23
Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform Gal Hartman 1,2 , Tina M. Niemi 3 , Gideon Tibor 2,4 , Zvi Ben-Avraham 1 , Abdallah Al-Zoubi 5 , Yizhaq Makovsky 4 , Emad Akawwi 5 , Abdel-Rahman Abueladas 5 , and Rami Al-Ruzouq 5 1 Department of Geophysics and Planetary Sciences, Tel Aviv University, Tel Aviv, Israel, 2 Israel Oceanographic and Limnological Research, Haifa, Israel, 3 Department of Geosciences, University of Missouri at Kansas City, Kansas City, Missouri, USA, 4 Department of Marine Geosciences, University of Haifa, Haifa, Israel, 5 Department of Surveying and Geomatics, Al-BalqaApplied University, Salt, Jordan Abstract The northern Gulf of Elat/Aqaba is located in the transition between the deep marine basins of the gulf and the shallow onland basins of the Arava Valley. Interpretation of 500 km of high-resolution seismic reection data collected across the northern shelf reveals the tectonic structure and evolution of this transition. Six NNE-trending faults and one E-W trending transverse fault are mapped. Slip rates are calculated based on measured offsets and age determination based on a radiocarbon-calibrated sedimentation rate and a Quaternary age model. The most active fault is the Evrona Fault that absorbs most of the left lateral slip within the basin with an average sinistral slip rate of 0.7 ± 0.3 mm/yr through the Late Pleistocene and 2.33.4 mm/yr during the Holocene. Two intrabasin faults east of the Evrona Fault that have been inactive for the last several tens of thousands of years were mapped, and motion from these faults has likely transferred to the Evrona Fault. The basin is anked on the west by the Elat Fault and on the east by the Aqaba Fault. Both faults are marked by large bathymetric escarpments. Based on displaced seismic reectors, we calculate a Holocene vertical slip rate of 1.0 ± 0.2 and 0.4 ± 0.1 mm/yr for the Elat and Aqaba Faults, respectively. The geometry, slip rates, and slip history of the northern Gulf of Elat/Aqaba faults show that during the Late Pleistocene several intrabasin faults became dominant across the basin but that during the Holocene the Evrona Fault accommodates most of the strike slip. 1. Introduction The basins of the Gulf of Elat/Aqaba (GEA) (Figure 1) developed along the southern part of the Dead Sea Transform (DST). The ~1000 km plate boundary extends from the Red Sea in the south to the East Anatolian fault system at the collision zone between the Arabian and the Eurasia plates in the north [e.g., Quennell, 1984]. The DST separates the Arabian Plate from the Levantine Subplate and has been tectonically active since the Miocene with an accumulated sinistral offset of ~105 km [e.g., Quennell, 1959; Freund et al., 1968, 1970; Bartov et al., 1980; Garfunkel et al., 1981]. In the Early Pleistocene the compression stress eld in the region shifted from a N to NW strike [Gomez et al., 2007] inducing a southward increase of extension across the southern and central segments of the DST [Ben-Avraham et al., 2005, and references therein]. Seismic refraction data along the western margin of the GEA [Ginzburg et al., 1981] suggest that the crust thins from the north (~35 km) to the south (~27 km) indicating a southward increase in extension toward the Red Sea that likely controls the structural history of the gulf transform basins. The GEA is a long, narrow, and deep (~180 km × 20 km × 5 km) structural depression (Figure 2) composed of three major basins (Dakar, Aragonese, and Elat) formed between left stepping, left lateral strike-slip faults [Ben Avraham et al., 1979a, 1979b; Ben Avraham, 1985]. The northern basin, the Elat Basin, formed between the mainly strike-slip fault on the east and the predominantly normal faulting on the west [Ben Avraham, 1985, 1992]. The fault along the eastern side of the Elat Basin is the northern continuation of the Aragonese fault that partially ruptured in the 1995 Mw 7.3 Nuweiba earthquake [Dziewonski et al., 1997; Pinar and Türkelli, 1997; Klinger et al., 1999, 2000c; Hofstetter et al., 2003]. Ehrhardt et al. [2005] using multichannel seismic reection data suggested that the Aragonese fault bifurcates at the north end of the Elat Basin (Figure 3). The eastern branch bends to the right with localized compression and presumably continues northward to become the Aqaba Fault. The western branch extends diagonally north across the basin to become the Elat-Evrona fault zone. However, the seismic reection lines in Ehrhardt et al. [2005] are widely spaced and HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1 PUBLICATION S Journal of Geophysical Research: Solid Earth RESEARCH ARTICLE 10.1002/2013JB010879 Key Points: The study reveals six intrabasin north trending faults and one transverse fault The Evrona strike-slip fault occupies most of the Holocene left lateral slip The northern GEA basin is in an advanced stage of evolution Correspondence to: G. Hartman, [email protected] Citation: Hartman, G., T. M. Niemi, G. Tibor, Z. Ben-Avraham, A. Al-Zoubi, Y. Makovsky, E. Akawwi, A.-R. Abueladas, and R. Al-Ruzouq (2014), Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform, J. Geophys. Res. Solid Earth, 119, doi:10.1002/2013JB010879. Received 5 DEC 2013 Accepted 14 NOV 2014 Accepted article online 17 NOV 2014

Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

  • Upload
    rami

  • View
    215

  • Download
    1

Embed Size (px)

Citation preview

Page 1: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Quaternary tectonic evolution of the Northern Gulfof Elat/Aqaba along the Dead Sea TransformGal Hartman1,2, Tina M. Niemi3, Gideon Tibor2,4, Zvi Ben-Avraham1, Abdallah Al-Zoubi5,Yizhaq Makovsky4, Emad Akawwi5, Abdel-Rahman Abueladas5, and Rami Al-Ruzouq5

1Department of Geophysics and Planetary Sciences, Tel Aviv University, Tel Aviv, Israel, 2Israel Oceanographic andLimnological Research, Haifa, Israel, 3Department of Geosciences, University of Missouri at Kansas City, Kansas City,Missouri, USA, 4Department of Marine Geosciences, University of Haifa, Haifa, Israel, 5Department of Surveying andGeomatics, Al-Balqa‵ Applied University, Salt, Jordan

Abstract The northern Gulf of Elat/Aqaba is located in the transition between the deep marine basins ofthe gulf and the shallow onland basins of the Arava Valley. Interpretation of 500 km of high-resolution seismicreflection data collected across the northern shelf reveals the tectonic structure and evolution of thistransition. Six NNE-trending faults and one E-W trending transverse fault are mapped. Slip rates are calculatedbased onmeasured offsets and age determination based on a radiocarbon-calibrated sedimentation rate anda Quaternary age model. The most active fault is the Evrona Fault that absorbs most of the left lateral slipwithin the basin with an average sinistral slip rate of 0.7 ± 0.3mm/yr through the Late Pleistocene and2.3–3.4mm/yr during the Holocene. Two intrabasin faults east of the Evrona Fault that have been inactive forthe last several tens of thousands of years were mapped, and motion from these faults has likely transferredto the Evrona Fault. The basin is flanked on the west by the Elat Fault and on the east by the Aqaba Fault. Bothfaults are marked by large bathymetric escarpments. Based on displaced seismic reflectors, we calculate aHolocene vertical slip rate of 1.0 ± 0.2 and 0.4 ± 0.1mm/yr for the Elat and Aqaba Faults, respectively. Thegeometry, slip rates, and slip history of the northern Gulf of Elat/Aqaba faults show that during the LatePleistocene several intrabasin faults became dominant across the basin but that during the Holocene theEvrona Fault accommodates most of the strike slip.

1. Introduction

The basins of the Gulf of Elat/Aqaba (GEA) (Figure 1) developed along the southern part of the Dead SeaTransform (DST). The ~1000 km plate boundary extends from the Red Sea in the south to the East Anatolianfault system at the collision zone between the Arabian and the Eurasia plates in the north [e.g., Quennell, 1984].The DST separates the Arabian Plate from the Levantine Subplate and has been tectonically active sincethe Miocene with an accumulated sinistral offset of ~105 km [e.g., Quennell, 1959; Freund et al., 1968, 1970;Bartov et al., 1980; Garfunkel et al., 1981]. In the Early Pleistocene the compression stress field in the regionshifted from a N to NW strike [Gomez et al., 2007] inducing a southward increase of extension across thesouthern and central segments of the DST [Ben-Avraham et al., 2005, and references therein]. Seismic refractiondata along the western margin of the GEA [Ginzburg et al., 1981] suggest that the crust thins from the north(~35 km) to the south (~27 km) indicating a southward increase in extension toward the Red Sea that likelycontrols the structural history of the gulf transform basins.

The GEA is a long, narrow, and deep (~180 km × 20 km × 5 km) structural depression (Figure 2) composed ofthree major basins (Dakar, Aragonese, and Elat) formed between left stepping, left lateral strike-slip faults[Ben Avraham et al., 1979a, 1979b; Ben Avraham, 1985]. The northern basin, the Elat Basin, formed betweenthe mainly strike-slip fault on the east and the predominantly normal faulting on the west [Ben Avraham,1985, 1992]. The fault along the eastern side of the Elat Basin is the northern continuation of the Aragonesefault that partially ruptured in the 1995Mw 7.3 Nuweiba earthquake [Dziewonski et al., 1997; Pinar and Türkelli,1997; Klinger et al., 1999, 2000c; Hofstetter et al., 2003]. Ehrhardt et al. [2005] using multichannel seismicreflection data suggested that the Aragonese fault bifurcates at the north end of the Elat Basin (Figure 3).The eastern branch bends to the right with localized compression and presumably continues northwardto become the Aqaba Fault. The western branch extends diagonally north across the basin to become theElat-Evrona fault zone. However, the seismic reflection lines in Ehrhardt et al. [2005] are widely spaced and

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 1

PUBLICATIONSJournal of Geophysical Research: Solid Earth

RESEARCH ARTICLE10.1002/2013JB010879

Key Points:• The study reveals six intrabasinnorth trending faults and onetransverse fault

• The Evrona strike-slip fault occupiesmost of the Holocene left lateral slip

• The northern GEA basin is in anadvanced stage of evolution

Correspondence to:G. Hartman,[email protected]

Citation:Hartman, G., T. M. Niemi, G. Tibor,Z. Ben-Avraham, A. Al-Zoubi, Y. Makovsky,E. Akawwi, A.-R. Abueladas, andR. Al-Ruzouq (2014), Quaternary tectonicevolution of the Northern Gulf ofElat/Aqaba along the Dead SeaTransform, J. Geophys. Res. Solid Earth, 119,doi:10.1002/2013JB010879.

Received 5 DEC 2013Accepted 14 NOV 2014Accepted article online 17 NOV 2014

Page 2: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

have relatively poor resolution. Detailsof how slip is transferred across thenorthern gulf where it narrows toapproximately 6 km wide, its evolution,and transitions to the onshore EvronaBasin are still unclear.

