to increase with increasing TOC content (Fig. 3.10a). OI values vary between

10 and 242 mg CO2/g TOC with an average of 57 mg CO2/g TOC. Highest OI values are recorded in samples with low TOC content (Table 3.2). Tmax values vary between 405 and

434°C and have an average of 418°C. Figure 3.10b shows a decrease in Tmax with increasing

TS. PI values vary between 0.01 and 0.1 with an average of 0.02.

Source rock potential index (SPI) (Demaison and Huizinga, 1991) was calculated for this source rock using the average S1+S2 of the analysed samples and a rock density of 2.5 g/cm³, resulting in SPI = 9.7 t HC/m².

Fig. 3.9. Triangular plot of original sediment composition (OM, silicate, carbonate) calculated after Littke (1993).

- 54 -

Chapter III Upper Cretaceous source rocks of south Lebanon

3.7.5 Organic petrology

The dominance of marine liptinite macerals, specifically liptodetrinite, is clearly seen in figure 3.11a-f. This dominance was seen in all analysed samples. Terrestrial macerals were also found, however, in very small amounts, in addition to some immature solid bitumen (Fig. 3.11a-f). With the exception of sample 59, no more than 47 vitrinite macerals were found (Table 3.3). The number of vitrinites identified seem to be independent of the TOC content (Table 3.3). Random vitrinite reflectance values vary between 0.32 and 0.44 % and have an average of 0.36 % (Table 3.3).

(a)

(b)

Fig. 3.10. Rock-Eval data for a selection of samples from the lower Campanian-lower

Maastrichtian succession at Hasbayya; (a) plot of TOC vs. S2; (b) plot of HI vs. Tmax. - 55 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Foraminifera show a great abundance in all the analysed samples (Fig. 3.11). Additionally, pyrite framboids were also identified in small amounts and in sizes that rarely exceed 10 µm

(Fig. 3.11a-f) except for samples 4, 5 and 62 which show much larger pyrite crystals (>100 µm)

(Fig. 3.11g&h).

Solid bitumen reflectance measurements on samples 67 and 68 show a reflectance of 0.10 % and 0.11 %, respectively. Calculations of equivalent vitrinite reflectance according to Jacob

(1989) result in 0.46 % and 0.47 %, respectively, and according to Schoenherr et al. (2007) result in 0.33 % and 0.34 %, respectively.

The volume % of OM, inferred from the area % of OM (observed in macerals), does not correlate with the measured TOC content (Fig. 3.12). Most of the analysed samples plot above the TOC-OM correlation line inferring an important contribution of submicroscopic

(unstructured) OM to the total OM (Fig. 3.12).

3.7.6 Organic geochemistry

Ratios of different biomarkers were calculated and used for the interpretation of depositional environment and thermal maturity of the studied rock succession. A dominance of short chain n-alkanes (nC14 – nC19) relative to long chain n-alkanes (nC27 – nC31) is seen in all analysed samples (Fig. 3.13). Calculated terrigenous/aquatic ratios, TAR = (nC27 +nC29 +C31)/(nC15

+nC17+C19) (Peters et al., 2005) vary between 0.02 and 0.81 (Table 3.4). The carbon preference index (CPI25-31) varies between 0.65 and 7.24 (Table 3.4). Pristane/phytane (Pr/Ph) ratios range from 0.36 to 0.95 (Table 3.4). Pr/nC17 and Ph/nC18 ratios are plotted in figure 3.14 and compared to results from north Lebanon reported by Bou Daher et al. (2014). Identified steranes and hopanes are listed in Table 3.5. Figures 3.15 and 3.16 indicate a great similarity with respect to hopanes and steranes of lower Campanian, upper Campanian, and lower Maastrichtian. The

C27, C28, and C29 regular steranes relative abundance is similar in almost all samples indicating

- 56 -

Chapter III Upper Cretaceous source rocks of south Lebanon a phytoplanctonic source of the OM with a minor influence of higher plants (Fig. 3.17). The steranes isomers are dominated by ααα20R and βαα20R configurations in all samples.

Norhopane (C29)/hopane (C30) ratio varies between 0.2 and 0.8.

Fig. 3.11. Photomicrographs in reflected white light (a, c, e, g, h) and in fluorescence mode incident light (b, d, f) of samples 44 (a, b), 59 (c, d), 34 (e, f), 4 (g), and 5 (h). Liptinite under arrow, vitrinite in white rectangle, semi-fusinite in white ellipse, “P” for pyrite, “F” for foraminifera, “B” for immature solid bitumen. - 57 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Pyrolysate composition for 6 samples analysed by CP-Py-GC-MS is shown in partial chromatographs in figure 3.18. The pyrolysates of all the analysed samples are dominated by organic sulphur compounds, specifically thiophene and 2-methylthiophene (Fig. 3.18). Ratios of thiophene/benzene, 2-methylthiophene/toluene, and 2,3-dimethylthiophene/1,3- dimethylbenzene were calculated and all show a positive correlation with TOC (Fig. 3.19).

Fig. 3.12. Plot of TOC (wt%) vs. visible OM (macerals) (vol.%). The density of marine OM (ρOM) is assumed to be 1.1 g/cm3, the rock density (ρrock) 2.5 g/cm3, and the percentage of carbon in marine OM (C%) 70% (Littke, 1993).

3.8 Discussion

3.8.1 Depositional environment

The nannofossil datings of the studied Upper Cretaceous rock succession in south Lebanon revealed thickness variation in several units compared to Upper Cretaceous section reported in north Lebanon (Hawie et al., 2013b; Bou Daher et al., 2014). The Campanian unit is approximately 110 m thicker in south Lebanon compared to north Lebanon (Hawie et al.,

- 58 -

Chapter III Upper Cretaceous source rocks of south Lebanon

2013b; Bou Daher et al., 2014). This thickness variation might be due to the preexisting topography possibly due to the early pulses of the Syrian Arc deformation (Walley, 1998; Brew et al., 2001; Hawie et al., 2013b; Nader, 2014) as reported in Israel (Edelman-Furstenberg,

2009; Almogi-Labin et al., 2012). The observed thickness variation for the lower Maastrichtian, ca. 160 m in Hasbayya compared to ca. 10 m in north Lebanon (Hawie et al., 2013b), could be the result of different depositional thicknesses and/or different erosional amounts (thicker eroded thicknesses from the lower Maastrichtian in north Lebanon compared to south

Lebanon).

The TOC content can serve as a first proxy for evaluation of paleoproductivity and preservation potential. High TOC contents as those reported in the studied rock succession indicate either high paleoproductivity at moderate to high preservation potential due to oxygen-depleted or anoxic bottom water conditions, or very good preservation potential due to anoxic bottom water at moderate to high bioproductivity (Demaison and Moore, 1980). High productivity and increased OM flux to the sediments lead to enhanced oxygen-depletion due to oxidation of organic matter in deeper water, thus enhancing preservation (Reimers and Suess, 1983;

Pedersen and Calvert, 1990). A partly productivity-controlled depositional system is indicated by the observed increase in HI with increasing TOC (Fig. 3.10a). In silled basins in which limited water circulation leading to anoxic bottom water controls organic matter deposition, such a clear correlation would not be expected, because all organic matter would be well preserved and hydrogen-rich. In contrast, the Cretaceous sequence onshore Lebanon is characterized by higher hydrogen content (as indicated by HI values) in the more TOC-rich intervals; the latter represent the optimum conditions for organic matter preservation probably at high sedimentation rates (Littke, 1993).

The increase in the amount of foraminifera and other skeletal grains with increasing TOC (Fig.

3.5) could indicate a productivity controlled source rock deposition. Enhanced continental shelf - 59 -

Chapter III Upper Cretaceous source rocks of south Lebanon productivity is usually triggered and controlled by nutrient supply. The lack of correlation between TOC and terrestrial macerals such as vitrinite (Table 3.3), and the very small amount of Fe might indicate a weak terrestrial nutrient source and a strong upwelling-related nutrient source. The presence of submicroscopic OM (Fig. 3.12) might also suggest deposition of OM in an upwelling zone, since these deposits are usually characterized by abundant submicroscopic OM (<1 µm; Littke and Sachsenhofer 1994; Lückge et al., 1996).

Fig. 3.13. Typical aliphatic fraction total ion chromatogram for (a) lower Maastrichtian (b) upper Campanian (c) lower Campanian. - 60 -

Chapter III Upper Cretaceous source rocks of south Lebanon

The calculated original sediment composition (Fig. 3.9), showing an increase in original OM with increasing silicate content, indicates that carbonate is a diluent to OM and that calcareous plankton is not the main contributor to the OM. In marine, carbonate-dominated environments, high silicate contents reflect commonly enhanced nutrient supply either derived from terrestrial sources such as river systems or in silica enriched cool water upwelling systems (Littke 1993).

Thus the positive correlation provides further evidence for a productivity-driven depositional system. An exception are Maastrichtian samples, for which this clear relationship is not valid.

Maastrichtian sediments are generally poor in organic matter (Fig.3. 6a,b) and might reflect open ocean conditions without any strong oxygen-depletion of bottom waters.

Fig. 3.14. Plot of Pr/nC17 vs. Ph/nC18 for selected lower Campanian-lower Maastrichtian samples from Hasbayya indicating slightly more reducing depositional bottom water conditions compared to Chekka (Bou Daher et al., 2014) (Chapter II).

The n-alkane (TAR<1) as well as the regular sterane distribution (Figs. 3.16, 3.17) show a dominance of algal and phytoplankton-derived marine OM over terrestrial OM (Table 3.4).

This dominance seems to be similar in all the analysed samples, with only small variations in the terrestrial input especially in the Maastrichtian (Fig. 3.13). The same pattern has been

- 61 -

Chapter III Upper Cretaceous source rocks of south Lebanon deduced from the petrographic analysis (Fig. 3.11b,d), showing a clear dominance of marine liptinite compared to other maceral groups, thus confirming a marine depositional environment with very little terrestrial influence, typical for carbonate-dominated continental margin OM rich deposits (e.g. Sachse et al., 2011, 2014; Bou Daher et al., 2014).

Fig. 3.15. Hopane distribution for selected (a) lower Maastrichtian (b) upper Campanian (c) lower Campanian samples (see Table 3.5 for a list of identified hopanes).

The low Pr/Ph ratios (<1) as well as the Pr/nC17 and Ph/nC18 (Fig. 3.14) ratios suggest deposition under reducing bottom water conditions that were slightly more oxygen depleted than in case of the Upper Cretaceous in north Lebanon (Bou Daher et al., 2014). This difference is consistent with the TOC difference between both sections: only up to 5% TOC is recorded in northern Lebanon (Bou Daher et al., 2014), indicating a higher primary productivity and thus also higher oxygen-depletion in Hasbayya, southern Lebanon (up to 11.6 wt. % TOC; Table

3.1). The occurrence of bedded cherts and phosporite deposits is usually associated with high organic productivity (Arthur and Jenkyns, 1981; Hein and Obradovis, 1989). The higher primary productivity in Hasbayya is also reflected by the presence of a 20 m thick chert bed and phosphorite nodules compared to much thinner chert bands (10-20 cm) in Chekka (Hawie et al., 2013b; Bou Daher et al., 2014). This considerable difference in primary productivity is - 62 -

Chapter III Upper Cretaceous source rocks of south Lebanon most probably due to a more proximal paleoposition at Hasbayya (Fig. 3.2), which is also suggested by larger terrestrial macerals (up to 50 µm) compared to Chekka (up to 10 µm) (Fig.

3.11) (Bou Daher et al., 2014). A similar increase in productivity from relatively distal sections

(e.g. Shefela Basin) to relatively proximal sections (e.g Zin Basin) (Fig. 3.1) accompanied by an increase in bottom waters oxygen deficiency is reported in Israel (Bein et al., 1990; Almogi-

Labin et al, 1993, 2012; Edelman-Furstenberg, 2008, 2009). This trend is also observed in

Quaternary and modern day analogues (e.g. offshore Peru, Oman and Pakistan; Lückge et al.,

1996; Paulmier & Ruiz-Pino, 2009; Thamdrup et al., 2012)

Results from the Ocean Drilling Program (Leg 160, Hole 967E) show that the Campanian cherts and phosphorites, that are present in the Levant margin (Almogi-Labin et al, 1993, 2012;

Edelman-Furstenberg, 2008, 2009; Hawie et al., 2013b; Bou Daher et., al 2014), are absent from the Eratosthenes seamount (Fig. 3.1). This was attributed to the paleoposition of the

Eratosthenes seamount beyond the marginal zone of upwelling and high biological productivity

(Robertson, 1998b). Only a rare occurrence of OM rich Campanian sediments was recorded in

Hole 967E, and was interpreted to reflect short lived drop in the oxygen minimum zone (OMZ) to intersect the seafloor of the Eratosthenes seamount during periods of highest biological productivity in the Late Cretaceous (Robertson, 1998b).

Fig. 3.16. Steranes distribution for selected (a) lower Maastrichtian (b) upper Campanian (c) lower Campanian samples (see Table 3.5 for a list of identified steranes).

- 63 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Whether completely anoxic bottom water conditions were ever established remains, however, questionable. The amounts of Ni and V (Fig. 3.8a) as well as the V/(Ni+V) ratio (Fig. 3.8b) indicate reducing conditions that were not completely anoxic (Barwise, 1990; Tribovillard et al., 2006). This is further supported by the low norhopane (C29)/hopane (C30) (<1) (Table 3.4) and the absence of C33–C35 homohopanes (Fig. 3.15) (Peters et al., 2005). Additionally, the mono-aromatic fraction of 5 samples with high TOC was analysed with GC-MS and no isoreniaratane was detected in any of the samples, suggesting that anoxic conditions in the photic zone were not established (Peters et al., 2005).

Fig. 3.17. Regular steranes ternary diagram for selected lower Campanian-lower Maastrichtian samples and the Hasbayya solid bitumen, indicating a dominance of phytoplanktonic OM in all samples, and a rather uniform composition.

The TS/TOC ratio is often used as a proxy for evaluation of depositional environment (Berner,

1984; Bein et al., 1990). In completly anoxic marine environments, ratios tend to plot above the

“normal marine line” (Fig. 3.7a), because pyrite is already formed by bacterial sulphate reduction within the water column and at the sea floor. This process leading to very high

- 64 -

Chapter III Upper Cretaceous source rocks of south Lebanon

TS/TOC ratios is not possible in oxygenated bottom water, even if oxygen concentration is low.

For our sample set, the positive correlation between TS and TOC (Fig. 3.7a) indicates a marine depositional environment where the intensity of sulfate reduction depends on the availability and abundance of metabolizable OM (Littke et al., 1991; Lückge et al., 1996; Lallier-Vergès et al., 1993). However, the analysed samples plot under the “normal marine line” (Berner, 1984).

This trend is often observed in OM rich marine carbonates due to iron shortage (e.g. Sachse et al., 2011; Bou Daher et al., 2014). Figure 3.7b shows that most of the samples analysed plot under the pyrite line and thus have a shortage of reactive iron relative to reduced sulfate. The excess reduced sulfate would then partly be incorporated into the OM as organic sulphur. This organic sulphur would have implications on the kerogen type as well as kinetics of hydrocarbon generation as discussed below. The larger pyrite size (Fig. 3.11g&h) and the relatively high

TS/TOC ratio in some OM poor upper Campanian and lower Maastrichtian samples (4, 5, and

62; see Fig. 3.11g,h) suggest a late diagenetic pyrite origin for these exceptional samples.

Accordingly, these samples were not used for calculation of original sediment composition (Fig.

3.9) due to their diagenetic overprint. Possibly excess hydrogen sulphide migrating upward from deeper sediments caused this diagenetic precipitation.

In summary most geochemical indicators suggest oxygen-deficient but not completely anoxic bottom water conditions during deposition, i.e. a productivity controlled organic matter accumulation similar to present-day upwelling system offshore Peru rather than a silled basin similar to the present-day Black Sea. Additionally, comparing the OM content, facies, and depositional conditions reported in the studied rock succession in Hasbayya to results reported in regional equivalent sections studied by several authors (Table 3.6) (Bein et al., 1990; Almogi-

Labin et al., 1993, 2012; Abed et al., 2005; Edelman-Furstenberg, 2009; Bou Daher et al., 2014) suggests an increased intensity of productivity and thus preservation towards the inner shelf

(present day Negev, central Israel, and NW Jordan) reflected by higher OM content and more

- 65 -

Chapter III Upper Cretaceous source rocks of south Lebanon reducing depositional conditions. Consequently, we propose a conceptual depositional model, illustrated in figure 3.20, which suggests a decreasing source rock quality and a potential change in kerogen type towards the deeper parts of the basin.

Fig. 3.18. CP-Py-GC-MS partial chromatographs showing pyrolysate composition of 6 samples with varying TOC content. (a) Sample #60 (b) sample #22 (c) sample #17 (d) sample #6 (e) sample #5 (f) sample # 51. “B” Benzene, “T” Thiophene, “To” Toluene, “MT” Methylthiophene, “DMT” Dimethylthiophene, “DMB” Dimethylbenzene, “TMT” Trimethylthiophene, “EMT” Ethylmethylthiophene.

- 66 -

Chapter III Upper Cretaceous source rocks of south Lebanon

3.8.2 Petroleum potential, maturity, and kerogen type

Thickness, TOC content, and HI of the studied succession indicate excellent petroleum source rock properties (Espitalié et al., 1977; Langford and Blanc-Valleron, 1990). The Rock-Eval pyrolysis-based kerogen classification indicates oil prone Type II kerogen derived from marine

OM thus matching the microscopic observations (Fig. 3.11) and the molecular data (Fig. 3.13).

Maturity parameters, such as VRr and Tmax, indicate that the analysed source rocks have not reached thermal maturity for hydrocarbon generation from Type II kerogen. This is further supported by the dominance of the thermally unstable, biologically derived, 20R isomers and absence of the thermally stable 20S configuration of C29 regular steranes (Fig. 3.16), and

CPI25-31 that is generally high (>1) (Table 3.4) (Peters et al., 2005).

Fig. 3.19. Plot of thiophene/benzene, 2-methylthiophene/toluene, and 2,3- dimethylthiophene/1,3-dimethylbenzene vs. TOC indicating an increase in organic sulphur with increasing TOC

- 67 -

Chapter III Upper Cretaceous source rocks of south Lebanon

The TOC-TS-Fe values discussed above suggest a high organic sulphur content. This is confirmed by the CP-Py-GC-MS data (Fig. 3.18) showing a high abundance of thiophenes relative to other non-sulphur compounds, suggesting a Type IIS kerogen. Furthermore, the positive correlation between TOC values and thiophene/benzene, 2-methylthiophene/toluene, and 2,3-dimethylthiophene/1,3-dimethylbenzene (Fig. 3.19) suggest a higher amount of organic sulphur with increasing TOC. This would have important implications on petroleum generation kinetics, as it is well known that sulphur-rich tend to have wider activation energy distributions with a low mean value (Tissot et al., 1987; di Primio and Horsfield, 1996). This effect is clearly seen in figure 3.10b, showing a decrease in Tmax with increasing TS and HI implying that samples with high TS (i.e. high TOC) would produce hydrocarbons at considerably lower temperatures (~60°C). This fact explains the presence of large amounts of immature solid bitumens in the area of Hasbayya (Connan and Nissenbaum, 2004).

Fig. 3.20. Conceptual depositional model (not to scale) of the Upper Cretaceous organic matter rich carbonates of the east Mediterranean.

- 68 -

Chapter III Upper Cretaceous source rocks of south Lebanon

3.9 Conclusions

The data presented in this study provide important information concerning the depositional environment and paleogeographic setting at the time of deposition of the investigated Upper

Cretaceous source rocks. Lateral changes throughout the region show an increasing primary productivity coupled with increasing oxygen deficiency from distal sections, e.g. Chekka (Bou

Daher et al., 2014) and Hasbayya, to relatively more proximal, e.g. Shefela Basin, Zin Valley, and the Negev (Bein et al., 1990; Almogi-Labin et al, 1993, 2012; Edelman-Furstenberg, 2008,

2009) (Figs. 3.1, 3.2). Intervals with high TOC reflect high primary productivity correlating also with high organic sulphur contents leading to a shift from Type II to Type IIS kerogen.

