17.01.2015 Views

The Levant Basin Offshore Israel: Stratigraphy, Structure, Tectonic ...

The Levant Basin Offshore Israel: Stratigraphy, Structure, Tectonic ...

The Levant Basin Offshore Israel: Stratigraphy, Structure, Tectonic ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>, <strong>Tectonic</strong> Evolution<br />

and Implications for<br />

Hydrocarbon Exploration<br />

Michael Gardosh, Yehezkel Druckman,<br />

Binyamin Buchbinder and Michael Rybakov<br />

Revised Edition<br />

GSI/4/2008<br />

GII 429/328/08<br />

April 2008


<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>, <strong>Tectonic</strong> Evolution<br />

and Implications for<br />

Hydrocarbon Exploration<br />

Michael Gardosh 1 , Yehezkel Druckman 2 ,<br />

Binyamin Buchbinder 2 and Michael Rybakov 1<br />

1<br />

- Geophysical Institute of <strong>Israel</strong>, P.O.B. 182, Lod 71100, <strong>Israel</strong>, miki@gii.co.il<br />

2 - Geological Survey of <strong>Israel</strong>, 30 Malkhe <strong>Israel</strong> St., Jerusalem 95501, <strong>Israel</strong>,<br />

ydruckman@gsi.gov.il, b.buchbinder@gsi.gov.il<br />

Prepared for the Petroleum Commissioner,<br />

Ministry of Infrastructure<br />

Revised Edition<br />

GSI/4/2008<br />

GII 429/328/08<br />

April 2008


Abstract<br />

Contents<br />

Page<br />

1. Introduction…………………………………………………………………………...... 5<br />

2. <strong>Tectonic</strong> Setting………………………………………………………………………… 9<br />

3. Data Sets………………………………………………………………………………… 13<br />

3.1 Well data……………………………………………………………………………. 13<br />

3.2 Seismic data…………………………………………………………………………. 14<br />

3.3 Gravity and magnetic data sets…………………………………………………….... 15<br />

4. <strong>Stratigraphy</strong>…………………………………………………………………………….. 15<br />

4.1 General……………………………………………………………............................ 15<br />

4.2 <strong>The</strong> Permo-Triassic section………………………………………………………..... 16<br />

4.3 <strong>The</strong> Jurassic section………………………………………………………………..... 21<br />

4.4 <strong>The</strong> Cretaceous section……………………………………………………………… 29<br />

4.5 <strong>The</strong> Tertiary section………………………………………………...……………….. 35<br />

5. Interpretation of Seismic Data………………………………………………………..... 43<br />

5.1 Seismic horizons…………………………………………………………………….. 43<br />

5.2 Time and depth maps………………………………………………………………… 58<br />

6. Interpretation of Gravity and Magnetic Data…………………………………………. 67<br />

6.1 Gravity maps…………………………………………………………………………. 67<br />

6.2 Magnetic maps……………………………………………………………………..... 70<br />

6.3 Gravity and magnetic modeling (2D)………………………………………………… 71<br />

6.4 Discussion…………………………………………………………………………….. 73<br />

7. <strong>Tectonic</strong> Evolution of the <strong>Levant</strong> <strong>Basin</strong> and Margin………………............................. 74<br />

7.1 Rifting stage……………………………………………………….............................. 74<br />

7.2 Post-rift, passive margin stage…………………………………….............................. 80<br />

7.3 Convergence stage…………………………………………………………………… 83<br />

8. Erosion and Mass Transport on the <strong>Levant</strong> Slope……………………………………. 89<br />

8.1 Rifting and passive margin stage…………………………………………………….. 89<br />

8.2 Convergence stage…………………………………………………………………… 94<br />

9. Hydrocarbon Potential…………………………………………………………………. 100<br />

9.1 Trap typs and hydrocarbon plays…………………………………............................. 100<br />

9.2 Source rocks and petroleum systems………………………………………………… 103<br />

10. Summary……………………………………………………………………………….. 105<br />

Acknowledgments……………………………………………………………………… 107<br />

References……………………………………………………………………………….. 108


List of Figures<br />

Page<br />

Fig. 1.1 Main physiographic and tectonic elements of the <strong>Levant</strong> region……………………...6<br />

Fig. 1.2 <strong>Offshore</strong> drilling history and main hydrocarbon occurrences…………………………7<br />

Fig. 1.3 Location of seismic lines, wells, and stratigraphic cross-sections……………………..8<br />

Fig. 2.1 Schematic section across the <strong>Levant</strong> <strong>Basin</strong> and margin, from the Eratosthenes<br />

Seamount to the Dead Sea Rift………………………………………………………...10<br />

Fig. 4.1 Triassic stratigraphy and marine onlaps, <strong>Israel</strong>………………………………………..18<br />

Fig. 4.2 Jurassic stratigraphy and marine onlaps, <strong>Israel</strong>………………………………………..22<br />

Fig. 4.3 Cretaceous stratigraphy and marine onlaps, <strong>Israel</strong>…………………………………….30<br />

Fig. 4.4 Tertiary stratigraphy and marine onlaps, <strong>Israel</strong>………………………………………..36<br />

Fig. 4.5 SE-NW cross-section from the Gevim-1 to the Yam West-1 well…………………….37<br />

Fig. 4.6 N-S cross-section from the Asher Yam-1 to the Yam West-1 well……………………38<br />

Fig. 5.1 Interpreted seismic profile showing the correlation of seismic horizons to the<br />

Yam-2 and Yam-West-1 wells…………………………………………………………44<br />

Fig. 5.2 Synthetic seismograms of the Yam West-1 and Yam Yafo-1 wells…………………...45<br />

Fig. 5.3 Interpreted, E-W oriented seismic profile through the Yam-2 and<br />

Yam West-1 wells………………………………………………………………...........47<br />

Fig. 5.4 Interpreted, composite seismic profile, generally oriented in a N-S direction along<br />

the shallow part of the basin…………………………………………………………...48<br />

Fig. 5.5 Interpreted seismic profile, oriented in a NE-SW direction showing the correlation<br />

of seismic horizons in the deep part of the basin……………………………………….49<br />

Fig. 5.6 Interpreted seismic profile showing the transition area from the elevated<br />

Mesozoic thrust belt to the deep Tertiary basin………………………………………..51<br />

Fig. 5.7 Interpreted composite seismic profile through the Hannah-1 well……………………..52<br />

Fig. 5.8 Interpreted, NE-SW regional seismic profile showing extensional and contractional<br />

structures in the central part of the basin…...…………………………………………...53<br />

Fig. 5.9 Interpreted NW-SE seismic profile from the Eratosthenes Seamount to the<br />

<strong>Levant</strong> margin……….…………………………………………………………………55<br />

Fig. 5.10 Interpreted E-W seismic profile showing extensional and contractional structures<br />

in the northeastern part of the basin…………………………………………………..56<br />

Fig. 5.11 Interpreted SW-NE seismic profile showing the southern part of the Jonah Ridge…..57<br />

Fig. 5.12 Depth map of the near top basement, Brown horizon………………………………....59<br />

Fig. 5.13 Depth map of the Middle Jurassic, Blue horizon……………………………………...60<br />

Fig. 5.14 Depth map of the Middle Cretaceous, Green horizon…………………………………61


Fig. 5.15 Depth map of the base Oligocene, Red horizon……………………………………...62<br />

Fig. 5.16 Depth map of the top Lower Miocene, Cyan horizon………………………………..63<br />

Fig. 5.17 Depth map of the base Messinian, Purple horizon…………………………………...64<br />

Fig. 5.18 Depth map of the top Messinian, Violet horizon……………………………………..65<br />

Fig. 5.19 Depth map of the sea floor……………………………………………………………66<br />

Fig. 6.1 Revised gravity anomaly maps of the <strong>Levant</strong> <strong>Basin</strong>………… ………………..……..68<br />

Fig. 6.2 Revised magnetic anomaly maps of the <strong>Levant</strong> <strong>Basin</strong>………………………………....69<br />

Fig. 6.3 2D Gravity and magnetic modeling of the <strong>Levant</strong> <strong>Basin</strong>………………………………72<br />

Fig. 7.1 Neotethyan rifting stage - inferred Triassic to Early Jurassic horst and graben<br />

system of the <strong>Levant</strong> region……………………………………………………………75<br />

Fig. 7.2 Neotethyan rifting stage - Early Mesozoic extensional structures in the<br />

<strong>Levant</strong> <strong>Basin</strong> (seismic examples)...……………………………………………………77<br />

Fig. 7.3 Passive margin stage- isopach map and geologic section through the <strong>Levant</strong> margin....81<br />

Fig. 7.4 Convergence stage - Syrian Arc fold structures of the <strong>Levant</strong> region………………….84<br />

Fig. 7.5 Convergence stage - contractional structures in the <strong>Levant</strong> <strong>Basin</strong> (seismic examples)...86<br />

Fig. 8.1 Convergence stage- the Tertiary fill of the <strong>Levant</strong> <strong>Basin</strong>...............................................91<br />

Fig. 8.2 Tertiary submarine canyon and channel system of the <strong>Levant</strong> margin<br />

(seismic examples)…………………………………………………..………..…….….93<br />

Fig. 8.3 Tertiary submarine incision and deposition of clastic sediments in the <strong>Levant</strong> <strong>Basin</strong>…97<br />

Fig. 9.1 Schematic section through the <strong>Levant</strong> <strong>Basin</strong> showing potential hydrocarbon<br />

traps in the Phanerozoic fill…………………………………………….…………….101<br />

List of Tables<br />

Page<br />

Table 3.1- Summary of well data used in the present study……………………………………..14<br />

Table 4.1- Summary of Jurassic formation tops and thicknesses in offshore wells…...………...29<br />

Table 5.1- Correlation of interpreted seismic horizons to chrono and lithostratigraphic<br />

unit tops……..……………………………………………………………………..….46


Abstract<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> is a deep and long existing geologic structure located in the eastern<br />

Mediterranean Sea. <strong>The</strong> southern part of the basin hosts a world-class hydrocarbon province<br />

offshore the Nile Delta. Recent discoveries of biogenic gas and various oil shows indicate that<br />

the central part of the basin, offshore <strong>Israel</strong> has significant hydrocarbon potential. To further<br />

investigate this area, which is generally under-explored, a basin-scale study was initiated by the<br />

petroleum commissioner in the <strong>Israel</strong>i Ministry of Infrastructure. In this study new geophysical<br />

data is integrated with regional and local geologic data to reconstruct the basin history and to<br />

identify favorable hydrocarbon plays.<br />

<strong>The</strong> reconstructed basin history shows that the <strong>Levant</strong> <strong>Basin</strong> was shaped in several main<br />

tectonic stages. Early Mesozoic rifting resulted in the formation of an extensive graben and<br />

horst system, extending throughout the <strong>Levant</strong> onshore and offshore. Late Jurassic to Middle<br />

Cretaceous, post-rift subsidence was followed by the formation of a deep marine basin in the<br />

present-day offshore and a shallow-marine, carbonate dominated margin and shelf near the<br />

Mediterranean coastline and further inland. A Late Cretaceous and Tertiary convergence phase<br />

resulted in inversion of Early Mesozoic structures and the formation of extensive, Syrian Arctype<br />

contractional structures throughout the <strong>Levant</strong> <strong>Basin</strong> and margin. <strong>The</strong> Tertiary convergence<br />

was further associated with uplift, widespread erosion, slope incision and basinward sediment<br />

transport. A Late Tertiary desiccation of the Mediterranean Sea was followed by deposition of a<br />

thick evaporitic blanket that was later covered by a Plio-Pleistocene siliciclastic wedge. <strong>The</strong><br />

Phanerozoic basin-fill ranges in thickness from 5-6 km on the margin to more than 15 km in the<br />

central part of the basin<br />

A variety of potential, structural and stratigraphic traps were formed in the <strong>Levant</strong> <strong>Basin</strong><br />

during the three main tectonic stages. Possible hydrocarbon play types are: Triassic and Lower<br />

Jurassic fault-controlled highs and rift-related traps; Middle Jurassic shallow-marine reservoirs;<br />

Lower Cretaceous deepwater siliciclastics; the Tertiary canyon and channel system and<br />

associated deepwater siliciclastics found in either confined or non-confined settings; Upper<br />

Cretaceous and Tertiary Syrian Arc folds; and Mesozoic and Cenozoic isolated, carbonate<br />

buildups on structurally elevated blocks.<br />

Shallow gas discoveries in Pliocene sands and high-grade oil shows found in the<br />

Mesozoic section indicate the presence of source rocks and appropriate conditions for<br />

hydrocarbon generation in both biogenic and thermogenic petroleum systems. <strong>The</strong> size, depth<br />

and trapping potential of the <strong>Levant</strong> <strong>Basin</strong> suggest that large quantities of hydrocarbons can be<br />

found offshore <strong>Israel</strong>.


1. Introduction<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> occupies the eastern Mediterranean Sea (Fig. 1.1). Its southern part,<br />

offshore Egypt hosts the prolific Nile Delta hydrocarbon province where extensive exploration<br />

and production activity is taking place. <strong>The</strong> increase in demand and rising price of oil in recent<br />

years has promoted a growing interest in the less known and under-explored, central and<br />

northern parts of the basin, offshore <strong>Israel</strong>, Lebanon and Cyprus.<br />

Petroleum exploration in the eastern Mediterranean has a long history and has so far yielded<br />

modest success. <strong>Offshore</strong> exploration started in the late 1960's and early 1970's with a series of<br />

wells drilled by Belpetco (Fig. 1.2). <strong>The</strong> seven Belpetco wells targeted structural culminations<br />

on the shallow shelf of <strong>Israel</strong> and northern Sinai, but all were found dry. <strong>The</strong>se early wells,<br />

however, provided important information and established the initial geologic model of the<br />

eastern Mediterranean as a deep marine basin that was formed during the Early Mesozoic, and<br />

continued to develop west of an extensive shallow-marine shelf.<br />

<strong>The</strong> next exploration campaign during the mid 1970's to mid 1980's resulted in more<br />

success. Several wells were drilled on shallow structures offshore Sinai (Fig. 1.2). Light oil was<br />

found in Early Cretaceous sandstone in the Ziv-1 (drilled by Oil Exploration Limited) and the<br />

Mango-1 (drilled by Total) wells. Mango-1 initially produced 41-50 0 API gravity oil at a rate of<br />

about 10,000 Bbl/day; however, performance rapidly decreased and production was stopped.<br />

A series of four wells were drilled offshore <strong>Israel</strong> by Isramco during the late 1980's to<br />

late 1990's (Fig. 1.2). <strong>The</strong>se boreholes that targeted structural culminations were typically<br />

deeper than earlier wells and most of them reached a Middle Jurassic stratigraphic level at depths<br />

of more than 5000 m. All the Isramco wells encountered various oil and gas shows. Yam-2 and<br />

Yam Yafo-1 tested 44-47 0 API gravity oil at a rate of 500-800 Bbl/day from Middle Jurassic<br />

limestone, although no commercial production was established.<br />

Exploration activity in the <strong>Levant</strong> <strong>Basin</strong> underwent a significant resurgence since 1999-<br />

2000 when several, large gas fields were discovered at a shallow depth within Pliocene sands<br />

west of the towns of Ashqelon and Gaza (Fig. 1.2). Gas reserves in the Noa/MariB/Gaza Marine<br />

fields are estimated at about 3 TCF. Gas is presently produced from the MariB field and is<br />

transported onshore to various <strong>Israel</strong>i consumers.<br />

<strong>The</strong> successful exploration campaign of the early 2000's promoted the acquisition of<br />

geophysical data throughout the entire eastern Mediterranean area. <strong>The</strong> new geophysical data<br />

sets include about 12,000 km of regional 2D seismic reflection lines (Fig. 1.3) and several, high<br />

resolution 2D and 3D surveys, as well as new gravity and magnetic measurements.<br />

<strong>The</strong> petroleum commissioner in the <strong>Israel</strong>i Ministry of Infrastructure initiated a regional<br />

analysis of the <strong>Levant</strong> <strong>Basin</strong> and margin. <strong>The</strong> study was carried out by professional teams from<br />

5


Eurasian Plate<br />

Cyprian Arc<br />

Palmyra<br />

Eratosthenes<br />

Seamount<br />

A<br />

Mediterranean<br />

Sea<br />

<strong>Levant</strong> <strong>Basin</strong><br />

Study<br />

area<br />

Dead<br />

Sea<br />

Nile Delta<br />

Negev<br />

B<br />

Western<br />

Desert<br />

African<br />

Plate<br />

Sinai<br />

200 km<br />

- Main Fault<br />

Zone<br />

Fig. 1.1- Main physiographic and tectonic elements of the <strong>Levant</strong> region. A-B indicates the<br />

location of the schematic cross-section in Fig. 2.1.<br />

6


Mediterranean Sea<br />

Asher Yam-1<br />

Qishon Yam-1<br />

Foxtrot-1<br />

Drilling year<br />

1969-71<br />

1976-77<br />

~1985<br />

1989-99<br />

Delta-1<br />

Item-1<br />

Romi-1<br />

2000-03<br />

1999-2003<br />

Hana-1<br />

Yam Yafo-1<br />

<strong>Israel</strong><br />

Gas Field<br />

Oil Field<br />

Noa-Or<br />

Yam W.-2<br />

Yam W.-1<br />

Joshua-1<br />

Echo-1<br />

Yam-1/2<br />

Bravo-1<br />

Mari<br />

Gaza<br />

Marine<br />

Nir<br />

Helez-Kokhav<br />

Mango-1<br />

Dag-1<br />

Aneb-1<br />

Til-1<br />

Gaza<br />

Ziv-1<br />

Gal-1<br />

Sinai<br />

Sadot<br />

Fig. 1.2- <strong>Offshore</strong> drilling history and main hydrocarbon occurrences. <strong>The</strong> Noa-Or, Marie,<br />

Nir and Gaza Marine fields are recent gas discoveries in Pliocene deepwater sands.<br />

7


the Geophysical Institute of <strong>Israel</strong> and Geological Survey of <strong>Israel</strong>. Its main emphasis was to<br />

integrate the vast amount of geological information from the <strong>Levant</strong> margin onshore with the<br />

high-quality geophysical data sets that were recently acquired in the basin offshore. <strong>The</strong> final<br />

results include depth maps of key stratigraphic horizons, stratigraphic cross-sections, and<br />

regional tectonic and paleogeographic maps. <strong>The</strong> results enabled the reconstruction of a regional<br />

geologic framework for the two areas, traditionally studied in different scales and techniques.<br />

2. <strong>Tectonic</strong> Setting<br />

Wide-angle deep refraction profiles show significant variations in seismic velocities and<br />

depth to the Moho from the inland part of <strong>Israel</strong> on the east to the Mediterranean Sea on the west<br />

(Makris et al., 1983; Ginzburg and Ben-Avraham, 1987, Ben-Avraham et al., 2002). <strong>The</strong><br />

velocity variations (summarized in Fig. 2.1) are associated with a change from a ~30 km thick,<br />

low velocity continental-type crust inland, to a ~10 km thick oceanic- or intermediate-type crust<br />

offshore (Garfunkel, 1998, Ben-Avraham et al, 2002; Gardosh and Druckman, 2006). This<br />

fundamental geophysical observation and its interpretation are milestones in the research of the<br />

<strong>Levant</strong> area. <strong>The</strong>y unequivocally demonstrate that the present-day marine basin is a deeply<br />

rooted geologic structure associated with large-scale plate motions.<br />

<strong>The</strong> Paleozoic section found in outcrops and wells in southern and central <strong>Israel</strong><br />

comprises continental to shallow marine deposits. Evidence from adjacent countries further<br />

indicates that during most of the Paleozoic the <strong>Levant</strong> region formed part of the super-continent<br />

Gondwana and was located inland, several hundred kilometers south of the Paleotethys Ocean<br />

(Garfunkel, 1998, Weissbrod, 2005). <strong>The</strong> Paleozoic landscape was dominated by a series of<br />

region-wide swells and intervening basins that were hundreds of kilometers in diameter<br />

(Weissbrod, 2005). <strong>The</strong> area of the eastern Mediterranean formed part of this landscape and no<br />

deep marine basin was in existence at that time.<br />

<strong>The</strong> Late Permian, Triassic and Early Jurassic section of the <strong>Levant</strong> region is dominated<br />

by shallow-marine carbonate and occasionally siliciclastic rock. Significant thickness variations<br />

in the Triassic and Early Jurassic strata are identified in well and seismic data from southern and<br />

central <strong>Israel</strong> (Fig. 2.1) (Goldberg and Friedman, 1974; Druckman, 1974, 1977; Freund et al.,<br />

1975; Druckman, 1984; Bruner, 1991; Druckman et al., 1995a; Garfunkel, 1998, Gelberman,<br />

1995). <strong>The</strong> Palmyra trough in central Syria (Fig. 1.1) contains an unusually thick Permian to<br />

Triassic section. This range of phenomena indicates extensive vertical motions, continental<br />

9


km<br />

0<br />

10<br />

20<br />

Eratosthenes Seamount <strong>Levant</strong> <strong>Basin</strong><br />

Dead Sea<br />

<strong>Basin</strong><br />

A B<br />

Coastline<br />

Messinian Evap.<br />

Paleocene to Miocene<br />

Cret.-Jur.<br />

Cretaceous<br />

6 km/s<br />

Volcanic<br />

Oceanic/Intermediate<br />

crust <br />

6.7 km/s<br />

7 km/s<br />

8 km/s<br />

Plio-Pleistocene<br />

Dead Sea Transform<br />

10<br />

Upper Continental<br />

crust<br />

Continental<br />

crust<br />

Precambrian<br />

Trias-Perm<br />

6.15 km/s<br />

6.5 km/s<br />

10<br />

Jurassic and older<br />

20<br />

30<br />

75 km<br />

30<br />

Moho<br />

Moho<br />

Fig. 2.1- Schematic section across the <strong>Levant</strong> <strong>Basin</strong> and margin, from the Eratosthenes Seamount to the Dead Sea Rift. <strong>The</strong> section shows<br />

the main structural elements, stratigraphic units and crustal seismic velocities (after Garfunkel, 1998). See Fig. 1.1 for location.


eakup and rifting of the Gondwanian shallow-shelf (Freund et al., 1975; Garfunkel and Derin,<br />

1984, Garfunkel, 1998).<br />

A series of horst and graben structures (Fig. 2.1) evolved in several pulses during the<br />

uppermost Paleozoic and Early Mesozoic periods, accompanied by widespread magmatic<br />

activity. <strong>The</strong>se regional extensional pulses resulted in separation of the Tauride and<br />

Eratosthenes continental blocks from the Gondwanaland craton and the formation of a deep<br />

marine basin in the area of the present-day eastern Mediterranean (Bein and Gvirtzman,1977;<br />

Garfunkel and Derin, 1984; Garfunkel, 1998; Robertson, 1998) (Fig. 2.1). <strong>The</strong> basin was<br />

connected to the extensive Neotethys Ocean that developed at the same time further to the north<br />

and west.<br />

<strong>The</strong> Early Mesozoic rifting activity is considered by most authors to be the cause of the<br />

thinner crust identified within the <strong>Levant</strong> <strong>Basin</strong> (Fig. 2.1). <strong>The</strong> composition of the crust in the<br />

southeastern Mediterranean Sea is still under debate. Some researchers postulate an oceanic<br />

lithosphere that was formed when the initial rifting evolved into full scale sea-floor spreading<br />

(Makris et al., 1983; Ginzburg and Ben-Avraham, 1987; Ben-Avraham and Ginzburg, 1990).<br />

Others assume only stretched and intruded crust of continental origin that was formed during<br />

intra-plate rifting (Cohen et al., 1988, 1990; Hirsch et al., 1995; Gardosh and Druckman 2006).<br />

Rifting was followed by post-rift cooling and subsidence of the newly formed or<br />

modified lithosphere in the <strong>Levant</strong> <strong>Basin</strong> (Garfunkel and Derin, 1984; ten Brink 1987; Garfunkel<br />

1998). Starting from the Middle Jurassic a paleo-depositional hinge belt developed along the<br />

eastern margin of the basin. <strong>The</strong> hinge belt that was originally identified in deep wells along the<br />

present-day <strong>Israel</strong>i coastline (Gvirtzman and Klang, 1972; Bein and Gvirtzman ,1977) separated<br />

a vast shallow-marine platform on the east from a deep marine basin on the west. Several<br />

studies suggest that the Mesozoic shelf-edge extended near the entire southeastern Mediterranean<br />

coastline from northern Egypt to western Lebanon (Figs. 1.1, 2.1) (Walley, 1998, 2001).<br />

<strong>The</strong> Late Cretaceous to Early Tertiary section is dominated by hemi-pelagic to pelagic<br />

sedimentation, marking the termination of the extensive, shallow-marine carbonate deposition on<br />

the <strong>Levant</strong> margin. <strong>The</strong> lithologic transition reflects a major drowning event and ecological<br />

changes associated with plate re-arrangement and the beginning of the closure of the Neotethys<br />

Ocean (Sass and Bein, 1982; Almogi-Labin et al., 1993).<br />

<strong>The</strong> Late Cretaceous convergence of the Eurasian and Afro-Arabian plates initiated a<br />

northward-dipping subduction zone in the southern Neotethys Ocean, in the present-day areas of<br />

Cyprus and southern Turkey (Fig. 1.1) (Robertson, 1998). Far-field stresses related to the onset<br />

of subduction may have caused compression of the <strong>Levant</strong> passive margin and triggered the<br />

formation of anticlines and synclines throughout the <strong>Levant</strong> region, known as the 'Syrian Arc'<br />

11


(Krenkel, 1924) or '<strong>Levant</strong>ide' (Picard, 1959) fold belt. Well and seismic data suggest that the<br />

Syrian Arc deformation is generally associated with reactivation of Early Mesozoic normal faults<br />

in a reverse motion (Fig. 2.1) (Freund et al., 1975; Druckman et al., 1995). Evidence from <strong>Israel</strong><br />

and Lebanon further indicates that the Syrian Arc deformation that started in the Late Cretaceous<br />

continued through the Tertiary (Eyal and Reches, 1983; Walley, 1998, 2001).<br />

<strong>The</strong> continued plate convergence during the Oligocene and Early Miocene resulted in the<br />

formation of a subduction zone along the present Cypraen Arc plate boundary (Fig.<br />

1.1)(Robertson, 1998). In the <strong>Levant</strong> region plate motion was associated with uplifting east of<br />

the <strong>Levant</strong> coast and updoming of the Arabian Shield south of it. This activity is reflected by a<br />

regional unconformity surface that is identified at the upper part of the Eocene strata, which<br />

separates the older carbonate section from the overlying Oligo-Miocene siliciclastics of the<br />

Saqyie Group (Picard, 1943; Ball and Ball, 1953). <strong>The</strong> deposition of Oligo-Miocene, coarsegrained<br />

clastic units is associated with widespread erosion and basinward transport of sediments<br />

through an extensive subaerial and submarine drainage system that developed on the <strong>Levant</strong><br />

slope (Gvirtzman and Buchbinder, 1978; Druckman et al., 1995b).<br />

A thick, Upper Miocene series of halite and anhydrite (Fig. 2.1) is associated with a huge<br />

sea-level drop and desiccation of the Mediterranean Sea known as the Messinian Salinity Crisis<br />

(Hsu et al., 1978, Gvirtzman and Buchbinder, 1978). <strong>The</strong> Messinian evaporites filled a 1-2 km<br />

deep topographic depression located west of the <strong>Levant</strong> coast (Tibor et al., 1992). Gradual sealevel<br />

rise during the Plio-Pleistocene resulted in renewed sediment transport and basinward<br />

progradation of a fine-grained siliciclastic wedge on the <strong>Levant</strong> shelf.<br />

<strong>The</strong> breakup of the Arabian Craton by the end of the Oligocene brought into existence the<br />

Dead Sea plate boundary. This transform fault is several thousands of kilometers long,<br />

connecting the spreading center of the Red Sea with the Taurus collision zone in southeastern<br />

Turkey (Fig. 1.1)(Freund et al., 1970; Garfunkel, 1981). <strong>The</strong> <strong>Israel</strong>i segment is characterized by<br />

several, deep continental depressions, such as the Dead Sea basin, in which thick (up to 8-10 km)<br />

Miocene and Plio-Pleistocene sections accumulated (Fig. 2.1) (Kashai and Crocker, 1987;<br />

Gardosh et al., 1997).<br />

12


3. Data Sets<br />

3.1 Well Data<br />

<strong>The</strong> well data used in this study are from 43 wells located within the basin or on its<br />

margin (Fig. 1.3). Stratigraphic subdivisions of these wells were taken from the lithostratigraphic<br />

data base of oil and gas wells drilled in <strong>Israel</strong>, compiled by Fleischer and Varshavsky (2002)<br />

(Table 3.1). Formation tops were used as control points in the construction of seismic depth<br />

maps (Table 3.1, Figs. 5.11-18). Synthetic seismograms and time-converted wireline logs of 14<br />

wells, most of them located offshore and some near the coast, were used for correlation of well<br />

and seismic data (Table 3.1, Figs. 5.1, 5.2). <strong>The</strong> lithostratigraphy and biostratigraphy of selected<br />

wells were thoroughly reviewed and were used to construct two, detailed stratigraphic crosssections<br />

(Figs. 4.5, 4.6). Additional well and outcrop data from the <strong>Levant</strong> margin were<br />

analyzed and were integrated into four composite sequence stratigraphic schemes of the Triassic,<br />

Jurassic, Cretaceous and Cenozoic (Figs. 4.1, 4.2, 4.3, 4.4).<br />

Chronostrat.<br />

Top<br />

Precamb.<br />

()<br />

Basement<br />

In Middle<br />

Jurassic<br />

Albian to<br />

Turonian<br />

Senon.<br />

to<br />

Eocene<br />

Lower<br />

Miocene<br />

Upper<br />

Miocene<br />

(Base<br />

Messin.)<br />

Upper<br />

Miocene<br />

(Top<br />

Messin.)<br />

Seismic<br />

horizon Brown Blue Green Red Cyan Purple Violet<br />

Well names:<br />

Well<br />

Abbrv.<br />

Andromeda-1 AND1<br />

2149/Ziqim 2127/Mav<br />

Asher Yam-1 ASYM1 ~7800 ~2100 479/Yagur<br />

479<br />

479 479 479<br />

Ashqelon-2 ASHQ2 3900/Barnea 2234/TY<br />

2040/Av 1303/BG<br />

1120/Ziqim 1063/Mav<br />

Ashqelon-3 ASHQ3 3400/Barnea 2098/TY<br />

1955/Av 1245/BG<br />

1009/Ziqim 961/Mav<br />

Atlit-1 Deep ATLT1 ~8500 1850/ in M. 45/Negba<br />

45 45<br />

Haifa<br />

45 45<br />

Bat Yam-1 BYAM1 2432/TY<br />

2197/Av 2087/BG<br />

1479/AshCl 1479/NPVulc<br />

Bravo-1 BRV1 2939Daliyya<br />

2197/Av 1627/BG<br />

1380/Ziqim 1350/Mav<br />

Cesarea-3 CSR3 2750/in M. 304/Bina<br />

184<br />

Haifa<br />

184/MtSc 184 184<br />

Canusa-9 CAN9<br />

885/Av 730/BG<br />

393/Ziqim 353/Mav<br />

Canusa-10 CAN10<br />

192/Av 132/BG<br />

132 132<br />

Delta-1 DLT1 1416/Negba<br />

1133/Av 1062/BG<br />

1062 1062<br />

Echo-1<br />

ECO1<br />

1972/BG<br />

1972 1807/Mav<br />

Foxtrot-1 FXT1 ~7800 2100/in M. 247/Negba<br />

247<br />

Haifa<br />

247 247 247<br />

Gaash-2 GSH2 ~6100 2950/in M. 772/Daliyya<br />

772<br />

Haifa<br />

772 772 772<br />

Gad-1 GAD1 2503/BG 1926 1848/Mav<br />

Gaza-1 GAZ1 1657/TY<br />

1631/Av 1367/BG<br />

Gevim-1 GEV1 4588 2177/<br />

567/Daliyya<br />

Shederot<br />

102/Av 32/BG<br />

Hannah-1 HAN1 3933/<br />

Gevar'am 3933 2900<br />

1367 1255/Mav<br />

32 32<br />

2763/ 1917/Mav<br />

13


Hadera-1 HAD1<br />

878/Av 477/BG<br />

416/Ziqim 416<br />

Haifa-1 HAIF1 ~6800 1670/in M. -40/Yagur<br />

Haifa<br />

-40 -40 -40 -40<br />

HLZD1<br />

1918/<br />

Helez Deep-1<br />

5916 Shederot 393/Yagur 393 393 393 393<br />

Hof Ashdod-1 HASD1 2560/TY<br />

2351/Av 2178/BG<br />

1849/Ziqlag 1729/Mav<br />

Item-1 ITM1 1007/Daliyya<br />

777/Av 727/BG<br />

687/Ziqim 687<br />

Jaffa-1 JAF1 2315/TY<br />

2057/MtSc 1472/BG<br />

1227/Ziqim 1204/Mav<br />

Joshua-2 JOSH1<br />

2502/Av 2305/AshCl<br />

2230/Ziqlag 2077/Mav<br />

Mari-3 MARI3 2021/Ziqim 1924/Mav<br />

Netanya-1 NAT1 1223/Bina<br />

1053/MtSc 1053 1053<br />

1053<br />

Netanya-2 NAT2 1315/Negba<br />

1315 1187/AshCl<br />

1022/Ziqlag 918/Mav<br />

Nir-1 NIR1 2053/Afridar 2004/Mav<br />

Nissanit-1 NIS1 ~6200 3065/ 1642/TY<br />

919/Ziqim 919<br />

Shederot<br />

1602/Av 1035/BG<br />

Noa-1 NOA1 1960/Mav<br />

Noa South-1 NOA1S 1936/Mav<br />

Nordan-4 NORD4<br />

1320/BG<br />

1134/Ziqim 1055/Mav<br />

Or-1 OR1 1973/Mav<br />

Or South-1 OR1S 1984/Mav<br />

QishonYam-1 QISHY1 ~1850<br />

1596/ 1590/BG<br />

1550/ 1180/Mav<br />

Romi-1 ROM1 1428/Mav<br />

Shiqma-1<br />

SHQ1<br />

Yam-2 YAM2 ~5500 2717/Negba<br />

Yam West-1 YAMW1 4900/<br />

2848/TY<br />

Shederot<br />

Yam West-2 YAMW2 2495/<br />

Yam Yafo-1 YAMY1 5350/Jr<br />

Undivided<br />

Time converted, used for<br />

calibration of seismic data<br />

~xxxx= not penetrated, depth is estimated<br />

1960/Av<br />

2123/Av<br />

2395/Av<br />

1352/BG<br />

2095/Lakhis<br />

2216/BG<br />

Gevar'am 2438/Av 2028/BG<br />

2380/Daliyya<br />

1774/Av 1694/BG<br />

1352 1192/Mav<br />

2096 1772/Mav<br />

1731/Ziqim<br />

1714/Ziqim<br />

1557/Ziqim<br />

1657/Mav<br />

1663/Mav<br />

1524/Mav<br />

Table 3.1- Summary of well data used in the present study. <strong>The</strong> depth of chronostratographic<br />

and lithostratigraphic tops in the wells is in meters below MSL. <strong>The</strong>se depth values were used<br />

for calibration of seismic data and depth maps. Abbreviations of lithostratigraphic units are: TY-<br />

Talme Yafe Formation, Mt Sc-Mount Scopus Group, AV-Avedat Group, BG- Bet Guvrin<br />

Formation, AshCl- Ashdod Clastics, Mav- Mavqiim Formation.<br />

3.2 Seismic Data<br />

<strong>The</strong> seismic data set used in this research includes two, large marine surveys (A and B)<br />

that were carried out during the year 2001in the framework of hydrocarbon exploration activity<br />

offshore <strong>Israel</strong> (Fig. 1.3). 98 multi-channel, 2D seismic reflection lines, totaling 12,000 km were<br />

interpreted (Fig. 1.3).<br />

14


<strong>The</strong> A survey covers most of the <strong>Levant</strong> <strong>Basin</strong> from the <strong>Israel</strong>i shallow shelf to the<br />

submarine Eratosthenes Seamount, some 200 km northwest. <strong>The</strong> B survey extends in a more<br />

densely spaced grid from near the coastline to about 120 km offshore (Fig. 1.3). <strong>The</strong> long cable,<br />

small group interval and high-energy source used in the two surveys resulted in high-quality<br />

seismic data sets that display better vertical resolution and deeper penetration than older,<br />

offshore seismic vintages.<br />

Two onshore Vibroseis lines, extending along the coastline from Haifa in the north to the<br />