In this study we interpret new high-resolution seismic reflection datacollected across the entire northern GEAshelf in both Israeli and Jordanian waters.This study is the first to image thestructure of the Dead Sea Transform onthe east side of the gulf and to collectdata continuously across the entire headof the gulf. Data from this study providesome of the first seismic reflection datato address deformation styles and ratesin the transition from the deep marineElat Basin to the subaerial basins ofthe Arava Valley along the northerntermination of the GEA. Models for howdeformation is transferred between adeep (several kilometers of sediment)and a shallow (<1 km of sediment) basinalong a transform fault are still not wellunderstood. This paper analyzes thegeometry of newly identified faults onthe shelf of the GEA. Furthermore, weexamine the slip rates of faults and therole of intrabasin faults in the partitioningof slip across the basin. The northern GEAshelf is uniquely positioned to provideinsight into the evolution of lateQuaternary faulting and processes oftransform basin formation.

2. Previous Work

The Gulf of Elat/Aqaba (GEA) has been mapped as a series of pull-apart basins [Ben Avraham et al., 1979a,1979b]. The seafloor morphology of the northern GEA was investigated in the geophysical studies of Recheset al. [1987], Ben Avraham and Tibor [1993], Ehrhardt et al. [2005],Makovsky et al. [2008], and Tibor et al. [2010].The width of the western shelf of the northern GEA varies between 400 and 800m with slopes rangingfrom 6° to >13°. The eastern shelf of the northern GEA is narrower (<400m) with slopes greater than13° (Figure 3), or is absent where a steep (>45°) slope continues from the shore. Morphological analyses[Tibor et al., 2010] show that the head of the gule is subdivided into three domains: the Elat and the AqabaSubbasins that are separated by the north-south trending Ayla High.

The shelf edge of the northern GEA is characterized by a prominent set of submarine terraces [Reches et al.,1987; Makovsky et al., 2008; Tibor et al., 2010]. These terraces mark the southern boundary of the northernshelf of the GEA. Its depth varies from 70 to 80 meters below sea level (mbsl) along the Aqaba subbasin to~100 mbsl along the Ayla High and 110–120 mbsl along the Elat subbasin (Figure 3). In the Aqaba Subbasinthe ~1300m wide northern shelf dips at ~4°. To the west, toward the Elat subbasin, the width of theshelf increases to ~1800m and its slope slightly decreases to ~3°. Tibor et al. [2010] suggested that thesemorphological differences imply a varied pattern of modern sedimentation: the shelf along the AqabaSubbasin is currently eroding and the shelf-slope break is retreating.

Figure 1. Generalized tectonic settings of the Dead Sea fault or transform(DST) [modified after Le Béon et al., 2010, and references therein]. The DSTseparates the Arabian Plate from the Levantine Subplate. It is connectedto the Arabian-African plate boundary of the Red Sea to the south and tothe East Anatolian Fault to the north. The DST is seismically active, andearthquakes’ epicenters between the years 1969–2009 are marked.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 2

Page 3: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Previous studies have suggested severallocations for the main transform faultacross the northern GEA. Garfunkel [1970]traced a fault he named the Elat Faultalong the western Elat Sabkha and thewestern coast of the northern GEA. Byextrapolating the fault line, he speculatedthat the Elat Fault continues southwardand becomes submarine (Figure 3).Shaked et al. [2004, 2010] argued thatcatastrophic earthquake events causedsubmergence and burial of reefs alongthe western coastline of Elat. Theysuggested that slip along a segment ofthe western boundary normal fault(Figure 3) caused subsidence of 1.8m intwo seismic events in the past 5 ka. Recheset al. [1987] and Ben Avraham and Tibor[1993] observed a prominent change inthe slope angle at ~100m mbsl along thewestern shelf of the northern GEA. Theysuggested that the change in the slopeangle is shaped by the main transformfault that they also named the Elat Fault(Figure 3). Ben Avraham and Tibor [1993]suggested that the main transformcrosses from the easternmargin along thenorthern edge of the rhomb-shaped ElatBasin to the western margin of thenorthern GEA in an angular fashion(Figure 3). Ehrhardt et al. [2005] mappedthe main transform fault crossing thenorthern GEA from the eastern to thewestern margin in a curving path(Figure 3). Makovsky et al. [2008] mapped~30m of sinistral offset across a linearterrace using high-resolution bathymetryand subbottom profiles about 1 kmsoutheast of the northwestern shore ofthe GEA. They suggested that this terracewas a fringing reef that evolved duringthe last rise in sea level at ~11.5 ka,accommodating an average sinistral sliprate of 2.7± 1.5mm/yr. This slip rate ledthem to suggest that the fault connectsnorthward with the colinear Evrona Faultas themain segment of the DST (Figure 3).

North of the GEA, the DST passes along the Arava (Wadi ’Arabah) Valley. Mapping of offset surface features,aided by field observations, aerial photo analysis, and paleoseismic trenches [e.g., Zak and Freund, 1966;Garfunkel et al., 1981; Ginat et al., 1998; Klinger et al., 2000a, 2000b; Niemi et al., 2001; Le Béon et al., 2010, 2012]and recent gravimetric and magnetic studies [Haberland et al., 2007; ten Brink et al., 2007] revealed that theDST traverses the central Arava Valley predominantly as a single, almost continuous, sinistral strike-slip fault.Modeling of gravity and magnetic data [ten Brink et al., 1999, 2007], and seismic profiles [Frieslander, 2000]

Figure 2. (a) Generalized tectonic settings of the GEA (modified afterBen Avraham [1985]). The main transform fault segments as suggestedby Ben Avraham [1985] are marked with thick lines. The bathymetricdeeps of the GEA are circled and noted. The GEA contains three trans-form basins (marked by light blue), of which the Aragonese-Arnona Basinis bound by strike-slip faults as a true pull-apart basin (marked by darkerblue). Note that the main transform fault is marked on the eastern side ofthe asymmetric Elat Basin. The main transform fault north of the ElatBasin is mapped as suggested by Ben Avraham and Tibor [1993] andEhrhardt et al. [2005]. The Evrona and Timna basins are mapped fromgravimetric and magnetic data [ten Brink et al., 1999; Frieslander, 2000;ten Brink et al., 2007]. The red dashed box at the northern GEA outlines thestudy area. The red star marks the epicenter of the 1995 Mw 7.2 Nuweibaearthquake [Hofstetter et al., 2003]. (b) Schematic models of the deepsection of the basins (modified after Ben Avraham [1985] and ten Brink et al.[1999]). The marine basins of the GEA are deeper (3–5 km) than theon-land basins of the southern Arava Valley (<1 km).

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 3

Page 4: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

revealed separate longitudinal basins along the Arava Valley. Frieslander [2000] suggested that the southernArava Valley comprises two longitudinal basins, Evrona and Timna (Figure 2a). Gravity and magnetic studiessuggest that these shallow basins along the southern Arava Valley reach depths of 0.5–1.5 km [ten Brink et al.,1999, 2007], much shallower than the deep marine basins of the GEA (Figure 2b).

In the area of Evrona Sabkha, 15 to 25 km north of Elat, the Evrona Fault splays into a transtensionalcomplex [Garfunkel, 1981; Amit et al., 2002]. A set of comprehensive paleoseismic studies found evidencefor at least 15 earthquakes of magnitude M> 6 since the Late Pleistocene, partitioned over normal faultsthat border the basin and oblique-slip faults within the Evrona Sabkha [Gerson et al., 1993; Amit et al., 1995,1996; Enzel et al., 1996; Porat et al., 1996, 1997; Shtivelman et al., 1998; Amit et al., 1999, 2002; Zilbermanet al., 2005].

The Elat Sabkha (Figure 3), located onland along the northwest coast of the GEA, was partially mapped inprevious studies [e.g., Garfunkel, 1970; Frieslander, 2000]. Wachs and Zilberman [1994] suggested that theEvrona Fault zone continues southward crossing the Elat Sabkha around the hotels district. Rotstein et al.[1994] investigated seismic reflection data and interpreted a vertical deformation band that extendedseveral hundred meters wide along the eastern part of Elat hotels district. Frieslander [2000] interpreted adistinct subvertical discontinuity in the sediment in the same area. The seismic reflection profiles inFrieslander [2000] show a sedimentary basin that is more than 2 km deep beneath the city of Elat and theElat Sabkha.

Figure 3. Geological map of the study area overlaid with the outline of major faults suggested in previous studies(modified after Garfunkel [1970], Reches et al. [1987], Ben Avraham and Tibor [1993], Frieslander [2000], Niemi andSmith [1999], Slater and Niemi, [2003], Shaked et al. [2004], Ehrhardt et al. [2005], and Makovsky et al. [2008]).AF = Aragonese fault.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 4

Page 5: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

A prominent bathymetric escarpment bounds the eastern margin of the northern GEA and is the inferredlocation of the Aqaba Fault [Garfunkel, 1970]. The Aqaba Fault is shown on published maps following theeastern shoreline of the gulf [e.g., Quennell, 1959; Garfunkel, 1981; Ben Avraham, 1985]. Ehrhardt et al. [2005]suggested that most of the transform motion is transferred to the western margin of the northern GEA,and therefore, the Aqaba Fault carries little strike-slip motion. Based on analyses of aerial photos, Garfunkel[1970] suggested that the Aqaba fault continues onshore from the northeast corner of the GEA throughthe town of Aqaba and to the east into the Edom Mountains along the valley of Wadi Yutim (Figure 3).Niemi and Smith [1999] and Slater and Niemi [2003] constrained the location of the Aqaba fault betweenNW trending transverse faults (originally mapped by Garfunkel [1970]) and the alluvial fan of Wadi Yutim tothe east (Figure 3). Mapping the locations of faults on the shelf and quantifying their deformation should helpresolve how slip is transferred across the northern GEA basin.

3. Data Acquisition and Processing

This study is based on the analysis of data acquired in the northern GEA during two high-resolution marinegeophysical surveys. The first survey (MERC I) was conducted from 29 October to 16 November 2006 andincluded the acquisition of 261 seismic lines with a total length of 426 km (Figure 4). The survey was carriedout in two phases: deep survey (water depths of 10–700m) onboard the R/V Etziona and a shallow phase(water depths of less than 100m) onboard a fishing vessel called Danny-Boy. The second survey (MERC II)included 47 additional seismic lines (~70 km) collected during 2–8 February 2010 onboard the R/VMediterranean Explorer.

The high-resolution geophysical survey used a GeoResources © GeoSparker200 seismic source generated byelectrical spark at energies of 300–1000 J and seismic frequencies of 200–3000Hz. During the deep survey,the system used two hydrophone arrays (channels), one towed parallel to the sparker, and the otheroffset about 50m attempting to receive more seismic energy across steep slopes. In the shallow survey onlyone hydrophone array was towed parallel to the sparker. A seismic penetration of more than 100m wasachieved in the soft sediments, with a vertical resolution of less than 1m.

During the second marine seismic survey, a denser grid (100m spacing) of seismic profiles of the northernshelf of the GEA (Figure 4) was acquired using the GeoSparker seismic source with an electrical energy outputof 500 J and frequency range between 200 and 3000Hz. These data were recorded with a 24 bit digital

Figure 4. Location map of the high-resolution seismic reflection lines of the 2010 survey (marked by dark gray) and the 2006survey (marked by light gray) along the northern shelf of the Gulf of Elat/Aqaba. The location of core H02 discussed in the textis shown.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 5

Page 6: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

minitrace system. Data gaps in the previous 2006 data set were filled in during this survey. During our secondsurvey, we were permitted to collect data continuously from east to west across the head of the gulf andacross the international marine border between Jordan and Israel. Furthermore, fish aquaculture cages thatwere present in Israeli waters had been removed thus allowing us to acquire new data in this area.