The geochemical data assembled for Chekka (Bou Daher et al., 2014), and Hasbayya, and in

Israel (Bein et al., 1990; Almogi-Labin et al, 1993, 2012) suggest open ocean conditions with oxygen-deficient but not anoxic bottom waters. These observations allow us to suggest a conceptual depositional model for the Upper Cretaceous source rocks of the eastern

Mediterranean whereby decreasing source rock properties and a shift in kerogen type toward the deeper parts of the Levant Basin are suggested. The TOC content of these Upper Cretaceous rocks in the deep basin, where these rocks could be part of a conventional thermogenic petroleum system, would then be controlled by the amount and quality of OM reaching the sediment water interface, which is in turn controlled by the intensity and extent of the oxygen minimum zone (OMZ) and the basin’s bathymetry.

- 69 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Table 3.1 List of all analysed samples showing values of TOC, TIC, TS, age, and GPS coordinates. *Height above base was measured in the field by means of a Jakob’ staff (base of each section is 0). **Two parallel traverses were made across the organic rich succession in order to sample more outcrops (see Fig. 3.3). ***Solid bitumen samples were collected from around the old Hasbayya asphalt mines. Sample TOC TIC TS Age GPS Coordinates Stratigraphic height 1 (wt%)0.2 (wt%)11.3 (wt%)0.0 Lower Maastrichtian N33°26'14.48" E35°41'17.09" above495.6 base (m)* 2 0.2 11.1 0.0 Lower Maastrichtian N33°26'13.94" E35°41'17.70" 474.6 3 0.2 10.9 0.0 Lower Maastrichtian N33°26'13.72" E35°41'17.74" 470.4 4 0.8 9.8 0.4 Lower Maastrichtian N33°26'13.45" E35°41'18.00" 460.6 5 2.0 10.4 0.6 Lower Maastrichtian N33°26'13.13" E35°41'18.26" 446.6 6 2.2 10.8 0.2 Upper Campanian N33°26'12.05" E35°41'22.70" 387.8 7 9.4 9.3 1.7 Upper Campanian N33°26'12.05" E35°41'22.70" 387.8 8 1.0 11.2 0.1 Upper Campanian N33°26'12.05" E35°41'22.70" 387.8 9 7.3 10.3 1.1 Upper Campanian N33°26'11.78" E35°41'22.96" 373.8 10 0.3 12.0 0.0 Upper Campanian N33°26'09.72" E35°41'26.02" 306.6 11 0.5 12.0 0.0 Upper Campanian N33°26'09.37" E35°41'26.53" 291.2 12 5.6 10.9 0.6 Upper Campanian N33°26'09.30" E35°41'26.65" 287.0

13 0.4 11.9 0.0 Upper Campanian N33°26'09.11" E35°41'27.39" 267.4 14 2.5 11.6 0.3 Upper Campanian N33°26'09.11" E35°41'27.39" 267.4 15 0.5 11.8 0.0 Upper Campanian N33°26'08.81" E35°41'27.58" 260.4 16 0.4 11.7 0.0 Upper Campanian N33°26'08.78" E35°41'27.63" 257.6

17 3.3 11.4 0.3 Upper Campanian N33°26'08.75" E35°41'27.79" 253.4 Traverse 1** 18 3.1 11.3 0.3 Upper Campanian N33°26'08.72" E35°41'27.84" 249.2 19 0.2 11.1 0.0 Upper Campanian N33°26'08.55" E35°41'28.04" 240.8 20 5.3 10.1 0.7 Upper Campanian N33°26'08.40" E35°41'28.11" 236.6 21 0.3 11.6 0.0 Lower Campanian N33°26'07.82" E35°41'28.25" 225.4 22 3.5 10.9 0.5 Lower Campanian N33°26'07.63" E35°41'28.49" 215.6 23 6.7 9.4 0.9 Lower Campanian N33°26'06.29" E35°41'30.24" 159.6 24 4.2 9.6 0.7 Lower Campanian N33°26'06.09" E35°41'30.39" 154.0 25 5.6 10.8 0.7 Lower Campanian N33°26'05.29" E35°41'30.53" 135.8 26 4.9 10.8 0.6 Lower Campanian N33°26'04.67" E35°41'31.49" 109.2 27 6.9 9.8 1.3 Lower Campanian N33°26'04.71" E35°41'31.77" 102.2 28 0.2 12.0 0.0 Lower Campanian N33°26'03.27" E35°41'33.51" 47.6 29 0.2 12.0 0.0 Santonian N33°26'02.40" E35°41'34.90" 0* 30 7.5 9.8 1.0 Lower Maastrichtian N33°26'19.07" E35°41'26.65" 358.4 31 9.8 9.6 1.5 Upper Campanian N33°26'16.05" E35°41'28.92" 259.0 32 11.6 9.3 1.6 Upper Campanian N33°26'15.81" E35°41'29.46" 245.0 33 5.1 10.7 0.8 Upper Campanian N33°26'15.64" E35°41'29.45" 233.8 34 8.5 9.3 1.6 Upper Campanian N33°26'15.69" E35°41'29.59" 232.0 35 8.6 9.9 1.4 Upper Campanian N33°26'15.41" E35°41'29.75" 229.0 36 6.2 10.0 0.9 Upper Campanian N33°26'15.22" E35°41'29.90" 226.8 37 6.3 10.2 1.1 Upper Campanian N33°26'14.51" E35°41'30.81" 196.0 38 0.4 11.4 0.0 Upper Campanian N33°26'14.51" E35°41'30.81" 196.0

39 0.5 11.6 0.0 Upper Campanian N33°26'14.09" E35°41'31.18" 180.0 40 3.6 11.2 0.5 Upper Campanian N33°26'13.73" E35°41'31.54" 165.0 41 4.4 10.8 0.7 Lower Campanian N33°26'13.23" E35°41'32.24" 150.0 42 7.9 9.8 0.9 Lower Campanian N33°26'13.08" E35°41'32.41" 134.4

43 3.8 10.8 0.4 Lower Campanian N33°26'11.63" E35°41'34.85" 56.0 Traverse 2** 44 6.7 9.0 1.3 Lower Campanian N33°26'11.46" E35°41'35.15" 49.0 45 6.3 9.3 1.2 Lower Campanian N33°26'11.28" E35°41'35.29" 40.0 46 6.9 9.3 1.2 Lower Campanian N33°26'10.77" E35°41'35.40" 32.0 47 3.8 11.1 0.7 Lower Campanian N33°26'10.44" E35°41'35.41" 23.8 48 7.3 10.4 1.1 Lower Campanian N33°26'10.09" E35°41'35.48" 13.0 49 0.3 11.5 0.0 Lower Campanian N33°26'09.99" E35°41'35.60" 7.0 50 4.6 10.4 1.0 Lower Campanian N33°26'09.92" E35°41'35.73" 1.4 51 1.8 11.6 0.2 Lower Campanian N33°26'09.81" E35°41'35.83" 1.0 52 4.5 10.6 1.0 Lower Campanian N33°26'09.76" E35°41'35.78" 0* 53 0.3 11.9 0.0 Lower Maastrichtian N33°26'17.80" E35°41'17.84" 54 0.4 10.2 0.0 Lower Maastrichtian N33°26'17.07" E35°41'14.37" 55 0.2 10.7 0.1 Lower Maastrichtian N33°26'13.98" E35°41'12.31" 56 0.3 8.5 0.0 Upper Paleocene N33°26'18.16" E35°41'05.21" 57 0.2 9.3 0.0 Barren possibly Eocene N33°26'21.48" E35°40'57.35" 58 0.2 9.2 0.0 Barren possibly Eocene N33°26'22.26" E35°40'55.95" 59 9.4 9.1 1.3 Lower Maastrichtian N33°26'21.08" E35°41'27.66" 60 7.5 9.4 1.2 Lower Campanian N33°24'13.16" E35°39'44.91" 61 9.0 9.5 1.5 Upper Campanian N33°24'19.36" E35°39'21.34" 62 0.4 10.7 0.4 Upper Campanian N33°24'23.41" E35°39'23.42" 63 5.5 10.4 0.8 Campanian N33°24'14.79" E35°39'16.10" 64 1.9 10.8 0.3 Lower Maastrichtian N33°26'20.27" E35°41'25.19" 65 5.7 9.5 0.9 Lower Campanian N33°24'31.01" E35°40'07.36" 66 7.2 9.1 1.5 Lower Campanian N33°24'31.01" E35°40'07.36" 67 21.7 0.2 2.5 Solid Bitumen*** N33°24'14.79" E35°39'16.10" 68 63.0 0.0 6.9 Solid Bitumen*** N33°24'15.70" E35°39'23.45"

- 70 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Table 3.2 Rock-Eval pyrolysis data for 33 samples from the lower Campanian-lower Maastrichtian in Hasbayya and two solid bitumen samples (67 &68). (HI = S2*100/TOC; OI = S3*100/ TOC). TOC S1 (mg S2 (mg S3 (mg HI (mg OI (mg Tmax PI Sample (wt HC/g HC/g CO /g HC/g CO /g 2 (°C) 2 (S1/(S1+S2)) %) rock) rock) rock) TOC) TOC) 5 2.0 0.09 10.49 1.44 427 530 73 0.01 6 2.2 0.08 10.35 1.74 425 469 79 0.01 7 9.4 0.93 73.81 2.98 406 786 32 0.01 8 1.0 0.04 3.32 1.22 431 323 119 0.01 11 0.5 0.05 1.48 1.28 432 275 238 0.03 14 2.5 0.15 10.95 2.74 423 431 108 0.01 15 0.5 0.05 1.37 1.29 433 257 242 0.04 17 3.3 0.14 18.36 1.70 417 555 51 0.01 18 3.1 0.18 13.27 3.16 434 429 102 0.01 22 3.5 0.19 21.66 1.81 419 612 51 0.01 23 6.7 0.42 46.33 1.53 419 692 23 0.01 25 5.6 0.46 41.21 1.64 412 740 29 0.01 27 6.9 0.47 49.13 1.95 413 715 28 0.01 30 7.5 0.51 45.94 1.97 409 610 26 0.01 31 9.8 4.89 75.13 3.04 405 770 31 0.06 32 11.6 7.67 82.98 2.76 413 715 24 0.08 33 5.1 2.68 37.15 1.88 413 734 37 0.07 34 8.5 7.60 72.25 2.54 414 849 30 0.10 35 8.6 4.34 74.55 2.76 412 872 32 0.06 36 6.2 3.74 45.98 1.59 416 746 26 0.08 40 4.0 0.68 24.46 1.81 414 609 45 0.03 42 7.9 0.52 51.64 1.87 415 654 24 0.01 43 3.8 0.18 22.44 1.88 422 591 50 0.01 44 6.7 0.43 48.41 1.61 418 724 24 0.01 45 6.2 0.33 45.47 1.75 421 728 28 0.01 47 3.8 0.20 29.31 1.47 414 776 39 0.01 48 7.3 0.62 53.18 1.60 413 726 22 0.01 51 1.8 0.07 6.37 1.28 421 356 72 0.01 59 9.4 0.61 57.46 2.91 412 613 31 0.01 60 7.5 0.67 54.50 1.74 411 725 23 0.01 61 9.0 1.11 61.47 3.61 408 685 40 0.02 64 1.9 0.09 8.17 2.20 426 430 116 0.01 65 5.7 0.43 36.83 2.79 412 648 49 0.01 67 21.7 5.80 155.16 2.18 423 715 10 0.04 68 63.0 23.90 463.19 4.92 425 735 8 0.05

- 71 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Table 3.3 Random vitrinite reflectance (VRr%) for nine lower Campanian-lower Maastrichtian samples in Hasbayya.

TOC Sample Age VRr% s* n** (wt%) 64 1.9 Lower Maastrichtian 0.37 0.06 28 59 9.4 Lower Maastrichtian 0.32 0.02 100 9 7.3 Upper Campanian 0.39 0.05 42 18 3.1 Upper Campanian 0.39 0.06 47 34 8.5 Upper Campanian 0.44 0.03 31 42 7.9 Lower Campanian 0.32 0.03 28 44 6.7 Lower Campanian 0.32 0.03 36 52 4.5 Lower Campanian 0.35 0.05 44 60 7.5 Lower Campanian 0.33 0.03 29

Table 3.4 Biomarker ratios determined in this study. (Pr = pristane; Ph = phytane; TAR = Terrigenous Aquatic Ratio; CPI = Carbon Preference Index).

TOC Pr/ Ph/ norhopane (C )/ Sample (wt. Pr/Ph TAR CPI 29 nC nC 25-31 hopane (C ) %) 17 18 30

6 2.2 0.64 0.77 1.49 0.22 1.59 0.42 9 7.3 0.68 0.48 0.92 0.27 2.03 0.36 18 3.1 0.95 1.47 1.98 0.78 2.99 0.39 24 4.2 0.73 0.51 1.48 0.28 7.24 0.56 25 5.6 0.79 2.08 0.61 0.13 0.65 0.68 27 6.9 0.67 0.84 1.92 0.81 3.27 0.20 30 7.5 0.76 0.35 0.69 0.08 2.58 0.30 40 3.6 0.71 0.36 0.61 0.02 0.91 0.44 42 7.9 0.73 0.90 2.22 0.63 2.16 0.40 44 6.7 0.78 1.06 2.40 0.33 1.89 0.47 45 6.3 0.88 0.69 1.48 0.35 2.60 0.62 52 4.5 0.36 0.42 1.43 0.14 1.45 0.40 59 9.4 0.85 0.53 0.86 0.50 2.54 0.33 65 5.7 0.49 0.25 0.98 0.34 3.75 0.80 67 21.7 0.54 68 63.0 0.43

- 72 -

Chapter III Upper Cretaceous source rocks of south Lebanon

Table 3.5 List of identified hopanes and steranes in this study along with their m/z.

Number Component (m/z) Hopanes 1 18α(H)-22,29,30-Trisnorneohopane (Ts) 370 2 17α(H)-22,29,30 Trisnorhopane (Tm) 370 3 17α(H),21β(H)-30-Norhopane 398 4 18α(H)-Norneohopane 396 5 Hop-(17)21-ene 410 6 17β(H),21α(H)-30-Norhopane 398 7 17α(H),21β(H)-Hopane 412 8 Neohop-13(18)ene 410 9 17β(H),21β(H)-30-Norhopane 398 10 17α(H),21β(H)-Homohopane 426 11 17β(H),21β(H)-Hopane 412 12 17β(H),21β(H)-Homohopane 426 13 (22S)-17α(H),21β(H)-Dihomohopane 440 14 (22R)-17α(H),21β(H)-Dihomohopane 440

Steranes A (20R)-5β(H),14α(H),17α(H)-Cholestane 372 B (20R)-5α(H),14β(H),17β(H)-Cholestane 372 C (20S)-5α(H),14β(H),17β(H)-Cholestane 372 D (20R)-5α(H),14α(H),17α(H)-Cholestane 372 E (20R)-24-Methyl-5β(H),14α(H),17α(H)-Cholestane 386 F (20R)-24-Methyl-5α(H),14β(H),17β(H)-Cholestane 386 G (20S)-24-Methyl-5α(H),14β(H),17β(H)-Cholestane 386 H (20R)-24-Methyl-5α(H),14α(H),17α(H)-Cholestane 386 I (20R)-24-Ethyl-5β(H),14α(H),17α(H)-Cholestane 400 J (20R)-24-Ethyl-5α(H),14α(H),17α(H)-Cholestane 400

Table 3.6 Summary table of key source rock parameters of the Campanian-Maastrichtian source rock in Lebanon, Israel, and Jordan (after Bein et al., 1990; Abed et al., 2005; Schneider-Mor et al., 2012; Bou Daher et al., 2014; Alsenz et al. 2015) (*unrealistically high value).

Lebanon Israel Jordan Chekka Hasbayya Pama quarry Shefela Basin Zin Basin TOC (wt. %) 0.02-5.7 0.2-11.6 1-18 0-19.4 0.7-25.5 0.23-22.25 0.39- TS (wt. %) 0-1.5 0-1.7 0.1-2.9 0.36-4.82 0.31-7.79 4.59 Pr/Ph 0.6-1.3 0.4-1.0 0.1-0.7 0.4-0.8 0.2-1.3 0.4-1.2 693- HI (mgHC/gTOC) 413-578 257-872 - 981-1316* - 1175

OI (mgHC/gCO2) 46-98 8-242 - 7-25 6-71 -

- 73 -

Chapter IV 3D thermal history and maturity modelling

Chapter IV

3D THERMAL HISTORY AND MATURITY MODELLING OF THE LEVANT BASIN AND ITS EASTERN MARGIN, OFFSHORE- ONSHORE LEBANON

Keywords: Thermal maturity, burial history, hydrocarbon generation kinetics, petroleum systems.

4.1 Abstract

In the last decade the East Mediterranean Levant Basin has become a frontier hydrocarbon province. Several gas discoveries have been recorded in Miocene reservoirs offshore Israel and seismic data suggest promising prospective plays in deeper intervals throughout the basin.

Source rock quality, quantity, and distribution as well as thermal history are hitherto not well constrained, especially in the deep offshore. In this study we present new source rock information from Jurassic to Paleocene intervals that prove the presence of several gas prone and oil prone potential source rocks along the eastern margin of the Levant basin onshore

Lebanon. None of the analysed onshore source rocks have reached sufficient thermal maturity for oil generation. However, 3D thermal history and maturity modelling indicate that these source rocks have reached thermal maturity in the offshore basin. Several potential petroleum systems have been suggested, including an Upper Cretaceous-Oligo-Miocene biogenic and thermogenic system in the deep basin, a Jurassic-Cretaceous system along the margin, and a

Permian-Triassic system in the onshore. Sensitivity analysis in the poorly calibrated offshore basin showed a large uncertainty with respect to the depth of the oil and gas window and suggested an important effect of the depth of the lithospheric-asthenospheric boundary on the thermal history of the basin. The different scenarios tested in this study showed that the

- 74 -

Chapter IV 3D thermal history and maturity modelling thickness of the biogenic zone below the Messinian salt would vary between 700 and 1500 m in the deep basin offshore Lebanon.

4.2 Introduction

The growing global demand for oil, natural gas, and other sources of energy has moved exploration to challenging and complex provinces. The East Mediterranean region is one of those frontier provinces. Recent gas discoveries in Oligocene and Miocene sandstone reservoirs offshore Israel and Cyprus (e.g.Zohr, Tamar, Dalit, Leviathan, Karish, Tanin, Dolphin, and

Cyprus-A) (www.nobleenergyinc.com; Esestime et al., 2016) have proven the hydrocarbon potential of the Levant Basin. The source of this gas is still unclear, although it has been reported to be of biogenic origin comprising 99% methane (Needham et al., 2013). Oil discoveries in

Mesozoic structures off the coast of Israel were also recorded in drilling campaigns during the

1990s (Gardosh, 2013). Additionally, new seismic data acquired in the last decade revealed the large thickness of the sedimentary column and the presence of direct hydrocarbon indicators

(DHI) thus suggesting the presence of working thermogenic petroleum systems rendering the under-explored Levant Basin one of the most promising hydrocarbon provinces in the region.

Onshore, the Palmyra Basin in Syria produces oil and gas from Early Cretaceous and Middle

Triassic reservoirs, respectively (Nader, 2014). Gas production is also ongoing from

Carboniferous sandstone (Markada Formation) in several fields onshore Syria (Lucic et al.,

2013). In Lebanon, seven exploration wells have been drilled between 1947 and 1966 penetrating only down to the Upper Jurassic which is the oldest surface-exposed rock formation.

Drilling failed to encounter commercial oil and gas volumes in Lebanon, probably due to (i) meteoric water washing affecting the Jurassic and post-Jurassic rock succession and (ii) off structure position of drilled wells (Nader, 2014). The pre-Jurassic succession might, however, include a working petroleum system similar to the Triassic petroleum system in Syria and sealed

- 75 -

Chapter IV 3D thermal history and maturity modelling by the Kurrachine evaporites presumed to be present in Lebanon (Beydoun and Habib, 1995;

Brew et al., 2001; Nader, 2014). These sediments are thought to be the extension of the Upper

Triassic evaporites known in the Palmyra basin (Renouard, 1955; Beydoun and Habib, 1995;

Brew et al., 2001; Nader, 2014).