Gaza area in the south, were also incorporated into the seismic data set (Fig. 1.3). Time-migrated<br />

profiles were interpreted on a seismic interpretation workstation using Paradigm-Epos3 software.<br />

Four seismic lines of the A survey were reprocessed in the GII. <strong>The</strong> primary purpose of<br />

the reprocessing was to produce pre-stacked depth-migrated profiles and to extract seismic<br />

interval velocities that were used in depth conversion of time maps.<br />

3.3 Gravity and Magnetic Data Sets<br />

Gravity and magnetic data were acquired in the B seismic survey (Fig. 1.3). A total of<br />

320,000 stations were measured along the seismic lines at intervals of about 20 m. <strong>The</strong> accuracy<br />

of the data (estimated on the crossing locations) is about a few nanoTesla for magnetic<br />

measurements and 0.1mGal for the gravity measurements.<br />

<strong>The</strong> raw gravity data were filtered and Eotvos, Free Air and Bouguer corrections were<br />

applied. <strong>The</strong> IGRF was subtracted from the magnetic data. Both data sets were incorporated in<br />

the regional gravity and magnetic grids of the <strong>Levant</strong> area compiled by Rybakov et al, (1997)<br />

(Figs. 6.1, 6.2).<br />

4. <strong>Stratigraphy</strong><br />

4.1 General<br />

<strong>The</strong> <strong>Levant</strong> area is located on the northwestern margins of the Arabo-Nubian Shield.<br />

This shield was shaped during the Late Precambrian Pan-African Orogeny at about 850-700 Ma.<br />

It underwent cratonization at ~630-600 Ma and was uplifted and peneplaned at 532 Ma<br />

(Garfunkel, 1980; Beyth and Heimann, 1999). <strong>The</strong> overlying sedimentary cover forms a<br />

northwest-thickening wedge, ranging from about 2 km in southern <strong>Israel</strong> through ca 7.5 km<br />

along the Mediterranean coastal plain and to ca 15 km in the eastern Mediterranean basin.<br />

Throughout the Phanerozoic the <strong>Levant</strong> area formed a transitional zone between the<br />

continental domain in the southeast and the marine domain of the Tethys and Neotethys oceans<br />

15


to the west. <strong>The</strong> dominant sedimentary facies east and south of the present-day Mediterranean<br />

coast were thus epicontinental, consisting of alternating, shallow-marine carbonate, shale, and<br />

sandstone, with abundant coarse-grained components in the south. Slope and basinal<br />

sedimentation was dominant in the west, across the coastal plain and in the marine basin at times<br />

when this area subsided.<br />

<strong>The</strong> Phanerozoic sedimentary section of the <strong>Levant</strong> is characterized by abundant coastal<br />

onlap-offlap cycles of different orders (Figs. 4.1 to 4.4). <strong>The</strong>se cycles reflect recurring<br />

transgressions and regressions of the Tethys and Neotethys oceans onto the Arabian Craton.<br />

Relative high and low sea-level stands were controlled both by global eustatic changes and by<br />

local, vertical tectonic movements. <strong>The</strong> latter are evident by the substantial gaps in the<br />

sedimentary records in the Paleozoic, Early Jurassic and Early Cretaceous successions.<br />

4.2 <strong>The</strong> Permo-Triassic Section<br />

<strong>The</strong> earliest Phanerozoic sediments in the central and southern part of <strong>Israel</strong> are of Early<br />

Permian age (Weissbrod, 1969b, Eshet, 1990). Late Carboniferous to Early Permian uplift of the<br />

Negev and central <strong>Israel</strong> led to the erosion of the older Paleozoic section down to the<br />

Infracambrian Zenifim Formation (Garfunkel, 1988). Pre-Permian sections that escaped this<br />

erosion may have been locally preserved, e.g., the 300 m thick section of Cambrian sandstones<br />

and carbonates that are found in the southernmost tip of the country (Elat region; Weissbrod,<br />

1969b). No Paleozoic section was penetrated in any wells in northern <strong>Israel</strong>.<br />

<strong>The</strong> Permo-Triassic shallow-marine section accumulated in a large, N-S oriented<br />

embayment, roughly coinciding with today's Judean highlands. <strong>The</strong> section thickens gradually to<br />

the north and thins to the east and west (Weissbrod, 1969a; Druckman, 1974; Druckman and<br />

Kashai, 1981; Druckman et al., 1995a; Korngrin, 2004).<br />

<strong>The</strong> western flank of this embayment is delineated by a 'basement high' that was penetrated<br />

in the Helez Deep-1, Gevim-1, and the Bessor-1 boreholes, located in the southern coastal plain<br />

(Fig. 4.5). Its position in the southern coastal plain may explain the thickening of the Permo-<br />

Triassic north and east of it. This 'basement high' was fractured and downfaulted to the west in<br />

several pulses, during the Late Triassic to Middle Jurassic times (Fig. 4.5). <strong>The</strong> Permo-Triassic<br />

stratigraphy was not penetrated, and is therefore unknown west of the ' basement high'.<br />

Gvirtzman and Weissbrod (1984) termed this feature the 'Helez High' and considered it as part of<br />

a regional 'geoanticline.'<br />

<strong>The</strong> Lower Permian Sa'ad Formation overlies the Precambrian basement and<br />

Infracambrian section and is found in wells in the Negev area and southern coastal plain. Its<br />

northernmost occurrence is in the David-1 borehole, in central <strong>Israel</strong>. <strong>The</strong> Sa'ad Formation<br />

16


consists of a 70-180 m thick fluvio-deltaic section (Weissbrod, 1969a). <strong>The</strong> deltaic sandstones<br />

may provide favorable reservoir properties.<br />

<strong>The</strong> Upper Permian (Thuringian) consists of the Arqov Formation. <strong>The</strong> formation is<br />

found in wells in the Negev area and southern coastal plain where it comprises a 110-250 m<br />

thick section of shallow-marine carbonate, sandstone and shale. <strong>The</strong> unit has source, reservoir<br />

and seal potential. Noteworthy is its great petrographic similarities with the gas producing Khuff<br />

Formation in the Arabian Gulf area (Derin and Gerry, 1981; Weissbrod, 2005).<br />

<strong>The</strong> Permian-Triassic boundary is marked by an hiatus of the Induan (lowermost Triassic<br />

stage). <strong>The</strong> boundary is marked by red clay, a characteristic fungal spike and reworked Lower<br />

Paleozoic microfossils (Eshet, 1990; Benjamini et al., 2005).<br />

<strong>The</strong> Triassic section accumulated in a large, shallow-marine embayment oriented in an N-S<br />

direction. <strong>The</strong> section thickens from about 1000 m in the Negev to over 2600 m in the northern<br />

part of the country (the Devorah-2a well; Druckman and Kashai, 1981). <strong>The</strong> thickness of the<br />

Triassic section in the Helez Deep-1, Gevim-1, and Bessor-1 wells located on the ‘Helez High' is<br />

reduced to a few tens to several hundred meters (Fig. 4.5) (Druckman, 1974, 1984; Druckman et<br />

al., 1994, 1995a); whereas the Gaash-2 and Atlit-1 deep wells located 70 and 130 km north of<br />

Helez Deep-1, respectively, penetrated over 1150 m of the Middle and Upper Triassic sections<br />

(Derin and Gerry, 1981, Korngrin, 2004). <strong>The</strong> thickness variations in the Triassic section reflect<br />

a regional northward dip of the basement and local vertical motions.<br />

<strong>The</strong> Triassic section comprises a series of onlaps reflecting ten, stacked depositional<br />

cycles, which are described in the following sub-sections and are shown in Figure 4.1.<br />

4.2.1 <strong>The</strong> early Olenekian cycle (Yamin Formation)<br />

<strong>The</strong> Yamin Formation represents the first cycle. <strong>The</strong> cycle consists of some 120 to 155 m<br />

of nearshore and shallow-marine shale and carbonates and its time of deposition spans ~ 5 m.y.<br />

4.2.2 <strong>The</strong> early Scythian to early Anisian cycle (Zafir Formation)<br />

<strong>The</strong> Zafir Formation represents the second Triassic cycle. <strong>The</strong> cycle consists of some<br />

175 to 320 m of shallow marine carbonate, calcareous shale and sandstone and its time of<br />

deposition spans ~ 3.5 m.y.<br />

4.2.3 <strong>The</strong> early Anisian cycle (Ra'af Formation)<br />

<strong>The</strong> Ra'af Formation represents the third cycle. This cycle consists of carbonate<br />

mudstones, packstones, and oolitic and oncolitic grainstones. Its thickness ranges from 80 m in<br />

the south to 170 m in northern <strong>Israel</strong> (Devora-2a well), and its time of deposition spans ~1.5 m.y.<br />

17


Ma<br />

205<br />

Jurassic<br />

201.9<br />

HETTANGIAN<br />

205.7<br />

L MID. U<br />

offshore<br />

inland<br />

Sea level (m)<br />

0 50<br />

Haq et al.<br />

1987;1988<br />

RHAETIAN<br />

210<br />

209.6<br />

SEVATIAN<br />

ALAUNIAN<br />

10<br />

215<br />

220<br />

225<br />

230<br />

235<br />

240<br />

245<br />

TRIASSIC<br />

LOWER MIDDLE UPPER<br />

SCYTHIAN<br />

NORIAN<br />

220.7<br />

CARNIAN<br />

227.4<br />

LADINIAN<br />

234.3<br />

ANISIAN<br />

241.7<br />

OLENEKIAN<br />

244.8<br />

INDUAN<br />

LACIAN<br />

TUVALIAN<br />

JULIAN<br />

LONGO-<br />

BARDIAN<br />

FASSANIAN<br />

ILLYRIAN<br />

PELSONIAN<br />

BITHYNIAN<br />

AEGEAN<br />

SPATHIAN<br />

SMITHIAN<br />

DIENERIAN<br />

GRIS-<br />

BACHIAN<br />

Erez cgl.<br />

Zafir Saharonim Mohila<br />

Shefayim<br />

Ra’af<br />

Yamin<br />

Upper Mbr.<br />

Lower Mbr.<br />

Limestone Mbr.<br />

Lst. Gypsum Mbr.<br />

Lst. Marl Mbr.<br />

Fossiliferous Lst. Mbr.<br />

Gevanim M. - U. Mbrs.<br />

Gevanim L. Mbr.<br />

2<br />

1<br />

6<br />

5<br />

4<br />

9<br />

3<br />

8<br />

7<br />

Figure 4.1-Triassic stratigraphy and marine onlaps, <strong>Israel</strong><br />

18


Late diagenetic dolomitization affected the carbonates, mainly in the south. In the southernmost<br />

occurrence at Har Arif, the carbonates are sandy towards the top of the Ra'af Formation,<br />

probably representing the highstand progradation stage.<br />

4.2.4 <strong>The</strong> late Anisian to early Ladinian cycle (Gevanim and Saharonim formations)<br />

<strong>The</strong> lower member of the Gevanim Formation reflects a lowstand wedge. This clastic<br />

wedge, was deposited in the Negev area only within a time-span of less than 1 m.y. It is<br />

composed of fluvial to fluvio-deltaic coarse- to medium-grained sandstone and shale with<br />

abundant plant remains. <strong>The</strong> unit thins from 160 m in the Makhtesh Ramon area, to 25 m in the<br />

Sherif-1 well, 40 km northwards (Druckman, 1974). <strong>The</strong> thinning of the unit marks the onset of<br />

vertical movements related to the Early Mesozoic rifting (Freund et al., 1975).<br />

Several tens of thousands of barrels of medium to high grade oil were produced from the<br />

lower Gevanim sandstones and the Raaf carbonates in the Zuk Tamrur-3 and Emmunah-1 wells,<br />

located in the western margins of the Dead Sea.<br />

<strong>The</strong> Late Anisian - Early Ladinian middle and upper members of the Gevanim Formation<br />

mark the transgressive systems tract of the cycle. <strong>The</strong> maximum flooding stage was reached<br />

during the deposition of the burrowed carbonate mudstones topping the Fossiliferous Limestone<br />

Member at the lower part of the Saharonim Formation (found in outcrop in Makhtesh Ramon in<br />

southern <strong>Israel</strong>). <strong>The</strong> unit consists of fluvio-deltaic and nearshore sandstone and shale passing<br />

upward into alternating, meter thick, marlstone and storm beds of molluscan limestone. This<br />

part of the cycle spans ~3.5 m.y. (Druckman, 1974, 1976; Benjamini et al., 1993).<br />

4.2.5 <strong>The</strong> Ladinian to early Carnian cycles (Saharonim Formation)<br />

<strong>The</strong> middle and upper parts of the Saharonim Formation comprises the fifth, sixth and<br />

seventh depositional cycles. <strong>The</strong>ir deposition lasted some ~3.8 m.y. <strong>The</strong>se units consist of<br />

ammonite bearing limestones, shell beds, oncolitic and oolitic grainstones, stromatolite beds,<br />

burrowed mudstones, flat-pebble conglomerates, dolostones, marls and gypsum beds, all<br />

deposited in tidal to shallow marine environments (Druckman, 1969, 1976; Benjamini et al.,<br />

1993). <strong>The</strong> maximum flooding stage of each cycle is marked by a burrowed carbonate mudstone<br />

that can be traced throughout the entire Negev. <strong>The</strong> Saharonim Formation displays a significant<br />

thickness change, from 260 m in the south (Ramon-1 well) to 760 m in the northern part of <strong>Israel</strong><br />

(Devorah-2 well).<br />

19


4.2.6 <strong>The</strong> Carnian cycles (Mohilla Formation)<br />

<strong>The</strong> Mohilla Formation comprises the eighth and ninth cycles. <strong>The</strong> lower boundary of<br />

the eighth cycle is identified by a sharp transition from the marine burrowed mudstone of the<br />

Saharonim Formation to the supratidal dolomicrites at the base of the lower member of the<br />

Mohilla Formation. <strong>The</strong> onset of the ninth cycle is marked by evaporite deposition in the upper<br />

member of the Mohilla Formation. <strong>The</strong> termination of the cycle is marked by the upper Carnian<br />

skeletal and stromatolite limestone.<br />

<strong>The</strong> Mohilla Formation displays a substantial northward thickening from 200 m in the<br />

Ramon outcrop to 1000 m in the Devora-2a well (Druckman and Kashai, 1981). During its<br />

depositional time span (~ 5 m.y.) a marked differentiation in lithologies took place, i.e., thick<br />

accumulations of anhydrites (150 m in the Negev and 800 m in the northern part of <strong>Israel</strong>) were<br />

deposited in distinctive basins trending in a SW–NE direction. <strong>The</strong> basins were separated by<br />

dolomitic sections of reduced thickness (50 m in the Negev and 200 m in the north), indicating<br />

elevated swells between the basins. <strong>The</strong>se facies and thickness changes indicate differential rates<br />

of subsidence apparently caused by vertical movements (Druckman, 1974; Druckman and<br />

Kashai, 1981). Normal faulting of several hundred meters during Carnian times resulted in the<br />

accumulation of the Erez Conglomerate on the foot-wall of the Helez fault (Druckman, 1984;<br />

Eshet, 1990). <strong>The</strong> Carnian vertical movements mark the second onset of extensional tectonics of<br />

the Early Mesozoic rifting phase.<br />

4.2.10 <strong>The</strong> Norian cycle (Shefayim Formation)<br />

<strong>The</strong> Shefayim Formation represents the last Triassic cycle. <strong>The</strong> unit was recognized only<br />

in the Atlit-1 and Gaash-2 wells in the northern coastal plain (Gvirtzman, 1981). <strong>The</strong> duration of<br />

the Shefayim cycle is estimated to be ~8.5 m.y. It is characterized by a reef facies consisting of<br />

sponges and other encrusting organisms, unidentified corals, and oolite beds (Derin et al., 1981;<br />

Korngrin and Benjamini, 2001; Korngrin, 2004). In the Atlit-1 well the Shefayim section reaches<br />

1150 m (including about 270 m of Jurassic magmatic intrusions).<br />

<strong>The</strong> Norian sequence was probably eroded in most of <strong>Israel</strong> during the latest Triassic to<br />

earliest Jurassic sea-level drop (Fig. 4.1). This global lowstand is marked by a notable<br />

unconformity surface of low relief, representing a matured peneplain surface. <strong>The</strong> hiatus<br />

between the Norian Shefayim Formation and the Pliensbachian Ardon Formation spans ~15 m.y.<br />

This unconformity is known throughout the Middle East. It is associated with complete<br />

exposure of Triassic terrain to karstic dissolution, laterization and the development of a thick<br />

paleosol referred to as the Mishhor Formation (Goldberg and Friedman, 1974).<br />

20


4.3 <strong>The</strong> Jurassic Section<br />

In southern <strong>Israel</strong>, the Jurassic succession consists of ~1200 m of continental to nearshore<br />

sandstone, shale and carbonate. It gradually thickens to over 3700 m of shelf carbonates in<br />

central <strong>Israel</strong> and thins to ~2200 m in northwestern <strong>Israel</strong>. South of Makhtesh Ramon the<br />

Jurassic section is entirely truncated by the Valanginian erosive unconformity. Significant<br />

thickness variations, mainly in the Lower Jurassic sections, have been related to normal faulting<br />

associated with rifting that eventually opened the Neotethys Ocean and established the<br />

continental margin of the Arabian Craton (see chapter 7). Accretion of reefs and carbonate<br />

platforms mark the shelf edge, located close to the present-day Mediterranean shoreline. In the<br />

offshore, more than 1,650 m of shelf, slope and basinal sediments accumulated.<br />

<strong>The</strong> Jurassic section was partially penetrated in six wells only in the Mediterranean<br />

offshore: Delta-1A, Yam-2, Yam Yafo-1, Yam West-1, Foxtrot-1 and Asher Yam-1(Fig. 1.3,<br />

Table 3.1). Correlations between these wells and the onshore wells are displayed in the two cross<br />

sections (Figs. 4.5, 4.6) and in the seismic profiles. <strong>The</strong> biostratigraphy of the Jurassic section<br />

offshore has been a source of confusion: Different fossil groups used for dating the samples (e.g.,<br />

pollen and spores, dinoflagelates versus benthic foraminifera) often produced contradicting<br />

results. Additional difficulties arose from the lack of core samples and from the occurrences of<br />

allochthonous rock constituents associated with mass-transport.<br />

<strong>The</strong> description of the Jurassic succession presented in the following chapters is largely<br />

based on the well-sampled section found in outcrops and wells in the Negev area. <strong>The</strong> Jurassic<br />

section comprises a series of onlaps reflecting seven, stacked depositional cycles that are shown<br />

in Figure 4.2.<br />

4.3.1 <strong>The</strong> Pliensbachian cycle (Ardon Formation)<br />

<strong>The</strong> Ardon Formation is the earliest of the Jurassic cycles (Maync,1966; Derin and<br />

Reiss,1966; Goldberg and Friedman, 1974, Perelis-Grossowicz et al., 2000). <strong>The</strong> cycle overlies<br />

the Mishhor paleosol and consists of shallow marine, peritidal carbonates with minor amounts of<br />

evaporites. Its time of deposition spans ~ 5 m.y. (Buchbinder and leRoux, 1993). <strong>The</strong> unit<br />

thickens from a few tens of meters in Makhtesh Ramon to nearly 800 m in the Hilal well in<br />

northeastern Sinai, and almost 1000 m in the Helez Deep-1A well in the southern coastal plain.<br />

<strong>The</strong>se thickness variations are associated with an Early Jurassic faulting episode that<br />

occurred simultaneously with extensive volcanism found in the subsurface in the Carmel area.<br />

<strong>The</strong>re, a thick 2500 m pile of basalts and pyroclastics, termed the Asher Volcanics, accumulated<br />

in a fault-controlled depression oriented in a NW-SE direction (Gvirtzman and Steinitz, 1983;<br />

Garfunkel, 1989; Gvirtzman et al. 1990). <strong>The</strong>se volcanics have an intraplate basalt affinity<br />

21


Ma<br />

140<br />

150<br />

160<br />

170<br />

180<br />

190<br />

200<br />

210<br />

CRETACEOUS<br />

JURASSIC<br />

TRIASSIC<br />

UPPER<br />

MIDDLE<br />

LOWER<br />

UPPER<br />

(NEOCOMIAN)<br />

MALM<br />

DOGGER<br />

LIAS<br />

KEUPER<br />

Valanginian<br />

137.0<br />

Berriasian<br />

144.2<br />

Tithonian<br />

150.7<br />

Kimmeridgian<br />

154.1<br />

Oxfordian<br />

159.4<br />

Callovian<br />

164.4<br />

Bathonian<br />

169.2<br />

Bajocian<br />

176.5<br />

Aalenian<br />

180.1<br />

Toarcian<br />

189.6<br />

Pliensbachian<br />

195.3<br />

Sinemurian<br />

201.9<br />

Hettangian<br />

205.7<br />

Rhaetian<br />

209.6<br />

Norian<br />

Middle L Mid. Upper L Upper<br />

M. L L Upp.<br />

Delta Yam<br />

Low. Gevar’Am<br />

A r s u f<br />

L M. U Lower Upper Lower Upp. Lower Middle Upper L Mid U Lower Upper L M. Upp. Lower M. U L Upp. Upp.<br />

Undivided Kidod<br />

allocthonous section<br />

<strong>Offshore</strong> Inland<br />

Haifa<br />

Bay<br />

H a i f a<br />

Nirim<br />

Nirim<br />

Haluza<br />

Be’er Sheva, Bikfaya<br />

Nir Am<br />

eroded<br />

eroded<br />

Brur<br />

Shederot<br />

Barnea<br />

Kidod<br />

Matmor<br />

Zohar<br />

Sherif<br />

Rosh Pinna Upp. Inmar<br />

Qeren<br />

Low. Inmar<br />

A r d o n<br />

4<br />

Daya<br />

S h e f a y i m<br />

N a h a<br />

Karmon<br />

Mish’hor<br />

Salima<br />

l<br />

eroded<br />

3<br />

r<br />

S a ’ a<br />

5<br />

Tayasir<br />

Ramon Laccolith<br />

Ardon dykes<br />

1<br />

6<br />

7<br />

2<br />

Asher Volcanics<br />

133<br />

140<br />

150<br />

160<br />

170<br />

180<br />

190<br />

200<br />

210<br />

0m 50 100 150<br />

Sea level<br />

Haq et al.<br />

1987;1988<br />

Fig. 4.2 - Jurassic stratigraphy and marine onlaps, <strong>Israel</strong>.<br />

<strong>The</strong> blue bar denotes theTop Bathonian seismic horizon.<br />

22


(Dvorkin and Kohn, 1989). <strong>The</strong> Asher Volcanics range in age from 206 to 189 Ma (Steinitz et al,<br />

1983; Segev, 2000)<br />

4.3.2 <strong>The</strong> Aalenian to early Bajocian cycle (Inmar Formation)<br />

<strong>The</strong> Inmar Formation comprises the second Jurassic cycle (Perelis-Grossowicz et al.,<br />

2000). <strong>The</strong> Lower Member of the Inmar Formation consists of fluvial and shoreface sandstone<br />

with minor amounts of mudstone, passing upwards to the shallow-marine, inner-shelf carbonate<br />

of the Qeren Member. <strong>The</strong> latter marks the maximum flooding surface of the cycle. <strong>The</strong> Upper<br />

Member of the Inmar Formation consists of fluvial and fluvio-deltaic sandstone, siltstone and<br />

mudstone, marking a highstand, progradational systems tract. This part of the cycle passes into<br />

the Rosh-Pinna Formation (Derin, 1974), which represents the distal part of the highstand<br />

progradation wedge in the coastal plain and northern part of the country. In the coastal plain and<br />

central and northern <strong>Israel</strong> the Inmar and the Daya formations merge distally to form a ~1500 m<br />

thick carbonate platform of the Nirim Formation. This thick platform is built of mudstones,<br />

peletal and oolitic packstones, wackestones and grainstones. <strong>The</strong> unified cycle lasted ~5 m.y.<br />

<strong>The</strong> Aalenian-Lower Bajocian sequence displays significant thickness variations in the<br />

Negev, ranging from 0 in the NE to over 900 m in the NW (Goldberg and Friedman, 1974;<br />

Druckman, 1977). <strong>The</strong>se variations indicate the most significant vertical movement related to<br />

the Early Mesozoic rifting phase that is identified in southern.<br />

4.3.3 <strong>The</strong> late Bajocian cycle (Daya and Barnea formations)<br />

<strong>The</strong> Daya and Barnea formations form the third Jurassic cycle. <strong>The</strong> cycle begins with<br />

transgressive, shallow-marine carbonate and sandstone of the Daya Formation. <strong>The</strong> thickness of<br />

the unit changes gradually from ~100 m in Makhtesh Ramon to nearly 250 m in the coastal<br />

plain. In several wells near the present coastline the Daya Formation either passes laterally to or<br />

is overlain by the deeper-marine Barnea Formation. <strong>The</strong> latter unit consists of thinly bedded<br />

sponge rhaxes, spicules and radiolarian packstones, interpreted to reflect the maximum flooding<br />

conditions of the cycle. <strong>The</strong> duration of the Daya-Barnea cycle is ~5.5 m.y.<br />

4.3.4 <strong>The</strong> Bathonian cycle (Sherif and Shederot formations)<br />

<strong>The</strong> Sherif and Shederot formations and an 'Undivided' Middle Jurassic unit found in the<br />

offshore, comprises the fourth cycle. This cycle probably represents the smallest Jurassic onlap.<br />

In the Negev the unit consists of nearshore sandstones and peritidal and subtidal carbonates of<br />

the Sherif Formation. <strong>The</strong>se may denote the lowstand to transgressive systems tracts of the cycle.<br />

In the central and northern coastal plain it comprises the oolitic shoals of the Shederot<br />

23


Formation, which probably correspond to the highstand systems tract. <strong>The</strong> inland part of the<br />

Shederot-Sherif cycle thickens from ~100 m in the south to about 500 m in the coastal plain (Fig.<br />

4.5).<br />

In the offshore wells, the fourth Jurassic cycle is found only in the lower part of Yam<br />

West-1 (4829 m to T.D.) and Yam Yafo-1 (5033 m to T.D.). <strong>The</strong> presumed Bathonian<br />

succession in Yam West-1 consists of more then 421 m (bottom not reached at the T.D.) of<br />

oolitic pelletal and intraclast grainstones and skeletal grainstones. <strong>The</strong> grainstones are stacked in<br />

20-50 m thick packages separated by a few meters of thick gray shale. On the Gamma Ray and<br />

Resistivity logs a total of 14 such packages were defined. At the base of the section three<br />

coarsening upward cycles (40-50 m thick) were detected. <strong>The</strong>se are followed by five finingupward<br />

(bell shaped), small-scale cycles (10-30 m thick). <strong>The</strong> upper part of the succession<br />

consists of five coarsening-upward cycles (25-40 m thick). <strong>The</strong> oolitic grainstones are interpreted<br />

as shoals, deposited in a platform setting that existed during the Bathonian in the Yam West-1<br />

area.<br />

In the Yam Yafo-1 well, the lowermost 752 m is dominated by siltstones and shale,<br />

interlayered with carbonates . <strong>The</strong> carbonate beds consist of skeletal and oolitic grainstones and<br />

packstones. Microconglomerates composed of all the above lithologies were detected in ditch<br />

samples. A 60 m thick mudstone interval rich in poriferan spicules and rhaxes was recorded<br />

between 5260 and 5320 m. Following Conway (written comm., 2006) a Bathonian age is<br />

assigned to this interval. <strong>The</strong> lithologic characteristics are different from the Shederot Formation<br />

found in the Yam West-1 well or in the onshore. <strong>The</strong> strata in the lower part of the Yam Yafo-1<br />

well are therefore considered as an 'Undivided' Middle Jurassic unit of a presumed Bathonian<br />

age. <strong>The</strong> existence of microconglomerates and the large amounts of clastics suggest that the<br />

'Undivided' unit in Yam Yafo-1, unlike the Shederot Formation in Yam West, may represent an<br />

allochthonous section derived in part from the nearby shelf (Druckman et al., 1994; Gardosh,<br />

2002).<br />

<strong>The</strong> Bathonian unconformity marked by the Blue horizon is correlated to the top of the<br />

Shederot Formation in Yam West-1 and to the middle part of the 'Undivided' unit in the Yam<br />

Yafo-1 well. <strong>The</strong> age definition of these offshore sections and their relation to the Shederot<br />

cycle onshore is problematic due to contradicting biostratigraphic ages obtained from different<br />

fossil groups.<br />

An age of Aalenian to Early Bajocian in the Yam West-1 well is indicated by the<br />

occurrence of the fossils Gutnicella gr. cayexi between 5080 and 5180 m and Timidonella sarda<br />

between 5180 and 5250 m. A similar age is indicated in the Yam Yafo-1 by the fossils<br />

“Mesoendothyra” sp. between 5338 and 5368 m and a badly preserved specimen of Timidonella<br />

24


sarda at 5770 m (Perelis-Grossowicz, 2006, written comm.). In the Yam West-1 well Perelis-<br />

Grossowicz further inferred a Late Bajocian-Early Bathonian hiatus of ~6-7 m.y. at a depth of<br />

4910 m. <strong>The</strong> ostracode Fabanella ramonensis, which is known from the Pliensbachian Ardon<br />

Formation in the Negev, was found at the very bottom of the Yam West-1 and Yam Yafo-1 wells<br />

together with Gomphocythere sp. known from the Bathonian of the Paris basin.<br />

In contrast to the biostratigraphic ages described above, Conway (written comm., 2006)<br />

assigned an age of "not older than Bathonian" to the lower 200 m in the Yam West-1 and Yam<br />

Yafo-1 wells. Conway based his age determination on the occurrence of the dynocyst<br />

Systematofora penicillata.<br />

<strong>The</strong> mixed occurrence (Pliensbachian and Aalenian –Bajocian) of foraminifera and<br />

ostracoda faunas in the Yam West-1 and Yam Yafo-1 wells suggestes that they are reworked in<br />

younger strata of Bathonian age, as indicated by the dynocyst assemblage. <strong>The</strong> correlation of the<br />

oolitic offshore interval in Yam West-1 with the onshore Bathonian Shederot Formation is<br />

further supported by the fact that the Bajocian Barnea Formation which underlies the Shederot<br />

Formation onshore was not encountered in the Yam West-1 well, as would be the case if the<br />

oolitic succession in this well were indeed of Aalenian –Early Bajocian age (Fig. 4.5).<br />

4.3.5 <strong>The</strong> early Bathonian to Callovian cycle (Karmon, Zohar, Matmor and Brur<br />

formations)<br />

<strong>The</strong> Karmon, Zohar, Matmor and Brur formations comprise the fifth cycle. <strong>The</strong> Karmon<br />

shale is a terrigeneous unit several tens of meters thick found in wells of the southern coastal<br />

plain below the Zohar limestone. It is interpreted as the lowstand systems tract of the cycle that<br />

was deposited following a sea-level drop and exposure of the underlying Shederot shelf (Fig.<br />

4.5) (Gardosh, 2002). <strong>The</strong> base of the Karmon shale is an upper Bathonian unconformity surface<br />

that is correlated to the Blue seismic horizon (Fig. 4.2). <strong>The</strong> overlying Zohar and Matmor<br />

formations in the Negev area consist of shallow-marine, low-energy carbonate interlayered with<br />

minor amounts of shale and sandstone. In the coastal plain the equivalent strata are the oolitic,<br />

oncoid, skeletal grainstones of the Brur Formation. <strong>The</strong> three shallow-marine units are<br />

interpreted as the highstand systems tract of the cycle. <strong>The</strong> strata of the fifth cycle attain<br />

thicknesses of 160-180 m in the northern Negev and ~ 100 m in the coastal plain. An<br />

unconformity surface associated with karst phenomena found at the top of the Brur Formation<br />

marks the end of the cycle (Buchbinder, 1979).<br />

In the offshore Yam-2, Yam West-1 and Yam Yafo-1 wells the Zohar Formation ranges<br />

in thickness from 119 to 138 m. It consists of oolitic, pelletal and intraclast and skeletal<br />

grainstones, mudstones and packstones made of poriferan spicules, and a few intercalations of<br />

25


dark gray shale with radiolaria. <strong>The</strong> Zohar Formation conformably overlies the 'Undivided'<br />

Middle Jurassic unit in the Yam Yafo-1 borehole and the Bathonian Shederot Formation in the<br />

Yam West-1 borehole.<br />

Based on a revision of the foraminiferal biostratigraphy in the three offshore wells,<br />

Perelis-Grossowicz (written comm., 2006) updated the age of the Zohar to uppermost Bathonian<br />

to lower Callovian. Conway (written comm., 2006) assigned a Callovian age, similar to the age<br />

assigned to the Zohar and Matmor formations onshore (Hirsch, 2005).<br />

Over 2 Bcm of thermogenic gas was produced from fractured limestone of the Zohar<br />

Formation in the Dead Sea area (Zohar, Kidod and Qanaim fields). Sub-commercial quantities<br />

of high-grade oil were also recovered from oolitic grainstones of the Zohar Formation in the<br />

offshore Yam Yafo-1 and Yam-2 wells.<br />

4.3.6 <strong>The</strong> Haifa Formation<br />

<strong>The</strong> Haifa Formation is a thick and uniform carbonate succession of Bajocian to<br />

Callovian age (Derin, 1974) that is found in wells in the northern coastal plain and offshore area<br />

(Fig. 4.6). This succession is equivalent to the 3 rd , 4 th and 5 th cycles of the central and southern<br />

parts of <strong>Israel</strong>. <strong>The</strong> Bathonian unconformity marked by the Blue horizon is correlated to the top<br />

of the Lower Haifa Member (Fig. 4.6)<br />

4.3.7 <strong>The</strong> Oxfordian cycle (Kidod Formation)<br />

<strong>The</strong> Kidod Formation forms the sixth depositional cycle. <strong>The</strong> unit consists of dark gray to<br />

black shale, with minor intercalations of carbonate mudstone and packstone. It ranges in<br />

thickness from 0 to 200 m and its deposition spans ~2.6 m.y. Its maximum thickness is found<br />

along a NE-SW trending belt, extending from Mt. Hermon towards the northwestern Negev. In<br />

most of the Negev the unit was entirely truncated by the Valanginian unconformity. In the<br />

northern coastal plain and northern <strong>Israel</strong> it is entirely absent. In the offshore Yam West-1, Yam-<br />

2 and Yam Yafo-1 wells the unit reappears and ranges in thickness from 170 to 340 m. In these<br />

wells it consists of gray, green and reddish shale and a few interlayers of poorly sorted<br />

grainstones containing intraclasts, pellets, forams, and echinoids. <strong>The</strong> Kidod shale is generally<br />

barren of fauna, though a few radiolaria, foraminifera, molluscs, echinoids and ostracodes were<br />

found. In the offshore the unit conformably overlies the Zohar Formation and underlies the Delta<br />

Formation. This predominantly transgressive unit marks the termination of the Middle Jurassic<br />

carbonate platform offshore <strong>Israel</strong>, while onshore, carbonate platform conditions were later<br />

reestablished.<br />

26


An Oxfordian age was assigned to the Kidod Formation (Druckman et al., 1994; Gill et al.,<br />

1995; Conway, 2006, written comm.). Perelis-Grossowicz (2006, written comm.) attributed a<br />

Middle Callovian – Oxfordian age to the Kidod Formation, whereas Derin et al. (1990) assigned<br />

it an Oxfordian –Kimmeridgian age. In the Hermon and Maghara outcrops an early to middle<br />

Oxfordian age was assigned, based on ammonite zonation (Hirsch, 2005).<br />

<strong>The</strong> distribution pattern of the Kidod Formation in the offshore wells and in central <strong>Israel</strong><br />

and its absence along the coastal area may be attributed to submarine erosion or non deposition<br />

along the upper slope during late Oxfordian times. However, subaerial erosion cannot entirely be<br />

ruled out. This interpretation differs from that proposed by Derin, (1974), who related the<br />

absence of the Kidod Formation in the coastal plain to interfingering with carbonates of the Nir<br />