The Paradigm© software package was used to process the high-resolution seismic data. The processingprocedure started with geospatial data correction that include converting the geographic coordinates to Zone36 Universal Transverse Mercator projection in WGS84 datum and converting offset correction to the commonmidpoint of each trace. All seismic traces were vertically corrected using normal moveout with water acousticvelocity (1530m/s) and a static tidal correction. The data were filtered with a band-pass filter of 200–3000Hz,based on frequency analysis and a F-K filter to remove AC noise (50Hz and its multiples) in some of the seismicprofiles. A spherical gain was added in order to recover near-true amplitudes of the subsurface reflections.In addition, the seismic data that were acquired in the 2010 survey contained a long-period swell. In order toeliminate the swell from the seafloor and substrata reflections, a swell filter was applied.

The profiles presented in this study include a water depth scale that was calculated using an averagemeasured acoustic velocity of the water of 1530m/s, determined many times during the multibeam sonarsurvey [Sade et al., 2008]. The depth of the sedimentary layers was not correctly scaled because of a lack ofinformation on seismic velocities throughout the sediment column (applying velocity analysis is notapplicable for constant offset seismic data). The seismic velocity of the sediment is higher than that of thewater, so subbottom reflectors are actually deeper than is apparent in the seismic profiles.

The processed multibeam bathymetry and backscatter [Sade et al., 2008; Tibor et al., 2010] were analyzed inArcView© (ESRI) software to produce slope values and hillshade images of the seafloor and were integratedwith the seismic data in Paradigm© package.

The seismic profiles were interpreted in their 3-D context using the Paradigm© package. Seismic attributeswere analyzed in OpendTect© software. The seismic sections are interpreted to represent several units bymanually selecting the unit boundaries. The resulting horizons are interpolated using the Paradigmsoftware’s minimum curvature algorithm that allows the development of structural time maps of eachseismic unit. Isochron maps are contour maps that display the variation in two-way time (TWT) between thehorizons that bound the interpreted seismic unit. A fault boundary is introduced and contours generatedon each side thus allowing the lateral offsets of each isochron to bemeasured across the fault. In addition, thelateral and vertical offsets of interpreted linear structures representing relict fringing reef are measuredacross these faults. Finally, vertical offsets of interpreted horizons are measured across several faults.

4. Seismic Stratigraphy

We classify the reflectivity pattern of the seismic units (Table 1) based on seismic attributes of the internalreflections including amplitude level, dips, and reflector geometry (layered versus chaotic), as well asspatial continuity and stratigraphic relations (onlapping, downlapping, toplapping, truncating, etc.). Thisclassification is guided by the geological interpretation of the sedimentary properties and history of each unitdiscussed below. In a predominantly clastic shelf environment, such as the northern GEA shelf, seismicreflection amplitudes are controlled primarily by porosity and grain size changes [e.g., Bachrach and Mukerji,2004]. Laterally coherent, high-reflection amplitudes imply large changes of these properties across aninterface. While strong scattering indicates heterogeneity of these properties within the sedimentary unit.This internal heterogeneity is characterized by interbedded layers of sediment with different grain sizeand porosity. The scale of this interbedding is within the resolution of the seismic data (tens of centimeters).In contrast, low reflectivity and low scattering exhibits homogeneity of these properties through thesedimentary section.

The graphical shape of the reflectors provides information about the morphology and style of clastic shelfsedimentary deposits. Subparallel and flat reflectors normally represent well-sorted, layered sedimentdeposited, presumably, during periods of silt and sand deposition on the continental shelf, while chaotic orwavy reflectors likely represent poorly sorted sediment deposited in a higher flow regime with deltaic orfan delta sedimentation [Mitchum and Vail, 1977] (Table 1). This interpretation of sediment properties is basedon reflector geometry and awaits verification by sediment analysis of borehole data.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 6

Page 7: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Several mounded structures were observed in the seismic profiles, protruding from the surroundingsedimentary horizons. We interpret these as relict fringing reefs based on the following: (1) these moundswere characterized by high-amplitude reflection with seismic blanking below (Table 1) and horizontalreflections underneath the mounded structures appeared to be pulled-up, suggesting that these structureshave a higher seismic velocity than the surrounding sediment and are inferred to be rocky structures;(2) underwater observations made by divers on two such structures exposed at the seafloor at depths of~20m and ~60m revealed the presence of fossil coral that in several locations are covered with live corals,sponges, and seaweed [Makovsky et al., 2008; Hartman, 2012], and (3) these structures are long, narrow, andhave relatively flat tops similar to the fringing reefs that develop along coasts in close proximity to the seasurface today. All these characteristics suggest that the mounded structures represent relict fringing coralreefs (Table 1). Reefs with a more isolated structure, such as R6, are likely patch reefs.

Using the principles of sequence stratigraphy [e.g., Vail et al., 1977; Posamentier et al., 1988; Catuneanu et al.,2009], we reconstruct a relative history of sedimentary controls on the northern GEA shelf.Makovsky et al. [2008]similarly reconstructed the Holocene transgression, and we verify and extend back in time. Three distinctsubaerial unconformities define three seismic sequence boundaries SB0, SB1, and SB2. These sequenceboundaries divide the seismic stratigraphic section into four packages: the lower, middle, and upper packages,and the strata below these three packages. The sequence packages are further divided into seismic units(U0–U9). Reef systems (R1–R6) situated within the stratigraphic units are interpreted as well [Hartman, 2012].

Figure 5 correlates the GEA seismic sequence boundaries to eustatic changes in sea level. The relative baselevel is drawn relying on the relationship between the reflectors of the units (e.g., onlap and truncation). Theassumption that fringing reef developed during deceleration in sea level change [e.g., Montaggioni, 2005,and reference therein] or during sea level extrema (highstand or lowstand), led to the apparent steps alongthe relative base level curve during times when fringing reefs were generated. The environment ofsedimentation and climate are suggested based on the seismic stratigraphy analysis, the attributes of thereflectors within the units, and the spatial geometry of the units. The age model integrates the results of thisstudy with previous records of Quaternary sea level curve and climate [e.g., Shackleton and Opdyke, 1973;Fairbanks, 1989; Alley et al., 1997; Moustafa et al., 2000; Arz et al., 2003; Siddall et al., 2003, 2008; Hazan et al.,2005; Montaggioni, 2005; Arz et al., 2006, 2007; Bookman et al., 2006].

Siddall et al. [2003] developed a sea level curve of the Red Sea for the past 470 ka using oxygen-isotope datafrom planktic foraminifera from three sediment cores (Figure 5). These data show that sea level was at orabove the 60m isobath for most of the marine oxygen isotope stage 5 (MIS5; 130–80 ka). Red Sea levels mayhave reached above the 60m isobaths during early MIS3 briefly in several peaks, most notably at 52 ka[Siddall et al., 2003; 2008; Rohling et al., 2008]. Other data from the northern Red Sea presented in Arz et al.[2007] suggest that levels may not have reached as high as these isobaths. All of these sea level records show

Table 1. Seismic Characterizations of the Different Types of Seismic Units

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 7

Page 8: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

subsequent declining sea levels to the �120m glacial maximum in MIS2. Our data show that the R4.1 reefformed in a deglaciation cycle, where sea level rose to the ~60m isobath and then continued to rise toshallow water depths represented by deposition of unit U3 and the R4.2 reef. The R4.1 reef could not havedeveloped in MIS3, which was followed by regressive seas. Furthermore, our data show that the R4.1 reef

Figure 5. (a) Chronostratigraphic analysis of the northern shelf of the GEA and (b) its Quaternary climatic, environmentaland age model. The chronostratigraphic analysis is presented by applying a Wheeler diagram [Wheeler, 1958] to theinterpretation of the seismic stratigraphy of this study combined with the interpreted reefs. The suggested properties ofthe seismic units (U3 to U9) are illustrated in column 2, based on the analysis of the seismic attributes of the reflections(detailed in section 3). The Wheeler diagram (column 3) shows the units and the reefs at their relative time of depositionand generation. The interpreted sequence boundaries (SB0, SB1, and SB2) are shown as well. The relative base level isdrawn relying on the relationship between the reflectors of the units (e.g., on-lap and truncation). (c) Red Sea level curvefrom Siddall et al. [2003]. MIS =Marine oxygen isotope stage. RG = Reef generation period during a sea level stillstand.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 8

Page 9: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

formed between Sequence BoundarySB1 and SB2. SB1 and the overlyingU3 sediments mark the sea leveltransgression that reach watersshallower than 20m and is earlier thanthe Holocene because it is below SB2.Red Sea and global eustatic sea levelcurves indicate that this correlates toMIS5 and the 125 ka highstand. Earliertransgressive and regressive cyclesmarked by reefs R1–R3 and Units U1 andU2 are visible on the seismic reflectionprofiles and likely correlate to earlierglacial periods (Figure 5).

Our age model (Figure 5) is supportedby results from a 4.3m core (H02 inFigure 4) collected on the GEA shelf in25m water depth [Galloway, 2011]. A

sedimentation rate of 1.24mm/yr for the upper seismic stratigraphic units (U9-U8) was calculated (Figure 6).At the core study location, the top of seismic unit U6 lies at 11.6ms and corresponds to a depth of 10.4m(using a velocity of 1800m/s). Using the sedimentation rate of 1.24mm/yr calculated for U9-U8 deposition,the age of the top of the U6 is likely ~8400 year. Large uncertainties exist when extrapolating a sedimentationaccumulation rate to lower units that may have a different mode of deposition. Whereas Units U9–U7appears conformable, we have estimated the age of the top of U6 with the average sedimentation rate fromU9-U8. These will help determine the rate of motion on faults (see section below).

5. Fault Geometry

The northern GEA is cut along its northern shelf by several faults that offset shelf deposits and relict fringingreefs (Figure 7). These faults are interpreted along the seismic profiles based on displacement of the reflectors

Figure 7. Interpreted faults along the northern shelf of the Gulf of Elat/Aqaba. The location of each fault was picked fromthe seismic profiles, and each fault interpretation is marked by a grey dot. The fault lines were interpolated between thesedata points and projected onto the bathymetric image of the seafloor.

Figure 6. Calibrated and marine-reservoir-corrected radiocarbon ages onforaminifera tests collected from core H02 on the northern shelf of theGEA in proximity to the Evrona Fault [after Galloway, 2011]. These datayield a sedimentation rate of 1.24mm/yr.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 9

Page 10: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

(Figures 8–12). Our maps show the faults as we map them on the shelf, but they likely extend to deeper waterdepths where they are marked by linear seafloor scarps on the bathymetric data [Tibor et al., 2010]. Ourseismic reflection data had limited penetration over the steep slopes in water depths >100m.

Six major faults trending NNE-SSW and deforming the deposits of the shelf weremapped. These faults divide thenorthern GEA into three structural blocks; the Elat Subbasin on the west, the Ayla High in middle, and the AqabaSubbasin to the east (Figure 8). These three internal blocks were previously identified based on bathymetricdata [Tibor et al., 2010]. In addition, one E-W striking fault was mapped across the Aqaba Subbasin (Figure 7).

