Open this page in a new tab

DST Geomechanics

Structural Styles Along the Dead Sea Fault - Ben-Avraham et al (2012)

Introduction

Introduction

Source Image Description
Ben-Avraham et al (2012) Fig. 17.1 - DST Tectonics
The Dead Sea fault, which spans around 1000 km (Fig. 17.1), is an active left lateral, strike-slip (transform) plate boundary. This feature forms the northern part of the Syrian-African rift system, which extends for >6000 km from Turkey to Mozambique. The fault formed as a consequence of late-Cenozoic breakup of the Afro-Arabian continent and the subsequent drift of Arabia away from Africa. Older structures resulting from mild compression during the Senonian to Mio¬cene are evident throughout the region. These east-northeast to northeast trending and east to west trending folds and faults compose the Syrian-Arc system. The Dead Sea fault obliquely crosses and offsets laterally these structures and is hence, younger. Recent continental breakup was accompanied by a marked change in the behaviour of the region - the Syrian-Arc deformation virtually ceased, the region was uplifted, especially near the newly formed plate boundaries and widespread igneous activity took place. In particular, this resulted in the uplift of the margins of the Dead Sea fault system as well as subsidence along much of its length forming the pronounced physiography seen today (a profound valley, bordered by high mountainous scarps and occupied by deep basins). The fault valley is for the most part, bordered by high escarpments commonly known as the Western and Eastern Boundary faults.

In terms of plate tectonics, the Dead Sea fault is considered to be a left-lateral transform plate boundary with an extensional component, separating the Arabian Plate and the Sinai sub-plate. Activity along the fault is thought to have started in the middle Miocene, when a transition of motion took place from opening in the Gulf of Suez to transcurrent displacement along the Dead Sea fault. The total amount of left lateral slip is estimated at about 105 km. History of motion is not well constrained because young markers cannot be reliably matched across the Fault.

Calculation of regional plate kinematics provides an instantaneous rate of motion along the Fault boundary of about 5-6 mm/yr for the South Sinai triple Junction. This average rate is similar to the average slip computed from the ratio between the entire 105 km slip to ca. 17-20 Ma of tectonic activity (presuming that the overall slip rate has remained approximately uniform during the history of the Fault). More recent GPS measurements provide a minimum slip rate of 3.7 ± 1.1 mm/yr for the three years between 1996 and 1999.

The fault is divided into two sections — the southern section from the Gulf of Elat to south of Lebanon, and a section continuing northward from Lebanon to the Taurus mountain range in Turkey (Garfunkel, 1981). The fault valley is character¬ized by large-scale topographic and structural asymmetries (Wdowinski and Zilberman, 1996). Along the southern section, particularly in the northern Arava valley, the fault's eastern shoulder is flexed upwards towards the axis, resembling an uplifted shoulder, while the western side is flexed down towards the axis, resembling a wide arch. In addition, the eastern side is for the most part topographically higher than the western side. The two sections of the Dead Sea fault differ by a reversal in the large-scale asymmetry; in the southern section (Gulf of Elat to south of Lebanon), the eastern side is usually higher than the western side, while in the northern section (north of Lebanon), the western side is mostly higher than the eastern one.

Along strike variations in topography and structure subdivide the southern Dead Sea fault into five segments. From south to north, these are the Gulf of Elat, Arava, Dead Sea and the Sea of Galilee and the Hula Valley (Fig. 17.1) (Garfunkel, 1981). Deep pull-apart basins, formed between left-stepping fault segments of the Dead Sea fault, can be found in topographically lower areas. Lengths of the individual pull apart basins range from 15 to 50 km and are 5-20 km wide, while their depths range between 5 and 15 km. Marginal normal faults usually border these depressions, while transverse or oblique normal faults define their southern and northern limits. Structural saddles typically occur between the basins and there, transpression often takes place. The crystalline basement underlying the larger pull-aparts is thinner than normal, which accounts for basin subsidence. The basins tend to become shorter and younger in age north¬wards along the fault. Asymmetry occurs in most of the basins, interpreted by Ben-Avraham and Zobak (1992) as indicating that they are bounded by transform segments on one side and sub-parallel normal faults on the other. In these cases, the simple pull-apart model cannot be applied in a straightforward way. Compressional features can be found along many sub-basins, suggesting that the tectonic regime is more complex than specified in the simple models. Both asymmetry and compressional features will be examined in the basins along the southern Dead Sea fault, from the south to the north.

The extensional regime combined with the dominant lateral motion along the Dead Sea fault resulted in the formation of a series of pull-apart basins (or rhomb-shaped grabens). Basins are separated by structural saddles, which formed as a result of local stress fields. In many places, transverse faults cross the saddle, thus separating the basins tectonically as well as structurally.

Descriptions of Individual Sections
Gulf of Elat (aka Gulf of Aqaba)

Source Image Description
Ben-Avraham et al (2012) Fig. 17.2 - Gulf of Elat
The Gulf of Elat is the largest depression along the Dead Sea fault (~180 km long and up to 20 km wide) and occupies its southernmost part. The interior of the gulf comprises three elongated coalescing en-echelon pull apart basins, which strike N20°-25°E. Undulations in the floors of the basins produce several distinct deeps: Tiran and Dakar in the southern depression, Amona and Aragonese in the central depression and the Elat Deep in the northern depression. The Hume Deep, south of the Straights of Tiran, links the Gulf of Elat with the axial depression of the Red Sea. The northern part of the gulf has a relatively simple bathym-etry and is dominated by a single flat-bottomed basin with one deep, the Elat Deep, which is the largest and shallowest in the gulf. Only this part of the gulf is symmetrical in cross section with respect to the bathymetry. Although the bathymetry of the Elat Deep is relatively flat and simple, its underlying structure suggests that it was originally composed of at least two separate deeps. The central part of the gulf is the narrowest and deepest and is occupied by a basin, which crosses the axis of the gulf diagonally from northeast to southwest. The southern part of the gulf is the widest and occupies the eastern side of the gulf. The internal structure of the gulf of Elat is dominated by intra-basinal longitudinal faults and by oblique faults that define these sub-basins. In the central part of the gulf, the older sediments in the deep depressions and on much of the marginal blocks are considerably distorted into arches and open folds, raising the possibility of local transpression (Fig. 17.2A). However, some of these features may also be the expression of underlying salt (or shale) diapirs. It is interesting to note that this is the only area where there seems to be evi¬dence for strike slip motion on both sides of the basin (Ben-Avraham, 1985). The sub-basins are separated by structural saddles, which also internally exhibit folding, flexure and uplift suggesting transpression (Fig. 17.2A).

At the southern extremity of the Gulf of Elat, local transpression is evident by the uplift of the Tiran Island block on which Quaternary reefs are found at elevations exceeding 500 m above sea level. An additional feature observed in Figs. 17.2A and 17.2B is the existence of asymmetry within the sub-basins of the Gulf of Elat. The transition from the Arnona Deep (Fig. 17.2A) to the Elat Deep (Fig. 17.2B) is accompanied by a change in lateral asymmetry. A possible explanation may be that the basins are bounded on one side by a strike-slip master fault and on the other by a predominantly normal boundary fault. Where motion along en-echelon segments of the main fault switches from one segment to another (i.e., from one side of a basin to the other), asymmetry in the basin structure is also expected to switch sides. According to this interpretation, asymmetry within the basins suggests that in the southern and central portion of the gulf, active strike-slip motion is taking place primarily on the western boundary faults, while in the northern portion it is taking place primarily on the eastern boundary fault (Ben-Avraham, 1985). In addition, it can be seen from Fig. 17.2A that the Arnona Deep is bounded on the east by a large compressional feature.

Arava Valley

Source Image Description
Ben-Avraham et al (2012) Fig. 17.3 - Arava
The Arava valley is a morphotectonic depression connecting the Gulf of Elat and the Dead Sea (Fig. 17.3A). The transition from the Gulf of Elat to the Arava Valley is marked by a structural saddle (Fig. 17.1), where mild compression is apparent by upwarping of Neogene and younger sediments. During the Pliocene, this saddle was high and wide enough to allow a major drainage line to flow west¬ward across the Dead Sea fault into the Negev. In the Arava Valley itself, a few small transpressional uplifts developed. The small size of these features may indicate that they are young and formed as a result of recent rearrangements of the active fault strand.

This segment of the Dead Sea fault is characterized by a series of elongated, en-echelon tectonic sub-basins. The left stepping arrangement of the basins (Evrona, Ya'alon, Zofar and Shizaf; Frieslander, 2000; Fig. 17.3) coincides with the junction of E-W dextral faults in the central Sinai-Negev area, which predate the left lateral motion on the Dead Sea fault. The depths of these sub-basins, which are filled with clastic sediments, vary from several hundred meters to a few kilometres. The basins are bounded by sub-parallel, longitudinal boundary faults. In general, the boundary faults are high-angle strike-slip faults with normal displacement. In places, due to the left lateral strike slip motion, reverse faults, collapse antithetic features and compressional 'flower structures' were formed (Fig. 17.3). The pattern of these basins subdivides the Arava valley into two low structural segments, separated by an elevated saddle (Frieslander, 2000).

North of this saddle, the Zofar and Shezaf basins are separated by two listric faults with large displacements. The nature of features of the fault on the east side of this segment of the Arava indicate that it is the main strike slip fault (Garfunkel et al., 1981). The asymmetric fill in these sub-basins may well be related to the fact that strike-slip faulting took place mainly along the eastern boundary fault. In the shallowest (Zofar) basin, the seismic data indicate a regional dip toward the north (Frieslander, 2000).

As in the Gulf of Elat, reversal in asymmetry is also observed in the Zofar sub-basin. Figure 17.3B shows a classic example of a strike-slip fault, where horizons on both sides of the fault, dip towards it. Figure 17.3C, located around 15 km to the north, shows a distinctly different picture. Instead of reflectors of the Zofar basin dipping westward towards the fault, they dip to the east and away from the fault. This reversal in dip direction is the only known example along the Arava valley where reversal in asymmetry occurs within a sub-basin (instead of between sub-basins), and may indicate that the Zofar basin can be structurally sub-divided.

Dead Sea Basin

Introduction

Source Image Description
Ben-Avraham et al (2012) Fig. 17.4 - Dead Sea Basin
The Dead Sea basin is ~110 km long and 15-17 km wide extending from the structural saddle of the Central Arava to near Jericho, with around 10 km of Neogene to recent sediment fill (Ben-Avraham, 1997; Garfunkel and Ben-Avraham, 1996; Ginzburg and Ben-Avraham, 1997). The Dead Sea basin is divided into two sub-basins, which are evident in the deep structure and are separated by the Lisan Peninsula — a large buried salt diapir (Fig. 17.4A). The two basins are thought to be divided by a large oblique normal fault. The two main strands of the Dead Sea fault in this area are the Jericho fault, which borders the northern sub-basin on the west and the Arava fault, which borders the southern sub-basin on the east. The two faults overlap in the central part of the Dead Sea basin where it is bordered by two en-echelon strike-slip faults. Today, the northern basin is occupied by a lake (known simply as the Dead Sea) while the southern basin, once covered by a very shallow lake (part of the Dead Sea), is now used by the Dead Sea Works (Israel) and by the Arab Potash Company (Jordan) as evaporation ponds.

Southern Dead Sea Basin

Source Image Description
Ben-Avraham et al (2012) Fig. 17.4 - Dead Sea Basin
Listric faults are a prominent feature along the Dead Sea fault valley. One of the best examples is the Amazyahu fault in the southern Dead Sea basin (Fig. 17.4B). This fault detaches on top of the Sedom salt formation. Mainly westward, salt flowage has formed the Mt. Sedom diapir, and perhaps also northward flowage formed the Lisan diapir. Increase in thickness of most of the fill layers on the downthrown side of the fault indicates continuous salt-flow and fault activity from the Early Pleistocene to the Holocene (ten Brink and Ben-Avraham, 1989).

Structurally, the southern Dead Sea basin is divided into three step blocks — the rim blocks, intermediate blocks and deep-sunken block (Kashai and Croker, 1987). Recently, compilation of seismic profiles from Jordan and Israel (Fig. 17.4C) has provided a more complete picture across the Dead Sea basin (Al-Zoubi et al., 2002). Tectonically, these structural steps seem to result from some extension in east-west and northeast-southwest directions combined with the predominant lateral motion (Garfunkel, 1981; Garfunkel and Ben-Avraham, 1996). The northern Dead Sea basin One of the most striking features of the southern Dead Sea basin is the existence of the large Mt. Sedom salt diapir. This structure is exposed on the surface and extends northwards and southwards in the subsurface along the edge of the intermediate block (Fig. 17.4C). The initiation of this salt diapir is related to faulting, increased overburden and subsequent upward flow of salt along the Sedom fault (e.g., ten-Brink and Ben-Avraham, 1989; Zak, 1967).

Northern Dead Sea Basin

Source Image Description
Ben-Avraham et al (2012) Fig. 17.4 - Dead Sea Basin
The sediments infilling the northern basin are about 6 km in thickness compared to 12 km in the southern basin (Ginzburg and Ben-Avraham, 1997). The northern Dead Sea basin can be divided into two sections - the Dead Sea (lake) and the adjacent land of the lower Jordan valley. The lake is comprised of a number of smaller sub-basins, which are separated by structural saddles (Lazar et al., 2006). The most striking feature of the lake area is the presence of salt diapirs, such as the En Geddi diapir (Fig. 17.4D) and other smaller salt structures, which occupy the western margin (Neev and Hall, 1979).

On land, deformation of the valley fill on the surface north of the Dead Sea and the moderate dip of the main fault in the subsurface indicate that current motion on the Jericho fault is transpressional. In the subsurface, a distinct structure termed the Kalia monocline (Fig. 17.4E) has long been taken as evidence for compression along this segment of the fault. However, this point is currently debated.

Little is known of the structure of the region between the northern shore of the Dead Sea and the Sea of Galilee. This area includes a number of uplifted structures in which the valley fill dips steeply (Garfunkel, 1981; Garfunkel et al., 1981). Due to poor seismic coverage and quality, it is impossible to tell with any degree of certainty, if they were formed as a result of compressional forces or salt movement. However, velocity-depth sections modelled using seismic refraction data seem to indicate the presence of salt bodies.

Sea of Galilee

Source Image Description
Ben-Avraham et al (2012) Fig. 17.5A - Sea of Galilee
Ben-Avraham et al (2012) Fig. 17.5 B and C - Sea of Galilee
Situated at 200 m below mean sea level, the Sea of Galilee (Lake Kinneret) is around 5 km wide and ca. 20 km long (Fig. 17.5A). The basin is characterized by a negative gravity anomaly (Ben-Avraham et al., 1996), indicating a ~8 km thick sedimentary fill under the central part of the lake (thinning to ~5 km towards the southern extension of the basin). Around the lake, Plio-Pleistocene outcrops of volcanic rocks are present. The Sea of Galilee was divided by Ben-Avraham et al. (1996) into two structural units. The southern half is a pull-apart that extends southwards from the lake, whereas the northern half is thought to have been formed by rotational opening and transverse normal faulting. A later seismic reflection survey (Hurwitz et al., 2002) has shown that the southern basin, delimited by two longitudinal faults, extends under much of the lake. Seismic sections show that the deepest part of the basin is located well south of the deepest bathymetric depression implying that the later is an actively subsiding young feature. The eastern boundary fault extends to the northern part of the lake, while at least on the surface the western fault does not cross the northern section.