The East Mediterranean offshore, onshore, and intermediate margin seem to include several promising petroleum systems. The thermal history of the region, as well as the distribution, quality, and maturity of major source rocks are yet unknown. Thus we constructed a 3D thermal history model (TemisFlow™, v. 2013.2) covering an area of 315 x 315 km (grid resolution 5x5 km) including the Levant Basin, margin, and onshore. New source rocks kinetic data were produced and used in the model in order to assess the maturity and the timing of hydrocarbon generation of major source rocks, and to discuss the implications on potential thermogenic petroleum systems in the study area.

4.3 Geological setting

An extensional rift phase along the northern margin of Gondwana started in the late Paleozoic and resulted in the formation of the Levant and the Palmyra basins (Fig. 4.1) (Garfunkel, 1989;

Hawie et al., 2013a; Montadert et al., 2014; Nader, 2014). Successive rift pulses affected the

Levant Basin until the Late Jurassic (Gardosh et al., 2010; Hawie et al., 2013a). A calm post- rift cooling and subsidence phase prevailed until the Late Cretaceous (Hawie et al., 2013a). In the Palmyra basin, subsidence started in the Late Permian and continued till the Late Cretaceous

(Ponikarov, 1966; Chaimov et al., 1992). During the Late Cretaceous, the closure of the

Neotethys Ocean and the collision of the Afro-Arabian and Eurasian plates started. The compressional regime persisted until the Miocene leading to the emplacement of the Syrian Arc fold belt and the inversion of Mesozoic extensional structures in the Palmyra basin (Ponikarov,

1966; Chaimov et al., 1992; Robertson, 1998; Walley, 1998). The two major phases of Syrian

- 76 -

Chapter IV 3D thermal history and maturity modelling

Arc folding occurred in the Coniacian (Late Cretacous) and Late Eocene (Garfunkel, 1998;

Walley, 1998). The initiation of the Red Sea rifting during the Oligocene/Miocene resulted in the propagation of a series of strike-slip faults northward forming the Levant Fracture System which extends from the Gulf of Aqaba to the Taurus Mountains (Beydoun, 1999). The Levant

Fracture System is a sinistral fault system that includes the N-S striking Dead Sea segment, the

NNE-SSW central segment (Yammouneh Fault and fault splays) forming a restraining bend in

Lebanon, and the N-S Ghab Fault segment (Ghalayini et al., 2014).

Fig. 4.1. Regional map of the east Mediterranean showing major structural elements. Green square marks the modelled area. PS1-PS8 are pseudo wells created for calibration, yellow pseudo wells have vitrinite reflectance data, green pseudo wells have temperature data. Orange circle marks the position of 1D extract reported in fig. 12&13 (Modified after Hawie et al., 2013a; Bou Daher et al., 2014; Ghalayini et al., 2014).

- 77 -

Chapter IV 3D thermal history and maturity modelling

4.4 Lithostratigraphic framework

The onshore Jurassic – Quaternary stratigraphy was constrained based on field investigations and biostratigraphic studies (Müller et al., 2010; Hawie et al., 2013b; Nader 2014; Bou Daher et al., 2015). The offshore Jurassic – Quaternary stratigraphy was postulated mostly based on seismic interpretation (Hawie et al., 2013a). The pre-Jurassic stratigraphy was inferred from regional correlations based on published data from neighbouring countries (Brew et al., 2001;

Nader, 2003; Gardosh et al., 2008; Naylor et al., 2013).

The Mesozoic lithostratigraphic succession of the Levant Basin and that of onshore Lebanon realm are relatively similar and mostly dominated by carbonates (Fig. 4.2) (Nader, 2014). In the late Cretaceous, the Afro-Arabian and Eurasian plate convergence and the emplacement of a flexural basin resulted in differential subsidence and deposition of hemipelagic/pelagic and clastic Neogene basinal fill while the marginal realms remained predominated by carbonate platforms (Fig. 4.2) (Hawie et al., 2013a&b). Sedimentation in the basin was almost uninterrupted while in the margin and onshore several erosion events are recorded (Müller et al., 2010; Hawie et al., 2013b). The latest ongoing erosion event is related to the uplift of Mount

Lebanon and Anti-Lebanon, reaching its acme during the middle/late Miocene as a result of a transpressive regime at the Lebanese segment of the Levant Fracture System (Beydoun, 1999;

Gomez et al., 2006; Hawie et al., 2013a). This event led to the exposure of the Jurassic cores of the Lebanese mountains while deposition continued in the adjacent topographic lows; e.g.

Bekaa valley and the coastal areas (Fig. 4.1) (Khair et al., 1997; Hawie et al., 2013a; Nader

2014). The erosional thicknesses were estimated based on regional correlations and on maturity data when available (Fig. 4.2) (Brew et al., 2001; Nader, 2003; Gardosh et al., 2008; Naylor et al., 2013).

- 78 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.2. Stratigraphic chart showing the onshore sedimentary facies and their extrapolation into the offshore basin, the main petroleum system elements, and the major tectonic events (modified after Hawie et al., 2013a).

- 79 -

Chapter IV 3D thermal history and maturity modelling

Several volcanic episodes are recorded along the eastern margin of the Levant basin (Fig. 4.2).

Late Jurassic – Early Cretaceous alkaline volcanism is found in north Lebanon and completely absent in south Lebanon (Dubertret, 1955; Nader, 2014). Wilson (1992) and Garfunkel (1992) attributed this regional intermittent volcanism to mantle plume activity in the Levant region.

Another episode of volcanism occurred in the late Cenozoic (Fig. 4.2). These alkaline volcanics are linked to open cracks related to the Levant Fracture System permitting local deep decompression and magma ascent (Adiyaman and Chorowicz, 2002). Abdel-Rahman and

Nassar (2004) attributed the Pliocene alkali basalts of northern Lebanon to a transtensional regime at the junction of the Lebanese restraining bend and the Ghab segment of the Levant

Fracture System.

4.5 Materials and methods

4.5.1 Source rocks

26 samples were collected from potential source rock intervals within the Kimmeridgian,

Neocomian (Lower Cretaceous), Albian, Cenomanian, and Upper Paleocene in different outcrops and shallow boreholes onshore Lebanon. TOC, TIC and Rock-Eval pyrolysis were performed on all samples using a LiquiTOC II and a Rock-Eval VI device, respectively.

Additionally, Campanian – lower Maastrichtian Rock-Eval data is taken from Bou Daher et al.

(2014, 2015). Total sulphur (TS) content was measured using a Leco S200 analyzer (detection limit 20 ppm; error < 5%).

Vitrinite reflectance measurements were conducted on 9 Neocomian coaly shales and two

Cenomanian calcareous shales. Sample cuttings were prepared according to guidelines described by Taylor et al. (1998). Random vitrinite reflectance was measured at a magnification of 500x in a dark room with a Zeiss Axio Imager microscope for incident light equipped with a tungsten-halogen lamp (12V, 100W), a 50X/0.85 Epiplan-NEOFLUAR oil immersion objective - 80 -

Chapter IV 3D thermal history and maturity modelling and a 546 nm filter, and using Zeiss immersion oil with refraction index ne =1.518; 23°C. For calibration, a leuco-saphire mineral standard (0.592%) was used. Details of sample preparation and microscopic equipment are described in Littke et al. (2012). Vitrinite reflectance data for the Campanian rocks was taken from Bou Daher et al. (2014, 2015).

Bulk hydrocarbon generation kinetics were determined for 5 samples. Kerogen was isolated from the samples by elimination of minerals using non-oxidizing acids treatment under a continuous nitrogen flow (Durand and Nicaise, 1980; Behar et al., 2008), followed by a soxhlet extraction for one hour using DCM to eliminate the bitumen fraction. A Rock-Eval VI apparatus was used to conduct the experiments. A pyrolysis run was realized on a portion of each sample to ensure that the S1 peak was successfully eliminated. XRD analysis was also performed prior and post acid treatment to insure the elimination of mineral matter. Bulk hydrocarbon generation rates were then measured at five different heating rates (2, 5, 10, 15, 25°C/min) starting at 200°C for 15 min and reaching to 700°C.

The bulk hydrocarbon generation curves were then used to calculate mean activation energy

(Ea) and a pre-exponential factor (A) for every sample using the least squares method according to Schenk et al. (1997). A discrete model with a spacing of 2kcal/mol between consecutive activation energies was adopted, and an IFPEN software, GeoKin, was used to calculate the activation energy distributions using the previously calculated pre-exponential factor (A).

4.5.2 3D data set

Regional isopach maps for the Levant basin and margin representing the Jurassic to Plio-

Quaternary units were provided by Hawie (2014) who constructed the maps based on available literature (Ponikarov, 1966; Ukla, 1970; Dubertret, 1975; Brew et al., 2001; Hardenberg and

Robertson, 2007; Gardosh et al., 2008; Powell and Moh’d, 2011; Zilberman and Calvo, 2013;

Hawie et al., 2013a&b). The thicknesses of the pre-Jurassic units were also inferred from data

- 81 -

Chapter IV 3D thermal history and maturity modelling published in neighbouring countries (Brew et al., 2001; Nader, 2003; Gardosh et al., 2008;

Naylor et al., 2013). However, for simplicity, constant values as shown in Fig. 4.2 were used for the pre-Jurassic units, except where specific data was available.

Lithological information was based on published data (e.g. Dubertret, 1975; Saint-Marc, 1972,

1974; Bou-Dagher Fadel and Clark, 2006; Gardosh et al., 2008; Bowman, 2011; Hawie et al.,

2013a&b) according to the lithostratigraphic framework illustrated in Fig. 4.2.

Bathymetry maps were defined using sedimentological and biostratigraphical investigation in the onshore and seismic facies interpretation in the offshore (Hawie et al., 2013a&b; Hawie,

2014).

4.5.3 Boundary conditions

As an upper thermal boundary, the bottom water temperature (for the offshore) and the surface temperature (for the onshore) and its evolution through the basin history was derived from

Wygrala (1989). The lower boundary condition and its evolution through time were determined using a McKenzie type crustal model with the TemisFlow™ “Advanced basement” tool. This crustal model was constructed based on the assumption that the crust flooring the Levant basin is attenuated continental rather than oceanic (Beydoun, 1977; Khair et al., 1993, 1997; Makris et al., 1983; Netzeband et al., 2006; Segev et al., 2006). Present day crustal thicknesses were set to 8 km under the basin’s central axis and 24 km under the onshore realm based on seismic refraction data presented in Netzeband et al. (2006). The depth of the lithospheric- asthenospheric boundary (LAB) with the thermal value of 1333°C was adjusted to provide a best fit with the available calibration data and was set at 90 km under the stretched continental crust in the basin and 115 km under the onshore. Pre-rift crustal thickness was assumed to be

24 km. Thus a Beta factor map was calculated and three rifting events were introduced in the

- 82 -

Chapter IV 3D thermal history and maturity modelling

Permian, Triassic, and Early Jurassic. The heat flow (HF) at the base of the sediments was then produced as an output of the crustal model.

(a)

(b)

Fig. 4.3. (a) Temperature calibration in wells Delta 01 and Tamar, and pseudo well PS8. (b) Vitrinite reflectance calibration in pseudo wells (see Fig. 4.1 for location). (Color coded stratigraphy based on international chronostratigraphic chart)

- 83 -

Chapter IV 3D thermal history and maturity modelling

4.5.4 Vitrinite reflectance, transformation ratio, expulsion

The calculation of vitrinite reflectance was based on the EASY Ro% kinetic after Sweeney and

Burnham (1990). Transformation ratios were calculated using the measured kinetics for each reported source rock, and using the following formula:

(퐼푛푖푡푖푎푙 퐻퐶 푃표푡푒푛푡푖푎푙 − 퐶푢푟푟푒푛푡 퐻퐶 푃표푡푒푛푡푖푎푙) × 100 푇푟푎푛푓표푟푚푎푡푖표푛 푅푎푡푖표 % = 퐼푛푖푡푖푎푙 퐻퐶 푃표푡푒푛푡푖푎푙

Expulsion was calculated based on the assumption that a hydrocarbon saturation threshold of

20% is needed for effective hydrocarbon expulsion from all source rock facies. The amount of expelled hydrocarbon is also a function of TOC and source rock thickness. The TOC values used for expulsion calculations are averages of the values reported in Table 4.1. Except for the

Campanian – lower Maastrichtian source rocks, where a lower TOC value (1%) was assumed in the offshore basin, while a higher value (5%) was used for the onshore based on the depositional model suggested by Bou Daher et al. (2015). The thickness of the source rocks in the offshore was assumed to be similar to what has been observed in the onshore and reported in the results below.

4.6 Results and discussion

4.6.1 Calibration

Calibration data used in this study consists of borehole temperature data and vitrinite reflectance data (Fig. 4.3 a&b). Temperature data for Delta 01 and Tamar wells was derived from Dubille and Thomas (2012). Due to the scarcity of calibration points in many parts of the basin, especially the offshore, temperature data from calibration wells was used not only at the wells location but also for calibration in pseudo wells in areas of the basin with similar crustal thickness, lithologies, salt thickness, and bathymetry to the location of the wells to which the

- 84 -

Chapter IV 3D thermal history and maturity modelling calibration data belongs (Fig. 4.1). Temperature data from the nearby Nir-Am field (SW-Israel)

(Shalev et al., 2013) was similarly used for calibration in pseudo well PS8 (Fig. 4.1). Vitrinite reflectance data (Fig. 4.3b) was measured on outcrop and near surface samples collected from

Mount Lebanon (Table 4.1). 1D burial analysis showed that those rocks have reached their deepest burial during the middle Miocene while their adjacent equivalents in the Bekaa valley and along the coastal areas have reached their deepest burial at present day due to continuous sedimentation since the Miocene. However, the increase in the calculated VRr for all layers between the middle Miocene to present day is small where the Miocene - Quaternary rock succession is not of a great thickness (< 300m). Thus the vitrinite reflectance data collected from Lower Cretaceous, Cenomanian, and Campanian rocks on and around Mount Lebanon were used for calibration in several pseudo wells (Fig. 4.1) where the Miocene - Quaternary rock succession is not thicker than 300 m (Fig. 4.3b). The use of surface collected vitrinite reflectance data for calibration in pseudo wells is further explained in Noeth et al. (2001; 2002).

4.6.2 Source rocks

The Rock-Eval data (Table 4.1; Fig. 4.4) proved the presence of several oil prone and gas prone source rocks, with varying bulk hydrocarbon generation kinetics (Fig. 4.5), along the eastern margin of the Levant basin. The Paleocene contains around 30 m of calcareous source rocks showing a wide range of HI and OI values, varying from 100 to 508 mg HC/g TOC and from

96 to 249 mg CO2/g TOC, respectively (Table 4.1; Fig. 4.4). The variation observed in HI and

OI values of Paleocene source rocks can be due to changes in the relative input of marine and terrestrial organic matter and/or due to oxidation of marine organic matter resulting in a kerogen type varying between Type II and Type III. The Campanian – lower Maastrichtian contains around 150 m of calcareous source rocks that show a wide range of HI and OI values, varying from 257 to 872 mg HC/g TOC and from 22 to 242 mg CO2/g TOC, respectively (Table 4.1;

Fig. 4.4). The variation observed in the HI and OI values of the Campanian – lower - 85 -

Chapter IV 3D thermal history and maturity modelling

Maastrichtian source rocks is most probably due to differences in bottom water redox conditions at the time of deposition (Bou Daher et al., 2014, 2015). The Campanian – lower

Maastrichtian source rocks are of marine origin and plot in the Type I and Type II field in figure

4.4. Detailed geochemical and petrographic results confirms the presence of Type IIS kerogen especially in the TOC-rich intervals of the Campanian – lower Maastrichtian source rocks (Bou

Daher et al., 2015).

Fig. 4.4. Pseudo Van-Krevelen diagram showing the type and the organic content of the source rocks sampled onshore Lebanon.

The Cenomanian includes also marine calcareous source rocks showing very high HI and relatively low OI values (Table 4.1; Fig. 4.4), which might be the result of deposition under anoxic conditions. The vertical extent of the Cenomanian source rocks in Lebanon is restricted to few beds (3-4 m) only, but this source rock might have a larger extent in the offshore as it might be related to an ocean anoxic event (Kolonic et al., 2002; Lüning et al., 2004; Sachse et al., 2011). The Albian source rock shows a poor quality (Type III/IV kerogen) and a low TOC

- 86 -

Chapter IV 3D thermal history and maturity modelling

(Table 4.1; Fig. 4.4). The Neocomian (Lower Cretaceous) consists of coaly shales of moderate source rock quality (Table 4.1; Fig. 4.4) interbedded within the sandstones of the Chouf formation. The total thickness of the organic rich beds in the Neocomian can reach up to 12 m.

Kimmeridgian source rocks onshore Lebanon consists of thin (< 0.5 m) lenses of shales with very high TOC, moderate HI and low OI within a generally reefal carbonate formation (Table

4.1; Fig. 4.4). However, very prolific Upper Jurassic source rocks are known to exist in the

Middle East (Murris 1980; Ayres et al., 1982). All the source rocks analysed from onshore

Lebanon are immature (Table 4.1).

The activation energy distribution for the Paleocene source rock shows a typical type II bell shaped distribution (Fig. 4.5a). The Campanian source rock shows a major difference in the activation energy distribution between Campanian 1 and Campanian 2 samples (Fig. 4.5b&c) which can be attributed to the presence of Type II (Campanian 1) and Type IIS (Campanian 2) kerogen. A depositional model for the Campanian-lower Maastrichtian source rocks was proposed by Bou Daher et al. (2015) suggesting that Type II kerogen with high activation energies would be expected for the Campanian source rocks in the basin, while along most of the onshore Type IIS kerogen with lower activation energies is more abundant. The Lower

Cretaceous and Upper Jurassic coaly shales show a wide spectrum of activation energies tailing up to 66 kcal/mol which is observed in many type III kerogen bearing source rocks (Fig.

4.5d&e) (Schenk et al., 1997). The presence of these reported source rocks in the offshore part of the study area is an open question. However, molecular and isotopic data from gaseous hydrocarbons encountered in the Mesozoic – Cenozoic section at the continental margin offshore Israel suggest the presence of several genetic systems (Feinstein et al., 2002).

- 87 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.5. (a-e) Bulk hydrocarbon generation kinetics of five source rock samples and (f) their calculated transformation ratio for a heating rate of 1°C/Ma.

4.6.3 Heat Flow

The modelled basal heat flow (HF) showed different trends in the basin, margin, and onshore areas (Fig. 4.6). In the basinal realm, an elevated HF during the Permian - Early Jurassic is the result of rifting pulses that occurred during that period. The HF values in the basin decrease slowly during the post rift cooling and subsidence phase until the late Eocene/early Oligocene when a fast decrease in HF occurs, most probably as a result of fast sedimentation (Fig. 4.6a).

Along the margin, the HF trend is relatively similar to the basinal area, but with lower calculated

- 88 -

Chapter IV 3D thermal history and maturity modelling

HF values for the Permian - Early Jurassic and slightly higher ones during the last 40 Ma (Fig.

4.6b), which is due to a lower β factor and a lower sedimentation rate along the margin, respectively. In the onshore area, our model shows relatively constant values through time, varying slightly above 50 mW/m2 with an increase in the last 10 Ma due to uplift and erosion of Mount Lebanon (Fig. 4.6c). The modelled present day HF in this study shows higher values in the onshore than in the offshore, which is in accordance with the few data available on the present day surface heat flow. Jiménez-Munt et al. (2006) reported HF values ranging from 40 to 60 mW/m2 with an increase from the thin continental crust flooring the basin to the thicker crust in the onshore. Makris and Stobbe (1984) also reported a similar trend from 0.9 HFU (38 mW/m2) in the basin to 1.6 HFU (67 mW/m2) in the onshore. Many other authors reported a similar decrease in the HF values from the onshore to the offshore as well as from the southern margin (southern Israel) northwards (northern Israel) (Eckstein and Simmons 1978; Roded et al., 2013; Shalev et al. 2013; Schütz et al., 2014). This trend is most probably the result of a decrease in the thickness of the radiogenic continental crust and an increase in the less radiogenic sedimentary cover. The resulting average present day geothermal gradient varies from 19°C/km in the offshore basin to around 25°C/km in the onshore area.