Two NNE trending faults are traced along the margins of the northern GEA. The western margin is flanked bya down-to-the-east, dip-slip fault that vertically displaces reflectors ~10ms (Figure 9). These reflectors are

Figure 8. Cross section along the northern GEA shelf comprised of four seismic profiles (DE10, W05a, E05a, and DA21;marked by a dashed line). This cross section shows the major faults that divide the northern GEA into the Elat subbasin,Ayla horst, and Aqaba subbasin. The marginal Elat and Aqaba Faults have a normal component, and they displace thereflectors of the Upper Stratigraphic Package. The Evrona Fault is a strike-slip fault with a component of normal slip. Boththe Ayla and East-Ayla faults have a normal component and neither displaces the reflectors of the upper depositionalsequence. The vertical scale is shown in two-way travel time (TWT) in milliseconds (ms).

Figure 9. Seismic profiles DE10 and DA21 and their interpretation of the major NNE-trending normal faults along themargins of the northern Gulf of Elat/Aqaba. The vertical scale is shown in two-way travel time (TWT) in milliseconds (ms).The reflectors are offset down toward the basin, across the Elat Fault and the Aqaba Fault. These faults displace the deposits ofthe last stratigraphic sequence, which correlates to the Holocene transgression.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 10

Page 11: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

interpreted as part of the SB2 subaerial unconformity and the deposition package that overlies it. No faultscarp was observed on the seafloor. Rather, the seafloor above the fault is covered by a live patch reef. We callthis the Elat Fault because it was previously mapped along this trend by Reches et al. [1987] and Ben Avrahamand Tibor [1993] (Figure 3). Another predominantly NNE trending normal fault is mapped along the easternGEA margin (Figures 7 and 9). A vertical offset of ~4ms of the SB2 subaerial unconformity suggestsdown-to-the-west, dip slip since the Holocene transgression. This fault is traced along the northward extensionof the bathymetric escarpment that was previously called the Aqaba Fault by Ben Avraham [1985] and Ehrhardtet al. [2005] (Figure 3).

The Elat Subbasin is bounded to the east by the prominent Evrona Fault seen on the bathymetry and theseismic profiles (Figures 7 and 10). Based on the seismic reflection profiles and the bathymetry, the EvronaFault is mapped as a single strand across most of the shelf. At a few locations, a second short fault strandis observed parallel to and west of the main fault (Figures 7 and 10). All reflectors visible on the seismicprofiles, as well as the seafloor, are offset by the Evrona Fault. The seafloor scarp is visible as a linear step inthe bathymetry.

We identify andmeasure several different piercing lines across the Evrona Fault. An E-W linear segment of theR5 fossil fringing reef in ~60m to ~70m water depth (Figure 11) is clearly visible on the multibeambathymetric data [Sade et al., 2008; Tibor et al., 2010]. The R5 reef is sinistrally offset by 28 ± 6m across the

Figure 10. Seismic profiles W05, W06, and W07 and their interpretation of the Evrona Fault. The units along both sides of the Evrona Fault are folded down to thefault. A change in the thickness was also observed. The Evrona Fault deformed the reflectors up to the seafloor. In some seismic profiles (e.g., W05) two branches ofthe Evrona Fault are observed.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 11

Page 12: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Evrona Fault in agreement with the estimate of Makovsky et al. [2008]. A much older relict reef (R4.1) that isburied below the surface sediment and R5 reef was also used as a piercing line. By measuring the trend ofthe buried R4.1 reef identified in seismic reflection profiles, we estimate a left lateral offset of 62 ± 13m acrossthe Evrona Fault for this feature (Figure 11).

In the Evrona Subbasin, the lens-shaped seismic unit U6 was deposited landward of the R5 reef and overliesseismic unit U5. U5 is a time-transgressive unit of nearshore deposits (a transgressive lag) that marks the rapidearly Holocene sea level rise above the SB2 sequence boundary. The overlying seismic unit U6 consists ofscattered high-amplitude, wavy, and chaotic reflections derived from a mixed sedimentary source withseveral localized, subhorizontal reflectors. U6 thins out both in its deeper part (~60 mbsl) and along itsshallower part (~30mbsl). The lenticular shape is derived from the primary depositional environment within adeltaic to fan delta setting where proximal and distal regions receive less sediment compared to the maindepocenter. Deposition of U6 ceased as environmental conditions shifted resulting in the formation ofoverlying reef R6.

Because of the unique shape of the deposit, seismic unit U6 can be used to measure offset across the EvronaFault. We constructed a map of the thickness of unit U6 measured in two-way travel time (an isochron map).In constructing the U6 isochron map, a boundary was introduced along the trace of the Evrona Fault, andcontours are generated independently on each side of the plane using the automated software. Figure 12shows the isochron map of U6. Offset of each of the isochrons was measured across the Evrona Fault onthe shoreward side of the deposit and on the slope side yielding an average offset of all measurements of33 ± 29m (Table 2). However, contours on the slopeward edge of the unit U6 show that the deposit isdissected in several regions by postdepositional erosion as seen in lines W06 and W07 (Figure 10). Thisleads to the wide variance of the slopeward isochron offset measurements of U6. Given this observation,the average offset of the shoreward edge of U6 which is largely buried under units U7–U9 measures 29±7mand is a more reliable estimation of the displacement of U6.

We also measure a component of normal displacement along the offset R4.1 and R5 reefs. Both reefs arepresumed to have developed close to the sea surface and thus were originally level and formed at a constantelevation. Both the segments of R4.1 and R5 are offset by ~12m down-to-the-west toward the Elat Subbasin.These data indicate that the dip-slip motion occurred after growth of the R5 reef (Figure 12).

The central block of the northern GEA shelf, the Ayla High, is bounded by the Evrona Fault to the west and by anewly discovered fault to the east that we call the Ayla Fault (Figures 8 and 13). Reflectors across theNNE trending Ayla Fault show normal, down-to-the-east offset (Figure 13). East of the Ayla Fault, the layers of thelower to middle stratigraphic packages are concave upward forming a small basin. The small basin is boundedon the east by another fault with down-to-west offsets. We call this the East Ayla Fault (Figures 8 and 13).

Figure 11. Sinistral offsets and depths of Reef 4.1 (R4.1) and Reef 5 (R5) along the Evrona Fault. (a) The trend of R5 ismarked bya green line and is delineated based on the bathymetric data (see shaded background). The trend of R4.1 marked by a blueline is the interpreted location from the seismic reflection profiles. Both trends of the reefs were picked from the top ofthe structures. Depths of the top of the reefs aremarked by the same colors as the lines of the reefs’ trends. (b) Enlarged sketchof the reefs and the measured lateral offset across the Evrona Fault (marked by a red line). Because of the uncertainty in thelocation of the trends, an estimated 20% of error is added due to flexibility in outlining the reefs’ trend. As a result sinistraloffsets of 62 ± 13m for R4.1 and 28± 6m for R5 are estimated.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 12

Page 13: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Both faults are vertical and show significant mismatch of sedimentary layer thickness across the faultssuggesting strike slip. The basin likely formed by transtension between these two oblique-slip faults. Both theAyla and East Ayla Faults are buried beneath the SB2 sequence boundary. They offset only the reflectors of themiddle and lower stratigraphic packages and are capped by the Upper Stratigraphic Package (Figures 8 and 13).

Figure 12. (a) Isochron map of seismic unit U6 across the northern Gulf of Elat/Aqaba shelf showing the location of theinterpreted seismic reflection lines. Thickness is shown in milliseconds (ms). (b) The isochron map of U6 across theEvrona Fault mapped with a 2ms interval reveals that the unit was sinistrally offset. The measured offsets are shown inTable 2. The shaded relief seafloor map is generated from multibeam bathymetric data [Sade et al., 2008].

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 13

Page 14: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Because the R4.1 and R5 have beeneroded from this section of theshelf, there are no piercing lines tomeasure the amount of slip on theAyla or East Ayla faults.

Another fault zone is traced in theseismic profiles within the Aqabasubbasin. This NE trending fault zonewith two parallel traces is newlydescribed and named here the WestAqaba Fault (Figures 8 and 13). Thefault is active and can be mapped tothe seafloor. The stratigraphicthickness of three units, U3, U5, andU6,

Figure 13. Seismic profiles E05a and M2W005 and their interpretation of the Ayla, East Ayla, and West Aqaba faults alongthe northern Gulf of Elat/Aqaba shelf. The Ayla Fault and the East Ayla Fault are normal faults that bound a small graben.The faults are not presently active and are buried beneath the Sequence Boundary SB2 thatmarks the Holocene transgression.The West-Aqaba Fault appears to be an active strike-slip fault zone. The faults intersect the seafloor, and the thickness ofseismic units changes appreciably across the faults indicating lateral slip. The vertical scale is shown in two-way travel time(TWT) in milliseconds (ms).

Table 2. A Summary of the Measured Offsets of the Isochrons of SeismicUnit U6a

Isochron (ms) Shoreward Offset (m) Slope Side Offset (m)

2 19 03 19 444 38 595 38 836 35 667 31 358 29 59 21 1110 25 1911 30 57

aAverage of all measurements: 33 ± 29m; average of shoreward mea-surements: 29 ± 7m.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 14

Page 15: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

changes across the fault strands. We interpret these data to indicate that the West Aqaba Fault is a strike-slipfault, probably with left lateral slip.

Seismic profiles oriented N-S along the Aqaba Subbasin revealed an E-W trending fault with a normalcomponent of slip (Figure 14). The reflectors on the northern side are offset downward by ~2ms. The fault,named here the E-W fault, was traced between the West Aqaba fault and the East Ayla Fault (Figure 8).

The structural block, the Ayla High, separates the Elat and Aqaba Subbasins. The Ayla High is bounded bythe Evrona and Ayla Faults, both with an apparent normal component of slip. The dip-slip motion on theEvrona Fault can be measured on the R5 reef as described above. The bathymetry of the Ayla High isshallower with respect to the Elat and Aqaba Subbasins. Although the fringe reefs R4.1 and R5 are leftlaterally offset by the Evrona Fault between the Elat Subbasin and the Ayla High, no significant offset wasobserved to the east of the Ayla High along these reefs within the Aqaba subbasin. The tops of R4.1 and R5are inclined in the Ayla High to the east by ~8ms and ~2ms (respectively) over a distance of ~1000m(Figure 11), suggesting that the Ayla High was tilted eastward by ~0.46° and ~0.11° (respectively) since thesefringing reefs developed.

6. Fault Slip Rates

The interpretation of seismic reflection data reveals a complex system of faults at the northern shelf of theGEA with strike, oblique, and normal slip. All of the faults are active showing Holocene displacements exceptthe Ayla and East Ayla Faults. We calculate the average slip rate of each fault based on offset reflectorsand piercing lines and the age model proposed in Figure 5 and described in section 4 of this paper.

The rate of subsidence across the Elat and Aqaba normal faults is estimated based on offsets of reflectorsrepresenting the upper sequence boundary SB2 (Figure 7). The SB2 reflector truncates reflectors below itand is interpreted as the result of subaerial erosion. Reflectors above SB2 represent marine sedimentationduring the most recent sea level rise. Hence, the erosional surface of SB2 was shaped in the Holocenetransgression and is covered by sediment when sea level rose to present levels. The SB2 surface and theoverlying sediments of units U5 and U6 are vertically offset an equal amount implying that the total measuredoffset developed since the early Holocene approximately 9–10 ka. The sedimentary fill is thicker on the basinside of each fault indicating syntectonic deposition. Using an age of 9–10 ka for the SB2 surface, we calculate avertical slip rate of 1.0± 0.2mm/yr and 0.4± 0.1mm/yr for the Elat and Aqaba faults, respectively (Figure 15).