Am Reef. Derin (1974) further attributed the thick accumulation of the Kidod in the NE-SW<br />

trending belt to backreef lagoonal conditions. <strong>The</strong> proposed model herein indicates open marine<br />

conditions, which accords better with the rich ammonite fauna found in the Hermon and<br />

Maghara outcrops. <strong>The</strong> widespread distribution of Oxfordian shale in the Middle East better<br />

supports an open marine model for the Kidod shale rather than a shelf lagoon as proposed by<br />

Derin (1974).<br />

4.3.8 <strong>The</strong> late Oxfordian to Tithonian cycle<br />

<strong>The</strong> upper Oxfordian to Tithonian strata comprise the seventh cycle. This cycle, which<br />

lasted about 12 m.y., is assumed to represent the maximum onlap extension of all Jurassic cycles.<br />

It corresponds to the highest Jurassic global sea-level stand (Haq et al., 1988). <strong>The</strong> cycle<br />

comprises the Beer Sheva, Haluza, Nahal Saar, Haifa Bay , Bikfaya and Salima formations<br />

onshore <strong>Israel</strong> and southern Lebanon, and the Delta and Yam formations in the offshore wells.<br />

In the onshore it consists of 200 to 400 m of shallow marine carbonate mudstones, packstones,<br />

coral and stromatoporid reefs, and oolitic and oncolithic shoals (Nir Am, Be'er Sheva, Nahal<br />

Saar, Bikfaya formations). <strong>The</strong>se carbonates are overlain by peritidal sandstones, siltstones and<br />

shales interlayered with low–energy, subtidal carbonates and numerous, locally developed oolitic<br />

shoals (in the Haluza and Salima formations). <strong>The</strong>se Upper Jurassic units have escaped<br />

truncation by the Valanginian unconformity due to their low structural position (Eliezri, 1964).<br />

In the offshore wells the Oxfordian to Tithonian sequence is represented by the Delta and<br />

Yam formations. <strong>The</strong> Delta Formation varies in thickness from 148 to 240 m. It consists of<br />

redeposited reefoidal limestones, oolitic and pelletal grainstones, skeletal grainstones, burrowed<br />

shale and minor amounts of fine to medium-grain, moderately sorted quartz sandstone. Twelve<br />

meters of the Delta Formation were cored in the Delta-1A well. <strong>The</strong> cores contain mainly<br />

burrowed black shale and rip-up clasts, with occasional 10-20 cm thick layers consisting of<br />

27


centimeter-scale redeposited, oolitic and skeletal grainstones with fine to coarse lithoclasts. A<br />

few fining-upward cycles several centimeters thick were also observed. Three distinct volcanic<br />

bodies, 15-20m thick are found within the Delta Formation, in the Delta-1 well.<br />

Friedman et al. (1971) interpreted a depositional slope environment for the Delta<br />

Formation. In the Yam-2 and Yam West-1 wells, three stacked fining-upward cycles are<br />

observed on the Gamma Ray and Resistivity logs that possibly represent submarine, migrating<br />

mid-fan channel fill. <strong>The</strong>se fining-upward cycles are poorly developed in the Delta-1A well and<br />

absent in the Yam Yafo-1 well. An Oxfordian to Kimerridgian age was assigned to the Delta<br />

Formation offshore (Derin et al., 1988, 1990; Moshkovitz, in Derin et al., 1988; Druckman et al.,<br />

1994; Gill et al., 1995; Perelis-Grossowicz and Conway, 2006, written comm.). <strong>The</strong> unit<br />

conformably overlies the Kidod Formation and underlies the Yam Formation.<br />

<strong>The</strong> Yam Formation ranges in thickness from 100 to 387 m. It consists of dark, finely<br />

laminated shale and siltstone generally barren of faunal elements but occasionally containing<br />

some foraminifera and ostracodes, skeletal, oolithic and pelletal packstones and fine to coarsegrained<br />

grainstones, subangular to subrounded quartz sandstone, poorly sorted (Druckman et al.,<br />

1994; Gill et al., 1995). In the Delta-1 well volcanic fragments were also recorded (Derin et al.,<br />

1990). <strong>The</strong> Yam Formation conformably overlies the Delta Formation and underlies the Arsuf<br />

Formation. Its age was determined as Tithonian, however a Kimmeridgian age for its lower part<br />

is possible (Derin et al., 1988, 1990; Moshkovitz in Derin et al., 1988; Druckman et al., 1994;<br />

Gill et al., 1995; Perelis-Grossowicz and Conway, 2006, written comm.).<br />

During the Late Jurassic a distinct shelf-edge was established separating the shallow shelf<br />

to the east from a slope and basin to the west. <strong>The</strong> Delta and Yam formations are thus interpreted<br />

as slope and basin deposits that were deposited during Late Jurassic lowstands. <strong>The</strong> absence of<br />

these units west of the shelf-edge is related to either: (a) non deposition on the Late Jurassic<br />

upper slope; or (b) erosion of both the upper slope and shelf deposits during the Valanginian<br />

unconformity.<br />

Onshore, there is a hiatus of ~ 23 m.y. above the Upper Jurasic strata, from the Berriasian<br />

until the lowermost Aptian. <strong>Offshore</strong>, however, continuous sedimentation took place across the<br />

Jurassic-Cretaceous boundary. It is possible that Berriasian sediments were deposited onshore<br />

and were later eroded during latest Berriasian to Early Valangian sea-level drop (see next<br />

chapter).<br />

28


Delta 1A<br />

Yam 2<br />

Yam Yafo 1<br />

Yam West 1<br />

Formation<br />

Top<br />

Thickness<br />

Top<br />

Thickness<br />

Top<br />

Thickness<br />

Top<br />

Thickness<br />

m.<br />

m.<br />

m.<br />

m.<br />

m.<br />

m.<br />

m.<br />

m.<br />

Arsuf<br />

3905<br />

200<br />

4263<br />

171<br />

4127<br />

63<br />

3953<br />

109<br />

Yam<br />

4105<br />

100<br />

4434<br />

263<br />

4190<br />

387<br />

4062<br />

210<br />

Delta<br />

4205<br />

218+<br />

4697<br />

198<br />

4577<br />

148<br />

4272<br />

240<br />

Kidod<br />

4895<br />

340<br />

4725<br />

170<br />

4512<br />

198<br />

Zohar<br />

5235<br />

135+<br />

4895<br />

138<br />

4710<br />

119<br />

Shederot/Undefined<br />

5033<br />

752+<br />

4829<br />

421+<br />

Total Jurassic<br />

518+<br />

1107+<br />

1658+<br />

1297+<br />

Table 4.1- Summary of Jurassic formation tops and thicknesses in offshore wells.<br />

4.4 <strong>The</strong> Cretaceous Section<br />

4.4.1 Berriasian to early Valanginian cycle (Arsuf Formation)<br />

A global sea-level drop started in the Berriasian and culminated in the early Valanginian<br />

(Fig. 4.3) (Haq et al., 1988). At the same time, a large-scale uplift of the landmass took place,<br />

east and south of the present-day shoreline. This uplift was caused by heating related to an<br />

upwelling of a mantle plume associated with the Tayasir Volcanics and the shallow Ramon<br />

intrusions (Garfunkel and Derin, 1984; Gvirtzman et al., 1998). Consequently, intensive and<br />

widespread erosion took place. <strong>The</strong> intensity of the erosion increased from north to south,<br />

eventually resulting in the removal of most of the Phanerozoic section in the southern Negev<br />

(Eliezri, 1964; Weissbrod, 1969a; Druckman, 1974; Garfunkel and Derin, 1984).<br />

<strong>The</strong> Arsuf Formation (a new formation name proposed herein) ranges in thickness from<br />

63 to 200 m. (Table 4.1). It consists of carbonate debris interbedded with thin shale<br />

intercalations. <strong>The</strong> carbonate debris is apparently allochthonous, representing reworked<br />

carbonate platform sediments, e.g., skeletal, oolitic, pelletal, grainstone or packstone,<br />

occasionally sandy. <strong>The</strong> formation is characterized by a 'bow' shaped gamma ray pattern,<br />

reflecting a gradual decrease of mud content towards the crest of the 'bow' and an increase of<br />

mud again towards its top. This pattern characterizes waxing and waning of sedimentation on a<br />

mixed sand-mud fan (Emery and Myers, 1996). <strong>The</strong> sandy component of the Arsuf fan, however,<br />

is made of carbonate debris.<br />

<strong>The</strong> age of the Arsuf Formation is latest Tithonian to Berriasian. However, an early<br />

Valanginian age cannot be ruled out (Fig. 4.3) (Derin et al., 1988, 1990; Moshkovitz, in Derin et<br />

al., 1988; Druckman et al., 1994; Gill et al., 1995). Reworked Jurassic faunal elements are<br />

29


LOWER CRETACEOUS UPPER CRETACEOUS<br />

85.8<br />

Coniacian<br />

89.0<br />

Turonian<br />

93.5<br />

Cenomanian<br />

98.9<br />

Albian<br />

112.2<br />

Aptian<br />

121.0<br />

Barremian<br />

127.0<br />

Hauterivian<br />

132.0<br />

Valanginian<br />

137.0<br />

Berriasian<br />

144.2<br />

Upp. Lower Mid. Upper L Upper Lower Upper Lower Upper Lower Upper Lower Middle Upper Lower M Up Low Mid Up L M Up L<br />

Item Yagur Kamon<br />

Kefira<br />

Glauconite horizon<br />

Uza<br />

T a l m e - Y a f e<br />

G e v a r ’ a m<br />

A r s u f<br />

<strong>Offshore</strong><br />

Daliyya<br />

Negba<br />

Y a k h i n i<br />

R a m a<br />

Coarse Cgl,<br />

Yavneh Shale<br />

Helez<br />

Telamim<br />

Upp. Hidra<br />

Hevyon<br />

Couches a Knemiceras<br />

Qatana<br />

En Qinya<br />

“Zumoffen”<br />

Tamun<br />

L.C.-3<br />

Deragot<br />

Low. Hidra<br />

Ein el Assad “Blanche” Zuweira<br />

Nebi Sa’id<br />

Top Jurassic<br />

Gres de Base<br />

H a t i r a<br />

Inland<br />

Senonian drowning<br />

Upper Nezer Zihor<br />

Gerofit, up. Bina<br />

Shivta “Vroman”<br />

“Clastic<br />

Drowning<br />

Low. Ora Unit”<br />

Mukhraqa Sakhnin, Tamar Up. Ora<br />

Avnon<br />

N. Oren<br />

Deir Hanna<br />

Zafit<br />

En Yorqeam<br />

Distal slope<br />

fan sands<br />

Continental<br />

Kesalon<br />

event<br />

Kohal, Chouf<br />

Hermon<br />

Tayasir, Hermon<br />

Carmel<br />

Ramon<br />

Ramon<br />

Judea Gr.<br />

K u r n u b G r .<br />

85 Sea level (Haq et al., 1987, 1988)<br />

0 50 100 200<br />

90<br />

95<br />

100<br />

105<br />

110<br />

115<br />

120<br />

125<br />

130<br />

135<br />

140<br />

145<br />

Fig. 4.3 - Cretaceous stratigraphy and marine onlaps, <strong>Israel</strong>.<br />

Selected continental deposits are delineated by dotted lines<br />

<strong>The</strong> green bar denotes Top Turonian seismic horizon.<br />

30


occasionally present. This unit conformably overlies the Yam Formation and underlies the<br />

Gevar’am Formation. <strong>The</strong> redeposition of Jurassic microfauna in the Berriasian to early<br />

Valanginian sediments supports the interpretation that the main erosion phase took place during<br />

this lowstand (Druckman et al., 1994).<br />

4.4.2 Valanginian to Aptian cycle (Gevar'am Formation)<br />

<strong>The</strong> pronounced uplift and related sea-level drop of late Berriasian - early Valangian<br />

times resulted in intensive erosion and transportation of the erosional products towards the sea.<br />

<strong>The</strong> transportation conduit persisted along the sea floor, eventually resulting in the incision of a<br />

large submarine canyon (Gevar'am Canyon), which in the coastal plain area, reaches a width of<br />

ca 30 km and depth of ca 1000 m (Cohen, 1976). <strong>The</strong> canyon is filled by marine shales<br />

(Gevar'am Formation) of Valangian to Barremian age with a few conglomerate and sandstone<br />

intercalations.<br />

Sandstone intercalations up to 15-20 m thick were encountered in the Yam-2, Yam West-<br />

1 and Yam West-2 offshore wells. <strong>The</strong>y are referred to as "Lower Sand" of Hauterivian age and<br />

"Upper Sand"of Barremian age (Fig. 4.3) (Isramco Report ). Most of these intercalations display<br />

a serrated log motif, indicating sand/clay interchange, which characterizes a 'channel levee'<br />

system of proximal slope fan (Emery and Myers, 1996). <strong>The</strong> lower sand in the Yam West-1 well<br />

displays a blocky gamma-ray motif that probably represents a sand lobe that breached the<br />

channel's levee (crevasse splay). <strong>The</strong> interpretation the offshore sand beds further suggestes that<br />

the basin floor should have been located westward of the Yam wells.<br />

4.4.3 Hauterivian to early Aptian cycle (Helez and Telamim formations)<br />

<strong>The</strong> (siliciclastic-rich) Helez Formation (up to 240 m thick) represents a gradual<br />

encroachment of sea onto the shelf, probably in Barremian times (Fig. 4.3). This rise culminated<br />

in the Lower Aptian and is reflected by the oolite-rich Telamim (270 m)/Ein el Assad (Muraille<br />

de Blanche) carbonate platform and the Zuweira marine intercalation within the largely<br />

continental Hatira sandstone of the Negev (Fig. 4.3) (Weissbrod et al., 1990; Lewy and<br />

Weissbrod, 1993). In the offshore, the Helez -Telamim cycle is represented by the Gevar'am<br />

Formation, especially by its upper part (Fig. 4.3). Inland, the base-level rise during deposition of<br />

the Helez Formation resulted in the proximal accretion of continental, fluvial deposits of the<br />

Kohal Formation (up to 400 m thick; Shilo and Wolf, 1989), the sandstones below the Nabi Said<br />

Formation in the Qiryat Shemone (Galilee) area and the Gres de Base or Chouf Formation<br />

(Walley, 1998) in Lebanon.<br />

31


<strong>The</strong> time span of the Helez -Telamim deposition was ca 10 m.y., thus encompassing a<br />

second-order depositional cycle. <strong>The</strong> cycle terminates abruptly and is replaced by a pelagic shale<br />

interval (Yavneh Shale, ca 50 m thick) marked by the LC-3 electrolog marker (Fig. 4.3). This<br />

change reflects a drowning phase which hampered platform growth. In northern <strong>Israel</strong> the<br />

platform carbonate of the Ein el Assad Formation shows a pronounced deepening towards the<br />

uppermost part of the formation (Bachmann and Hirsch, 2006), presumably paralleling the L.C.-<br />

3 drowning phase. A global ecological event responsible for the demise and drowning of<br />

carbonate platforms in the Mediterranean area at the end of the lower Aptian (Castro and Ruiz-<br />

Ortiz, 1995; Wissler et al., 2003) could have been also responsible for the termination of the<br />

Telamim/Ein el Assad platform (Bachmann and Hirsch, 2006).<br />

<strong>The</strong> overlying Hidra Formation in the Galilee proximal area begins with oolitic marls of<br />

open lagoon environment (Bachmann and Hirsch, 2006). It is portrayed as a late highstand<br />

progradation stage of the Telamim - Ein el Assad cycle (Fig. 4.3). In the middle part of the Hidra<br />

Formation, 10 m of recurring intercalations of fluvial sediments and ferruginous crusts signal a<br />

pronounced sequence boundary and lowstand systems tract deposits.<br />

A notable sandstone horizon overlain by a dolomite bed was identified by Grader and<br />

Reiss (1958), directly above the Yavneh Shale. In their composite log of the Lower Cretaceous<br />

in the Helez area this horizon is termed the "Main Shale Break" (M.S.B. electrolog marker). A<br />

correlation is proposed between this sandstone bed and the coarse conglomerate at the base of<br />

the Aptian-Albian Talme Yafe-Yakhini cycles (see below). This lowstand event is also<br />

responsible for the prominent erosion surface, which the Talme Yafe sediments overlie.<br />

4.4.4 Late Aptian cycle (upper Hidra to top Limestone Member of the Rama Formation)<br />

<strong>The</strong> next cycle (in proximal areas of northern <strong>Israel</strong> and the Negev) includes the upper<br />

Hidra, the "Zumoffen" Limestone Member and the Deragot marine intercalation within the<br />

Hatira sandstone in the Negev (Fig. 4.3). Its time span is ca 7 m.y. Along the coastal Hinge Belt<br />

it is represented by a carbonate platform facies of the Yakhini Formation (Fig 4.3). West of the<br />

Hinge Belt it is represented by the Talme Yafe prism of carbonate debris. <strong>The</strong> Talme Yafe<br />

Formation unconformably overlies different units from the Aptian Telamim Formation to Middle<br />

Jurassic formations, with coarse conglomerate breccia at its base (Bein and Weiler, 1976). <strong>The</strong><br />

reasons for this intense erosion are not clear. Unlike the truncation by the Gevar'am Formation,<br />

which was associated with a tectonic uplift, the truncation at the base of the Talme Yafe<br />

Formation could have been caused by the pronounced sea-level fall in middle Aptian times,<br />

which is clearly displayed in Haq's curve (Fig. 4.3).<br />

32


4.4.5 Albian cycle<br />

<strong>The</strong> restricted environment at the top of the "Zumoffen" Limestone Member of the lower<br />

part of the Rama Formation reflects a sequence boundary (Fig. 4.3) (Bachmann and Hirsch,<br />

2006). <strong>The</strong> next cycle in proximal areas includes the Rama Formation (Couches a Knemiceras)<br />

of mixed platform and basinal carbonate and the Yagur Formation, representing the highstandprogradation<br />

stage (Fig. 4.3). A parallel section was also found in the Judea Mountains<br />

(Ramallah area; Shachnai, 1969). <strong>The</strong> time range of this cycle is ca 13 m.y. In the more proximal<br />

areas of the Negev, the well-known sandy glauconite zone just below the Late Albian carbonate<br />

of the Hevyon Member (Judea Group) signifies the maximum flooding interval of the Albian<br />

cycle. In the coastal plain and the offshore distal areas the sedimentation of the Yakhini and the<br />

Talme Yafe formations continued.<br />

4.4.6 Cenomanian cycle<br />

An erosional unconformity at the top of the Yagur Formation (Fig. 4.3) (Folkman, 1969;<br />

Kafri, 1986; Lipson-Benitah et al., 1997) is marked by conglomerates and breccia, calcrete and<br />

dedolomitization, indicating aerial exposure (Folkman, 1969) and is thus designated as a<br />

sequence boundary. Extensive volcanism products overlie this surface in the Carmel and Umm el<br />

Fahm (Segev et al., 2002). In the northern Negev, this surface truncates the rudistid buildup at<br />

the top of the Hevyon Member, which contains the foraminifer Preaealveolina, thus indicating<br />

an early Cenomanian age (Lewy and Weissbrod, 1993). <strong>The</strong> overlying cycle begins with a thin,<br />

ca 0.5 m glauconitic carbonate bank with Pycnodonte vesiculosa and with early Cenomanian<br />

ammonites, representing part of the transgressive systems tract of the En Yorqeam – Zafit subcycle<br />

(Fig. 4.3). In the Jerusalem and Hebron hills the top of the Kesalon Formation represents<br />

the same phase (Fig. 4.3) (Braun and Hirsch, 1994).<br />

<strong>The</strong> substantial abnormal thickness (2700 m) of the Cenomanian Item Formation in the<br />

Item-1 well offshore (reexamination by S. Lipson Benitah, 2006) is attributed to a submarine<br />

canyon incised into the Albian sediments of the Talme Yafe Formation as a result of the Kesalon<br />

lowstand event (Fig. 4.3). Cenomanian, mixed planktic and benthic foraminifera were found in<br />

the Item Canyon fill. Notably, both the Kesalon event and the younger Mukhraqa channeling<br />

events (see below) are associated with volcanic eruptions in the Carmel, at the base of the Isfiye<br />

Chalk (=Negba Formation) and the Mukhraqa Formation (Segev et al., 2002).<br />

This early Cenomanian erosional surface was also described in Jordan by Abed (1984)<br />

and by Schulze et al. (2003) as sequence boundary CeJo1 at the top of Member b of the Naur<br />

Formation. <strong>The</strong> time span of the emergence which caused the unconformity was brief and was<br />

33


soon followed by extensive sea level rise (represented by the chalky Negba Formation),<br />

suppressing carbonate platform development, except in the Judean hills (partial drowning).<br />

<strong>The</strong> strata of the Cenomanian cycle are mostly basinal, pelagic sediments such as chalks<br />

with chert beds, e.g., the Deir Hanna Formation in the Galilee, the chalk complex of the Carmel<br />

and the Negba Formation of the coastal plain and the offshore areas (Fig. 4.3). This chalk event<br />

indicates a drowning phase suppressing and eventually replacing the Albian carbonate platform.<br />

Inland, the cycle terminates with the prograding platform carbonate of the Sakhnin, Weradim<br />

and Tamar formations. In the Negev, the cycle is divided into two transgressive-regressive<br />

cycles: (1) En Yorqeam/Zafit, (2) Avnon/Tamar (Fig. 4.3).<br />

Bet Oren– Atlit incised canyon of Middle/Late Cenomanian age- In the Carmel region,<br />

pelagic distal ramp deposits of the upper Cenomanian (Junediya Formation) are overlain by the<br />

lower part of the Late Cenomanian Mukhraqa Formation (age emendation by Buchbinder et al.,<br />

2000). Various types of transport phenomena characterize the latter unit, and are interpreted as<br />

distal bioclastic material shed off the prograding highstand deposits of the Late Cenomanian.<br />

‘Reefal’ sediments of this unit were reported in an E-W trending erosive channel 10 km long<br />

which breaches the shelf edge near Atlit (Bein, 1977, figs. 2, 3). This channel may have been<br />

initiated as a large-scale slope failure. <strong>The</strong> erosion cuts down to early Cenomanian sediments.<br />

<strong>The</strong> age of the "reefal" Mukhraqa facies was emended to Late Cenomanian (instead of<br />

Turonian), based on the presence of caprinids (Buchbinder et al., 2000). <strong>The</strong> time of incision of<br />

this channel should, therefore, be late Middle Cenomanian to pre or early Late Cenomanian.<br />

Presumably, it is related to Haq's (1988) major sequence boundary Ce3/Ce4 at 94 m.y.<br />

(recalibrated to Gradstein et al., 1995) following a 60 m sea-level fall.<br />

4.4.7 First Turonian cycle<br />

This cycle begins with a major Late Cenomanian drowning phase which resulted in the<br />

demise of upper Cenomanian platforms (Tamar, Weradim, Sakhnin) followed by different time<br />

durations of condensation and hiatuses (Fig. 4.3) (Buchbinder et al., 2000). Sedimentation was<br />

later resumed in hemi-pelagic facies (Fig. 4.3) (Ora, Yirqa and Daliyya) and finally carbonate<br />

platforms were reestablished in proximal areas: Shivta; "Vroman" (Fig. 4.3). <strong>The</strong> cycle was<br />

abruptly terminated by a pronounced sea-level fall, resulting in the exposure of the shelf and the<br />

development of a system of rivers and overbanks, depositing quartzose sandstone and shale of<br />

the "Clastic Unit" (Fig. 4.3) (Sandler, 1996). In distal areas (coastal plain and Mediterranean<br />

offshore) basinal sediments of the Daliyya Formation either accumulated, or were replaced by<br />

condensed sections and hiatuses.<br />

34


4.4.8 Middle Turonian to late Coniacian cycle (upper Bina, Gerofit and Zihor<br />

formations)<br />

<strong>The</strong> cycle begins in proximal areas with platform carbonate of the Gerofit and upper<br />

Nezer formations (Fig. 4.3). A deepening trend developed towards Early Coniacian times and in<br />

late Coniacian the platform facies was abruptly replaced by pelagic-rich beds.<br />

4.4.9 <strong>The</strong> Santonian to Paleocene basinal sediments (Mount Scopus Group)<br />

<strong>Basin</strong>al-type sedimentation started in the late Coniacian and persisted inland throughout<br />

the Santonian to Paleocene times. <strong>The</strong>se sediments are represented by the Mount Scopus Group<br />

(Flexer, 1968) and are dominated by pelagic chalk both in far proximal areas and in the offshore.<br />

<strong>The</strong>y contain also marl, chert and phosphate. Sedimentation took place in the framework of an<br />

upwelling system that affected the southern coasts of the Tethys Ocean along a belt<br />

encompassing today's North Africa and the east <strong>Levant</strong> margins (Almogi-Labin et al., 1993). <strong>The</strong><br />

widespread pelagic sedimentation reflects a major drowning of the shallow-marine Judean<br />

platform that occurred due to oceanographic, ecological change, probably as a consequence of<br />

re-arrangement of the plates bordering the Mediterranean Tethys. Several eustatic lowstands at<br />

that time (Flexer and Honigstein, 1984; Gardosh, 2002) did not effect the extensive penetration<br />

of the sea far inland and thus could not be expressed in the onlap curve (Fig. 4.3).<br />

<strong>The</strong> Mount Scopus section contains organic-rich carbonate intervals. <strong>The</strong>se bituminous<br />

rocks were deposited in the highly productive environment of small morphotectonic basins<br />

(Syrian Arc type synclines) and contain up to 20% of Type II kerogen (Tannenbaum and<br />

Aizenshtat, 1985; Gardosh et al., 1997). <strong>The</strong> Senonian bituminous rocks show geochemical<br />

correlation to asphalt, oil and gas accumulations found near the Dead Sea Lake and are<br />

considered to be the source of hydrocarbons that were generated within the Dead Sea <strong>Basin</strong><br />

(Tannenbaum and Aizenshtat, 1985; Gardosh et al., 1997).<br />

4.5 <strong>The</strong> Tertiary Section<br />

4.5.1 General<br />

Facies patterns of the Lower and Middle Eocene are predominantly pelagic. Although<br />

seemingly they contain shallower neritic interludes of nummolitic carbonates, the mixture with<br />

pelagic elements indicates a mass transport origin in a deeper basinal setting (Benjamini, 1993)<br />

beginning in latest Early Eocene times (Buchbinder et al., 1988).<br />

Remarkably, the drowning conditions that started in the Sennonian persisted during most<br />

of the Tertiary; for about 70 m.y. and a true carbonate platform was re-established for a brief<br />

time span (ca 1 m.y.) only in Middle Miocene times (the Langhian Ziqlag Formation).<br />

35


Ma<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

SERIES STAGES<br />

PLIOCENE<br />

MIOCENE<br />

OLIGOCENE<br />

EOCENE<br />

Mid.L.<br />

Early<br />

Zancl.<br />

MIDDLE UPPER<br />

LOWER<br />

UPPER<br />

LOWER<br />

UPPER<br />

MIDDLE<br />

Piace. G<br />

2.13<br />

3.22<br />

3.95<br />

MP3<br />

4.50 MP2<br />

Tortonian<br />

11.20<br />

Serravallian<br />

14.80<br />

Langhian<br />

N9<br />

N8<br />

16.40<br />

N7 Gl. insueta<br />

17.3<br />

N6 Cx. stainforthi<br />

18.8<br />

Burdigalian<br />

20.52<br />

Aquitanian<br />

23.8<br />

Chattian<br />

41.3<br />

Lutetian<br />

BIOSTRATIGRAPHY<br />

MP6<br />

MP5<br />

MP4<br />

5.11 MP1<br />

5.33<br />

Messinian<br />

7.12 N17<br />

N16<br />

N15<br />

N14<br />

N11-13<br />

N10<br />

P22<br />

N3<br />

Bet Guvrin<br />

Shoreline<br />

4.37<br />

Judean Hills<br />

Slumped complex<br />

Turbidite sands<br />

Cgl + sandstones, offshore<br />

Canyons’ incision Bet Nir cgl.<br />

Gr.f.peripheroronda Bet Ziqlag<br />

Pb. glomerosa Guvrin<br />

Catapsydrax<br />

dissimilis<br />

21.5<br />

Pr. kugleri<br />

Gg. ciperoensis<br />

ciperoensis<br />

27.1<br />

Lakhish +<br />

b<br />

Ashdod clastics<br />

28.5<br />

a Pr. opima opima<br />

LS prograding wedge<br />

P20,N1<br />

30.6<br />

Rupelian<br />

P19 Gg. ampliapertura<br />

32.0 Canyons’ incision<br />

P18 Cg. chipolensis/ Bet Eroded<br />

Guvrin<br />

Fiq+ L. Ha’on<br />

Ps. micra<br />

33.7<br />

P17 33.8 Canyons’ incision<br />

Gr. cerroazulensis<br />

Priabonian P16<br />

Mostly eroded<br />

35.3<br />

Har Aqrav<br />

37.0 P15 Gk. semiinvoluta<br />

Bet Guvrin<br />

38.4<br />

Qeziot<br />

Bartonian P14 T. rohri<br />

P13 O. beckmannii 40.5<br />

Maresha<br />

P12 M. lehneri<br />

P21,N2<br />

N5<br />

N4<br />

Gt. inflata<br />

Gt. Crassaformis<br />

Ss. Seminolina<br />

Gt. puncticulata/margaritae<br />

Gr. margaritae<br />

Ss. Seminolina acme<br />

Gr. humerosa<br />

8.30<br />

Gr. acostaensis<br />

Gr.menardii<br />

Gr.mayeri<br />

Gs. ruber<br />

Y a f o<br />

Upp.<br />

Ziqim<br />

<br />

Mavqiim Evap.<br />

Pattish Reef<br />

Lower Ziqim<br />

Submarine canyons fill<br />

Mid. Ziqim<br />

Gaza Sands<br />

(Turbidites)<br />

Eroded<br />

Upp.<br />

Ha’on<br />

SEA LEVEL<br />

Miller et al. 2005<br />

-50 0 50 m<br />

Haq et al., 1977, 1978<br />

0 50 100 m<br />

Ma<br />

5<br />

10<br />

15<br />

20<br />

25<br />

30<br />

35<br />

40<br />

Fig. 4.4 - Tertiary stratigraphy and marine onlaps, <strong>Israel</strong>.<br />

<strong>The</strong> colored bars denote seismic horizon.<br />

36


Rosh Pinna<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

10500 m<br />

<br />

0 5 10 km<br />

Top Basement<br />

revised edition<br />

10500 m<br />

Hiatus, erosional unconformity, fault Mid. Miocene<br />

Top Bathonian<br />

10000<br />

Undivided Permo-Triassic<br />

<br />

Formation contact<br />

Base Mavqiim<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

Top Turonian<br />

10000<br />

Sea floor<br />

Base Yafo<br />

Base Oligocene<br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

9500<br />

Basement<br />

9500<br />

9000<br />

<br />

Seismic Horizons<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geophysical Institute<br />

of <strong>Israel</strong><br />

LEGEND<br />

<br />

9000<br />

8500<br />

Undivided Permo-Triassic<br />

<br />

8500<br />

8000<br />

<br />

Undivided Early Jurassic<br />

Undivided Permo-Triassic<br />

<br />

<br />

Basement<br />

<br />

8000<br />

7500<br />

<br />

<br />

<br />

<br />

7500<br />

7000<br />

Undivided Early Jurassic<br />

<br />

Undivided Permo-Triassic<br />

<br />

7000<br />

Undivided Early Jurassic<br />

6500<br />

Barnea (Bajocian)<br />

Barnea (Bajocian)<br />

Undivided Early Jurassic<br />

<br />

<br />

6500<br />

Shederot (Bathonian)<br />

6093 m<br />

6000<br />

Karmia (Permian)<br />

6000<br />

Dorot Schist<br />

Shederot (Bathonian)<br />

Triassic<br />

Erez Cgl.<br />

U. Triassic<br />

Karmia Shale<br />

(U. Perm.-L. Triass.)<br />

Gevim Qz.Porph.<br />

184 m.y<br />

6000<br />

Kidod (Oxfordian)<br />

Zohar (Callovian)<br />

Barnea (Bajocian)<br />

Barnea<br />

Shederot (Bathonian)<br />

5500<br />

Delta (Kimmeridgian)<br />

Shederot<br />

5370 m<br />

Zohar (Callovian)<br />

5500<br />

5500<br />

5250 m<br />

Zohar (Callovian)<br />

<br />

Kidod (Oxfordian)<br />

Arsuf (Berriasian-Valanginian)<br />

Yam (Tithonian)<br />

5000<br />

Zohar<br />

Arsuf<br />

Yam<br />

Delta<br />

Kidod<br />

5000<br />

Kidod (Oxfordian)<br />

Delta (Kimmeridgian)<br />

5000<br />

Yam (Tithonian)<br />

Delta (Kimmeridgian)<br />

Mishhor<br />

5000<br />

Mohilla<br />

(Carnian)<br />

Erez Cgl.<br />

U. Triassic<br />

5000<br />

Arsuf (Berriasian-Valanginian)<br />

Yam (Tithonian)<br />

Arsuf (Berriasian-Valanginian)<br />

4500<br />

Mishhor<br />

Basement<br />

4500<br />

4620 m<br />

Sa’ad (L. Perm.)<br />

4500<br />

Gevar’am (Haut.-Aptian)<br />

4204 m<br />

4076 m<br />

Arqov (U. Permian)<br />

4000<br />

Gevar’am (Haut.-Aptian)<br />

4500<br />

4500<br />

37<br />

Sand<br />

4500<br />

Sa’ad<br />

Sand<br />

Arqov<br />

4000<br />

Sand<br />

Sand<br />

4000<br />

Gevar’am (Haut.-Aptian)<br />

4000<br />

4000<br />

Nirim<br />

(Pliensbach.<br />

-L. Bajoc.)<br />

Mohilla<br />

Arsuf<br />

Nirim<br />

(Pliensbachian-L. Bajocian)<br />

4000<br />

Saharonim (Ladinian)<br />

4000<br />

Saharonim<br />

4000<br />

Sand<br />

3500<br />

(Pliensbach.)<br />

(Pliensbach.)<br />

Sand<br />

TalmeYafe (Albian)<br />

En Zetim (Senonian) + Taqiye (Paleocene)<br />

Sand<br />

Talme Yafe (Albian)<br />

Barnea (Bajocian)<br />

Mishhor<br />

3500<br />

3500<br />

3500<br />

3500<br />

3500<br />

Ardon<br />

Ardon<br />

3500<br />

Talme Yafe (Albian)<br />

3500<br />

En Zetim (Senonian) +<br />

Taqiye (Paleocene)<br />

Lower Inmar<br />

Lower Inmar<br />

3000<br />

3000<br />

TalmeYafe (Albian)<br />

3000<br />

3000<br />

3000<br />

3000<br />

Bet Guvrin (U. Eocene - M. Miocene)<br />

En Zetim +<br />

Taqiye<br />

Talme Yafe<br />

Daya<br />

Qeren<br />

Barnea (Bajocian)<br />

Daya<br />

Qeren (Aalenian-L. Bajocian)<br />

3000<br />

Avedat Grp. (L.- M. Eocene)<br />

En Zetim (Senonian) + Taqiye (Paleocene)<br />

3000<br />

Avedat Grp. (L.- M. Eocene)<br />

Upper Inmar (Bajocian)<br />

Upper Inmar<br />

Avedat Grp. (L.- M. Eocene)<br />

2500<br />

Daya (Bajocian)<br />

Daya<br />

2500<br />

Bet Guvrin (U. Eocene - M. Miocene)<br />

2500<br />

2500<br />

2500<br />

2500<br />

2500<br />

Ziqim (M.-U. Miocene)<br />

T.Y. Conglomerate<br />

Shederot<br />

Mavqiim (Messinian)<br />

Ziqim (M.-U. Miocene)<br />

Mavqiim (Messinian)<br />

Ziqim (Mid.-U. Miocene)<br />

Bet Guvrin (U. Eocene - M. Miocene)<br />

Gevar’am (Haut.-Aptian)<br />

2500<br />

Shederot (Bathon.)<br />

Shederot<br />

2000<br />

2000<br />

2000<br />

2000<br />

2000<br />

2000<br />

Zohar<br />

Kidod (Oxf.)<br />

Zohar (Callovian)<br />

Karmon<br />

2000<br />

Kidod<br />

Zohar<br />

2000<br />

Beer Sheva<br />

Mavqiim (Messinian)<br />

Gevar’am<br />

1500<br />

1500<br />

1500<br />

1500<br />

1500<br />

Telamim<br />

Noa<br />

Talme Yafe (Albian)<br />

1500<br />

Telamim (U.<br />

Helez (U. Haut.-Barrem.)<br />

Barrem.-Aptian)<br />

Helez<br />

1500<br />

1500<br />

Yakhini<br />

Telamim<br />

Yakhini (U. Apt.- Albian)<br />

1000<br />

1000<br />

1000<br />

1000<br />

1000<br />

1000<br />

Yagur<br />

1000<br />

Yafo (Pliocene) Yafo (Pliocene)<br />

Yakhini<br />

1000<br />

Yagur (Albian)<br />

K.B. 26.0 m<br />

coast line<br />

K.B. 13.2 m<br />

K.B. 14.9 m<br />

K.B. 51.3 m<br />

K.B. 62.2 m<br />

R.T. 138.4 m<br />

GR Acoustic<br />

GR Acoustic<br />

Acoustic<br />

SP Acoustic<br />

SP Acoustic<br />

GR Acoustic<br />

A’ 0<br />

Datum m.s.l.<br />

0 m A<br />

Bet Guvrin<br />

Yafo (Pliocene)<br />

Adulam Adulam<br />

500<br />

500 m<br />

500<br />

500<br />

500<br />

500<br />

500 Yagur<br />

500<br />

Negba (U. Cenomanian)<br />

Dalyya (Turon.)<br />

En Zetim+Taqiye<br />

(Sen.-Pal.)<br />

Yam West 1<br />

Yam-2<br />

Ashqelon-2<br />

Talme Yafe-4<br />

Helez Deep-1A<br />

Gevim-1<br />

NW SE


(Tithonian)<br />

Negba<br />

0<br />

0<br />

0<br />

0<br />

0<br />

B’<br />

0<br />

500<br />

1000<br />

1500<br />

2000<br />

2500<br />

3000<br />

3500<br />

4000<br />

4500<br />

5000<br />

5500<br />

6000<br />

6500<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

revised edition<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

0 10 20 km<br />

Hiatus, erosional unconformity, fault Mid. Miocene<br />

Top Bathonian<br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

Formation contact<br />

Base Mavqiim<br />

Top Turonian<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geophysical Institute<br />