4.6.4 Burial history and maturity

The different geodynamic events experienced by the Levant basin and the onshore Levant region have been recorded differently in the sedimentary successions of different parts of the study area, particularly during the Cenozoic. Since the middle Eocene, the northwestern part of the Arabian plate has been emerging out of the water (Bar et al., 2011), resulting in several regional erosion events (Hawie et al., 2013a), while the nearby Levant basin has been rapidly subsiding and accumulating more than 5 km of sediments (Hawie et al., 2013b). This differential vertical movement had an important effect on the thermal history and source rock maturity of the study area. To illustrate this effect four burial history diagrams were extracted - 89 -

Chapter IV 3D thermal history and maturity modelling from four different areas of the model (Fig. 4.7). Up untill the Late Cretaceous, the study area underwent a relatively similar burial history with only slightly higher thermal maturity in the offshore basin due to higher heat flow resulting from rifting pulses at that time.

Fig. 4.6. Basal heat flow trends for the offshore basin, margin, and onshore Lebanon.

The Cenozoic burial history, however, differs considerably in the onshore from the offshore

(Fig. 4.7a-d). A fast burial in the deep offshore basin during the last 60 Ma (Fig. 4.7a) resulted in the thermal maturation of all Mesozoic and early Cenozoic source rocks. A similar fast burial occurred along the margin, but with a shallower depth due to the thinner Neogene cover,

- 90 -

Chapter IV 3D thermal history and maturity modelling resulting in the thermal maturation of most of the Mesozoic with the exception of the Upper

Cretaceous (Fig. 4.7b). In the onshore, deepest burial on Mount Lebanon was reached in the middle Miocene leading to a halt in maturation since that time (Fig. 4.7c). In the Bekaa valley and along the coastal area and topographic lows onshore, sedimentation continued during the uplift of Mount Lebanon throughout the Miocene, Pliocene, and Quaternary resulting in deepest burial at present day (Fig. 4.7d). With the exception of the Jurassic in the Bekaa valley reaching the early oil window, only Triassic and older rocks have reached the petroleum generation stage onshore, while the younger rocks are still immature.

The effect of the thick Cenozoic cover in the Levant basin on thermal maturity of potential source rocks can be clearly seen in maturity maps of calculated vitrinite reflectance (Fig. 4.8).

Permian and Triassic source rocks seem to have entered the oil window everywhere in the study area during the Late Cretaceous (Fig. 4.8a). The Permian and Triassic source rocks experienced deep burial and maturation reaching the wet gas window in most of the offshore basin during the mid to late Eocene and the dry gas window during the Oligo-Miocene. Since the late

Miocene, the Permian and Triassic source rocks entered the overmature stage in most of the offshore Levant basin (Fig. 4.8a). The maturity of Permian and Triassic source rocks decreases gradually along the margin towards the onshore where it has remained within the oil window throughout the Cenozoic, with the exception of some local deeper burial (e.g. in the Bekaa valley) where the Permian source rocks have reached the wet gas window.

- 91 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.7. (a) Extracted burial history and calculated vitrinite reflectance from (a) the deepest offshore, (b) margin, (c) uplifted mount Lebanon, and (d) Bekaa valley.

Upper Jurassic to Lower Cretaceous source rocks entered the oil window in most of the offshore basin during the mid to late Eocene, the wet gas window during the late Miocene and the dry gas stage at present day (Fig. 4.8b). Only in the deepest part of the offshore basin are the Upper

Jurassic and Lower Cretaceous source rocks overmature (Fig. 4.8b). Along the margin, these source rocks vary from early oil window to late oil window, while in the onshore they remain immature or in the very early oil window with slightly higher maturities in the Bekaa valley.

The Upper Cretaceous Cenomanian source rocks entered the oil window during the late Eocene, the wet gas window in the northern offshore basin during the mid to late Miocene, and the dry gas stage at present day (Fig. 4.8c). Along the margin, the Cenomanian source rocks are at - 92 -

Chapter IV 3D thermal history and maturity modelling present day in the early oil window (Fig. 4.8c). The Campanian and Paleocene source rocks have entered the oil window in the offshore basin during the late Oligocene (Fig. 4.8d). In the northern segment of the offshore Levant basin the Campanian and Paleocene source rocks reached the wet gas window during the late Miocene while they remained in the oil window in the rest of the offshore basin. Campanian source rock have reached the dry gas window in the deepest part of the basin offshore Lebanon (Fig. 4.8d). Along parts of the margin both source rocks have just entered the oil window, while in the onshore they are still immature. Figures

8e&f show two E-W and NNE-SSW cross sections, respectively, with calculated vitrinite reflectance that reflect the very different maturation trends along the onshore and offshore.

Fig. 4.8a. Calculated vitrinite reflectance maps for the Lower Triassic source rocks at different time steps.

- 93 -

Chapter IV 3D thermal history and maturity modelling

It should be noted that the maturity stages mentioned in this chapter are standard translations of calculated vitrinite reflectance into petroleum generation stages and do not consider the different kerogen types. Therefore type IIS bearing source rocks in the onshore area might have generated some oil although they are classified here as thermally immature.

Fig. 4.8b. Calculated vitrinite reflectance maps for the Kimmeridgian source rocks at different time steps.

- 94 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.8c. Calculated vitrinite reflectance maps for the Cenomanian source rocks at different time steps.

4.6.5 Transformation ratio and expulsion

The very different source rock maturation trends in the Levant basin, margin, and onshore imply very different potential petroleum systems in each of these realms due to different timing and depth of hydrocarbon generation. In order to assess the petroleum generation timing and potential of a certain source rock maturity information alone is not sufficient, but transformation ratios and expulsion information are also needed.

- 95 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.8d. Calculated vitrinite reflectance maps for the Campanian source rocks at different time steps.

Transformation ratio and expulsion maps (Fig. 4.9 & 4.10) have been extracted for several potential source rocks. The Upper Jurassic Kimmeridgian source rocks started generating and expelling hydrocarbons in most of the offshore basin during the very Late Cretaceous and early

Paleocene, have reached 100% transformation ratio and expelled a maximum of 210 kgHC/m2 by the late Miocene (Fig. 4.9a & 4.10a). Along the margin, hydrocarbon expulsion from the

Kimmeridgian source rock continues to present day (Fig. 4.10a). The Lower Cretaceous

Neocomian source rocks show a similar trend, with a slightly younger onset of hydrocarbon

- 96 -

Chapter IV 3D thermal history and maturity modelling generation and expulsion, however, with larger amounts of expelled hydrocarbons, reaching

2000 kgHC/m2 due to the higher thickness of the Neocomian source rock (Fig. 4.9b & 4.10b).

In the onshore, limited hydrocarbon generation and expulsion have occurred from the

Kimmeridgian source rocks in some of the coastal areas and inland topographic lows since the late Miocene (Fig. 4.9a & 4.10a), while the Neocomian source rocks did not generate any hydrocarbons onshore (Fig. 4.9b & 4.10b). The generated Kimmeridgian hydrocarbons onshore have most probably been washed by meteoric water which have been invading the Jurassic and the overlying rock succession since the emergence of Mount Lebanon in the late Cenozoic

(Nader and Swennen, 2004).

Fig. 4.8e. Cross section AA’ showing calculated vitrinite reflectance at present day (See fig. 1 for location)

- 97 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.8f. Cross section BB’ showing calculated vitrinite reflectance at present day (See fig. 1 for location)

Fig. 4.9a. Transformation ratio maps for the Kimmeridgian source rock at different time steps.

- 98 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.9b. Transformation ratio maps for the Neocomian source rock at different time steps.

Campanian and Paleocene source rocks have very similar hydrocarbon generation and expulsion histories starting in the late Oligocene in the deepest part of the offshore basin, and remaining restricted to the offshore basin while almost no hydrocarbon generation and expulsion have occurred from those source rocks along the margin or in the onshore (Fig. 4.9 c&d & Fig. 4.10 c&d). The Campanian source rock has, however, expelled much more hydrocarbons than the Paleocene source rock due to its higher thickness (Fig. 4.10 c&d). The transformation ratio and expulsion maps of the Campanian source rock were calculated using the kinetics of Type II kerogen (Campanian 1) (Fig. 4.5b). Another simulation was run using the kinetics of Type IIS kerogen (Campanian 2) (Fig. 4.5c) which resulted in much earlier onset of hydrocarbon generation from Campanian source rocks in the offshore Levant basin and

- 99 -

Chapter IV 3D thermal history and maturity modelling considerable generation and expulsion along the margin, while in the onshore only minor hydrocarbon generation occurred in the Bekaa valley. Based on the conceptual depositional model proposed by Bou Daher et al. (2015), suggesting that Type II kerogen with high activation energies would be expected for the Campanian source rocks in the basin and margin, while along most of the onshore Type IIS kerogen with lower activation energies is more representative, we suggest that the scenario reported in Fig. 4.9c & Fig. 4.10c is to our best knowledge the representative case for the Levant basin and margin. The second scenario, however, is validated for the onshore realm through the presence of large amounts of immature solid bitumen south of the Bekaa valley which is typical for early stages of hydrocarbon generation from Type IIS kerogen (Bou Daher et al., 2015).

Fig. 4.9c. Transformation ratio maps for the Campanian source rock at different time steps.

- 100 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.9d. Transformation ratio maps for the Paleocene source rock at different time steps.

4.6.6 Petroleum systems

The results presented and discussed above illustrate the quality and quantity of known source rocks and the timing of hydrocarbon generation from key source rock intervals in different realms of the study area and indicate that different potential source rocks can be active at different times in the offshore, margin, and onshore. However, for a successful petroleum system to be present several other factors need to come in place. These include migration paths, reservoir quality, trapping, sealing, and seal integrity. In the Permian-Triassic succession, rock units with source, reservoir, and seal potential have been reported in Israel (Gardosh et al.,

2008) and in Syria (Brew et al., 2001; Lucic et al., 2010). Triassic source rocks have been - 101 -

Chapter IV 3D thermal history and maturity modelling proved to charge Middle Triassic, Middle Jurassic, and Upper Cretaceous reservoirs in Syria

(Abboud et al., 2005). In the deepest part of the offshore Levant basin, hydrocarbon generation from Permian-Triassic source rocks occurs between 90-34 Ma and petroleum can accumulate in tilted rift blocks (Fig. 4.11). Along the margin hydrocarbon generation from the Permian-

Triassic starts at later stages (75Ma) and continues in the onshore till present day. Onshore structures, such as the Qartaba box fold (Nader, 2014) can be an excellent trap for Permian-

Triassic hydrocarbons sealed by the Upper Triassic anhydrites that are expected to be present in Lebanon, as an extension of the Kurrachine anhydrites of the Palmyra basin. Such evaporite deposits could protect the Permian-Triassic succession in Lebanon from water washing and karstification that seems to have affected the Jurassic and younger rock units (Renouard, 1955;

Beydoun, 1977; Beydoun and Habib, 1995; Brew et al., 2001; Lucic et al., 2010). Hydrocarbons from Jurassic and Lower Cretaceous source rocks were generated in the deepest basin between

75-24 Ma and might have accumulated in reactivated Mesozoic structures along the margin

(Fig. 4.11) (Ghalayini et al., 2014). Jurassic and Lower Cretaceous rocks exposed onshore have excellent reservoir potential (Nader, 2014) and can be sealed by the Upper Jurassic

(Kimmeridgian) shales, marls, and volcanics and the Lower Cretaceous (Albian) marls, respectively (Nader 2014). Oil has been reported in Lower Cretaceous sandstones in the Mango field offshore Sinai and in the Euphrates Graben (OAPEC Annual Report, 2002; Gardosh et al.,

2008; Nader, 2014). Oil and gas shows have been also reported from Middle Jurassic limestones

Yam-2 and Yam Yafi-1 wells off the coast of Israel (Gardosh et al., 2008). One very promising structure that can trap Jurassic and Lower Cretaceous hydrocarbons is the box fold Ile du

Palmier off the coast of Tripoli (Fig. 4.1). The Jurassic and Lower Cretaceous reservoirs in this structure would be protected from water washing by the distance from the uplifted inland intake area (Beydoun, 1977; Nader and Swennen, 2004).

- 102 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.10a. Expulsion maps for the Kimmeridgian source rock at different time steps.

Upper Cretaceous source rocks are only mature in the deep basin, and in the deepest part of the basin they produced hydrocarbons between 34 and 16 Ma. Upward vertical migration from

Upper Cretaceous (particularly Campanian) source rocks in the offshore basin would be impeded by the presence of a very good Paleocene seal. Thus migration would most probably occur laterally into marginal pinchouts and reactivated Mesozoic structural traps (Fig. 4.11).

Lower Maastrichtian calcareous turbidites were reported along the coast in Northern Lebanon

(Hawie et al., 2013b). These can be potential reservoirs for Upper Cretaceous hydrocarbons, sealed by the Paleocene marls. Paleocene and Eocene source rocks produce hydrocarbons at slightly younger ages and might be charging Oligocene and Miocene reservoirs in offshore - 103 -

Chapter IV 3D thermal history and maturity modelling structures. Oligocene potential source rocks are in the oil window at present day and have started to generate hydrocarbons at around 6Ma and thus can charge Oligocene and Miocene reservoirs in young structures (Fig. 4.11), such as the NNE-SSW trending anticlines affecting the Oligocene-Miocene succession in the offshore basin (Ghalayini et al., 2014). Disseminated terrestrial organic matter in the upper Oligocene has been proved to be an important source rock offshore the Nile Delta (Villinski, 2013).

Fig. 4.10b. Expulsion maps for the Neocomian source rock at different time steps.

- 104 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.10c. Expulsion maps for the Campanian source rock at different time steps.

4.6.7 Sensitivity and uncertainty

The thermal history of a sedimentary basin is controlled by many parameters, most of which are often very poorly constrained, particularly in frontier basins such as the Levant basin.

Testing the sensitivity of the system to some of those parameters can help in quantifying the effects they have on the temperature of the basin and thus on source rocks maturation. In this study we have extracted a 1D model from the deepest part of the offshore basin and tested for the sensitivity of the present day geothermal gradient to lithologies variation, crustal thickness, and depth of the lithospheric-asthenospheric boundary (LAB). The pre-Messinian Cenozoic - 105 -

Chapter IV 3D thermal history and maturity modelling lithology was varied from carbonate dominated (low radiogenic heat production, high thermal conductivity) to siliciclastic dominated (high radiogenic heat production, low thermal conductivity). The resulting difference in the present day geothermal gradient between the tested scenarios affects the pre Miocene succession shifting the temperatures by up to 18°C

(Fig. 4.12a). This temperature variation seems to have no effect on the Miocene succession, most probably due to the chimney effect exerted by the high thermal conductivity of the thick

Messinian salt (Fig. 4.12a).

Fig. 4.10d. Expulsion maps for the Paleocene source rock at different time steps.

- 106 -

Chapter IV 3D thermal history and maturity modelling

The present day crustal thickness was varied from 8 km to 20 km (Netzeband et al., 2006). The resulting effect on the present day geothermal gradient increases with depth, reaching a difference of 28°C between end member scenarios for the deepest intervals (Fig. 4.12b). The effect of crustal thickness is minor for shallow intervals (Fig. 4.12b).

Fig. 4.11. Schematic diagram summarizing the potential petroleum systems of the Levant basin offshore, margin, and onshore Lebanon.

The depth of the LAB was varied from 80 km to 120 km (Jiménez-Munt et al., 2006; Segev et al., 2006) which had a major effect on the present day geothermal gradient (Fig. 4.12c). The effect of these three modified parameters on source rock maturation and petroleum systems is summarized in Figure 4.13, showing a significant impact on the expected depth of the oil and the gas windows. The thickness of the pre-Messinian biogenic zone also varies between 700 m

- 107 -

Chapter IV 3D thermal history and maturity modelling and 1500 m indicating that under any scenario the pre-salt Miocene in the deep offshore basin would have a biogenic gas potential in a zone of at least 700 m, assuming a biogenic zone from zero to 80°C (Fig. 4.13).

Fig. 4.12a. Extracted 1D model from the deepest offshore basin showing the sensitivity of the system to changes in lithology (See figure 1 for location) (Color coded stratigraphy based on international chronostratigraphic chart) (Note that the first 2300m depth are water column). - 108 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.12b. Extracted 1D model from the deepest offshore basin showing the sensitivity of the system to changes in crustal thickness (See figure 1 for location) (Color coded stratigraphy based on international chronostratigraphic chart) (Note that the first 2300m depth are water column).

- 109 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.12c. Extracted 1D model from the deepest offshore basin showing the sensitivity of the system to changes in depth of the LAB (See figure 1 for location) (Color coded stratigraphy based on international chronostratigraphic chart) (Note that the first 2300m depth are water column).

In an underexplored basin such as the Levant basin, uncertainties related to hydrocarbon generation timing can raise not only from uncertainty in thermal history, but also from several - 110 -

Chapter IV 3D thermal history and maturity modelling other parameters. TOC values, source rock presence, quality and thickness, hydrocarbon generation kinetics, and lateral changes in organofacies are parameters that have an important impact on hydrocarbon generation and eventually on petroleum systems. Further uncertainties that should be considered when assessing potential petroleum systems in the Levant Basin, particularly offshore, are related to reservoir quality, migration paths, and seal integrity. A catastrophic event that had an enormous impact on seal integrity and petroleum systems is the

Messinian salinity crisis. The fast drop of sea level by around 2000 m resulted in a drop in pressure, an increase in temperature, and fast salt deposition. The surface temperature change during the Messinian is minor and has little effect on the petroleum systems. The hydrostatic pressure drop, however, would result in higher pore pressure and possible seal fracturing within the shallow biogenic system (Wygrala et al., 2014).

4.7 Conclusions

The results presented in this study proved the presence of several potential petroleum source rocks along the eastern margin of the Levant basin and suggest different prospective working petroleum systems in the offshore, margin, and onshore Lebanon.

An Upper Cretaceous-Oligo-Miocene mixed thermogenic and biogenic system is expected in the offshore. In the pre-Messinian salt succession, the thickness of the biogenic zone can vary between 700 and 1500 m throughout the offshore basin, rendering the NNE-SSW offshore anticlines which affect the Oligo-Miocene succession an excellent target for biogenic and thermogenic hydrocarbons. The biogenic hydrocarbon potential decreases towards the margin and onshore due to the lack of appropriate seal.

- 111 -

Chapter IV 3D thermal history and maturity modelling

Fig. 4.13. Extracted 1D model from the deepest offshore basin showing minimum and maximum geothermal gradients and their effect on maturity and petroleum systems. (See figure 1 for location) (Color coded stratigraphy based on international chronostratigraphic chart) (Note that the first 2300m depth are water column).

A Jurassic-Lower Cretaceous petroleum system is expected along the margin, where Lower

Cretaceous sandstones and Upper Jurassic carbonates, which have excellent reservoir quality in the onshore, could be charged with oil and gas from Jurassic and Lower Cretaceous shales.

- 112 -

Chapter IV 3D thermal history and maturity modelling

A Permo-Triassic petroleum system is expected in the onshore, where Lower Triassic reservoirs sealed by Upper Triassic evaporites in Miocene structures could be filled with gas from Permian

(and deeper) source rocks.

Kinetic results derived from Campanian source rocks showed a variation of bulk hydrocarbon generation kinetics attributed to lateral variation of organofacies within the same source rock.

These lateral source rock quality changes have an important impact on the assessment of hydrocarbon generation particularly in a frontier basin. Thus a detailed understanding of a source rock’s depositional environment at the sampling sites is essential for a better source rock extrapolation at a basin scale.