Figure 14. Seismic profiles N09 and N25 trending N-S along the northern shelf of the GEA, situated in the Aqaba subbasin.A normal fault trending E-W is observed in these sections. The E-W Fault shifts downward the reflectors and the units acrossits northern side. The fault offsets the units up to the seafloor.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 15

Page 16: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Our mapped trace of the offshore Evrona Fault is in agreement with previous suggestions of this fault trendbased on deeper seismic reflection profiles to the north and south [Frieslander, 2000; Ehrhardt et al., 2005] andonland field mapping of the fault scarps [e.g., Garfunkel, 1970; Amit at al., 1999]. Makovsky et al. [2008]estimated an average slip rate of 2.7 ± 1.5mm/yr for the Evrona Fault across the northern GEA shelf based ona 30 ± 10m offset of the relict fringing reef (identical with our R5) that developed during the Younger Dryasstillstand (11.5 ± 2 ka). Here we estimate slip rates for the Evrona Fault based on offsets of three structures(reefs R4.1 and R5, and the shoreward thickness of unit U6). Reefs R4.1 and R5 are both vertically offset by~12m (Figure 11).

Figure 15. A map of the major faults (black lines) that divide the northern GEA and their estimated slip rates (internalframes), overlaid on the curvature of the seafloor. Faults are marked by solid lines where they were interpreted from theseismic data across the shelf. The continuation of the faults southward is suggested based on their extrapolation fromthe curvature analysis of the bathymetric data and in similarity to the fault outlines that were previously suggested [Recheset al., 1987; Ben Avraham and Tibor, 1993; Ehrhardt et al., 2005; Tibor et al., 2010]. Their continuation northward is suggestedbased on their extrapolation from presettlement (1945) aerial photographs. The graphs of average slip rates and tiltrates as a function of the averaging time-span are shown for their relevant faults. Average slip rates are marked by blacklines and calculated uncertainty indicated by gray boxes.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 16

Page 17: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Reef R4.1 is left laterally offset by 62 ± 13m. Following our age model (Figure 5), R4.1 likely developed duringthe last interglacial (MIS 5) that spanned 130–80 ka [e.g., Siddall et al., 2003, 2004]. We cannot refine the agefurther as sea levels in the Red Sea were above the 60m isobaths for most of this interval. There is also a smallprobability that the R4.1 reef developed during the penultimate deglaciation (Termination II). Siddall et al.[2006] show that there is a stillstand at about the 60m isobath in the Red Sea around 132–134 ka duringTermination II similar to the last deglaciation and the Younger Dryas climatic event. We estimate that the R4.1reef likely correlates to the MIS5 transgression. Using these MIS5 age determinations, we calculate an averagelate Pleistocene slip rate for the Evrona fault of 0.7 ± 0.3mm/yr since the R4.1 generation.

The R5 is a fringe reef that provides a piercing line across the Evrona Fault and is offset 28 ± 6m. The R5 reefdeveloped in the last deglaciation (Termination I). We interpret the R5 reef at the 60–70m isobaths tohave grown during the Younger Dryas stillstand (12.8–11.5 ka). These data indicate a Holocene slip rates of2.3 ± 0.6mm/yr.

The unit U6 developed above the Holocene transgressive sequence boundary (SB2) and the unit U5transgressive lag. As the slope side the unit U6 displays postdepositional erosion, we use the offsets of theshoreward isochrons to measure an average offset of 29±7m (Figure 12 and Table 2). As U6 is clearly youngerthan the R5 reef, this offset range represents a maximum value. The range of offsets is a product of both thelimits of the methodology of the isochron map generation and the natural variability of the sedimentarydepositional environment. As we have nomethod for statistically trimming themaximum offsets, we report thefull range here. Our chronostratigraphic analysis indicates that U6 was deposited during a wetter climaticperiod that began at 9.25 ka [Arz et al., 2003] until an arid event of 8.4–8 ka [e.g., Alley et al., 1997]. An age of8.4–8 ka for the cessation of deposition of the unit U6 is in agreement with our age-depth model for theoverlying U9-U8 seismic units based on core analyses (Figure 6) [Galloway, 2011]. Using the 29±7m offset andan age for the top of U6 of 8.4–8 ka yields a slip rate of 3.5± 0.9mm/yr since deposition of U6. Given theuncertainty in the age determination of U6, we can also calculate a maximum limiting age of 9.25 for the unit.Using this maximum limiting age and the offset of 29±7m, a slip rate of 3.13± 0.76mm/yr is calculated.Although the U6 isochron piercing lines are less certain than the R5 fringe reef, the U6 slip rate is in agreementwith the rate calculated from the offset of the R5 reef. An average estimation of the U6 offset is 3.4± 1.0.

The congruent normal offset of both R4.1 and R5 suggests that the dip slip along the Evrona Fault occurredafter the development of reef R5. Since R5 developed until 11.5 ka, the Evrona fault has an averagesubsidence rate of ~1mm/yr during the Holocene (Figure 15). The resulting sinistral and normal slip ratesalong the Evrona Fault indicate an increasing slip rate from the Late Pleistocene to the early Holocene.We therefore suggest that the Evrona Fault is relatively young and that its fault history developed after thelate Pleistocene. The Evrona Fault currently appears to accommodate a significant part of the slip of the DSTfault system across the shelf. North of the GEA shelf, the Evrona Fault was mapped [Garfunkel, 1970;Frieslander, 2000] along the eastern side of the hotel district of the city of Elat.

East of the Ayla High, the Ayla and the East Ayla faults bound an internal sag depression and are likely inactiveoblique-slip faults. Both these faults terminate upward at SB2. This interpretation implies that the Ayla andthe East Ayla Faults had been inactive during the Holocene or even for a longer period during subaerialexposure before the last marine transgression.

The West Aqaba Fault has a strike-slip component that could not be estimated due to the limitation of thedata of this study. This fault offsets the reflectors of the upper sequence boundary SB2 (Figure 13), suggestingthat it had been active during the Holocene. This is the only fault for which a predominant strike-slipcomponent is inferred within the Aqaba Subbasin. TheWest Aqaba fault has two interpreted branches, whichmay represent a flower structure of a relatively young strike-skip fault [e.g., Harding and Lowell, 1979]. Thisfault is traced only in a limited zone; hence, its role in the tectonic framework of the northern GEA is unclear.

The Ayla High is a relatively young, mid-to-late Quaternary structure. Unit 3 represents prograding lagoonalsediment deposited along the Ayla High and retrograding nearshore sediment within the subbasins,suggestive of higher subsidence within the Elat and Aqaba subbasins during mid-Late Quaternary.Furthermore, R4.1 and R5 are tilted eastward by ~0.46° and ~0.11° (Figure 11), respectively, based on our agemodel estimated tilting rates of ~4°/my and ~9°/my (Figure 15), respectively. These results emphasize theincreasing tectonic role of the Ayla High and the Evrona Fault during the Late Quaternary and particularly in

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 17

Page 18: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

the Holocene. A possible cause of this tilting and itsincreasing rate is the change in fault activity thatbound the Ayla High (i.e., Evrona and Ayla faults). TheEvrona Fault along the western side of the Ayla High isin a process of increasing activity and the Ayla Faultalong the eastern side of the Ayla High has not beenactive at least since the beginning of the Holocene. Therelative uplift of the western Ayla High is compatiblewith the model of strain buildup across normal faultthat was suggested by Stein et al. [1988]. In this modelthe layers in the upper block are inclined upward inproximity to the fault, similar to the eastward tilting ofthe Ayla High. Therefore, the activation of these faultsand slip rate differences imply that the tilting of theAyla High is a result of a balance between the slipacross the Evrona and Ayla Faults.

A normal fault (Figure 14) oriented east-west wasmapped in this study between the West Aqaba Faultand the East Ayla Fault (Figure 15). This fault probablybelongs to an east-west to NW trending system oftransverse faults mapped by Garfunkel [1970] andNiemi and Smith [1999] (Figure 3). This system ofnormal faults is connected to the Aqaba Fault thatdiverts eastward on land, similarly to the horsetail splaymodel [e.g., Twiss and Moores, 1992].

7. Discussion

The most basic model proposed to explain transformfault basins is the pull-apart basin [e.g., Carey, 1958;Quennell, 1959; Burchfiel and Stewart, 1966; Freund et al.,1968; Mann et al., 1983]. It is a first-order rigid platemodel where transform basins form at bends or stepsalong strike-slip faults. The pull-apart model predicts asymmetrical basin where the two longitudinal sides arebounded by predominantly strike-slip faults, and theother two sides are bounded by predominantly normaltransverse faults (Figure 16). Field observations alongtransform basins are commonly inconsistent with thesimple pull-apart model. Many transform basins areasymmetrical, both longitudinally and laterally [e.g., BenAvraham, 1992; Armijo et al., 2002].

Several models suggest that the development of theinternal structure of transform basins is related to theevolutionary stage of the basin [e.g.,McClay and Dooley,1995; Rahe et al., 1998; Sims et al., 1999]. McClay andDooley [1995] and Rahe et al. [1998] examined the

development of the internal structure in transform basins in sandbox analog models using dry sand torepresent the brittle crust above a horizontal detachment horizon. Their models showed an initial formation ofrhombic basins bounded by normal faults (Figure 16b). In advanced stages, intrabasinal deformation developscharacterized by an oblique cross-basin fault zone that links the main strike-slip faults of both sides of thebasin. This model suggests that the Evrona Fault on the GEA shelf is located above the main crustal fault.The Ayla, East Ayla, and West Aqaba Fault zones may be accommodating part of the plate boundary motion as

Figure 16. Schematic illustrations of models of basin evo-lution along left lateral strike-slip fault. (a) The pull-apartbasin model (modified after Garfunkel [1981]); a 3-D view ofa rhomb-shaped basin develops along a left bend of astrike-slip fault. The two longitudinal sides of the basin aredefined by strike-slip faults, and the two latitudinal sides aredefined by normal faults. (b) The evolutionary model isillustrated in three phases in an orthogonal view (modifedafter McClay and Dooley [1995], Rahe et al. [1998], andWu et al. [2009]). In the first phase a rhomb-shaped basindevelops. Then a cross-basin strike-slip fault evolves, andthe strike-slip strain partitions between the marginalfaults and the intrabasin fault. In the last phase themarginal faults became inactive, and most of the strike-slip strain is along the cross-basin fault.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 18

Page 19: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

the slip is transferred across from the east to the west side of the basin. The Ayla and East Ayla faults bound abasin on the shelf that formed from localized extension. Both faults deform late Pleistocene sediment butare truncated by the SB2 sequence boundary. Slip is mainly transferred across the northern GEA shelf by theWest Aqaba Fault zone and transverse normal faults.

One offshore E-W transverse fault was documented in the seismic reflection data of this study (Figure 14).In the Aqaba area, field mapping has shown several NW to E-W striking faults [Niemi and Smith, 1999;Slater and Niemi, 2004,Mansoor et al., 2004]. These transverse, normal fault likely act to further transfer strainacross the pull-apart basin as the basin grows northward and subsidence increases to the south.

