Bulletin of the Seismological Society of America, Vol. 101, No. 1, pp. 39–67, February 2011, doi: 10.1785/0120100097

Episodic Behavior of the Section of the Dead Sea Fault Inferred from a 14-ka-Long Integrated Catalog of Large Earthquakes by Matthieu Ferry,* Mustapha Meghraoui, Najib Abou Karaki, Masdouq Al-Taj, and Lutfi Khalil

Abstract The continuous record of large surface-rupturing earthquakes along the Dead Sea fault brings unprecedented insights for paleoseismic and archaeoseismic research. In most recent studies, paleoseismic trenching documents the late Holocene faulting activity, while tectonic geomorphology addresses the long-term behavior (>10 ka), with a tendency to smooth the effect of individual earthquake rupture M >7 events ( w ). Here, we combine historical, archaeological, and paleoseismic investigations to build a consolidated catalog of destructive surface-rupturing earth- quakes for the last 14 ka along the left-lateral Jordan Valley fault segment. The 120- km-long fault segment limited to the north and the south by major pull-apart basins (the Hula and the Dead Sea, respectively) is mapped in detail and shows five subseg- ments with narrow stepovers (width < 3 km). We conducted quantitative geomor- phology along the fault, measured more than 20 offset drainages, excavated four trenches at two sites, and investigated archaeological sites with seismic damage in the Jordan Valley. Our results in paleoseismic trenching with 28 radiocarbon datings and the archaeoseismology at Tell Saydiyeh, supplemented with a rich historical seis- mic record, document 12 surface-rupturing events along the fault segment with a mean interval of ∼1160 yr and an average 5 mm=yr slip rate for the last 25 ka. The most complete part of the catalog indicates recurrence intervals that vary from 280 yr to 1500 yr, with a median value of 790 yr, and suggests an episodic behavior for the Jordan Valley fault. Our study allows a better constraint of the seismic cycle and re- lated short-term variations (late Holocene) versus long-term behavior (Holocene and late Pleistocene) of a major continental transform fault.

Introduction The occurrence of large earthquakes on continental berman et al., 2005; Haynes et al., 2006). Paleoseismic faults holds crucial questions on their physical and mechan- studies along the Jordan Valley fault (JVF) revealed the ical characteristics, their size, and their time distribution in episodic activity from a long-term earthquake record terms of magnitude and frequency. Recent field investiga- (50 ka) in the Lisan lacustrine deposits (El-Isa and Mustafa, tions in paleoseismology and archaeoseismology along the 1986; Marco et al., 1996; Migowski et al., 2004) and from Dead Sea fault (DSF) show evidence of historical coseismic the correlation between cumulative stream offsets and 48-ka- surface rupturing at the Sicantarla Tell in Turkey (Amik long paleoclimatic fluctuations (Ferry et al., 2007). This epi- basin; Altunel et al., 2009), the Al-Harif Roman aqueduct sodic activity is expressed not only by periods of earthquake in Syria (Meghraoui et al., 2003), the Lebanese restraining clusters affecting a single segment but also by a sequence of bend (Gomez et al., 2003; Daëron et al., 2004; Nemer and earthquakes on different segments during a short period of Meghraoui, 2006; Daëron et al., 2007; Nemer et al., 2008), time (Ambraseys, 2004; Sbeinati et al., 2005). The long-term the Jordan Valley gorge and Hula basin (Ellenblum et al., record of past earthquakes contributes to better understand the faulting behavior and stability of segment boundaries 1998; Marco et al., 2003; Marco et al., 2005), the Jordan (Sieh, 1996), as well as fault interactions during earthquake Valley (Reches and Hoexter, 1981), and the Wadi Araba (Zil- sequences (Stein et al., 1997). Although earthquake-induced soft-sediment deformations were largely studied in the Lisan *Also at Institut de Physique du Globe, Strasbourg, France. formation, earthquake surface ruptures associated with the

39 40 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

JVF needed a detailed paleoseismic and archaeoseismic system exhibits a relatively simple geometry with large study in order to document the earthquake sequence and pull-apart basins distributed along strike (the Red Sea, the related seismic cycle on a single fault segment. Dead Sea, the Hula basin, the Ghab basin, and the Amik The DSF forms the boundary between the African and basin) and a single major restraining bend at its center Arabian plates and accommodates ∼1 cm=yr of relative left- (Mounts Lebanon and Anti-Lebanon). It is composed of lateral strike-slip motion (Quennell, 1981; Fig. 1). The fault eight major segments (Fig. 1a), all of which are capable

Figure 1. (a) General map of the system. Numbers are geological slip rates (in black) and geodetic strain rates (in white). Sources: Klinger et al. (2000); Niemi et al. (2001); Meghraoui et al. (2003); Reilinger et al. (2006); Ferry et al. (2007). Pull-apart basins: ab, Amik basin; gb, Ghab basin; hb, Hula basin; ds, Dead Sea. Major fault segments: EAF, East Anatolian fault; AF, Afrin fault; KF, Karasu fault; JSF, Jisr Shuggur fault; MF, Missyaf fault; YF, Yammouneh fault; ROF, Roum fault; RAF, Rachaya fault; SF, Serghaya fault; JVF, Jordan Valley fault; WAF, Wadi Araba fault. (b) Detailed map of the JVF segment between the Sea of and the Dead Sea. The segment itself is organized as six 15-km to 30-km-long right-stepping subsegments limited by 2-km to 3-km-wide transpressive relay zones. The active trace of the JVF continues for a further ∼10 km northward into the Sea of Galilee (SG) and ∼20 km southward into the northern Dead Sea (DS). The color version of this figure is available only in the electronic edition. Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 41 of producing large destructive earthquakes and surface fault- ing as documented by an extended historical record (e.g., Guidoboni et al., 1994; Ambraseys and Jackson, 1998; Sbei- nati et al., 2005; Ambraseys, 2009). However, other than the M 1995 w 7.3 Aqaba earthquake, none of the main segments has released a large earthquake in the last eight to nine cen- turies (Fig. 2). The lack of seismicity and elapsed time since the most recent historical earthquakes suggest that a tectonic loading has been accumulating along most of the DSF. Global Positioning System plate velocities along the fault system suggest 2:5–6 mm=yr (Fig. 1a; also, McClusky et al., 2003; Wdowinski et al., 2004; Reilinger et al., 2006; Gomez et al., 2007; Le Beon et al., 2008; Alchalbi et al., 2010) that are comparable to 4–7 mm=yr geological slip rates (Garfun- kel et al., 1981; Ginat et al., 1998; Klinger et al., 2000; Niemi et al., 2001; Meghraoui et al., 2003; Gomez et al., 2003; Daëron et al., 2004; Akyüz et al., 2006; Ferry et al., 2007; Karabacak et al., 2010; Sbeinati et al., 2010) measured at 2-to-100-ka time scales. It implies 3–5 m of slip deficit for the different segments and suggests an increasing potential for destructive events in the near future. After a geology, tectonic geomorphology, and seismicity setting, we present the paleoseismic investigations with four trenches across the fault and the archaeoseismic studies at 11 different sites, with a specific focus on the Tell Saydiyehda- mages. Our results yield an integrated catalog of faulting events and related large earthquakes for the last 14 ka. The earthquake catalog of faulting events that includes both clus- tering and quiescence periods sheds light on the distribution of interseismic periods and the associated tectonic-loading process. The long-term faulting behavior of the JVF and its relationship to neighboring segments is also discussed.