of <strong>Israel</strong><br />

Sea floor<br />

Base Yafo<br />

Base Oligocene<br />

Seismic Horizons<br />

LEGEND<br />

<br />

Mohilla<br />

Mohilla (Carnian)<br />

6531 m<br />

(Carnian)<br />

6000 m<br />

5787 m<br />

<br />

Shefayim<br />

Barnea<br />

<br />

<br />

(Norian)<br />

6000<br />

Mohilla (Carnian)<br />

Shefayim (Norian)<br />

Barnea<br />

Shefayim<br />

Shederot<br />

Barnea<br />

(Bajocian)<br />

Barnea<br />

(Bajocian)<br />

5500<br />

Undivided (Aalenian-Bathonian)<br />

Kidod Zohar<br />

5500<br />

Shederot<br />

Zohar (Callovian)<br />

<br />

Asher Volcanics<br />

5500<br />

5250 m<br />

Kidod<br />

(Oxfordian)<br />

Delta<br />

Delta<br />

Kidod<br />

Zohar<br />

Zohar<br />

Delta<br />

Kidod<br />

Zohar<br />

Kidod<br />

Shefayim (Norian)<br />

<br />

Shederot (Bathonian)<br />

Yam<br />

Shederot (Bathonian)<br />

5000<br />

Zohar (Callovian)<br />

5000<br />

Delta (Kimmeridgian)<br />

5000<br />

5000<br />

Arsuf<br />

Zohar (Callovian)<br />

Kidod (Oxfordian)<br />

Yam (Tithonian)<br />

Kidod (Oxfordian)<br />

Yam<br />

Arsuf<br />

(Berriasian-Valanginian)<br />

Arsuf<br />

Yam<br />

Delta<br />

Asher Volcanics<br />

Nirim<br />

Mohilla<br />

Yam<br />

Delta (Kimmeridgian)<br />

Delta<br />

(Kimmeridgian)<br />

4500<br />

4500<br />

4423 m<br />

4500<br />

Mohilla (Carnian)<br />

4500<br />

Yam<br />

Yam (Tithonian)<br />

Asher Volcanics<br />

Asher V.<br />

Arsuf<br />

Arsuf<br />

Arsuf (Berriasian-Valanginian)<br />

Shefayim<br />

(Norian)<br />

Arsuf (Berriasian-Valanginian)<br />

4000<br />

4000<br />

4000<br />

Shefayim<br />

4000<br />

Gevar’am<br />

(Hauterivian<br />

Sand<br />

4000<br />

Gevar’am<br />

(Hauterivian-Aptian)<br />

<br />

Gevar’am (Hauterivian-Aptian)<br />

-Aptian)<br />

Nirim<br />

Asher Volcanics<br />

Sand<br />

3500<br />

Sand<br />

3500<br />

3500<br />

Item (Cenomanian)<br />

3500<br />

Nirim<br />

3500<br />

En Zetim (Senonian) + Taqiye(Paleocene)<br />

Sand<br />

Talme Yafe<br />

Negba<br />

Talme Yafe<br />

(Albian)<br />

Talme Yafe<br />

T.Y.<br />

En Zetim+Ghareb<br />

+Taqiye<br />

3000<br />

3000<br />

3000<br />

Nirim<br />

3000<br />

<br />

<br />

Negba (U. Cenomanian)<br />

Negba<br />

Yaqum (U. Hauterivian-Barremian)<br />

Bet Guvrin (U. Eocene - M. Miocene)<br />

Dalyya (Turonian)<br />

Dalyya<br />

Avedat<br />

2500<br />

2500<br />

2500<br />

Yakhini (U. Aptian-Albian)<br />

Telamim (U. Barremian-Aptian)<br />

Haifa<br />

2500<br />

Mavqiim (Messinian)<br />

Item (Cenomanian)<br />

Gevar’am (Haut.-Aptian)<br />

Nirim<br />

(Aalenian-L. Bajocian)<br />

3000<br />

Talme Yafe (Albian)<br />

<br />

Ni<br />

Bet Guvrin (U. Eocene - M. Miocene)<br />

Sand<br />

(Bajocian-Callovian)<br />

(Bajocian-Callovian)<br />

Haifa<br />

Haifa<br />

38<br />

E.Z.+Taq.<br />

En Zetim+Ghareb+Taqiye<br />

Sand<br />

Avedat Grp. (L.- M. Eocene)<br />

Ziqim<br />

(Bajocian-Callovian)<br />

2500<br />

(M.-U. Miocene)<br />

Haifa<br />

Haifa<br />

2153 m<br />

Yagur (Albian)<br />

Mavqiim (Messinian)<br />

2020 m<br />

Ziqim (M. Miocene)<br />

2000<br />

2000<br />

En Zetim (Senon.)+Taqiye (Paleocene)<br />

Bet Guvrin<br />

2000<br />

Negba<br />

2000<br />

2000<br />

En -Zetim<br />

2000<br />

1800 m<br />

Negba<br />

Telamim<br />

Yaqum (Up. Haut.-Barrem.<br />

Yaqum<br />

YAFO (Pliocene)<br />

Ziqim (M. Miocene)<br />

1500<br />

1500<br />

1500<br />

Avedat (L.-M. Eocene)<br />

(Senonian)<br />

Bet Guvrin<br />

Telamim<br />

1500<br />

YAFO (Pliocene)<br />

En Zetim+Ghareb+Taqiye (Senonian-Paleocene)<br />

Yakhini<br />

(Up. Barrem.-Aptian)<br />

Yaq<br />

1500<br />

1500<br />

(Oligo.-L. Mio.)<br />

Ziqim<br />

Yagur (Albian)<br />

Yakhini<br />

1000<br />

1000<br />

1000<br />

Talme Yafe (Albian)<br />

(Up. Apt.-Albian)<br />

Te<br />

Mavqiim<br />

(debris<br />

flows)<br />

(U. Cenomanian)<br />

Dalyya<br />

1000<br />

1000<br />

Ya<br />

YAFO<br />

Yagur (Albian)<br />

1000<br />

Dalyya<br />

Talme Yafe<br />

T.Y.<br />

YAFO (Pliocene)<br />

500<br />

500<br />

Bet Guvrin<br />

(Oligo.-L. Mio.)<br />

500<br />

Negba<br />

500<br />

500<br />

(Pliocene)<br />

500<br />

YAFO<br />

(Pliocene)<br />

YAFO (Pliocene)<br />

Datum m.s.l.<br />

0<br />

K.B. 26.0 m<br />

GR Acoustic<br />

GR Acoustic<br />

K.B. 11.0 m<br />

GR Resisitivity<br />

SP Acoustic<br />

K.B. 26.0 m<br />

K.B. 12.7 m<br />

SP Acoustic<br />

K.B. 12.7 m<br />

K.B. 21.0 m<br />

GR Resistivity<br />

K.B. 21 m<br />

GR Acoustic<br />

B<br />

Yam West-1<br />

Yam Yafo-1<br />

Delta-1A<br />

Asher Atlit Deep- 1<br />

Foxtrot-1<br />

Qishon Yam-1<br />

Asher Yam-1<br />

S N


4.5.2 Late Middle Eocene to Late Eocene cycle<br />

In proximal areas of the northern Negev a major unconformity was found between the<br />

Middle Eocene chalks of the Maresha Formation and the hemi-pelagic marls of the Qeziot<br />

Formation (Fig. 4.4). Benjamini (1979, 1984) indicated that the maximum hiatus between the<br />

two formations spans the P10 to P14 biostratigraphic zones. <strong>The</strong> Qeziot and the Bet Guvrin (its<br />

equivalent in the coastal plain and the Shefela areas, Fig. 4.4) formations are included in the<br />

Saqiye Group (Gvirtzman, 1970.). <strong>The</strong> latter signify the "Upper Clastic division of Ball and Ball<br />

(1953). Occasional mass transported conglomerates are found especially in distal areas. In the<br />

Shefela, the Bet Guvrin Formation is of the Upper Eocene semiinvoluta zone (P15). <strong>The</strong> latest<br />

Eocene (P16) zone is missing. Siman-Tov (1991) reported on the rare occurrence of the latest<br />

Eocene T. cerroazulensis zone in Ofaqim and in Tel-Nagila. In other Upper Eocene locations<br />

this zone was truncated by the following major Early Oligocene unconformity. <strong>The</strong> Har Aqrav<br />

limestone (Fig. 4.4) which overlies the Qeziot Formation in the western Negev (Benjamini,<br />

1980) apparently represents the highstand progradation phase of the Late Eocene cycle.<br />

Markedly, the hemi-pelagic facies of the Late Eocene is primarily uniform, both inland and far<br />

offshore.<br />

Unlike other locations, the Eocene succession in the synclinorial area of Ramat Hagolan<br />

(Michelson and Lipson-Benitah, 1986) shows an uninterrupted Early to Late Eocene<br />

biostratigraphic succession from P9 to P15. <strong>The</strong>refore, Michelson and Lipson-Benitah (1986)<br />

regarded the Upper Eocene (P15) chalk of the Ramat Hagolan as the Maresha Formation (Fig.<br />

4.4).<br />

4.5.3 Latest Eocene - earliest Oligocene erosion phase<br />

A major unconformity is found between the Late Eocene and Early Oligocene<br />

depositional cycles (Fig. 4.4). It could have been related to eustatic sea-level fall at 34 m.y.<br />

(shown by Haq et al., 1988 and Miller et al., 2005). Local uplift movements associated with<br />

Syrian Arc contraction (see chapter 7) or with doming of the Arabian Shield prior to the rifting<br />

of the Red Sea (Garfunkel, 1988) further intensified the erosional processes. A regional<br />

unconformity surface (peneplain) developed in the Negev and the Judea Mountains (Picard,<br />

1951; Garfunkel and Horowitz, 1966). Zilberman (1992, 1993) named this feature “the upper<br />

erosion surface” and pointed out that it postdates Upper Eocene deposits and predates Miocene<br />

continental deposits of the Hazeva Formation. <strong>The</strong> presence of this erosion surface in the Judean<br />

hills indicates that the anticlinorial backbone of the country had already emerged at that time.<br />

39


Buchbinder et al. (2005) matched the first incision of canyons on the <strong>Levant</strong> margin to<br />

the Early/Late Oligocene boundary (Fig. 4.4). Druckman et al. (1995b) matched it to the pre P19<br />

– ampliapertura Zone (Fig. 4.4). <strong>The</strong> main incision in the Ashdod-1 well (Buchbinder et al.,<br />

2005, fig. 7) may be located at the bottom of the "Lower Sands", i.e., either between P19/P20, or<br />

within the P20, P21 zones (Fig. 4.4). Alternatively, it may be put at the base of the Lower<br />

Oligocene (P18) conglomerates. Obviously, the biostratigraphic resolution is not refined enough<br />

to determine the precise age of this erosion event because of intense contamination and<br />

redeposition in the cutting samples. However, for the sake of convenience and simplicity in<br />

interpretation of the seismic data the initial phase of the canyon's incision was placed at the base<br />

of the Oligocene sedimentary package, and this surface is represented as the Red horizon in the<br />

seismic profiles (Fig. 4.4, see chapter 5).<br />

4.5.4 Early Oligocene cycle<br />

Patchy outcrops of lowermost Oligocene sediments (Zone P18, Fig. 4.4) are scarce and<br />

are sparsely spread from the Golan to the northern Negev. <strong>The</strong>y are more common in boreholes<br />

along the Mediterranean coastal plain and offshore areas. <strong>The</strong>y invariably consist of deepwater<br />

pelagic, marly chalks, and their aerial distribution largely follows the distribution of the Upper<br />

Eocene deposits, except for their absence in the southern Negev (Buchbinder et al., 2005). <strong>The</strong>y<br />

thus indicate that most of <strong>Israel</strong> except the southern and central Negev was covered by sea in<br />

earliest Oligocene times.<br />

4.5.5. Late Rupelian-earlyChattian lowstand wedge and transported clastics<br />

Sediments of P19-P21 zones (ampliapertura to opima opima, Fig. 4.4) have a limited<br />

inland distribution, especially the P19 Zone. Both are of hemi-pelagic facies, however, sediments<br />

of the opima opima Zone penetrate more inland to the Lower Shefela region (Lakhish<br />

Formation), displaying slumped blocks, debris flows and occasional sandy calcareous turbidites<br />

(Lakhish outcrops). In the Lakhish Formation the mass transported sediments are covered by a<br />

few meters of in situ large-foraminifera limestone representing a lowstand prograding wedge<br />

(Buchbinder et al., 2005). More than 300 m of sand and conglomerates of this age were<br />

penetrated in the Ashdod wells in the coastal plain. It is most likely that coarse-grained clastics<br />

could have been transported several tens of kilometers further to the west into the <strong>Levant</strong> <strong>Basin</strong>.<br />

40


4.5.6 Late Oligocene (P22) – Early Miocene (N4) cycles<br />

<strong>The</strong>se cycles represent the last major inland transgressions (Fig. 4.4). <strong>The</strong>y are separated<br />

by an erosional unconformity of moderate magnitude. <strong>The</strong> onlap of the Early Miocene cycle<br />

reached across the rift valley to the Golan area (Fig. 4.4, upper Ha'on Member, Buchbinder et al.,<br />

2005). However, their hemi-pelagic sediments were largely eroded in the hilly backbone of the<br />

country in later periods. <strong>The</strong> facies is ubiquitously pelagic or hemi-pelagic (Bet Guvrin type),<br />

except in the most proximal outcrop in the Golan.<br />

4.5.7 <strong>The</strong> Burdigalian lowstand (N5-N7)<br />

Sediments of this interval are limited to the coastal plain and offshore area. Canyon<br />

incision was renewed and mass transported deposits occasionally accumulated, e.g., the Gaza<br />

sands (Fig. 4.4; Martinotti, 1973).<br />

4.5.8 Early Middle Miocene cycle (Ziqlag Formation)<br />

This cycle is unique in the sense that it shows a development of a carbonate platform<br />

(Ziqlag Formation, Fig. 4.4) unconformably abutting the western flank of the Judean anticline,<br />

in times of dominant muddy deposition (Buchbinder et al., 1993; Buchbinder and Zilberman,<br />

1997). However, its life span was short (ca 1.5 m.y.). <strong>The</strong> high sea level at that time facilitated<br />

the development of the carbonate platform by curbing siliciclastic supply. Hemi-pelagic<br />

deposition of the Bet Guvrin facies, however, persisted in the basin (Fig .4.4). On the other hand,<br />

a thick (250 m) succession of conglomerate, consisting of limestone, chalk and chert pebbles<br />

apparently derived from Cretaceous to Tertiary formations, was found in the Nahal Oz-1 well in<br />

the Afiq Canyon (Druckman et al., 1995b). Its distribution probably spread towards the offshore<br />

continuation of the Gaza Canyon.<br />

4.5.9 <strong>The</strong> Middle Miocene Seravallian crisis<br />

<strong>The</strong> Ziglag episode terminated abruptly when the sea level dropped and erosion and<br />

canyon incision resumed, leaving behind their products: conglomerates and overbank deposits<br />

(Bet Nir conglomerate, Buchbinder et al., 1986). This happened despite the relatively high global<br />

sea level. Apparently there is evidence for a climatic, ecological and salinity crisis in the<br />

Mediterranean during Serravallian times, and as a result, the Mediterranean sea-level curve<br />

departed from that of the global sea level (see Martinotti, 1990; Buchbinder et al., 1993). This<br />

lowstand episode produced a distinct seismic reflector in the basin, namely the Cyan horizon<br />

(Fig. 4.4, chapter 5). Small onlap episodes of limited landward penetration are indicated by the<br />

lower and middle Ziqim "mini" cycles (Fig. 4.4).<br />

41


4.5.10 Late Miocene – Tortonian to Messinian cycle (N16-N17)<br />

Although of limited inland penetration, sediments of this cycle are relatively common in<br />

outcrops (Buchbinder and Zilberman, 1997) and especially in boreholes in the coastal plain (e.g.,<br />

the Ashdod wells, Fig. 4.4). <strong>The</strong> cycle is marked by the development of coral (mostly Porites)<br />

and algal reefs of the Pattish Formation (Fig. 4.4). However, in the offshore extension of the<br />

Afiq (Gaza) Canyon, hemi-pelagic muds are found, interbedded with mass transported sands,<br />

conglomerates of chert pebbles.<br />

4.5.11 <strong>The</strong> Messinian Mavqiim cycle<br />

<strong>The</strong> Messinian cycle is composed of a thick evaporitic series found throughout the<br />

Mediterranean region (Hsu et al., 1973a,b; Neev et al., 1976; Ryan, 1978). <strong>The</strong> deposition of<br />

Mavqiim evaporites (Fig. 4.4), which are composed mainly of halite and some anhydrite,<br />

followed a major sea-level drop associated with the closure and desiccation of the Mediterranean<br />

Sea (Hsu et al.1973a,b; Hsu et al., 1978, Gvirtzman and Buchbinder, 1978). <strong>The</strong> area of<br />

evaporite accumulation extended throughout the <strong>Levant</strong> basin to about 20-40 km west of the<br />

present-day coastline (Gardosh and Druckman, 2006). <strong>The</strong> Messinian evaporites most likely<br />

filled a pre-existing topographic depression (Tibor et al., 1992) and are not associated with<br />

significant, local tectonic movements. However, during short highstand episodes the brine<br />

covered the slope and penetrated through deep canyons further inland. Remnants of the<br />

Messinian brine in the form of a few tens of meters of thick halite and anhydrite beds were<br />

encountered in various wells in the coastal areas of <strong>Israel</strong> (Druckman et al., 1995b). On the<br />

seismic data the base of the Messinian cycle is marked by the Purple horizon (Fig. 4.4)<br />

4.5.12 Plio-Pleistocene cycles<br />

<strong>The</strong> abrupt sea-level rise in the Early Pliocene terminated the evaporite deposition and<br />

resulted in the deposition of hemi-pelagic clays and marls of the Plio-Pleistocene Yafo<br />

Formation (Fig. 4.4) reaching over 1000 m in thickness. <strong>The</strong> unconformity surface at the top of<br />

the Mavqiim evaporites and the base of the Yafo Formation is marked on the seismic data by the<br />

Violet marker (also known as the M seismic marker; Neev et al. 1976) (Fig .4.4, chapter 5).<br />

Sedimentation during the lowermost Pliocene succession was highly condensed. After<br />

deposition of a few tens of meters of hemi-pelagic marl in a period of ca 1 m.y., a significant<br />

sea-level drop occurred, estimated to correspond with the 4.37 m.y. sequence boundary of<br />

Hardenbol et al. (1998) (correlated with Haq et al.'s (1988) 4.2 m.y. sequence boundary). This<br />

lowstand resulted in the deposition of gas-bearing sands of turbidite origin in a basin floor or<br />

lower slope setting (Druckman, 2001; Oats, 2001). On land, this lowstand event is correlated<br />

42


with a boulder conglomerate separating between the Pliocene sequences I and II of Buchbinder<br />

and Zilberman (1997) in the Beer Sheva area.<br />

Hemi-pelagic deposition was resumed during the following highstand and the rate of<br />

sedimentation gradually increased, eventually resulting in a progradational topset-clinoform<br />

pattern, clearly portrayed in the seismic sections (Figs. 5.1, 5.3).<br />

<strong>The</strong> next significant events expressed in the seismic records are chaotic intervals of<br />

slump deposits of Late Pliocene and Pleistocene times (Frey-Martinez et al., 2005). Most notable<br />

is the large scale-slumping zone, ca 150 m thick, of Late Pliocene times, which may also be<br />

related to a lowstand phase.<br />

5. Interpretation of Seismic Data<br />

5.1 Seismic Horizons<br />

<strong>The</strong> regional, 2D seismic profiles of the A and B surveys show a continuous series of<br />

seismic reflection that is 5 to 6 sec thick near the <strong>Israel</strong>i coast and reaches up to 10 sec (15-17<br />

km) in the far offshore (Figs. 5.1, 5.3, 5.8, 5.9). This series comprises the entire sedimentary<br />

rock section that accumulated and was preserved within the <strong>Levant</strong> <strong>Basin</strong> during the<br />

Phanerozoic. <strong>The</strong> Phanerozoic basin-fill is subdivided into eight distinct seismic units that are<br />

separated by seven seismic horizons (Fig. 5.1, Table 5.1). Each of these horizons is interpreted<br />

as an unconformity surface that bounds a regional, basin-scale depositional unit. All the<br />

interpreted surfaces mark significant changes in seismic character within the basin-fill (Fig. 5.1).<br />

Some are recognized also by truncation or onlapping and downlapping reflections (Figs. 5.1,<br />

5.3). <strong>The</strong> interpreted seismic markers extend throughout the entire basin; therefore they are not<br />

correlated to individual seismic reflections but rather to the tops (or bottoms) of separable<br />

seismic packages.<br />

Well control on the seismic interpretation is possible only in the relatively shallow,<br />

eastern part of the basin. In this area the horizons show good fit to the tops of lithostratigraphic<br />

units identified in the wells (Figs. 5.1, 5.2). <strong>The</strong> correlation of seismic units from the shallow to<br />

the deep part of the basin, particularly for the deeper Brown, Blue, Green and Red horizons, is<br />

hampered by extensive faulting and tilting and is therefore somewhat speculative.<br />

43


44<br />

W<br />

Yam West-1 (proj. 2km SW) Yam-2 E<br />

AC. AC.<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-L. Miocene<br />

TWT (ms)<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

<br />

Permian- M. Jurassic<br />

<br />

Precambrian ()<br />

Crystalline<br />

Basement<br />

5 km<br />

Fig 5.1- Interpreted seismic profile showing the correlation of seismic horizons to<br />

the Yam-2 and Yam-West-1 wells, in the southeastern part of the basin.<br />

<strong>The</strong> lithologic column below the TD of the Yam West-1 well is inferred.


Yam Yafo-1 Yam West-1<br />

AC.<br />

AC.<br />

TWT(ms)<br />

Fig 5.2 - Synthetic seismograms of Yam West-1 and Yam Yafo-1 showing the correlation of the seismic reflections to<br />

lithostratigraphic tops in the wells. <strong>The</strong> synthetic seismograms were prepared from acoustic (AC.) and density logs,<br />

and a zero-phase, normal 25 Hz Ricker wavelet<br />

45


Seismic Horizon Chronostratigraphic Top Lithostratigraphic Top<br />

Violet Late Messinian (Late Miocene) Mavqiim (M horizon)<br />

Purple Early Messinian (Late Miocene) Ziqim/Ziqlag/AshdodClst.(N horizon)<br />

Cyan Langhian (early Middle Miocene) Bet Guvrin/Lakhish<br />

Red Eocene Avedat<br />

Green Albian to Turonian Talme Yafe/Yagur/Negba/Daliyya<br />

Blue Bathonian Shederot/ within 'undivided' M. Jurassic<br />

Brown Precambrian Crystalline basement<br />

Table 5.1- Correlation of interpreted seismic horizons to chrono- and lithostratigraphic unit tops.<br />

5.1.1 <strong>The</strong> Brown horizon<br />

<strong>The</strong> brown horizon is the deepest seismic marker interpreted in this study. It is correlated<br />

throughout the basin with the transition from a chaotic and reflection-free seismic character<br />

below to a continuous, parallel to divergent, high-amplitude reflection series above (Fig. 5.1).<br />

No direct well control is possible for this deep marker in the offshore. We interpret the transition<br />

as characterizing a major acoustic boundary between non-layered, magmatic and metamorphic<br />

complexes and the overlying, layered Paleozoic-Mesozoic sedimentary interval. <strong>The</strong> Brown<br />

horizon is therefore considered as the 'near top of the crystalline basement' of an assumed<br />

Precambrian age.<br />

<strong>The</strong> transition in seismic character is more evident in the southeastern corner of the study<br />

area, near the Yam West-1 well, where it is found at a relatively shallow depth of about 5-6 sec<br />

(Figs. 5.1, 5.3). In this area the seismic boundary possibly correlates to the top of the<br />

Precambrian to Infrcambrian basement rocks, penetrated by the Helez Deep-1 and the Gevim-1<br />

wells, located about 15 km east of the coastline (Fig. 1.3). <strong>The</strong> resolution of the seismic data in<br />

the offshore generally decreases below 7-8 sec due the reduction of the seismic energy.<br />

However, the transition between the chaotic seismic character and the overlying, high-amplitude<br />

partly continuous reflection series is identifiable at the bottom of many seismic profiles that<br />

cross the central, deep part of the basin (Figs. 5.8, 5.9).<br />

5.1.2 <strong>The</strong> Blue horizon<br />

<strong>The</strong> Blue horizon is correlated to a distinct change in seismic character from highamplitude;<br />

continuous reflection series below to discontinuous, low-amplitude occasionally<br />

shingled reflections above (Fig. 5.1 at 4500 ms between Yam-2 and Yam West-1 wells). <strong>The</strong><br />

change in seismic character is most evident in the southeastern area, though it is observed also in<br />

46


TWT(ms)<br />

W<br />

AC. AC.<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Precambrian ()<br />

Crystalline Basement<br />

Fig 5.3- Interpreted, E-W oriented seismic profile through the Yam-2 and Yam West-1 wells,<br />

showing the area of transition from the elevated, Mesozoic thrust belt on the east<br />

to the deep, Tertiary basin on the west.<br />

E<br />

47


48<br />

W E S N W<br />

E<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

TWT(ms)<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Precambrian ()<br />

Crystalline Basement<br />

Fig 5.4- Interpreted, composite seismic profile, generally oriented in a N-S direction along the shallow<br />

part of the basin, showing the continuity of seismic horizons from Yam West-1 (south) to<br />

Yam Yafo-1(north).


S N<br />

Plio-Pleistocene<br />

Messinian<br />

TWT(ms)<br />

M.-U. Miocene<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Precambrian ()<br />

Crystalline Basement<br />

Fig 5.5- Interpreted seismic profile, oriented in a NE-SW direction showing the correlation of<br />

seismic horizons in the deep part of the basin.<br />

49


the central and western parts of the basin where the horizon is found at the top of a relatively<br />

continuous high-amplitude seismic event below less continuous and lower amplitude reflections<br />

(Figs. 5.3, 5.4, 5.5, 5.8, 5.9, 5.11).<br />

<strong>The</strong> correlation of the Blue horizon from Yam West-1 to Yam Yafo-1 presents some<br />

difficulty. At Yam West-1 well it is correlated to a high-amplitude event found at the lower part<br />

of the well near the top of the Bathonian Shederot Formation (Fig. 5.2). At the Yam Yafo-1 well<br />

it is correlated to a high-amplitude event within an undivided Jurassic section of poorly defined<br />

age (probably Aalenian to Bathonian) (Fig. 5.2). Furthermore, there is a significant difference<br />

between the two wells in the lithology of the Middle Jurassic section. It is therefore unclear<br />

whether the seismic event represents the same chronostratigraphic unit in this area.<br />

<strong>The</strong> change in character below and above the Blue horizon is, however, identifiable and<br />

most likely signifies a basin-wide seismic boundary. According to our stratigraphic analysis of<br />

the Jurassic section (chapter 4.3) the Blue horizon may correlate to the transition from a Lower<br />

Jurassic, shallow-marine carbonate section to a Middle-Upper Jurassic deep-marine section that<br />

was deposited during the passive margin stage.<br />

5.1.3 <strong>The</strong> Green horizon<br />

<strong>The</strong> Green horizon corresponds to the top of a distinct high-amplitude reflection package<br />

below lower-amplitude reflections, chaotic and reflection-free zones (Figs. 5.1 to 5.8). At the<br />

eastern part of the basin the marker is an onlapping surface (Fig. 5.1). Well control shows that<br />

the Green horizon correlates to the tops of several Middle Cretaceous units. <strong>The</strong>se are the<br />

Albian Talme Yafe and Yagur formations, the Cenomanian Negba Formation and the Turonian<br />

Daliyya Formation (Fig. 5.2) (See also Table 3.1). In most wells of the study area the marker is<br />

overlain by strata of the Upper Cretaceous Mount Scopus Group (Figs. 5.1, 5.2, Table 3.1).<br />

Throughout inland <strong>Israel</strong> a Middle Cretaceous unconformity is evident by a depositional<br />

hiatus and an abrupt lithologic change between shelf and slope carbonates of Albian to Turonian<br />

age (Judea and Talme Yafe formations) and pelagic marl and chalk of Senonian to Paleogene age<br />

(Mount Scopus and Hashefela groups) (Gvirtzman and Reiss 1965; Flexer 1968, Flexer et al.,<br />

1986). <strong>The</strong> change in seismic character corresponds to this well-recognized regional surface that<br />

is associated with drowning of the Middle Cretaceous carbonate platform (Sass and Bein, 1982;<br />

Almogi-Labin et al., 1993).<br />

<strong>The</strong> correlation of the shallow, eastern part of the basin to the deep, western part requires<br />

a certain degree of jump-correlation across conspicuous sets of reverse faults (Figs. 5.3, 5.8). No<br />

well control is available in the far offshore. However, the identification of the Green marker at<br />

the top of a continuous, high-amplitude reflection series in the deep basin (Fig. 5.5) is well<br />

50


NW SE<br />

AC.<br />

Plio-Pleistocene<br />

M.-U. Miocene<br />

Messinian<br />

TWT(ms)<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Precambrian ()<br />

Crystalline Basement<br />

Fig 5.6- Interpreted seismic profile showing the transition area from the elevated<br />

Mesozoic thrust belt to the deep Tertiary basin. <strong>The</strong> Red horizon is a base<br />

Oligocene unconformity (correlated to the Yam West-2 well) that truncates the<br />

Senonian-Eocene section at the tops of Syrian Arc fold structures.<br />

51


52<br />

NW SE SW<br />

NE<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

TWT(ms)<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Fig 5.7- Interpreted composite seismic profile through the Hannah-1 well. <strong>The</strong> Red horizon<br />

is a base Oligocene unconformity that truncates the Cretaceous, Senonian and Eocene section.<br />

In the Hannah-1 well the Oligo-Miocene fill overlies the Lower Cretaceous Gevar’am Formation.


NE SW<br />

AC.<br />

Plio-Pleistocene<br />

TWT(ms)<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Precambrian ()<br />

Crystalline Basement<br />

Jonah<br />

Ridge<br />

Fig 5.8- Interpreted, NE-SW regional seismic profile showing extensional and<br />

contractional structures in the central part of the basin.<br />

53


supported by the overall continuous seismic character of this unit as well as by the assumed<br />

vertical offsets on the reverse faults.<br />

5.1.4 <strong>The</strong> Red horizon<br />

<strong>The</strong> Red horizon is an unconformity surface that is characterized throughout the eastern<br />

part of the basin by truncation, incision and onlapping (Figs. 5.3, 5.6, 5.7). Well control shows<br />

that the Red horizon correlates to an Early Oligocene unconformity at the top of the Eocene<br />

Avedat Group or older stratigraphic units (Fig. 5.2, Table 3.1). In most of the wells of the study<br />

area the marker is overlain by strata of the Oligo-Miocene Saqiye Group.<br />

<strong>The</strong> correlation of seismic profiles to the Yam West-2 and Hannah-1 well further<br />

demonstrate the amount of erosion associated with this seismic marker. In Yam West -1 the Red<br />

horizon is found at the top of an eroded, 30 m thick Eocene section (Fig. 5.6). At the Hannah-1<br />

well the marker is correlated to the base of a deep canyon, incised within the Lower Cretaceous<br />

Gevar'am Formation, filled with Oligocene to Lower Miocene strata (Fig. 5.7). From these two<br />

wells the marker is carried across the top of several eroded structures down to the deep basin,<br />

where it merges with a series of high-amplitude, continuous seismic reflections at a depth of<br />

about 5-6 sec (Figs. 5.6, 5.7).<br />

It should be noted that the present interpretation of the Red horizon within the deep basin<br />

is a revision of the results presented by Gardosh and Druckman (2006). This revision was made<br />

following the integration of the critical information from the Hannah-1 well that was not<br />

available at the time of the previous interpretation. <strong>The</strong> revised interpretation suggests that a<br />

significant part of the basin-fill is younger than was assumed.<br />

<strong>The</strong> base Saqiye unconformity is recognized throughout the inland part of <strong>Israel</strong>. It is<br />

associated with an Oligocene uplift event that resulted in erosion and the development of an<br />

extensive drainage system and transport of sediments into the <strong>Levant</strong> <strong>Basin</strong> (see chapter 4.5).<br />

<strong>The</strong> incision and truncation on the <strong>Levant</strong> slope associated with the Red horizon comforms with<br />

the Early Oligocene paleogeography. <strong>The</strong> conformable, high-amplitude Red event within the<br />

basin is probably a condensed section at the top of the Upper Cretaceous to Eocene strata.<br />

5.1.5 <strong>The</strong> Cyan horizon<br />

<strong>The</strong> Cyan horizon is an unconformity surface that is characterized by a minor degree of<br />

truncation, incision and onlapping (Figs. 5.1, 5.3, 5.6). At the Yam West-1 and Yam Yafo-1<br />

wells the marker is a low-amplitude event that is correlated to the top of the Oligocene-Lower<br />

Miocene Bet Guvrin Formation (Fig. 5.2, Table 3.1). <strong>The</strong> correlation is carried from the wells to<br />

the deep part of the basin where the Cyan horizon is identified as the top of several continuous,<br />

54


NW SE<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

TWT(ms)<br />

Oligocene-L. Miocene<br />

M. Jurassic-Turonian<br />

Senonian-Eocene<br />

Permian- M. Jurassic<br />

Precambrian ()<br />

Crystalline Basement<br />

Jonah<br />

Ridge<br />

Leviathan<br />

High<br />

Eratosthenes<br />

High<br />

Fig 5.9- Interpreted, NW-SE seismic profile from the Eratosthenes Seamount to the <strong>Levant</strong> margin.<br />

<strong>The</strong> profile shows a series of horst and graben structures, associated with Early Mesozoic rifting.<br />

55


56<br />

W E<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

TWT(ms)<br />

Oligocene-L. Miocene<br />

M. Jurassic-Turonian<br />

Senonian-Eocene<br />

Permian- M. Jurassic<br />

Jonah<br />

Ridge<br />

Carmel- Caesarea<br />

Block<br />

Fig 5.10- Interpreted E-W seismic profile showing extensional and contractional structures in the<br />

northeastern part of the basin. <strong>The</strong> bordering fault of the Carmel-Caesarea block may<br />

have originated as a normal fault during the Early Jurassic and was reactivated in a<br />

reverse motion during the Tertiary.