Seismic data from theM 7.3 Nuweiba earthquake of 22 November 1995, the largest modern earthquake to berecorded on the DST, and the precursor earthquakes and aftershocks provide a model for the pattern ofactive faults in the GEA. The Nuweiba earthquake ruptured the submarine Aragonese Fault that borders thewest side of the Aragonese Basin and the east side of the Elat Deep (Figure 3) in the GEA in two or threesubevents from south to north [Dziewonski et al., 1997; Pinar and Türkelli, 1997; Klinger et al., 1999, 2000c;Hofstetter et al., 2003]. These seismic studies showed that fault motion on the main shock was predominatelyleft lateral strike slip, whereas the subevents had large components of normal slip. The 1995 rupture roughlyaligns between the location of the southern precursor earthquake swarms in 1993 and the northernearthquake swarm 1983 [Klinger et al., 1999]. Earthquake focal plane mechanisms of the largest event ofthe 1993 data indicate that rupture occurred on a N32°W trending, normal fault [Dziewonski et al., 1994;Pinar and Türkelli, 1997].

Focal plane mechanisms of the 1995 Nuweiba aftershock sequence show a complex pattern of dip sliporiented E-W or NW across the basin, normal fault motion parallel to the margins of the GEA and close to oronshore, and left lateral motion within the basin away from the main mapped fault [Hofstetter et al., 2003].The E-W and NW striking, dip slip likely ruptured cross faults in the submarine pull-apart basins that act toaccommodate subsidence and lengthening of the basin similar to transverse fault mapped in the Aqabasubbasin in this study. Basin perpendicular extension was also observed with oblique slip in the main 1995shock, subevents, and in aftershocks [Klinger et al., 2000; Hofstetter et al., 2003]. We have shown that theEvrona strike-slip fault on the GEA shelf has a significant component of normal slip suggesting that it may bethe end of a rupture segment.

The basin margin normal faults in the deep marine Elat and Aragonese Basins are analogous to the Elat andAqaba Faults in the north in the GEA basin that produce subsidence on faults oriented parallel to the mainstrike-slip fault. Linear ground cracks along the Egyptian coast of Sinai observed after the Nuweibaearthquake appear to represent sympathetic slip along preexisting basin marginal faults and not coseismicslip [Klinger et al., 1999]. Onshore ground deformation parallel to the main rupture is likely caused bygravitational sliding perhaps also during the aftershocks [Baer et al., 2002].

The most prominent faults projecting on the topography and the GEA bathymetry are the Elat and Aqabamarginal faults. These faults define the borders of the basin and they generate steep slopes along themargins of the northern GEA. The influence of these faults on the general shape of the basin predominatesand these marginal faults are active since the inception of the basin up to the present. The Elat and AqabaFaults are likely normal faults that parallel the northern GEA basin similar to other transtensional basins alongthe DST, including the Dead Sea and the Aragonese and Dakar Basins in the GEA. Strike slip has not beendocumented on the Elat and Aqaba Faults, but this motion cannot be ruled out. Vertical slip rates are high onthese faults indicating that these faults accommodate rapid subsidence of the GEA basin.

A possible explanation of the coexistence of strike-slip faults such as the Evrona, Ayla, East Ayla, and WestAqaba faults between marginal Elat and Aqaba normal faults was provided by a numerical elastic strainmodel [Bowman et al., 2003]. This model describes the physics of fault partitioning along oblique, strike-slipfaults. It predicts that a region that contains an undergoing deep oblique, strike-slip fault should be expectedto accommodate motion on predominantly strike-slip and dip-slip faults at the surface. The displacementsare partitioned between strike slip at the center and dip slip at the margins.

The pattern of transform basin evolution from dominant marginal faults to partitioned intrabasin faults,concluding with a dominant intrabasin fault had been suggested by other studies of continental transforms.Zhang et al. [1989] mapped a prominent diagonal fault that crosses the entire Dayinshui basin of northern

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 19

Page 20: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

China and accommodates some of the motion from its bounding faults. They defined it as an extinctpull-apart basin. A study of a similar system of transform basins in the Sea of Marmara [Rangin et al., 2004]revealed strain localization through time from a system of pull-apart basins to a main propagated transformfault in an inactive pull-apart basin. Along the DST, Schattner and Weinberger [2008] mapped a intrabasinal,left-lateral diagonal fault within the Hula Basin developing during a phase when the pull-apart mechanismsno longer prevailed. Marco [2007] suggested that the faulting along the DST began in the early-middleMiocene over a wide (~50 km) zone and became localized by the end of the Miocene. Later during thePlio-Pleistocene, new faults were formed and another cycle of faulting localization began. The northern shelfof the GEA appears to be a location in an intermediate stage of development where distributed deformationtranslates motion onto the Evrona Fault from other intrabasinal faults.

The results of this study support the model of basin evolution along strike-slip fault, where the stage ofevolution is predicted by fault geometry and slip. Whereas the initial phase of transform basin evolution ischaracterized by a rhomb-shaped graben bounded by marginal dominant faults, the intermediate phasehas intrabasin faults partitioning strain on various types of faults. The final stage shows localization of thestrike-slip strain to a main intrabasin fault.

Our seismic reflection data show that most of the GEA intrabasin faults became active at least prior to three tofour sea level cycles, several hundreds of thousands of years ago. However, no offset is observed alongthe reflectors of the SB2 sequence boundary across the Ayla and East Ayla Faults. These faults are inferred tohave been inactive at least since the last sea level rise or since the beginning of the last subaerial exposure(Pre-MIS5). Since the last marine transgression (i.e., MIS 5, ~130–80 ka), the average sinistral, slip rate of theEvrona Fault has increased from ~0.7 ± 0.3mm/yr to between approximately 2.3 and 3.4mm/yr, and theaverage rate of tilting of the Ayla High has increased from 4°/my to 9°/my. These calculated rates of faultactivity may potentially be attributed to the narrowing of the transform strain from the margins of the basinand from several intrabasin faults including the Ayla and East Ayla faults to a main intrabasin fault, the EvronaFault that accommodates most of the plate boundary slip.

Regional and local slip rates on various segments of DST system have been measured by several methods.Quaternary slip rates in the range of 2 to 8mm/yr were estimated based on offset surface features alongsegments of the Arava and EvronaF [Zak and Freund, 1966; Garfunkel, 1981; Ginat et al., 1998; Klinger et al.,2000a, 2000b; Niemi et al., 2001; Makovsky et al., 2008; Le Béon et al., 2010, 2012]. Geodetic estimates of thecurrent horizontal plate motion across the DST are in the range of 3.7 to 7.5mm/yr [Wdowinski et al., 2004;Mahmoud et al., 2005; Ostrovsky, 2005; Gomez et al., 2007; Le Béon et al., 2008; Al Tarazi et al., 2011; Sadeh et al.,2012] and are in agreement with the geological rates. This estimation is in agreement with the model ofDeves et al. [2011] that find that more than half of the deformation in the Dead Sea region can localize onkinematically stable throughgoing, strike-slip faults while the rest have to remain distributed.

Taking into account these previous results and our calculated slip rate for motion on the Evrona Fault, itappears that slip rates on this fault may have increased from the Late Pleistocene to the present. The changein the slip rate of the Evrona Fault over the late Quaternary may be related to fault motion migrating froman adjacent fault strand such as the Ayla or East Ayla to the Evrona Fault. Conversely, motionmay have shiftedfrom basin margin to intrabasinal structures as suggested for other faults in the DST system [e.g.,Marco et al.,1996; Amit et al., 2002].

The large, dip-slip component on the Evrona Fault suggests that the fault on the shelf is located near thesouthern end of the segment. This is in agreement with the mapping of Ehrhardt et al. [2005] and inferencesfrom the bathymethric mapping [Tibor et al., 2010] that show the Evrona Fault in the deeper portion of thenorthern GEA to curve from the east and the Aragonese Fault to the west side of the basin.

8. Conclusions

The Quaternary geological evolution of the GEA has been affected by sedimentary and tectonic processes.The results of this study revealed six intrabasin, NNE trending faults and one transverse fault. The activity andslip measurements across the faults imply that a major strike-slip fault, the Evrona Fault, occupies most of theHolocene left lateral slip within the basin. The onshore extension of the Evrona Fault is located along theeastern side of the hotels district of Elat, reflecting a significant earthquake hazard to this highly populated

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 20

Page 21: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

area. The average sinistral slip rate across this fault increased from 0.7 ± 0.3mm/yr in the late Pleistocene tobetween 2.3 and 3.4mm/yr in the Holocene possibly with the addition of slip from parallel faults. Thelatter rate is comparable to other geologic and geodetic studies of the plate boundary velocity. Furthermore,two other intrabasin faults have been inactive over the last several tens of thousands of years. The changein rate and location of active faulting may reflect migration of fault motion from Ayla and East Ayla strands tothe Evrona Fault during time.

The results of this study support a basin evolution model along strike-slip faults, where the stage of evolutionis determined by fault geometry and sense of fault motion. The northern GEA basin is situated between thedeep rhomb-shaped basins of the GEA to the south and the shallow basins, mostly occupied by a singletransform fault of the Arava Valley to the north. The basin of the northern GEA is still in an intermediatephase of evolution, where slip is partitioned onto marginal normal faults and strike-slip and oblique-slipintrabasinal faults.

ReferencesAl Tarazi, E., J. Abu-Rajab, F. Gomez, W. Cochran, R. Jaafar, and M. Ferry (2011), GPS measurements of near-field deformation along the

southern Dead Sea Fault System, Geochem. Geophys. Geosyst., 12, Q12021, doi:10.1029/2011GC003736.Alley, R. B., P. A. Mayewski, T. Sowers, M. Stuiver, K. C. Taylor, and P. U. Clark (1997), Holocene climatic instability: A prominent, widespread

event 8200 yr ago, Geology, 25, 483–486.Amit, R., J. B. J. Harrison, and Y. Enzel (1995), Use of soils and colluvial deposits in analyzing tectonic events—The Southern Arava Rift, Israel,

Geomorphology, 12, 91–107.Amit, R., J. B. J. Harrison, Y. Enzel, and N. Porat (1996), Soils as a tool for estimating ages of Quaternary fault scarps in a hyperarid environment

—The southern Arava Valley, the Dead Sea Rift, Israel, Catena, 28, 21–45.Amit, R., E. Zilberman, N. Porat, and Y. Enzel (1999), Relief inversion in the Avrona Playa as evidence of large-magnitude historical earthquakes,

southern Arava Valley, Dead Sea Rift, Quat. Res., 52, 76–91.Amit, R., E. Zilberman, Y. Enzel, and N. Porat (2002), Paleoseismic evidence for time dependency of seismic response on a fault system in the

southern Arava Valley, Dead Sea rift, Israel, Geol. Soc. Am. Bull., 114, 192–206.Armijo, R., B. Meyer, S. Navarro, G. King, and A. Barka (2002), Asymmetric slip partitioning in the Sea of Marmara pull-apart: A clue to

propagation processes of the North Anatolian Fault?, Terra Nova, 14, 80–86.Arz, H. W., F. Lamy, J. Pätzold, P. J. Müller, and M. Prins (2003), Mediterranean moisture source for an Early-Holocene humid period in the

Northern Red Sea, Science, 300, 118–121.Arz, H. W., F. Lamy, and J. Pätzold (2006), A pronounced dry event recorded around 4.2 ka in brine sediments from the northern Red Sea,