Geological Setting

The north–south-trending DSF transform (Fig. 1)is made of a transtensional system to the south (including the Hula, Dead Sea, and Gulf of Aqaba pull-apart basins), the Lebanese restraining bend (the Yammouneh, Rachaya, Seghaya, and Roum faults) in the middle, and a strike-slip system to the north (the Missyaf fault and the Ghab pull-apart basin). The seismicity described by Aldersons et al. (2003) suggests that the lower crust has a brittle behavior below 20 km and possibly as deep as 32 km. This is supported by thermomechanical modeling (Petrunin and Sobolev, 2006), which suggests the brittle part of the cold lithosphere beneath the Dead Sea basin may be locally as thick as 27 km. Figure 2. Seismicity of the Dead Sea Transform system. M ≥4 Similarly, heat-flow measurements indicate relatively low Instrumental events with from 1964 to 2006 (IRIS Data Management Center; see Data and Resources section) in filled values not typical of a rifting region (Ben-Avraham et al., circles. Background seismicity is very scarce and mainly restricted M 1978). Furthermore, Ryberg et al. (2007) image subvertical to the Lebanese Bend and the Jordan Valley. The 1995 w 7.3 major faults and deep sedimentary basins along the Wadi Aqaba earthquake and aftershock swarm dominate the seismicity Araba segment. Although the level of background seismicity of the Red Sea basin. Historical events with I0 ≥ VII (Ambraseys M<5 and Jackson, 1998; Sbeinati et al., 2005) in open circles. Apart from is low ( ), the seismotectonic characteristics along the M the 1927 w 6.2 earthquake, no significant event has plate boundary show a systematic pattern of strike-slip focal occurred along the JVF since A.D. 1033 (see text for details). mechanisms and some normal faulting solutions (Salamon The color version of this figure is available only in the electronic et al., 2003). The DSF is a typical continental transform fault edition. 42 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil and exhibits a narrow deformation zone dominated by strike- In the middle section of the JVF, at the Tell Saidiyeh area, slip faulting that affects a thick cold crust. The post-Miocene small stream offsets and abandoned beheaded channels left-lateral offset along the fault is estimated to be ∼45 km, provide an ideal site for faulting event characterization and which is consistent with the accumulation of more than 8 km slip rate calculations (Fig. 3d). We performed a detailed of clastic, carbonatic, and evaporitic sediments in the Dead microtopographic survey (Fig. 3d) of the site in order to study Sea basin (Quennel, 1984; ten Brink, 1993; Ginzburg and the cumulative offsets along the fault revealed by the drainage Ben-Avraham, 1997; Bartov et al., 2006). At a large scale, system. These stream channels expose the fault and exhibit a the DSF is highly segmented, and the JVF section is limited conspicuous shear zone with evidence of recent faulting by the Lebanese restraining bend to the north and the large (Fig. 4b). On the eastern block, the drainage system consists Dead Sea pull-apart basins to the south. of a small catchment area to the north and a single linear stream (E1) flowing from the east with a 60° N–80° N direction that Qt Tectonic Geomorphology along cuts into an abandoned alluvial terrace ( 0). Because of the the Jordan Valley Fault catchment area, the northern bank of that stream has been eroded and modified, and only the southern bank is adequately The active JVF is made of five 15-to-30-km-long subseg- preserved. Against the fault, stream E1 flows into a marshy ments (Fig. 1b; Al-Taj, 2000; Malkawi and Alawneh, 2000; water hole that may be partly man-made based on an existing Ferry et al., 2007) limited by relatively small (2-to-3-km- depression. On the western block, two strongly incising wide) transpressive and transtensive relay zones. Using aerial gullies flow westward to the valley. The northern one (W1) (1:25000 scale) and satellite photographs (SPOT-5, Landsat continues west of the water pond (E1) along a 60° N direction 7, Google Earth), field investigations, and offset measure- and displays a subtle offset of 7  0:5 m. The southern one ments, we mapped in detail a total of 120 km of the active (W2), while being very well expressed, has no counterpart fault trace from the Hula basin to the Dead Sea. The active east of the fault where a small catchment area has been formed fault trace is visible within the valley and cuts through the by regressive erosion. Hence, the only possible source for W2 former Lake Lisan (65–18 ka B.P.) and clastic Damya is E1, making W2 a beheaded remnant left-laterally offset by (18 ka B.P. to present) deposits. The geomorphology of the 114  5 m. Because W1 and W2 cut into the upper surface of valley shows regionally flat Lisan lacustrine terraces display- the Lisan formation, they necessarily are younger than its ing an average slope of ∼0:1° toward the present-day Dead ultimate deposits; that is, they are younger than 25 ka B.P. Sea. The very fine-grained Lisan sediments (mostly varve- (Abed and Yaghan, 2000). Furthermore, a trench exposure like detritus and aragonite), combined with a semiarid (see the Paleoseismology section, Fig. 4a, and Fig. 5) reveals climate produce classical badland morphology, where mate- channel deposits related to Qt0, which have been subse- rial is eroded away through a dense dentritic gully network. quently radiocarbon dated at 19,700–16,800 B.C. (see sample The hydrographic system provides clear markers for the L-02 AkR fraction in Table 1). Considering that sample L-02 study of left-lateral cumulative offsets along the fault trace originates from the middle of the stratigraphic section (see (Ferry et al., 2007; Ferry and Meghraoui, 2008; Klein, 2008). Fig. 5), the associated age is a minimum value for the empla- The JVF displays complex transtensional features with cement of the channel, which suggests a minimum age of 22 ka numerous 100-to-300-m-long and 50-m-wide pull-apart for W1 and W2. Additionally, because the drainage source is basins (Fig. 3) separated by right-stepping en echelon ruptures limited to the surface of the late Lisan terrace, we may adopt (fig. 3 in Ferry et al., 2007). Located in the southern section of the approach developed by Ferry et al. (2007) in the same re- the JVF, the impressive Ghor Katar badland area is made of gion and assume the inception of that drainage was triggered numerous stream incisions that expose ∼50-m-thick Damya by an abrupt lake-level drop of Lake Lisan. The present eleva- and Lisan lacustrine units in remarkable cliffs (Abed and tion of the terrace at Tell Saidiyeh is ∼255 m below sea level, Yaghan, 2000). The fault is well visible in the many incisions which, from the lake-level curve of Bartov et al.(2002) and of Ghor Katar before it enters the flat lacustrine terrace. inferences by Ferry et al. (2007), yield an age of 21–25 ka Located 2.5 km south of Ghor Katar, the Ghor Kabed B.P. for the latest level drop below that elevation and indepen- pull-apart system (Fig. 3c) affects the Damya and Lisan dently confirms our age inference. Taking into account the dat- terraces and illustrates the pattern of active faulting in the ing of channels and terraces, we infer that the total left-lateral valley. The accumulation of late Pleistocene and Holocene offset of 114  0:5 m has been accumulated during the last deposits in the Ghor Kabed pull-apart depressions constitutes 22–25 ka. The resulting long-term average slip rate is 4:9  a good record of past faulting events. Indeed, a detailed 0:3 mm=yr for that period, which is in good agreement with a microtopographic survey (1–2-m resolution) illustrates the previous geological slip rate obtained from 20 offset streams basin morphology and related north–south-trending 6-m- (Ferry et al., 2007). high fault scarps with 4–15° slope cutting through the middle of the depocenter. The clear fault scarp morphology and Instrumental Seismicity active alluvial and lacustrine sedimentary processes present a good potential for paleoseismic trenching at this site (see The Dead Sea fault exhibits scarce instrumental seis- the Paleoseismology section). micity with mostly low to moderate earthquakes (M<6, Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 43

Figure 3. Geomorphology of the Jordan Valley fault. (a) Central section of the JVF (see location on Fig. 1b), showing drainage (outlined) that is systematically left-laterally displaced at the passage of the fault. The active trace of the JVF is pointed out by the arrows. (b) Geo- morphology of the Tell Saidiyeh site from a high-resolution total station topographic survey (contour spacing 0.5 m). South of the archae- ological tell (located ∼100 m to the north, see inset in Fig. 8b), the morphology displays a recent terrace strath (Qt0) affected and left-laterally displaced by the fault. The southern edges (dashed lines) of streams serve as piercing points because they are less likely to be eroded than the northern ones in a left-lateral setting. Stream E1 flows westward along the southern edge of Qt0 and is displaced by 7  0:5 m across the fault. Stream W2 is a beheaded remnant of E1 and displays 114  5 m of offset. A minimum emplacement age of 22 ka for W2 yields an average slip rate of 4:9 mm=yr for that period (see text for details). Solid rectangles represent trenches T3 and T4 (see text for descriptions), which display faulting evidence for the last 17 ka. Blanked areas could not be surveyed due to the presence of agricultural and military facilities. (c) Geomorphology of the Ghor Kabed site from a high-resolution total station topographic survey. The eastern fault strand shows a linear and continuous geometry with a gentle slope (the steep slope visible to the north is artificial), while the western strand displays a steeper slope and a left-step geometry. Two trenches were excavated at that site: T1 on the central strand north of the depression, and T2 on the eastern strand southeast of the depression (see logs in Fig. 5). Height curve spacing is 0.25 m. (d) The Ghor Kabed site displays a very subtle morphology that could not have been fully deciphered without high-resolution topography. The color version of this figure is available only in the electronic edition. 44 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