SW NE<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

TWT(ms)<br />

Oligocene-L. Miocene<br />

c<br />

b<br />

a<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

Jonah<br />

Ridge<br />

Fig 5.11- Interpreted SW-NE seismic profile showing the southern part of the Jonah Ridge. <strong>The</strong> ridge<br />

is composed of three units: (a) an Early Jurassic horst block, (b) a chaotic seismic package<br />

interpreted as a volcanic cone with highly dipping slopes, of Late Cretaceous to Early<br />

Tertiary age and, (c) a series of parallel, high-amplitude reflections interpreted as a carbonate<br />

buildup of Lower Miocene age.<br />

57


high amplitude reflections below less reflective zones at a depth of about 4 to 5 sec (Figs. 5.5,<br />

5.8, 5.9). In the deep basin the Cyan event is probably a condensed section at the top of the<br />

Oligocene to Lower Miocene strata.<br />

5.1.6 <strong>The</strong> Purple horizon<br />

<strong>The</strong> Purple horizon is the base of the Messinian evaporites. In wells at the eastern part of<br />

the basin it is correlated to a high-amplitude event near the top of the Ziqim Formation (Fig 5.2).<br />

Throughout the deep basin it is clearly recognized by the conspicuous change in seismic<br />

character from the underlying finely layered Middle-Upper Miocene strata to the overlying<br />

chaotic salt layer (Figs. 5.5, 5.8, 5.9, 5.10). <strong>The</strong> base Messinian unconformity, also known as the<br />

N horizon, is an Upper Miocene erosion surface associated with the huge drop of the<br />

Mediterranean sea-level and subsequent deposition of the Messinian evaporites (Hsu et al.,<br />

1973a,b; Finetti and Morelli, 1973; Ryan, 1978).<br />

5.1.7 <strong>The</strong> Violet horizon<br />

<strong>The</strong> Violet horizon is the top of the Messinian evaporites. In the deep basin the surface is<br />

clearly recognized by the conspicuous change in seismic character from the underlying chaotic<br />

salt layer to the finely layered Plio-Pleistocene strata (Figs. 5.5, 5.7, 5.8, 5.9). In the eastern part<br />

of the basin it is correlated to a high-amplitude event near the top of the Mavqiim Formation or<br />

to an erosion surface at the base of the Plio-Pleistocene section. In this area the Messinian<br />

evaporite layer is either very thin or completely missing and the Purple and Violet horizons<br />

merge (Figs. 5.3, 5.6). <strong>The</strong> top Messinian unconformity, also known as the M horizon, reflects a<br />

major lithofacies boundary associated with marine transgression and deposition of the hemipelagicYafo<br />

Formation.<br />

5.2 Time and Depth Maps<br />

<strong>The</strong> seven seismic markers described above were interpreted on the seismic profiles of<br />

the A and B surveys throughout the study area (Fig. 1.3). A set of two-way-time maps was<br />

initially produced. <strong>The</strong>se were converted to depth maps, layer by layer (layered cake method),<br />

through the multiplication of time and interval velocity grids (Figs. 5.12 to 5.19).<br />

<strong>The</strong> reliability of time-to-depth conversion is commonly dependent on the quality of the<br />

velocity data. <strong>The</strong> main source of velocity information is check-shot surveys measured in wells.<br />

Well information in the study area is limited to the eastern part of the basin. <strong>The</strong>refore, in a<br />

previous study Gardosh and Druckman (2006) applied a simplistic approach, using single<br />

interval velocity values.<br />

58


3480000 3500000 3520000 3540000 3560000 3580000 3600000 3620000 3640000 3660000 3680000<br />

N<br />

Legend<br />

Seismic Lines<br />

GEV 1<br />

Cross sections<br />

Well (used for time & depth calibration)<br />

Well (used for depth calibration)<br />

Fault<br />

A’<br />

Depth Contours (m)<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geoph ysical Institute<br />

of <strong>Israel</strong><br />

B’<br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

revised edition<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

0 50 Km<br />

A<br />

460000 480000 500000 520000 540000 560000 580000 600000 620000 640000 660000 680000<br />

Fig 5.13- Depth map of the Middle Jurassic, Blue horizon (meters below MSL).<br />

B<br />

0<br />

(m)<br />

600<br />

2130<br />

2840<br />

3550<br />

4260<br />

4970<br />

5680<br />

6390<br />

7100<br />

7810<br />

8520<br />

9230<br />

9940<br />

700000<br />

60


460000 480000 500000 520000 540000 560000 580000 600000 620000 640000 660000 680000<br />

A<br />

B<br />

N<br />

0 (m)<br />

670<br />

1340<br />

2010<br />

2680<br />

3350<br />

4020<br />

4690<br />

5360<br />

6030<br />

6700<br />

7370<br />

8040<br />

700000<br />

61<br />

Legend<br />

Seismic Lines<br />

Cross sections<br />

GEV 1<br />

Well (used for time & depth calibration)<br />

Well (used for depth calibration)<br />

Fault<br />

Depth Contours (m)<br />

A’<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geoph ysical Institute<br />

of <strong>Israel</strong><br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

revised edition<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

0 50 Km<br />

B’<br />

3480000 3500000 3520000 3540000 3560000 3580000 3600000 3620000 3640000 3660000 3680000<br />

Fig 5.14- Depth map of the Middle Cretaceous, Green horizon (meters below MSL).


B<br />

N<br />

0 (m)<br />

600<br />

1200<br />

4800<br />

3480000 3500000 3520000 3540000 3560000 3580000 3600000 3620000 3640000 3660000 3680000<br />

62<br />

Legend<br />

1800<br />

Seismic Lines<br />

GEV 1<br />

Cross sections<br />

Well (used for time & depth calibration)<br />

2400<br />

3000<br />

Well (used for depth calibration)<br />

Fault<br />

A’<br />

3600<br />

Depth Contours (m)<br />

4200<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geoph ysical Institute<br />

of <strong>Israel</strong><br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

revised edition<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

0 50 Km<br />

B’<br />

A<br />

5400<br />

6000<br />

6600<br />

7200<br />

7800<br />

460000 480000 500000 520000 540000 560000 580000 600000 620000 640000 660000 680000<br />

Fig. 5.15- Depth map of the base Oligocene, Red horizon (meters below MSL).<br />

700000


B<br />

N<br />

0 (m)<br />

440<br />

3520<br />

3480000 3500000 3520000 3540000 3560000 3580000 3600000 3620000 3640000 3660000 3680000<br />

63<br />

Legend<br />

880<br />

Seismic Lines<br />

1320<br />

GEV 1<br />

Cross sections<br />

Well (used for time & depth calibration)<br />

1760<br />

Well (used for depth calibration)<br />

Fault<br />

A’<br />

2200<br />

2640<br />

Depth Contours (m)<br />

3080<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geoph ysical Institute<br />

of <strong>Israel</strong><br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

revised edition<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

0 50 Km<br />

B’<br />

A<br />

3960<br />

4400<br />

4840<br />

5280<br />

460000 480000 500000 520000 540000 560000 580000 600000 620000 640000 660000 680000<br />

Fig 5.16- Depth map of the top Lower Miocene, Cyan Horizon (meters below MSL).<br />

700000


460000 480000 500000 520000 540000 560000 580000 600000 620000 640000 660000 680000<br />

B<br />

N<br />

350<br />

(m)<br />

2800<br />

700000<br />

3480000 3500000 3520000 3540000 3560000 3580000 3600000 3620000 3640000 3660000 3680000<br />

64<br />

0<br />

Legend<br />

700<br />

Seismic Lines<br />

1050<br />

GEV 1<br />

Cross sections<br />

Well (used for time & depth calibration)<br />

1400<br />

Well (used for depth calibration)<br />

Fault<br />

A’<br />

1750<br />

2100<br />

Depth Contours (m)<br />

2450<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geoph ysical Institute<br />

of <strong>Israel</strong><br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

revised edition<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

0 50 Km<br />

B’<br />

A<br />

3150<br />

3500<br />

3850<br />

4200<br />

Fig 5.17- Depth map of the base Messinian, Purple horizon (meters below MSL).


460000 480000 500000 520000 540000 560000 580000 600000 620000 640000 660000 680000<br />

B<br />

N<br />

0 (m)<br />

210<br />

1470<br />

1680<br />

700000<br />

3480000 3500000 3520000 3540000 3560000 3580000 3600000 3620000 3640000 3660000 3680000<br />

65<br />

Legend<br />

420<br />

Seismic Lines<br />

630<br />

GEV 1<br />

Cross sections<br />

Well (used for time & depth calibration)<br />

840<br />

Well (used for depth calibration)<br />

Fault<br />

A’<br />

1050<br />

Depth Contours (m)<br />

1260<br />

<strong>The</strong> Ministry of National Infrastructures<br />

Geological Survey of <strong>Israel</strong><br />

<strong>The</strong> Geoph ysical Institute<br />

of <strong>Israel</strong><br />

Gardosh M., Druckman Y., Buchbinder B. and Rybakov M.,<br />

<strong>The</strong> <strong>Levant</strong> <strong>Basin</strong> <strong>Offshore</strong> <strong>Israel</strong>:<br />

<strong>Stratigraphy</strong>, <strong>Structure</strong>,<strong>Tectonic</strong> Evolution<br />

and Implications for Hydrocarbon Exploration<br />

revised edition<br />

Report GSI/4/2008, GII Report 429/328/08<br />

Jerusalem, April 2008<br />

0 50 Km<br />

B’<br />

A<br />

1890<br />

2100<br />

2310<br />

2520<br />

Fig 5.18- Depth map of the top Messinan, Violet Horizon (meters below MSL).


To improve the velocity control, in the present study we used two additional sources of<br />

velocity information: (a) stacking velocities that were extracted from the time processing of<br />

selected, regional lines from the A survey; and (b) horizon velocity analysis performed during<br />

the process of Pre-Stack Depth Migration of four lines. An interval velocity map was produced<br />

for each of the interpreted seismic horizons by integrating these velocities with the available<br />

check-shot data from the wells. Most of the interval velocity maps show east to west variations<br />

that are associated with increased depth of burial and degree of compaction from the margin into<br />

the basin. <strong>The</strong> resulting depth maps reflect these lateral velocity variations and are therefore<br />

considered more reliable than the previous versions of Gardosh and Druckman (2006). <strong>The</strong><br />

depth maps were further controlled by the correlation of the various depth grids to the<br />

corresponding lithostratigraphic units in the 43 wells located within the study area (Fig. 1.3,<br />

Table 3.1).<br />

6. Interpretation of Gravity and Magnetic Data<br />

6.1 Gravity Maps<br />

<strong>The</strong> raw gravity data measured during the acquisition of the B seismic survey were<br />

filtered and corrected for Eotvos, Free Air and Bouguer correction. <strong>The</strong> data were incorporated<br />

in the regional grid of the <strong>Levant</strong> area previously compiled by Rybakov et al. (1997). <strong>The</strong><br />

combined Bouguer gravity map (Fig. 6.1a) shows a good fit with the new and old data sets.<br />

Negligible inconsistencies observed close to the tie line (white polygon) are explained by the<br />

poor gravity coverage of the B survey area.<br />

<strong>The</strong> Bouguer gravity map (Fig. 6.1a) shows that a significant part of the area is occupied<br />

by high gravity values that reach up to 125 mgal. <strong>The</strong> composition of the crust underlying the<br />

Mediterranean is the subject of much discussion. Makris and Wang (1994) suggest an oceanic<br />

crust using seismic refraction, gravity and magnetic data, while Knipper and Sharaskin (1994)<br />

suggest that the crust is continental, based on their analysis of the tectonic history of the region.<br />

<strong>The</strong> high Bouguer gravity values are in agreement with the hypothesis that the crust underlying<br />

this part of the <strong>Levant</strong> <strong>Basin</strong> is dense. It is therefore either oceanic or intermediate in<br />

composition. Negative gravity values are typical for the continental crust of the African and<br />

Arabian plates.<br />

Within the positive Bouguer gravity province a local gravity high is noticeable in the<br />

northern part of the basin (Fig. 6.1a). It should be noted that this gravity high was not highlighted<br />

in the previous gravity maps (Makris and Wang, 1994), probably due to poor gravity coverage in<br />

67


Y<br />

X<br />

a<br />

b<br />

Fig 6.1- Revised gravity anomaly maps of the <strong>Levant</strong> <strong>Basin</strong>. New gravity data set acquired<br />

in the area of Survey B (white polygon) is compiled with the gravity map of the<br />

<strong>Levant</strong> region (Rybakov et al.,1997): (a) Bouguer gravity, (b) Residual Bouguer<br />

gravity-second order polynomial. X-Y is the modeled section in Fig. 6.3.<br />

68


Y<br />

X<br />

a<br />

b<br />

Fig 6.2- Revised magnetic anomaly maps of the <strong>Levant</strong> <strong>Basin</strong>. New magnetic data set acquired<br />

in the area of Survey B (white polygon) is compiled with the magnetic map of the<br />

<strong>Levant</strong> region (Rybakov et al.,1997): (a) total magnetic intensity, (b) total magnetic<br />

intensity after reduction to pole. X-Y is the modeled section in Fig. 6.3.<br />

69


this area. <strong>The</strong> positive province appears to be bounded to the SW and SE by two branches of<br />

relatively reduced gravity (zero Bouguer gravity close to the southern coastal plain of <strong>Israel</strong>)<br />

(Fig. 6.1a). <strong>The</strong> wide range of gravity values is probably associated with crustal changes;<br />

however the presence of the thick, light density sedimentary prism of the Nile River is an<br />

additional factor for the low gravity province in the south. <strong>The</strong> gravity steps and larger horizontal<br />

gradients in the southeastern part of the basin appear to represent a more complex pattern of<br />

near-surface density contrasts than in the southwestern area.<br />

<strong>The</strong> effect of the shallower anomalies was enhanced by removing the regional deep trend<br />

up to the third order polynomial. <strong>The</strong> second order polynomial, shown in Figure 6.1b, is almost<br />

free of the deep crustal heterogeneities and shows the distribution of several high and low<br />

density blocks. <strong>The</strong> northern anomaly is further enhanced and is interpreted as a NE-SW<br />

trending, high density, shallow crustal block. It is limited to the SW by a gravity low, possibly<br />

associated with the effect of the light Nile prism. An additional more complex, but generally<br />

NE-SW trending, shallow crustal block is interpreted in the southern part of the basin. <strong>The</strong><br />

narrow, elongated gravity low along the <strong>Israel</strong>i coast may be associated with the effects of the<br />

Plio-Pleistocene prism, or with deeper crustal heterogeneities that at this stage are not well<br />

resolved.<br />

6.2 Magnetic Maps<br />

<strong>The</strong> pattern of the magnetic anomalies (Fig. 6.2a) shows they are distributed randomly<br />

throughout the region, rather than associated with distinct, large-scale features as in the gravity<br />

anomaly map. We therefore assume that the magnetic anomalies are associated with local<br />

magmatic features and not with the deep crustal structures.<br />

<strong>The</strong> low inclination of the Earth's magnetic field in the study area results in misplacement<br />

of the location and direction of the magnetic anomalies. An RTP (Reduction-To- Pole)<br />

calculation removes anomaly asymmetry caused by inclination and position of the anomalies<br />

above the causative magnetic bodies. It converts data which were recorded in the inclined<br />

Earth's magnetic field to what they would have been if the magnetic field was vertical. <strong>The</strong> RTP<br />

calculation was performed by the IGRF program (http://swdcdb.kugi.kyoto-u.ac.jp/igrf/)<br />

assuming the midpoint of the area at 35E, 34N, the total intensity of the Earth's magnetic field is<br />

44,763nT, its inclination is 49.6° and declination 2.7°. We also assume that the remanent<br />

magnetism is small compared to the induced magnetism<br />

<strong>The</strong> RTP map (Fig. 6.2b) preserves the overall magnetic pattern of the area; however it<br />

shows a shift of the positive and negative magnetic anomalies towards the NW. <strong>The</strong> accuracy<br />

of the RTP map is supported by the good correlation between the small, circular positive<br />

70


anomaly found in the center of the area and a high density block of similar pattern observed on<br />

the residual gravity anomaly map (Fig. 6.1b).<br />

<strong>The</strong> large positive magnetic anomaly found in the NW corner of the map (Fig. 6.2b) is<br />

the edge of the pronounced Eratosthenes magnetic anomaly (Ben-Avraham et al., 1976) located<br />

further to the NW. A large positive anomaly located in the SE part of the map (Fig. 6.2b) is<br />

associated with the Hebron magnetic anomaly (Rybakov et al., 1995) located inland further to<br />

the SE. Both anomalies are considered to be associated with Mesozoic volcanic bodies<br />

(Kempler, 1998; Rybakov et al., 1995).<br />

An elongated NE trending positive magnetic anomaly intersects the NW trending Carmel<br />

magnetic anomaly near the coastline (Fig. 6.2b). <strong>The</strong> Carmel anomaly is associated with Jurassic<br />

basalts found in several wells near the coastline (Asher Volcanics) (Garfunkel, 1989; Gvirtzman<br />

et al., 1990). <strong>The</strong> NE trending anomaly in the center of the basin is possibly associated with a<br />

Jurassic or younger magnetic source.<br />

6.3 Gravity and magnetic modeling (2D)<br />

A regional geologic cross-section compiled from the seismic depth maps was used for<br />

modeling of the gravity and magnetic fields (Fig. 6.3). <strong>The</strong> section crosses the basin in a NW-SE<br />

direction more or less perpendicular to most of the anomalies observed in the gravity and<br />

magnetic maps (Figs. 6.1, 6.2). Calculation of the gravity and magnetic curves was carried out<br />

using Geosoft software.<br />

Figure 6.3a shows the results of gravity modeling. <strong>The</strong> observed curve is taken from the<br />

Bouguer gravity map in Figure 6.1b. <strong>The</strong> modeled depth to the Moho is based on Ben-Avraham<br />

et al. (2002). <strong>The</strong> densities used for modeling are taken from well data. <strong>The</strong> basement density<br />

assumed in the central and western part of the section is 2.9 gr/mm 3 , corresponding to an oceanic<br />

or transitional type crust, whereas at the southeastern part it is assumed to be 2.75 gr/mm 3 ,<br />

corresponding to a continental type crust (Ben-Avraham et al., 2002).<br />

<strong>The</strong> good fit between the observed and calculated curves in the center of the basin<br />

supports the structural interpretation. A significant misfit is observed in the southeastern part of<br />

the section (Fig. 6.3a). <strong>The</strong> higher values of the calculated curve in this area can be explained in<br />

several ways: (a) the top of the crystalline basement is deeper than assumed; (b) the crystalline<br />

basement is either lighter; or (c) thicker than assumed (~22 km). <strong>The</strong> depth to the top of the<br />

crystalline basement in this area is relatively well constrained by the seismic data. Furthermore,<br />

in the Helez Deep-1 well located several tens of kilometers south of the edge of the section, the<br />

basement rocks were encountered at a depth of about 6 km, in agreement with the present<br />

structural interpretation. <strong>The</strong> density of the basement rocks in this well is 2.77 gr/mm 3 , similar<br />

71


Y<br />

X<br />

a<br />

b<br />

c<br />

Fig 6.3- 2D Gravity and magnetic modeling of the <strong>Levant</strong> <strong>Basin</strong> (see Figs. 6.1 and 6.2 for location<br />

of the section): (a) observed and calculated Bouguer gravity, (b) observed and calculated<br />

magnetic intensity assuming continuous magnetic basement (light blue line), and<br />

(c) observed and calculated magnetic intensity assuming separate magnetic bodies (blue polygons).<br />

72


to the value used for modeling. <strong>The</strong> most likely explanation for the misfit is therefore a thicker<br />

basement in the southeastern part of the section (Fig. 6.3a). This explanation is supported by a<br />

new deep refraction test (Netzeband et al., 2006), suggesting that the depth to the Moho near the<br />

coastline is 27 km, about 5 km more than previously assumed by Ben-Avraham et al. (2002).<br />

A misfit between the observed and calculated curves is found also in the northern edge of<br />

the section (Fig. 6.3a). <strong>The</strong> higher calculated values may suggest a shallower basement than<br />

interpreted from the seismic data. However, it should be noted that the reliability of the<br />

observed gravity data in this area is reduced due to the smaller number of measurements.<br />

Figures 6.3b and 6.3c show the results of the magnetic modeling. <strong>The</strong> observed curve is<br />

taken from the total magnetic anomaly map in Figure 6.2a. Two calculations were made. In the<br />

first one we assumed a magnetic basement extending along the entire basin (Fig. 6.3b). An<br />

optimal fit for this curve is reached when the top of the magnetic basement modeled is<br />

significantly shallower than the structural one (Fig. 6.3b). Additionally, the regional data<br />

suggests that the crystalline basement in the area is generally non-magnetic (Rybakov et al,<br />

1999). <strong>The</strong>refore, we consider the model in Figure 6.3b as unrealistic.<br />

An alternative option is presented in Figure 6.3c. In this calculation we assume the<br />

presence of several highly magnetized bodies at a shallow depth within the sedimentary section.<br />

<strong>The</strong> southeastern body may correspond to the northern edge of the Hebron anomaly, which is<br />

assumed to be associated with Jurassic volcanics (Fig. 6.3c). A shallow magmatic body in this<br />

area may correspond to the Tertiary National Park Volcanics. A second magnetic body, located<br />

in the center of the basin is possibly associated with extrusive igneous rocks at the top of the<br />

Jonah Ridge (Fig. 6.3c). Several small bodies in the northwestern part of the section may<br />

correspond to the Eratosthenes high located nearby (Fig. 6.3c). It should be noted that the<br />

geometry and properties of these bodies as presented in Figure 6.3c are only tentative.<br />

6.4. Discussion<br />

<strong>The</strong> gravity and magnetic anomaly maps (Figs. 6.1, 6.2) and the modeled section (Fig.<br />

6.3) shed new light on the deep structure of the <strong>Levant</strong> <strong>Basin</strong>. <strong>The</strong> high Bouguer gravity values<br />

in the central and northern part of the basin (Fig. 6.1a) indicate the presence of a dense crust, in<br />

the range of 2.9 gr/mm 3 . <strong>The</strong> residual map (Fig. 6.1b) suggests that the distribution of the<br />

positive gravity anomalies is related to several distinct blocks oriented NE-SW. <strong>The</strong>se blocks<br />

are interpreted as fault-controlled basement highs (Fig. 6.3a) that were formed during Early<br />

Mesozoic rifting (Garfunkel and Derin, 1984; Gardosh and Druckman, 2006). Rifting was<br />

probably associated with extrusive volcanism as well as with intrusion of mafic rocks at a lower<br />

73


crustal level. <strong>The</strong> combination of extension and magmatic intrusion resulted in the modification<br />

of the old continental crust and the formation of a thinned and heavy crust in the <strong>Levant</strong> <strong>Basin</strong>.<br />

<strong>The</strong> NE-SW negative anomaly extending along the coastline is related to the transition<br />

from the intruded, heavy crust of the basin into the lighter (~2.75 gr/mm 3 ), continental crust of<br />

the margin of the Arabian plate (Ben-Avraham et al., 2002). It is possible that the depth to the<br />

Moho at the eastern edge of the basin is deeper than previously assumed and is in the range of 27<br />

km, as recently proposed by Netzeband et al. (2006).<br />

<strong>The</strong> magnetic anomaly maps (Figs. 6.2a, b) and the modeled section (Fig. 6.3a, b) reveal<br />

the presence of several highly magnetized bodies. <strong>The</strong> distribution the magnetic anomalies is not<br />

directly related to the deep crustal structure of the <strong>Levant</strong> <strong>Basin</strong> indicated by the gravity anomaly<br />

maps. <strong>The</strong> modeled section suggests that these bodies are found within the sedimentary section<br />

and they are interpreted as extrusive volcanics. <strong>The</strong> Carmel anomaly at the northeastern edge of<br />

the basin is most likely associated with the Early Jurassic Asher Volcanics (Gvirtzman et<br />

al.,1990) found in the Atlit-1 well and other boreholes in the Haifa area. An Early Jurassic age is<br />

also assumed for the Hebron anomaly (Rybakov et al., 1995) at the southeastern edge of the<br />

basin and the Eratosthenes anomaly at its northwestern edge.<br />

<strong>The</strong> origin of the elongated, NE-SW oriented anomaly in the central part of the basin is<br />

less well understood. Its northeastern edge intersects the Carmel magmatic anomaly. Its<br />

southwestern edge is correlated to a positive gravimetric anomaly and a Mesozoic basement<br />

high, known as the Jonah structure (Folkman and Ben-Gai, 2004). <strong>The</strong> modeled section shows<br />

that the Jonah magnetic anomaly is found at a relatively shallow depth within rocks of<br />

Cretaceous to Tertiary age (Fig. 6.3c). It is speculated that this feature is related to Cretaceous<br />

and Tertiary eruptive episodes that are probably associated with a deeply rooted magmatic<br />

chamber.<br />

7. <strong>Tectonic</strong> Evolution of the <strong>Levant</strong> <strong>Basin</strong> and Margin<br />

7.1 Rifting Stage<br />

7.1.1 Late Paleozoic to Early Mesozoic extensional structures inland<br />

Rifting activity in the <strong>Levant</strong> region started in the Late Permian and continued in several<br />

pulses through the Triassic and Early to Middle Jurassic. Significant vertical movements and<br />

differential block motions took place during this period in northern Sinai, <strong>Israel</strong> and Syria (Fig.<br />

7.1)(Goldberg and Friedman, 1974; Druckman, 1974,1977; Freund et al., 1975; Druckman,<br />

1984; Gelberman and Kemmis, 1987; Bruner, 1991; Druckman et al., 1995a; Garfunkel, 1998).<br />

74


75<br />

Fig 7.1- Neotethyan rifting stage- inferred Triassic to Early Jurassic horst and graben system of the <strong>Levant</strong> region. Inserted section<br />

shows the structural relations across the Helez fault, on the southern coastal plain (after Gardosh and Druckman, 2006).


Attesting to these motions are the thickness and facies changes of the Late Paleozoic and<br />

Mesozoic strata found in wells and outcrops in the region (for details see chapter 4).<br />

<strong>The</strong> structural configuration revealed by this wide range of phenomena is an extensive<br />

graben and horst system extending east of the Mediterranean coastline (Fig. 7.1). Some of the<br />

faults that bound these structures are identified in seismic reflection lines (Gelbermann, 1995;<br />

Druckman et al., 1995b), while others are inferred from well data.<br />

A major structural low extends in a NW-SE direction from northern Sinai to central <strong>Israel</strong><br />

(the Hilal and Judea grabens)(Fig. 7.1). Towards the north this structure splits into two<br />

segments, the Asher graben trending to the NW and the southern edge of the Palmyra Trough<br />

trending to the NE. <strong>The</strong> Judea-Asher graben system is bounded on the west by three<br />

downstepping horst blocks: the Gevim high in the south and the Gaash and Maanit highs in the<br />

north (Fig. 7.1). <strong>The</strong> Helez fault is the eastern boundary of a structural low extending near the<br />

coastline west of the Gevim high (Fig. 7.1). Other, smaller scale structures are found in the<br />

southern Dead Sea area (Hemar and Massada highs) and probably also below the Syrian Arc<br />

structures of the northern Negev (Fig. 7.1)(Freund et al., 1975).<br />

<strong>The</strong> timing of activity of this horst and graben system is broadly defined. Part of the<br />

system was active during the Permian (Garfunkel, 1998). Evidence for this initial stage in the<br />

inland <strong>Levant</strong> area is sparse due to limited well data. A significant tectonic phase took place<br />

during the Middle and Late Triassic (Anisian and Carnian) followed by a relatively quiet period<br />

during the latest Triassic (Garfunkel, 1998). A third and possibly the most extensive phase took<br />

place during the Early to Middle Jurassic (Liassic to Bathonian). <strong>The</strong> large amount of vertical<br />

offset during this phase is evident on the Helez fault (Fig. 7.1insert) (Gardosh and Druckman,<br />

2006), and abrupt thickness changes in the Liassic Ardon and Inmar formations (Goldberg and<br />

Friedman, 1974; Druckman 1977; Buchbinder and le Roux, 1992).<br />

<strong>The</strong> Early Jurassic tectonic phase was followed by extensive extrusive magmatism in<br />

northern <strong>Israel</strong> (Asher Volcanics). Evidence for magmatic activity during this time, probably of<br />

more intrusive nature, is found also in various wells in central and southern <strong>Israel</strong> (Garfunkel,<br />

1989; Rybakov et al., 1955). It is estimated that parts of the inland rift system were active at<br />

different rates during different times.<br />

7.1.2 Late Paleozoic to Early Mesozoic extensional structures in the <strong>Levant</strong> <strong>Basin</strong><br />

<strong>The</strong> horst and graben system identified inland is observed on the seismic data of the<br />

<strong>Levant</strong> <strong>Basin</strong> offshore. <strong>The</strong> details of this system are more easily recognized in the central and<br />

western part of the basin than in the eastern part, where the sedimentary section was strongly<br />

affected by younger, contractional deformation (Figs. 5.8, 5.9).<br />

76


W<br />

E<br />

TWT(ms)<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

a<br />

Precambrian ()<br />

Crystalline Basement<br />

Jonah Ridge<br />

W<br />

E<br />

TWT(ms)<br />

M. Jurassic-Turonian<br />

Permian- M. Jurassic<br />

b<br />

Precambrian ()<br />

Crystalline Basement<br />

Yam High<br />

Fig 7.2- Neotethyan rifting stage- Early Mesozoic extensional structures in the<br />

<strong>Levant</strong> <strong>Basin</strong>; (a) the northeastern part of the Jonah Ridge and (b) the<br />

southern part of the Yam High (see Fig. 7.1). Normal faults found east<br />

and west of the Yam High were reactivated as reverse faults.<br />

77


<strong>The</strong> most prominent structures are two basement highs, the Jonah Ridge (Folkman and<br />

Ben-Gai, 2004) and Leviathan High (Figs. 5.8, 5.9, 7.1). A third structure, the Eratosthenes<br />

High, is partly covered by the seismic data and only its southeastern part is observed (Figs. 5.9,<br />

7.1). <strong>The</strong> Jonah and Leviathan highs are structures 15 to 30 km wide and 80 to 100 km long<br />

trending in a NE-SW direction (Figs. 5.11, 5.12, 7.1). In the two structures the near-topbasement<br />

Brown horizon and the Middle Jurassic Blue horizon are markedly elevated from their<br />

surroundings. <strong>The</strong> identification of these horizons within the structures is somewhat tentative<br />

due to the correlation of seismic events across faults. However, the depth to the top of the<br />

basement level as derived from the seismic interpretation and depth maps is supported by the<br />

results of the gravity modeling presented in Chapter 6 (Fig. 6.3a). <strong>The</strong> Leviathan basement high<br />

is also observed as a large positive anomaly on the residual Bouguer gravity map (Fig. 6.1).<br />

<strong>The</strong> two structures are bounded by sets of high-angle normal faults with vertical offsets<br />

in the range of several hundred meters to a few kilometers. <strong>The</strong> structural lows on the two sides<br />

of the structures are characterized by dipping events, forming asymmetric grabens (Figs. 5.9,<br />

7.2a). <strong>The</strong> grabens are filled with 4-8 km thick sections of Permian to Middle Jurassic strata.<br />

Several smaller basement highs, about 10 km wide and 20 km long trending in a NE-SW<br />

direction are identified south and north of the Jonah and Leviathan ridges (Figs. 5.12, 5.13, 7.1).<br />

A larger structure is partly revealed by the seismic data in the southeastern part of the basin. <strong>The</strong><br />

Yam High is identified by a series of deep-seated normal faults found west of the Yam-2 and<br />

Bravo-1 wells (Figs. 7.1, 7.2b). <strong>The</strong> near-top-basement level is downfaulted to the west, forming<br />

a set of down-to-the-basin fault blocks. <strong>The</strong> thickness of the Brown to Blue interval shows a<br />

marked increase from about 1 km in the east to about 3 to 4 km in the west (Fig. 7.2b). <strong>The</strong> Yam<br />

High probably extends several tens of kilometers to the NE (Fig. 7.1). Its eastern boundary is<br />

marked by a series of faults located near the Yam-2 and Bravo-1 wells (Fig. 7.2b). <strong>The</strong>se are<br />

interpreted as normal faults that were later reactivated in a reverse motion.<br />

Unlike the situation inland, the timing of activity of the horst and graben system within<br />

the basin is only broadly defined due to the uncertainty associated with the age of the seismic<br />

markers. <strong>The</strong> asymmetric grabens near the Jonah and Leviathan ridges suggest syntectonic<br />

deposition through the Late Paleozoic to Middle Jurassic periods. Some of the faults bounding<br />

these structures appear to have been active also during post Middle Jurassic time (Fig. 7.2a),<br />

however, it is assumed that this late activity was only minor and during the Late Jurassic the<br />

structures were well developed and formed submarine topographic highs. <strong>The</strong> faults on the<br />

western edge of the Yam High do not offset the Middle Jurassic horizon, thus suggesting a Late<br />

Paleozoic to Triassic age of this structure (Fig. 7.2b).<br />

78


<strong>The</strong> near-top-basement to Middle Jurassic interval shows significant thickness variations<br />

across the basin. It is about 3 km thick near the Mediterranean coastline and the Eratosthenes<br />

and Leviathan highs and about 5 to 8 km thick in the central part of the basin (Figs. 5.8 to 5.10).<br />

<strong>The</strong> middle part of the basin termed here the Central <strong>Levant</strong> Rift (Fig. 7.1) is interpreted as the<br />

main depocenter associated with the Late Paleozoic to Middle Jurassic rifting activity. Intra-rift<br />

highs such as the Jonah Ridge were formed within this large depocenter. Localized lows may<br />

have been partly filled with a thick volcanic series (Asher Volcanics), as suggested by Gardosh<br />

and Druckman (2006).<br />

7.1.3 Neotethyan rifting in the <strong>Levant</strong> area<br />

<strong>The</strong> onshore and offshore horst and graben system was formed under an extensional<br />

regime associated with major plate motions and the opening of the Neotethys Ocean in the<br />

<strong>Levant</strong> area and north of it (Garfunkel and Derin, 1984; Garfunkel, 1998). <strong>The</strong> <strong>Levant</strong> rift<br />

system was bounded by the Arabian Massif to the southeast and the Eratosthenes continental<br />

block to the west and thus was probably separated from the main body of the Neotethys Ocean<br />

that extended north of the Eratosthenes Seamount and was later consumed underneath Cyprus<br />

and southern Turkey (Garfunkel, 1998, 2004; Robertson, 1998). Four main rifting episodes are<br />

recognized: during the Permian; early Middle Triassic; Late Triassic; and Early to Middle<br />

Jurassic. <strong>The</strong> amount of extension during each of these phases is not well established<br />

<strong>The</strong> general strike of the normal faulting in the <strong>Levant</strong> <strong>Basin</strong> is NE-SW; accordingly, we<br />

interpret an extension in a NW-SE direction perpendicular to the strike of the faulting. <strong>The</strong><br />

extension may have been accommodated by NW-SE trending transform faults within the basin<br />

(Fig. 7.1). Garfunkel and Derin (1984) and Garfunkel (1998) postulated a similar strike-slip<br />

fault along northern Sinai to explain the separation and northward motion of the Tauride and<br />

Eratosthenes blocks. An apparent lateral offset in the northern edge of the Jonah ridge and the<br />

southern edge of the Leviathan High (Fig. 7.1) may be explained by strike–slip faulting.<br />

<strong>The</strong> structural scheme presented above for the Neotethyan rifting is not in accord with<br />

extension and spreading of the eastern Mediterranean in an N-S direction accompanied by a<br />

transform fault along the eastern Mediterranean shoreline, as proposed by Dewey et al. (1973),<br />

Bein and Gvirtzman (1977), Robertson and Dixon (1984) and Stampfli et al. (2001).<br />