- 113 -

Chapter IV 3D thermal history and maturity modelling

Table 1. Summary of source rocks data gathered from onshore Lebanon. (*Sample selected for kinetic experiments; **Data from Bou Daher et al., 2014&2015 in chapters II &III)

S1 (mg/g S2 (mg/g S3 (mg/g HI (mgHC/g OI (mgCO /g Sample # Age TOC (wt. %) TIC (wt. %) TS (wt. %) 2 PI (S1/(S1+S2)) Tmax (°C) VRr (%) GPS rock) rock) rock) TOC) TOC) 1 Upper Paleocene 0.9 8.6 0.7 0.09 4.00 1.32 444 146 0.02 424 - N 34°19' 47.65" E 35°45’49.19" 2 Upper Paleocene 0.8 7.1 0.5 0.07 2.95 1.22 361 150 0.02 433 - N 34°19' 47.65" E 35°45’49.19" 3 Upper Paleocene 1.3 9.1 0.7 0.06 6.38 1.21 508 96 0.01 430 - N 34°19' 47.65" E 35°45’49.19" 4 Upper Paleocene 1.4 7.6 0.8 0.05 6.17 1.31 450 96 0.01 428 - N 34°19' 47.65" E 35°45’49.19" 5* Upper Paleocene 1.3 7.1 0.7 0.05 3.24 1.88 245 142 0.02 426 - N 34°19' 47.65" E 35°45’49.19" 6 Upper Paleocene 1.0 7.3 1.0 0.05 3.09 1.31 319 135 0.01 431 - N 34°19' 47.65" E 35°45’49.19" 7 Upper Paleocene 1.0 7.6 1.0 0.04 3.44 1.21 355 125 0.01 431 - N 34°19' 47.65" E 35°45’49.19" 8 Upper Paleocene 0.5 9.1 0.5 0.03 1.21 1.25 235 244 0.02 431 - N 34°19' 47.65" E 35°45’49.19" 9 Upper Paleocene 0.5 8.7 0.9 0.03 0.75 1.18 154 242 0.04 424 - N 34°19' 47.65" E 35°45’49.19" 10 Upper Paleocene 0.5 7.9 1.3 0.04 0.49 1.22 100 249 0.07 416 - N 34°19' 47.65" E 35°45’49.19" 0.41 (n=16; 11 Cenomanian 1.2 12.8 - 0.14 10.68 1.16 890 97 0.01 405 N 34°10' 04.86" E 35°45’25.10" s=0.026) 0.42 (n=7; 12 Cenomanian 0.9 12.2 - 0.17 7.56 1.19 840 132 0.02 405 N 34°10' 04.86" E 35°45’25.10" s=0.032) 13 Albian 0.4 10.0 - 0.01 0.17 1.34 42 334 0.06 423 - N 33°32' 05.84" E 35°34’51.23" 14 Albian 0.5 5.8 - 0.04 0.42 1.27 83 254 0.09 416 - N 33°32' 05.84" E 35°34’51.23" 0.54 (n=100; 15 Neocomian 37.6 0.0 4.7 1.04 16.31 18.42 43 49 0.06 415 N 33°37'59.8" E 35°55’21.2" s=0.059) 0.444 (n=100; 16 Neocomian 10.8 1.5 1.0 0.15 10.63 4.25 99 39 0.01 418 N 33°33'00.7" E 35°33’02.2" s=0.039) 0.513 (n=100; 17 Neocomian 31.3 1.4 1.1 0.30 17.78 11.48 57 37 0.02 418 N 33°33'00.7" E 35°33’02.2" s=0.055) 0.655 (n=100; 18 Neocomian 2.4 0.0 1.0 0.15 0.45 1.89 19 80 0.25 415 N 33°33'00.7" E 35°33’02.2" s=0.051) 0.47 (n=100; 19 Neocomian 24.2 1.5 1.1 1.09 47.67 9.06 197 37 0.02 416 N 33°33'00.7" E 35°33’02.2" s=0.057) 0.481 (n=100; 20 Neocomian 8.3 1.3 0.3 0.19 5.10 4.60 61 55 0.04 423 N 33°33'00.7" E 35°33’02.2" s=0.045) 21 Neocomian 0.3 0.2 0.1 0.03 0.27 1.12 81 335 0.10 429 - N 33°36'11.2" E 35°38’24.2" 22 Neocomian 3.9 0.0 0.3 0.16 2.55 1.48 66 38 0.06 425 - N 33°36'11.2" E 35°38’24.2" 0.608 (n=100; 23 Neocomian 14.0 0.0 1.8 0.34 5.16 13.33 37 95 0.06 429 N 33°36'11.2" E 35°38’24.2" s=0.042) 0.603 (n=100; 24 Neocomian 62.0 0.0 5.2 1.48 79.56 31.83 128 51 0.02 415 N 34°04' 06.2" E 35°51’10.3" s=0.032) 0.514 (n=100; 25* Neocomian 2.3 0.2 0.2 0.06 5 1.55 215 66 0.01 433 N 34°04' 06.2" E 35°51’10.3" s=0.078) 26* Kimmeridgian 53.3 3.5 - 2.61 158.41 16.29 297 31 0.02 400 - N 33°54' 28.5" E 35°36’23.6" Campanian - lower 0.5 - 11.6 3.65 - 12.76 0.08 - 1.7 0.04 - 7.67 1.37 - 82.98 0.85 - 4.92 257 - 872 405 - 434 0.31 - 0.44 22 - 242 (63) 0.01 - 0.1 (0.02) Maastrichtian** (3.8) (10.0) (0.79) (0.75) (25.8) (1.78) (559) (422) (0.35) 0.33 (n=33; Campanian 1* 1.9 10.8 0.6 0.11 10.36 1.32 542 69 0.01 427 N 34°19’46.90" E 35°45’21.83" s=0.072) 0.33 (n=29; Campanian 2* 7.5 9.4 1.2 0.67 54.50 1.74 725 23 0.01 411 N 33° 24'13.16" E 35° 39'44.91" s=0.026)

- 114 -

Chapter V Conclusion

Chapter V

5.1 Summary and conclusions

The conclusions of this thesis can be divided into three major parts. The first part is related to the detailed characterization of the Upper Cretaceous source rocks and the learning outcome that can be applied not only in the Levant basin, but also in any basin with similar source rocks.

The second part is related to the potential petroleum source rocks of the east Mediterranean.

And the third is related to thermal history, maturity, and potential petroleum systems of the east

Mediterranean and the gained knowledge from the conducted analytical and numerical modelling work.

5.1.1 Upper Cretaceous (Campanian – Maastrichtian) source rock

The Upper Cretaceous (Campanian – Maastrichtian) source rocks analysed in this study show lateral and vertical variations in their quantity and quality between low TOC and HI to high

TOC and HI. This variation was also accompanied by a variation in their hydrocarbon generation kinetics from Type II to Type IIS. The data produced in this thesis along with the data from published literature allowed the construction of a conceptual depositional model (Fig.

3.20) that explains the observed variation and allows a better prediction of the distribution, quality, and quantity of the Upper Cretaceous source rocks on basin scale away from the actual sampling locations. The constructed depositional model suggests deposition of the Upper

Cretaceous source rocks under a productivity belt that extended along the eastern margin of the

Levant basin. Due to the increased primary productivity and associated microbial respiration, an oxygen minimum zone (OMZ) was established. The OMZ impinged the inner shelf in the area of present day central and southern Israel and north-western Jordan resulting in deposition of extremely organic matter rich Campanian – Maastrichtian carbonates under an anoxic

- 115 -

Chapter V Conclusion depositional environment. In deeper outer shelf areas (e.g. present day Lebanon) (Fig. 3.2 &

3.20) the OMZ only reached the sediment-water interface at times of peak primary productivity while most of the time the outer shelf remained in a dysoxic zone resulting in deposition of rocks with variable organic matter content and preservation. This further suggests a decrease in the organic matter content towards distal basinal realms.

Fig. 5.1. Transformation ratio maps for the Campanian source rock at present day using

(a) Type II kinetics and (b) Type IIS kinetics (See Chapter IV). Transformation ratio through time extracted from two synthetic wells in the deep offshore and in the margin from (c) scenario Type II and (d) Type IIS.

This decrease in TOC is accompanied by a shift in hydrocarbon generation kinetics into higher activation energies. The switch to higher activation energies with decreasing TOC is explained by two processes. The first is degradation of sinking organic matter in oxygenated water leading to oxidation of labile compounds and thus lower HI values and relatively more refractory

- 116 -

Chapter V Conclusion organic matter in distal area with weaker primary productivity. The second and main reason for the variation in hydrocarbon generation kinetics is due to the shift between Type II and Type

IIS kerogen as a function of TOC (Fig. 3.19 & 4.5b & c). Figure 3.19 shows a relative increase in organic sulphur with increasing TOC. This trend is often observed in iron poor organic matter rich marine carbonates like the Campanian – Maastrichtian rock succession investigated here.

In this kind of depositional environments, the reduced sulphate, during the organic matter- fuelled bacterial sulphate reduction, reacts first with the available reactive iron to form pyrite, and once the available reactive iron is depleted, the excess reduced sulphate reacts with the organic matter. Thus with increasing organic matter the excess reduced sulphate increases and eventually more sulphur gets incorporated in the organic matter gradually shifting the hydrocarbon generation kinetics into lower activation energies (Fig. 4.5 b & c) due to the weaker C-S bonds compared to the C-C bonds.

This lateral shift in organofacies has an important impact on the assessment of hydrocarbon generation throughout the basin. Simulations of hydrocarbon generation showed a difference of approximately 15 Ma between the “critical moment” (50% transformation ratio) at the basin’s geological heating rate (Fig. 5.1). Figure 4.5f shows a difference of around 30°C between the times of peak hydrocarbon generation between the two end member kinetics of the

Campanian source rock. This is approximately equivalent to 1.5 km difference in burial depth.

The vertical variation observed in the Campanian – Maastrichtian source rocks has an important implication that should be considered when assessing hydrocarbon generation in basin modelling studies. Figure 5.2 shows a decrease in Rock-Eval Tmax with increasing TOC which in turn positively correlates with organic sulphur (Fig. 3.19). This shift in laboratory derived

Tmax reflects a shift in geological Tmax. Additionally a shift in the onset and ending of hydrocarbon generation would be expected, as seen in figure 4.5f. This implies that for a source

- 117 -

Chapter V Conclusion rock similar to the one analysed here, the oil and gas windows would be much wider than usually assumed in basin modelling studies.

Fig. 5.2. Plot of Tmax vs. TOC of all the analysed Santonian – Maastrichtian samples showing a decrease in Tmax with increasing TOC as a result of the gradual shift from Type

II to Type IIS at higher TOC values.

5.1.2 Source rocks of the east Mediterranean

Several potential oil prone and gas prone source rocks have been identified within the outcropping rock succession along the eastern margin of the Levant basin onshore Lebanon.

These organic matter rich rocks include Upper Jurassic (Kimmeridgian), Lower Cretaceous

(Neocomian), Albian, Cenomanian, Campanian – lower Maastrichtian, and upper Paleocene

- 118 -

Chapter V Conclusion

(Fig. 4.4). All of the analysed source rocks onshore Lebanon are thermally immature. However, these source rocks, if present in the offshore Levant basin and its distal margin can be part of several potential petroleum systems. Among these source rocks, the Campanian – lower

Maastrichtian is the most prolific source rock in the onshore area, however, as discussed above, its quality and quantity is expected to decrease towards the offshore realm. Ongoing work on the source rocks of the Eratosthenes Continental Block (ECB) at RWTH Aachen show very minor presence of Campanian-Maastrichtian source rocks, thus supporting the suggested model.

Other potential source rocks are postulated based on data from nearby countries. Among those source rocks are the Middle Jurassic organic rich limestones penetrated in coastal wells in

Israel. Lower Jurassic and Triassic syn-rift organic matter rich deposits were encountered in wells onshore Israel, and are expected to have a better quality in the deep offshore. Younger source rocks have also been suggested, including Eocene, Oligocene, Miocene, and Pliocene.

Eocene source rocks on Eratosthenes sea mount show fair to good source rock quality.

5.1.3 Petroleum systems of the Levant Basin

The study area includes three realms that reacted differently to the various geodynamic events, particularly during the Cenozoic. This differential geological evolution of the offshore Levant basin, the onshore, and the intermediate eastern margin lead to very different potential petroleum systems.

In the whole Levant area, most of the trap forming events are of late Cenozoic age, particularly

Eocene, early Miocene, and late Miocene. The maturity of potential source rocks is, however, very different. Thus structures of the same age can be charged by different source rocks in the offshore, margin, and onshore.

- 119 -

Chapter V Conclusion

In the offshore Lebanon, Permian and Triassic source rocks have been in the gas window since the Eocene. Upper Jurassic and Lower Cretaceous have been in the oil window since the Eocene and in the gas window since the middle Miocene. Upper Cretaceous and Paleocene source rocks have been in the oil window since the late Oligocene – early Miocene. The Oligocene is at present day in the early oil window only in the deepest parts of the basin while in the rest it remains in the biogenic zone (< 80°C) as is the case for the Miocene and younger succession.

Due to the excellent sealing potential that is present in the Upper Cretaceous – Eocene rock succession, the offshore Levant basin might be divided into two different pre-salt petroleum systems: an Oligo-Miocene biogenic system with some early thermogenic potential, and a deeper isolated thermogenic petroleum system.

Along the margin, the source rock maturity gradually decreases and the Mesozoic rock succession becomes mostly in the oil window, rendering the Cretaceous and Jurassic rocks successions along the margin more oil prone as compared to the offshore realm.

In the onshore, due to the water washing affecting the Jurassic and younger strata, the only potential petroleum system is expected in the pre-Jurassic rock succession which at present day is in the oil and early gas window.

The thermal history of sedimentary basins is controlled by many parameters that are often poorly constrained. Thus uncertainties are always present in every basin modelling study, especially in an area with scarce public calibration data like the Levant area. Sensitivity analysis in the offshore Levant basin, where no calibration data is available, showed that the depth of the oil and gas windows can vary considerably resulting in a huge impact on potential petroleum systems. This has also showed that under any scenario, 700 m to 1500 m of the pre-salt Miocene succession in the offshore Levant basin is in the biogenic zone.

- 120 -

Chapter V Conclusion

5.2 Outlook

This thesis targeted several analytical and numerical modelling aspects that each of which can be further addressed in more details. A number of follow up research topics are suggested below:

- The Upper Cretaceous (Campanian – Maastrichtian) organic matter rich rocks exposed

and sampled in Lebanon form an excellent candidate to study the contemporaneous

depositional conditions and the upwelling system at the southern Tethys margin.

Additional analysis can include organic and inorganic carbon isotopes, nitrogen, and

oxygen isotopes, to better assess the depositional environment and the early diagenetic

processes that affected the investigated rock succession.

- Further analysis including sulphur isotopes and compositional kinetics on the Upper

Cretaceous (Campanian – Maastrichtian) source rocks can help better understand the

processes of sulphur incorporation in the organic matter and the effect this has on the

generated hydrocarbon products.

- A model is as good as its input data. The constructed 3D petroleum system model of the

Levant area can be enhanced using refined depth horizons that can delineate actual

structures and with that fluid flow simulations can be realistically achieved.

- Better understanding of the deep lithospheric structures of the Levant basin and its

margins using a combination of seismic, gravity, and magnetic data can considerably

enhance temperature calculations throughout the basin.

- Integration of forward stratigraphic modelling, using detailed seismic facies

interpretation as a constraint (in the absence of well data), with petroleum system

modelling can provide better estimation of reservoirs and seals distribution.

- 121 -

Chapter V Conclusion

- Quantitative modelling of source rocks distribution on a basin scale using modern day

analogues to calculate variations in marine source rock quality as a function of

paleoproductivity, sedimentation rates, paleobathymetry, and distance from the coast.

- 122 -

References

6. References

Abboud, M., Philip, R.P., Allen, J., 2005. Geochemical correlation of oils and source rocks from central and NE Syria. Journal of Petroleum Geology, 28, 203-218.

Abed, A.M., Arouri, K., Boreham, C., 2005. Source rock potential of the phosphorite- bituminous chalk-marl sequence in Jordan. Marine and Petroleum Geology, 22, 413 – 425.

Abdel-Rahman, A-F.M., Nassar, P., 2004. Cenozoic volcanism in the Middle East: petrogenesis of alkali basalts from northern Lebanon. Geological Magazine, 141 (5), 545–563.

Abou Shagar, S., 2000. Time-temperature and organic maturation relations at some wells at the northern Nile Delta, Egypt. Journal of African Earth Sciences, 30 (1), 169-181.

Adiyaman, Ö., Chorowicz, J., 2002. Late Cenozoic tectonics and volcanism in the northwestern corner of the Arabian plate: a consequence of the strike-slip Dead Sea fault zone and the lateral escape of Anatolia. Journal of Volcanology and Geothermal Research, 117, 327 – 345.

Ala, M.A., Moss, B.J., 1979. Comparative petroleum geology of southeast Turkey and northeast Syria. Journal of Petroleum Geology, 1 (4), 3 – 27.

Almogi-Labin, A., Bein, A., Sass, E., 1993. Late Cretaceous upwelling system along the southern Tethys margin (Israel): interrelationship between productivity, bottom water environments, and organic matter preservation. Paleoceanography, 8, 671 – 690.

Almogi-Labin, A., Ashckenazi-Polivoda, S., Edelman-Furstenberg, Y., Benjamini, C., 2012. Anoxia-dysoxia at the sediment-water interface of the southern Tethys in the late Cretaceous: Mishash Formation, southern Israel. In: Anoxia: Evidence for Eukaryote Survival and Paleontological Strategies. In: Altenbach, A.V., Bernhard, J.M. and Seckbach, J. (Eds.), Cellular Origin, Life in Extreme Habitats and Astrobiology, 21, 553 – 572.

Alsenz, H., Illner, P., Ashckenazi-Polivoda, S., Meilijson, A., Abramovich, S., Feinstein, S., Almogi-Labin, A., Berner, Z., Püttmann, W., 2015. Geochemical evidence for the link between sulfate reduction, sulfide oxidation and phosphate accumulation in Late Cretaceous upwelling system. Geochemical Transactions, 16:2. DOI 10.1186/s12932-015-0017-1

Alsharhan, A.S., Salah, M.G., 1997. A common source rock for Egyptian and Saudi hydrocarbons in the Red Sea. American Association of Petroleum Geologists Bulletin, 70, 977 – 1002. al Sandouk-Lincke, N.A., Schwarzbauer, J., Volk, H., Hartkopf-Fröder, C., Fuentes, D., Young, M., Littke, R., 2013. Alteration of organic material during maturation: A pyrolytic and infrared spectroscopic study of isolated bisaccate pollen and total organic matter (Lower Jurassic, Hils Syncline, Germany). Organic Geochemistry, 59, 22 – 36.

- 123 -

References

Arthur, M.A., Jenkyns, H.C., 1981. Phosphorites and paleoceanography. Oceanologica Acta. 4 Supplement, 83 – 96.

Ashckenazi-Polivoda, S., Edelman-Furstenberg, Y., Almogi-Labin, A., Benjamini, C., 2010. Characterization of lowest oxygen environments within ancient upwelling environments: benthic foraminifera assemblages. Palaeogeography Palaeoclimatology Palaeoecology, 289, 134 – 144.

Ashckenazi-Polivoda, S., Abramovich, S., Almogi-Labin, A., Schneider-Mor, A., Feinstein, S., Püttmann, W., Berner, Z., 2011. Paleoenvironments of the latest Cretaceous sequence, Southern Tethys, Israel, as an integral part of the prevailing upwelling system. Palaeogeography Palaeoclimatology Palaeoecology, 305, 93 – 108.

Barwise, A.J.G., 1990. Role of nickel and vanadium in petroleum classification. Energy and Fuels, 4, 647 – 652.

Bar, O., Gvirtzman, Z., Feinstein, S., Zilberman, E., 2011. Late Tertiary subsidence of the Levant margin:Distinction between sedimentary load and tectonics. Geological Survey of Israel 20/2011, pp. 33.

Behar, F., Lorant, F., Lewan, M.D., 2008. Role of NSO compounds during primary cracking of a Type II kerogen and a Type III lignite. Organic Geochemistry, 39, 1–22.

Bein, A., Amit, O., 1982. Depositional environements of the Senonian chert, phosphorite and oil shale sequence in Israel as deduced from their organic matter composition. Sedimentology, 29, 81 – 90.

Bein, A., Almogi-Labin, A., Sass, E., 1990. Sulfur sinks and organic-carbon relationships in Cretaceous organic-rich carbonates — implications for evaluation of oxygen poor depositional- environments. American Journal of Science, 290, 882 – 911.