Figure 4. Field photographs of the Tell Saidiyeh and Ghor Kabed sites showing the shear zones (thick lines). (a) Oblique view of trench T4 showing Lisan units affected by faulting and liquefaction in the foreground and sand and silt units in the background. (b) The main shear zone is outlined by a 1-cm-thick layer of crushed sand (thick lines) and displays numerous oriented pebbles. (c) Seismites affecting aragonite and detritus layers from Lisan units as observed at a nearby roadcut. (d) Main shear zone in trench T2 affecting Lisan (massive clay) and Damya (clastic) units. Thin lines are stratigraphic contacts, thick lines are faults, and dashed lines mark the bottom of the trench. (e) Main shear zone in trench T1. (Legend as in part d.) Letters in parentheses in (d) and (e) refer to units in Figure 5. The color version of this figure is available only in the electronic edition. psdcBhvo fteJra alyScino h edSaFault Sea Dead the of Section Valley Jordan the of Behavior Episodic

Figure 5. Trench logs with lithological descriptions. (a) Trench T1 shows a distributed pattern of vertical faults that may be resolved within the uppermost layers but cannot be followed through massive clay units of Lisan age. Radiocarbon dates suggest the most recent event occurred before A.D. 1490–1800. (b) Trench T2 displays a main fault zone filled with breccia that have been ruptured afterward and documents the most recent event, radiocarbon dated after A.D. 560–660. Combined, these observations suggest two surface-rupturing events occurred at Ghor Kabed between A.D. 560 and A.D. 1800, which may be related to the A.D. 749 and A.D. 1033 events. (c) The exposure of T3 is mainly composed of Lisan sediments. A series of fine- gained colluvial and alluvial units overlays Lisan clays and provides insight on recent events. (d) Trench T4 is originally a road cut that was noticeably extended and cleaned. It is oriented ∼45° to the fault, which widens the deformation zone. This exposure provides the bulk of the paleoseismic dataset. See text for details. (e) Correlations of stratigraphic sections of trenches. The geological formations of Lisan and Damya are common basement-bottom units for trenches. Erosion processes (tilde lines) have major effects on soft sediments, and Trench T3 shows a significant hiatus of the Damya formation. The correlation between alluvial and lacustrine deposits and the related radiocarbon dating (see also Table 1 and Fig. 7) illustrate the different recent 45 depositional environments at trench sites. The color version of this figure is available only in the electronic edition. (Continued) 46 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

(e) a AD 1660-1950 b AD 1690-1920 cd (AD 87-319) ~~~~~~e 1610-1410 BC f 5470-5110 BC g h 5060-4770 BC 7500-4400 BC i1 ~

11460-11270 BC i2 11600-10910 BC

j 10200-8800 BC (17655-16475 BC) k 1 ~ k2

a AD 1490-1800 a modern l c AD 560-660 d 450-790 BC b1 a1-3 m e 1320-1520 BC b2 b c n f c d d AD 1490 - 1640 ~ ~~~~~~ ~~~~~~ e g 11790-11310 BC f 12170-11720 BC h ~~~~~~ e o Damya i Damya j 11200-12000 BC k f g Lisan ~~~~~~ l g

Lisan 14700-16300 BC p T3 Damya T1 T2

q Lisan

1 m T4

Figure 5. Continued.

Fig. 2a) mainly concentrated along the Lebanese Bend and Amman, Ramallah, , the Mount of Olives (“a mile the Jordan Valley fault with three recorded moderate east of ”), Reineh (), As Salt, and Jericho M M earthquakes (11 July 1927 w 6.2; 23 April 1979 b 5.2; (Fig. 6 for location). The event was recorded at more than M and 2 February 2004 L 5.2). However, the DSF is capable 100 seismological stations throughout the world, and its of producing large destructive events, as attested by the 22 epicenter was located within the northern basin of the Dead M November 1995 w 7.3 Aqaba earthquake (Hofstetter, Sea (Shapira et al., 1993). On the basis of reinterpreted 2003) and by large historical events (Fig. 2b; Ambraseys historical documents, Avni et al. (2002) confirm these find- and Jackson, 1998; Sbeinati et al., 2005). ings and describe a seiche wave in the Dead Sea that pleads The 11 July 1927 event is the most destructive earth- for either a submarine landslide or a displacement of bathy- quake to strike the region in the last century, with a known metry along a surface rupture. Submersible images of the toll reaching 285 people killed and ∼1000 injured. Wide- bottom of the Dead Sea (Lazar and Ben-Avraham, 2002) spread destruction was documented by Willis (1928) in show an apparently fresh and sharp scarp continuing the Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 47

Figure 6. Archaeology of the Jordan Valley. White squares, main populated areas cited in historical documents; white dots, archae- ological sites visited and reappraised in this study; gray dots are archaeological sites not studied here (lack of evidence and/or available literature) but of potential interest for future studies; gray squares, paleoseismic sites; black squares, geomorphological sites studied by Ferry et al. (2007). The color version of this figure is available only in the electronic edition.

Jordan Valley fault into the Dead Sea. It should be considered Sea basin. The 2 February 2004 event (epicenter 31.69° N, M that the 1927 w 6.2 earthquake may have produced surface 35.58° E) was felt strongly in Jordan, , Palestine, and rupture locally with tenuous displacement, in the range of a Syria and produced slight damage in Jordan and Israel, injur- few tens of centimeters (Wells and Coppersmith, 1994). ing a total of 20 persons (Jordan Seismological Observatory, The 23 April 1979 event (epicenter 31.24° N, 35.46° E) 2004). The combined analysis of aftershock distribution and was extensively instrumentally and macroseismically studied fault plane solution suggests a normal fault branch perpendi- by Arieh et al. (1982). It reached Io ˆ IV–V MSK (though cular to the JVF, probably associated with the Dead Sea isoseismal lines are available for Israel only) in the Jordan Val- pull-apart (Al-Tarazi et al., 2006). One may notice that the ley and according to Arieh et al. (1982), its focal mechanism instrumental seismicity does not reflect the level of active de- points to a 20° N-striking border fault of the northeastern Dead formation of the JVF and its potential for large earthquakes. 48 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