An important question regarding the nature of the Neotethyan rifting processes is whether<br />

emplacement of new oceanic crust took place in the <strong>Levant</strong> area. Apart from faulting no<br />

pronounced disruption or conspicuous lateral variations in the seismic properties of either the<br />

upper part of the basement or the Late Paleozoic to Middle Jurassic interval are observed in the<br />

seismic data set. <strong>The</strong> basement layer in the central part of the basin does not show any<br />

79


characteristics of oceanic crust such as described in other passive continental margins (Klitgord<br />

and Hutchinson, 1988) or in the central part of the Mediterranean Sea (Finetti, 1985; Avedik et<br />

al., 1995). <strong>The</strong> normal faults and basement highs associated with the rifting stage are well<br />

preserved. Although a considerable amount of Triassic and Jurassic volcanics may exist within<br />

the basin, their equivalent units in the onshore area (Asher Volcanics) do not show MORB<br />

characteristics associated with sea-floor spreading. Thus, the occurrence of spreading and<br />

introduction of new oceanic crust cannot be supported by the present data.<br />

In view of the above discussion, it is suggested that the Early Mesozoic rifting that<br />

started on the northern edge of Gondwana may not have reached the spreading phase in the area<br />

of the present <strong>Levant</strong> <strong>Basin</strong> (Gardosh and Druckman, 2006). A recent analog for this scenario is<br />

the northern part of the Red Sea. <strong>The</strong>re, the rift is continental with only a nucleation of an<br />

oceanic spreading center and an early magmatic phase. An oceanic spreading center has<br />

developed only in the southern and central parts of the Red Sea (Martinez and Cochran 1988;<br />

Bohannon and Eittreim 1990; Cochran, 2001).<br />

<strong>The</strong> thinned crust in the center of the <strong>Levant</strong> <strong>Basin</strong> (Fig. 2.1) may be explained in two<br />

ways: (a) extension and stretching on basement-involved sets of normal faults; (b) magmatic<br />

underplating and thermal erosion at the base of the crust (Hirsch et al., 1995). It should be noted<br />

that the amount of extension on the Early Mesozoic faults appears to be limited as most of them<br />

are high-angle normal faults and lack the listric character that is often observed in other rifted<br />

continental margins. <strong>The</strong> high-density of the thinned crust that is interpreted from seismic<br />

velocities (Fig. 2.1) (Ben-Avraham et al., 2002) and is indicated by the modeling of gravity data<br />

in the present study (Fig. 6.3c) may be associated with intrusion of mantle material into the old<br />

Pan-African continental crust of northern Gondwana (Hirsch et al., 1995).<br />

7.2 Post-Rift, Passive Margin Stage<br />

Rifting and normal faulting activity in the <strong>Levant</strong> by and large ceased during Middle to<br />

Late Jurassic (Garfunkel, 1998). A post-rift, passive margin stage was initiated at this time as a<br />

result of crustal cooling and thermal subsidence (Garfunkel and Derin, 1984; Garfunkel, 1989;<br />

ten Brink, 1987). A faster rate of subsidence in the central part of the basin compared to that on<br />

its margins may have initiated the formation of the 'Mesozoic depositional Hinge-Belt' along the<br />

Mediteranean coast.<br />

<strong>The</strong> Hinge-Belt is characterized by a pronounced facies change in the Mesozoic section,<br />

from shallow water carbonates and sandstones in the east to fine-grained, pelagic and hemipelagic<br />

carbonates and siliciclastics in the west (Derin, 1974; Bein and Gvirtzman, 1977; Flexer<br />

et al., 1986). This pronounced facies change, which was detected in many onshore and offshore<br />

80


UP.<br />

LO. MIDDLE<br />

LOWER MIDDLE<br />

UP.<br />

Carbonate Platforms<br />

PLIOCENE - PLEISTOCENE<br />

MIDDLE -UPPER TERTIARY<br />

UPPER CRETACEOUS<br />

TO LOWER TERTIARY<br />

MID JUR.-<br />

MID CRET.<br />

LOWER<br />

JURASSIC<br />

TRIASSIC<br />

PALEOZOIC <br />

Fig 7.3- Passive margin stage- (b) isopach map of the Middle Jurassic to<br />

Middle Cretaceous interval, and (a) geologic section through the<br />

<strong>Levant</strong> margin showing the two dominant depositional styles:<br />

backstepping and aggrading, shallow marine carbonate platforms<br />

on the east; and deepwater, siliciclastic and carbonate turbidite<br />

systems on the west (from Gardosh, 2002). Blue and green lines<br />

denote seismic horizons.<br />

TRIASSIC<br />

PERMIAN<br />

PRECAMBRIAN<br />

Deepwater Turbidites<br />

a<br />

b<br />

81


wells (Fig. 7.3), indicates the development of a deep marine basin in the <strong>Levant</strong> area, bordered<br />

by a slope and a shallow-marine shelf.<br />

<strong>The</strong>re is some ambiguity regarding the time of initiation of this passive margin-type,<br />

depositional profile. Our stratigraphic analysis shows that all the Triassic as well as the first four<br />

Jurassic depositional cycles (Pliensbachian to Bathonian) are generally dominated by shallowmarine-type<br />

rocks. However, the allochthonous Bathonian section in the offshore Yam Yafo-1<br />

well (Fig. 4.2) contains mass flows that were obviously deposited in a deeper marine<br />

environment (see chapter 4.3). It is therefore assumed that the allochthonous succession in the<br />

Yam Yafo-1 well indicates an initial stage of basin evolution within a predominantly platform<br />

setting.<br />

A deep marine depositional environment was well established throughout the <strong>Levant</strong><br />

basin since Oxfordian to Kimmeridgian times. <strong>The</strong> corresponding rock interval in the offshore<br />

Yam Yafo-1, Yam-2 and Yam West-1 well (Delta Formation) is composed of carbonate mud of<br />

deep marine origin with transported material (Fig. 7.3) (Derin et al., 1990; Druckman et al.,<br />

1994, Gardosh, 2002).<br />

<strong>The</strong> deepening of the basin and the change in depositional environments is further<br />

indicated by the change in seismic character delineated by the Blue horizon. A relatively<br />

continuous high-amplitude reflection series, interpreted as shallow-marine strata below the Blue<br />

horizon changes upward to a discontinuous, low-amplitude, occasionally shingled seismic facies<br />

(between the Blue and Green horizons). <strong>The</strong> latter is interpreted as fine-grained strata of deepmarine<br />

origin with mass transported components (Fig. 5.1).<br />

<strong>The</strong> strata architecture along the passive margin was significantly affected by eustatic<br />

changes. Gradual sea-level rise resulted in backstepping and aggradation of the carbonate<br />

platforms, whereas fast sea-level rise resulted in drowning of the carbonate platforms. Sea-level<br />

drops were associated with bypass of the shelf and deposition of siliciclastic and carbonate<br />

gravity flows on the slope and in the basin. <strong>The</strong>se are remarkably demonstrated by the<br />

siliciclastic, turbidite system of the first two Cretaceous depositional cycles (Berriasian to<br />

Hauterivian and Hauterivian to Lower Aptian; Fig. 4.3).<br />

<strong>The</strong> passive margin succession reaches a thickness of 2500 to maximal 3500 m (Fig.<br />

7.3a). Its depocenter extends from the present coastline area to about 70 km westward. In the<br />

central part of the basin and near the Eratosthenes High the thickness of this succession is<br />

reduced to several hundreds up to 2000 m (Fig. 7.3a). In this distal part of the basin the rate of<br />

accumulation of deepwater sediments was significantly slower than near the margin.<br />

Thick accumulations of the passive margin stage (between the Blue and Green horizons)<br />

are located adjacent to the Early Mesozoic horsts (east and west of the Jonah Ridge in Fig. 7.2a).<br />

82


It is assumed that these areas continued to act as localized depocenters, accumulating clastic<br />

material that was shed from the nearby submarine topographic highs. <strong>The</strong> tops of the Jonah,<br />

Leviathan and Eratosthenes structures may have reached shallow water level and hosted isolated<br />

carbonate buildups of Middle-Late Jurassic and Middle Cretaceous age.<br />

7.3 Convergence Stage<br />

7.3.1 Late Cretaceous to Tertiary contractional structures inland<br />

<strong>The</strong> termination of post-rift, passive margin stage coincides with the onset of a regional<br />

contractional deformation phase associated with the formation of the 'Syrian Arc' (Krenkel,<br />

1924) or the '<strong>Levant</strong>id' (Picard, 1959) fold belt. This structural element is an 'S' shape mountain<br />

belt extending from the Palmyra Mountains in Syria to the Lebanon and Anti-Lebanon<br />

Mountains; through the Judea Mountains and Negev anticlines and into the northern Sinai<br />

anticlines (Fig 7.4) (Picard, 1943, 1959; Ball and Ball, 1953; Bentor and Vroman, 1954, 1960; de<br />

Sitter, 1962; Gvirtzman, 1970; Bartov, 1974; Neev et al., 1976; Horowitz, 1979; Eyal and<br />

Reches, 1983; Lovelock, 1984; Beydoun, 1988; McBride et al., 1990). <strong>The</strong> fold belt extends<br />

further to the southwest below the thick Mio-Pliocene sediments of the Nile Delta (Aal et al.,<br />

2000) and into the Western Desert of Egypt (Said, 1990).<br />

<strong>The</strong> Syrian Arc belt consists of a series of surface and subsurface anticlines that strike in<br />

an ENE, NE and NNE direction. Most of the structures are open flexures and folds, in some<br />

areas forming broad anticlinorial upwarps. Several characteristic geometries are identified: the<br />

folds of the central and northern Negev are all strongly asymmetric with steep flanks on the<br />

southeast, while the Hebron anticline is strongly asymmetric on the west (Fig. 7.4). Other<br />

structures, such as the Fari'a anticline (Fig. 7.4), are roughly symmetric box folds with steeply<br />

dipping flanks on both sides (Flexer et al., 2005).<br />

<strong>The</strong> Syrian Arc folds are frequently associated with reverse faulting. <strong>The</strong> concept of<br />

reverse faults at depth, as first suggested by de Sitter (1962), was substantiated by several deep<br />

oil wells in southern and central <strong>Israel</strong> in which strata repetitions were discovered. Freund et al.<br />

(1975) suggested that these reverse faults originated as normal faults during the Early Mesozoic<br />

rifting stage. Compelling evidence for inversion of the older extensional structures has been<br />

supplied from various parts of the Syrian Arc fold belt in onshore <strong>Israel</strong> and Syria (Davis, 1982;<br />

Gelbermann and Kemmis, 1987; Bruner, 1991; Best et al., 1993, Chaimov et al., 1993;<br />

Druckman et al., 1995a).<br />

<strong>The</strong> timing of the Syrian Arc deformation inland is debated, but the main range of ages<br />

given by various authors is Turonian to Neogene (Picard, 1943; Bentor and Vroman, 1954, 1960;<br />

Freund, 1965; Bartov, 1974; Eyal and Reches, 1983). Walley (1998), who summarized<br />

83


84<br />

Fig 7.4 – Convergence stage-Syrian Arc fold structures of the <strong>Levant</strong> region. Combined depth map of the Middle Cretaceous level:<br />

onshore after Fleischer and Gafsou (2003) top Judea map; and offshore depth map of the Green horizon in present study. Note Syrian Arc<br />

I and II folding phases. <strong>The</strong> Syrian Arc II structures offshore are projected from the top Lower Miocene level.


NW<br />

SE<br />

Plio-Pleistocene<br />

Senonian-<br />

Eocene<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-<br />

L. Miocene<br />

2<br />

1<br />

TWT (ms)<br />

M. Jurassic-<br />

Turonian<br />

a<br />

12.5 km<br />

NW<br />

SE<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-<br />

L. Miocene<br />

2<br />

1<br />

TWT(ms)<br />

Senonian-<br />

Eocene<br />

b<br />

M. Jurassic-<br />

Turonian<br />

85


NW<br />

SE<br />

TWT(ms)<br />

Plio-<br />

Pleistocene<br />

Messinian<br />

M.-U. Miocene<br />

Oligocene-<br />

L. Miocene<br />

2<br />

1<br />

Senonian<br />

-Eocene<br />

c<br />

M. Jurassic-<br />

Turonian<br />

NW<br />

SE<br />

Plio-<br />

Pleistocene<br />

Messinian<br />

TWT(ms)<br />

M.-U. Miocene<br />

Oligocene-<br />

L. Miocene<br />

2<br />

Senonian<br />

-Eocene<br />

M. Jurassic-<br />

Turonian<br />

d<br />

Fig 7.5- Convergence stage- contractional structures in the <strong>Levant</strong> <strong>Basin</strong>: (a) profile from the<br />

southeastern part of the basin showing two folding phases, 1= Syrian Arc I (Senonian) and<br />

2=Syrian Arc II (Oligo-Miocene); (b) profile from the southeastern part of the basin showing<br />

Syrian Arc I+II folds, note the onlapping of the Oligo-Miocene strata on the tilted, faulted and<br />

folded block apparently deformed during Syrian Arc II folding phase; (c) profile from the eastern<br />

part of the basin showing Syrian Arc I+II folds; (d) profile from the northeastern part of the basin<br />

showing Syrian Arc II folds<br />

86


observations on the Syrian Arc system from Lebanon and other parts of the <strong>Levant</strong>, suggested<br />

two main episodes of deformation: Syrian Arc I during Coniacian to Santonian times; and Syrian<br />

Arc II during Late Eocene to Late Oligocene times. Mimran (1984) found a 'post Neogene' age<br />

for the latest folding phase in the Fari'a anticline (Fig. 7.4), while Freund (1965) infers a post<br />

Pliocene continuation of folding in some structures. Eyal (1996), who studied the properties of<br />

Syrian Arc folds and related structures throughout <strong>Israel</strong> and Sinai, suggested that the Syrian Arc<br />

stress field in the <strong>Levant</strong> area may have persisted into the Miocene and possibly to the present.<br />

7.3.2 Late Cretaceous to Tertiary contractional structures in the <strong>Levant</strong> <strong>Basin</strong><br />

Syrian Arc type contractional deformation is observed in the offshore seismic data<br />

throughout the <strong>Levant</strong> <strong>Basin</strong> and margin. Two folding episodes are identified, and following<br />

Walley (1998, 2001), these are termed here Syrian Arc I and II (Fig. 7.5a).<br />

Syrian Arc I structures are predominant in the eastern part of the basin some 50 to 70 km<br />

west of the coastline (Fig. 7.4). <strong>The</strong> dominant deformational style in this area is of highamplitude<br />

and short wave length anticlines and synclines accompanied by high-angle thrust<br />

faults (Figs. 5.3, 7.5a). <strong>The</strong> deformation affects the Middle Cretaceous unconformity (Green<br />

horizon) and older strata; therefore a Senonian age for this folding is inferred (Figs. 5.3, 5.6,<br />

7.5a).<br />

<strong>The</strong> time and style of deformation of the Syrian Arc I structures offshore appear to be<br />

similar to the folds of the northern Negev inland. <strong>The</strong> size of the folds ranges from 10 to 30 km<br />

in length and 5 to 10 km in width. <strong>The</strong>ir height ranges from several hundred to more than 1000<br />

m and their flanks dip 10 0 -30 0 . Many of the folds are asymmetric with steep flanks on the east or<br />

southeast (Fig. 7.5a, b).<br />

<strong>The</strong> thrust faults dip 65-75 0 and are traced down into the basement. Most of the faults<br />

offset up to the Middle Cretaceous unconformity (Green horizon), although rarely, they also<br />

penetrate the overlying Tertiary strata (Figs. 5.3, 5.6, 7.5a, b). In some areas evidence for<br />

inversion of the older, extensional structures is observed. For example, the reverse fault west of<br />

the Yam West-1 well may have been a normal fault associated with down-to-the-basin, Early<br />

Mesozoic faulting (Figs. 5.3, 7.2b). Many of the faults resemble positive flower structures and<br />

contain supporting limbs; thus a certain amount of wrenching or transpression is inferred.<br />

Syrian Arc II contractional deformation in the offshore is more complex and contains two<br />

elements. <strong>The</strong> first element is a series of low-amplitude long wave folds that are in some places<br />

superimposed on the older Syrian Arc I structures (Figs. 7.5a, c). <strong>The</strong> deformation affects the<br />

Lower Miocene unconformity (Cyan horizon) and older strata; therefore a Middle Miocene age<br />

of this folding phase can be inferred. Syrian Arc II folds are found throughout the basin and also<br />

87


in its deep part east of the Eratosthenes high (Fig. 7.4). Reverse faulting and inversion are<br />

generally not associated with this style of deformation; however, some of the structures in the<br />

deep part of the basin are located above deep-seated basement highs, possibly suggesting some<br />

reactivation of the Early Mesozoic horst and graben system (Fig. 5.9). <strong>The</strong> time span of the<br />

Syrian Arc II deformation phase is not well constrained. Folding could have started in the Late<br />

Eocene, as proposed by Walley (1989, 2001), and could have continued through the Early to<br />

Middle Miocene. <strong>The</strong> base Messinian unconformity (Purple horizon) is generally not affected or<br />

very mildly deformed (Figs. 7.5a-d). <strong>The</strong>refore an upper age limit for the Syrian Arc II<br />

deformation in the <strong>Levant</strong> <strong>Basin</strong> is the latest Miocene.<br />

A second element associated with the Syrian Arc II deformation is uplifting and tilting.<br />

Figure 7.5b shows an uplifted block at the western end of the Syrian Arc I fold belt (Fig. 7.4).<br />

<strong>The</strong> Middle Cretaceous unconformity (Green horizon) is tilted westward and the uppermost part<br />

of the structure is elevated some 3000 m from the basin floor (western part of the profile in Fig.<br />

7.5b). <strong>The</strong> Oligocene to Lower Miocene section onlaps the elevated structure on the east; thus<br />

indicating filling of a negative relief (Fig. 7.5b). Uplifting and onlapping of Neogene strata<br />

characterize the entire western edge of the Syrian Arc I fold belt in the eastern part of the basin<br />

(Fig. 7.4 and seismic examples in Figs. 5.3, 5.6, 5.8,). <strong>The</strong> age of this event is estimated as<br />

Oligocene to Upper Miocene (pre-Messinian).<br />

At the northeastern part of the basin a younger deformation is observed. In the Carmel-<br />

Cesarea area the Mesozoic and Cenozoic section is uplifted near the coastline and the thin Plio-<br />

Pleistocene section on top of the structure thickens significantly west of it (Fig. 5.10). It is<br />

therefore suggested that in this area the uplifting continued through the Plio-Pleistocene.<br />

7.3.3 Neotethyan convergence in the <strong>Levant</strong> area<br />

It is widely accepted that the Syrian Arc contractional deformation of the <strong>Levant</strong> area is<br />

associated with the collision of the African-Arabian and Eurasian plates (Reches and Eyal, 1983;<br />

Garfunkel, 1998; Walley, 1998, 2001; Flexer et al., 2005). A late Early Cretaceous convergence<br />

initiated a northward-dipping subduction zone within the southerly Neotethys oceanic basin<br />

(Robertson, 1998) that eventually progressed to collision and accretion in the present-day area of<br />

Cyprus and southern Turkey. This activity is manifested at a lower intensity by the Syrian Arc<br />

contractional structures of the <strong>Levant</strong>.<br />

Reches and Eyal (1983) and Eyal (1996) described the Syrian Arc stress field (SAS)<br />

associated with this deformation. According to these authors the SAS regime has been active<br />

from the Turonian to the present. This suggestion is generally supported by the data from the<br />

<strong>Levant</strong> <strong>Basin</strong>. <strong>The</strong> long-term contraction is marked by phases of more intense deformation.<br />

88


<strong>The</strong> Senonian Syrian Arc I phase is characterized by localized folding, partly controlled by the<br />

location of the deep-seated, Early Mesozoic normal faults. Walley (1998) proposed that this<br />

phase reflects an ocean-ocean collision in the southern Neotethys. Thickness and facies<br />

variations of the Senonian strata in the <strong>Levant</strong> area are relatively limited, therefore it is<br />

speculated that no significant relief was formed between the <strong>Levant</strong> margin on the east and the<br />

deep basin on the west during the Syrian Arc I deformation.<br />

<strong>The</strong> Syrian Arc II phase that started in the Paleogene is possibly related to the<br />

progression of convergence and to a continent-continent collision on the southern Neotethys<br />

margin (Walley, 1998, 2001). <strong>The</strong> wide extent of this deformation that was previously identified<br />

only in several locations inland is revealed in the offshore data (Fig. 7.4). It is assumed that<br />

many of the older, Syrian Arc I folds onshore and offshore were reactivated while new structures<br />

were continuously being formed. <strong>The</strong> offshore seismic data further indicates that the Syrian Arc<br />

II contractional deformation continued through the Miocene (Eyal, 1996) and did not end in the<br />

Late Oligocene as proposed by Walley (1998, 2001).<br />

An important element of the Syrian Arc II deformation shown by the offshore seismic<br />

data is uplifting and tilting on the <strong>Levant</strong> margin (Fig. 7.5b). Walley (1998) proposed that the<br />

main uplifts of the Lebanon and the Anti-Lebanon mountains occurred during this phase. It is<br />

assumed that uplifting of the mountainous backbone of <strong>Israel</strong> also took place during the Oligo-<br />

Miocene and resulted in erosion, westward transport of clastic sediments and huge<br />

accumulations within the deep part of the <strong>Levant</strong> <strong>Basin</strong>. Uplifting continued locally in the<br />

Carmel-Caesarea area during the Plio-Pleistocene, possibly in conjunction with the activity on<br />

the nearby segment of the Dead Sea Transform (Yagur fault) (Ben-Gai and Ben-Avraham,<br />

1995).<br />

8. Erosion Processes and Mass Transport on the <strong>Levant</strong> Slope<br />

8.1 Rifting and Passive Margin Stages<br />

Erosion and transport of sediments from elevated areas in the south and east to the basin<br />

in the west and northwest have been taking place throughout the Mesozoic and Cenozoic history<br />

of the region. <strong>The</strong> earliest evidence for coarse, clastic sediment accumulation is found in the<br />

Helez Deep-1 well in the southern coastal plain of <strong>Israel</strong> (Fig. 1.3). <strong>The</strong> Erez Conglomerate<br />

(Druckman, 1984) is a few hundred meters thick, built of polimictic carbonate breccia of lower<br />

Late Triassic age (Fig. 4.1). Druckman (1984) described this unit as a fault breccia that was<br />

eroded from a nearby horst block (Gevim High, Fig. 7.1) and accumulated on the downthrown<br />

side of the Helez fault. <strong>The</strong> Triassic to Middle Jurassic horst and graben morphology could have<br />

89


accommodated a well-developed drainage system, transporting fine and coarse siliciclastics.<br />

During lowstands, e.g., between the Jurassic cycles 3 and 4 (Fig. 4.2), the Inmar sands may have<br />

been transported from the distal Sinai and Negev hinterland toward the basin in the north and<br />

west<br />

During the passive margin stage deepwater turbidities accumulated on the slope and<br />

within the basin, west of the shallow-marine shelf (Fig. 7.3)(Gardosh, 2002). Thin oolitic and<br />

micro-conglomerate beds found in the Upper Jurassic cycle offshore (Yam-2, Yam West-1 and<br />

Yam Yafo-1 wells; Derin et al., 1990; Druckman et al., 1994) are the products of submarine,<br />

mass transport processes. <strong>The</strong> distribution and paleogeographic extent of these units are not fully<br />

resolved.<br />

<strong>The</strong> Lower Cretaceous Gevar'am Formation reflects an extensive deepwater turbidite<br />

system that developed on the <strong>Levant</strong> slope during a prolonged lowstand (Fig. 4.3). This unit<br />

comprises a thick series of hemi-pelagic dark shale found in various onshore and offshore wells.<br />

In the southern coastal plain it fills a 1 km deep canyon, incised within the Jurassic shelf (Cohen,<br />

1976). <strong>The</strong> Gevar'am Canyon was the main conduit for submarine turbidite flows that<br />

transported significant amounts of quartztoze sands into the basin from the elevated area on the<br />

southeast. Several sand beds up to 20 m in thickness, found within the Gevar'am shale offshore<br />

(Yam-2 and Yam West-1 wells, Fig. 7.3), are interpreted as distal slope fans and questionably<br />

basin floor fans that are associated with the Gevar'am Canyon and channel system (Gardosh,<br />

2002). Lower Cretaceous deepwater slope and basin floor fans are likely to be found in other<br />

parts of the southern and central <strong>Levant</strong> <strong>Basin</strong> (e.g., Mango-1 well, Fig. 1.2).<br />

<strong>The</strong> Middle Cretaceous section of the <strong>Levant</strong> margin and basin, although dominated by<br />

carbonate strata, show evidence of clastic sediment transport. <strong>The</strong> Aptian-Albian slope deposits<br />

of the Talme Yafe Formation contain coarse-grained carbonate breccias found in various wells in<br />

the coastal plain (Cohen, 1971; Bein and Weiler, 1976). <strong>The</strong>se were probably eroded from the<br />

shelf and deposited by gravity flows on the slope during short term Aptian –Albian lowstands<br />

(Gardosh, 2002).<br />

Another Middle Cretaceous erosional feature of the <strong>Levant</strong> slope is the Item Canyon.<br />

<strong>The</strong> Item Formation is a 2000 m thick carbonate series of Cenomanian age (Fig. 4.3), found in<br />

the Item-1 well (Fig. 1.3). <strong>The</strong> origin of this unit was previously not fully understood. <strong>The</strong> new<br />

offshore seismic data provides evidence for erosion at the base of the unit, indicating deposition<br />

within a deep submarine slope canyon (Fig. 4.6). <strong>The</strong> geometry of the Item Canyon is at present<br />

not well resolved because of the later Syrian Arc deformation.<br />

A pronounced middle Turonian lowstand associated with karstic phenomena and<br />

quartztose sandstone deposition is well documented in the shallow-marine platform from the<br />

90


Pliocene-Pleistocene<br />

Messinian<br />

(Evaporites)<br />

M.-U. Miocene<br />

Oligocene-L. Miocene<br />

Senonian-Eocene<br />

M. Jurassic-Turonian<br />

Permian-M. Jurassic<br />

Crystalline Basement<br />

a<br />

Fig 8.1- Convergence stage-the Tertiary fill of the <strong>Levant</strong> <strong>Basin</strong>: (a) Stratigraphic<br />

section through the <strong>Levant</strong> <strong>Basin</strong> and Margin, and (b) isopach map of the<br />

Oligocene to Upper Miocene interval. <strong>The</strong> great thickness of the<br />

Tertiary section in the basin (more then 3 km) resulted from tectonic<br />

uplift, eustatic sea-level falls and intense erosion and mass transport<br />

across the <strong>Levant</strong> slope. See location of the section on the isopach map.<br />

b<br />

Yam Yafo-1<br />

91


-Long distance transport of<br />

Nubian sandstone<br />

a<br />

- Long distance transport of<br />

Nubian, Hazeva and<br />

Hordos sandstone<br />

b<br />

92


- Short distance transport of<br />

Hazeva and Hordos sandstone<br />

c<br />

- Short distance transport of<br />

Hazeva sandstone<br />

- Long distance transport of<br />

Nilotic sands<br />

<br />

d<br />

Fig 8.2- <strong>The</strong> Tertiary, submarine canyon and channel system of the <strong>Levant</strong> margin, shown on the<br />

depth maps of: (a) base Oligocene, (b) top Lower Miocene (c) base Messinian and (d) top<br />

Messinian. <strong>The</strong> main directions of siliciclastic sediment supply are marked with black arrows.<br />

93


Negev to the Galilee area (Fig. 4.3) (Sandler, 1996). It is possible that Turonian sands bypassed<br />

the shelf and were transported westward by submarine turbidite flows to the slope and basin<br />

areas, although these rocks were not encountered yet in any offshore wells<br />

8.2 Convergence stage<br />

8.2.1 Submarine canyon morphology<br />

Extensive erosion and incision on the <strong>Levant</strong> slope, followed by the accumulation of<br />

several kilometers thick Tertiary fill within the basin took place during the convergence stage<br />

(Fig. 8.1). <strong>The</strong>se phenomena reflect the combined effects of tectonic uplifting and tilting of the<br />

<strong>Levant</strong> margin and the mountainous backbone of <strong>Israel</strong>, the uplifting of the Arabian Shield and<br />

eustatic sea-level drops (Fig. 4.4). A highly developed submarine canyon and channel system<br />

evolved during the Early Oligocene and persisted till the Late Pliocene (Figs. 8.2, 8.3). <strong>The</strong><br />

system included several main conduits generally oriented in NE-SW and E-W directions. <strong>The</strong>se<br />

are, from south to north: the Afiq, Ashdod, Hadera, Caeserea, Atlit and Qishon canyons (Fig.<br />

8.2). Some of these erosional features were previously identified in onshore wells (the Afiq and<br />

Ashdod canyons, Gvirtzman and Buchbinder, 1978; Druckman et al., 1995b; Buchbinder and<br />

Zilberman, 1997). <strong>The</strong> offshore seismic data reveal the extent of the Tertiary drainage system<br />

along the entire eastern part of the <strong>Levant</strong> <strong>Basin</strong> up to 70 km west of the present-day coastline.<br />

<strong>The</strong> system is built of superimposed canyon incisions revealed by the topography of the base<br />

Oligocene, Lower Miocene, base Messinian and top Messinian unconformities (Figs. 8.2a-d).<br />

<strong>The</strong> major canyons are more or less vertically stacked and are straight to slightly<br />

curvilinear. Secondary, subsequent channels confined by the morphology of pre-existing fold<br />

structures flowed into the main canyons. <strong>The</strong> Afiq and Ashdod canyons are deeply incised. <strong>The</strong><br />

depth of the Afiq Canyon is up to 1500 m onshore (Druckman et al., 1995b) and several hundred<br />

meters offshore (Fig. 8.3b). <strong>The</strong> Ashdod Canyon cuts a 700 m-deep gorge into Mesozoic strata<br />

some 40 km offshore (Fig 8.3a). <strong>The</strong> incisions in the northern channels (Hadera, Caesarea, Atlit<br />

and Qishon) are limited to the upper slope, near the present-day coastline.<br />

8.2.2 Oligocene to Lower Miocene canyon system<br />

<strong>The</strong> Oligocene tectonic activity and sea-level falls resulted in considerable shedding of<br />

clastic material into the <strong>Levant</strong> <strong>Basin</strong>. Denudation was extensive, exposing Nubian sandstone<br />

terrain in far proximal areas. Widespread transport of clastic material initiated a system of<br />

submarine canyons, depicted by the Red seismic horizon (Fig. 8.2a). Oligocene continental<br />

deposits are absent in <strong>Israel</strong>. Apparently, clastic material was not stored inland but was directly<br />

transported to the basin.<br />

94


Approximately 250 m of Oligocene coarse-grained clastics (sandstones and<br />

conglomerates, named "Ashdod Clastics" ) were penetrated in the Ashdod wells, located in the<br />

southern coastal plain where the Ashdod Canyon meets the present-day shoreline (Fig. 8.2a)<br />

(Buchbinder and Siman-Tov, 2000; Buchbinder et al., 2005). <strong>The</strong> age of the Ashdod Clastics is<br />

not strictly defined. It may range from upper Rupelian to Lower Chattian (zones P19 to P21 in<br />

Fig. 4.4; Buchbinder et al., 2005). Core #2 from the Ashdod-2 well reveals a conglomerate<br />

consisting of boulders, imbricated pebbles (mostly of Cretaceous limestones and dolostones) and<br />

bioclastic sandstones of turbidite origin. Two intervals of highly porous sandstones, 20 m to 50m<br />

thick, were detected in the Ashdod Clastics section. <strong>The</strong>ir Spontaneous Potential (SP) and<br />

Gamma-Ray (GR) log response exhibit blocky to fining-upward patterns. Other parts of the<br />

section show a serrated log motif, reflecting interbded conglomerates, sandstones and marls<br />

(Buchbinder et al., 2005). <strong>The</strong> clastic succession is interpreted as canyon fill or proximal slope<br />

fan, deposited within the submarine, Ashdod Canyon that was formed during the lower<br />

Oligocene (Figs. 4.4, 8.2a)<br />

Although no significant sandy deposits were encountered in onshore wells along the Afiq<br />

Canyon (Fig 8.2a), Martinotti, (1981) reported a Cretaceous boulder within hemi-pelagic middle<br />

Oligocene sediments in the Gaza-1 well. Apparently, sands either bypassed the present-day,<br />

onshore part of the Afiq canyon and were deposited offshore, or were eroded during the repeated<br />

canyon entrenchments in the Miocene.<br />

According to new findings from the Qishon-Yam-1 well (Fig. 1.2) (Schattner et al.<br />

2006), the Qishon-Yizrael valley in the north part of the country was formed already in<br />

Oligocene times. <strong>The</strong>refore, clastic sediments from the Damascus- Ein Gev <strong>Basin</strong>s could have<br />

been transported through the Qishon and Atlit Canyons to the <strong>Levant</strong> <strong>Basin</strong> (Fig 8.2a) (see<br />

below).<br />

In the offshore area, high-amplitude, discontinuous seismic events that were identified<br />

within Oligocene canyons may be interpreted as stacked, migrating channel systems (Fig 8.3a,c).<br />

This seismic pattern probably indicates transported, coarse clastic material. It is highly probable<br />

that clastic material was further transported into the basin and was deposited as sandy basin-floor<br />

fans. On the seismic data these may correspond to high-amplitude seismic events, which were<br />

identified above the Red horizon near the distal outlets of the main canyons (Fig. 8.3d).<br />

<strong>The</strong> Lower Miocene is characterized by a prolonged lowstand (~6 my) indicated by the<br />

absence of marine Burdigalian sediments inland (Fig. 4.4). This lowstand occurred during a<br />

period of global eustatic rise of sea-level (Haq et al., 1977; 1978, Fig. 4.4) and may thus indicate<br />

a tectonic uplift (Syrian Arc II phase). <strong>The</strong> lowstand was accompanied by intensive erosion and<br />

transportation of clastic material into the basin. A Lower Miocene sand interval, 17 m thick, was<br />

95


SW<br />

NE<br />

Ashdod Canyon<br />

TWT(ms)<br />

a<br />

SW<br />

NE<br />

Afiq Canyon<br />

TWT(ms)<br />

b<br />

96


W<br />

E<br />

TWT(ms)<br />

1<br />

c<br />

SW<br />

NE<br />

TWT(ms)<br />

2<br />

2<br />

d<br />

Fig 8.3- Tertiary submarine incision and deposition of clastic sediments in the <strong>Levant</strong> <strong>Basin</strong>:<br />

(a) the incised Ashdod Canyon; (b) the incised Afiq Canyon; (c) high-amplitude,<br />

discontinuous seismic reflections within the Hadera channel (1), interpreted<br />

as stacked, migrating channel systems; and (d) high-amplitude seismic reflections<br />

near the mouth of the Ashdod Canyon(2), interpreted as basin-floor fan lobes.<br />

97


encountered in the Gaza-1 well, in the southern coastal plain (Figs.1.3, 4.4). It may represent the<br />

proximal tail of much more extensive accumulation in a distal basin position. <strong>The</strong> seismic data<br />

shows high-amplitude events in the upper part of the Oligocene to top Lower Miocene interval,<br />

below the Cyan horizon (Figs. 8.3b,d). <strong>The</strong>se seismic events may correspond to clastic<br />

accumulations in a distal basin position.<br />

Notably, continental clastic sediments of the Hazeva and Hordos formations started to<br />

accumulate in proximal inland areas in Early Miocene times. <strong>The</strong> Hordos sedimentation in the<br />

Galilee initiated before 17 my (Shaliv, 1991) and the Hazeva Group in the Negev started to<br />

accumulate before 20 my (Calvo and Bartov, 2001). <strong>The</strong> Hazeva continental deposition was<br />

occasionally interrupted by Syrian Arc or other tectonic movements which may correspond to<br />

the significant erosional truncation event of the Zefa Member (Calvo and Bartov, 2001). This<br />

event could have resulted in a long distance transportation of clastic material from southern<br />

<strong>Israel</strong> into the basin through the Afiq and Asdod Canyons (Fig. 8.2a,b).<br />

In the north, the Nukev, En Gev and Hordos sands (of the Golan area) could have also<br />

been remobilized and transported through the Qishon Valley to the <strong>Levant</strong> basin. According to<br />