The Zemah anticline, a large compressional feature south of the Sea of Galilee in the Kinarot valley (Fig. 17.5B), is at least 4 km long, 2500 m wide and reaches depths of at least 4 km. Figure 17.5B (Zurieli, 2002) clearly hows that the eastern limb of this structure is thicker than the western one (by almost 50%).The anticline is bounded by two faults; the one to the east reach the surface as opposed to the one in the west, which exhibits reverse fault geometry. Other faults can be found at the base of the feature. The Zemah-1 well was drilled on top of the structure and penetrated 4300 m of graben fill consisting of clas-tics, evaporites, intrusive and igneous rocks, with only relatively thin layers of salt present. An additional compressional structure the Ubeidiya structure) is evi¬dent in Fig. 17.5B around 3.5 km west of the Zemah anticline. This element is much smaller (around 800 m wide) and oriented to the northwest. Both elements are thought to have been formed sometime between 0.8 and 1.4 Myr. Rotstein et al. (1992) have suggested that reverse faulting in the Zemah anticline indicates that it is a compressional feature unrelated to igneous activity or salt diapirsm. Between these two compressional structures, an asymmetrical basin, which is the continuation of the Sea of Galilee syncline, is present. Rotstein et al. (1992) have also suggested that the basin in the southern Sea of Galilee and the Kinarot Valley was formed by two overlapping strike-slip faults. Zurieli (2002) proposed that the compressional features can be explained by assuming that the basin is a hybrid pull-apart, that is bordered on both sides by faults with different rates of lateral motion. The Golan Heights lie to the east of the Sea of Galilee and their internal structure is governed by the adjacent northern Dead Sea fault and the Palmyride folding. The Golan Heights can be divided into two separate sections by the Yehudiyya block, or the Yehudiyya transtensional zone (Shulman et al., 2004). The zone is part of a distinct extensional region, or sub-basin, which is confined by:

  1. a branch of the Jordan Fault (Fig. 17.5C) in the west (the Meshoshim fault)
  2. the southwest-northeast trending Sheikh-Ali fault, in the east and northeast
  3. by the Sea of Galilee in the southwest. The Neogene activation of the Sheikh-Ali fault, which branches to the northeast off the north-south trending Dead Sea fault, was the most significant tectonic phase that shaped the structural framework of the Golan Heights.
Within the Yehudiyya transtensional zone, the Qazrin structure, which indicates compression (Shulman et al., 2004), is a relatively wide (5 km across), deep, asymmetrical uplifted dome (Fig. 17.5C). This feature, located within a larger transtensional zone, exhibits structural elements that resemble features found in the southwest trending Palmyrides (short-wavelength anticlines with steeply dipping forelimbs and more shallowly dipping backlimbs). The North-South orientation of the Qazrin structure, which differs from the general trend of the Palmyrides, can be attributed to lateral movement along the Sheikh-Ali fault during later stages of movement (Shulman et al., 2004). The southwestern limb of the Qazrin structure terminates at the Jordan fault — the main strike-slip in the northern segment of the Dead Sea fault. This indicates that the fault is a young feature, which developed in the area after the formation of the Qazrin structure.

Hula Basin

Source Image Description
Ben-Avraham et al (2012) Fig. 17.6 - Hula Basin
The Hula Valley is the northernmost pull-apart basin along the southern section of the Dead Sea fault. The basin developed between left-stepping segments of the Dead Sea fault: the Jordan and the Rachaya fault in the east and the Yammu-neh fault in the west (Heimann, 1990), which extends northward to the Zagros-Taurus zone of plate convergence (Fig. 17.6).

The eastern structural boundary of the Hula Valley forms a complex fault pattern, reflecting a westward migration of tectonic activity during the Pleistocene (Heimann, 1990). The youngest branch is the Azaz fault, separates the western margins of the basalt plateau of the Golan Heights from the sedimentary fill of the valley (Fig. 17.6).

The Azaz fault is a sinistral strike—slip fault consisting of several segments arranged in a right-stepping en-echelon pattern striking NNE (Heimann, 1990). Small push-up structures, expressed as small hills, occur between neighboring segments (Fig. 17.6). Both the Rachaya fault and the Azaz fault are considered to be currently active.

Northern Segment of the Dead Sea Fault

Source Image Description
Ben-Avraham et al (2012) Fig. 17.7 - Ghab Basin
The northern segment of the Dead Sea fault

The structure and kinematics of the Dead Sea Transform are more complex and obscure to the north (Walley, 1998). While about 25 km of left lateral displacement in Lebanon and western Syria is generally accepted, there is some debate and claims of an offset of 80 km or more near the Syrian-Turkish border. A lateral offset of ca. 100 km across Lebanon has been suggested on the basis of structural correlations across the Dead Sea fault. Resolution of this question must await a comprehensive comparison of the structure and stratigraphy on both sides of this segment of the Dead Sea fault, similar to what was done along its southern segment (e.g., Freund et al., 1970). Cenozoic deformation is distributed across a range of faults and folds from the Levant coast across to the Palmyrides of Syria. The timing and significance of these structures is highly controversial. Within Lebanon, the Dead Sea Transform is represented by some or all of an array of fault strands. Most large-scale reviews of plate boundary continuity consider one of these structures, the Yammouneh Fault, to be the main strand. This fault maps along a NNE-SSW trend, implying a general restraining bend geometry to the transform.

The Ghab basin

The Ghab basin is located on the active, yet poorly understood, northern segment of the Dead Sea fault system. It formed during the Plio-Quaternary as a result of a complex step-over zone along the fault. Subsidence occurred along cross-basin and transform-parallel faults in two asymmetric depocenters. The larger depocenter in the south of the basin is asymmetric towards the east, the margin along which most active transform displacement apparently occurs. This suggests that, at the latitude of the Ghab Basin, most of the lateral movement on the Dead Sea fault is accommodated on the eastern bounding fault of the basin.In contrast, the second, smaller depocenter to the north is somewhat asymmetric toward the western bounding fault (Brew, 2001)(Fig. 17.7).

Discussion
Discussion

Motion along the Dead Sea fault is not pure strike-slip and the direction of the plate boundary changes several times resulting in areas of transtension and transpression. This is evident by the variable morphology and structure, which is characterized by extensional, compressional and asymmetrical features. These features vary in size, from the large-scale, which define the general structure of the valley, to the small-scale, which define the internal structure. The Dead Sea fault valley itself is a large-scale transtensional feature, which formed as a result of oblique strike-slip motion along the fault. The extensional component of motion is largely responsible for the formation of pull-apart basins that occupy its length. Compression is caused by a step to the right of the left lateral strike-slip master fault and results in structural saddles, which separate the main basins along the valley. Smaller compressional uplifts divide several of the major pull-aparts into sub-basins and also occur along the fault valley. However, it is often hard to determine whether these intra-basinal features are salt-related. Asymmetry is evident in the structure and topography of the highlands, which border the valley and in the basins that lie within. The large-scale asymmetry across the Dead Sea fault can perhaps be explained by the combined effect of normal faulting and isostatic uplift on existing pre-rift topography (Wdowinski and Zilberman, 1996).

Numerous examples of basin asymmetry and compression features have been presented from locations worldwide, indicating that these are not local phenom¬ena limited to the Dead Sea fault valley. The Cariaco basin in Venezuela, which crosses the El Pilar fault, shows many similarities to the Dead Sea basin, displaying a clear asymmetry towards the main fault (Ben-Avraham and Zoback, 1992). Additional examples exist along other continental transform faults and rift valleys, such as the Motagua fault system in Guatemala, Lake Baikal, Lake Tanganyika in the East African rift system and the north Anatolian fault in Turkey. Transform-normal extension may help explain the small-scale asymmetries observed within the basins themselves. This seemingly contradicts the basic assumption that pull-aparts are formed as a result of overlapping strike-slip faults. However, it is the extensional component that arises from such fault arrangements that may well be responsible for resolving this 'conflict' ('leaky transform', Garfunkel, 1981). It is interesting to note that very little or no mantle uplift exists under the Dead Sea fault in the Arava valley. This may explain the large negative gravity anomalies associated with the fault that reflect primarily the thick sedimentary sequences including salt layers, which are abundant along the length of the valley.

Source Image Description
Ben-Avraham et al (2012) Fig. 17.1 - DST Tectonics
Ben-Avraham et al (2012) Fig. 17.2 - Gulf of Elat
Ben-Avraham et al (2012) Fig. 17.3 - Arava
Ben-Avraham et al (2012) Fig. 17.4 - Dead Sea Basin
Ben-Avraham et al (2012) Fig. 17.5A - Sea of Galilee
Ben-Avraham et al (2012) Fig. 17.5 B and C - Sea of Galilee
Ben-Avraham et al (2012) Fig. 17.6 - Hula Basin
Ben-Avraham et al (2012) Fig. 17.7 - Ghab Basin

Geology and Evolution of the Southern Dead Sea Fault ... - Ben-Avraham et al (2008)

Introduction

Introduction

  • This study focuses on the southern segment based mainlyon the wealth of geophysical data. Owing to transtention caused by oblique-slip and the overlapping of en-echelon fault strands, a series of pull-apart basins were formed along the fault’s length. These basins are long and deep-reaching in places more than 10 km deep.They are characterized by extensional, compressional, and asymmetrical structures varying in size from large-scale (defining the general structure of the Dead Sea fault valley) to small-scale (defining the internal structure). This study examines the internal structure of thesebasins from south to north and summarizes the state of knowledge to date.

  • Today there is a marked difference between the geology of areas facing each other across the Dead Sea transform — a result of lateral motion that juxtaposed areas that were originally far from each other. Matching of numerous markers across the transform indicates a left-lateral offset of approximately 105 km (Quennell 1959, Bartov 1974).

  • ... these studies indicate an average slip rate of 5–7 mm year−1 in the past 5 Ma.

  • Igneous activity — overwhelmingly basaltic volcanism — occurred on a regional scale, mostly east of the transform (Giannerini et al. 1988, Garfunkel 1989, Steinitz and Bartov 1991, Mouti et al. 1992, Ilani et al. 2001). The volcanics occur in a wide belt that broadly follows the transform, but a series of relatively small occurrences exist within the transform. In places the transform appears to have been a preferred site of magma upwelling along the northern two-thirds of its length. Volcanism began locally 18–20 Ma ago near Tiberias and in the Damascus region; however, most volcanics were extruded since 5-12 Ma.

    The structure along the Dead Sea Fault varies considerably, and this is expressed in its physiography. Structurally, it can be divided into two segments: south and northof lat. 33° 10' N. The southern segment is marked by an almost continuous valley, most of which is underlain by a series of deep basins that are separated by less pronounced saddles (Garfunkel 1981, Garfunkel and Ben-Avraham 2001). These structures are controlled by longitudinal en-echelon faults on which lateral motion takes place. These are flanked by normal faults that produced the morphologic boundaries of the valley. Transverse faults are also present, but less conspicuous. In map view, the segment has an arcuate shape, and the strike of the major longitudinal faults varies gradually from approximately 25° NE in the south to nearly N-S in the north, which is compatible with the kinematics of Arabia-Africa plate separation. The margins of this segment were hardly deformed during the development of the Dead Sea Fault, except in the area to the west of lat. 32° N (Samaria and Galilee).

    The northern segment is quite different. It comprises well-expressed longitudinal faults with variable strikes. Valleys are developed only along parts of this segment. In Lebanon, the main fault extends along a high-standing area—the topographically highest area along the transform. The flanks of this segment were considerably deformed during the development of the transform. Given the changes in strike of theDead Sea Fault, continuing lateral motion had to deform its flanks. In particular, the fault bends some 30° clockwise in Lebanon. This would be expected to produce substantial transpression, which is considered to be the cause of the great uplifting of its flanks (Quennell 1959), and their intense faulting, which involved great rotations of individual fault blocks about vertical axes (Ron et al. 1984, Ron 1987). East of the transform the compression led to folding.

Descriptions of Individual Sections
Gulf of Elat (aka Gulf of Aqaba)

Introduction

Source Image Description
Ben-Avraham et al (2008) Fig. 2 - Gulf of Elat
The southernmost basins along the Dead Sea Fault are situated in the Gulf of Elat (or Aqaba). The Gulf is the largest depression along the fault, some 180 km long and 20 km wide and is comprised of three elongated, coalescing, en-echelon, actively subsiding pull-apart basins, which strike N20°-25°E (Figure 2a) (Ben Avraham et al. 1979, Ben Avraham 1985, Ben Avraham & Garfunkel 1986, Ben-Avraham & Tibor 1993). It is flanked by metamorphic-plutonic rocks that reach heights of more than 1000 m above sea level. The basins and structural undulations within them are expressed in the seafloor as a series of distinct deeps: the Tiran and Dakar deeps in the southern basin, the Arnona and Aragonese deeps in the central basin, and the Elat Deep in the northern basin (Figure 3a). The Gulf of Elat is linked to the axial depression of the Red Sea via the Hume Deep, which lies to the south of the Straits of Tiran.

The Gulf of Elat is bordered on the east and on the west by normal faults, which are responsible for steep escarpments and submarine slopes. Gravity anomalies indicate that thick, >6—8 km sedimentary sections exist within the main basins.

The internal structure of the Gulf is dominated by intrabasinal longitudinal and oblique faults, which define the limits of the internal pull-aparts (Ben Avraham et al. 1979; Garfunkel & Ben-Avraham 1996, 2001). The sedimentary basin fill was affected by syn-depositional deformation. In the southern half of the gulf the fill is distorted, arched and folded, whereas in the northern basin the structure is rather simple (Ben-Avraham et al. 1979). Small, diapir-like domal structures in the south may be the result of mobility of middle-late Miocene evaporites extending from the Red Sea into the Gulf.

The active structures that define the large pull-aparts are probably relatively young, having formed during the last stages of lateral motion along the Dead Sea Fault (Garfunkel & Ben-Avraham 2001). However, the basin margins, especially in the southwest, are probably remnants of older (probably Miocene) pull-apart basins. This is suggested by the presence of the diapir-like structures, which indicates the presence of Miocene evaporites at depth. This suggests that a basin already existed during this period. The western border fault truncates some marginal strike-slip faults, indicating that the present structure was superimposed on older faults that were active during the early history of the Dead Sea Fault.

Shallow Section

Source Image Description
Ben-Avraham et al (2008) Fig. 2 - Gulf of Elat
The basins within the gulf reflect the complexity of faulting resulting from deeper processes and shifts in plate motion. The northern section is dominated by the relatively flat-bottom Elat Deep (Figure 2a ,b), marking the northern active pull apart, which is the largest and shallowest in the gulf. At 900 m water depth, it is approximately 50 km long, up to 25 km wide, and is bounded by longitudinal marginal faults. Evidence for faulting at its northern and southern edges has also been reported (Ben-Avraham & Garfunkel 1986, Ben-Avraham & Tibor 1993). This is the only area in the Gulf of Elat that is symmetrical in cross section with respect to the bathymetry. However, asymmetry prevails in the subbottom. The sedimentary section here wedges out to the west and truncates abruptly to the east (Ben-Avraham 1985) (Figure 2b). Although the bathymetry of the Elat Deep is quite flat and simple, the underlying structure suggests that it is composed of at least two elongated subbasins (Ben-Avraham 1985). The thickness of the syn-sedimentary unit within the Elat Deep suggests that the latest phase of its formation began in the early Pliocene (Ehrhardt et al. 2005). This may be related to a change in the motion of the plates bordering the Dead Sea fault, which took place at that time (Garfunkel 1981, Joffe & Garfunkel 1987). Although this probably was not an abrupt change, it must have influenced the deformation along the length of the Dead Sea Fault (Garfunkel 1981).

The central part of the gulf is the narrowest and deepest and is occupied by the central pull-apart basin. It is divided into two deeps, the Aragonese Deep (Figure 2a,c) (water depth of 1830 m) and the Arnona Deep (1550 m water depth), which are separated by a N-NE- to S-SW-trending uplifted and folded structural saddle. The sediments within the basin are considerably distorted into arches and open folds (e.g., the Arnona Deep is bounded on the east by a large fold that appears to be compressional feature), raising the possibility of local transpression. In the cen-tral basin, strike slip motion may occur on both sides of the basin and the major fault strands are arranged in an en-echelon pattern (Ben-Avraham 1985), although no displaced markers have been detected to prove this.

The southern basin of the Gulf of Elat is the widest and occupies its eastern side. As with the other basins, this one is also divided into two distinct and separate deeps: the Dakar Deep (Figure 2a,d) (1100 m water depth) and the southern-most Tiran Deep (1300 m). Some of the deformation observed in the southern basin may be the expression of underlying salt (or shale) diapirs. The transition from the northern basin to the southern basin is accompanied by a change in lateral asymmetry. This could be explained by the fact that on one side, the basins are bounded by a strike-slip master fault and on the other by a predominantly normal boundary fault. It is possible that asymmetry in the basin structure switches sides when motion along segments of the main fault switches from one segment to another (i.e., from one side of a basin to the other). In the southern basin the deeps are bounded on the west by a strike-slip master fault and on the east by a predominantly normal boundary fault, whereas in the north, the situation is reversed. The Straits of Tiran separate the Gulf of Elat from the Red Sea. Local transpression is evident by the uplifted island of the Tiran block, where Quaternary reefs can be found at elevations of 500 m above sea level (Goldberg & Beyth 1991). To the south, another deep is present: the Hume Deep.

The northern basin occupies most of the width of the gulf and is located centrally along the axis of the gulf. This is not the case for the central and southern basins. North of the Elat Deep lies the gulf's head. Here, the major strand of the Dead Sea Fault is on the western margin of the basin (Garfunkel 1981, Reches et al. 1987, Ben-Avraham & Tibor 1993). The shift from the eastern side of the Elat Deep to the western side of the gulf's head is achieved by a rather complicated fault system.