Historical Seismicity the Mount of Olives, southeast of Jerusalem, without a clear causative link. The historical seismicity relies on inscriptions and docu- ments from Greek, Hebrew, Roman, Byzantine, Arabic, and Additionally, numerous other earthquakes have been felt Ottoman times (Guidoboni et al., 1994; Ambraseys, 2009). A in the Jordan Valley in historical times but may not be con- wealth of testimonies, invoices, and reports are available for sidered as candidates for surface-rupturing events along the last two millennia and document one of the most complete the JVF: historical catalogs to date. However, it should be noted that the • One may consider a candidate earthquake in A.D. 418 (not region was not evenly populated at any time in the past, with A.D. 419, as justified by Ambraseys, 2009). However, densely populated areas along the Mediterranean coast and evidence is very weak because contemporaneous chroni- the shores of the Sea of Galilee and the Dead Sea and barren clers (Marcellinus Comes, Philostorgius; see also Ambra- areas in the Negev desert and Wadi Araba. This situation seys, 2009) and archaeological investigations (Meyers induces a systematic bias into intensity maps by (a) shifting et al., 1976) describe earthquake damage in the north of epicenters northward and (b) restricting the extent of felt tes- Galilee region. There is no mention of major damage or timonies. Furthermore, magnitudes derived from historical victims in Jerusalem, Jericho, and Palestinian villages studies are usually very delicate to assess and carry a large located in the Dead Sea vicinity in A.D. 418. (and undefined) uncertainty, especially for older events. • The A.D. 363 earthquake is better documented and consists Recent findings by Katz and Crouvi (2007) show that anthro- of a sequence of two shocks on 18 and 19 May (Ambra- pogenic tali of archaeological origin have very poor geotech- seys, 2009). Although a large number of sites in Palestine, nical characteristics and may amplify seismic shaking. including Jerusalem, were damaged and a sea wave was M >8 observed in the Dead Sea, further south half of Petra Previous works suggest that S historical earthquakes possibly occurred in the Jordan Valley (Ambraseys and Jack- was razed to the ground, and localities near the Red Sea son, 1998). However, the length of fault segments and were badly damaged (Niemi and Mansoor, 2002). ∼250 thickness of the seismogenic crust suggest that M 7.2–7.4 The wide region of damage and the -km-long Wadi w Araba fault from the Dead Sea to the Gulf of Aqaba is a reasonable maximum magnitude in this region. to the south may well be the site of two major shocks Based on historical seismicity catalogs and recent litera- of A.D. 363. ture, we list here the main earthquake events that occurred in • Historian Flavius Josephus (A.D. 37–100) vividly describes the Jordan Valley (Fig. 2b): an event in 31 B.C. in the Jewish Wars: “For in the early • Event ZH (A.D. 1033, 10 Safar 425 A.H.): According to Ibn spring, an earthquake shock killed an infinite number of Al-Jawzi (A.D. 1113–1200), on that morning the walls of cattle and 30 thousand people; but the army was unharmed, Jerusalem crumbled down during construction, and the because it was camped in the open” (Guidoboni et al., 1994, cities of Jerusalem, Ramallah, Jericho, Nablus, and pp. 173–174). A critical and exhaustive reappraisal of that were heavily damaged (Abou Karaki, 1987). This event by Ambraseys (2009) suggests that Flavius Josephus’ event was felt throughout Judea and possibly as far away as account—the only coeval source—is greatly exaggerated, Egypt and Syria, and it produced a sea wave along the with a number of casualties larger than the actual population Mediterranean coast (Fig. 6 for locations; Poirier and of the region. The author also concludes that previously Taher, 1980; Ambraseys et al., 1994). reported damage to archaeological structures is actually • Event YH (A.D. 749): Theophanes (A.D. 760–818), a his- spurious and not associated with earthquake faulting. In torian whose work constitutes one of the main sources for summary, a small to moderate earthquake probably oc- that period, narrates “a powerful earthquake in Palestine, curred in 31 B.C. but did not produce significant damage along the river Jordan and throughout Syria, and countless to buildings or surface rupture. As mentioned by Ambraseys thousands of people were killed, and churches and mon- (2009, p. 101), “The reappraisal of the available data reveals asteries also collapsed, especially in the desert near the nothing more than that the 31 B.C. earthquake in Judaea that Holy City [Jerusalem]” (Ambraseys, 2009, p. 232). This caused damage and loss of life, which Josephus grossly ex- event has been intensively studied by numerous authors, aggerates. There is no evidence that Jerusalem was affected due in great part to inconsistencies between calendars and the destruction or damage of other historical sites in Ju- (Abou Karaki, 1987; Tsafrir and Foerster, 1992). daea is conjectural and cannot be tested on archeological • Event XH (759 B.C.): This earthquake produced great ground. The association of the earthquake with a fault break destruction and many casualties in Judea, Samaria, and at Khirbet Qumran seems to me untenable and I can find no Galilee (Guidoboni et al., 1994). A thorough reappraisal justification for the addition of Diospolis to towns affected of this event by Ambraseys (2005) indicates that few con- and the dating of the event to A.D. 31 (Guidoboni et al., temporary accounts are available for this event, the earliest 1994; Guidoboni, 1989).” one being the Book of Amos. The first detailed description • Previous studies consider an event in 64 B.C. that would is given by Zachariah around 520 B.C. (i.e., ∼240 years have damaged the Temple in Jerusalem and been felt later) and suggests that a large landslide developed on throughout the region and as far away as Antioch (south- Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 49

east Turkey). However, the critical reinterpretation by sand (unit b in T1 and T2) visible mainly in the shear zones Karcz (2004) strongly suggests an opposite situation, with overlain by scattered caliche and organic soil (unit a in T1 an event source near Antioch (probably along the Hacipasa and T2). fault or the Karasu fault) and related seismic shaking felt as far as Jerusalem. This is supported by the fact that only Trench 1 minor damage was reported in the Jordan Valley. Indeed, In trench T1 (Fig. 5a), ruptures are distributed over the this event is most likely the 65 B.C. Antioch earthquake section east of the main fault zone (Fig. 4e) and affect Lisan that claimed some 170,000 lives (Guidoboni et al., and Damya deposits. All upward fault terminations corre- 1994; Sbeinati et al., 2005). spond to the base of the present-day plow unit (unit a) and do not show clear indications for a chronology. However, Paleoseismology at the contact between Lisan/Damya and Holocene deposits, the faulted units correspond to a narrow fissure filled by In order to establish a correlation between seismites and pieces of unit b. Unit a, which covers the shear zone and historical large events, Marco et al. (1996), Ken-Tor et al. corresponds to an organic soil, has been dated at A.D. 1490– (2001), and Migowski et al. (2004) have taken advantage 1800, postdating the most recent faulting event ZT1, which of varvelike deposits in the Dead Sea around the southern possibly corresponds to the A.D. 749 or the A.D. 1033 tip of the JVF and the northern tip of the Wadi Araba fault. earthquakes. These authors claim an almost complete record of M>5:5 earthquakes for the last 50 ka. However, due to erosion Trench 2 and depositional hiatuses, Ken-Tor et al. (2001) could not identify the A.D. 1033 and A.D. 749 events in their varve In trench T2, the surface rupture consists of several fault section. These events, as well as the 31 B.C. and 759 B.C. branches in a 2.5-m-wide shear zone (Fig. 4d and Fig. 5b) events, were identified by Migowski et al. (2004) in a that shows ∼1 m of total apparent vertical separation, with different varve section. Furthermore, Migowski et al. the western block being the footwall. Here, abutting relation- (2004) identify three additional events in ∼1100 B.C., ships permit the identification of four events: ∼2100 B.C., and ∼2700 B.C. (labeled events 36, 42, and • Event ZT2: The most recent event observed in trench T2 is 43) for which historical data must be supplemented with ar- associated with surface ruptures that affect finely lami- chaeology and paleoseismology. nated unit b2, unit b1, and possibly c, d, and e with fault Paleoseismic studies bring evidence for surface ruptures splays terminating at the base of the top unit a. This event is that can be correlated with historical and prehistorical events. necessarily younger than unit b1, radiocarbon-dated While historical and instrumental data do not point to a A.D. 560–660, and may be associated with the historical specific seismic source, active faulting studies and paleoseis- A.D. 749 earthquake and/or the A.D. 1033 earthquake. mology may provide a direct observation of the causative • Event YT2: The event is attested by the formation of a fault. In order to perform successful paleoseismological inves- 1.5-m-wide flower structure filled with breccia (b1) and tigations, we carefully selected trench sites to ensure an opti- stratified silty clay (unit b2). In case unit b1 is a fissure mal expression of faulting events, a continous and detailed fill, the event would have taken place shortly before the sedimentary record, and material suitable for age determina- deposition of unit b1 (i.e., shortly before A.D. 560–660). tions. In the following subsections of this paper, we describe However, if unit b1 is composed of preexisting layers four trench exposures (Fig. 5), for which deposit chronologies affected by this event, it may then have occurred after are constrained by 28 radiocarbon samples (Table 1 the deposition of unit b1, which would naturally point and Fig. 7). to the A.D. 749 earthquake. In that latter case, event ZT2 Two trenches were dug across the fault scarps that limit would correspond to the A.D. 1033 earthquake. the northernmost pull-apart basin along the fault at Ghor • Event XT2: This event is documented by two fault splays Kabed (Fig. 3). Trenches are east–west-trending and dug affecting unit d up to the base of unit c, as well as by a across the eastern fault section (trench T1) and across the ∼40-cm-long vertical liquefaction dyke affecting unit d western fault section (trench T2). The trenches expose the with its source in the underlying sandy unit f. The scarcity fault zone and related lacustrine (Lisan) and clastic (Damya of available datable material does not allow us to date that formation and Holocene) stratigraphic units. The stratigraphy event accurately other than earlier than sixth century A.D. in the two trenches (Fig. 5) is similar and made of (1) laminated • Event WT2: The oldest event that may be observed in gray detritus (unit l in T1 and f in T2) visible at the trench trench T2 is attested by a Y-shaped rupture affecting units bottom, which can be correlated with the upper Lisan forma- g, f, and the base of unit e at the eastern end of the trench. tion; (2) a succession of 1.2-m-thick intercalated sandy This event could be contemporaneous with the deposition and detritus layers (units k–g in T1) that corresponds to the of unit e (Damya formation). Damya formation; (3) silt and sand units (units f–cinT1 and f, d, and c in T2) that belong to the Holocene pull-apart The burial of unit b1 and related shear zone by unit deposits; and (4) mixed units with detritus, silty-sand, and a in the two trenches (Fig. 5e) indicates a bracket of 50 Table 1 Radiocarbon Dating of Samples Collected in Trenches T1 and T2 (Ghor Kabed Site) and T3 and T4 (Tell Saidiyeh Site)*