Quat. Res., 66, 432–441.Arz, H. W., F. Lamy, A. Ganopolski, N. Nowaczyk, and J. Pätzold (2007), Dominant Northern Hemisphere climate control over millennial-scale

glacial sea-level variability, Quat. Sci. Rev., 26(3–4), 312–321.Bachrach, R., and T. Mukerji (2004), The effect of texture and porosity on seismic reflection amplitude in granular sediments: Theory and

examples from a high-resolution shallow seismic experiment, Geophysics, 69, 1513–1520.Baer, G., G. Shamir, D. Sandwell, and Y. Bock (2002), Crustal deformation during 6 years spanning the Mw 7.2 1995 Nuweiba earthquake,

analyzed by interferometric synthetic aperture radar (InSAR), Isr. J. Earth Sci., 50, 9–22.Bartov, Y., G. Steinitz, M. Eyal, and Y. Eyal (1980), Sinistral movement along the Gulf of Aqaba—Its age and relation to the opening of the Red

Sea, Nature, 285, 220–222.Ben Avraham, Z. (1985), Structural framework of the Gulf of Elat (Aqaba), Northern Red-Sea, J. Geophys. Res., 90, 703–726, doi:10.1029/

JB090iB01p00703.Ben Avraham, Z. (1992), Development of asymmetric basins along continental transform faults, Tectonophysics, 215, 209–220.Ben Avraham, Z., and G. Tibor (1993), The northern edge of the Gulf of Elat, Tectonophysics, 226, 319–331.Ben Avraham, Z., G. Almagor, and Z. Garfunkel (1979a), Sediments and structure of the Gulf of Elat (Aqaba) Northern Red-Sea, Sediment. Geol.,

23, 239–267.Ben Avraham, Z., Z. Garfunkel, G. Almagor, and J. K. Hall (1979b), Continental breakup by a leaky transform—Gulf of Elat (Aqaba), Science, 206,

214–216.Ben-Avraham, Z., M. Lazar, U. Schattner, and S. Marco (2005), The dead sea fault and its effect on civilization, in Lecture Notes in Earth Sciences:

Perspectives in Modern Seismology, edited by F. Wenzel, pp. 145–167, Springer, Heidelberg, Germany.Bookman, R., Y. Bartov, Y. Enzel, and M. Stein (2006), Quaternary lake levels in the Dead Sea basin: Two centuries of research, Geol. Soc. Am.

Spec. Pap., 401, pp. 155–170, doi:10.1130/2006.2401(10).Bowman, D., G. King, and P. Tapponnier (2003), Slip partitioning by elastoplastic propagation of oblique slip at depth, Science, 300, 1121–1123.Burchfiel, B., and J. Stewart (1966), ‘Pull-apart’ origin of the central segment of Death Valley, California, Geol. Soc. Am. Bull., 77, 439–442.Carey, S. (1958), The tectonic approach to continental drift, a Symposium, Hobart, Tasmania, Geology Department, Univ. of Tasmania.Catuneanu, O., et al. (2009), Towards the standardization of sequence stratigraphy, Earth Sci. Rev., 92, 1–33.Deves, M., G. C. P. King, Y. Klinger, and A. Agnon (2011), Localised and distributed deformation in the lithosphere: Modelling the Dead Sea

region in 3 dimensions, Earth Planet. Sci. Lett., 308, 172–184, doi:10.1016/j.epsl.2011.05.044.Dziewonski, A. M., G. Ekstrom, and M. P. Salganik (1997), Centroid-moment tensor solutions for October–December 1995, Phys. Earth Planet.

Inter., 101, 1–12.Dziewonski, A. M., G. Ekstrom, and M. P. Salganik (1994), Centroid-moment tensor solutions for July–September 1993, Phys. Earth Planet.

Inter., 83, 165–174.Ehrhardt, A., C. Hubscher, Z. Ben-Avraham, and D. Gajewski (2005), Seismic study of pull-apart-induced sedimentation and deformation in

the Northern Gulf of Aqaba (Elat), Tectonophysics, 396, 59–79.Enzel, Y., R. Amit, N. Porat, E. Zilberman, and B. J. Harrison (1996), Estimating the ages of fault scarps in the Arava, Israel, Tectonophysics, 253,

305–317.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 21

AcknowledgmentsThis study was funded by the U.S.Agency for International DevelopmentMiddle East Regional Cooperation pro-gram under contracts TAU-MOU-05-M25-04 and TAU-MOU-08-M29-036. Wethank the crews of the R/V Etziona,Mediterranean-Explorer, and Danny-boy,and the Jordanian military and IsraeliNavy for coordinating and helping withthe marine survey. Analyses of the seis-mic data was facilitated by Paradigm©and OpendTect©, which provided uswith their software. We appreciate theassistance of GeoResources which pro-vided us with the seismic equipment,training, and support. We thank AmotzAgnon, an anonymous reviewer, andthe Associate Editor for their criticalcomments, which help us revise andimprove this paper.

Page 22: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Fairbanks, R. G. (1989), A 17,000-year glacio-eustatic sea level record—Influence of glacial melting rates on the Younger Dryas event anddeep ocean circulation, Nature, 342, 637–642.

Freund, R., I. Zak, and Z. Garfunkel (1968), Age and rate of the sinistral movement along the Dead Sea Rift, Nature, 220, 253–255.Freund, R., Z. Garfunkel, I. Zak, M. Goldberg, T. Weissbro, and B. Derin (1970), The shear along the Dead Sea Rift, Philos. Trans. R Soc. London

Ser. A-Math. Phys. Sci., 267, 107–130.Frieslander, U. (2000), The structure of the Dead Sea transform emphasizing the Arava, using new geophysical data [in Hebrew, English

Abstract], PhD thesis, pp. 43–53, Hebrew Univ., Jerusalem, Israel.Galloway, J. (2011), Late Holocene paleoclimate reconstruction of the Northern Gulf of Aqaba using foraminifera as a proxy, MS thesis,

Geosciences Dept., UMKC, 75 p.Garfunkel, Z. (1970), The tectonics of the western margins of the southern Arava [in Hebrew, English Abstract], PhD thesis, p. 96,

Hebrew Univ., Jerusalem, Israel.Garfunkel, Z. (1981), Internal structure of the Dead Sea leaky transform (rift) in relation to plate kinematics, Tectonophysics, 80, 81–108.Garfunkel, Z., I. Zak, and R. Freund (1981), Active faulting in the Dead Sea rift, Tectonophysics, 80, 1–26.Gerson, R., S. Grossman, R. Amit, and N. Greenbaum (1993), Indicators of faulting events and periods of quiescence in desert alluvial fans,

Earth Surf. Processes Landforms, 18, 181–202.Ginat, H., Y. Enzel, and Y. Avni (1998), Translocated Plio-Pleistocene drainage systems along the Arava fault of the Dead Sea transform,

Tectonophysics, 284, 151–160.Ginzburg, A., J. Makris, K. Fuchs, and C. Prodehl (1981), The structure of the crust and upper mantle in the Dead Sea rift, Tectonophysics, 80,

109–119.Gomez, F., G. Karam, M. Khawlie, S. McClusky, P. Vernant, R. Reilinger, R. Jaafar, C. Tabet, K. Khair, and M. Barazangi (2007), Global Positioning

Systemmeasurements of strain accumulation and slip transfer through the restraining bend along the Dead Sea fault system in Lebanon,Geophys. J. Int., 168, 1021–1028.

Haberland, C., N. Maercklin, D. Kesten, T. Ryberg, C. Janssen, A. Agnon, M. Weber, A. Schulze, I. Qabbani, and R. El-Kelani (2007), Shallowarchitecture of the Wadi Araba fault (Dead Sea Transform) from high-resolution seismic investigations, Tectonophysics, 432, 37–50.

Harding, T. P., and J. D. Lowell (1979), Structural styles, their plate tectonic habitats and hydrocarbon traps in petroleum provinces, AAPG Bull.,63, 1016–1058.

Hartman, G. (2012), Quaternary evolution of a transform basin: The northern Gulf of Eilat/Aqaba, PhD dissertation, 108 pp., Tel Aviv Univ.,Tel Aviv, Israel.

Hazan, N., M. Stein, A. Agnon, S. Marco, D. Nadel, J. F. W. Negendank, M. J. Schwab, and D. Neev (2005), The late quaternary limnologicalhistory of Lake Kinneret (Sea of Galilee) Israel, Quat. Res., 63, 60–77.

Hofstetter, A. (2003), Seismic observations of the 22/11/1995 Gulf of Aqaba earthquake sequence, Tectonophysics, 369(1-2), 21–36.Klinger, Y., L. Rivera, H. Haessler, and J.-C. Maurin (1999), Active faulting in the Gulf of Aqaba: New knowledge from theMw 7.3 earthquake of

22 November 1995, Bull. Seismol. Soc. Am., 89(4), 1025–1036.Klinger, Y., J. P. Avouac, N. Abou Karaki, L. Dorbath, D. Bourles, and J. L. Reyss (2000a), Slip rate on the Dead Sea transform fault in northern

Araba valley (Jordan), Geophys. J. Int., 142, 755–768.Klinger, Y., J. P. Avouac, L. Dorbath, N. A. Karaki, and N. Tisnerat (2000b), Seismic behaviour of the Dead Sea fault along Araba valley Jordan,

Geophys. J. Int., 142, 769–782.Klinger, Y., R. Michel, and J.-P. Avouac (2000c), Co-seismic deformation during theMw 7.3 Aqaba earthquake (1995) from ERS-SAR interferometry,

Geophys. Res. Lett., 27(22), 3651–3654, doi:10.1029/1999GL008463.Le Béon, M., Y. Klinger, A. Abdel Qader, A. Agnon, L. Dorbath, G. Daer, J.-C. Ruegg, O. Charade, and O. Mayyas (2008), Slip rate and locking

depth from GPS profiles across the southern Dead Sea Transform, J. Geophys. Res., 113, B11403, doi:10.1029/2007JB005280.Le Béon, M., Y. Klinger, M. Al-Qaryouti, A.-S. Mériaux, R. C. Finkel, A. Elias, O. Mayyas, F. J. Ryerson, and P. Tapponnier (2010), Early Holocene

and Late Pleistocene slip rates of the southern Dead Sea Fault determined from10Be cosmogenic dating of offset alluvial deposits,

J. Geophys. Res., 115, B11414, doi:10.1029/2009JB007198.Le Béon, M., Y. Klinger, A.-S. Mériaux, M. Al-Qaryouti, R. C. Finkel, O. Mayyas, and P. Tapponnier (2012), Quaternary morphotectonic mapping

of the Wadi Araba and implications for the tectonic activity of the southern Dead Sea fault, Tectonics, 31, TC5003, doi:10.1029/2012TC003112.Mahmoud, S., R. Reilinger, S. McClusky, P. Vernant, and A. Tealeb (2005), GPS evidence for northward motion of the Sinai Block: Implications

for E. Mediterranean tectonics, Earth Planet. Sci. Lett., 238, 217–224Makovsky, Y., A. Wunch, R. Ariely, Y. Shaked, A. Rivlin, A. Shemesh, Z. Ben Avraham, and A. Agnon (2008), Quaternary transform kinematics

constrained by sequence stratigraphy and submerged coastline features. The Gulf of Aqaba, Earth Planet. Sci. Lett., 271, 109–122.Mann, P., M. R. Hempton, D. C. Bradley, and K. Burke (1983), Development of pull-apart basins, J. Geol., 91, 529–554.Mansoor, N. M., T. M. Niemi, A. Misra (2004), A GIS-Based assessment of liquefaction potential of the city of Aqaba, Jordan, Environ. Eng. Geosci.,

10(4), 297–320Marco, S. (2007), Temporal variation in the geometry of a strike-slip fault zone: Examples from the Dead Sea Transform, Tectonophysics, 445,

186–199.Marco, S., M. Stein, A. Agnon, and H. Ron (1996), Long term earthquake clustering: A 50,000 year paleoseismic record in the Dead Sea graben,

J. Geophys. Res., 101, 6179–6192, doi:10.1029/95JB01587.McClay, K., and T. Dooley (1995), Analog models of pull-apart basins, Geology, 23, 711–714.Mitchum, R. M. J., and P. R. Vail (1977), Seismic stratigraphy and global changes of sea level, Part 7: Stratigraphic interpretation of seismic

reflection patterns in depositional sequences, in Seismic Stratigraphy—Applications to Hydrocarbon Exploration, vol. 26, edited byC. E. Payton, pp. 135–144, Memoir of the American Association of Petroleum Geologists, Tulsa.