Deep Section

Source Image Description
Ben-Avraham et al (2008) Fig. 3 - Gulf of Elat
The southern third of the Gulf of Elat exhibits a rather smooth magnetic field in contrast to the rest of the gulf. The trend and amplitude of the magnetic anomalies over this part of the gulf differ vastly from those on land. Here, the gulf is comprised of almost a single magnetic anomaly, whose wavelength is long — approximately 20 km — and trends parallel to the strike of the gulf, NNE. However, the anomalies on land are of much shorter wavelengths and trend WNW. On land, igneous and metamorphic outcrops cause a disturbed magnetic field. In the gulf, the northern half of the gulf is more anomalous than the southern half. The trend of the anomalies in the north differs from those on land. One possibility is that slight counter-clockwise rotation took place relative to magnetic bodies on land. In general, the magnetic field does not comply with the location of the basins or deeps, suggesting that they are perhaps not deep-rooted features. Anomalies in the southern section of the Gulf indicate a depth of more than 10 km to the magnetic basement, indicating that in this area, a large vertical offset separates the basement on land from that under the gulf. Geophysical evidence (see below) suggests that the southern Gulf of Elat is subjected to different crustal processes than those acting in other parts (Ben-Avraham 1985). On land, several large magnetic anomalies extend into the Gulf of Elat. However, none of them extend from one coast to the other, across the entire gulf, supporting geological evidence for strike-slip motion between the Arabian and Sinai plates.

Two distinct lows are present in the free-air anomaly map of the gulf — the first low encloses the central basin and northern part of the southern basin, whereas the second low encloses the northern basin (both at contour —150 mGal). The two lows are separated by a sill separating the northern and central basins. According to the Bouguer anomaly map, which also indicates two lows (the —110 mGal contour enclosing the northern basin and only a relatively small area in the divide zone between the central and southern basins), the northern one is larger than the southern one (Ben-Avraham 1985).

Data from magnetic, gravity, and refraction surveys can be interpreted as showing a similarity between the deep structure of southern Gulf of Elat and the Red Sea (Ben-Avraham 1987). In this area, the magnetic anomalies are rather smooth and trend NNE parallel to the strike of the gulf. Free-air anomalies are less negative in the southern basin, indicating more compensation. Seismic refraction data from the western margin were interpreted (Ginzburg et al. 1981) to show a considerable thinning of the crust in its southern third (from 35 km in the north to 27 km in the south). This can be interpreted as showing that the crust in the southern third of the gulf is approximately 8 km thinner than in the north. High heat flow values (a mean of 93 mW m-2) in the southern part of the gulf toward the Red Sea support this interpretation (Ben-Avraham & Von Herzen 1987). These geophysical data, thus, indicate that in this area, transition from the structure of the Red Sea to that of the Dead Sea Fault occurs. Crustal thinning indicates mantle upwelling preceding spreading. Ben-Avraham (1987) suggested that crustal spreading activity is propagating from south (the Red Sea) to the north.

Alternatively, the refraction data can be interpreted as recording the crustal structure of the western margin of the gulf, as it indicates high velocities at the surface (characteristic of basement rocks exposed along the seismic profile). However, the seismic reflection data from the gulf and the gravity data (Ben Avraham 1985) show a >6—8-km-thick young fill within the gulf, and therefore require that along the gulf, the crystalline crust is thinner than along its margins and flanks. This can be interpreted as a result of the transtensional nature of the motion of the plates flanking the gulf (Garfunkel 1981). However, the envisaged plate separation is much less than in the Red Sea, so according to this interpretation, the major change in crustal structure would be south of the Straits of Tiran, approximately where the gravity anomaly becomes distinctly positive.

The central basin is narrower, more deformed, and more seismically active than the southern basin, although seismic activity may be a short-term phenomenon. Bouguer gravity values are higher than the adjacent gravity lows, indicating the presence of a thinner crust. Active tectonic processes probably affect both the shallow and deep crust in this area. It has been suggested that the tip of spreading propagation may be located presently in the central basin (Ben-Avraham 1987).

The northern third of the gulf is separated from the central basin by a structural sill that strongly influences the seismic activity, as indicated by the seismic moment release (Figure 3). Most of the aftershocks of the 1995 Gulf of Aqaba earthquake occurred in the southern basin, as predicted by the complex fault patterns in the area (Ben-Avraham & Garfunkel 1986). However, the majority of the moment release occurred along the structural sill (Hofstetter et al. 2003). The Elat Deep is flat and the bathymetry is reflected by the Bouguer values. Gravity and structural data indicate that the northern basin of the gulf continues on land into the Arava Valley.

On both sides of the Gulf of Elat and the southern Arava Valley, marginal faults, which are parallel to the Dead Sea Fault, can be found in zones approximately 20 km wide (e.g., Bender 1968; Eyal et al. 1980, 1981). The approximately 20-Ma-old Red Sea dike system and the Miocene Hazeva Formation are left-laterally displaced by these faults. A cumulative offset of more than 20 km can be found between both sides of the Dead Sea Fault in the Arava Valley. The marginal faults are truncated by the border faults of the Gulf of Elat, and thus predate them. Thus, it would seem that the basin began to form as a result of motion along the marginal faults but the present-day structure was formed when motion became concentrated within the gulf.

Arava Valley

To the north of the gulf's head lies the Arava Valley (Supplemental Figure 2), a 160-km-long morphotectonic depression connecting the Gulf of Elat and the Dead Sea (Bentor et al. 1965, Ben-Avraham et al. 1979). The southern part of the Arava Valley is marked by a structural saddle, resulting from mild compression (Garfunkel 1981, Garfunkel et al. 1981, Garfunkel & Ben-Avraham 2001, Basson et al. 2002). This compression can clearly be seen in upwarping Neogene and younger sediments. According to Ginat & Avni (1994) the saddle, located 50-70 km north of the Gulf of Elat, was high and wide enough to allow a major drainage line to flow westward across the Dead Sea Fault into the Negev during the Pliocene. In the southern section of the Arava, just north of the Gulf of Elat, compressional features developed in the sedimentary fill (Basson et al. 2002). These structures are young and formed as a result of recent changes in the geometry of the longitudinal master faults.

The Arava Valley is characterized by a series of elongated, en-echelon (left-stepping), tectonic subbasins filled with clastic sediments, the Evrona, Ya'alon, and Zofar basins, and the Shizaf Basin north of the Zofar Basin (Supplemental Figure 2a) (Bartov et al. 1998, Frieslander 2000). These basins are bound by subparallel high-angle strike-slip longitudinal faults with normal displacement and range in depths from several hundred meters to a few kilometers and are filled with elastic sediments. The southern ends of the Yaalon and Zofar basins seem to be aligned with the inter¬section of E-W Themed and Paran dextral faults in the central Sinai-Negev and the Dead Sea Fault. These E-W faults predate left-lateral motion along the Dead Sea Fault, as they themselves are displaced laterally. As such, they may have influenced the younger structure along the Arava.

To the north, the Zofar and Shezaf basins are delimited by two NW-SE striking listric faults with large displacements. In this area, the main strike-slip fault, the Arava Fault, is located on the eastern side of the valley (Garfunkel et al. 1981, Garfunkel 1981, Atallah 1992). Predominance of strike-slip motion along the eastern boundary fault may explain the asymmetry in the sedimentary fill of these subbasins. The Zofar Basin is bounded on the west and east by the Zofar and Arava faults, respectively. Whereas motion along the Arava Fault is left-lateral strike-slip, the young motion along the Zofar Fault is predominantly normal (Bartov et al. 1998, Frieslander 2000).

Seismic data indicates that the shallow Zofar Basin dips toward the north (Frieslander 2000). In addition, a unique reversal in asymmetry occurs (Supplemental Figure 2b,c), where horizons within the southern part of the basin dip toward the west, whereas in the northern section, horizons dip to the east. This is the only known example along the Arava Valley where reversal in asymmetry occurs within a subbasin (instead of between subbasins), and may indicate that the Zofar Basin can be structurally subdivided.

The Dead Sea

Introduction

Source Image Description
Ben-Avraham et al (2008) Fig. 4a - DTM Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4b - Geologic X-section
Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4c - Geologic X-section
Dead Sea Basin
Probably the most studied of the basins along the fault, the Dead Sea Basin, is the lowest continental depression (around 418 m below mean sea level) and is one of the largest pull-apart basins on Earth. In part, it is occupied by the Dead Sea, with a salinity of around 330 g kg-1. The basin extends from a structural saddle in the central Arava to north of Jericho (Neev & Emery 1967, Freund et al. 1970, Garfunkel 1981, ten Brink et al. 1993, Lazar et al. 2006). The basin is about 150 km long and between 15 to 17 km wide, with around 10 km of Neogene to recent sediment fill (Garfunkel & Ben-Avraham 1996, Ben-Avraham 1997, Ginzburg & Ben-Avraham 1997). The basin formed around 18-15 Ma ago (Garfunkel 1981, Horowitz 1987, Garfunkel & Ben Avraham 1996) between the left-stepping Arava Fault to the southeast and the Jericho Fault on the northwest (Figure 4). It is usually divided into two main basins, a northern basin and a southern basin, separated by the Lisan Peninsula, which is probably bounded to the north and south by large, deep-rooted, oblique normal faults. The Lisan Peninsula is a large buried salt diapir, acting as a buffer zone between the northern and southern basins.

The northern basin is occupied by a terminal lake known as the Dead Sea. The southern basin was once covered by a very shallow lake (part of the Dead Sea), which is now used by the Dead Sea Works (Israel) and by the Arab Potash Company (Jordan) as artificial evaporation ponds. Crustal studies indicate that a thick sedimentary fill characterizes the two basins. In the northern basin, this thick sedimentary fill reaches approximately 6 km and thins out toward the northern shore of the lake. The fill of the southern basin is somewhat greater, reaching a thickness of more than 14 km (Ginzburg & Ben-Avraham 1997). This indicates that the area is still subsiding.

According to stratigraphy, the Dead Sea Basin was already an accentuated depression during the early and middle Miocene (Horowitz 1987; Garfunkel & Ben-Avraham 1997, 2001). Clastic sediments of the Miocene Hazeva Formation were deposited by a river system that flowed across the Dead Sea Basin and indicate that during this period (late Miocene), sedimentation and subsidence kept pace and that the flanks of the basin did not form topographic barriers (Garfunkel 1997).

During the Pliocene or perhaps somewhat earlier, a thick evaporitic series consisting mainly of halite (the Sedom Formation) was deposited in the central section of theDead Sea Basin, while the southern section stopped subsiding (Zak 1967, Horowitz 1987). These evaporites were deposited in a marine embayment that extended from the Mediterranean to the Dead Sea area fault valley (Zak 1967). Sometime during the Pliocene, the connection with the sea was cut and the valley became a land-locked depression in which lakes of varying sizes developed according to climatic fluctuations.Since then, sedimentation has lagged behind subsidence, leaving a deep topographic depression composed of fluvial and lacustrine clastics and evaporates (Zak 1967, Gardosh et al, 1997).

Shallow Section

Introduction

Source Image Description
Ben-Avraham et al (2008) Fig. 4a - DTM Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4b - Geologic X-section
Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4c - Geologic X-section
Dead Sea Basin
Similar to the Gulf of Elat, the structure of the Dead Sea Basin is controlled by normal border faults and longitudinal intrabasinal faults, which take up most of the lateral motion. The Dead Sea is also divided into a number of subbasins, which are separated by active transverse faults (Figure 4a). However, in contrast to the Gulf of Elat, the subbasins are not arranged en-echelon. This is due to the fact that the Eastern and Western boundary faults limit the extent of transverse development. It seems that the Dead Sea Basin was delimited throughout its entire history by two left-stepping faults, and this basic configuration did not change.

The southern Dead Sea Basin

Source Image Description
Ben-Avraham et al (2008) Fig. 4a - DTM Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4b - Geologic X-section
Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4c - Geologic X-section
Dead Sea Basin
The southern Dead Sea Basin is divided into at least two distinct subbasins (ten-Brink & Ben-Avraham 1989, Ben-Avraham 1997): from the Iddan Fault in the central Arava to the Amazyahu Fault and from the Amazyahu Fault to the Boqeq Fault (Figure 4a). The subsurface of the southern basin is characterized by the presence of a thick layer of evaporates and by the presence of salt diapirs. Faulting, increased overburden, and subsequent upward flow of salt along active faults led to the formation of a number of diapirs — the Lisan and Sedom salt diapirs (e.g., Zak 1967, ten-Brink & Ben-Avraham 1989). The Lisan Peninsula separates the northern and southern basins, whereas the Mt. Sedom Diapir towers above the southern basin on its western margins. The Sedom diapir is the only exposed salt diapir in the Dead Sea Basin. The initiation of the Sedom salt diapir is related to faulting, increased overburden, and subsequent upward flow of salt along the Sedom Fault (e.g., Zak 1967, ten-Brink & Ben-Avraham 1989, Larsen et al. 2002). Increase in thickness of most of the fill layers on the downthrown side of the Sedom Fault indicates continuous salt flow and fault activity from the Early Pleistocene to the Holocene (ten Brink & Ben-Avraham 1989, Gardosh et al. 1997). The Sedom Diapir is currently rising. The calculated uplift rates show that the salt diapir is divided into two sections, with each section having a different uplift rate. Measurements were found to be 8.3 ± 0.3 and 6.9 ± 0.3 mm year-1 for the northern and southern blocks, respectively (Pe'eri et al. 2004).

The northern subbasin in the southern Dead Sea Basin is also the deepest and is bordered by faults that seem to be vertical — the transverse Boqeq Fault in the north, the transverse Amazyahu Fault in the south, the longitudinal Sedom Fault in the west, and the longitudinal Gur Safi Fault to the east (Figure 4b,c). The faults on the east and west accommodate strike-slip motion (van Eck & Hofstetter 1990), indicating an overlap in this motion. This led to a perfectly symmetrical (E-W and N-S) subbasin, which comprises a full graben (Figure 4b). On the surface, it is 30 km long and 18 km wide, while at depth it is 20 km long and 13 km wide. This subbasin is the deepest in the Dead Sea Basin. It has recently been hypothesized that this subbasin was the first to form during the early stages of the development of the Dead Sea Fault (Ben-Avraham & Schubert 2006). Growth was dictated by the tilting and collapse of blocks from the eastern and western margins.

Ben-Avraham & Schubert (2006) suggested that this deep subbasin is a unique basin along the entire Dead Sea Fault. According to some models for the formation of the Dead Sea Fault (Ben-Avraham & Lyakhovsky 1992, Lyakhovsky et al. 1994), it may have formed as a result of simultaneous propagation of two fracture zones at its northern and southern ends toward each other. The southern fault propagated northward owing to the opening of the Red Sea, whereas the northern fault propagated southward as a result of motion along the east Anatolian Fault and the collision of the Arabian and Eurasian plates. The tips of the two faults overlapped in the area of the southern subbasin and veered to join each other. This created an isolated piece of heavy lithosphere, which proceeded to sink into the underlying asthenosphere, creating a deep "drop down" subbasin at the surface. Continuing oblique strike-slip motion along the master fault led to the formation of additional pull-aparts to the north and south of this deep subbasin. This theory agrees with the theory recently proposed (Lazar et al. 2006) that the subbasins that make up the Dead Sea Basin developed together (but not necessarily simultaneously). A smaller subbasin, the Iddan Basin (depth to basement estimated at 10 km), lies to the south of the "drop down" subbasin. Additional subbasins may exist further south. However, their south¬ward extent would be limited by converging boundary faults. An alternative model (Garfunkel 1981, 1997) suggests that the basin formed where the fracture of the lithosphere was somewhat irregular, which led to local transtension and minor separation that eventually led to the development of en-echelon faults. Subsequently, this basin formed by becoming longer, as is theorized for all pull-apart basins that form between en-echelon strike slip faults.

The new data and interpretation of the southern and northern Dead Sea Basins (e.g., Ben-Avraham & Schubert, 2006, Lazar et al. 2006) also disagrees with the theory that the observed subsidence in the basin is the result of necking of the lower crust over a longer area than that defined by brittle deformation (Al-Zoubi & ten-Brink 2002). According to this theory, thinning of the overburden in the central part of the basin, and possibly thinning of the lower crust owing to N-S stretching, may drive lower crustal flow. Al-Zoubi & ten-Brink (2002) also suggested that lateral shear along the Dead Sea transform increases the likelihood that lower crustal flow is a significant factor in the subsidence of the Dead Sea basin. However, relocation of earthquake epicenters (Aldersons et al. 2003) indicates that the lower crust is prob¬ably brittle. Aldersons et al. (2003) revealed that the majority of well-constrained microearthquakes (ML less than 3.2) occurred much deeper than previously expected, exhibiting continuous focal depths down to the Moho (located at a depth of 32 km) (Supplemental Figure 3). Sixty percent of the earthquakes they examined nucleated at depths of 20-30 km. According to their study, the upper mantle appeared to be aseismic during the period they examined (14 years). In addition, refraction studies (e.g., Ginzburg & Ben-Avraham 1997) indicate the presence of a sharp step between the southern and northern basins, which is in keeping with a brittle lower crust, which is more easily broken and faulted, lending strength to the "deep basin" theory (Ben-Avraham & Schubert 2006). However, this structure needs further investigation, as the refraction line may have crossed from a marginal step in the north to the deeper central part of the basin in the south (see below).