Sample Laboratory Depth Amount of Carbon Radiocarbon Calibrated Trench Name Code (m) Material Fraction† (AoC, in mg) δ13C(‰) Age (B.P.) Date 2σ Range‡ Observations T1 Ka-T3-C04 KIA 24314 0.15 Sprig AkR 5.4 16:06 280  30 1490 1800 T1 Ka-T3-C05 KIA 24315 1.15 Charcoal HA 2.9 25:38 11600  50 12000 11200 T1 Ka-T3-S02 KIA 24317 0.45 Carbonate Carb 1.2 7:92 3160  30 1520 1320 T1 Ka-T3-S03 KIA 24318 0.25 Carbonate Carb 0.8 8:52 2500  20 790 450 T2 Ka-T2-1N KIA 24305 0.5 Charcoal AkR 5.3 26:16 1440  20 560 660 T2 Ka-T2-2N KIA 24306 0.15 Peat AkR 5.8 23:63 Modern - - T2 Ka-T2-6N KIA 24309 1.5 Dust layer AkR 0.3 24:85 14570  250 16300 14700 Low carbon content; subject to contamination T2 Ka-T2-9N KIA 24311 0.2 Organic mat AkR 4.7 25:67 Modern - - T3 Tol-01 KIA29719 0.3 Charcoal pieces in sand AkR 3.4 23:55 335  20 1490 1640 T3 Tol-14 KIA29720 0.2 Wood pieces in AkR 0.2 Modern - - Mixture of prebomb sandy soil and postbomb carbon T4 L-02 KIA29707 3.1 Carbonate pieces Carb 1.5 4:35 15920  70 17655 16475 Age difference between in sandy soil Carb and AkR fractions T4 L-02 KIA29707 3.1 Carbonate pieces AkR 0.2 24:59 16950 ‡ 570=  530 19700 16800 not statistically significant; in sandy soil confirms quality of the date T4 L-06 KIA29715 2.1 Charcoal dust in sand AkR 0.5 29:81 11630  120 11790 11310 T4 L-07 KIA29708 2.2 Charcoal dust in sand AkR 0.6 28:40 12010  100 12150 11720 Age difference between T4 L-07 KIA29708 2.2 Charcoal dust in sand HA 2.4 25:36 12125  50 12170 11880 AkR and HA fractions not statistically significant; Khalil L. and Al-Taj, M. Karaki, Abou N. Meghraoui, M. Ferry, M. confirms quality of the date T4 L-14 KIA29701 0.3 One small seed AkR 0.4 21:41 118  55 1660 1950 14C age plateau; calibration of charcoal piece is undetermined T4 L-15 KIA29702 0.3 Seed (pumpkin) AkR 4.5 25:87 Modern - - Mixture of prebomb embedded in the soil and postbomb carbon T4 L-21 KIA29717 1.9 Charcoal dust in sand HA 2.9 26:00 11455  45 11460 11270 T4 L-22 KIA29718 1.9 Charcoal pieces and AkR 0.8 29:20 11040  70 11150 10910 dust in sand T4 L-22 KIA29718 1.9 Charcoal pieces and HA 4.1 24:54 11560  50 11600 11320 dust in sand T4 Tbc-04 KIA30479 2.1 Sand with charcoal AcidR 0.3 27:31 9940  180 10200 8800 Low carbon content; pieces subject to contamination T4 Tbc-16 KIA29709 0.2 Sand with plant AkR 3.8 23:34 50  25 1690 1920 Age difference between T4 Tbc-16 KIA29709 0.2 Sand with plant HA 5.0 23:54 85  20 1690 1919 AkR and HA fractions not statistically significant; confirms quality of the date T4 Tbc-16 KIA29709 0.2 Small snail Carb 2.5 4:83 1825  30 87 319 Significantly older snail shell, probably reworked T4 Tbc-18 KIA 30480 1.7 Sand with charcoal AkR 0.04 n.a. 7000  700 7500 4400 Low AoC, normalized to pieces specially prepared OxII-targets T4 Tbc-23 KIA29711 0.9 Bulk sandy soil AkR 0.6 23:87 6015  55 5060 4770 Low carbon content; subject to contamination

(continued) psdcBhvo fteJra alyScino h edSaFault Sea Dead the of Section Valley Jordan the of Behavior Episodic Table 1 (Continued) Sample Laboratory Depth Amount of Carbon Radiocarbon Calibrated Trench Name Code (m) Material Fraction† (AoC, in mg) δ13C(‰) Age (B.P.) Date 2σ Range‡ Observations T4 Tbc-24 KIA29712 0.7 Bulk sandy soil AkR 0.6 26:82 6320  60 5470 5110 Low carbon content; subject to contamination T4 Tbc-26 KIA29714 0.3 Bulk sandy soil AkR 1.7 24:81 3210  35 1610 1410

*Detailed measurements presented here give proper insights regarding the quality (amount of carbon should be larger than 1 mg) of collected samples, the possibility of modern contamination (depth and fraction), and the adequacy of calibration (δ13C). †AkR, alkali residue; HA, humic acids; Carb, carbonate; AcidR, acid residue.. ‡ The 2σ calibrations were performed using the OxCal 3.10 software (Bronk Ramsey, 1995) and the INTCAL04 calibration curve from Reimer et al. (2004). 51 52 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

2900 BC2300 BC 1150 BC759 BC AD 749AD 1033

Events TT4 ZT2 BT4 V X YT4 Y UT4 T4 WT4 T4 T2 Samples AT4 CT4 ST4 ZT3

AD 1490-1800 BC 450-790 BC 1320-1520 Trench 1 Trench BC 11200-1200 AD 560-660 T2 BC 14700-16300

T3 AD 1490-1640 AD 1690-1920 AD 1690-1919 Phase Tbc-16 AD 1660-1950

AD 87-319 BC 1610-1410

BC 5060-4770

BC 5470-5110

BC 7500-4400

BC 10200-8800

Trench 4 Trench BC 11460-11270 BC 11150-10910

Phase L21-22 BC 11600-11320

BC 11790-11310 depositional hiatus BC 12170-11880 BC 12150-11720 pdf of calibrated date Phase L-07

BC 17655-16475 pdf of event BC 19700-16800

Phase L-02 hist./arch. event

20000CalBC 15000CalBC 10000CalBC 5000CalBC CalBC/CalAD

Calibrated date

Figure 7. Distribution of radiocarbon dates used in trenches 1–4 with inferred events, know historical earthquakes, and inferred archaeoseismic events. Gray boxes indicate depositional hiatuses where no date could be determined. All dates given herein and in Table 1 correspond to 2σ (95.4%) intervals on these probability density functions (pdf). Event pdfs are modeled for a Gaussian distribution on the basis of inferred uncertainties defined in Table 3.