Montaggioni, L. F. (2005), History of Indo-Pacific coral reef systems since the last glaciation: Development patterns and controlling factors,Earth Sci. Rev., 71, 1–75.

Moustafa, Y. A., J. Pätzold, Y. Loya, and G. Wefer (2000), Mid-Holocene stable isotope record of corals from the northern Red Sea, Int. J. EarthSci., 88, 742–751.

Niemi, T. M., and A. M. Smith (1999), Initial results of the southeastern Wadi Araba, Jordan geoarchaeological study: Implications for shifts inLate Quaternary aridity, Geoarchaeology, 14, 791–820.

Niemi, T. M., H. W. Zhang, M. Atallah, and J. B. J. Harrison (2001), Late Pleistocene and Holocene slip rate of the Northern Wadi Araba fault,Dead Sea Transform, Jordan, J. Seismol., 5, 449–474.

Ostrovsky, E. (2005), The G1 geodetic-geodynamic network: Results of the G1 GPS surveying campaigns in 1996/1997 and 2001/2002,Geophys. Inst. of Israel.

Pinar, A., and N. Türkelli (1997), Source inversion of the 1993 and 1995 Gulf of Aqaba earthquakes, Tectonophysics, 283, 279–288.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 22

Page 23: Quaternary tectonic evolution of the Northern Gulf of Elat/Aqaba along the Dead Sea Transform

Porat, N., A. G. Wintle, R. Amit, and Y. Enzel (1996), Late Quaternary earthquake chronology from luminescence dating of colluvial and alluvialdeposits of the Arava Valley, Israel, Quat. Res., 46, 107–117.

Porat, N., R. Amit, E. Zilberman, and Y. Enzel (1997), Luminescence dating of fault-related alluvial fan sediments in the southern Arava Valley,Israel, Quat. Sci. Rev., 16, 397–402.

Posamentier, H. W., M. T. Jervey, and P. R. Vail (1988), Eustatic controls on clastic deposition I—Conceptual framework, in Sea-Level Changes:An Integrated Approach, edited by C. K. Wilgus et al., SEPM Spec. Publ., 42, 109–124.

Quennell, A. M. (1959), Tectonics of the Dead Sea rift. 20th Int. Geol. Congr. Assoc. Serv. Geol. Afr. 1956, Mexico, 385–405.Quennell, A. M. (1984), The Western Arabia rift system, in The Geological Evolution of the Eastern Mediterranean, edited by J. E. Dixon and

A. H. F. Robertson, Geol. Soc. London Spec. Publ., 17, 775–788.Rahe, B., D. A. Ferrill, and A. P. Morris (1998), Physical analog modeling of pull-apart basin evolution, Tectonophysics, 285, 21–40.Rangin, C., X. Le Pichon, E. Demirbag, and C. Imren (2004), Strain localization in the Sea of Marmara: Propagation of the North Anatolian Fault

in a now inactive pull-apart, Tectonics, 23, TC2014, doi:10.1029/2002TC001437.Reches, Z., J. Erez, and Z. Garfunkel (1987), Sedimentary and tectonic features in the northwestern Gulf of Elat, Israel, Tectonophysics, 141, 169–180.Rohling, E. J., K. Grant, C. Hemleben, M. Kucera, A. P. Roberts, I. Schmeltzer, H. Schulz, M. Siccha, M. Siddall, and G. Trommer (2008), New

constraints on the timing of sea level fluctuations during early to middle marine isotope stage 3, Paleoceanography, 23, PA3219,doi:10.1029/2008PA001617.

Rotstein, Y., U. Frieslander, and Y. Bartov (1994), Detailed seismic imaging of the Dead Sea transform in Elat. Geophysical Institute of Israel(in Hebrew, English Abstract).

Sade, A., J. K. Hall, G. Tibor, T. M. Niemi, Z. Ben-Avraham, A. A. Al-Zoubi, G. Hartman, E. Akawwi, A.-R. Abueladas, and G. Amit (2008), Themultinational bathymetric survey: Northern Gulf of ‵Aqaba/Elat Poster, Isr. J. Earth Sci., 57, 139–144.

Sadeh, M., Y. Hamiel, A. Ziv, Y. Bock, P. Fang, and S. Wdowinski (2012), Crustal deformation along the Dead Sea Transform and the CarmelFault inferred from 12 years of GPS measurements, J. Geophys. Res., 117, B08410, doi:10.1029/2012JB009241.

Schattner, U., and R. Weinberger (2008), A mid-Pleistocene deformation transition in the Hula basin, northern Israel: Implications for thetectonic evolution of the Dead Sea Fault, Geochem. Geophys. Geosyst., 9, Q07009, doi:10.1029/2007GC001937.

Shackleton, N. J., and N. D. Opdyke (1973), Oxygen isotopes and paleomagnetic stratigraphy of equatorial Pacific core V 28–238: Oxygenisotope temperature and ice volume on 10

5year and 10

6year scale, Quat. Res., 3, 33–35.

Shaked, Y., A. Agnon, B. Lazar, S. Marco, U. Avner, and M. Stein (2004), Large earthquakes kill coral reefs at the northwest Gulf of Aqaba, TerraNova, 16, 133–138.

Shaked, Y., B. Lazar, S. Marco, v Stein, and A. Agnon (2010), Late Holocene events that shaped the shoreline at the northern Gulf of Aqabarecorded by a buried fossil reef, Isr. J. Earth Sci., 58, 355–368.

Shtivelman, V., U. Frieslander, E. Zilberman, and R. Amit (1998), Mapping shallow faults at the Evrona playa site using high-resolutionreflection method, Geophysics, 63, 1257–1264.

Siddall, M., E. J. Rohling, A. Almogi-Labin, C. Hemleben, D. Meischner, I. Schmelzer, and D. A. Smeed (2003), Sea level fluctuations during thelast glacial cycle, Nature, 423, 853–858.

Siddall, M., D. A. Smeed, C. Hemleben, E. J. Rohling, I. Schmelzer, and W. R. Peltier (2004), Understanding the Red Sea response to sea level,Earth Planet. Sci. Lett., 225, 421–434.

Siddall, M., E. Bard, E. J. Rohling, and C. Hemleben (2006), Sea-level reversal during Termination II, Geology, 34(10), 817–820.Siddall, M., E. J. Rohling, W. G. Thompson, and C. Waelbroeck (2008), Marine isotope stage 3 sea level fluctuations: Data synthesis and new

outlook, Rev. Geophys., 46, RG4003, doi:10.1029/2007RG000226.Sims, D., D. A. Ferrill, and J. A. Stamatakos (1999), Role of a ductile decollement in the development of pull-apart basins: Experimental results

and natural examples, J. Struct. Geol., 21, 533–554.Slater, L., and T. M. Niemi (2003), Ground-penetrating radar investigation of active faults along the Dead Sea Transform and implications for

seismic hazards within the city of Aqaba, Jordan, Tectonophysics, 368, 33–50.Stein, R. S., G. C. P. King, and J. B. Rundle (1988), The growth of geological structures by repeated earthquakes. 2. Field examples of

continental dip-slip faults, J. Geophys. Res., 93, 13,319–13,331, doi:10.1029/JB093iB11p13319.ten Brink, U. S., M. Rybakov, A. S. Al-Zoubi, M. Hassouneh, U. Frieslander, A. T. Batayneh, V. Goldschmidt, M. N. Daoud, Y. Rotstein, and J. K. Hall

(1999), Anatomy of the Dead Sea Transform; does it reflect continuous changes in plate motion?, Geology, 27, 887–890.ten Brink, U. S., M. Rybakov, A. S. Al-Zoubi, and Y. Rotstein (2007), Magnetic character of a large continental transform: An aeromagnetic

survey of the Dead Sea Fault, Geochem. Geophys. Geosyst., 8, Q07005, doi:10.1029/2007GC001582.Tibor, G., T. Niemi, Z. Ben-Avraham, A. Al-Zoubi, R. Sade, J. Hall, G. Hartman, E. Akawi, A. Abueladas, and R. Al-Ruzouq (2010), Active tectonic

morphology and submarine deformation of the northern Gulf of Eilat/Aqaba from analyses of multibeam data, GeoMar. Lett., 30, 561–573.Twiss, R. J., and E. M. Moores (1992), Structural Geology, W. H. Freeman and Co., New York.Vail, P. R., R. M. Mitchum Jr., and S. Thompson (1977), Seismic stratigraphy and global changes of sea level, Part 3, relative changes of sea level

from coastal onlap, in Stratigraphic Interpretation of Seismic Data, edited by C. E. Payton, Mem. Am. Assoc. Pet. Geol., 26, 63–82.Wachs, D., and E. Zilberman (1994), Preliminary evaluation of the seismic hazard in the Elat area [In Hebrew], Geology Survey of Israel,

Jerusalem, p. 60.Wdowinski, S., Y. Bock, G. Baer, L. Prawirodirdjo, N. Bechor, S. Naaman, R. Knafo, Y. Forrai, and Y. Melzer (2004), GPS measurements of current

crustal movements along the Dead Sea Fault, J. Geophys. Res., 109, B05403, doi:10.1029/2003JB002640.Wheeler, H. E. (1958), Time stratigraphy, Am. Assoc. Petrol. Geol. Bull., 42, 1047–1063.Wu, J. E., K. McClay, P. Whitehouse, and T. Dooley (2009), 4D analogue modelling of transtensional pull-apart basins, Mar. Pet. Geol., 26,

1608–1623.Zak, I., and R. Freund (1966), Recent strike-slip movements along the Dead Sea rift, Isr. J. Earth Sci., 15, 33–37.Zhang, P., B. C. Burchfiel, S. Chen, and Q. Deng (1989), Extinction of pull-apart basins, Geology, 17, 814–817.Zilberman, E., R. Amit, N. Porat, Y. Enzel, and U. Avner (2005), Surface ruptures induced by the devastating 1068 AD earthquake in the

southern Arava valley, Dead Sea Rift, Israel, Tectonophysics, 408, 79–99.

Journal of Geophysical Research: Solid Earth 10.1002/2013JB010879

HARTMAN ET AL. ©2014. American Geophysical Union. All Rights Reserved. 23