In E-W cross section, the southern Dead Sea Basin is divided into three struc-tural step blocks: the rim block, the intermediate block, and the deep-sunken block (Kashai & Croker 1987, Al-Zoubi et al. 2002). E-W and NE-SW extension combined with predominant lateral motion is responsible for the formation of these blocks (Garfunkel 1981, Garfunkel & Ben-Avraham 1996).

The northern Dead Sea Basin

Source Image Description
Ben-Avraham et al (2008) Fig. 5 - Dead Sea Basin
sub-basins and
Gravity Map
The sedimentary fill of the northern basin, along the available seismic refraction profile, is somewhat thinner than that of the southern basin, 6—8 km at its deepest as opposed to approximately 14 km in the south (Ginzburg & Ben-Avraham 1997). The northern Dead Sea Basin can be divided into two sections based on morphology rather than structure. The southern section of the northern basin is comprised of a terminal hypersaline lake known as the Dead Sea. The northern section stems from the lake northward on land into the lower Jordan Valley. Although morphology dictates this division, structurally, the northern Dead Sea Basin is divided into a number of smaller subbasins in the deep subsurface. As with the southern Dead Sea Basin, these subbasins are separated by structural saddles and transverse faults, and grow shallower and narrower to the north (Figure 5). The deepest of these, which is located in the lake from the En Geddi Fault in the south to the Kalia Fault in the north, was termed the Arnon sink (Neev & Hall 1979) and is considered to mark the area of fastest late Pliocene and Pleistocene subsidence (Garfunkel & Ben-Avraham 2001).

The Kalia fault is an active transverse fault. It was first mapped in seismic cross sections by Lazar et al. (2006). Its trend and activity were later confirmed by a 5.1 Mb earthquake that occurred on the northeastern side of the lake and not on the Jericho fault as expected (e.g., Gardosh et al. 1990, Rotstein et al. 1991). Depth of the epicenter was calculated to be approximately 20 km (Al-Tarazi et al. 2006), in-dicating that this is a deep-rooted fault. This fault separates the Arnon sink (or basin) from the joint lake-land Kalia Basin (Lazar et al. 2006). This subbasin continues until around Jericho. The presence of transverse faults bounding subbasins disproves the theory that the northern end of the Dead Sea Basin sags toward the center (ten-Brink et al. 1993).

The transverse fault that bounds the Kalia Basin to the north (Lazar et al. 2006) is not the final bounding fault of the northern Dead Sea Basin (Figure 5), as hypothesized by Kashai & Croker (1987), ten-Brink et al. (1993), and Garfunkel & Ben-Avraham (2001), who suggested that the basin terminates near Jericho. To the north, another smaller subbasin, the Jericho Basin, was found, and north of this is the Fazael Basin, which is probably the last of the subbasins comprising the larger Dead Sea Basin. In this area, as in the south, the eastern and western boundary faults converge, preventing the development of additional subbasins to the north.

As stated above, the division into subbasins separated by active transverse faults fits with the theory that initially the deep subbasin in the southern Dead Sea Basin was formed, and later additional subbasins developed to the south and north owing to motion along the main Dead Sea Fault. It is also possible that converging boundary faults, both in the south and north, prevented the expansion of these subbasins to the north and south, but also limited their E-W development. Activity may occur on any of the transverse faults and subsidence is not limited to a single depocenter.

Deep Section

Source Image Description
Ben-Avraham et al (2008) Fig. 4a - DTM Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4b - Geologic X-section
Dead Sea Basin
Ben-Avraham et al (2008) Fig. 4c - Geologic X-section
Dead Sea Basin
The magnetic map of the northern Dead Sea Basin is smooth with few anomalies (Frieslander & Ben-Avraham 1989). It can be divided into two distinct parts, separated in the vicinity of the En-Geddi Fault (Figure 4). To the north, the field is smooth and trends N-S to NW. To the south the contours change trend and become much more disturbed. Small basaltic bodies are present at depth and some are the continuation of basaltic flows on land east of the basin (Ben-Avraham 1997). The magnetic anomalies on the land west of the Dead Sea extend uninterrupted over the lake (Frieslander & Ben-Avraham 1989, Ram 1989), indicating mostly normal faulting along the western side of the northern basin. However, contours are discontinuous across the eastern margin of the basin, suggesting predominate strike-slip motion resulting in major lithological changes across the transform.

Gravity studies indicate that the Dead Sea Basin becomes shallower and narrower to the north and south. A gradual decrease in the Bouguer gravity anomaly from both ends of the basin led ten Brink et al. (1993) to suggest that the basin sags toward its deepest part in the center. These data also show fault blocks, several kilometers in width, along the western side of the basin (the eastern side being occupied by a wide graben). These blocks were interpreted as evidence for passive collapse into the deepening graben (ten Brink et al. 1993). Gravity also suggests that the Moho is not significantly elevated under the basin, with deformation being limited to the crust. The division into subbasins is also evident in the gravity data.

A wide-angle seismic reflection-refraction experiment was conducted in the Dead Sea Basin and the areas area on land to the north (Ginzburg & Ben-Avraham 1997). The calculated profile revealed significant differences between its northern and south¬ern parts, which are separated by a substantial basement fault, downthrown to the south by 4-5 km. This fault may separate the northern and southern basins (although the strike of the fault is not well constrained, and it may also be a longitudinal fault along the basin margin). As a result of this faulting and intense sedimentation, a 6-8-km-thick sedimentary sequence was deposited in the north, and a 14-km-thick sequence was deposited in the south. Analysis of the recorded seismograms indicated the presence of two large Pliocene salt diapirs in the young basin fill. The southern diapir was interpreted as being part of the Lisan Diapir. Other than these two salt diapirs, it would seem that the refraction indicates that little salt is present in the fill of the Dead Sea Basin.

These differences between north and south may indicate that the northern basin is younger than the southern one or that greater subsidence was prevalent in the southern basin (Ben-Avraham & Lazar 2006). These data do not support the idea of a gradual sagging of the basin from north and south toward its center (as proposed by ten Brink et al. 1993 and Al-Zoubi & ten-Brink 2002).

Sea of Galilee-Kinneret Basin

Introduction

Source Image Description
Ben-Avraham et al (2008) Fig. 6 - Tectonic Map
Sea of Galilee
and Golan Heights
The Sea of Galilee (Lake Kinneret) is located at the northern section of the Kinneret-Beit Shean Basin (Figure 6a). The Sea of Galilee is a freshwater lake 12 km at its widest point and approximately 20 km long. Its surface is approximately 210 m below mean sea level and has a maximum depth of 46 m (Ben-Avraham et al. 1990). Owing to high sedimentation rates of 2-7 mm year (e.g., Serruya 1973, Inbar 1976) the lake floor is relatively flat. The basin began to form during the Neogene (Picard 1943, Schulmann 1962, Garfunkel & Ben-Avraham 2001) as a result of N-S motion along the Dead Sea Fault. This remains the major fault trend in the basin. However, a secondary NW-SE- to W-E-trending fault system composed of branching faults began to develop during the Middle Miocene on the western side of the basin (Shaliv 1991). This secondary system resulted in counterclockwise rotation (Ron et al. 1984) and reactivation of older systems, creating complicated structures and fault patterns. Farther north, additional faults branch off from the Dead Sea Fault (Ron et al. 1984, Heimann 1990) in the region of the southernmost Lebanon and Mount Hermon (Dubertret 1955, 1966; Walley 1988). North of the Sea of Galilee lies the Hula Valley, the northernmost pull-apart basin along the southern section of the Dead Sea Fault (Supplemental Appendix).

The structure of the Sea of Galilee is complex and results from the intersection of the two fault systems in the area. Superposition of vertical displacements perpendicular or oblique to the N-S transform fault created complicated structures in the area. Owing to Plio-Pleistocene basalt flows and intrusions varying in thickness, structural interpretation is difficult.

Shallow Section

Source Image Description
Ben-Avraham et al (2008) Fig. 6 - Tectonic Map
Sea of Galilee
and Golan Heights
Active tectonics in the Sea of Galilee is expressed by young faulted, tilted, and flexed sediments along the basin margins and in the deepest section of the lake (Ben-Avraham et al. 1981, Hurwitz et al. 2002, Reznikov et al. 2004). High values of heat flow of 93.4 mW m-2 (Ben-Avraham et al. 1978) and earthquake-induced surface ruptures from near the lake (Marco et al. 1997, 2003) are indicative of active faulting. Com¬pressional structures are found to the north and south of the basin (Rotstein & Bartov 1989, Rotstein et al. 1992). Plio-Pleistocene volcanic outcrops are present around the basin. The basin is divided into two distinct sections, although transi¬tion from one to another is still somewhat debated. The southern section, a deep pull-apart bordered by N-S-trending faults, is a continuation of the Kinnarot Valley (Figure 6a), which lies to the south (Ben-Avraham et al. 1996). Rotstein et al. (1992) suggested that the basin in the Kinnarot Valley was formed by two overlapping strike-slip faults. This section developed in the Miocene and was filled with more than 4 km of evaporites, igneous, and clastic rocks (Marcus & Slager 1985, Shaliv 1991).

Direct evidence for the subsurface composition of the Kinenret basin is provided by shallow water wells and the Zemah-1 (4249 m) well located south of the lake in the Kinnarot Valley (Marcus & Slager 1985). Thick strata composed of limestones, marls, conglomerates, salt, and igneous rocks are evident in the well.

Seismic profiles indicate that in the subsurface, the western flank of the basin is broken by a number of faults with downthrow toward the axis of the basin and to the southeast (Figure 6a). This seismically active, NW-SE-trending secondary fault system extends from the Galilee into the basin (Hurwitz et al. 2002 and references therein). It has been suggested 3%-6% of the total lateral slip has been shifted from the main strike-slip fault to these normal faults since the Pliocene (Hurwitz et al. 2002).

The structure of the deep graben beneath the Sea of Galilee is dominated by N-S-trending longitudinal marginal faults (Hurwitz et al. 2002). The eastern marginal fault is probably a continuation of the on land marginal fault in the south (Rotstein et al. 1992). The sedimentary section under the lake wedges toward this marginal fault. The southwestern marginal fault is also an extension of the on land fault to the south of the lake. To the north of the lake, the boundary fault zone was identified as a zone of intense deformation by Rotstein & Bartov (1989). This zone may coincide with a topographic terrace stemming for the Golan Heights. In addition, a longitudinal median step fault separating the shallow western section of the basin from the deep eastern section was observed in the central part of the lake (Reznikov et al. 2004). The eastern boundary fault extends to the northern part of the lake, whereas, at least on the surface, the western fault does not cross the northern section (Figure 6a).

Transverse faults delimiting the basin to the south and to the north have been inferred from seismic studies (Reznikov et al. 2004, and references therein). A central transverse fault dividing the basin into two may also be present (Ben-Avraham et al. 1996). This fault presumably separates the deep southern subbasin (containing at least 5–6 km of postrift sediments, including volcanic and intrusive evaporates) from a half-graben structure in the northeastern part of the lake. The southern sub-basin is delimited in the south by another transverse fault. A boundary fault with a large vertical throw delimits the northern subbasin in the east and a zone of complex deformation delimits it to the west.

Deep Section

Source Image Description
Ben-Avraham et al (2008) Fig. 6 - Tectonic Map
Sea of Galilee
and Golan Heights
The Sea of Galilee can be divided into two distinct domains based on the magnetic anomaly map: the magnetically quiet and smooth central section of the lake and the magnetically disturbed margins, with the anomalies continuing southward on land (Ben-Avraham et al. 1980, Ginzburg & Ben-Avraham 1986). The majority of the anomalies are elongated, have different trends, and some can be correlated with basalt outcrops. A few local narrow and steep gradient anomalies are probably caused by very shallow bodies or by faulting. Along the western margin of the lake, the magnetic anomaly zone continues southward on land to a distance of approximately 3 km,where the trend changes direction to NW-SE. This probably forms the southern boundary of the Sea of Galilee graben (Ginzburg & Ben-Avraham 1986). The SSW-trending magnetic anomaly of the eastern margin of the lake meets the western anomaly approximately 5 km south of the Sea of Galilee. No correlation has been found between the magnetic field and the bathymetry (Ben-Avraham et al. 1990), probably owing to the high sedimentation rates in the lake (2–7 mm year−1). Thus, according to Ben-Avraham et al. (1980), the magnetic anomalies in the Sea of Galilee are probably the result of lithological variability and structures within the sediments and basement.

Gravity data (Figure 6b) also indicate that the Sea of Galilee can be divided into two distinct sections (Ben-Avraham et al. 1996): The southern section is narrow and bordered by two N-S-trending faults, which were found by seismic reflection (Hurwitz et al. 2002). The largest gravity anomaly is located over this structure, which is thus the deepest part of the basin. It lies south of the bathymetric low, implying that it is an actively subsiding young feature.

The large negative gravity anomaly that characterizes the southern subbasin, indicates an ~8-km-thick sedimentary fill under this part of the lake, which thins to ~5 km toward its southern extension (Ben-Avraham et al. 1996). Gravity models (Figure 6b) also suggest that this subbasin is a full graben, indicating that strike-slip motion may be occurring simultaneously on both the western and eastern boundary faults (Ben-Avraham et al. 1996). This full graben may extend southward into the Kinnarot Valley. The southern subbasin is probably a pull-apart, which formed as a result of the transform motion along the main faults.

The wider northern section of the basin was probably formed as a result of interactions between the main N-S and secondary NW-SE- to E-W-trending fault systems (Ben-Avraham et al. 1981), i.e., rotational opening and transverse normal faulting. The northeastern part of the basin is occupied by an asymmetrical (to the east) half-graben structure, which differs from the deep asymmetrical subbasin of the southern section (Ben-Avraham et al. 1996, Hurwitz et al. 2002, Reznikov et al. 2004). The spatial extent of this asymmetrical graben is unclear (Ben-Avraham et al. 1996, Hurwitz et al. 2002, Reznikov et al. 2004) and is not considered to be a pull-apart basin (Ben-Avraham et al. 1996). Bathymetrically, this is the deepest part of the lake and thus, probably the most actively subsiding area. In addition, heat flow values in the lake are distinctly higher than on the transform flanks, indicating emplacement of hot material below the pull-apart basin (Garfunkel & Ben-Avraham 1996).

Discussion
Discussion

While the dominant motion along the Dead Sea Fault is left-lateral slip, transtensional and transpressional features are often present. Thus the motion is not purely strike-slip. In detail, the structure probably changed as a result of reorganization of plate boundary. Therefore, the structure is characterized by extensional, compressional, and asymmetrical structures varying in size from large scale (defining the general structure of the Dead Sea Fault Valley) to small scale (defining the internal structure).

Oblique strike-slip motion is largely responsible for the formation of the physiographic valley and of the pull-apart basins along the southern part of the Dead Sea Fault. This oblique plate separation resulted in the creation of collapsed structures and features evident in the many lows, i.e., pull apart basins, along the valley floor. Compressional features are caused by a step to the right of the main left-lateral strike-slip fault strands resulting in the formation of transverse structural saddles that separate the different basins. Within the pull-aparts, smaller compressional uplifts and saddles divide them into distinct subbasins.

Another characteristic feature of the Dead Sea Fault Valley is asymmetry, both in the large scale, as exhibited in the differing topography of the eastern and western-flanking cliffs, and the small scale, i.e., within the basins and subbasins. Large-scale asymmetry could have been influenced by the combined effect of nor-mal faulting and isostatic uplift on existing topography (Wdowinski & Zilberman 1996).

On the smaller scale, asymmetry within the pull-apart basins can be explained by transform-normal extension. Overlapping strike-slip faults, which are respon¬sible for the formation of these basins, together with rearrangement of plate motions, can result in the formation of an extensional component [the "leaky transform," termed by Garfunkel (1981)]. This asymmetry is not a local phenomenon, with other pull-apart basins along continental rift zones clearly showing asymmetry toward the main fault. Examples may be found in the Cariaco Basin in Venezuela (Ben-Avraham & Zoback 1992), the Motagua fault system in Guatemala, Lake Baikal, Lake Tanganyika in the East African rift system, and the North Anatolian Fault in Turkey.