A.D. 560–1800 of the last two faulting movements in Trench 3 the pull-apart area. Our interpretation is that the two post- Trench T3 was excavated perpendicular to the fault sixth century faulting events may be correlated with the across a 2-m-high scarp used for agricultural purposes A.D. 749 and 5 December 1033 large earthquakes in the Jordan Valley (Abou Karaki, 1987; Ambraseys and Jackson, (Fig. 3b). It exhibits a shallow network of subvertical faults 1998). spread over 2 m and affecting Lisan deposits overlain with a <60 Two further excavations were opened at Tell Saydiyeh thin ( cm) succession of Holocene units. The bottom of outside of the archaeological perimeter (see Fig. 3 for the trench (Fig. 5c) is composed of laminated aragonite and location), across the main fault scarp and into an alluvial detritus (unit g) typical of Lisan deposits. Stratigraphy is terrace. marked by the conspicuous aragonite layers and may be Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 53 resolved with difficulty when only detritus is present. Unit g possible perturbations (e.g., ancient cliff collapse causing is overlain with three groups of fine-grained deposits. A first artificial deformation, contamination of potential radiocar- group is located on the central section of the trench and is bon samples, dense vegetation on the top surface). Trench composed of (1) a ∼20-cm-thick layer of siltstone (unit f), T4 (Fig. 5d) displays a very well-expressed fault zone affect- (2) a 5-to-15-cm-thick brown paleosol (unit e), and (3) an ing late Pleistocene and Holocene units. West of the main 8-to-25-cm-thick layer of cemented silty sand (unit d). To the shear zone (FZ1), the deepest unit (q) is composed of finely west, a second unit truncates all other units and is composed laminated aragonite and very fine gray clayey sand that may of reworked material where remains of a wooden-soled boot, be attributed to the Lisan. Unit q is strongly affected by soft- rust-encrusted hinges, and large lumps of charcoal were sediment deformation, liquefaction (unit r), and minor fault- found. It is unclear whether this anthropic unit is backfill ing (lower part of unit q). The overlying unit p is composed from a small prior excavation or fissure fill. Indeed, the edges of massive clay with occasional pods of gravelly sand and of the unit do not display obvious signs of shearing, and no displays deformation bands. Following the detailed descrip- apparent vertical displacement is observed across that zone. tion by Abed and Yaghan (2000) of late Quaternary deposits The whole section is covered with a 20–30-cm-thick plow in the region, that specific stratigraphic contact may be attrib- zone (unit b). At the toe of the scarp, all units are truncated uted to the transition between Lisan (unit q) and Damya (unit by a ∼1-m-wide man-made channel composed of two layers p) dated by Abed and Yaghan at 16–15 ka B.P.. Unit p is of clay (units a3 and a2) at the bottom and a 40-cm-thick overlain with a series of alluvial units (n, m, and l) that dis- layer of clean gravels (unit a1) on top. play upward reverse grading. Unit n is composed of gray Because the scope of our study is limited to the most coarse sand with occasional pebbles and grades into sand recent continuous record of events, we opted to focus on against the fault zone. Unit m is orange-yellow coarse sand Holocene deposits and did not investigate Lisan units in with occasional large pebbles. Unit l is a clast-supported peb- details. Thus, we could identify two events in trench T3. ble conglomerate that displays strong imbrication. The group lies unconformably against unit p along an erosional surface • Event ZT3: This event affected unit e in two places, which and forms the northwestern edge of an alluvial channel. It is are east and west of a large modern root (see Fig. 5c). The itself overlain with a thin red paleosoil (unit c) that caps the absence of unit e west of the two fault splays suggests that main shear zone, an irregular dark clay unit (b), and a matrix- vertical displacement was larger than 15 cm on each splay. supported sandy conglomerate (unit a) that truncates all units Event ZT3 has likely occurred shortly before the deposition and forms the surface. of unit d dated A.D. 1490–1640 and probably corresponds Southeast of the main shear zone, Pleistocene units are to the A.D. 1033 earthquake. represented by a limited remnant of unit p that is observed • Event YT3: The oldest event recognized in trench T3 is at the southeastern-most end of the trench. Its top surface marked by the faulting of unit f, the oldest non-Lisan unit is erosional and overlain with a yellowish sandy clay unit (unit observed here. It has likely occurred between the deposi- o) that does not appear in the northern block. Unit o is overlain tion of units f and e. However, because event ZT3 cut with a regular 10-cm-thick unit (h) composed of yellow silt through the whole thickness of unit e while event YT3 with carbonate nodules that sits immediately underneath affects it partially, we assume that event YT3 occurred the topsoil unit. Units p, o, and h are affected by conjugate closer to the deposition of unit e and event ZT3 closer to faults that display more than 50 cm of apparent vertical dis- the deposition of unit d. placement. Northwestward, unit o crops out at the base of the It should be noted that the artificial fill unit at the wes- trench and is affected by a series of minor faults. The top sur- ternmost end of trench T3 may provide evidence for an extra face of unit o is deeply cut into by subsequent units. Unit n event. Indeed, considering the magnitude of the 1927 earth- forms a wedge against unit o that we interpret as the eastern quake, its location (Avni et al., 2002), as well as apparently edge of the channel described previously. At the same level, recent surface breaks at the bottom of the Dead Sea (Lazar the central part of the trench displays a different picture. and Ben-Avraham, 2002), a surface expression cannot be Indeed, the lowermost deposit is unit m—units p, o, and n ruled out for that event, with as much as 10–20 cm of co- do not crop out there and must be deeply buried—and is over- seismic displacement. It follows that the artificial fill unit lain with unit l, which pinches out against a major fault splay could be an associated fissure fill. (FZ3) to the southeast. It is then overlain with coarse-to-fine cemented sand units k2, k1, and j that exist only there. The common erosional top surface of units n and j is overlain Trench 4 by a group of sandy channel units (i1 and i2) that cut into units Trench T4 (see Fig. 3 for location) is actually a cut o and n and are overlain by unit h. Those three units extend realized during leveling works to extend a nearby field, as northwestward against the main shear zone (FZ1). While units mentioned by the field owner, and has a northwest–southeast n, i2, and i1 cannot be clearly followed inside of the shear trend (i.e., oblique to the fault). It was widened and cleaned, zone, a 50-cm-long section of unit h can be observed within and the first meter of material (perpendicularly to the expo- it and is vertically offset by ∼40 cm. Unit h is overlain with a sure surface) was removed from the whole section to avoid group of subhorizontal fine-grained units (g–d), composed of 54 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil varying amounts of yellow to brown cemented silt and clay. All units in the central section from m to e display minor Capping the fault, unit c may be followed eastward for offsets. Unit d caps the rupture and forms the event ∼30 cm. It is then confused with the topsoil. horizon. Event YT4 occurred between the deposition of The different units are affected by a complex network of units e and d and may be dated by samples Tbc-23 and faults and fissures. The main fault zone (FZ1 in Fig. 5d)is Tbc-26 (Table 1). This yields a wide window of occurrence 30–70 cm wide and displays densely packed gravels (prob- between 5060 B.C. and 1410 B.C. ably incorporated from unit l) with the long axis of pebbles • Event XT4: This event is marked by a fan-shaped network of oriented subvertically along the main shearing direction splays located immediately southeast of FZ2. Three splays (Fig. 4b). Where it affects sandy units, the main fault zone affect units h, g, and f and produce ∼5 cm of cumulated (FZ1) is outlined by a ∼1-cm-thick band of white pulverized vertical throw. They are consistently capped by a thin, silty sand. Three supplementary fault splays can be traced from clay unit. Event XT4 can be dated by bulk soil samples the bottom up to the shallowest units and are named FZ2, Tbc-23 and Tbc-24. The stratigraphically lower sample FZ3, and FZ4. They are ∼1 cm wide and filled with a 23 is dated 5060 B.C.–4770 B.C. and appears to be slightly brown-red silty clayey material that may originate from younger than sample 24 (dated 5470 B.C.–5110 B.C.), thus the uppermost soil units. Besides, a wealth of minor splays suggesting an age inversion. However, samples 23 and 24 and numerous fissures affect the central part of the section, only produced 0.6 mg of carbon and may therefore be between FZ1 and FZ3. It should be noted that the width of subject to contamination. Because samples are bulk soil, the fault zone is only apparent due to the obliquity of the contamination is probably related to exposure, manipula- trench with respect to the fault’s direction. Though not fol- tion, and storage conditions and should then be associated lowing standards of paleoseismic trenching, this situation with a rejuvenation process. This would imply that the does provide a better exposure and extended wall surface actual age of samples may be slightly older than the mea- to collect more observations and samples for dating. sured ones. In summary, this suggests that sample 23 should be somehow older and points to an occurrence shortly Starting with ZT4 as the most recent event, we identified a set of eight to nine surface-rupturing events affecting before the deposition of the thin, silty clay unit (i.e., between 5470 B.C. and 5000 B.C.). Event XT4 may also be documen- alluvial Holocene and Damya units, as well as three older ted by a fault splay located immediately west of FZ3, which events affecting Lisan deposits (see circled letters in Fig. 5d). affects the limit between units g and f. However, the fault The lack of visible stratification in the massive clays of unit p termination could not be pinpointed. prevents us from identifying deformation features and recon- • Event WT4: This event is documented by a fault splay structing the corresponding part of the faulting history. located ∼1 m northwest of FZ3. This splay affects all units • Event ZT4: This most recent event is illustrated by three up to unit h and is capped by unit g. Its occurrence time major splays (FZ1, FZ3, and FZ4) that affect the whole may be bracketed by samples Tbc-23/Tbc-24 and Tbc-18 stratigraphic section up to 20–30 cm below the present- and yields a window between 7500 B.C. and 5080 B.C. day surface. Vertical displacement can only be resolved Considering the stratigraphic position of event WT4 with on FZ3, where it reaches ∼5 cm. That rupture does not respect to samples, we propose the actual date is closer affect the shallowest units b and c. It has likely occurred to the age of sample Tbc-18 and is therefore probably com- after the deposition of unit d and before the deposition of prised between 7500 B.C. and 5500 B.C. unit c, thus yielding a time window between A.D. 87 and • Event VT4: This event is located on FZ3, a fault splay that A.D. 1920 (Table 1) and pleading for a historical event. broke during event ZT4. However, the bottom limit of unit i1 Because the A.D. 87 lower bracket is based on a snail shell displays about twice as much displacement as the bottom of that is significantly older than the surrounding soil, we unit h, thus suggesting that an event took place between the consider that the event occurred significantly closer to the deposition of units i1 and h. This yields a probable time of upper bracket; that is, more likely after ∼A.D. 500. occurrence between 11,600 B.C. and 4400 B.C. Considering However, from the available radiocarbon datings alone, the possible rejuvenation of sample Tbc-18 (total amounts it is not possible to decide if this exposed fault has experi- of carbon [AoC] of 0.04 mg; see Table 1) and the intermedi- enced rupture in A.D. 749 or A.D. 1033 or both. Alterna- ate stratigraphic position of event VT4, we estimate the tively, one may argue that unit c (dated A.D. 1660–1950) occurrence date between 10,900 B.C. and 7500 B.C. exhibits noticeable warping across the main fault zone with • Event UT4: This event is located along the same fault zone an apparent vertical deformation of ∼25 cm and that unit b as event XT4 but ∼0:8 m deeper. In that part of the section, thickens at the toe of the related scarplet into what may be a the faulting pattern is somehow more complex and more colluvial wedge. Age and dimensions of those features mature. Because the top limit of unit k2 is strongly sheared M ∼ 6 ∼60 correspond to a recent w earthquake, such as the and cumulatively offset by cm, we infer that an older 1927 Palestine earthquake. This interpretation is supported event took place there after the deposition of unit k1. Fault by the occurrence of a modern fissure fill unit in T3. terminations are unclear in that particular part but affect • Event YT4: This event is interpreted from small (a few some features that we correlate with the bottom limit of centimeters) displacements affecting units along FZ2. unit i2. Upward, the splays cannot be resolved, and it is Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 55