Berger, W.H., 1979. Impact of deep-sea drilling in paleoceanography. In: M. Talwani, W.Hay and W.B.F. Ryan (Editors), Deep Drilling Results in the Atlantic Ocean: Continental Margins and Paleoenvironment. Maurice Ewing Ser., 3, 297 – 314.

Berner, R.A., 1984. Sedimentary pyrite formation: an update. Geochimica et Cosmochimica Acta, 48, 605 – 615.

Beydoun, Z.R., 1977. Petroleum prospects of Lebanon: Re-evaluation. American Association of Petroleum Geologists Bulletin, 61, 43-64.

Beydoun, Z.R., 1988. The Middle-East: Regional geology and petroleum resources. Beaconsfield, Scientific Press Ltd..

Beydoun, Z.R., Habib, J.G., 1995. Lebanon revisited: New insights into Triassic hydrocarbon prospects. Journal of Petroleum Geology, 18 (1), 75 - 90.

- 124 -

References

Beydoun, Z.R., 1999. Evolution and development of the Levant (Dead Sea Rift) transform system: a historical-chronological review of a structural controversy. In: Mac Niocaill, C., Ryan, P.D. (Eds.), Continental Tectonics, Geological Society, London, Special Publications. 164, 239 – 255.

BouDagher-Fadel, M., Clark, G-N., 2006. Stratigraphy, paleoenvironment and paleogeography of Maritime Lebanon: A key to Eastern Mediterranean Cenozoic history. Stratigraphy, 3, 81- 118.

Bou Daher, S., Nader, F.H., Strauss, H., Littke, R., 2014. Depositional environment and source- rock characterisation of organic- matter rich Upper Turonian - Upper Campanian carbonates, Northern Lebanon. Journal of Petroleum Geology, 37, 1 – 20.

Bou Daher, S., Nader, F.H., Müller, C., Littke, R. 2015. Geochemical and petrographic characterization of Campanian-Lower Maastrichtian calcareous petroleum source rocks of Hasbayya, South Lebanon. Marine and Petroleum Geology, 64, 304 – 323.

Bouillon, S., Boschker, S.T. 2005. Bacterial carbon sources in coastal sediments: a review based on stable isotope data of biomarkers. Biogeosciences Discussions, 2 (5), 1617 – 1644.

Bowman, S.A. 2011. Regional seismic interpretation of the hydrocarbon prospectivity of offshore Syria. GeoArabia, 16, 95-124.

Bray, E.E., Evans, E.D. 1961. Distribution of n-paraffins as a clue to recognition for source beds. Geochimica et Cosmochimica Acta, 22, 2 –15.

Brew, G., Barazangi, M., Al-Maleh, A.K., Sawaf, T., 2001. Tectonic and geologic evolution of Syria. GeoArabia, 6 (4), 573 – 615.

Böcker, J., Littke, R., Hartkopf-Fröder, C., Jasper, K., Schwarzbauer, J., 2013. Organic geochemistry of Duckmantian (Pennsylvanian) coals from the Ruhr Basin, western Germany. International Journal of Coal Geology, 107, 112 – 126.

Calvert, S.E., Bustin, R.M., Ingall, E.D., 1996. Influence of water column anoxia and sediment supply on the burial and preservation of organic carbon in marine shales. Geochimica et Cosmochimica Acta, 60 (9), 1577 – 1593.

Chaimov, T., Barazangi, M., Al-Saad, D., Sawaf, T., Gebran, A., 1992. Mesozoic and Cenozoic deformation inferred from seismic stratigraphy in the southwestern intracontinental Palmyride fold-thrust belt, Syria. Bulletin Geological Society of America, 104, 704 – 715.

Challands, T.J., Armstrong, H.A., Maloney, D.P., Davies, J.R., Wilson, D., Owen, A, W., 2009. Organic-carbon deposition and coastal upwelling at mid-latitude during the Upper Ordovician (Late Katian): A case study from the Welsh Basin, UK. Palaeogeography, Palaeoclimatology, Palaeoecology, 273, 395 – 410.

- 125 -

References

Connan, J., Nissenbaum, A., 2004. The organic geochemistry of the Hasbeya asphalt (Lebanon): comparison with asphalts from the Dead Sea area and Iraq. Organic Geochemistry, 35, 775 – 789.

Dean, K.E., Arthur, M.A., Claypool, G.E., 1986. Depletion of 13C in Cretaceous marine organic matter: source, diagenetic, or environmental signal? Marine Geology, 70, 119 – 157.

Demaison, G.J., Moore, G.T., 1980. Anoxic environments and oil source bed genesis. American Association of Petroleum Geologists Bulletin, 64, 1179 – 1209.

Demaison, G., Huizinga, B., 1991. Genetic Classification of Petroleum Systems. American Association of Petroleum Geologists Bulletin, 75, 1626 – 1643.

Dercourt, J., Gaetani, M., Vrielynck, B., Barrier, E., Biju Duval, B., Brunet, M.F., Cadet, J.P., Crasquin, S., Sandulescu, M. (eds), 2000. Atlas Peri-Tethys, Palaeogeographical maps, 24 maps and explanatory notes: I-XX. CCGM/CGMW, Paris, 269 pp.

Dinur, D., Spiro, B., Aizenshtat, Z., 1980. The distribution and isotopic composition of sulfur in organic-rich sedimentary rocks. Chemical Geology, 31, 37 – 51. di Primio, R., Horsfield, B., 1996. Predicting the generation of heavy oils in carbonate/evaporitic environments using pyrolysis methods, in B. Horsfield, G. Claypool, N. Telnaes, J. A. Curiale, S. Larter, and K. Oygard, eds., Proceedings of the 17th International Meeting on Organic Chemistry: Part III: Origin of natural gases; petroleum geochemistry, impact of organic geochemistry on exploration; migration and expulsion of oil and gas: Oxford, Pergamon, Organic Geochemistry, v. 24, no. 10 – 11, p. 999 – 1016.

Dubertret, L., 1955. Carte géologique du Liban au 1/200000 avec notice explicative: Beirut, République Libanaise, Ministère des Travaux Publiques, 74 pp.

Dubertret, L., 1975. Introduction à la carte géologique au 1/50000 du Liban. Notes et Mémoires sur le Moyen-Orient, 23, 345 – 403.

Dubille, M., Thomas D., 2012. Petroleum System Assessment of the Offshore Lebanon Thorugh 3D Basin Modeling. Presentation, Lebanese International Petroleum Exploration Forum (LIPE), Beirut, Lebanon.

Dunham, R.J., 1962. Classification of carbonate rocks according to their depositional texture. In Ham, W.E., ed., Classification of Carbonate Rocks, American Association of Petroleum Geologists Memoir, 1. 108 – 121.

Durand, B., Nicaise, G., 1980. Procedure of kerogen isolation. In: Durand, B. (Ed.), Kerogen, Insoluble Organic Matter from Sedimentary Rocks. Editions Technip, Paris, pp. 13–34.

Eckstein, Y., Simmons, G., 1978. Measurement and interpretation of terrestrial heat flow from Israel. Geothermics, 6, 117 – 142.

- 126 -

References

Edelman-Furstenberg, Y., 2008. Macrobenthic community structure in a high productivity region: Upper Campanian Mishash Formation (Israel). Palaeogeography, Palaeoclimatology, Palaeoecology, 261, 58 – 77.

Edelman-Furstenberg, Y., 2009. Cyclic upwelling facies along the Late Cretaceous southern Tethys (Israel): taphonomic and ichnofacies evidence of a high productivity mosaic. Cretaceous Research, 30, 847 – 863.

Esestime, P., Hewitt, A., Hodgson, N., 2016. Zohr – A newborn carbonate play in the Levantine Basin, East-Mediterranean. First Break, 34, 87-93.

Espitalié, J., Laporte, J.L., Madec, M., Marquis, F., Leplat, P., Paulet, J., Boutefeu, A., 1977. Méthode rapide de caractérisation des roches mères de leur potential pétrolier et de leur degré d'évolution. Revue de l’Institut Français du Pétrole, 32 (1), 23 – 42.

Feinstein, S., Aizenshtat, Z., Miloslavsky, I., Gerli, N.P., Slager, J., McQuilken, J., 2002. Genetic characterization of gas shows in the east Mediterranean offshore of southwestern Israel. Organic Geochemistry, 33:12, 1401-1413.

Flexer A., Honigstein. A, 1984. The Senonian succession in Israel—lithostratigraphy, biostratigraphy and sea level changes. Cretaceous Research, 5, 302 – 312.

Folk, R.L., Ward, W., 1957. Brazos River bar: a study in the significance of grain-size parameters. Journal of Sedimentary Petrology, 27, 3 – 26.

Gardosh, M., 2013. Exploration in the Eastern Mediterranean, Offshore Israel – Ongoing Success and Future Challenges. American Association of Petroleum Geologists Search and Discovery Article #90161©2013 American Association of Petroleum Geologists European Regional Conference, Barcelona, Spain, 8-10 April 2013.

Gardosh, M., Druckman, Y., 2006. Seismic stratigraphy, structure and tectonic evolution of the Levantine basin, offshore Israel. In: A.H.F. Robertson and D. Mountrakis, (Eds.), Geological Society of London Special Publication, 260, 201 – 227.

Gardosh, M., Druckman, Y., Buchbinder, B., Rybakov, M., 2008. The Levant Basin offshore Israel: stratigraphy, structure, tectonic evolution and implications for hydrocarbon exploration. Geophysical institute of Israel, 429/328/08, 1 – 119.

Gardosh, M.A., Garfunkel, Z., Druckman, Y., Buchbinder, B., 2010. Tethyan rifting in the Levant Region and its role in Early Mesozoic crustal evolution. Geological Society of London Special Publication, 341, 9 – 36.

Garfunkel, Z., 1989. Tectonic setting of Phanerozoic magmatism in Israel. Israel Journal of Earth-Sciences, 38, 51 – 74.

- 127 -

References

Garfunkel, Z., 1992. Darfur-Levant array of volcanics – A 140-Ma-long record of hot spot beneath the African-Arabian continent and its bearing on Africa’s absolute motion. Israel Journal of Earth Sciences, 40, 135-150.

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

Geng, M., Duan, Z. 2010. Prediction of oxygen solubility in pure water and brines up to high temperatures and pressures. Geochimica et Cosmochimica Acta, 74, 5631 – 5640.

Ghalayini, R., J.-M. Daniel, C. Homberg, F. H. Nader, J. E. Comstock (2014), Impact of Cenozoic strike-slip tectonics on the evolution of the northern Levant Basin (offshore Lebanon). Tectonics, 33, doi:10.1002/2014TC003574.

Gomez, F., Khawlie, M., Tabet, C., Darkal, A.N., Khair, K., Barazangi, M., 2006. Late Cenozoic uplift along the northern Dead Sea transform in Lebanon and Syria. Earth and Planetary Science Letters, 241, 913-931.

Haq, B.U., 2014. Cretaceous eustasy revisited. Global and Planetary Change, 113, 44 – 58.

Haq, B.U., Hardenbohl, J., Vail, P.R., 1988. Mesozoic and Cenozoic chronostratigraphy and eustatic cycles. In: C.K. Wilgus, B.J. Hastings, G.G. Kendall, H. Posamentier, C.A. Ross and J.C. Van Wagoner (Eds.), Sea-level changes: an integrated approach. SEPM Special Publications, 42, 71 – 108.

Hardenberg, M.F., Robertson, A.H.F, 2007. Sedimentology of the NW margin of the Araban plate and the SWNE trending Nahr el Kabir half graben in northern Syria during the latest Cretaceous and Cenozoic. Sedimentary Geology, 201, 231-266.

Hawie, N., 2014. Architecture, geodynamic evolution and sedimentary filling of the Levant Basin: A 3D quantitative approach based on seismic data. PhD thesis, Universite Pierre et Marie Curie. pp. 250.

Hawie, N., Gorini, C., Deschamps, R., Nader, F.H., Montadert, L., Grajeon, D., Baudin, F, 2013a. Tectono-stratigraphic evolution of the northern Levant Basin (offshore Lebanon). Marine and Petroleum Geology, 48. 392 – 410.

Hawie, N., Deschamps, R., Nader, F.H., Gorini, C., Müller, C., Desmares, D., Hoteit, A., Granjeon, D., Montadert, L., Baudin, F., 2013b. Sedimentologic and stratigraphic evolution of northern Lebanon since the Late Cretaceous: implications on the Levant margin and basin. Arabian Journal of Geosciences (DOI: 10.1007/s12517-013-0914-5).

Hayes, J.M., Strauss, H., Kaufman, A.J., 1999. The abundance of 13C in marine organic matter and isotopic fractionation in the global biogeochemical cycle of carbon during the past 800 Ma. Chemical Geology, 161, 103 – 125.

- 128 -

References

Hein, J.R., Obradovic, J., 1989. Siliceous deposits of the Tethys and Pacific Regions. Springer- Verlag, New York, 244 pp.

Henson, F.R. S., 1951. Observations on the Geology and Petroleum Occurrences of the Middle East. 3rd World Petroleum Congress.

Hofmann, P., Ricken, W., Schwark, L., Leythaeuser, D., 2000. Carbon–sulfur–iron relationships and d13C of organic matter for late Albian sedimentary rocks from the North Atlantic Ocean: paleoceanographic implications. Palaeogeography, Palaeoclimatology, Palaeoecology, 163, 97 – 113.

Huc, A.Y. 1988. Sedimentology of organic matter. In: Frimmel, F.H. and Christman, R.F. (eds): Humic substances and their role in the environment, 215 – 243, John Wiley and Sons, Chichester.

Inan, S., Alp Ugur, F., Inan, T., Yalcin, M.N., Mann, U., 2010. Relationship between organic matter, sulphur and phosphate contents in upper cretaceous marine carbonates (Karabogaz Formation, SE Turkey): Implications for early oil generation. Journal of Petroleum Geology, 33 (4), 319 – 338.

Jacob, H., 1989. Classification, structure, genesis and practical importance of natural solid oil bitumen (migrabitumen). International Journal of Coal Geology 11, 65 – 79.

Jiménez-Munt, I., R. Sabadini, A. Gardi, G. Bianco, 2006. Active deformation in the Mediterranean from Gibraltar to Anatolia inferred from numerical modeling and geodetic and seismological data. J. Geophys. Res., 108(B1), doi:10.1029/ 2001JB001544, 2003.

Khair, K., Khawlie, M., Haddad, F., Barazangi, M., Chaimov, T., Seber, D., 1993. Gravity modeling and geologic setting and tectonics of the central segment of the Dead Sea fault zone. Geology, 21, 739-742.

Khair, K., Tsokas, G.N., Sawaf, T., 1997. Crustal structure of the northern Levant region: multiple source Werner deconvolution estimates for Bouger gravity anomalies. Geophysics Journal International, 128, 605 - 616.

Krenkel, E., 1924. Der Syriasche Bogen. Zentralbulletin für Mineralogie Geologie und Palaontologie 9, 274 – 281. Krenkel, E., 1924. Der Syriasche Bogen. Zentralbulletin für Mineralogie Geologie und Palaontologie, 10, 301 – 313.

Kruijs, E., Barron, E.J., 1990. Climate model prediction of paleoproductivity and potential source rock distribution. In: Huc, A.Y. (ed.): Deposition of organic facies. American Association of Petroleum Geologist. Studies in Geology, 30, 195 – 216.

Kolonic, S., Sinninghe Damsté, J.S., Böttcher, M.E., Kuypers, M.M.M., Kuhnt, W., Beckmann, B., Scheeder, G., Wagner, T., 2002. Geochemical characterization of Cenomanian/Turonian black shales from the Tarfaya Basin (SW Morocco). Journal of Petroleum Geology, 25, 325– 350. - 129 -

References

Lafargue, E., Marquis, F., Pillot, D., 1998. Rock-Eval 6 applications in hydrocarbon exploration, production and soils contamination studies. Oil and Gas Science and Technology, 53, 421 – 437.

Lallier-Vergès, E., Bertrand, P., Huc, A.Y., Tremblay, P., Bückel, D., 1993. Control of the preservation of organic matter by productivity and sulphate reduction in marine anoxic sediments: Kimmeridgian shales from Dorset (UK). Marine and Petroleum Geology, 10, 600 – 605.

Langford, F.F., Blanc-Valleron, M.M., 1990. Interpreting Rock-Eval Pyrolysis data using graphs of pyrolizable hydrocarbons vs. total organic carbon. American Association of Petroleum Geologist Bulletin, 74 (6), 799 – 804.

Leythaeuser, D., Welte, G.H. 1969. Relation between ditribution of heavy n-paraffins and coalification in Carboniferous coals from the Saar district, Germany. In: SCHENCK, P.A. and Havenaar, I. (eds): Advances in organic geochemistry, 1968, 429 – 442, Pergamon Press, Oxford.

Littke, R., 1993. Deposition, diagenesis and weathering of organic matter-rich sediments. In: Lecture Notes in Earth Sciences. Springer-Verlag, Berlin, 216 pp.

Littke, R., Klussmann, U., Krooss, B., Leythaeuser, D., 1991. Quantification of calcite-, pyrite- , and organic matter-loss due to weathering of Toarcian black shales and effects on kerogen and bitumen characteristics. Geochimica et Cosmochimica Acta, 55, 3369 – 3378.

Littke, R., Sachsenhofer, R.F., 1994. Organic petrology of deep sea sediments: a compilation of results from the Ocean Drilling Program and the Deep Sea Drilling Project. Energy and Fuels, 8, 1498 – 1512.

Littke, R., Lückge, A., Wilkes, H., 1998. In: Weaver, P.P.E., Schmincke, H.U., et al. (Eds.), Organic matter in Neogene sediments of the southern Canary Channel, Canary Islands (ODP Sites 955 and 956). Proceedings of the Ocean Drilling Program, 157, 361–372.

Littke, R., Urai, J.L., Uffmann, A.K., Risvanis, F., 2012. Reflectance of dispersed vitrinite in Palaeozoic rocks with and without cleavage: implications for burial and thermal history modeling in the Devonian of Rursee area, northern Rhenish Massif, Germany. International Journal of Coal Geology, 89, 41–50.

Lückge, A., Boussafir, M., Lallier-Vergѐs, E., Littke, R., 1996. Comparative study of organic matter preservation in immature sediments along the continental margins of Peru and Oman. Part I: results of petrographical and bulk geochemical data. Organic Geochemistry, 24, 437 – 451.

Lučić, D., Foršek, G., et al. 2003, Depositional environment and reservoir properties of Kurrachine Dolomite and Markada Formation in Central Palmyrides, Syria, Abstract, 22nd IAS Meeting of Sedimentology, Opatija, 110-111.

- 130 -

References

Lučić, D. Ivkoviċ, Z., Foršek, G., Takač, D., Bubniċ, J., Koch, G., 2010. Depositional sequences and palynology of Triassic carbonate-evaporite platform deposits in the Palmyrides, Syria in VAN BUCHEM, F. S. P., GERDES, K. D. & ESTEBAN, M. (eds) Mesozoic and Cenozoic Carbonate Systems of the Mediterranean and the Middle East: Stratigraphic and Diagenetic Reference Models. Geological Society of London Special Publications, 329, 43–63.

Lüning, S., Kolonic, S., Belhadj, E.M., Belhadj, Z., Cota, L., Baric, G., Wagner, T., 2004. Integrated depositional model for the Cenomanian–Turonian organic-rich strata in North Africa. Earth-Science Reviews, 64, 51–117.

Makris, J., Ben-Avraham, Z., Behle, A., Ginzburg, P., Gieze, P., Steinmetz, L., Whitmarsh, R.B., Eleftheriou, S., 1983. Seismic refraction profiles between Cyprus and Israel and their interpretation. Geophysical Journal, Royal Astronomical Society, 75, 575-591.

Makris, J., Stobbe, C., 1984. Physical properties and state of the crust and upper mantle of the Eastern Mediterranean Sea deduced from geophysical data. Marine Geology, 55, 347 - 363.

May, P.R., 1991. The Eastern Mediterranean Mesozoic Basin: evolution and oil habitat. American Association of Petroleum Geologist Bulletin, 75, 1215 – 1232.

Minster, T., Nathan, Y., Ravh, A., 1992. Carbon and sulfur relationships in marine Senonian organic-rich, iron-poor sediments from Israel - a case study. Chemical Geology. 97, 145 – 161.