Although it seems that motion along the Dead Sea Fault was initiated some 18 Ma ago (Garfunkel 1981, Garfunkel et al. 1981), the basins started to evolve later owing to changes in the fault pattern, sometimes probably related to slight shifts in the pole of rotation between the Arabian plate and the Sinai subplate. This introduced a component of oblique motion (Joffe & Garfunkel 1987), which led to the formation of the pull-aparts. An interesting question is the spatial evolution of the basins along the Dead Sea Fault, i.e., did the southern basins evolve before the northern ones, or did they all develop more or less simultaneously? Also, how did each basin evolve? For example, did the Dead Sea Basin develop from south to north as some researchers have suggested (e.g., ten-Brink & Ben-Avraham 1989), or from the center first and then to the north and south (Ben-Avraham & Schubert 2006), or simultaneously through the entire length (Lazar et al. 2006)?

One of the characteristics of the Dead Sea Fault is the deep seismicity. As demon¬strated by Aldersons et al. (2003, most microearthquakes nucleate at rather deep depths — some 20-30 km below the surface. In this respect, the Dead Sea Fault is quite different from the San Andreas Fault, where most earthquakes occur within the top 15 km of the subsurface, although some anomalously deep crustal earthquakes have been recorded at depths between 20-30 km (Bryant & Jones 1992). This may have to do with the differences in slip-rates between the two fault systems; the slip along the Dead Sea Fault is only approximately one tenth [4 mm year (Wdowinski et al. 2004)] the overall slip along the San Andreas Fault [-50 mm year-1 (van der Woerd et al. 2006)].

Another interesting characteristic is the low values of heat flow exhibited along the Dead Sea Fault, except for the Sea of Galilee, which is located within a large field of Tertiary volcanism and has a relatively high heat flow, and the southern Gulf of Elat, close to the Red Sea. The Dead Sea Basin exhibits the lowest heat flow values along the fault, 38 mW m-2 (Ben-Avraham et al. 1978). Recently, Forster et al. (2007) reported higher vales in southern Jordan. The deep seismically found by Aldersons et al. (2003) suggests that this may not represent the regional heat flux, but this issue requires further study.

The deep seismicity and low heat flow values suggest that the deformation might be brittle in the lower crust. This also supports recent claims (Ginzburg et al. 2007) that some of the transverse faults within the Dead Sea Basin are deep normal faults extending to the basement and no listric faults as previously assumed.

Source Image Description
Ben-Avraham et al (2008) Fig. 1 -
Ben-Avraham et al (2008) Fig. 2 -
Ben-Avraham et al (2008) Fig. 3 -
Ben-Avraham et al (2008) Fig. 4a -
Ben-Avraham et al (2008) Fig. 4b -
Ben-Avraham et al (2008) Fig. 4c -
Ben-Avraham et al (2008) Fig. 5 -
Ben-Avraham et al (2008) Fig. 6 -

Segmentation of the Levant Continental Margin ... - Ben-Avraham et al (2006)

The Levant continental margin is divided into two units. From south to north these are Negev, Judea-Samaria, major segments by the Carmel structure, which and Galilee-Lebanon (Figure 2b). Ben-Gai and Ben-Avraham extends from the Dead Sea fault into the eastern [1995] divided the Levant continental margin into two crustal Mediterranean.
...
On the basis of seismicrefraction [Ginzburg and Ben-Avraham, 1992; Ben-Avrahamet al., 2002], gravity [e.g., Ben-Avraham and Ginzburg,1990] and magnetics [Ben-Avraham and Ginzburg, 1986], the area onland was divided into several distinct crustal units. From south to north these are Negev, Judea-Samaria, and Galilee-Lebanon (Figure 2b). Ben-Gai and Ben-Avraham [1995] divided the Levant continental margin into two crustal segments north and south of the Cannel structure, which correspond to the division onland between Judea-Samaria and Galilee-Lebanon. This is clearly seen in the pattern of the magnetic field [Folkman, 1980; Ben-Avraham and Ginzburg, 1986; Rybakov et al., 2000] and suggests that these segments were probably formed through different breakup processes [Ginzburg et al., 1975; Neev and Ben-Avraham, 1977; Ben-Avraham and Hall, 1977; Ginzburg and Ben-Avraham, 1992]. Offshore, Ben-Avraham and Ginzburg [1990] distinguished between the crustal structure of the Levant Basin and the Eratosthenes seamount. The Levant Basin crustal unit is underlined by oceanic crust, covered by a 10-14 km thick sedimentary sequence [Ben-Avraham et al., 2002]. In this area only the upper 3 km of the subsurface are known in detail, offshore the southern seg¬ment of the Levant margin [Garfunkel, 1998]. However, little is known about the subsurface of the basin offshore the northern segment.
Description Image Source
Fig. 1 - DTM Map E Med. Ben-Avraham et al (2006)
Fig. 2 - Formation of Levant Basin Ben-Avraham et al (2006)
Fig. 3 - Seismic Survey Lines Ben-Avraham et al (2006)
Fig. 4 - Magnetic Intensity Anomaly Map Ben-Avraham et al (2006)
Fig. 5 - Bathymetry/Topography and
Top and Base of
Messinian Evaporitic Sequence
Ben-Avraham et al (2006)
Fig. 6 - Offshore Lebanon Profile Ben-Avraham et al (2006)
Fig. 7 - Profile 11 Ben-Avraham et al (2006)
Fig. 8 - Profile 1 Ben-Avraham et al (2006)
Fig. 9 - Profile S of Carmel Structure Ben-Avraham et al (2006)
Fig. 10 - Profile 4 Ben-Avraham et al (2006)
Fig. 11 - Profile S of Carmel Structure Ben-Avraham et al (2006)
Fig. 12 - Map of Major Tectonic Elements Ben-Avraham et al (2006)
Fig. 13 - Isopach depth map of
Messinian evaporitic sequence
Ben-Avraham et al (2006)
Fig. 14 - Shelf Marginal Wedge
S Lebanon vs. Central Israel
Depositional Model
Ben-Avraham et al (2006)
Fig. 15 - Shore Perpindicular Seismic Profiles
and one Shore Parallel Seismic Profile
Ben-Avraham et al (2006)
Fig. 16 - Seismic Profiles 3, 4, & 5 Ben-Avraham et al (2006)

Coastal Uplift

Figures

  • Fig 2c Bench elevations from Elias et al (2007)
  • Fig 1 Coastal uplift for the northern DST from Elias et al (2007)
Elias et al (2007) and previous researchers (Morhange et al, 2006) examined uplifted benches on the Lebanese coast between Sarafand and Tripolis; some in the vicinity of Tabarja (~20 km. NE of Beirut). They radiocarbon dated fossil Vermetids on the tops of these benches in order to estimate when the bench top was last in the sub tidal zone (which approximates mean sea level). They identified four uplifts from 3 or more [sizeable Mw = ~7.5] earthquakes in the past ca. 6-7 ka. They attributed the latest uplift (B1) to the 551 CE Beirut Quake while the earlier events (B2, B3, and B4) were no more precisely dated than between ~5000 BCE and 551 CE. Bench uplift on the earlier events (B2, B3, and B4) would likely have been due to uplift on the Mount Lebanon Thrust system - as was surmised for Event B1 and the 551 CE Beirut Quake.

Sivan et al (2010) performed a similar analysis over a broader geographical area. They measured elevations and radiocarbon dated fossil Vermetids on the tops of benches in order to estimate when the bench top was last in the sub tidal zone (which approximates mean sea level). They then subtracted out eustatic and glacio-hydro isostatic components of uplift to arrive at an estimate of tectonic uplift. Tectonic uplift showed a general trend of increasing in an northerly direction varying from within the range of error (±10 cm) at present day mean sea level in the northern Israeli coast to up to +390 cm at the Orontes north site in Turkey. They noted the following
This positive gradient in vertical tectonic displacement could have been explained by elastic bending of the plate. However, studies from the last several decades show that the region is extensively faulted and folded (e.g., Freund and Tarling, 1979; Beydoun, 1981; Ron, 1987; Walley, 1998; Ben-Avraham et al., 2006; Schattner et al., 2006; Elias et al., 2007; Carton et al., 2009). Therefore the gradient suggests a northward increase in brittle failure of the Levantine coast during the Late Holocene.
They divided the coast into structural segments which they described and interpreted tectonically in the table below:
Segment Description Interpretation
south of Galilee–Lebanon south of the Carmel fault The smallest vertical displacement values 1.5 cm (Habonim site, Table 1), calculated for northern Israel, correspond well with the zero displacement reported by Wdowinski et al. (2004) based on GPS measurements. Negligible vertical displacement was also reported based on archaeological evidence: the Galilee coast was stable for the last 3000– 4000 years (Sivan and Galili, 1999), while coastal sites in Caesarea indicate stability for the last 2000 years (Sivan et al., 2004). Since northern Israel was shown to be in isostatic equilibrium (Segev et al., 2006) the negligible vertical displacements suggest that it is tectonically stable (vertical displacements) during the last two millennia. Hence any measured change in relative sea level in this region stems conclusively from eustasy.
Galilee–Lebanon bounded by the Carmel and Roum faults (south and north) Further north along the coast of southern Lebanon (Galilee– Lebanon structural segment) slightly larger values are calculated for the vertical tectonic displacement, between 50 and 150 cm (sites ZireSaida, Ras Qantara, Hotel Mounes and Ras Abou Zeid, Table 1; Marriner and Morhange, 2005; Morhange et al., 2006). Along this segment, which is bounded by the Roum fault in the north, the topography becomes progressively more prominent (∼1000 m) northwards (Schattner et al., 2006 and references therein). Internal deformations produced by the nearby DSF are manifested by second order southwest trending dextral faults (Ron, 1987; Walley, 1998) which extend to the Levantine coast. Our results show that differential vertical displacement occurs along the coast of the Galilee–Lebanon segment during the Holocene
Lebanon Mountains consists of the Western Lebanon Flexure (Walley, 1998) of the highly elevated part of the Lebanese restraining bend. This segment is bounded to the west by the marine “Beirut–Tripoli thrust” (after Daëron et al., 2001) Much higher displacement values are calculated for the Lebanon Mountains segment north and east of the Roum fault. Both this segments and the Galilee–Lebanon are located along the Lebanese restraining bend, a right-step of the sinistral DSF along the NNE trending Yammunneh fault (Fig. 1). This slight divergence from the Nstriking axis of the DSF induces crustal overlap which is mainly absorbed by north-westward push of the Lebanon Mountains segment against the marine Beirut–Tripoli thrust (Schattner et al., 2006 and references therein). The highly elevated topography of this segment (∼3000 m) is not in isostatic equilibrium (Segev et al., 2006). It extends from the Lebanon and Anti-Lebanon mountains through the western Lebanon flexure to the coastline, where our results show high displacement values for the Holocene period, ranging between 60 and 340 cm (sites Phare, Palmier, Hannouch, Ras Koubba, Selaata, Ras Madfoun, Fidar sud, Nahr Ibrahim, North of Bouar, Safra, Tabarja, and South of Tabarja, Table 1; Sanlaville et al., 1997; Morhange et al., 2006).
Northern Lebanon a low laying topography juxtaposed from the south to the Cyprus arc convergent plate boundary Only one Dendropoma site was measured along the coast of the northern Lebanese segment (site Tell Soukas, Table 1; Fig. 1; Sanlaville et al., 1997). In this low topography segment the calculated vertical displacement is 41 cm. The site is located closely south to the intersection of the Levantine coast with the Larnaka ridge (part of the Cyprus arc convergence plate margin). North of the Larnaka ridge, however, displacements show higher values, ranging between 40 and 130 cm (sites Ras Ibn Hani, Ras el Karm (Ibn Hani) and Maksar, Table 1; Sanlaville et al., 1997). This jump in displacement values reflects the active convergence across the easternmost Cyprus arc, where the latter sites are overthrusted.
Western Syria consists a part of the triple junction between the DSF, East Anatolian fault and the Cyprus arc. Two main ridges of the arc deform the coasts of this segment — Latakia and Larnaka ridges (e.g., Kempler, 1998; Robertson, 1998). A similar change in the amount of vertical displacement is observed further north across the Latakia Ridge (part of the Cyprus arc convergence plate margin; Fig. 1). In Ras el Bassit site, south of the ridge, values range between 70 and 106 cm, while north of it vertical displacement extends between 142 and 360 cm — the highest values obtained along the entire margin (sites Orontes north, Table 1; Pirazzoli et al., 1991; Sanlaville et al., 1997). The northernmost sites (sites Guverdijne Kaya south, Table 1; Sanlaville et al., 1997) are also displaced and are located along the Kyrenia–Misis Ridge of the Cyprus arc. Here the values range between 85 and 230 cm, yet no comparable sites were sampled across the ridge
References

The Seismogenic Thickness in the Dead Sea Area - Alderson and Ben-Avraham (2014)

Source Image Description
Aldersons and Ben-Avraham (2014) Fig. 3.1 - Depth section of well-constrained seismicity
Aldersons and Ben-Avraham (2014) Fig. 3.2 - MW 5.3 earthquake of 11 February 2004
Aldersons and Ben-Avraham (2014) Fig. 3.3 - MW 5.3 earthquake of 11 February 2004
Aldersons and Ben-Avraham (2014) Fig. 3.4 - Depth distribution for the 188 earthquakes (1986–2001)
Aldersons and Ben-Avraham (2014) Fig. 3.5 - Depth distribution for the 188 earthquakes (1986–2001)
Aldersons and Ben-Avraham (2014) Fig. 3.6 - MW 6.3 earthquake of 11 July 1927
Aldersons and Ben-Avraham (2014) Fig. 3.7 - Isoseismal map (MMI) of the earthquake of 11 July 1927
Aldersons and Ben-Avraham (2014) Fig. 3.8 - Isoseismal map (MSK) of the earthquake of 11 July 1927
Aldersons and Ben-Avraham (2014) Fig. 3.9 - Preferred epicenter, and tentative causative fault
(Version 1) of the MW 6.3 earthquake of 11 July 1927
Aldersons and Ben-Avraham (2014) Fig. 3.10 - Preferred epicenter, and tentative causative fault
(Version 2) of the MW 6.3 earthquake of 11 July 1927
Aldersons and Ben-Avraham (2014) Fig. 3.11 - Ground fissures caused by the 1927 earthquake
Aldersons and Ben-Avraham (2014) Fig. 3.12 - Rheology of the Dead Sea Basin
Aldersons and Ben-Avraham (2014) Fig. 3.13 - Rheology of the Dead Sea Basin
Aldersons and Ben-Avraham (2014) Fig. 3.14 - Temperature distribution along the Dead Sea Fault
from the Sea of Galilee to the Gulf of Aqaba-Elat
Aldersons and Ben-Avraham (2014) Fig. 3.15 - Isoseismal map (MSK) of the earthquake of 11 July 1927
produced by kriging of Mode Intensities determined by Avni (1999)
Aldersons and Ben-Avraham (2014) Fig. 3.16 - Isoseismal map (MSK) of the earthquake of 11 July 1927
produced by kriging of Max Intensities determined by Avni (1999)
Aldersons and Ben-Avraham (2014) Fig. 3.17 - Isoseismal map (MSK) of the earthquake of 11 July 1927
produced by kriging of Modal Intensities determined by Avni (1999)
Aldersons and Ben-Avraham (2014) Table 3.1 - Velocity model Dead Sea 2013
Aldersons and Ben-Avraham (2014) Table 3.2 - parameters for the MW 5.3 earthquake of 11 February 2004
Aldersons and Ben-Avraham (2014) Table 3.3 - Time span and Seismogenic Thickness
Aldersons and Ben-Avraham (2014) Table 3.4 - Earthquake of 11 July 1927: main parameters
Aldersons and Ben-Avraham (2014) Table 3.5 - Earthquake of 11 July 1927.
Source parameters from spectral amplitudes
Aldersons and Ben-Avraham (2014) Table 3.5 - Earthquake of 11 July 1927.
Source parameters from spectral amplitudes

Stress evolution and seismic hazard of the Dead Sea Fault System - Heidbach and Ben-Avraham (2007)

Description Image Source
Fig. 1 - Recorded Seismicity Heidbach and Ben-Avraham (2007)
Fig. 2 - Assumed Historical
Quakes
Heidbach and Ben-Avraham (2007)
Fig. 3 - Assumed Historical
Quake Breaks
Heidbach and Ben-Avraham (2007)
Fig. 4 - Fault Segments (3D) Heidbach and Ben-Avraham (2007)
Fig. 5a - ∆CFS since
551 CE
South DSF
Heidbach and Ben-Avraham (2007)
Fig. 5b - ∆CFS since
551 CE
North DSF
Heidbach and Ben-Avraham (2007)
Fig. 5c - ∆CFS since
551 CE
Carmel Fault
Heidbach and Ben-Avraham (2007)
Fig. 5d - ∆CFS since
551 CE
Roum Fault
Heidbach and Ben-Avraham (2007)
Fig. 5e - ∆CFS since
551 CE
Yammouneh Fault
Heidbach and Ben-Avraham (2007)
Fig. 6 - Est. Fault
Stress in 2005
Heidbach and Ben-Avraham (2007)
Table 1 - Model parameters
for historical earthquakes
Heidbach and Ben-Avraham (2007)
Table 2 - Results of
ΔCFS analysis
Heidbach and Ben-Avraham (2007)

... faulting near the intersection of DSF and Carmel−Gilboa−Faria Fault System - Hamiel et al (2022)

Abstract

Crustal deformation and seismicity in the Levant region are mainly related to the plate-boundary Dead Sea Fault (DSF) and the intraplate Carmel-Gilboa-Faria Fault System (CGFS). The intersection between these two major fault systems is generally treated as an ~35-km-wide deformation belt stretched between the Faria and Gilboa Faults. Here, we present spatial and temporal analysis of faulting near this intersection. Our analysis is based on new geological mapping, new high-resolution airborne light detection and ranging (LiDAR) data, and seismic reflection profiles and indicates northward migration and localization of the intersection over time since the early Miocene. We discovered and mapped outcrops of Miocene, Pliocene, and Pleistocene rock units as well as faults and reconstructed the evolution of deformation. Three main tectonic phases were identified in this area covering the following periods: the early-middle Miocene, the late Mio¬cene-Pliocene, and the Quaternary. During the first phase, the DSF and the CGFS developed, and the CGFS faulted along a series of subparallel grabens and elongated NW-SE, between the southernmost Faria and the northernmost Gilboa faults, over a belt width of —35 km. During the second phase, deformation along the CGFS migrated northward and concentrated at an ~6-km-wide zone in the northern Faria Anticline. During the third stage, small-scale northward migration and localization of the deformation to a width zone of ~1-2 km at the southern boundary of the Beit She'an Valley occurred. Faults from the third phase reveal both sinistral and nor¬mal faulting. We propose that the currently active intersection between the DSF and the CGFS is located east of this localized deformation zone, near a right step of the DSF and the uplifted area of Tel Al-Qarn in the eastern Jordan Valley. We suggest that the northward migration and localization of this intersection are related to regional tectonic changes, spatial variations in the Sinai-Arabia Euler pole, and the localization of deformation along the DSF.