unclear whether they affect unit i1 or not. Several splays Events CT4, BT4, and AT4 all occurred during the deposi- seem to stop within unit i2 (see splays left of event symbol tion of Lisan units or shortly after, while they were still water U in Fig. 5d), suggesting event UT4 took place after the saturated. This dates all three events back to the end of the deposition of unit i2. This yields a time window between Lisan (shortly before 20 ka B.P.; Bartov et al., 2002). 11,600 B.C. and 10,200 B.C. Considering the stratigraphic To achieve the best possible characterization of events, location, we propose an event date between 11,500 B.C. we took advantage of outcropping site-wide and region-wide and 10,500 B.C. sedimentary formations (Fig. 5) and enhanced our analysis • Event TT4: This event is documented by a splay that using stratigraphic correlations across trenches at a given site originates from the main fault zone FZ1 and displays an and across sites. Hence, our four paleoseismic trenches, apparent dip of ∼45°. This splay strongly affects unit l opened at two sites along the central and southern sections (gravels) and displaces the bottom of unit k1 vertically by of the JVF, yield a total of 12 surface-rupturing events: 2 may 20 cm and its top surface by only a few cm, and it may be correlated to historical earthquakes (A.D. 1033 and be followed upward 50 cm into the base of unit j. The as- A.D. 749), and the remaining 10 are prehistoric. The oldest sociated event may hence be postdated by sample L-02. events identified (A, B, and C) are synchronous with the However, this sample has a relatively low AoC (mg), is latest Lisan units deposited between 17 ka B.P. and 20 ka composed of carbonate (susceptible to be reworked), and B.P. It should be noted that, due to sedimentary hiatuses, appears to be older than the stratigraphically lower and well- no event could be identified between the last historical earth- dated samples L-07 and L-06. Thus, we choose not to rely quakes and YT4 (5060 B.C.–1410 B.C.), thus leading to a on sample L-02. Consequently, we propose that event TT4 is significant gap in the record. Considering the proximity bracketed by samples L-21/L-22 and L-06, which yields an between the active trace of the JVF and archaeological sites, occurrence date between 12,060 B.C. and 10,910 B.C. the adequate time period, and the rich record, we propose to • Event ST4: A conspicuous group of minor fault splays rely on archaeoseismology to compensate for the incomplete affect a clearly defined subhorizontal layer within unit o paleoseismic data. in the southeast section of the trench, thus indicating that S event T4 occurred after the deposition of unit o. However, Archaeoseismology due to subsequent erosion of that unit and the emplacement of younger channel deposits, all fault splays have been cut Rift valleys throughout the world are common passage- and stop at the erosion surface. Considering that the older ways for human populations and are thus generally rich with channel composed of units n, m, and l has originally cut the archaeological heritage of former civilizations. The evo- into units o and p (and totally removed unit o from the lution and migration of human groups with respect to active western-most section) we may infer that event ST4 has ac- faults has hence long been established (King et al., 1994). tually occurred prior to the deposition of unit n. This strong Here, we first provide detailed evidence for earthquake-in- erosion process has erased part of the stratigraphic record duced destruction at Tell Saidiyeh, an archaeological site that and limits the availability of dating samples. Consequently, sits a few tens of meters from the active trace of the JVF event ST4 may only be defined as having occurred within (Fig. 3) and which has been studied by means of paleoseis- the same time windows as event TT4; that is, between mic excavations (trenches T3 and T4 in section 5). Addition- 10,910 B.C. and 12,060 B.C. It may be assumed that the ally, we present a critical reappraisal of published data about main channel fill is noticeably thicker than the exposed 20 archaeological sites scattered over the Jordan Valley and section, which would put event ST4 stratigraphically close neighboring regions (Fig. 3 and Fig. 6), which show varying to sample L-06. This would suggest that event ST4 oc- degrees of evidence for earthquake-related damage and curred closer to the lower end of the bracket; that is, possibly surface rupture. The archaeological evidence for presumably between 12,060 B.C. and 11,500 B.C. paleoearthquakes takes the form of destruction to buildings • Event CT4: This event is the only clear liquefaction event with frequent fires and consequent signs of site abandon- that was identified at that site. It is marked by a 0.5-by-1- ment. Thus, a widespread burnt layer with a noticeable m-large pocket of homogeneous, fine-grained, red, well- content of rubble, pottery shards, and ashes may be a good sorted carbonate sand surrounded by distorted detritus candidate for earthquake evidence. However, indication for and aragonite layers. Furthermore, the pocket truncates an an earthquake is generally reduced to signs of destruction older fault that may be attributed to an older event (see that can be instead related to regional wars, local raids, or Event BT4). even accidents (e.g., accidental fire caused by an unattended • Event BT4: This event is pointed out by a single fault splay oil lamp). In a recent review, Ambraseys (2006) underlines that cuts through Lisan units and was later truncated by the different caveats pertaining to the use of archaeological liquefaction from event CT4. evidence to identify past earthquakes and particularly to the • Event AT4: The oldest event visible in exposure T4 affects interpretation of toppled structures. We follow Ambraseys’s the lowermost Lisan laminae as a fan-shaped splay guidelines in making our interpretations of existing archae- network. All splays, as well as a nearby seismite feature, ological data from Tubb (1988 and 1998), Savage et al. are consistently truncated by a subsequent deposit. (2001, 2002, and 2003), and Franken (1989) and retain the 56 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