Moldowan, J.M., Sundaraman, P., Schoell, M., 1985. Sensitivity of biomarker properties of depositional environment and/or source input in the Lower Toarcian of SW Germany. Organic Geochemistry, 10, 915 – 926.

Montadert, L., Nicolaides, S., Semb P.H., Lie, Ø. 2014, Petroleum systems offshore Cyprus, in L. Marlow, C. Kendall and L. Yose, eds., Petroleum systems of the Tethyan region: American Association of Petroleum Geologist Memoir, 106, p. 301 – 334

Müller, C., Higazi, F., Hamdan, W., Mroueh, M., 2010. Revised stratigraphy of the Upper Cretaceous and Cenozic series of Lebanon based on nannofossils. Geological Society of London Special Publications, 341, 287 – 303.

Nader, F.H., 2000. Petrographic and geochemical characterization of the Jurassic-Cretaceous carbonate sequence of the Nahr Ibrahim region, Lebanon. Master’s thesis, American University of Beirut, Lebanon.

Nader, F.H., 2003. Petrographic and Geochemical Study of the Kesrouane Formation (Jurassic), Mount Lebanon: Implications on Dolomitization and Petroleum Geology. Unpublished PhD thesis, Katholieke Universiteit Leuven, Belgium.

Nader, F.H, 2011. The petroleum prospectivity of Lebanon: an overview. Journal of Petroleum Geology, 34 (2), 135 – 156.

Nader, F.H., 2014. The Geology of Lebanon. Scientific Press Ltd, United Kingdom, p. 108.

- 131 -

References

Nader, F.H., Swennen, R., 2004. Petroleum prospects of Lebanon: Some remarks from sedimentological and diagenetic studies of Jurassic carbonates. Marine and Petroleum Geology, 21, 427 - 441.

Nader, F.H., Swennen, R., Keppens, E., 2008. Calcitization/dedolomitization of Jurassic dolostones (Lebanon): results from petrographic and sequential geochemical analyses. Sedimentology, 55, 1467 – 1485.

Naylor, D., Al-Rawi, M., Clayton, G., Fitzpatrick, M.J., Green, P.F., 2013. Hydrocarbon potential in Jordan. Journal of Petroleum Geology, 36 (3), 205-236.

Needham, D., Hosler, J., Nowak, S., Christensen, C., Ffrench, J., 2013. The Tamar Field from Discovery to Production (Abstract). American Association of Petroleum Geologist Search and Discovery Article #90161©2013 American Association of Petroleum Geologist European Regional Conference, Barcelona, Spain, 8-10 April 2013.

Netzeband, G.L., Gohl, K., Hubscher, C.P., Ben-Avraham, Z., Dehghani, G.A., Gajewski, D., Liersch, P., 2006. The Levantine Basin crustal structure and origin. Tectonophysics, 418, 167- 188.

Noeth, S., Thomsen, R.O., Littke, R., 2002. A method for assessing statistical significance and uncertainties for calibration of 1-D thermal basin maturation models. American Association of Petroleum Geologist Bulletin, 86 (3), 417 – 431.

Noeth, S., Karg, H., Littke, R., 2001. Reconstruction of Late Paleozoic heat flows and burial histories at the Rhenohercynian-Subvariscan boundary, Germany. International Journal of Earth Sciences, 90, 234 – 256.

Noujaim, G., 1977. Pétrographie et environnements sédimentaires de l’Albien, Cénomanien, Turonien dans les environs Nord de Beyrouth (Liban): Diplôme de Docteur 3° cycle, Université Pierre et Marie Curie, Paris VI.

OAPEC Annual Report, 2002. Secretary General's Annual Report - Organization of Arab Petroleum Exporting Countries, 29, pp. 229.

Parrish, J.T., Curtis, R.L., 1982. Atmospheric circulation, upwelling, and organic-rich rocks in the Mesozoic and Cenozoic eras. Palaeogeography, Palaeoclimatology, Palaeoecology, 40, 31 – 66.

Paulmier, A., Ruiz-Pino, D., 2009. Oxygen minimum zones (OMZs) in the modern ocean. Progress in Oceanography, 80, 113–128.

Pedersen, T.F., Calvert, S.E. (1990). Anoxia versus productivity: what controls the formation of organic-carbon-rich sediments and sedimentary rocks? American Association of Petroleum Geologist Bulletin, 74, 454-466.

- 132 -

References

Peters, K.E., 1986 Guidelines for Evaluating Petroleum Source Rock Using Programmed Pyrolysis. American Association of Petroleum Geologist Bulletin, 70 (3), 318 – 329.

Peters, K.E., Moldowan, J.M., 1991. Effect of source, thermal maturity and biodegradation on the distribution and isomerisation of homohopanes in petroleum. Organic Geochemistry, 17, 47 – 61.

Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The biomarker guide. In: Biomarkers and Isotopes in Petroleum Exploration and Earth History, second ed., vol. 2. Cambridge University Press, Cambridge.

Perch-Nielsen, K., 1985. Mesozoic calcareous nanno- fossils. In: Bolli H. M., Saunders J. B., and Perch- Nielsen, K. (eds) Plankton Stratigraphy, Cambridge, 329 – 426.

Ponikarov, V.P., 1966. The Geology of Syria. In: Explanatory Notes on the Geological Map of Syria, Scale 1:200 000. Ministry of Industry, Syrian Arab Republic.

Powell J.H., Moh’d B.K., 2011. Evolution of Late Cretaceous to Eocene alluvial and carbonate platform sequences in Jordan. GeoArabia, 17(3):29–82.

Reimers, C.E., Suess, E. (1983). Spatial and temporal patterns of organic matter accumulation on the Peru continental margin. In: Thiede, J. and Suess, E. (eds): Coastal upwelling. Part B: Sedimentary records of ancient coastal upwelling, 311 – 346, Plenum Press, New York.

Renouard, G., 1955. Oil Prospects of Lebanon. American Association of Petroleum Geologist Bulletin, 39 (11), 2125 – 2169.

Robertson, A.H.F., 1998a. Mesozoic – Tertiary tectonic evolution of the Easternmost Mediterranean area; integration of marine and land evidence. In: Robertson, A.H.F., Emeis, K.C., Richter, K.C., Camerlenghi, A. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results. 160, 723 – 782.

Robertson, A.H.F., 1998b. Lithofacies evidence for the Cretaceous – Paleogene sedimentary history of Eratosthenes seamount, Eastern Mediterranean, in its regional tectonic context. In: Robertson, A.H.F., Emeis, K.C., Richter, K.-C., Camerlenghi, A. (Eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 160, 403 – 417.

Roded, R., Shalev, E., Katoshevski, D., 2013. Basal heat-flow and hydrothermal regime at Golan-Ajloun hydrological basins. Journal of Hydrology, 476, 200 – 211.

Sachse, V.F., Littke, R., Heim, S., Kluth, O., Schober, J., Boutib, L., Jabour, H., Perssen, F., Sindern, S., 2011. Petroleum source rocks of the Tarfaya Basin and adjacent areas, Morocco. Organic Geochemistry, 42, 209 – 227.

Sachse, V.F., Littke, R., Jabour, H., Schümann, T., Kluth, O., 2012. Late Cretaceous (Late Turonian, Coniacian and Santonian) petroleum source rocks as part of an OAE, Tarfaya Basin, Morocco. Marine and Petroleum Geology, 29, 35 – 49.

- 133 -

References

Saint-Marc, P., 1972. Le Crétacé inférieur et moyen du bord occidental du Jabal-Sannine (Liban). Notes et Mémoires sur le Moyen-Orient, 13, 217-226.

Saint-Marc P., 1974. — Étude stratigraphique et micropaléontologique de l’Albien, du Cénomanien et du Turonien du Liban. Notes et Mémoires sur le Moyen-Orient, 13, 342 p.

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 Medi- terranean area. Tectonophysics, 425, 137–157.

Schenk, H.J., Horsfield, B., Krooß, B., Schaefer, R.G., Schwochau, K. (1997) Kinetics of petroleum formation and cracking. In Petroleum and Basin Evolution, eds. D. H. Welte et al. Springer, Berlin, 535 pp.

Schneider-Mor, A., Alsenz, H., Ashckenazi-Polivoda, S., Illner, P., Abramovich, S., Feinstein, S., Almogi-Labin, A., Berner, Z., Püttmann, W., 2012. Paleoceanographic reconstruction of the late Cretaceous oil shale of the Negev, Israel: Integration of geochemical, and stable isotope records of the organic matter. Palaeogeography, Palaeoclimatology, Palaeoecology, 319 – 320, 46 – 57.

Schoenherr, J., Littke, R., Urai, J.L., Kukla, P.A., Rawahi, Z., 2007. Polyphase thermal evolution in Infra-Cambrian Ara Group (South Oman Salt Basin) as deduced by maturity of solid reservoir bitumen. Organic Geochemistry, 38, 1293 – 1318.

Schütz, F., Förster, H.J., Förster, A., 2014. Thermal conditions of the central Sinai Microplate inferred from new surface heat-flow values and continuous borehole temperature logging in central and southern Israel. Journal of Geodynamics, 76, 8 – 24.

Shalev, E., Lyakhovsky, V., Weinstein, Y., Ben-Avraham, Z., 2013. The thermal structure of Israel and the Dead Sea Fault. Tectonophysics, 602, 69 – 77.

Shanmugam, G., 1985. Significance of coniferous rain forests and related organic matter in generating commercial quantities of oil, Gippsland Basin, Australia. American Association of Petroleum Geologists Bulletin, 69, 1241 – 1254.

Sharland, P.R., Archer, R., Casey, D.M., Davies, R.B., Hall, S.H., Heward, A.P., Horbury, A. D., Simmons, M. D., 2001. Arabian plate sequence stratigraphy. GeoArabia Special Publication, 2.

Sissingh, W., 1977. Biostratigraphy of Cretaceous calcareous nannoplankton. Geol Mijnbouw. 56, 37 – 65.

Soudry, D., 2000. Carbonate-phosphate competition in the Negev phosphorites (southern Israel): a microstructural study, In: Marine Authigenesis 66. SEPM Special Publication 415 – 426.

- 134 -

References

Stein, R. 1986. Surface-water paleo-productivity as inferred from sediments deposited in oxic and anoxic deep water environments of the Mesozoic Atlantic ocean. Mitt. Geol.-Paläont. Inst. Univ. Hamburg, 60, 55 – 70.

Strauss, H., Desmarais, D.J., Hayes, J.M., Summons, R.E., 1992. The carbon-isotopic record. In: Schopf, J.W., Klein, C. (Eds.), The Proterozoic Biosphere, a Multidisciplinary Study. Cambridge Univ. Press, Cambridge, UK, 117–128.

Tannenbaum, E., Aizenshtat, Z., 1985. Formation of immature asphalt from organic-rich carbonate rocks-I, geochemical correlation. Organic Geochemistry, 8(2), 181 – 192.

Taylor, G.H., Teichmüller, M., Davis, A., Diessel, C.F.K., Littke, R., Robert, P., 1998. Organic Petrology. Borntraeger, Stuttgart, 704 pp.

Tannenbaum, E. and Lewan, M.D., 2003. Regional Upper Cretaceous (Senonian) petroleum source rocks in the Middle East and North Africa (abstract). American Association of Petroleum Geologist, Annual Meeting.

Thamdrup, B., Dalsgaard, T., Revsbech, N.P., 2012. Widespread functional anoxia in the oxygen minimum zone of the Eastern South Pacific. Deep-Sea Research I. 65, 36 – 45.

Thierstein, H.R., 1976. Mesozoic calcareous nannoplankton biostratigraphy of marine sediments. Marine Micropaleontology, 1, 325 – 362.

Tissot, B.P., Welte, D.H., Durand, B., 1987. The role of geochemistry in exploration, risk evaluation and decision making. Proceedings of the 12th World Petroleum Congress, 2, 99 – 112.

Tribovillard, N., Algeo, T.J., Lyons, T., Riboulleau, A., 2006. Trace metals as paleoredox and paleoproductivity proxies: An update. Chemical Geology, 232, 12 – 32.

Ukla, S., 1970. Subsurface Geology and Well Correlation in North and Central Lebanon. Master’s thesis. American University of Beirut, Lebanon, p. 125.

Villinski, J., 2013. Unusual but Effective Petroleum Systems – Disseminated Terrestrial Organic Matter of the Upper Oligocene as the Primary Source Rock, Offshore Nile Delta, Egypt (Abstract). American Association of Petroleum Geologist Search and Discovery Article #90161©2013 American Association of Petroleum Geologist European Regional Conference, Barcelona, Spain.

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.D., 2001. The Lebanon passive margin and the evolution of the Levantine Neotethys. In: P.A. Ziegler, W. Cavazza, A.H.F. Robertson and S. Crasquin-Soleau (Eds.), Peri-Tethys Memoir 6: Peri-Tethyan Rift / Wrench Basins and Passive Margins. Mémoire du Muséum national d’Histoire naturelle. Paris 86, 407 – 439.

- 135 -

References

Wilkin, R.T., Barnes, H.L., Brantley, S.L., 1996. The size distribution of framboidal pyrite in modern sediments: an indicator of redox conditions. Geochimica et Cosmochimica Acta, 60, 3897 – 3912.

Wilson, M., 1992. Magmatism and continental rifting during the opening of the South Atlantic Ocean: a consequence of Lower Cretaceous super-plume activity. In: B. C. Storey, T. Alabaster and R. J. Pankhurst (Eds.), Magmatism and the Causes of Continental Break-up. Geological Society of London Special Publications, 68, 241-255.

Wolfart, R., 1967. Geologie von Syrien und dem Libanon. Berlin, Gebruder Borntraeger.

Wygrala, B.P., 1989. Integrated study of an oil field in the southern Po Basin, Northern Italy. PhD thesis, University of Cologne, Germany.

Wygrala, B., Rottke, W., Kornpihl, D., Neumaier, M., Al-Balushi, A., Marlow, L., 2014. Assessment of Controlling Factors in Mixed Biogenic and Thermogenic Petroleum Systems – A Case Study from the Levantine Basin. Search and Discovery Article #10636, American Association of Petroleum Geologist Annual Convention and Exhibition, Houston, Texas, April 6-9, 2014.

Zilberman, E., Calvo, R., 2013. Remnants of Miocene fluvial sediments in the Negev Desert, Israel, and the Jordanian Plateau: Evidence for an extensive subsiding basin in the northwestern margins of the Arabian plate. Journal of African Earth Sciences, 82, 33-53.

- 136 -

Appendix

7. Appendix

13 Appendix 1. Organic geochemical data (TOC, TIC, TS and δ Corg) from the Turonian – upper Campanian succession at Chekka, northern Lebanon.

δ 13C Depth TOC TIC TS org Sample (‰ Age (m) (%) (%) (%) VPDB) 12/668 1.0 0.06 10.78 0.04 U. Camp. 12/669 3.0 0.03 11.61 0.02 U. Camp. 12/671 5.0 0.02 11.35 0.02 U. Camp. 12/673 7.0 0.03 10.94 0.04 U. Camp. 12/676 9.0 0.13 9.78 0.06 U. Camp. 12/680 11.0 0.02 11.33 0.04 U. Camp. 12/684 13.0 0.11 10.94 1.00 U. Camp. 12/688 15.0 0.25 9.12 0.59 -27.23 U. Camp. 12/692 17.0 0.11 9.97 0.55 U. Camp. 12/696 19.0 0.09 10.11 0.57 U. Camp. 12/700 21.0 0.02 10.81 0.04 U. Camp. 12/701 23.0 0.04 9.93 0.04 U. Camp. 12/702 27.0 0.03 10.84 0.04 U. Camp. 12/716 28.0 0.19 9.91 0.32 U. Camp. 12/724 32.0 0.07 11.04 0.21 U. Camp. 12/725 32.5 0.12 10.18 0.16 U. Camp. 12/726 33.0 0.18 9.64 0.62 -27.11 U. Camp. 12/728 34.5 0.15 9.75 0.40 U. Camp. 12/729 35.0 0.10 10.61 0.22 U. Camp. 12/731 36.0 0.11 10.15 0.35 U. Camp. 12/734 37.5 0.14 10.17 0.49 -27.59 U. Camp. 12/736 38.5 0.03 10.53 0.09 U. Camp. 12/741 41.0 0.02 11.01 0.06 U. Camp. 12/743 42.0 0.05 9.97 0.02 U. Camp. 12/748 44.5 0.05 9.91 0.03 U. Camp. 12/751 47.0 0.03 9.66 0.02 U. Camp. 12/752 48.0 0.04 10.29 0.04 U. Camp. 12/753 49.0 0.05 10.48 0.07 U. Camp. 12/754 52.0 0.16 10.23 0.26 U. Camp. 12/755 54.0 0.24 10.08 0.63 U. Camp. 12/757 55.5 0.28 10.00 0.38 -27.70 U. Camp. 12/759 56.5 0.48 9.98 0.07 U. Camp. 12/762 58.0 0.33 9.94 0.49 U. Camp. 12/764 59.0 0.12 11.04 0.23 U. Camp. 12/166 60.0 0.18 10.61 0.29 U. Camp. 12/169 61.5 0.20 10.38 0.89 U. Camp. 12/171 62.5 0.41 10.29 0.35 U. Camp. 12/173 63.5 0.25 8.63 0.33 U. Camp. 12/175 64.5 0.52 9.62 0.70 U. Camp. 12/177 65.5 0.92 10.59 0.43 -28.17 U. Camp. 12/180 67.0 0.07 4.56 0.03 U. Camp. 12/182 68.0 0.20 10.23 0.26 U. Camp. 12/184 69.0 0.28 8.70 0.31 U. Camp. 12/186 70.0 0.19 10.12 0.38 U. Camp. 12/187 71.0 0.21 9.40 0.33 U. Camp.

- 137 -

Appendix

12/189 72.0 0.29 10.40 0.28 U. Camp. 12/190 72.5 0.82 9.09 0.57 -28.02 U. Camp. 12/192 73.5 0.59 9.00 0.43 U. Camp. 12/194 74.5 0.33 10.15 0.34 U. Camp. 12/196 75.5 0.63 9.66 0.58 U. Camp. 12/199 77.0 1.05 6.99 0.62 U. Camp. 12/201 78.5 0.77 10.05 0.54 U. Camp. 12/203 79.5 1.53 9.30 0.66 -28.35 U. Camp. 12/206 81.0 1.12 10.38 0.43 U. Camp. 12/211 83.5 0.74 6.47 0.38 U. Camp. 12/213 84.5 1.74 11.91 0.53 U. Camp. 12/215 85.5 0.91 9.61 0.40 U. Camp. 12/217 86.5 1.12 10.02 0.45 -28.71 U. Camp. 12/219 87.5 1.60 9.37 0.69 U. Camp. 12/222 89.0 1.34 10.04 0.61 U. Camp. 12/225 90.5 1.69 8.99 0.70 U. Camp. 12/227 91.5 1.53 9.76 0.60 U. Camp. 12/229 92.5 1.57 8.95 0.63 -28.54 U. Camp. 12/231 93.5 1.59 8.97 0.65 U. Camp. 12/233 94.5 2.57 8.37 1.06 U. Camp. 12/235 95.5 2.03 8.17 0.74 U. Camp. 12/238 97.0 2.31 8.66 0.83 U. Camp. 12/240 98.0 1.51 5.65 0.64 U. Camp. 12/242 99.0 2.50 8.45 0.93 -28.53 U. Camp. 12/245 100.5 2.66 8.99 0.75 U. Camp. 12/247 101.5 2.70 7.67 0.88 U. Camp. 12/249 102.5 2.72 8.18 0.89 U. Camp. 12/251 103.5 1.97 6.07 0.76 U. Camp. 12/252 104.0 1.57 5.74 0.53 -28.64 U. Camp. 12/253 111.0 1.97 7.70 0.97 -28.51 U. Camp. 12/255 112.0 1.78 6.73 0.66 U. Camp. 12/258 114.0 0.82 4.17 0.48 -28.54 U. Camp. 12/260 115.0 1.27 9.13 0.55 U. Camp. 12/262 116.0 1.43 4.13 0.42 U. Camp. 12/264 117.0 1.79 8.01 0.58 U. Camp. 12/266 118.0 2.54 8.51 0.86 -28.65 U. Camp. 12/268 119.0 2.07 8.36 0.59 U. Camp. 12/270 120.0 2.33 8.55 0.82 U. Camp. 12/272 121.0 2.41 8.93 0.83 U. Camp. 12/274 122.0 2.38 9.20 0.80 U. Camp. 12/277 123.5 1.72 8.12 0.87 -27.99 U. Camp. 12/279 124.5 1.60 9.50 0.76 U. Camp. 12/281 125.5 1.78 8.98 0.88 U. Camp. 12/283 126.5 2.18 8.56 1.19 U. Camp. 12/286 128.0 2.11 3.65 0.85 U. Camp. 12/288 129.0 3.83 6.93 1.12 U. Camp. 12/290 130.0 2.34 8.39 0.96 -28.16 U. Camp. 12/006 131.0 2.37 9.11 0.87 -28.24 U. Camp. 12/007 131.5 1.85 9.46 0.89 U. Camp. 12/009 133.0 1.34 9.42 0.81 U. Camp. 12/011 134.0 2.02 8.91 0.86 U. Camp. 12/013 135.0 2.11 8.55 1.04 U. Camp. 12/014 135.5 2.15 4.86 0.87 U. Camp. 12/018 137.5 1.98 9.25 0.67 -27.69 U. Camp. 12/022 139.5 2.33 8.50 1.03 U. Camp.