Carmel-Gilboa-Faria Fault System (CGFS)

The CGFS, a major intraplate fault within the Sinai plate (Fig. 1), defines the boundary between two major tectonic blocks within the Sinai plate (e.g., Dembo et al., 2015; Gomez et al., 2020; Hamiel and Piatibratova, 2021). The CGFS is ~80 km long, branches from the DSF at the central part of the Jordan Valley segment, and continues in the ~NW direction to the northern tip of Mt. Carmel and farther northwestward into the Mediterranean continental shelf (e.g., Garfunkel and Almagor, 1984; Ben-Gai and Ben-Avraham, 1995; Hofstetter et al., 1996). The Carmel, Gilboa, and Faria Faults comprise the main fault segments of this system (Fig. 1B). The Carmel Fault, the northwesternmost fault segment of the CGFS, is an oblique, sinistral-normal fault (e.g., Ben-Gai and Ben-Avraham, 1995; Achmon and Ben-Avraham, 1997; Sadeh et al., 2012; Dembo et al., 2015). The Gilboa Fault in the north and the Faria Fault in the south are the two main easternmost segments of this system (Fig. 1). They are normal faults that dip in the NE direction (Hatzor and Reches, 1990; Shaliv eta., 1991). Recent geodetic studies (e.g., Sadeh et al., 2012; Hamiel et al., 2016, 2018; Gomez et al., 2020; Hamiel and Piatibratova, 2021) showed that the current strike-slip rate along the DSF decreases from ~5 mm/yr south of the intersection with the CGFS to ~4 mm/ yr north of it, which suggests that slip is transferred from the DSF to the CGFS and that an active triple junction is located at the intersec¬tion between them. These studies also calculated an oblique strike-slip motion of 0.5-1.6 mm/yr across main segments of the CGFS and proposed that the observed reduction in slip rate along the DSF is due to transfer of slip to the CGFS. Hamiel and Piatibratova (2021) demonstrated a relative motion of 0.8 ± 0.4 mm/yr, sub-parallel to the DSF, between the southern and central Sinai tectonic blocks, i.e., south and north of the CGFS. This result was found to be in a good agreement with the CGFS total slip rate vector and the observed reduction in slip rate along the DSF near the intersection with the CGFS.

The intersection between the DSF and the CGFS

The intersection between the DSF and the CGFS is generally treated as an ~35-km-wide deformation zone between the Faria and Gilboa Faults and the central Jordan Valley segment of the DSF (e.g., Shaliv et al., 1991; Segev et al., 2014; Dembo et al., 2015). The northern end of this intersection zone is defined by the Beit She'an Valley, which is located between the Gilboa Fault and the DSF (Fig. 1B). However, it is unclear whether the entire zone between the Gilboa and Faria Faults is active, and if not, where exactly this intersection occurs and how it evolved over time since its formation in the Miocene. In this study, we investigate the spatial and temporal variations in the deformational processes of this area and better locate the intersec¬tion of the DSF and CGFS. To accomplish these tasks, we performed an integrated geological and geophysical study of faulting in this region.

Geological Background and Tectonic Setting

The study area extends from the Faria Fault in the south, through the Faria Anticline, and to the southern part of the down-faulted basin of the Beit She'an Valley in the north (Fig. 1C). This area developed as an interaction zone between the NW-trending CGFS and the N-trending tectonic system of the DSF. The axis of the Faria Anticline is in the SSW—NNE direction. The rock formations exposed in this region belong to several stratigraphic groups, ranging from the Jurassic Arad Group to the Quaternary Dead Sea Group. While old rock formations, i.e., Jurassic and Lower Cretaceous rock formations, are exposed at the core of the Faria Anticline, younger rocks of the Upper Cretaceous to Paleocene Mount Scopus and the Eocene Avedat Groups are exposed along the eastern flank of the anticline, forming the western boundary of the Jordan Valley (Mimran, 1984; Shaliv et al., 1991). The Faria Anticline is traversed by several NW-trending faults that demonstrate vertical displacements ranging from 500 m to 800 m. The major faults are the Faria Fault in the south and the southern boundary of the Beit She'an Valley in the north. Between the Faria Fault and Beit She'an Valley, less prominent faults also oriented NW, such as the Buqea and Tayasir Faults (Fig. 1), are observed (Shaliv et al., 1991; Mimran et al., 2016).

The Faria Anticline developed during the late Turonian—early Eocene as part of the Syrian arc structures (Krenkel, 1924) and was reactivated during the Oligocene—early Miocene as part of a tectonic phase that followed plate separation along the Red Sea (Mimran, 1984; Shaliv et al., 1991). Later, during the Miocene, this region was faulted along the north-south-trending DSF and along the NW-trending CGFS to form a major intersection area (Freund et al., 1970; Garfunkel, 1981; Mimran, 1984; Shaliv et al., 1991; Rozenbaum et al., 2016). The simultaneous activity of faulting of the two systems is still ongoing and divides the Sinai Plate into two tectonic domains south and north of the CGFS (Ben-Avraham and Ginzburg, 1990; Hofstetter et al., 1996; Sadeh et al., 2012; Dembo et al., 2015; Gomez et al., 2020; Hamiel and Piatibra-tova, 2021). Early Miocene faulting was accompanied by extensive volcanic activity, which is termed the Lower Basalt (e.g., Schulman, 1962). Near the Gilboa Fault (Fig. 1B), this volcanic phase started at ca. 17.5 Ma (Shaliv et al., 1991). Normal faulting accompanying this phase probably occurred under the same tensile regime that triggered volcanic eruptions during the Lower Basalt period (e.g., Hatzor and Reches, 1990; Shaliv et al., 1991; Dembo et al., 2015). The deposition of the Hordos conglomerate during the early—middle Miocene on the eastern flank of the Faria Anticline marks the development of a deep tectonic basin along the Jordan Valley at that time. Most of the clastic materials deposited within this basin were derived from the eastern flank of the Faria Anticline and expose Late Cretaceous to Eocene rocks. Similar conglomerates were derived from the high, faulted flanks of the Faria, Buqea, and Tayasir Grabens, which indicates the formation of these grabens that dissected the axis of the Faria Anticline in the early Miocene (Bentor, 1961; Schulman and Rosenthal, 1968; Shaliv et al., 1991).

During the late Miocene, a new phase of normal faulting occurred. These faults, striking to the NW-SE direction, ruptured only the northern part of the Faria Anticline and enhanced the tectonic relief along pre-existing faults that originally developed in the early Miocene (Shaliv et al., 1991). One of the larger tectonic basins that developed at this stage is the Beit She'an basin, which accumulated a thick sequence of the late Miocene Bira Formation (ca. 10-7 Ma; Rozenbaum et al., 2019) and was followed by the deposition of the late Miocene to early Pliocene Gesher Formation (ca. 7-5 Ma; Rozenbaum et al., 2019).

During the late Miocene and the early Pliocene, regional extensional deformation occurred in northern Israel. This is evidenced by the extensive volcanic activity that formed the Cover Basalt Formation in northern Israel (Shaliv et al., 1991; Heimann et al., 1996). In the northeastern sector of our study area, near Manna Feiyad (Figs. 2A and 2C), several volcanic bodies of this phase are exposed and dated to 5.65-5.90 Ma (Shaliv et al., 1991; Dembo et al., 2015). A tectonic phase of faulting in the northern Faria Anticline occurred after the deposition of the Cover Basalt. This tectonic phase triggered the incision of the drain-age system on which the early Quaternary Wadi Malih Formation was deposited, especially at the outlet of the Wadi Malih drainage system to the southeastern part of Beit She'an Valley (Fig. 2), forming an extensive fan (Mimran, 1984; Shaliv et al., 1991; Mimran et al., 2016). Farther north, a similar fan was deposited at the outlet of Nahal Bezeq draining the southern flank of the Gilboa Mountains into the Beit She' an Valley (Figs. 1-2; Hatzor, 2000). The late Quaternary is characterized by tectonic faulting along the western margin of the Jordan Valley, followed by deep erosion and later deposition of the late Pleistocene Lisan Formation (e.g., Begin et al., 1974). Recent tectonics are demonstrated by the deformation of Lisan and younger sediments.

Based on the studies mentioned above, Mimran et al. (2016) published an updated version of the geological map of the northern part of the region. However, the southern part of our study area is only described by a regional scale (1:200,000) geological map (Sneh et al., 1998), which in some places does not correlate well with the map of Mimran et al. (2016). Furthermore, previous works (e.g., Schulman and Rosenthal, 1968; Mimran, 1984; Shaliv et al., 1991) and geological maps (Sneh et al., 1998; Mimran et al., 2016) correlate the Neogene sequences exposed in the study area to the Neogene sequence exposed ~30-50 km to the north, near the Sea of Galilee, which was described by Picard (1943) and Schulman (1962). This long-distance correlation ignores the large variety of local facies changes observed within the stratigraphic sequence studied and therefore may cause misleading correlations.

In the present work, we demonstrate that careful examination and revision of this correlation leads to a better understanding of the geological sedimentary sequence and sheds new light on the Neogene-Quaternary tectonic evolution of our study area and in particular on the DSF-CGFS intersection.

Discussion and Concluding Remarks

Our observations indicate dramatic changes in the tectonic setting of the intersection between the DSF and the CGFS since the late Mio cene. This change is manifested in both the stratigraphy and the faulting architecture of this intersection area. As described above and summarized here, there are several major differences between our new map and previous geological maps of this area: (1) based on lithological characteristics, dating, and field relations, large outcrops along the northern part of the Faria Anticline, which border the Beit She'an Valley and were previously mapped as undivided outcrops of the Hordos and Umm Sabune Formation of early-late Miocene Age (Shaliv et al., 1991; Mimran et al., 2016), were found to belong to younger Neogene 2016), were found to belong to younger Neogene and late Miocene–early Pliocene Gesher Formations. (2) Some of these previously mapped outcrops (Shaliv et  al., 1991; Mimran et  al., 2016) were found to be pseudo-conglomerates of pedogenic origin, locally known as “Nari” of Pliocene to early Quaternary age, which formed due to weathering and pedogenesis on the exposed outcrops. (3) New outcrops of the Quaternary Wadi Malih Formation were found and mapped. (4) Based on lithological characteristics and field relations, lage outcrops along and near the Faria Graben, which were previously mapped as unclassified Neogene–Quaternar units (Sneh et al., 1998), were found to belong to the early–middle Miocene Hordos Formation. (5) New outcrops of the Hordos Formation com-posed of polymictic conglomerates cemented by carbonates were found and mapped in the southern part of the Faria Anticline, along and between the Faria and Buqea Grabens. (6) New faults that rupture Neogene and Quaternary units were found and mapped, especially in the northern sector of the Faria Anticline. Most of them are normal faults, and two are oblique faults, which demonstrate a combination of sinistral and normal faulting. (7) Faults that rupture Pliocene and Quaternary units were only found along the northern part of the Faria Anticline and the south-ern part of the Beit She’an Valley. We show that since the late Miocene, the fragmentation of the Sinai Plate along the CGFS has been accompanied by northward migration and localization of deformaWe show that since the late Miocene, the fragmentation of the Sinai Plate along the CGFS has been accompanied by northward migration and localization of deformation to the southern boundary of the Beit She'an Valley. Figure 13 summarizes the spatial distribution of the main phases of tectonic evolution. During the early—middle Miocene, the CGFS was composed of an —35-km-wide deformation belt that stretched from the Faria to Beit She'an Valleys. This belt was dominated by normal faulting and created NW-trending grabens, such as Faria, Buqea, and Tayasir Grabens (Fig. 13). Then, during the late Miocene and the Pliocene, tectonic activ¬ity migrated northward to a deformation belt of —6 km in the northern Faria Anticline (Fig. 13). Since the late Pleistocene, the deformation has been localized to a zone of —1-2 km along the southern boundary of the Beit She'an Valley (Fig. 13). At this stage, both normal and sinistral faulting can be clearly identified (Figs. 6, 10, and 12). A major fault in this localized zone is the Wadi Malih Fault, which ruptured the late Pleistocene Lisan Formation (Figs. 3, 6, 8, and 13). It marks a clear lineament that crosses the Jordan Valley as it branches from the DSF (at the eastern side of the Jordan Valley), where a right-lateral step and a change in the strike of the DSF are observed near the uplifted (push-up) area of Tel Al-Qarn (e.g., Ferry et al., 2007; Figs. 2A and S1). This fault defines the current boundary between the Faria Anticline and Beit She' an Valtion to the southern boundary of the Beit She’an Valley. Figure 13 summarizes the spatial distribution of the main phases of tectonic evolution. During the early–middle Miocene, the CGFS was composed of an ~35-km-wide deformation belt that stretched from the Faria to Beit She’an Valleys.

We show that since the late Miocene, the fragmentation of the Sinai Plate along the CGFS has been accompanied by northward migration and localization of deformation to the southern boundary of the Beit She'an Valley. Figure 13 summarizes the spatial distribution of the main phases of tectonic evolution. During the early—middle Miocene, the CGFS was composed of an ~35-km-wide deformation belt that stretched from the Faria to Beit She'an Valleys. This belt was dominated by normal faulting and created NW-trending grabens, such as Faria, Buqea, and Tayasir Grabens (Fig. 13). Then, during the late Miocene and the Pliocene, tectonic activity migrated northward to a deformation belt of ~6 km in the northern Faria Anticline (Fig. 13). Since the late Pleistocene, the deformation has been localized to a zone of ~1-2 km along the southern boundary of the Beit She'an Valley (Fig. 13). At this stage, both normal and sinistral faulting can be clearly identified (Figs. 6, 10, and 12). A major fault in this localized zone is the Wadi Malih Fault, which ruptured the late Pleistocene Lisan Formation (Figs. 3, 6, 8, and 13). It marks a clear lineament that crosses the Jordan Valley as it branches from the DSF (at the eastern side of the Jordan Valley), where a right-lateral step and a change in the strike of the DSF are observed near the uplifted (push-up) area of Tel Al-Qarn (e.g., Ferry et al., 2007; Figs. 2A and S1). This fault defines the current boundary between the Faria Anticline and Beit She'an Valley and shows clear evidence of normal faulting in the subsurface data (Fig. 12B). Farther NW of our study area, the observed post-late Miocene faults are connected to the Gilboa Fault (Fig. 1; Shaliv et al., 1991; Sneh et al., 1998; Dembo et al., 2015).