Shaliv (1991) the similarities between the En-Gev sands and the continental sediments of the<br />

Damascus basin suggest that the two sites are, in fact, parts of the same basin. This indicates the<br />

existence of extensive source for siliciclastics in the north (Fig. 8.2.a,b)<br />

In summary, the Oligocene to Lower Miocene period is characterized by widespread<br />

transport in submarine canyon and channel system along the entire <strong>Levant</strong> margin. <strong>The</strong><br />

provenance of the clastic material was long distance sites of Nubian sandstone (which was<br />

already exposed in Sinai, Negev, and Trans-Jordan) during the early Oligocene phase, and both<br />

Nubian sandstone and remobilized Hazeva and Hordos sands during early Miocene times (Fig.<br />

8.2a,b).<br />

8.2.3 Middle to Upper Miocene canyon system<br />

<strong>The</strong> Serravallian to early Tortonian time is characterized by a long term lowstand<br />

(Serravallian crisis). This lowstand is associated with the Bet Nir Conglomerate found in<br />

outcrops inland; and probably also with 250 m thick conglomerate interval found in the Nahal<br />

Oz-1 well, located in a proximal part of the Afiq Canyon, at the southern coastal plain (Fig.<br />

8.2b). Druckman et al. (1995b) described in this part of the canyon erosion and slumping event<br />

of Langhian to Serravallian age (2 nd erosion; Druckman et al., 1995, figure 4). Further updip, the<br />

Beer Sheva Canyon was excavated at this time and later filled by Late Miocene sediments of the<br />

Pattish cycle (Fig. 4.4) (Buchbinder and Zilberman, 1997).<br />

98


<strong>The</strong> Serravallian lowstand is associated with submarine incision on the <strong>Levant</strong> slope,<br />

depicted by the Cyan seismic horizon (Fig. 8 2b). <strong>The</strong> canyon pattern closely followed the older,<br />

base Oligocene system although the degree of incision was reduced (Fig. 8.3a,c). Coarse clastic<br />

material was transported basinward through the Middle to Upper Miocene canyon system (Fig.<br />

8.2b). Discontinuous, high-amplitude seismic events found above the Cyan horizon (Fig. 8.3b)<br />

may be associated with channel-fill deposits and basin floor fans. Evidence from several wells<br />

recently drilled in the Afiq Canyon offshore support this interpretation. <strong>The</strong> provenance of the<br />

siliciclastic material was long distance sites of Nubian sandstone and remobilized Hazeva and<br />

Hordos sands (Fig. 8.2b).<br />

8.2.4 <strong>The</strong> Base-Messinain-evaporite canyon system<br />

Substantial subaerial and submarine denudation of the North African and <strong>Levant</strong><br />

continental margin took place during the Messinian salinity crisis and lowstand (Fig. 4.4) (Ryan,<br />

1978; Gvirtzman and Buchbinder, 1978). On the <strong>Levant</strong> slope the base Messinain canyon<br />

system is depicted by the Purple seismic horizon (Fig. 8.2c). <strong>The</strong> degree of incision at the base<br />

of the evaporite section is smaller then previous erosion phases (Fig. 8.3a,c) possibly due to<br />

reduced runoff during this arid period.<br />

<strong>The</strong> mountainous backbone of <strong>Israel</strong> was well developed at this time forming a<br />

topographic barrier for long distance transport. A possible source for siliciclastics could have<br />

been the Hazeva and Hordos sands that where eroded from elevated areas near the coast (Fig.<br />

8.2c). <strong>The</strong>re is no indication for clastic accumulation at the base Messinain level inland.<br />

However, evidence from several wells recently drilled in the Afiq Canyon offshore indicates<br />

some coarse clastic deposition.<br />

8.2.5 <strong>The</strong> top-Messinian-evaporite canyon system and the Pliocene gas-bearing sands<br />

A latest Miocene to Early Pliocene lowstand, associated with erosion of the Messinian<br />

evaporites is identified in the proximal part of the Afiq Canyon inland (Druckman et al., 1995b,<br />

5 th erosion in figure 4). Druckman et al. (1995b) noted that this lowstand is followed by fluvial<br />

erosion and flooding of the entire Mediterranean with fresh and brackish water during the socalled<br />

'Lago Mare' event (Rouchy and Saint Martin, 1992). In the <strong>Levant</strong> <strong>Basin</strong> the top<br />

Messinian erosion is depicted by the Violet seismic horizon (Fig. 8.2d).<br />

A short-term episode of marine deposition of the lowermost Pliocene was followed by<br />

another significant lowstand at 4.37 m.y. (Fig. 4.4). Deepwater sands that are found above the<br />

Mavkiim Formation in the offshore distal part of the Afiq canyon are separated from the<br />

99


evaporites by several tens of meters thick layer of hemipelagic marl; and are therefore related to<br />

the second, Early Pliocene lowstand of 4.37 m.y.<br />

<strong>The</strong> thick accumulations of deepwater turbiditic sands, termed the Lower Member of the<br />

Yafo Formation or the "Noa Sand" form the reservoir of the biogenic gas in the Noa, Mari and<br />

Gaza Marine fields (Fig. 1.2). <strong>The</strong>se prolific gas sands represent a basin floor fan environment<br />

within the Afiq and El Arish Canyons. Some of the sands form large mounds up to 250 m in<br />

thickness. <strong>The</strong> formation of the mounds is related to postdepositional, remobilization of the<br />

deepwater sand, due to fluid injection in overpressure conditions (Oats, 2001; Frey-Martinez et<br />

al., 2007).<br />

A likely source for the Lower Pliocene siliciclastics is the Miocene Hazeva sandstone<br />

that was exposed at that time and was eroded and transported to a short distance from the Negev<br />

and northern Sinai area (Fig. 8.2d). Another possible source is 'Nilotic' sands that were<br />

transported from the Nile Delta by longshore currents and eventually trapped and transported<br />

offshore in the Afiq canyon (Fig. 8.2d). <strong>The</strong> Yafo sands were so far encountered only in the<br />

southern part of the basin. This may be explained by the lack of adequate source of siliciclastics<br />

in the northern area during Pliocene times.<br />

9. Hydrocarbon Potential<br />

9.1 Trap Types and Hydrocarbon Plays<br />

<strong>The</strong> integrated analysis of the <strong>Levant</strong> <strong>Basin</strong> presented in this report indicates a high<br />

potential for hydrocarbon accumulation. Each of the main stratigraphic intervals described and<br />

mapped in this report contains various types of potential structural and stratigraphic traps.<br />

Suggested hydrocarbon plays within each interval are shown in Figure 9.1 and are briefly<br />

described at the following sub-sections.<br />

9.1.1 Permian to Middle Jurassic interval<br />

This period is characterized by shallow-marine deposition and the formation of extensive<br />

graben and horst systems across the <strong>Levant</strong> region. <strong>The</strong> main hydrocarbon plays suggested for<br />

this interval are Triassic and Jurassic fault-related traps (Fig. 9.1). Alluvial fans (Erez<br />

Conglomerate type) and other coarse-grained clastic deposits may be found near the main faults<br />

or within the Triassic and Jurassic grabens. Buried hill-type traps or carbonate buildups may be<br />

found at the tops of the Triassic- Lower Jurassic highs, sealed by fine-grained Upper Jurassic and<br />

100


W Biogenic gas<br />

discoveries<br />

E<br />

Plio-Pleistocene<br />

Messinian<br />

M.-U.Miocene<br />

Oligocene-L. Miocene<br />

M. Jurassic-Turonian<br />

Permian-M. Jurassic<br />

High-grade<br />

Oil shows<br />

Trap type:<br />

Alluvial fan Buried-hill<br />

Channel-fill<br />

Turbidite lobes and sheets<br />

Carbonate buildup<br />

Fig 9.1- Schematic section through the <strong>Levant</strong> <strong>Basin</strong> showing<br />

potential hydrocarbon traps in the Phanerozoic fill.<br />

Pliocene gas discoveries in the southern part of the basin<br />

and high-grade oil shows in the Mesozoic section<br />

indicate the existence of biogenic and thermogenic<br />

petroleum systems.<br />

101


Cretaceous strata. <strong>The</strong>se features, which are deeply buried within the basin, are located at a<br />

shallower depth (5-6 km) near the eastern margin.<br />

9.1.2 Middle Jurassic to Turonian interval<br />

This period is characterized by shallow-marine conditions during its early part, followed<br />

by the deposition of carbonate platforms on the margin and deepwater turbidites in the basin<br />

during the passive margin stage. <strong>The</strong> main hydrocarbon plays suggested for this interval are<br />

Lower Cretaceous deepwater fans (Fig. 9.1). <strong>The</strong> Early Cretaceous Gevara'am canyon was part<br />

of a submarine canyon and channel system incised on the continental slope in the southeastern<br />

<strong>Levant</strong> margin. Several sand layers that where deposited as deepwater fans at the distal part of<br />

this system were penetrated by offshore wells (e.g. Yam-2 and Yam West-1) (Gardosh, 2002).<br />

Similar sand bodies can be expected in other parts of the basin. It should be noted that Early<br />

Cretaceous sand bodies pre-dated the Syrian Arc deformation stage and therefore, are equally<br />

distributed on highs and lows.<br />

<strong>The</strong> early Middle Jurassic period is characterized by the accumulation of oolitic shoals<br />

and fine-grained carbonate debris. <strong>The</strong>se may be considered as an additional play type for this<br />

interval. Light oil shows were discovered in the Yam-2 and Yam Yafo-1 wells in shallowmarine,<br />

Middle Jurassic reservoirs trapped within Syrian Arc folds. Likewise the Lower<br />

Cretaceous fans, the distribution of Middle Jurassic porous intervals is not controlled by Syrian<br />

Arc deformation.<br />

9.1.3 Oligocene to Lower Miocene interval<br />

This period is characterized by contractional deformation and uplift that was followed by<br />

intense erosion and transport of clastic sediments into the basin. An extensive canyon and<br />

channel system developed during the Oligocene throughout the eastern part of the <strong>Levant</strong> <strong>Basin</strong><br />

(see chapter 8.2.1). <strong>The</strong> main hydrocarbon play suggested for this interval consists of Oligocene<br />

channel-fill and deepwater fans (Fig. 9.1). Potential taps are channel-fill units interpreted on the<br />

seismic data in up-dip position (confined setting), and basin floor fans interpreted on the seismic<br />

data at the distal end of the canyons (non-confined setting) (Fig. 8.3). Some of the fans may be<br />

found in structurally favored locations within Syrian Arc II type folds.<br />

A unique feature that developed during this period is found at the southern part of the<br />

Jonah Ridge (Figs. 5.11, 9.2). <strong>The</strong> seismic data show a large, layered mound located above a<br />

chaotic package interpreted as a Cretaceous to Early Tertiary volcanic cone, which is<br />

superimposed on a Triassic-Jurassic high (Fig. 5.11). This mound is interpreted as an atoll of<br />

possibly upper Early Miocene age, equivalent to the Ziqlag reef and shelf carbonates found on<br />

102


the margin. This and other Miocene carbonate buildups that may be found in the basin, are<br />

suggested as an additional play in the Oligocene to Lower Miocene interval (Fig. 9.1). A similar<br />

carbonate buildup was likewise interpreted by Aal et al. (2000) in the ultra-deepwater of the Nile<br />

Delta, offshore Egypt.<br />

9.1.4 Middle to Upper Miocene interval<br />

This period is characterized by the continuation of contractional deformation, erosion and<br />

sediment transport into the basin (see chapter 8.2.1). <strong>The</strong> main hydrocarbon play suggested for<br />

this interval consists of Miocene channel-fills and deepwater fans similar to the above described<br />

Oligocene play (Fig. 9.1). Combinated, stratigraphic and structural traps are expected to be<br />

found. Deepwater channels at the base of the Messinian salt are an additional play that should be<br />

considered in this interval (Fig. 9.1)<br />

9.1.5 Pliocene interval.<br />

Sediment transport into the basin through a submarine canyon and channel system took<br />

place during the Early Pliocene (see chapter 8.2.1). Pliocene deepwater fans charged with<br />

biogenic gas were discovered in the Afiq Canyon. This play should be further explored in other<br />

Pliocene canyons, particularily in the southern part of the basin (Figs. 8.2d, 9.1).<br />

9.2 Source Rocks and Petroleum Systems<br />

An important aspect of the <strong>Levant</strong> <strong>Basin</strong> hydrocarbon potential is the distribution and<br />

maturation level of source rocks. Producing fields and hydrocarbon shows found in the basin<br />

and on its margins indicate the existence of two types of petroleum systems: biogenic and<br />

thermogenic. <strong>The</strong> organic rich shales of the Plio-Pleistocene, Oligo-Miocene and Eocene all<br />

have biogenic gas potential (McQuilken, 2001). <strong>The</strong> Middle Miocene Qantara Formation is<br />

considered by Dolson et al. (2002) as a major Tertiary source in the offshore Nile Delta. <strong>The</strong><br />

Sadot and Shiqma gas fields (Fig. 1.2) found in the southern coastal plain are probably<br />

associated with an Oligo-Miocene source of biogenic gas. <strong>The</strong> origin of the gas found in the<br />

Noa, Mari and Gaza Marine fields is considered to be Miocene and Pliocene organic rich shale,<br />

although a particular source rock for the Pliocene biogenic gas could not be explicitly established<br />

(Feinstein et al., 2002). <strong>The</strong> great thickness and wide distribution of the hemipelagic Tertiary<br />

section, however, suggest good potential for biogenic gas generation throughout the <strong>Levant</strong><br />

<strong>Basin</strong>.<br />

Mesozoic source rocks comprise theromgenic petroleum systems. <strong>The</strong> Upper Cretaceous,<br />

organic rich marl of the Mount Scopus Group have excellent source properties in the inland part<br />

103


of <strong>Israel</strong> and generate oil and thermogenic gas in the Dead Sea basin (Tannenbaum and<br />

Aizenshtat, 1985). <strong>The</strong>rmal maturity modeling shows that the Upper Cretaceous section reaches<br />

maturation within the <strong>Levant</strong> <strong>Basin</strong> at depths greater than 4 km (Gardosh, 2002). <strong>The</strong>refore,<br />

Senonian strata should be considered as a potential source for oil and thermogenic gas in the<br />

deep part of the basin.<br />

<strong>The</strong> Lower Cretaceous Gevar'am shale has source rock properties and may have<br />

generated oil and thermogenic gas where maturity is reached (McQuilken, 2001). <strong>The</strong> Middle<br />

Jurassic Barnea Formation is the source of oil in the Helez field, onshore (Fig 1.2) (Bein and<br />

Soffer, 1987). <strong>The</strong> organic rich limestone of the Barnea Formation was so far penetrated in the<br />

southern coastal plain. If the organic facies of the Barnea Formation extends into the basin (Fig.<br />

4.5) it may serve as an important source of hydrocarbons.<br />

Lower Jurassic and Triassic continental to shallow-marine rocks were deposited<br />

throughout the <strong>Levant</strong> area during the Neotethyan rifting stage. Bein et al.(1984) found source<br />

properties in this section in various deep, onshore wells. It is assumed that the organic content in<br />

Lower Jurassic and Triassic strata was higher in the paleograbens offshore, where lacustrine to<br />

deeper marine conditions may have prevailed. High-grade oil shows found in the Middle<br />

Jurassic limestone in the Yam-2 and Yam Yafo-1 well are probably related to these source rocks.<br />

<strong>The</strong> timing of oil generation of the Early Mesozoic rock units is not well established.<br />

McQuilken (2001) estimated that Lower to Middle Jurassic source rocks are presently in the<br />

peak oil window to just within the gas window offshore. However, Gardosh (2002) estimated<br />

that the Triassic rocks reached the maturity window in Late Jurassic to Middle Cretaceous time.<br />

<strong>The</strong> later results further suggest that the primary migration of Early Mesozoic hydrocarbons may<br />

have taken place prior to the onset of Syrian Arc folding phase, and therefore they should be<br />

found in Early Mesozoic fault blocks and stratigraphic traps.<br />

Recently discovered oil in the onshore Meged wells, located northeast of Helez (Fig. 1.2),<br />

was related to Silurian source rocks (www.givot.co.il). This is the first occurrence of such oil in<br />

the <strong>Levant</strong> region. It may be hypothesized that Silurian source rocks have been preserved, and<br />

generated hydrocarbons within Neotethyan rift structures such as the Judea Graben onshore and<br />

the deep-seated grabens offshore (Fig 7.1).<br />

In summary, a wide spectrum of biogenic and thermogenic petroleum systems, ranging in<br />

age from Paleozoic to Plio-Pleistocene is found in the <strong>Levant</strong> <strong>Basin</strong>. <strong>The</strong> situation offshore<br />

<strong>Israel</strong> is probably similar to that offshore the Nile Delta where deep structures serve as focal<br />

points for vertical hydrocarbon migration, resulting in a mix of biogenic and thermogenic gases<br />

in shallow structural levels (Dolson et al., 2002, Feinstein et al., 2002).<br />

104


10. Summary<br />

Modern, geophysical data yielded new information on the structure and stratigraphy of<br />

the <strong>Levant</strong> <strong>Basin</strong>, offshore. <strong>The</strong> integration of these data with the vast amount of information<br />

from the <strong>Levant</strong> margin and inland part of <strong>Israel</strong> enables the reconstruction of a regional<br />

geologic scheme. Three distinct tectonic stages in the evolution of the region are documented:<br />

(a) Rifting stage<br />

(b) Post-rift passive margin stage<br />

(c) Convergence stage<br />

Rifting in the <strong>Levant</strong> region is related to the breakup of the Gondwana plate and the<br />

formation of the Neotethys Ocean system. Four extensional pulses are postulated: in the latest<br />

Paleozoic; Middle Triassic; Late Triassic; and Early Jurassic. <strong>The</strong>se pulses resulted in normal<br />

faulting in the range of several kilometers, magmatic activity and the formation of NE-SW<br />

oriented graben and horst systems. Four basement structures associated with the Early Mesozoic<br />

extension are found in the deep parts of <strong>Levant</strong> <strong>Basin</strong> (from east to west): the Yam, Jonah,<br />

Leviathan, and the Eratosthenes highs. <strong>The</strong> main structures inland are the Gevim, Gaash and<br />

Maanit highs and the Hilal, Judea and Asher grabens. <strong>The</strong> rifting in the <strong>Levant</strong> area reached an<br />

early magmatic stage. Although magmatic intrusion and stretching of the crust took place, no<br />

indications for sea-floor spreading and emplacement of new oceanic crust are found. Continental<br />

to shallow-marine depositional environments were dominant in the <strong>Levant</strong> region during the<br />

rifting stage.<br />

<strong>The</strong> post-rift stage is associated with cooling and subsidence that was probably more<br />

intense within the basin than on the eastern margin. <strong>The</strong> Late Jurassic to Middle Cretaceous<br />

section records the gradual formation of a passive-margin profile and the subsequent<br />

development of a deep marine basin bordered by a shallow-marine shelf. <strong>The</strong> passive-margin<br />

stage is characterized by recurring cycles of marine transgression and regression associated with<br />

relative sea-level changes. <strong>The</strong>se are reflected by marine onlaps, numerous unconformity<br />

surfaces, accumulation of carbonate platforms on the margin and mass transported deposition in<br />

the basin.<br />

<strong>The</strong> convergence stage is related to the closure of the Neotethyan Ocean system and the<br />

motion of the Afro-Arabian plate towards Eurasia. In the <strong>Levant</strong> region this motion is<br />

manifested by large-scale contractional deformation of the Syrian Arc fold belt. Two<br />

contractional phases are observed. A Late Cretaceous Syrian-Arc I phase is characterized by<br />

inversion of older normal faults and the development of asymmetric, high-amplitude folds that<br />

are found mostly near the eastern margin and further inland. An Oligo-Miocene Syrian-Arc II<br />

105


phase is characterized by the formation of low-amplitude folds throughout the basin, and<br />

uplifting and tilting of marginal blocks, at the eastern part of the basin and further inland.<br />

Teriary tectonic activity and eustatic sea-level falls scaused intense erosion and<br />

accumulation of several kilometers thick basin-fill. Fluvial systems flowed on the exposed shelf<br />

from elevated areas in the east. Submarine turbidity currents incised the upper slope and<br />

transported siliciclastics further into the basin. <strong>The</strong> Oligo-Miocene and Pliocene drainage<br />

system includes, from south to north, the Afiq, Ashdod, Hadera, Casarea, Atlit and Qishon<br />

canyons. <strong>The</strong> Afiq and Ashdod canyons are highly erosive and are incised several hundred<br />

meters in the continental slope. Other canyons are less erosive and their flow pattern was for the<br />

most part controlled by the topography of the paleoslope. <strong>The</strong> Tertiary drainage system was shut<br />

off temporarily during the Messinian salinity event and turned on again during post-evaporites,<br />

latest Miocene to Early Pliocene lowstands.<br />

<strong>The</strong> variety in tectonic styles and depositional patterns may provide favorable trapping<br />

conditions for hydrocarbons in the <strong>Levant</strong> <strong>Basin</strong>. Potential structural traps are associated with the<br />

extensional rift structures and the contractional Syrian Arc folds. Stratigraphic traps are<br />

associated with Triassic-Middle Jurassic shallow-marine, carbonate and siliciclastic reservoirs<br />

and Cretaceous and Tertiary deepwater turbidite systems.<br />

Shallow gas discoveries in Pliocene sands and high-grade oil shows found in the<br />

Mesozoic section indicate the presence of source rocks and appropriate conditions for<br />

hydrocarbon generation in both biogenic and thermogenic petroleum systems. <strong>The</strong> size, depth<br />

and trapping potential of the <strong>Levant</strong> <strong>Basin</strong> supports the conclusions that large quantities of<br />

hydrocarbons can be found offshore <strong>Israel</strong>.<br />

This study was focused on the stratigraphy, structure and tectonic evolution of the <strong>Levant</strong><br />

<strong>Basin</strong>. It is recommend to conduct specific studies on the following topics: (a.) further analysis<br />

of the Tertiary drainage systems in wells, outcrops and 3D seismic data; (b) further analysis of<br />

the Middle Jurassic to Middle Cretaceous reservoir rocks and depositional environments, (c.) 2D<br />

and 3D structural reconstruction of the basin; (d.) modeling the thermal history and migration<br />

paths for hydrocarbons; and (e.) study of Direct Hydrocarbon Indicators (DHI) in 2D and 3D<br />

seismic data.<br />

106


Acknowledgments<br />

<strong>The</strong> authors would like to thank B. Conway, L. Perelis-Grossowicz , R. Siman-Tov and<br />

S. Lipson-Benitah for reviewing paleontological data of selected Jurassic, Cretaceous and<br />

Tertiary intervals. We thank B. Cohen and N. Almog for drafting maps and geologic sections,<br />

and B. Katz for editing the manuscript. <strong>The</strong> help of R. Calvo in data organization is much<br />

appreciated. We thank Y. Ben-Gai and A. Sneh for fruitful discussions and P. Weimar for<br />

helpful remarks on the Tertiary drainage system.<br />

107


References:<br />

Aal, A.A., El Barkooky, A., Gerrits, M., Meyer, H., Schwander, M. and Zaki, H., 2000. <strong>Tectonic</strong><br />

evolution of the Eastern Mediterranean <strong>Basin</strong> and its significance for hydrocarbon prospectivity in<br />

the ultra-deepwater of the Nile Delta. <strong>The</strong> Leading Edge, 1086-1102.<br />

Abed, A. M., 1984. Emergence of Wadi Mujib (Central Jordan) during Lower Cenomanian time and its<br />

regional tectonic implications. Geol. Soc. London, Spec. Publ. 17, 213-216.<br />

Almogi-Labin , A., Bein, A. and Sass, E., 1993. Late Cretaceous upwelling system along the southern<br />

Tethys margin (<strong>Israel</strong>): interrelationship between productivity, bottom water environments and<br />

organic matter preservation. Paleoceanogr. 8, 671-690.<br />

Avedik, F., Nicolich, R., Hirn, A., Maltezou, F., McBride, J. and Cernobori, L. and the<br />

STREAMERS/PROFILES Group, 1995. Appraisal of a new, high-energy and low frequency<br />

seismic pulse generating method on a deep seismic reflection profile on the Central Mediterranean<br />

Sea. First Break 13, 277-290.<br />

Bachmann, M. and Hirsch, F., 2006. Lower Cretaceous carbonate platform of the eastern <strong>Levant</strong> (Galilee<br />

and the Golan Heights): stratigraphy and second-order sea-level change. Cretaceous Res. 27, 487-<br />

512.<br />

Ball, M.W. and Ball, D., 1953. Oil prospects of <strong>Israel</strong>. Am. Assoc. Petrol. Geol. Bull. 37. 1-113.<br />

Bartov, Y., 1974. A structural and paleogeographical study of the central Sinai faults and domes. Ph.D.<br />

<strong>The</strong>sis, Hebrew Univ., Jerusalem, (in Hebrew, English abstract).<br />

Bein, A., 1977. Shelf basin sedimentation: mixing and diagenesis of pelagic and clastic Turonian<br />

carbonates, <strong>Israel</strong>. J. Sediment. Petrol. 47, 382-391.<br />

Bein, A., S. Feinstein, A. Aizenshtat, and Y. Weiler, 1984. Potential source rocks in <strong>Israel</strong>: a<br />

geochemical evaluation, Isr. Geol Survey Rep. GSI/17/84 and Oil Exploration<br />

(Investment) Ltd., Rep. 84/15, 46 p.<br />

Bein, A. and Gvirtzman, G., 1977. A Mesozoic fossil edge of the Arabian plate along the <strong>Levant</strong> coastline<br />

and its bearing on the evolution of the eastern Mediterranean, In: Biju-Duval B.and Montadert, L.<br />

(eds.) Structural History of the Mediterranean <strong>Basin</strong>s: Editions Technip, Paris, 95-110.<br />

Bein, A. and Soffer, Z., 1987. Origin of oils in the Helez region, <strong>Israel</strong>: implications for exploration<br />

in the eastern Mediterranean, Am. Assoc..Petrol. Geol. Bull. 71, 65-75.<br />

Bein, A. and Weiler, Y., 1976. <strong>The</strong> Cretaceous Talme Yafe formation: a contour shaped sedimentary<br />

prism of calcareous detritus at the continental margin of the Arabian craton, Sedimentology 23, 511-<br />

532.<br />

Ben-Avraham, Z. and Ginzburg, A., 1990. Displaced terranes and crustal evolution of the <strong>Levant</strong> and the<br />

eastern Mediterranean. <strong>Tectonic</strong>s 9(4), 613-622.<br />

Ben-Avraham, Z., Ginzburg, A., Makris, J. and Eppelbaum, L., 2002. Crustal structure of the <strong>Levant</strong><br />

<strong>Basin</strong>, eastern Mediterranean. Tectonophysics 346, 23-43.<br />

Ben-Avraham, Z., Shoham,Y. and Ginzburg, A., 1976. Magnetic anomalies in the Eastern Mediterranean<br />

and the tectonic setting of the Eratosthenes Seamount. Geophys.J. Res., Astronom. Soc. 45, 105-<br />

123.<br />

Ben-Gai, Y. and Ben-Avraham, Z., 1995. <strong>Tectonic</strong> processes in the offshore northern <strong>Israel</strong> and the<br />

evolution of the Carmel structure. Marine Petrol. Geol. 12(5), 553-548.<br />

108


Benjamini, C., 1979, Facies relationships of the Avdat Group (Eocene) in northern Negev, <strong>Israel</strong>. Isr. J.<br />

Earth Sci. 28:47-69.<br />

Benjamini, C., 1980, <strong>Stratigraphy</strong> and Foraminifera of the Qezi'ot and Har Aqrav formations (latest<br />

Middle to Late Eocene) of the Western Negev, <strong>Israel</strong>. Isr. J. Earth Sci. 29:227-244.<br />

Benjamini, C., 1984, <strong>Stratigraphy</strong> of the Eocene of the Arava valley (eastern and southern Negev,<br />

southern <strong>Israel</strong>). Isr. J. Earth Sci. 33:167-177.<br />

Benjamini, C., 1993, Paleobathymetry of the Eocene of <strong>Israel</strong>. Isr. Geol. Soc., Annu. Meet., Arad, Abstr.<br />

p. 17.<br />

Benjamini, Ch., Druckman, Y. and Zak, I., 1993. Depositional cycles in the Ramon Group (Triassic) ,<br />

Makhtesh Ramon. Isr. J. Earth Sci. 42, 115-124.<br />

Benjamini, Ch., Hirsch, F.and Eshet, Y., 2005. <strong>The</strong> Triassic in <strong>Israel</strong>. In: Hall, J.K., Krasheninnikov,<br />

V.A.,Hirsch, F., Benjamini, Ch. and Flexer, A., (eds.) Geological Framework of the <strong>Levant</strong>. Vol. II,<br />

<strong>The</strong> <strong>Levant</strong>ine <strong>Basin</strong> and <strong>Israel</strong>. 331-360.<br />

Bentor, Y.K., and Vroman, A., 1954. A structural contour map of <strong>Israel</strong> (1;250,000), with remarks on its<br />

dynamic interpretation. Bull. Res. Council, Isr. 4, 125-235.<br />

Bentor, Y.K., and Vroman, A., 1960. <strong>The</strong> geological map of <strong>Israel</strong> (1:100,000). Sheet 16: Mount Sdom,<br />

explanatory text. Isr. Geol. Surv.<br />

Best, J.A., Barazangi, M., Al-Saad, D., Sawaf, T. and Gebran, A., 1993. Continental margin evolution of<br />

the Northern Arabian Platform in Syria. Am. Assoc. Petrol.Geol. Bull., 77, 173-193.<br />

Beydoun, Z.R., 1988. <strong>The</strong> Middle East: Regional Geology and Petroleum Resources. Scientific Press<br />

Ltd., Beaconsfield, 292 p.<br />

Beyth, M. and Heimann, A., 1999. <strong>The</strong> youngest igneous event in the crystalline basement in the<br />

Arabian-Nubian Shield, Timna igneous complex. Isr. J. Earth Sci. 48, 113-120.<br />

Bohannon, R.G. and Eittreim, S.L., 1991. <strong>Tectonic</strong> development of passive continental margins of the<br />

southern and central Red Sea with a comparison of Wilkes Land, Antarctica. <strong>Tectonic</strong>s, 198, 129-<br />

154.<br />

Braun, M. and. Hirsch, F. 1994. Mid Cretaceous (Albian-Cenomanian) carbonate platforms in <strong>Israel</strong>,<br />

Cuader. Geolog. Iber., 18, 59-81.<br />

Bruner, I. 1991. Investigation of the subsurface in the northern Negev, <strong>Israel</strong>, using seismic reflection<br />

techniques. Ph.D. <strong>The</strong>sis, Tel Aviv Univ. 105 p.<br />

Buchbinder, B., Sneh, A., Dimant, E., 1986. <strong>The</strong> Neogene Bet Nir Formation, a study of alluvial<br />

aggradation along the toe of the Judean monoclines. Isr. J. Earth Sci. 35, 183-191.<br />

Buchbinder, B., Benjamini, C., Mimran, Y. And Gvirtzman, G., 1988. Mass transport in Eocene<br />

pelagic chalk on the northwestern edge of the Arabian platform, Shefela area, <strong>Israel</strong>.<br />

Sedimentology. 35, 257-274.<br />

Buchbinder, B. and le Roux, J.P. 1993. Inner platform cycles in the Ardon Formation: Lower Jurassic,<br />

southern <strong>Israel</strong>. Isr. J. Earth Sci. 42, 1-16.<br />

109


Buchbinder, B., Martinotti, G.M., Siman-Tov, R. and Zilberman, E., 1993. Temporal and spatial<br />

relationships in Miocene reef carbonates in <strong>Israel</strong>. Paleogeog, Paleoclim., Paleoecol., 101, 97-116.<br />

Buchbinder, B. and Zilberman, E., 1997. Sequence stratigraphy of Miocene–Pliocene carbonate–<br />

siliciclastic shelf deposits in the eastern Mediterranean margin (<strong>Israel</strong>): effects of eustasy and<br />

tectonics. Sediment. Geol. 112, 7-32.<br />

Buchbinder, B., Benjamini, C. and Lipson-Benitah, S., 2000. Sequence development of Late<br />

Cenomanian–Turonian Carbonate ramps, platforms and basins of <strong>Israel</strong>. Cretaceous Res. 21, 813-<br />

843.<br />

Buchbinder, B. and Siman-Tov, R., 2000. Oligocene clastics in the Ashdod area: mass flow and turbidite<br />

deposits in a submarine-fan setting. Isr. Geol. Soc. Annu. Meet., Abstr. p. 25.<br />

Buchbinder, B., Calvo, R. and Siman-Tov, R., 2005. <strong>The</strong> Oligocene in <strong>Israel</strong>: A marine realm with<br />

intermittent denudatioin accompanied by mass-flow deposition. Isr. J. Earth Sci. 54(2), 63-85.<br />

Buchbinder, L., 1979. Facies, environment of deposition, and correlation of the Zohar (Brur Calcarenite),<br />

Karmon and Shederot Formations in the Ashdod area. Isr. Geol. Surv. Rep. OD/3/79.<br />

Buchbinder, L.G., Magaritz, M. and Goldberg, M., 1984. Stable isotope study of carstic related<br />

dolomitization: Jurassic rocks from the Coastal Plain, <strong>Israel</strong>. J. Sediment. Petrol. 54, 236-256.<br />

Calvo, R. and Bartov, Y. 2001. Hazeva Group, southern <strong>Israel</strong>: New observations and their implications<br />

for its stratigraphy, paleogeography and tectono-sedimentary regime. Isr. J. Earth Sci., 50: 71-99.<br />

Castro, J. M. and Ruiz-Ortiz, P. A, 1995., Early Cretaceous evolution of the Prebetic zone in northeast<br />

Alicante province: the Sierra de Seguili section. Cretaceous Res. 16, 573-598.<br />

Chaimov, T.A., Barazangi, M., Al-Saad, D., Sawaf, T. and Khaddouri, M. 1993. Seismic fabric and 3-D<br />

structure of the southwestern intracontinental Palmyride fold belt, Syria. Am. Assoc. Petrol.<br />

Geol. Bull. 77, 2032-2047.<br />

Cochran, J.R., 2001. Nucleation of an oceanic spreading center in a continental rift: <strong>The</strong> northern Red<br />

Sea. Rupturing of Continental Lithosphere in the Red Sea/Gulf of Suez Symposium, Sharm-el-<br />

Sheikh, Egypt, Abstr.<br />

Cohen, Z., 1971. <strong>The</strong> geology of the Lower Cretaceous in the Helez field, <strong>Israel</strong>, Ph.D. <strong>The</strong>sis, Hebrew<br />

Univ. Jerusalem, (in Hebrew, English abstr.), 98 p.<br />

Cohen, Z., 1976. Early Cretaceous buried canyon: influence on accumulation of hydrocarbons in the<br />

Helez oil field, <strong>Israel</strong>, Am. Assoc. Petrol. Geol. Bull. 60, 108-114.<br />

Cohen, Z., Flexer, A. and Kaptsan, V., 1988. <strong>The</strong> Pleshet basin, a newly discovered link in the peripheral<br />

chain of basins of the Arabian craton, J. Petrol. Geol. 11(4), 403-414.<br />

Cohen, Z., Flexer, A. and Kaptsan, V., 1990. <strong>The</strong> tectonic mosaic of the southern <strong>Levant</strong>: Implication for<br />

hydrocarbon prospects. J. Petrol.Geol. 13(4), 437-462.<br />

Davis, B.K., 1982. Structural inversion in Europe and <strong>Israel</strong> and its importance in oil exploration. Oil<br />

Exploration (Invest.) Ltd. Rep. OEL/82/04.<br />

Derin, B., 1974, <strong>The</strong> Jurassic of central and northern <strong>Israel</strong>, Ph.D. thesis, Hebrew Univ., Jerusalem, 152 p.<br />

Derin, B. and Gerry, E., 1981. Late Permian-Late Triassic stratigraphy in <strong>Israel</strong> and its significance to oil<br />

exploration in <strong>Israel</strong>. Symposium Abstr., Isr. Geol. Soc., Jerusalem,.9-10<br />

110


Derin, B., Lipson,S. and Gerry, E., 1981. Gaash 2, stratigraphic log 1:200. Isr. Inst. Petrol. Energy, Tel<br />