- 138 -

Appendix

12/025 141.0 1.15 9.57 0.51 -28.26 U. Camp. 12/027 143.5 2.62 8.28 0.86 U. Camp. 12/028 144.5 1.91 9.09 0.81 -28.70 U. Camp. 12/031 146.0 2.24 8.66 0.95 U. Camp. 12/033 147.0 2.29 9.05 0.73 -28.98 U. Camp. 12/034 147.5 1.76 8.46 0.82 U. Camp. 12/039 150.0 3.28 8.89 0.96 -28.65 U. Camp. 12/042 151.5 2.72 9.31 0.67 U. Camp. 12/044 152.5 2.56 9.66 0.65 U. Camp. 12/045 153.5 2.87 9.63 0.87 -28.24 U. Camp. 12/048 155.0 2.98 8.46 1.15 U. Camp. 12/050 156.0 1.32 10.35 0.42 U. Camp. 12/052 157.0 1.24 10.09 0.61 U. Camp. 12/053 157.5 3.42 9.16 0.98 -28.27 U. Camp. 12/056 159.0 2.37 8.59 0.65 U. Camp. 12/059 160.5 2.47 8.44 0.63 -28.23 U. Camp. 12/062 162.0 2.91 8.34 1.05 U. Camp. 12/065 163.5 2.78 8.99 1.01 -28.07 U. Camp. 12/068 165.0 2.21 8.56 1.01 U. Camp. 12/071 166.5 2.54 8.47 1.17 U. Camp. 12/073 167.5 1.38 10.30 0.49 U. Camp. 12/076 168.5 1.30 10.23 0.47 U. Camp. 12/078 169.5 1.61 10.51 0.43 -28.81 U. Camp. 12/080 170.5 0.87 10.37 0.35 U. Camp. 12/081 171.5 1.00 10.03 0.42 -28.21 U. Camp. 12/083 172.5 0.62 10.32 0.38 U. Camp. 12/086 174.0 0.61 10.83 0.30 U. Camp. 12/089 175.5 0.64 10.79 0.43 U. Camp. 12/092 177.0 1.26 9.62 0.64 -27.85 U. Camp. 12/093 177.5 1.11 10.10 0.52 U. Camp. 12/100 181.0 2.70 8.40 1.55 -28.30 U. Camp. 12/104 183.0 1.62 10.91 0.63 -28.56 U. Camp. 12/107 184.5 1.36 10.27 0.87 U. Camp. 12/110 186.0 2.17 10.26 0.75 U. Camp. 12/113 187.5 2.24 10.48 0.73 -28.98 U. Camp. 12/116 189.0 2.97 10.07 0.82 U. Camp. 12/119 190.5 1.91 10.75 0.59 -29.37 U. Camp. 12/121 191.5 2.48 10.17 0.79 L. Camp. 12/122 192.0 2.93 10.05 0.80 L. Camp. 12/123 192.5 3.23 9.65 0.84 L. Camp. 12/125 193.5 2.74 9.75 0.85 L. Camp. 12/128 195.0 2.53 9.78 0.82 -28.72 L. Camp. 12/130 196.0 3.51 8.95 1.20 L. Camp. 12/132 197.0 3.35 9.70 0.98 L. Camp. 12/134 198.0 5.18 9.27 1.13 L. Camp. 12/136 199.0 3.11 10.05 0.81 L. Camp. 12/137 199.5 3.57 9.93 0.77 L. Camp. 12/140 201.0 5.69 9.91 1.19 -28.74 L. Camp. 12/142 202.0 4.00 10.73 L. Camp. 12/143 202.5 4.79 8.93 1.10 L. Camp. 12/145 203.5 4.90 8.70 1.26 -29.25 L. Camp. 12/147 204.5 4.54 7.85 2.74 L. Camp. 12/148 205.0 5.58 7.88 1.37 L. Camp. 12/149 205.5 3.98 9.76 0.85 L. Camp. 12/150 206.0 5.35 8.78 1.02 L. Camp.

- 139 -

Appendix

12/151 206.5 4.23 7.99 1.15 L. Camp. 12/152 207.0 3.59 8.64 1.09 -29.01 L. Camp. 12/154 208.0 3.41 9.18 0.75 L. Camp. 12/156 209.0 2.40 9.92 0.59 -28.63 L. Camp. 12/158 210.0 3.61 9.71 0.82 L. Camp. 12/160 211.0 4.77 7.99 1.12 -28.74 L. Camp. 12/163 213.5 2.99 8.56 0.70 -28.49 L. Camp. 12/766 214.5 0.04 11.70 0.05 L. Camp. 12/768 216.0 0.04 12.29 0.03 L. Camp. 12/771 217.5 0.02 12.59 0.04 L. Camp. 12/775 219.5 1.11 11.38 0.27 -28.27 L. Camp. 12/777 220.5 0.44 11.36 0.13 L. Camp. 12/781 222.5 0.93 11.51 0.28 L. Camp. 12/785 224.5 0.69 11.85 0.18 L. Camp. 12/789 226.5 0.91 11.40 0.26 L. Camp. 12/791 227.5 1.49 11.42 0.28 -27.92 L. Camp. 12/794 229.0 0.24 12.15 0.12 L. Camp. 12/797 230.5 0.21 12.25 0.09 L. Camp. 12/801 232.5 0.55 11.92 0.21 -28.27 L. Camp. 12/805 234.5 0.50 11.82 0.20 L. Camp. 12/809 236.5 1.44 11.46 0.34 L. Camp. 12/813 238.5 0.51 11.55 0.18 L. Camp. 12/818 241.0 0.87 11.52 0.18 -28.20 L. Camp. 12/820 242.0 1.04 11.33 0.31 L. Camp. 12/822 243.0 1.09 11.05 0.44 L. Camp. 12/824 244.0 1.12 11.15 0.35 -28.53 L. Camp. 12/826 245.0 1.07 11.16 0.26 L. Camp. 12/828 246.0 0.79 11.18 0.29 L. Camp. 12/830 247.0 1.24 10.99 0.30 L. Camp. 12/832 248.0 0.97 11.31 0.28 L. Camp. 12/834 249.0 1.67 10.69 0.52 L. Camp. 12/837 249.5 1.00 10.90 0.45 L. Camp. 12/840 251.0 1.40 11.44 0.42 -29.10 L. Camp. 12/842 252.0 1.23 10.97 0.45 U. Sant. 12/844 253.0 1.30 10.86 0.54 U. Sant. 12/846 254.0 2.07 9.78 1.12 -28.05 U. Sant. 12/847 254.5 1.85 9.79 0.75 U. Sant. 12/850 256.0 0.73 11.06 0.49 U. Sant. 12/852 257.0 2.24 9.56 1.14 U. Sant. 12/854 258.0 1.01 10.71 0.55 -27.63 U. Sant. 12/856 259.0 0.06 10.77 0.08 U. Sant. 12/859 261.0 0.08 12.75 0.07 Turonian 12/863 263.0 0.13 12.75 0.10 Turonian 12/867 265.0 0.12 12.63 0.11 Turonian 12/871 267.0 0.14 12.65 0.10 -25.63 Turonian 12/875 269.0 0.14 12.65 0.08 Turonian 12/879 271.0 0.28 12.58 0.14 -26.41 Turonian 12/883 273.0 0.22 12.66 0.15 -25.73 Turonian 12/887 275.0 0.02 12.74 0.04 Turonian 12/891 277.0 0.07 12.75 0.03 Turonian 12/895 279.0 0.06 12.75 0.04 Turonian 12/903 283.0 0.03 12.70 0.05 Turonian 12/909 285.5 0.02 12.76 0.03 Turonian 12/913 287.5 0.01 12.74 0.00 Turonian 12/917 289.5 0.03 12.55 0.09 Turonian

- 140 -

Appendix

12/921 291.5 0.01 12.76 Turonian 12/925 293.5 0.02 12.75 0.01 Turonian

- 141 -

Appendix

Appendix 2. XRF data for Upper Cretaceous rocks in north Lebanon (see chapter II).

Pb Ba La Sb Mo W Se Nb Zr Y Sr Rb As Ga Zn Cu Sn Ni Cr V Co Fe Samples ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm % 12/018 < 10 334 17 <10 < 20 < 20 < 20 < 10 < 20 < 20 1127 < 10 < 10 < 10 74 35 < 20 48 64 25 < 20 0.51 12/042 <10 160 26 <10 < 20 < 20 <10 <10 < 20 25 1156 <10 <10 <10 82 44 < 20 58 60 26 < 20 0.40 12/081 <10 309 20 <10 < 20 < 20 <10 <10 < 20 < 20 1144 10 <10 <10 55 26 < 20 34 49 22 < 20 0.46 12/119 <10 76 22 <10 < 20 < 20 <10 <10 < 20 < 20 1124 <10 <10 <10 73 38 < 20 56 47 21 < 20 0.36 12/134 <10 86 24 <10 < 20 < 20 <10 <10 26 23 978 <10 <10 <10 124 52 < 20 88 75 46 < 20 0.49 12/147 <10 46010* <10 65 < 20 < 20 <10 <10 < 20 < 20 1114 <10 11 <10 117 46 47 88 115 201 < 20 0.64 12/158 <10 175 20 <10 < 20 < 20 <10 <10 < 20 < 20 1024 <10 <10 <10 103 35 < 20 61 75 28 < 20 0.37 12/166 <10 89 19 <10 < 20 < 20 <10 <10 < 20 < 20 1018 <10 <10 <10 30 12 < 20 25 32 < 20 < 20 0.35 12/189 <10 385 16 <10 < 20 < 20 <10 <10 < 20 22 1066 <10 <10 <10 37 18 < 20 23 36 < 20 < 20 0.33 12/264 <10 331 16 <10 < 20 < 20 <10 <10 < 20 < 20 1073 <10 <10 <10 84 36 < 20 55 48 25 < 20 0.33 12/764 <10 60 19 <10 < 20 < 20 <10 <10 < 20 < 20 963 <10 <10 <10 24 <10 < 20 <10 23 < 20 < 20 0.23 12/777 <10 752 <10 <10 < 20 < 20 <10 <10 < 20 < 20 811 <10 <10 <10 25 15 < 20 22 < 20 < 20 < 20 0.12 12/828 <10 908 15 <10 < 20 < 20 <10 < 10 < 20 < 20 963 <10 <10 <10 42 24 < 20 21 34 < 20 < 20 0.27 12/856 <10 < 50 17 <10 < 20 < 20 <10 < 10 < 20 < 20 250 <10 <10 <10 20 <10 < 20 25 < 20 < 20 < 20 0.05 12/903 <10 < 50 13 <10 < 20 < 20 <10 < 10 < 20 < 20 186 <10 <10 <10 <10 <10 < 20 21 < 20 < 20 < 20 0.03 *> calibrated range

Appendix 3. XRF data for Upper Cretaceous rocks in south Lebanon (see chapter III).

Pb Ba La Ce Hf W Sb Mo Se Nb Zr Y Sr Rb As Ga Zn Cu Ag Sn Ni Cr V Fe Co Samples ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm ppm % ppm 5 <10 475 19 < 50 <20 <20 <10 <10 <10 <10 <10 24 1467 <10 <10 <10 82 37 <10 <10 45 69 26 0.58 <20 14 <10 47 19 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1613 <10 <10 <10 75 22 <10 <10 35 52 22 0.14 <20 16 <10 67 23 < 50 <20 <20 <10 <10 <10 <10 24 <10 1697 <10 <10 <10 98 25 <10 <10 31 62 25 0.18 <20 17 <10 54 22 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1631 <10 <10 <10 94 25 <10 <10 65 62 23 0.13 <20 21 <10 56 21 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1560 <10 <10 <10 85 28 <10 <10 25 77 31 0.22 <20 22 <10 45 20 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1415 <10 <10 <10 94 25 <10 <10 68 62 25 0.22 <20 30 <10 155 19 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1210 <10 10 <10 159 41 <10 <10 103 104 < 20 0.25 <20 33 <10 44 14 < 50 <20 <20 <10 <10 <10 <10 <10 <10 963 <10 <10 <10 97 22 <10 <10 67 58 31 0.11 <20 34 <10 68 24 < 50 <20 <20 <10 <10 <10 <10 <10 25 1374 <10 10 <10 132 32 <10 <10 90 109 31 0.31 <20 42 <10 59 23 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1463 <10 <10 <10 141 35 <10 <10 95 105 36 0.28 <20 44 <10 51 23 < 50 <20 <20 <10 <10 <10 <10 25 <10 1214 <10 <10 <10 92 26 <10 <10 82 154 56 0.70 <20 51 <10 42 16 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1563 <10 <10 <10 52 13 <10 <10 44 66 23 0.35 <20 59 <10 111 21 < 50 <20 <20 <10 27 10 <10 <10 25 1389 <10 11 <10 197 50 <10 <10 116 103 25 0.26 <20 60 <10 55 24 < 50 <20 <20 <10 <10 <10 <10 26 22 1226 <10 11 <10 138 34 <10 <10 92 102 45 0.33 <20 61 <10 92 18 < 50 <20 <20 <10 <10 <10 <10 <10 <10 1008 <10 12 <10 144 42 <10 <10 112 111 30 0.33 <20 64 <10 1165* 14 < 50 <20 <20 <10 <10 <10 <10 <10 23 1290 <10 <10 <10 79 39 <10 <10 62 61 < 20 0.44 <20 65 <10 55 27 < 50 <20 <20 <10 <10 <10 <10 26 23 1205 <10 10 <10 116 36 <10 <10 93 110 51 0.50 <20 *> calibrated range

- 142 -

”يا راكباً زم َن الليمو ْن ح ّول، وك َل ما نهش َك القر ُن العشرون ب ِّدّل،

وتنص ْت إلى هدي ِّل الموجِّ في إزدحا ِّم المتوس ِّط ع ّجل، إني أرس ُل عب َر جها ِّز الالسلكي بعضاً إل َّي وبعضاً إللكي.“ (أُسامة)

From “Bennesbeh Labokra Shou?” (1978)

By Ziad Rahbani

- 143 -

“If you need to invoke your academic pedigree or job title for people to believe what you say,

then you need a better argument”

Neil deGrasse Tyson

- 144 -

SAMER BOU DAHER Date of birth: 14.08.88 Address: Templergraben 40, 52062 Aachen, Germany Mobile: +49 176 31294103 Email: [email protected]

EDUCATION Feb 12 - RWTH Aachen University, Germany expected March Ph.D. - Source rock characterization and petroleum generation modelling of the Levant 2016 Basin, onshore-offshore Lebanon: An integrated approach. German Academic Exchange Service (DAAD) 3 years Scholarship.

Jan 10 - Jan 12 Lund University, Sweden M.Sc. - Geology (with Distinction). ERASMUS MUNDUS 2 years Scholarship.

Sep 06 - Jun 09 American University of Beirut (AUB), Lebanon B.Sc. - Major in Geology, Minor in Business Administration

EXPERIENCE Feb 12 - RWTH Aachen University, Germany - Ph.D. Applied modern organic geochemistry and associated analytical techniques on a large number of source rock samples from onshore Lebanon in order to evaluate their poten- tial and estimate their offshore distribution into the Levant Basin. Integrated the analytical results in a 3D petroleum system model assessing the thermogenic hydrocarbon poten- tial of the Levant Basin and the effect of organofacies changes on the petroleum sys- tems. Analytical work included elemental analyses, Rock-Eval pyrolysis, organic petrology (maceral analyses and vitrinite reflectance), organic carbon isotopes and biomarker analyses (GC-FID, GC-MS, Cu-Py-GC-MS). Oct 15 - RWTH Aachen University, Germany - Teaching assistant Participated in the teaching of Petroleum System Modelling, Sedimentary Basin Dynam- ics, and Exploration/Production courses at the Institute of Geology and Geochemistry of Petroleum and Coal. Dec 13 - Apr 14 IFP Energies nouvelles, France - Intern Constructed a 3D petroleum system model of the Levant Basin and Margin, integrating forward stratigraphic modelling results as an approach to overcome the lack of data in a frontier basin. Performed kinetic experiments on a selection of samples using open sys- tem pyrolysis. Apr 11 - Jan 12 Lund University / Swedish Geological Survey (SGU), Sweden - Master thesis (ranked 1st) Project title: “Lithofacies analysis and heterogeneity study of the subsurface Rhaetian- Pliensbachian sequence in SW Skåne (Southern Sweden)”. The objective of the thesis was to evaluate the relative frequency of Rhaetian- Pliensbachian sandstone units in the Swedish part of the Danish basin and understand their lateral distribution within a sequence stratigraphic framework, for purposes of geo- thermal energy production as well as CO2 storage. The project included wireline logs in- terpretation, cores and cuttings description, and field investigation.

Oct 09 - Dec 09 Advanced Construction Technology Services (ACTS), Lebanon Participated in the logging of core samples, and supervised the core drilling in the Port of Beirut extension project.

AWARDS May 12 Received an award from the Swedish Geological Survey for the best master thesis in 2011, Sweden. PUBLICATIONS Bou Daher, S., Nader, F.H., Strauss, H., Littke, R., 2014. Depositional environment and source- rock characterisation of organic- matter rich Upper Turonian- Upper Campa- nian carbonates, Northern Lebanon. Journal of Petroleum Geology, 37, 1-20.* Bou Daher, S., Nader, F.H., Müller, C., Littke, R., 2015. Geochemical and petrograph- ic characterization of Campanian – Lower Maastrichtian calcareous petroleum source rocks of Hasbayya, South Lebanon. Marine and Petroleum Geology, 64, 304-323.* Bou Daher, S., Ducros, M., Michel, P., Hawie, N., Nader, F.H., Littke, R., in press. 3D thermal history and maturity modelling of the Levant Basin and its eastern margin, offshore -onshore Lebanon.* *Articles and associated work have been presented at GeoBerlin 2015, AAPG European regional confer- ence 2013 & 2015 (Barcelona & Lisbon), EGU 2015 (Vienna), AAPG GTW 2013 (Beirut), DGMK 2013 (Celle), GV-Sediment 2012 & 2013 (Hamburg & Tübingen), IMOG 2013 (Teneriffe), ILP 2013 (Marseille), AAPG-LIPE Northern Arabia Geoscience Conference and Exhibition 2014 (Beirut).

EXTRA-CURRICULAR ACTIVITIES  Member of Geländefahrrad Aachen e.V. (Mountain bike club Aachen)  Member of the American Association of Petroleum Geologists (AAPG) since 2008  Mentor for international students at Lund University (Sweden, Jan 11 - Jun 11)  Vice president of the Studierådet för Naturgeografer och Geologer (SNG) at Lund University (Study council for Geologists and Physical Geographers) (Sweden, Sep 10 - Jan 11)  President (2008) and vice president (2007) of the AUB Geology Student Society (Lebanon)  Vice president (2008) and treasurer (2007) of the AUB Camping and Hiking Club (Lebanon)

LANGUAGES English and Arabic, fluent; French, good; German, basic.

References available upon request