The localization of deformation near the Beit She'an Valley and the Gilboa Fault agrees with current GPS observations (Hamiel and Piatibra-tova, 2021). The results of this study highlight the significant deformation near the intersection of the DSF and the CGFS and suggest that most of the deformation in the eastern section of the CGFS occurs along the southern boundary of the Beit She'an Valley (i.e., near the Wadi Malih Fault and the Mehola Fault), and the current contribution to deformation of the Faria Fault is negligibly small. The localization of deformation near the Beit She'an Valley and the Gilboa Fault is also in agreement with paleomagnetic observations and mechanical models (Dembo et al., 2015). Dembo et al. (2015) showed localization of deformation near the Cannel and Gilboa Faults at sites younger than ca. 8 Ma. Previous studies suggest that dramatic changes in plate kinematics occurred in the Levant during the late Miocene—early Pliocene (e.g., Garfunkel, 1981; Joffe and Garfunkel, 1987; Marco, 2007). Such studies divided the deformation along the DSF and surrounding areas into two main phases: before and after —5 m.y. ago. At this transition stage, a shift in the location of the Sinai-Arabia Euler pole took place, leading to changes in the style and rate of deformation as well as the internal structure and localization of the DSF system (e.g., Garfunkel, 1981; Joffe and Garfunkel, 1987; Marco, 2007). Farther north of our study area, along the DSF and within the Sea of Galilee and Hula depressions, studies have shown that changes and reorganization of the main and marginal faults occurred at ca. 4-5 Ma (e.g., Heimann and Ron, 1993; Hurwitz et al., 2002; Schattner and Weinberger, 2008; Heimann et al., 2009; Matmon and Zilberman, 2017). Other studies suggest that major changes in the regional plate tectonics and the structure of the DSF occurred at ca. 10 Ma. Around this time, the collision of Arabian and Eurasian plates along the Bitlis suture initialized in SE Anatolia (e.g., McQuarrie and van Hinsbergen, 2013), the DSF was fully developed as a plate boundary (e.g., Gomez et al., 2020), and the tectonic activity along the intraplate Sinai—Negev Shear Zone terminated (Weinberger et al., 2020). Similar to these studies, in our study area, a major transition in tectonic deformation was found to have started at ca. 10 Ma, when deposition of the Bira Formation first began. Spatial analysis of the new geological map (Figs. 3, 8, and S2) shows the migration of post-late Miocene faulting from SW to NE, which deepened the Beit She'an Valley toward the NE. The volcanic activity in the SE sector of our study area (Fig. 3), which is dated to 5.65-5.90 Ma (Shaliv et al., 1991; Dembo et al., 2015), is probably related to the same tectonic phase that started during the late Miocene and continued until the Pliocene. The strike of most post-late Miocene faults is in the NW direction, some are in the NNW direction, and a small minority of faults are in the NE direction. Many of the NW faults are probably rejuvenated faults that initiated during the early—middle Miocene tectonic phase or even before, as indicated by the early stages of faulting along the CGFS (e.g., Shaliv et al., 1991; Segev et al., 2014). Finally, we propose that the northward migration and localization of the intersection between the CGFS and the DSF that has occurred since the late Miocene is probably related to the above-mentioned regional tectonic and stress field changes, the transition in the Sinai-Arabia Euler pole location, and the localization of deformation along the DSF. We also propose that our observations, along with current GPS measurements (e.g., Gomez et al., 2020; Hamiel and Piatibratova, 2021), suggest the fragmentation of the Sinai plate since the late Miocene.

Description Image Source
Fig. 1A - Hamiel et al. (2022)
Fig. 1B - Hamiel et al. (2022)
Fig. 1C - Hamiel et al. (2022)
Fig. 2A - Hamiel et al. (2022)
Fig. 2B - Hamiel et al. (2022)
Fig. 2C - Hamiel et al. (2022)
Fig. 2D - Hamiel et al. (2022)
Fig. 3 - Hamiel et al. (2022)
Fig. 4 A-F - Hamiel et al. (2022)
Fig. 4 G-H - Hamiel et al. (2022)
Fig. 5 - Hamiel et al. (2022)
Fig. 6 - Hamiel et al. (2022)
Fig. 7 - Hamiel et al. (2022)
Fig. 8 - Hamiel et al. (2022)
Fig. 9 - Hamiel et al. (2022)
Fig. 10 - Hamiel et al. (2022)
Fig. 11 - Hamiel et al. (2022)
Fig. 12A - Hamiel et al. (2022)
Fig. 12B-D - Hamiel et al. (2022)
Fig. 12C-E - Hamiel et al. (2022)
Fig. 13 A - Hamiel et al. (2022)
Fig. 13 B - Hamiel et al. (2022)

The Crustal Structure of the Dead Sea Transform - Weber et al (2022)

Abstract

To address one of the central questions of plate tectonics — How do large transform systems work and what are their typical features? — seismic investigations across the Dead Sea Transform (DST), the boundary between the African and Arabian plates in the Middle East, were conducted for the first time. A major component of these investigations was a combined reflection/refraction survey across the territories of Palestine, Israel and Jordan. The main results of this study are:

  1. The seismic basement is offset by 3–5 km under the DST
  2. The DST cuts through the entire crust, broadening in the lower crust
  3. Strong lower crustal reflectors are imaged only on one side of the DST
  4. The seismic velocity sections show a steady increase in the depth of the crust-mantle transition (Moho) from ~26 km at the Mediterranean to ~39 km under the Jordan highlands, with only a small but visible, asymmetric topography of the Moho under the DST.
These observations can be linked to the left-lateral movement of 105 km of the two plates in the last 17 Myr, accompanied by strong deformation within a narrow zone cutting through the entire crust. Comparing the DST and the San Andreas Fault (SAF) system, a strong asymmetry in subhorizontal lower crustal reflectors and a deep reaching deformation zone both occur around the DST and the SAF. The fact that such lower crustal reflectors and deep deformation zones are observed in such different transform systems suggests that these structures are possibly fundamental features of large transform plate boundaries

Tectonic setting and geology along the profile

The DST is a system of left-lateral strike-slip faults that accommodate the relative motion between the African and Arabian plates. Except for a mild compressional deformation starting about 80 Ma, the larger Dead Sea region has remained a stable platform since the early Mesozoic. Approximately 17 Ma, this tectonic stability was interrupted by the formation of a transform, the DST, with a total left-lateral displacement of 105 km until today (Quennell 1958; Freund et al. 1970; Garfunkel 1981).

The crystalline basement of the area represents the NW part of the Arabo-Nubian Shield (ANS) and consists of mainly juvenile Late Proterozoic rocks (Stoeser & Camp 1985; Stern 1994). Regarding both isotopic and chemical data of xenoliths, the involvement of older crustal material in the lower crust of the ANS seems improbable (Henjes-Kunst et al. 1990; Stern 1994; Ibrahim & McCourt 1995). There is a gradual transition from the continental crust of the ANS with thicknesses of 35–40 km (El-Isa et al. 1987; Makris et al. 1983; Al-Zoubi & Ben-Avraham 2002) to the crust of the eastern Mediterranean, that is assumed to be partly underlain by typical oceanic crust with thicknesses smaller than 10 km (Ginzburg et al. 1979a; Makris et al. 1983; Ben-Avraham et al. 2002), see also Fig. 1(b).

The Precambrian basement is usually overlain by an Infracambrian to Early Cambrian volcano-sedimentary succession of variable thickness. Whereas coarse-grained clastics (Saramuj and Elat conglomerates) are restricted to fault-bounded basins, fine-grained clastics, mostly consisting of arkosic sandstones and associated volcanic rocks (Zenifim Formation, Haiyala Volcaniclastic Unit and equivalent rock units) have been observed in large parts of the Israel and Jordan subsurface (Weissbrod & Sneh 2002). In boreholes close to the WRR and NVR profiles (Fig. 1b) the Zenifim Formation was determined to be several hundreds of metres thick (> 500 m in Maktesh-Qatan, MQ, and >2000 m in Ramon-1, R1), although its base has not been encountered in any of the Israeli boreholes.

The position of the study region at the NW edge of the ANS, i.e. at a passive continental margin since early Paleozoic times, is reflected in facies changes and varying sedimentary thicknesses along the seismic profiles (Fig. 2). The Phanerozoic along the northwestern part of the profile is dominated by Cretaceous and Tertiary rocks underlain by Jurassic, Triassic and Permian sequences that thin out towards the east. East of the DST, however, Permian to Triassic strata are missing and Lower Cretaceous rocks unconformably overlie Ordovician and Cambrian sandstones. The crystalline basement rocks (calc-alkaline granitoids and rhyolites) cropping out in the Jebel Humrat Fiddan area east of the DST are thought to be equivalent to the basement rocks of the Timna region in Israel, near Elat.

Previous seismic work in this area

Previous crustal-scale wide-angle reflection/refraction experiments in the study area (Fig. 1b) include that in Israel in 1977 (Ginzburg et al. 1979a), the onshore-offshore experiment between the northwestern end of the WRR profile and Cyprus in 1978 (Makris et al. 1983; Ben-Avraham et al. 2002) and that in Jordan in 1984 (El-Isa et al. 1987). However, there was no profile, which crossed the DST to provide a complete image across this structure. Moreover, deep seismic reflection data only exist from the area between the Dead Sea and the Mediterranean (Yuval & Rotstein 1987; Rotstein et al. 1987), and high-resolution seismics (Frieslander 2000) are mostly limited to the western part of the Arava Valley. Further south within the Afro-Arabian rift system seismic profiles, which cross the rift structures, have proved to be very useful to understand the crustal structure as, for example, the E-W profiles crossing the Kenya rift (Maguire et al. 1994; Braile et al. 1994; Birt et al. 1997). Seismic crustal investigations of the Afro-Arabian Rift system are summarized by Mechie & Prodehl (1988) and Prodehl et al. (1997).

Discussion

What is the crustal structure at the DST in the Arava Valley?

In deep crustal reflection data the Moho is commonly defined as the break-off of lower crustal reflectivity (Fig. 11b). The increase in Moho depth from ∼30 to 38 km, that is observed in the WRR data beneath the NVR profile, is more or less in accordance with the NVR data. A few small discrepancies between the reflection and refraction seismic data (as for example at the western end of the NVR line) are not unusual for coincident seismic reflection/refraction surveys (see e.g. Mooney & Brocher 1987). Note also that reflections from Moho depths typically have dominant frequencies of 12–13 Hz in the NVR data, in contrast to 6–7 Hz in the WRR data. At the same time the differences between the reflection and refraction Moho are in the range of measurement uncertainties (see also WRR modelling, Section 3.3).

In comparing the depths to the seismic basement with those obtained from previous experiments in the area, the depths obtained in this study west of the Arava Valley (Figs 10a and 11a) are smaller than those obtained by Ginzburg et al. (1979a). This may be due in part to somewhat lower average velocities used by Ginzburg et al. (1979a). On the other hand the depths to the seismic basement obtained here west of the Arava Valley are in good agreement with those obtained by Perathoner (1979). The larger value for the depth to the seismic basement beneath the coast obtained by Makris et al. (1983) from the offshore-onshore experiment in 1978 probably indicates that the depth to the seismic basement increases towards the coast to the NW of the region of ray coverage for the model obtained here. Beneath the Arava Valley itself, the depth to the seismic basement obtained here is greater than those obtained by Ginzburg et al. (1979a) and Perathoner (1979) from the N–S profile within the Arava Valley (Fig. 1b). This is probably due to the fact that although, for the N–S profile, the shot was in the Dead Sea, the recording stations were at the western side of the Arava Valley and sometimes on the shoulders of the Arava Valley. To the east of the Arava Valley the depth to the seismic basement obtained here is within 1 km of the depth obtained by El-Isa et al. (1987) from the N–S profile on the eastern shoulder of the Arava Valley (Fig. 1b). The depths to the boundary between the upper and lower crust obtained in this study (Figs 10b and 11a) agree to within 1 km with the depths to this boundary obtained from the previous experiments in this area (Fig. 1b; Ginzburg et al. 1979a; El-Isa et al. 1987).

With respect to the depths to the Moho, west of the Arava Valley there is agreement to within 3 km between the depths obtained here (Fig. 10c) and those obtained by Ginzburg et al. (1979a) and Makris et al. (1983). Beneath the Arava Valley itself the Moho depth obtained here agrees with that obtained by Perathoner (1979) from the N–S profile within the Arava Valley, but it is 5 km deeper than the value obtained by Ginzburg et al. (1979a). To the east of the Arava Valley the Moho depths obtained here are also about 5 km deeper than those obtained by El-Isa et al. (1987) from the N–S profiles on the eastern shoulder of the Arava Valley. In fact, the Moho depths obtained by El-Isa et al. (1987) from the N–S profiles on the eastern shoulder of the Arava Valley (Fig. 10c) are more in agreement with the depth at which strong reflections are observed in the NVR section (Fig. 11b). In order to find out where PmP reflections would occur in the record sections assuming that the Moho depths are as indicated by either the near-vertical incidence reflection section or by El-Isa et al. (1987), a model was constructed with Moho depths taken from the near-vertical incidence reflection section and El-Isa et al. (1987) and an average lower crustal velocity of 6.7 km s−1. Tracing rays through this model results in arrival times for the PmP reflection which are 0.8–1.0 s earlier than those for the preferred model shown here (Fig. 11a) in the record sections from the densely spaced data from the Arava Valley from shots 9 and 10 (Figs 8b and 9b). In these data there are no strong reflections 0.8– 1.0 s earlier than the PmP reflections shown. For this reason and as the data from the previous experiments in the area are sparser than the data along the WRR profile, it is thought that the Moho depths obtained here are more accurate than those obtained by El-Isa et al. (1987). Alternatively, the discrepancy may be, at least in part, due to strong 3-D variations of the Moho in the vicinity of the DESERT profile. Such variations are evident in the receiver-function data (A. Mohsend, personal communication, 2003).

A remarkable feature in Fig. 11(b) is a zone of high reflectivity at a depth of 28 km below the Jordan highlands between profile km 78 and 92. This is about the depth and position along the profile from which the possible phase, Pi2P in the WRR data is reflected, although the boundary associated with the Pi2P phase, if present, would have to occur for at least about 90 km under the profile (Fig. 10d) instead of just about 20 km as identified in the NVR data. This zone of high reflectivity in the NVR data might be due to a lithological contrast caused by underplating. In this case higher velocities of about 7.0 km s−1 would be expected to occur across the region of high reflectivity. If a 7.0 km s−1 layer of limited thickness is, in fact, present at about 30 km depth, this would only have a small effect on the estimated Moho depths, such that they would still be within the error limits of ±3 km given above (Section 3.3). Strong magmatic activity that occurred in the region both in Late Precambrian/Early Cambrian, Cretaceous and Neogene times could possibly have caused such a proposed underplating, or sill-like intrusions into the lower crust. Another possibility for the creation of such a high reflectivity is a zone of localized strain close to the base of the crust, which is in agreement with the conclusions of Sobolev (unpublished data) (see also later discussion) which show a zone of high shear deformation and possible lower crustal flow east of the transform. It is expected that a gravity analysis currently being carried out along the profile might give some further constraints for interpreting these reflectors. The high reflectivity at Moho depth west of the AF is in accordance with the eastern part of a deep seismic reflection line between the Mediterranean and the Dead Sea (Yuval & Rotstein 1987; Rotstein et al. 1987) that shows a similar reflectivity pattern of the crust as observed here.

Does the DST cut through the whole crust?—Yes

Imaging near-vertical structures by near-vertical seismic reflection techniques is difficult (e.g. Meissner 1996). It is, however, possible to get indirect evidence of the depth continuation of steeply dipping faults by the offset of crustal reflectors or an observed change in crustal reflectivity.

However, although the DST/AF is clearly recognized on satellite data as a rather straight line between the Red Sea and the Dead Sea (DESERT Team 2000) it cannot unambiguously be delineated in the Common Depth Point (CDP)section (Figs 11b and A5). There is no pronounced difference in crustal structure west and east of the Arava Fault and in the immediate vicinity of its surface trace sedimentary reflections are missing. The absence of sedimentary reflectors might be due to strong deformation of the rocks close to the fault, but could also be caused by the absorption of high frequencies (Fig. A2) in an area covered by sand dunes and alluvium.