Aviv, Rep. 14/81.<br />

Derin, B., Lipson, S. and Gerry, E., 1988. Reexamination of the Delta 1-1A well: biostratigraphy,<br />

lithostratigraphy, paleoenvironment; with recommendation for new bio- and lithostratigraphical<br />

subdivision in the Early Cretaceous and Late Jurassic. Isr. Inst. Petrol. Energy, Micropaleontol.<br />

Lab., Rep. 7/88, 22 p.<br />

Derin, B., Lipson, S. and Gerry, E., 1990. Microbiostratigraphy and environments of deposition of the<br />

Yam 2 offshore well, Isr. Inst. Petrol.Energy, Rep. 5/90, 24 p.<br />

Derin, B., and Reiss, Z., 1966. Jurassic microfacies of <strong>Israel</strong>,, Isr. Inst. Petrol, Tel Aviv, Spec. Publ., 43 p.<br />

de Sitter, L.U., 1962. Structural development of the Arabian Shield in Palestine. Geol.Minjb. 11, 116-<br />

124.<br />

Dewey, J.F., Pitman, W.C. III, Ryan, W.B.F. and Bonnin, J. 1973. Plate tectonics and the evolution of<br />

the Alpine system: Geol. Soc. Am. Bull. 84, 3137-3180.<br />

Dolson, J., Boucher, P., Dodd, T. and Ismail, J., 2002. Petroleum potential of an emerging giant gas<br />

province, Nile Delta and Mediterranean Sea off Egypt. Oil and Gas J., 100(20).<br />

Druckman,Y., 1969. <strong>The</strong> petrography and environment of deposition of the Triassic Saharonim<br />

Formation and the Dolomite Member of the Mohilla Formation in Makhtesh Ramon, central Negev<br />

(southern <strong>Israel</strong>). Isr. Geol. Surv. Bull. 49, 47 p.<br />

Druckman, Y., 1974. <strong>The</strong> stratigraphy of the Triassic sequence in southern <strong>Israel</strong>. Isr. Geol. Surv. Bull.<br />

64, 92 p.<br />

Druckman, Y., 1976, <strong>The</strong> Triassic in southern <strong>Israel</strong> and Sinai. A sedimentological model of marginal<br />

epicontinental marine environments. Ph.D. <strong>The</strong>sis, Hebrew Univ. Jerusalem, 188 p. (in Hebrew,<br />

English abstract).<br />

Druckman, Y., 1977. Differential subsidence during the deposition of the Lower Jurassic Ardon<br />

Formation in western Jordan, southern <strong>Israel</strong> and northern Sinai. Isr. J. Earth Sci. 26, 45-54.<br />

Druckman, Y., 1984. Evidence for Early-Middle Triassic faulting and possible rifting from the Helez<br />

Deep borehole and the coastal plain of <strong>Israel</strong>, In: Dixon, J.E. and Robertson, A.H.F. (eds.) <strong>The</strong><br />

Geological Evolution of the Eastern Mediterranean, Geol. Soc. London, Spec. Publ. 17, 203-212.<br />

Druckman, Y., 2001. <strong>The</strong> hydrocarbon potential of <strong>Israel</strong>. Isr. Geol. Soc. Meet. Elat, Abstr., p. 28.<br />

Druckman, Y. and Kashai, E., 1981, <strong>The</strong> Heletz Deep and Devora 2A boreholes and their implication to<br />

oil prospects in pre-Jurassic strata in <strong>Israel</strong>. Isr. Geol. Surv., Rep.OD/1/81, 23 p.<br />

Druckman, Y., Conway, B.H., Eshet, Y., Gill. D., Lipson-Benitah, S., Moshkovitz, S., Perelis-<br />

Grossowicz, L., Rosenfeld, A. and Siman-Tov, R., 1994. <strong>The</strong> stratigraphy of the Yam Yafo 1<br />

borehole, Isr. Geol. Surv., Rep. GSI/28/94, 27 p.<br />

Druckman, Y., Gill, D., Fleischer, L., Gelbermann, E. and Wolff, O., 1995a. Subsurface geology and<br />

structural evolution of the northwestern Negev, southern <strong>Israel</strong>. Isr. J. Earth Sci. 44, 115-136.<br />

Druckman, Y., Buchbinder, B., Martinotti, G. M., Siman Tov, R. and Aharon, P., 1995b. <strong>The</strong> buried Afiq<br />

Canyon (eastern Mediterranean, <strong>Israel</strong>): a case study of a Tertiary submarine canyon exposed in<br />

Late Messinian times. Marine Geol. 123, 167-185.<br />

111


Dvorkin, A., and Kohn, B. P., 1989. <strong>The</strong> Asher Volcanics, northern <strong>Israel</strong>: Petrography, mineralogy and<br />

alteration. Isr. J. Earth Sci. 38, 105-124.<br />

Eliezri, I., 1964. <strong>The</strong> Jurassic - Cretaceous unconformity in the Coastal Plain and the Negev. Isr. J. Earth<br />

Sci., 12.<br />

Emery, D. and Myers, K. J.,1996. Sequence <strong>Stratigraphy</strong>, Blackwell Science, Oxford, 297 p.<br />

Eshet, Y., 1990. Paleozoic-Mesozoic palynology of <strong>Israel</strong>, Palynological aspects of the Permian-Triassic<br />

succession in the subsurface of <strong>Israel</strong>. Isr. Geol. Surv. Bull. 81, 57 p.<br />

Eyal, Y., 1996. Stress field fluctuations along the Dead Sea rift since the Middle Miocene. <strong>Tectonic</strong>s,<br />

15(1), 157-170.<br />

Eyal, Y. and Reches, Z., 1983. <strong>Tectonic</strong> analysis of the Dead Sea rift region since the Late Cretaceous<br />

based on mesostructures. <strong>Tectonic</strong>s, 2, 167-185.<br />

Fiesntien, S., Aizenshtat, Z., Miloslavsky, I., Gerli, N.P., Slager, J. and McQuilken, J., 2002. Genetic<br />

characterization of gas shows in the east Mediterranean offshore of southwestern <strong>Israel</strong>. Organic<br />

Geochem. 33:12, 1401-1413.<br />

Finetti I., 1985. <strong>Structure</strong> and evolution of the central Mediterranean (Pelagian and Ionian Seas) In:<br />

Stanley, D.J.. and Wezel, F.-C. (eds.), Geological Evolution of the Mediterranean <strong>Basin</strong>. Springer<br />

Verlag, 215-230.<br />

Finetti I. and Morelli C., 1973. Geophysical exploration of the Mediterannean Sea. Boll. Geofis. Teor.<br />

Apl., 15, 263-344.<br />

Flexer, A., 1968. <strong>Stratigraphy</strong> and facies development of Mount Scopus Group (Senonian-Paleocene) in<br />

<strong>Israel</strong> and adjacent countries. Isr. J. Earth Sci. 17, 85-114.<br />

Flexer, A. and Honigstein, A., 1984. <strong>The</strong> Senonian Succesion in <strong>Israel</strong>- Lithostratigraphy, Biostratigraphy<br />

and Sea Level Changes. Cretaceous Research, 5, 303-312.<br />

Flexer, A., Rosenfeld, A., Lipson-Benitah, S. and Honigstein, A., 1986. Relative sea level changes<br />

during the Cretaceous in <strong>Israel</strong>. Am. Assoc. Petrol. Geol. Bull. 70, 1685-1699.<br />

Flexer, A., Hirsch, F. and Hall, J.K., 2005. <strong>Tectonic</strong> evolution of <strong>Israel</strong>. In: Hall, J.K., Krasheninnikov,<br />

V.A., Hirsch, F., Benjamini, Ch., Flexer, A. (eds.), Geological Framework of the <strong>Levant</strong>. Vol. II,<br />

<strong>The</strong> <strong>Levant</strong>ine <strong>Basin</strong> and <strong>Israel</strong>, Historical Productions-Hall, Jerusalem, 523-537.<br />

Fleischer, L. and Gafsou, R., 2003. Top Judea Group- digital structural map of <strong>Israel</strong> (1:200.000).<br />

Geophys. Inst. Isr. Rep. 753/312/03.<br />

Fleischer, L. and Varshavsky, A., 2002. A lithostratigraphic data base of oil and gas wells drilled in<br />

<strong>Israel</strong>. Ministry of National Infrastructure, Oil and Gas Unit, Rep. OG/9/02, 280 p.<br />

Frey-Martinez, J., Cartwright, J., and Hall, B., 2005. 3D seismic interpretation of slump complexes:<br />

examples fromthe continentalmargin of <strong>Israel</strong>. <strong>Basin</strong> Res. 17, 83–108.<br />

Frey-Martinez, J., Cartwright, J., and Hall, B. and M., Huuse, 2007. Clastic Intrusion at the Base of Deepwater<br />

Sands: A Trap-forming Mechanism in the Eastern Mediterranean. In: Hurst, A. and<br />

Cartwright, J. (eds), Sand Injectites: Implications for Hydrocarbon Exploration, AAPG Mem. 87.<br />

Folkman, Y., 1969, Diagenrtic dolomitization in the Albian-Cenomanian Yagur dolomite on Mt. Carmel<br />

(northern <strong>Israel</strong>. J. Sediment. Petrol. 39, 380-85.<br />

112


Folkman , Y. and Ben-Gai, Y., 2004. <strong>The</strong> “Jonah” buried seamount: Intrusive structure in the<br />

southeastern <strong>Levant</strong> <strong>Basin</strong>, offshore <strong>Israel</strong>. Isr. Geol. Soc. Annu. Meet. Abstr. p. 29.<br />

Freund, R., 1965. A model of the structural development of <strong>Israel</strong> and adjacent areas since Upper<br />

Cretaceous times. Geol. Mag. 102, 189-205.<br />

Freund, R., Garfunkel, Z., Goldberg, M., Weissbrod, T. and Derin, B., 1970. <strong>The</strong> shear along the Dead<br />

Sea rift, Philos. Trans. Royal Soc. London A 267, 105-127.<br />

Freund, R., Goldberg, M., Weissbrod, T., Druckman, Y. and Derin, B., 1975. <strong>The</strong> Triassic-Jurassic<br />

structure of <strong>Israel</strong> and its relation to the origin of the eastern Mediterranean. Isr. Geol. Surv. Bull.<br />

65, 26 p.<br />

Friedman, G. M., Barzel, A. and Derin, B., 1971. Paleoenvironments in the central coastal belt of<br />

northern and central <strong>Israel</strong> and their significance in the search for petroleum reservoirs. Isr. Geol.<br />

Surv., Rep. OD/1/71, 26 p.<br />

Gardosh, M., 2002. <strong>The</strong> sequence stratigraphy and petroleum systems of the Mesozoic, southeastern<br />

Mediterranean continental margin. Ph.D. thesis, Tel Aviv Univ. 159 p.<br />

Gardosh, M. and Druckman, Y., 2006. Seismic stratigraphy, structure and tectonic evolution of the<br />

<strong>Levant</strong>ine <strong>Basin</strong>, offshore <strong>Israel</strong>. In: Robertson, A. H. F. and Mountrakis, D. (eds.) <strong>Tectonic</strong><br />

Development of the Eastern Mediterranean Region. Geol. Soc., London, Spec. Publ. 260, 201-227.<br />

Gardosh, M., Kashai, E., Salhov, S., Shulman, H. and Tannenbaum, E., 1997. Hydrocarbon exploration in<br />

the southern Dead Sea basin. In: Niemi, T., Ben-Avraham, Z. and Gat, J. (eds.), <strong>The</strong> Dead Sea, the<br />

Lake and Its Setting. Oxford University Press, New York, 53-72.<br />

Garfunkel, Z., 1980. Contribution to the geology of the Precambrian of the Elat area. Isr. J. Earth Sci.,<br />

29(1-2), 25-40.<br />

Garfunkel, Z., 1981. Internal structure of the Dead Sea Leaky Transform (rift) in relation to plate<br />

kinematics. Tectonophysics 80, 81-108.<br />

Garfunkel, Z., 1988. <strong>The</strong> pre-Quaternary geology of <strong>Israel</strong>. In: Yom-Tov Y. and Tchernov, E. (eds.), <strong>The</strong><br />

Zoogeography of <strong>Israel</strong>, Jung Publishers, Dordrecht, 7-34.<br />

Garfunkel, Z,. 1989. <strong>Tectonic</strong> setting of Phanerozoic magmatism in <strong>Israel</strong>. Isr. J. Earth Sci. 38, 51-74<br />

Garfunkel, Z. 1998. Constraints on the origin and history of the eastern Mediterranean basin.<br />

Tectonophysics, 298, 5-35.<br />

Garfunkel, Z. and Derin, B., 1984. Permian-early Mesozoic tectonism and continental margin formation<br />

in <strong>Israel</strong> and its implications to the history of the eastern Mediterranean. In: Dixon, J., E. and<br />

Robertson, A.H.F. (eds.) <strong>The</strong> Geological Evolution of the Eastern Mediterranean. Geol. Soc.<br />

London, Spec. Publ. 17, 18-201<br />

Garfunkel, Z. and Horowitz, A., 1966. <strong>The</strong> upper Tertiary and Quaternary morphology of the Negev,<br />

<strong>Israel</strong>. Isr. J. Earth Sci. 15, 101-117.<br />

Gelbermann, E., 1995. Revised seismic structural and isopach maps, central and southern <strong>Israel</strong> Inst.<br />

Petrol. Res. Geophys, Rep. 201/22, 88-94.<br />

Gelbermann, E. and Kemmis, J.L.; 1987. Triassic to Late Cretaceous tectonic alignments in southern<br />

coastal plain in the northwestern Negev. Isr. Geol. Soc. Annu. Meet.. Abstr.,39-40.<br />

113


Gill, D., Conway, B.H., Eshet, Y., Lipson, S., Perelis-Grossowicz, L., Rosenfeld, A. and R. Siman-Tov,<br />

1995. <strong>The</strong> stratigraphy of the Yam West 1 borehole, Isr. Geol. Surv., Rep. GSI/13/95, 43 p.<br />

Ginzburg, A. and Ben Avraham, Z., 1987. <strong>The</strong> deep structure of the central and southern <strong>Levant</strong><br />

continental margin. Annales <strong>Tectonic</strong>ae 1, 105-115.<br />

Goldberg, M. and Friedman, G.M., 1974. Paleoenvironments and paleogeographic evolution of the<br />

Jurassic system in southern <strong>Israel</strong>. Isr. Geol. Surv. Bull. 61, 44 p.<br />

Grader, P. and Reiss, Z., 1958. On the Lower Cretaceous of the Heletz area. Isr. Geol. Surv. Bull. 16, 30<br />

p.<br />

Gradstein, F. M, Agterberg, F. P., Ogg, J. G, Hardenbol, J., Van Veen, P.,Thierry, J. and Huang, Z., 1995.<br />

A Triassic, Jurassic and Cretaceous time scale. In: Berggren, W.A., Kent, D.V., Aubry, M.-P. and<br />

Hardenbol, J. (eds.), Time Scales and Global Stratigraphic Correlation, SEPM - Soc. Sediment.<br />

Geol. Spec.Publ. 54, 95-126.<br />

Gvirtzman, G., 1970. <strong>The</strong> Saqiye Group (late Eocene to early Pleistocene) in the coastal plain and<br />

Hashephela regions, <strong>Israel</strong>. Isr. Geol. Surv. Rep. OD/5/67.<br />

Gvirtzman, G., 1981. <strong>The</strong> findings of the Ga'ash-2 well and their implications on oil exploration in NW<br />

<strong>Israel</strong>. Oil Exploration in <strong>Israel</strong>-A Symposium, <strong>The</strong> <strong>Israel</strong> Geological Society, Jerusalem, 29-31.<br />

Gvirtzman, G. and Buchbinder, B., 1978. <strong>The</strong> Tertiary history of the coastal plain and continental shelf of<br />

<strong>Israel</strong> and its bearing on the history of the Eastern Mediterranean. Initial Report Deep Sea Drilling<br />

Project, 42B, 1195-1222.<br />

Gvirtzman, G. and Klang, A., 1972, A structural and depositional hinge-line along the coastal plain of<br />

<strong>Israel</strong> evidenced by magneto-tellurics. Isr. Geol. Surv. Bull. 55, 18 p.<br />

Gvirtzman, G., Klang, A. and Rotstein, Y., 1990. Early Jurassic shield volcano below Mount Carmel:<br />

New interpretation of the magmatic and gravity anomalies and implication for Early Jurassic rifting.<br />

Isr. J. Earth Sci. 39, 149-159.<br />

Gvirtzman, G., and Reiss, Z., 1965. Stratigraphic nomenclature in the coastal plain and Hashefela regions.<br />

Isr. Geol. Surv. Rep. OD/1/65.<br />

Gvirtzman, G. and Steinitz, G., 1983. <strong>The</strong> Asher Volcanics – an Early Jurassic event in northern <strong>Israel</strong>.<br />

Isr. Geol. Surv. Current Res., 28-33.<br />

Gvirtzman, G. and Weissbrod, T., 1984. <strong>The</strong> Hercynian geanticline of Helez and the Late Paleozoic<br />

history of the <strong>Levant</strong>. In: Dixon, J.E. and Robertson, A.H.F. (eds.), <strong>The</strong> Geological Evolution of the<br />

Eastern Mediterranean, Geol. Soc. London, Spec. Publ. 17, 177-186.<br />

Gvirtzman, Z., Garfunkel, Z. and Gvirtzman, G., 1998. Birth and decay of an intra-continental swell:<br />

Early Creaceous tectonics of southern <strong>Israel</strong>. <strong>Tectonic</strong>s 17, 441-457.<br />

Haq, B.U., Hardenbol, J. and Vail, P.R. 1987. <strong>The</strong> chronology of the fluctuating sea level since the<br />

Triassic. Science, 235, 1156–1167.<br />

Haq, B.U., Hardenbol, J. and Vail, P.R., 1988, Mesozoic and Cenozoic chronostratigraphy and cycles of<br />

sea-level changes. In: Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross,<br />

C.A. and Van Wagoner, J.C. (eds.), Sea-level Changes: an Integrated Approach, SEPM - Soc.<br />

Sediment. Geol. Spec. Publ. 42, 71-108.<br />

114


Hardenbol, J., 1998. Cenozoic sequence chronostratigraphy (Chart 2). In: de Graciansky, P-C.,<br />

Hardenbol, J., Jacquin T. and Vail, P. (eds.), Mesozoic and Cenozoic Sequence <strong>Stratigraphy</strong> of<br />

European <strong>Basin</strong>s. SEPM - Soc. Sediment. Geol. Spec. Publ. 60.<br />

Hirsch, F., 2005. <strong>The</strong> Jurassic of <strong>Israel</strong>. In: Hall, J.K., Krasheninnikov, V.A.,Hirsch, F., Benjamini, Ch.,<br />

and Flexer, A. (eds), Geological Framework of the <strong>Levant</strong>. Vol. II <strong>The</strong> <strong>Levant</strong>ine <strong>Basin</strong> and <strong>Israel</strong>,<br />

361- 391.<br />

Hirsch, F., Flexer, A., Rosenfeld, A. and Yellin-Dror, A., 1995. Palinspastic and crustal setting of the<br />

eastern Mediterranean, J. Petrol. Geol. 18(2), 149-170.<br />

Horowitz, A., 1979. <strong>The</strong> Quaternary of <strong>Israel</strong>. Academic Press, New York.<br />

Hsu, K.J., Cita, M.B., and Ryan, W.B.F., 1973a. <strong>The</strong> origin of the Mediterranean evaporites. Initial<br />

Reports of the Deep Sea Drilling Project, 42, 1203-1232.<br />

Hsu, K.J., Ryan, W.B.F. and Cita, M.B.,1973b. Late Miocene desiccation of the Mediterranean. Nature<br />

242, 240-244.<br />

Hsu, K.J., Montadert, L., Bernoulli, D., Cita, M.B., Erickson, A., Garrison, R.G., Kid, R.B., Melieres, F.,<br />

Muller, C. and Wright, R., 1978. History of the Mediterranean salinity crisis. Initial Reports of the<br />

Deep Sea Drilling Project, 42A, 1053-1078.<br />

Kafri, U., 1986. <strong>The</strong> Late Albian to Early Cenomanian sedimentary break in northern <strong>Israel</strong>. Isr. J. Earth<br />

Sci. 35:211-219.<br />

Kashai, E. and Crocker, P., F., 1987. Structural geometry and evolution of the Dead Sea-Jordan Rift<br />

system as deduced from new subsurface data. Tectonophysics 141, 36-60.<br />

Kempler, D., 1998. Eratosthenes Seamount: the possible spearhead of incipient continental collision in<br />

the eastern Mediterranean. In: Robertson, A.H.F., Eneis, K.C., Richter, C. and A. Camerlenghi<br />

(eds.), Proceed. Ocean Drilling Prog., Scientific Results, 160, 709-721.<br />

Klitgord, K. D. and Hutchinson, D. R., 1988. U.S. Atlantic continental margin; structural and tectonic<br />

framework. In: Sheridan, R. E. and Grow, J. A. (eds.), <strong>The</strong> Atlantic Continental Margin, U.S. Geol.<br />

Soc. Am., <strong>The</strong> Geology of North America, I-2, 19-55.<br />

Knipper, A. L. and Sharaskin, A.,Y., 1994. <strong>Tectonic</strong> evolution of the western part of the Peri-Arabian<br />

Ophiolite Arc. In: Krasheninnikov, V., A., and Hall, J., K. (eds.), Geological <strong>Structure</strong> of the Northeastern<br />

Mediterranean (Cruise 5 of the research vessel Akademik Nikolaj Strakhov), 87-98,<br />

Historical Productions-Hall, Ltd., Jerusalem, 295-306.<br />

Korngrin, D. (2004). <strong>The</strong> Triassic in northern <strong>Israel</strong>. Ph.D. <strong>The</strong>sis, Ben Gurion Univ.<br />

Korngrin, D. and Benjamini, Ch., 2001. Upper Triassic reef facies in the Asher-Atlit 1 Borehole, northern<br />

<strong>Israel</strong>: Microfacies, cement stratigraphy and paleogeographic implications. Facies, 45,.1-24.<br />

Krenkel, E., 1924. Der Syrische Bogen. Zentralblatt Mineralogie 9, 274-281 and 10, 301-313.<br />

Lewy, Z. and Weissbrod , T., 1993. <strong>The</strong> stratigraphy of the Cretaceous in the Makhtesh Gadol, Isr. Geol.<br />

Soc., Annu. Meet,, Arad, Field Guide, 1-13 (in Hebrew).<br />

Lipson-Benitah, S., Almogi Labin, A. and Sass, E., 1997. Cenomanian biostratigraphy and<br />

paleoenvironment in the northwest Carmel Region, northern <strong>Israel</strong>. Cretaceous Res. 18:469-491.<br />

115


Lovelock, P.E.R. 1984. A review of the tectonics of the northern Middle East region. Geol. Mag. 121,<br />

577-587.<br />

Makris, J., Ben-Avraham, Z., Behle, A., Ginzburg, A., Giese, P., Steinmetz, L., Whitmarsh, R.B. and<br />

Eleftheriou, S., 1983. Seismic refraction profiles between Cyprus and <strong>Israel</strong> and their interpretation.<br />

Geophys. J. Royal Astronom. Soc. 75, 575-591.<br />

Makris, J. and Wang, J., 1994. Bouguer gravity anomalies of the Eastern Mediterranean Sea. In:<br />

Krasheninnikov, V., A., and Hall, J., K. (eds.), Geological <strong>Structure</strong> of the North-eastern<br />

Mediterranean (Cruise 5 of the research vessel Akademik Nikolaj Strakhov), Historical Productions-<br />

Hall, Ltd., Jerusalem, 87-98.<br />

Martinez, F. and Cochran, J.R., 1988. <strong>Structure</strong> and tectonics of the Red Sea: catching a continental<br />

margin between rifiting and drifting. Tectonophysics 150, 1-32.<br />

Martinotti, G.M., 1973. Miogypsina in the Gaza-1 well, coastal plain, <strong>Israel</strong>. Isr. Geol. Surv. Bull. 57, 1-6.<br />

Martinotti, G.M. 1981. An Oligocene unconformity and its interregional interest. Isr. Geol. Surv. Current<br />

Research 1981, 30-35.<br />

Martinotti, G. M., 1990, <strong>The</strong> stratigraphic significance of Globigerinoides ruber and Globigerinoides<br />

obliquus obliquus in the Mediterranean Middle Miocene. Micropaleontol., 36: 96-101.<br />

Maync,W., 1966. Microbiostratigraphy of the Jurassic of <strong>Israel</strong>. Isr. Geol. Surv. Bull. 40, 56 p.<br />

McBride, J.H., Barazangi, M., Best, J., Al-Saad, D., Sawaf, T., Al-Otri, M. and Gebran, A., 1990. Seismic<br />

reflection structure of intracratonic Palmyrid fold-thrust belt and surrounding Arabian Platform,<br />

Syria. Am. Assoc. Petrol. Geol. Bull. 74, 238-259.<br />

McQuilken, J., 2001. Petroleum geochemistry and basin modeling offshore <strong>Levant</strong> <strong>Basin</strong>, <strong>Israel</strong>. Isr.<br />

Geol. Soc. Meet., Elat, Abstr. p. 79.<br />

Michelson, Ch. and Lipson-Benitah, S., 1986. <strong>The</strong> litho and biostratigraphy of the southern Golan<br />

Heights. Isr. J. Earth Sci. 35: 221-240.<br />

Miller, K.J., Kominz, M.A., Browning, J.V., Wright, J.D., Mountain, G.S., Katz, M.E., Sugarman, P.J.,<br />

Cramer, B.S., Christie-Blick, N. and Pekar, S.F., 2005. <strong>The</strong> Phanerozoic record of global sea-level<br />

change. Science, 310:1293-1298.<br />

Mimran, Y., 1984. Unconformities on the eastern flank of the Fari'a Anticline, and their implications on<br />

the structural evolution of Samaria (central <strong>Israel</strong>). Isr. J. Earth Sci. 33(1-2), 1-11.<br />

Moshkovitz, S., 1988 in Derin, B., Lipson, S. and Gerry, E., 1988, Reexamination of the Delta 1-1A well:<br />

biostratigraphy, lithostratigraphy, paleoenvironment; with recommendation for new bio- and<br />

lithostratigraphical subdivision in the Early Cretaceous and Late Jurassic. Isr. Inst. Pet. Energy,<br />

Micropaleontol. Lab., Rep. 7/88, 22 p.<br />

Neev, D., Almagor, G., Arad, A., Ginzburg, A. and Hall, J. K., 1976. <strong>The</strong> geology of the southeastern<br />

Mediterranean. Isr. Geol. Surv. Bull. 68, 52 p.<br />

Netzeband, G. L., Hubscher, C. and Gajevski, D., 2006, <strong>The</strong> structural evolution of the Messinian<br />

evaporates in the <strong>Levant</strong>ine <strong>Basin</strong>. Marine Geol. 230, 249- 273.<br />

Oats, M., 2001. Origin and genesis of Lower Oliocene sand mounds, Afiq Canyon, <strong>Israel</strong>. Isr. Geol.<br />

Soc.Annu. Meet., Elat, Abstr. p. 89.<br />

116


Perelis-Grossowicz, L., Bassoulet, J.P., Hirsch, F. and Peri, M., 2000. Jurassic large foraminifera from<br />

<strong>Israel</strong>. Isr. Geol. Surv. Current Res, 12, 132-144.<br />

Picard, L., 1943. <strong>Structure</strong> and evolution of Palestine; with comparative notes on neighbouring countries.<br />

Hebrew Univ. Jerusalem, Geol. Dept., Bull. 4(2-4), 187 p.<br />

Picard, L., 1951. Geomorphology of <strong>Israel</strong>, Part I- <strong>The</strong> Negev. Isr. Geol. Surv. Bull. 1, 28 p.<br />

Picard, L. 1959. Geology and oil exploration of <strong>Israel</strong>. Bull.Res. Council Isr., G8, 1-30.<br />

Robertson, A.H.F., 1998. Mesozoic-Tertiary tectonic evolution of the easternmost Mediterranean area:<br />

integration of marine and land evidence. In: Robertson, A.H.F., Eneis, K.C., Richter, C. and<br />

Camerlenghi, A. (eds.), Proceed.Ocean Drilling Prog., Scientific Results, 160, 723-782.<br />

Robertson, A.H.F. and J.E. Dixon, 1984. Introduction: aspects of the geological evolution of the eastern<br />

Mediterranean, In: Dixon, J.E. and Robertson, A.H.F. (eds.), <strong>The</strong> geological evolution of the eastern<br />

Mediterranean, Geol. Soc. London, Spec. Publ. 17, 1-74.<br />

Rouchy, J.M. and Saint Martin, J.P., 1992. Late Miocene events in the Mediterranean as recorded by<br />

carbonate-evaporite relations. Geology, 20, 629-632.<br />

Rybakov, M., Fleischer, L. and Goldshmidt, V., 1995. A new look at the Hebron magnetic anomaly. Isr.<br />

J. Earth Sci., 44, 41-49.<br />

Rybakov, M., Goldshmit, V., & Rotstein, Y. 1997. New regional gravity and magnetic maps of the<br />

<strong>Levant</strong>. Geophys. Res. Lett. 24 , 33-36.<br />

Rybakov, M., Goldshmit, V., Rotstein, Y., Fleischer, L., and Goldberg, I., 1999. Petrophysical constraints<br />

on gravity/magnetic interpretation in <strong>Israel</strong>. Lead Edge, 18, 269-272.<br />

Ryan, W.B.F., 1978. Messinian badlands on the southeastern margin of the Mediterranean Sea. Marine<br />

Geol. 27, 349-363.<br />

Said, R., 1990. <strong>The</strong> Geology of Egypt. Balkema, Netherland, 734 p.<br />

Sandler, A., 1996. A Turonian subaerial event in <strong>Israel</strong>: karst, sandstone and pedogenesis. Isr. Geol. Surv.<br />

Bull. 85, 52 p.<br />

Sass, E. and Bein, A., 1982. <strong>The</strong> Cretaceous carbonate platform in <strong>Israel</strong>, Cretaceous Res. 3, 135-144.<br />

Schattner, U., Ben-Avraham, Z., Reshef, M., Bar-Am, G. and Lazar, M., 2006. Oligocene–Miocene<br />

formation of the Haifa basin: Qishon–Sirhan rifting coeval with the Red Sea–Suez rift System.<br />

Tectonophysics, 419, 1-12.<br />

Schulze, F., Lewy, Z., Kuss, J. and Gharaibeh, A., 2003. Cenomanian–Turonian carbonate platform<br />

deposits in west central Jordan. Int. J. Earth Sci., 92, 641-660<br />

Segev, A., 2000. Synchronous magmatic cycles during the fragmentation of Gondwana: radiometric ages<br />

from the <strong>Levant</strong> and other provinces. Tectonophysics 325, 257-277.<br />

Segev, A., Sass E., Ron, H., Lang, B., Kolodny, Y. and McWilliams, M., 2002. <strong>Stratigraphy</strong>,<br />

geochronology, and paleomagnetic constraints on Late Cretaceous volcanism in northern <strong>Israel</strong>. Isr.<br />

J. Earth Sci., 51(3-4), 297-309.<br />

Shachnai, E., 1969. Lower Cretaceous stratigraphy of the Bet El Mountains. Isr. J. Earth Sci. 18(3-4), p.<br />

169<br />

117


Shaliv, G., 1991. Stages in the tectonic and volcanic history of the Neogene basin in the Lower Galilee<br />

and the valleys. Isr. Geol. Surv. Rep. GSI/11/91,94 p. (In Hebrew).<br />

Shilo, A. and Wolff, O., 1989. <strong>The</strong> stratigraphy of the Cretaceous in <strong>Israel</strong> emphasizing the subsurface.<br />

PIC Rep. 209/223/89.<br />

Siman-Tov, R., 1991. <strong>The</strong> Turborotalia cerroazulensis Zone at Tel Nagila, southeastern Hashefela. Isr.<br />

Geol. Surv. Rep. GSI/38/91, 15 p.<br />

Stampfli, G.M., Mosar, J., Favre, P., Pillevuit, A. and Vannay J.-C., 2001. Permo-Mesozoic evolution of<br />

the western Tethyan realm: the Neotethys East-Mediterranean <strong>Basin</strong> connection.<br />

In: Ziegler, P.A., Cavazza, W., Robertson, A.H.F. and Crasquin-Soleau, S. (eds.), Peri-Tethys Memoir 6:<br />

Peri-tethyan Rift/Wrench <strong>Basin</strong>s and Passive Margins. Bull. Mus. natn. Hist. Nat. 186, 51-108.<br />

Steinitz, G., Gvirtzman, G. and Lang, B., 1983. Evaluation of K-Ar ages of the Asher volcanics, <strong>Israel</strong>.<br />

Isr. Geol. Surv. Current Res., 34-38.<br />

Tannenbaum, E. and Aizenshtat, Z., 1985. Formation of immature asphalt from organic-rich carbonate<br />

rocks-I, geochemical correlation. Org. Geochem. 8(2), 181-192.<br />

ten Brink, U., 1987. Quantitative basin analysis with applications to<br />

the <strong>Israel</strong>i continental margin. Report prepared for Oil Exploration {Investments}, 24 p.<br />

Tibor, G., Ben-Avraham, Z., Steckler, M. and Fligelman, H., 1992. Late Tertiary subsidence history of<br />

the southern <strong>Levant</strong> margin, eastern Mediterranean Sea, and its implications to the understanding of<br />

the Messinian event, J. Geophys. Res. 97(B12), 17593-17614.<br />

Walley, C., D., 1998. Some outstanding issues in the geology of Lebanon and their importance in the<br />

tectonic evolution of the <strong>Levant</strong>ine region. Tectonophysics 298, 37-62.<br />

Walley, C. H., 2001. <strong>The</strong> Lebanon passive margin and the evolution of the <strong>Levant</strong>ine Neo-Tethys. In:<br />

Ziegler, P.A., Cavazza, W., Robertson, A. H. F. and Crasquin-Soleau (eds.), Peri-Tethys Memoir 6:<br />

Peri-Tethyan Rift/Wrench <strong>Basin</strong>s and Passive Margins. Mem. Mus. natn. Hist. Nat., 186, 407-439<br />

Weissbrod, T., 1969a. <strong>The</strong> Paleozoic of <strong>Israel</strong> and adjacent countries: Part I – <strong>The</strong> subsurface Paleozoic<br />

stratigraphy of southern <strong>Israel</strong>. Isr. Geol. Surv. Bull. 47, 32 p.<br />

Weissbrod, T., 1969b. <strong>The</strong> Paleozoic of <strong>Israel</strong> and adjacent countries: Part II – <strong>The</strong> Paleozoic outcrops in<br />

southwestern Sinai and their correlation with those of southern <strong>Israel</strong>. Isr. Geol. Surv. Bull. 48, 23 p.<br />

Weissbrod, T., 2005. <strong>The</strong> Paleozoic in <strong>Israel</strong> and environs. In: Hall, J.K., Krasheninnikov, V.A., Hirsch,<br />

F., Benjamini, Ch. and Flexer, A. (eds.) Geological Framework of the <strong>Levant</strong>. Vol. II <strong>The</strong> <strong>Levant</strong>ine<br />

<strong>Basin</strong> and <strong>Israel</strong>, Historical Productions-Hall, 283-316.<br />

Weissbrod, T., Gvirtzman, G., Weinberger, G. and Sandler, A., 1990. Interfingering of marine and nonmarine<br />

Cretaceous in southern <strong>Israel</strong>. Cretaceous Field Conf., Jerusalem, Guide book, 65-97.<br />

Wissler, L., Funk, H. and Weissert, H., 2003, Response of Early Cretaceous platforms to changes in<br />

atmospheric carbon dioxide levels. Paleogeogr. Paleoclim. Paleoecolog. 200, 187-2005.<br />

Zilberman, E. 1992. Remnants of Miocene landscape in the central and northern Negev and their<br />

paleogeographical implications. Isr. Geol. Surv. Bull. 83, 54 p.<br />

Zilberman, E., 1993. <strong>The</strong> Mahmal graben and Nahal Hawwa: an introduction to the paleogeography of<br />

central Negev in the Miocene. Isr. J. Earth Sci. 42, 197-209.<br />

118

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!