Whereas a possible Moho offset has been proposed for the San Andreas Fault in northern California from deep crustal seismic reflection and refraction studies (Henstock et al. 1997), and for the DST north of the Dead Sea Basin from the analysis of gravity data (Ten Brink et al. 1990), there is no evidence for such an offset at the AF along the NVR profile. Nonetheless it is inferred that the AF reaches down to the mantle, changing into a broader deformation zone at mid-crustal level, because of the following reasons:

  1. At Moho depth an ~15 km wide zone (profile km 54 to 70 in Figs 11b and A5) beneath the surface trace of the AF is observed that lacks the strong reflectors observed farther to the west and the lower crustal reflectors observed to the east. From this it is derived that the fault zone becomes broader in the lower crust. The strong reflections beneath the AF at about 18 km depth are thought to occur at the upper/lower crust boundary and are linked to a velocity jump from 6.4 to 6.7 km s−1 (Fig. 11a). There is no good expression of this boundary elsewhere in the NVR data. This is taken as an indication for a zone of localized shear strain between the felsic upper and mafic lower crust related to the transform motion along the AF. Whereas Furlong et al. (1989) and Brocher et al. (1994) interpreted a similar subhorizontal surface at 15–20 km depth below the San Andreas Fault Zone in the San Francisco Bay area as a possible detachment zone of the San Andreas Fault, linking it to the Hayward/Calaveras fault system, Holbrook et al. (1996) interpreted it as corresponding to the top of the lower crust, acting as a whole to accommodate shear deformation in a broad zone. Here we propose a similar model for the DST as Holbrook et al. (1996).

  2. The small but visible, asymmetric topography of the Moho below the Arava Valley in the WRR model (profile km 130 to 170 in Fig. 11a) is also consistent with the NVR data. This is another piece of evidence for the AF cutting through the whole crust. Whereas a bending down of the Moho, or a ‘Moho keel’ has been put forward for some Paleozoic strike-slip regions in transpressional tectonic regimes (McBride 1994; Stern & McBride 1998), this coupled upward-downward structure of the Moho might be due to the transtensional character of the DST between the Red Sea and the Dead Sea (see also later discussion).

  3. Reflectors in the lowermost crust (25–32 km depth from profile km 55 to 70 in Fig. 11b) that dip away from the suspected fault zone and that are most pronounced east of the AF, might correspond to anisotropic fabrics developing along mylonitic shear zones, similar to the dipping reflectors beneath the Walls Boundary strike-slip fault in the northern British Caledonides (McBride 1994).
Our conclusion from these seismic data, namely that the AF reaches down to the mantle and changes into a broader deformation zone at mid-crustal level, agrees also with the results of the thermo-mechanical modelling of the DST (Sobolev, unpublished data). These results show that shear-deformation focuses in one or two major faults in a 20–40 km wide region in the upper crust with minimal strength, and that a broad mechanically weak decoupling zone extends vertically from the lower crust into the asthenosphere. Further evidence for such an extension of this decoupling zone in the asthenosphere comes from teleseismic SKS-splitting observations along the NVR profile (Rumpker et al. 2003). This deep reaching boundary layer is thought to accommodate the transform motion between the African and the Arabian plates (Fig. 1).

Does rifting/extension play an important role in the dynamics of the DST?—No

Some features in the near-surface structure of the Arava Valley, e.g. surface topographic expression, sedimentary fill and normal faults at the edge of the valley, resemble those of rift structures. However, the narrow, only ~10 km wide, shallow sedimentary basin mainly to the west of the AF (Fig. 2b), a seismic basement offset of 3 to 5 km on the eastern side of the Arava Valley and the small but visible, asymmetric Moho topography (~1.5 km) with a coupled upwarp-downwarp structure beneath the Arava Valley (Fig. 11), although possibly related to the slight extension across this part of the DST, are untypical for rift structures. For example, the southern portion of the Kenya rift, a classical continental rift, has been under extension since about 10 Ma, and the Moho there is uplifted 5–10 km causing considerable crustal thinning (Mechie et al. 1997). We therefore conclude that rifting–type deformation (fault perpendicular extension) did not play a dominant role in shaping the crustal structure of the DST. A thermo-mechanical model of the DST by Sobolev (unpublished data) confirms this by showing that the crustal structure of the DST results mainly from the geologically documented 105 km left-lateral displacement between the Arabian plate and the African plate (Figs 13a and b) placing lithospheric blocks with different crustal structures opposite each other. The modelling also supports the scenario that changes in surface and Moho topography and in crustal structure result from large, localized deformation accommodating the transform motion within a narrow zone crossing the entire crust. However, this process is combined with less than 4 km of fault-perpendicular extension (Garfunkel 1981; Sobolev et al. 2003 Fig. 13c). Thus the ‘rift component’ at the DST between the Dead Sea and the Gulf of Aqaba, defined as the ratio between fault perpendicular extension [4 km] and strike slip motion [105 km], is probably smaller than 4 per cent. This small extension nevertheless produces a large topography because the extension is localized within the narrow (20 km wide) upper mantle and lower crustal shear zones, where viscosity is reduced due to the high strain rate produced by the strike-slip motion (Sobolev et al. 2003), thus giving the Arava Valley the appearance of a rift valley.

Are there structural similarities between the DST and the SAF system despite their different geological history and setting?
—Yes and No

A comparison with the San Andreas Fault (SAF), another end-member of transform structures (see e.g. Holbrook et al. 1996; Bonner & Blackwell 1998), shows several differences, especially in the shallow structure, and some similarities in the deeper structure. Fault Zone Guided Waves from controlled source experiments in the Arava Valley (Haberland et al. 2003) sample the top few hundred meters of the Arava Fault and are best explained by a fault model with a narrow, only 3–10 m wide low-velocity zone. This thickness is much smaller than the typical width of 100 to 170 m of low-velocity zones in the SAF system (Li et al. 1990), and is possibly due to the smaller total slip on the DST (105 km) versus the slip of more than 350 km at the SAF, or it could be a local feature controlled by the young sediments in the area where the DESERT profile crosses the AF. In contrast to the SAF the Arava Fault under the DESERT profile acts as a localized fluid barrier separating a high- from a low-velocity block in the uppermost crust. This contrast is visible in the combined magnetotelluric sounding and high-resolution tomography of Ritter et al. (2003). Such a feature is remarkably different from active segments of the SAF, which typically show a conductive fault core acting as a fluid conduit (Unsworth et al. 2000).

If, however, deeper crustal and mantle structures are compared, it becomes apparent that both transform systems show deep reaching deformation zones (Sobolev, unpublished data, Rumpker et al. 2003 for the DST and Holbrook et al. 1996; Henstock et al. 1997; Unsworth et al. 1997; Silver 1996 for the SAF) accompanied by a strong asymmetry in subhorizontal lower crustal reflectors (lower crustal flow, sill-like intrusions?). We therefore suggest that these features are common features of continental transform plate boundaries.

Conclusions

Our study provides the first whole-crustal image across the Dead Sea Transform (DST), one of the Earth’s major transform faults. Under the Arava Fault (Fig. 11), the main fault of the southern DST system, the seismic basement is offset by several kilometres, but the Moho depth increases steadily from ~26 km at the Mediterranean to ~39 km under the Jordan highlands, except for a small but visible, asymmetric topography under the Arava Valley. The general trend of continuous Moho-depth increase is confirmed by the interpretation of potential field data (Al-Zoubi & Ben-Avraham 2002) and the results of a receiver functions study (A. Mohsen, personal communication, 2003). Based on the interpretation of the NVR data, we infer that the AF cuts through the crust, becoming a broad zone in the lower crust, and reaches down to the mantle. This agrees with the results of the thermo-mechanical modelling of Sobolev et al. (2003) and the SKS-splitting observations by Rumpker et al. (2003), which suggest that the DST cuts through the whole lithosphere, thus accommodating the motion between the African and the Arabian plates (Fig. 1). The lack of significant uplift of the Moho under the Arava Valley speaks strongly against a potential rift structure in this area. We therefore conclude that rift-type deformation (fault perpendicular extension) did not play a dominant role in the dynamics of the DST, a fact corroborated again by the results of Sobolev (unpublished data). Although the shallow structure at the DST differs significantly from the structure at the San Andreas Fault, the deep reaching deformation zones accompanied by a strong asymmetry in subhorizontal lower crustal reflectors appear to be similar for both fault zone systems. We therefore suggest that these deep features are common for large continental transform plate boundaries

Images

Normal Size
Description Image Source
Fig. 1a - Weber et al. (2004)
Fig. 1b - Weber et al. (2004)
Fig. 2 - Weber et al. (2004)
Fig. 11 - Weber et al. (2004)
Fig. 13 - Weber et al. (2004)
Magnified
Description Image Source
Fig. 1a - Weber et al. (2004)
Fig. 1b - Weber et al. (2004)
Fig. 2 - Weber et al. (2004)
Fig. 11 - Weber et al. (2004)
Fig. 13 - Weber et al. (2004)

Resolving the slip-rate inconsistency of the northern Dead Sea fault - Li et al (2024)

Estimated Geodetic Slip Rates

Fig. 2

Fault- parallel velocities and estimated geodetic slip rates.

(A to D) Profiles across the dSF at different locations showing the fault- parallel velocity (into direction N12°E) from stacking several BOI profiles (colored circles) in comparison with GnSS velocities (gray); see Fig. 1B for locations and fig. S2 for further profile results.

(E) estimated geodetic slip rates and locking depths along the dSF (bold squares, 1- sigma confidence limits; see table S1) compared with other geodetic estimates (other symbols) and the predicted Sinai- Arabia plate motion prediction (49) (dashed line). For display purposes, some of the symbols have been slightly shifted vertically from their exact locations of estimation.

Li et al. (2024)

Seismicity and proposed tectonic block configuration in the eastern Mediterranean

Fig. 4

Seismicity and proposed tectonic block configuration in the eastern Mediterranean.

  1. Seismicity (from eMSc) showing elevated offshore seismicity between Lebanon/israel and the cyprus arc. this diffuse seismicity zone appears to bound the Latakia-tartus microplate to the southwest, with the cyprus arc bounding it to the west, hatay triple junction in the north, and the northern dSF in the east.
  2. Green vectors are GnSS velocities as in Fig. 1c, with respect to stable Arabia, and the blue dashed line in (A) marks the block boundary proposed by Gomez et al. (36).

Li et al. (2024)

Notes

Reported fault slip rates, a key quantity for earthquake hazard and risk analyses, have been inconsistent for the northern Dead Sea fault (DSF). Studies of offset geological and archeological structures suggest a slip rate of 4 to 6 millimeters per year, consistent with the southern DSF, whereas geodetic slip- rate estimates are only 2 to 3 millimeters per year.

... The Jordan Valley slip rate decreases from the southern part, just north of the Dead Sea, to the northern part, south of the Sea of Galilee, from 4.3 to 3.5 mm/year (fig. S3). This suggests that a part of the slip rate is transferred from the DSF to the Carmel- Gilboa fault, which has been previously reported in GNSS studies estimating its slip rate in the range of 0.8 to 1.2 mm/year (14, 36). Within the restraining bend, the BOI deformation is not clearly associated with a single fault and we thus presume that the Yammouneh fault accounts for most of the observed deformation. Within uncertainties, the result is consistent with the reported block model rate of 3.8 mm/year (26) and with the lower bound of the Late Pleistocene/Holocene slip rate of 3.8 to 6.4 mm/year (24).

...Although most estimates of geological slip rates on the northern DSF are in the range of ~4 to 6 mm/year, there is considerable variability in the reported slip rates. A relatively low slip rate of 1.4 to 4.5 mm/year was reported by Searle et al. (34), a result criticized later by Westaway (50) because of inaccurate dating of lava flow units offset by the fault and because of mistaken alignment of piercing points. A faster slip rate of 4.9 to 6.0 mm/year was estimated in the Amik basin based on offset alluvial fans and three archeological sites (29, 31). These rates, however, are not representative of the slip rate on the northern DSF, as the Amik basin is within the Hatay triple junction and thus influenced by the East Anatolian Fault and on- land extensions of faults related to the Cyprus arc. The main geologic slip- rate results of the northern DSF come from the Misyaf segment of the fault, where paleoseismic trenching and archaeoseismic studies of the faulted Al- Harif Roman aqueduct in Syria by four earthquakes during the past 3500 years, each having an estimated offset of ~4 m, yield a slip rate of 4.6 mm/year, or even higher in the range of 4.9 to 6.3 mm/year (28, 33), depending on how the slip rate is calculated.

We postulate that the high estimated geological slip rate for the Misyaf segment could be elevated due to earthquake clustering during the past 3500 years. Earthquake clustering is widely reported in paleoseismological studies for strike- slip faults where long enough records have been retrieved (4, 5, 51, 52).

... The question remains about where the ~2 mm/year difference between the low geodetic slip rate on the northern DSF and the predicted Sinai- Arabia plate motion is accommodated. We explored whether the low geodetic rate could be related to interactions with the offshore Cyprus arc or due to viscoelastic earthquake cycle effects, but neither possibility is likely (see fig. S5). Crustal shortening (1 mm/year) within the Arabian plate over the Palmyrides fold and thrust belt has been suggested (35), along with north- south extension (1.5 mm/year) just north of the Lebanese restraining bend. T here is, however, scant evidence to support this hypothesis in the GNSS or BOI data (Figs. 1B and 2A). A north- south block boundary just offshore the Syrian coastal ranges has also been proposed in GNSS block modeling (36), representing strike-slip motion parallel to the northern DSF (blue dashed line in Fig. 4A). However, results of extensive offshore seismic imaging show no evidence for a strike- slip fault at this location (54).

Given the above, it appears that the northern Sinai plate moves separately from the main Sinai plate to the south. The Carmel- Gilboa and Roum faults splay off to the northwest from the DSF and into the Mediterranean, but no clear indication has been found for an offshore plate boundary extending from the coast to the Cyprus arc. Maps of the regional seismicity show distributed seismicity in the area from the coast of northern Israel and Lebanon extending to the Cyprus arc that is markedly stronger than the neighboring offshore seismicity north of Egypt and west of Syria (Fig. 4A). This seismicity suggests a broad zone of diffuse deformation in the eastern Mediterranean, similar to what has been found for some other diffuse oceanic plate boundaries (55, 56). We estimated deviatoric moment tensor solutions for magnitude 3.3 to 5.0 events in this zone and they support (fig. S8 and table S3) that this zone accommodates the relative SSW- NNE extension between the main Sinai plate and its part west of the northern DSF.

Our results thus indicate that the northern Sinai plate is a separate microplate, here named the Latakia-Tartus microplate. It is bounded by the Cyprus arc to the west, the Hatay triple junction to the north, northern DSF to the east, and the Lebanon restraining bend and the offshore diffuse seismicity zone to the south. The GNSS velocities in southern Hatay, southwest of the Amik basin, and in northwestern Syria, clearly show a systematic pattern of velocities different from the velocities to the north (northern Hatay and Anatolia), east (Arabian plate), and south (Fig. 4A). Together, the results resolve the slip- rate inconsistency along the northern DSF and show that the seismic hazard associated with this fault is notably lower than if it were moving at 4 to 6 mm/year. Still, with the last major earthquake on the northern DSF in 1170 (32), a slip deficit amounting to about 2.4 m of fault slip has accumulated. While this might appear to indicate that the northern DSF is ripe for another major earthquake, the recent clustering of characteristic 4- m offset events suggests that it might not be the case.

Why do the largest continental earthquakes nucleate on branch faults?



References
References

Begin, Z. B., J. N. Louie, S. Marco, and Z. Ben-Avraham (2005). Prehistoric seismic basin effects in the Dead Sea pull-apart. Geol. Survey of Israel, Jerusalem: 31.

Ben-Avraham, Z., et al. (2006). "Segmentation of the Levant continental margin, eastern Mediterranean." Tectonics 25(5).

Ben-Avraham, Z., et al. (2008). "Geology and Evolution of the Southern Dead Sea Fault with Emphasis on Subsurface Structure." Annual Review of Earth and Planetary Sciences 36(1): 357-387.

Ben-Avraham, Z., et al. (2012). Structural styles along the Dead Sea Fault. in Regional Geology and Tectonics: Phanerozoic Passive Margins, Cratonic Basins and Global Tectonic Maps. 2: 616-633.

Garfunkel, Z., Ben-Avraham, Zvi and Kagan, Elisa (2014). Dead Sea Transform Fault System: Reviews, Springer Netherlands, Dordrecht.

Hamiel, Y., et al. (2022). "Migration and localization of faulting near the intersection of the Dead Sea Fault and the Carmel−Gilboa−Faria Fault System." GSA Bulletin.

Heidbach, O. and Z. Ben-Avraham (2007). "Stress evolution and seismic hazard of the Dead Sea Fault System." Earth and Planetary Science Letters 257(1): 299-312.

Li, X., et al. (2024). "Resolving the slip-rate inconsistency of the northern Dead Sea fault." Science Advances 10: eadj8408.

Pirazzoli, P. A. (2005). "A review of possible eustatic, isostatic and tectonic contributions in eight late-Holocene relative sea-level histories from the Mediterranean area." Quaternary Science Reviews 24(18): 1989-2001.

Weber, M. et al. (2004). "The crustal structure of the Dead Sea Transform." Geophysical Journal International 156(3): 655-681.