most robust evidence to identify destruction events that may • Event YA: This event is attested at the neighboring sites of be attributed to past earthquakes. Tell Saidiyeh and Tell Deir’ Alla, where the occurrence of For example, Tell Saidiyeh (Fig. 8) is located almost an earthquake in the early twelfth century B.C. leaves no exactly at the center of the Jordan Valley (Fig. 6), which doubt for Franken (1989) and J. N. Tubb (personal comm.. suggests the site would probably experience intense ground- 2005). At Tell Al’ Umayri, ∼30 km east of the JVF, shaking in the case of a major earthquake on the JVF. Tell contemporary destruction associated with a burn layer rich structures are common throughout the Middle East and are in broken vessels is documented by Savage et al. (2001).It composed of layers of successive settlement (spreading over should be noted that this event may be part of the “ ” a few centuries to several millennia), which may pile up to well-documented twelfth century B.C. earthquake storm reach 30–50 m elevation above the base level. Tell Saidiyeh studied by many authors and summarized by Nur and Cline ∼1150 has been studied for the last 60 years and been the object of (2000). Proposed date: B.C.. • X ’ systematic excavations since 1964 (Tubb, 1988). The site has Event A: This event is documented at Tell Abu en-Ni aj produced a wealth of artifacts that document rather contin- by Savage et al. (2003) as a series of ash layers offset by a fault splay. According to our mapping of the JVF (Fig. 6; uous occupation from the Chalcolithic (fourth millen- Al-Taj, 2000; Ferry et al., 2007), Tell Abu en-Ni’aj appears ium B.C.) up to the Roman period (Tubb, 1998). The tell to be west and off the main active trace by ∼1 km. Savage itself is composed of two distinct mounds (Fig. 8), with et al. (2003) do not provide details about the amount or noticeably different histories: an upper main tell where intensity of deformation along that fault, and it is presently elaborate structures were discovered (e.g., an olive-pressing not possible to decide whether the observed offsets are complex and a water staircase; see Fig. 8d) and a lower tell related to coseismic fault slip. At Khirbet Iskander, located that was mostly used as a burial ground during the Omayyad ∼30 km southeast of the JVF (Fig. 6), Savage et al. (2003) period. A compilation of indications for strong perturbations describe a destruction layer in great detail (Table 2), where to Tell Saidiyeh (Table 2) point to two specific strata fire was attested by burnt stones, ash layers, and charred for which destruction is significant: stratum L2, dated grain. Well-preserved wooden beams and human remains ∼2900 B.C., and stratum XII, dated 1150–1120 B.C. For both suggest a sudden possibly earthquake-related catastrophe. strata, the author mentions widespread damage with intense Proposed date: ∼2300 B.C. burning, collapsed walls, and broken potteries (Fig. 8c). An • Event WA: At Tell Saidiyeh, Tubb (1988) documents dense additional event may be linked to stratum VI, dated to the destruction debris associated with fragmentary and dis- middle of the eighth century B.C., and may correspond to turbed architecture and indicates these are consistent with the 759 B.C. Jericho earthquake (Nur and Cline, 2000). ground-shaking-related damage (J. N. Tubb, personal Indeed, Tubb (1998, p. 126) mentions that “houses of Stra- comm., 2005). At the nearby Tell el-Fukhar site, Savage tum VI were knocked down and leveled in preparation for et al. (2003) mention evidence of intense destruction to another major building programme.” The reason for such walls and infer a possible earthquake-related origin. Pro- a drastic solution may be the prior intense destruction of posed date: ∼2900 B.C. the tell by an earthquake with no possibility to re-use da- Overall, our compilation of archaeological studies on maged buildings. Other strata show evidence for destruction tells in the vicinity of the JVF strongly suggests the occur- or abandonment but could not be related to seismic shaking. rence of four events: in 759 B.C., ∼1150 B.C., ∼2300 B.C., In parallel, we compile existing data for 11 archaeologi- and ∼2900 B.C.. Here, the archaeological record intersects cal sites in the vicinity of the JVF showing a potential for past the historical record, as attested by the 759 B.C. earthquake earthquake damage (Fig. 6 and Table 2). However, we con- described in written documents. Furthermore, the archaeo- sider earthquake-induced damage in archaeological sites only seismic event XA can be correlated with the paleoseismic if it is attested by a minimum of two sufficiently distant sites event YT4 and compensates for the sedimentary hiatus and sites with evidence for surface rupture. The analysis of observed in paleoseismic trenches (period ∼1500 B.C. to damage to archaeological sites allows us to identify four ∼4500 B.C.). It should be noted that indications for surface events likely associated with large earthquakes along the JVF: breaks affecting tells may not necessarily be considered as

• Event ZA: This event is attested at Tell Deir’ Alla (Franken, faulting evidence. Indeed, tells are artificial mounds made 1989) and Tell Saidiyeh (Tubb, 1988) for the middle of the of heterogeneous material that includes rubble, dirt, and eighth century B.C. Available descriptions lack details, and architectural remains (Fig. 8a). As such, they are very sen- it is probable that the corresponding destruction has been sitive to gravitational collapse, especially under seismic indirectly associated with the well-known 759 B.C. Zechar- shaking. Hence, for sites that are not directly located across iah’s earthquake (Nur and Ron, 1996) without further age the main active trace of the JVF, we consider surface defor- determination. At Tell Saidiyeh, damage is not directly as- mation as secondary evidence in the sense of McCalpin sociated with an earthquake, but it is rather the subsequent (1998). Nonetheless, such indications may provide a relevant massive leveling of the site that suggests a catastrophic chronological constraint for a seismic event. Thus, the event. Proposed date: 759 B.C.. mention by Franken (1989, p. 203) of “a victim ‰…Š found Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 57

Figure 8. Tell Saydiyeh archaeological site. (a,b) General view of the site and satellite imagery (inset) showing its relationship to the fault. The archaeological site is composed with a lower and an upper tell that were occupied at different periods. The original site was a limited mound that looked over the region and has grown with the successive addition of settlement layers, each related to a specific period. The obvious proximity of the JVF is marked by the active fault scarp. In (b), the satellite imagery shows the relationship between the active fault trace (thick line), the archaeological site (LT, lower tell; UT, upper tell), and geomorphology (white outline represents the extent of the microtopographic survey in Fig. 3b). (c) Open pit at the top of the tell showing conspicuous ∼5-cm-thick black burnt layers. Those layers contain broken pottery, charred wood, and ashes and are remnants of a widespread intense fire. (d) The twelfth century B.C. olive processing area (Palace) that displays signs of destruction (from Tubb, 1998). (e) Blocked doorway and broken vessel interpreted as a direct result of earthquake shaking (from Tubb, 1988). The color version of this figure is available only in the electronic edition. 58 Table 2 Compilation of Archaeological Evidence for Strong Perturbations at Archaeological Sites in the Vicinity of the Jordan Valley Fault

Indication of Archaeological Date of Event Proposed Cause Probability of Surface Documented Date Period (Inferred) Site Description from Archaeological Source of Destruction an Earthquake Rupture ? Mid-fourteenth ∼ A:D: 1350 Tell Hisban “There is evidence of earthquake collapse and fire for Earthquake High century A.D. the mid 14th-century destruction that preserved the contents of the store.” (p. 446)* ? Ayyubid-Mamluk A.D. 1170–1520 Tell Ya’amun “The mosaic floor of the east room is extensively Earthquake dented by collapsed wall stones, which suggests that use ended with destruction caused by an earthquake.” (p. 457)† ZH (749 A.D.) A.D. 749 Khirbet Yajuz “In area E, an earthquake that occurred in A.D. 748 is Earthquake High illustrated by the collapsed vaulted arches and the irregularities of the paved floor, which date to the Umayyad period.” (p. 448)‡ ? Mid-seventh ∼A:D: 650 Tell Hisban “After an earthquake in the mid seventh century A.D., Earthquake High century A.D. which was responsible for the collapse of the stone barrel vaults, the structure was reoccupied and used into the Abbasid period.” (p. 446)* ? Late Byzantine A.D. 490–640 Jerash “The pottery and glass under this tumbled wall section Earthquake showed that the collapse must have occurred during the Late Byzantine period, probably the result of an

earthquake that was responsible for the destruction Khalil L. and Al-Taj, M. Karaki, Abou N. Meghraoui, M. Ferry, M. of other city buildings in the sixth century.” (p. 458)† ZA (759 B.C.) Seventh–sixth 500–700 B.C. Tell Saydiyeh “A series of building phases (IIIB–IIIG) was defined Fire Low century B.C. below Pritchard’s Stratum III (now termed IIIA), the lowest of which shows architecture very similar to Stratum V, with similar evidence for burning.” (p. 130)§ ∼720 B:C: Tell Saydiyeh “These stalls were frequently found to contain equid Accidental fire? Average bones, and it seems likely that these represent the War? remains of unfortunate animals which had been abandoned to the fire which brought an end to Stratum V around 720 BC. The destruction of Stratum V might have been accidental, but it might also be attributed to the Assyrians, who were campaigning in this region at the time.” (p. 127)§ Eighth 700–800 B.C. Deir ’Alla “Phase M was destroyed by earthquake and fire.” Earthquake, fire century B.C. (p. 204)∥ Mid-eighth ∼750 B:C: Tell Saydiyeh “Towards the middle of the eigth century the houses of Anthropic Average century B.C. Stratum VI were knocked down and leveled in postearthquake? preparation for another major building programme.” (p. 126)§ (continued) psdcBhvo fteJra alyScino h edSaFault Sea Dead the of Section Valley Jordan the of Behavior Episodic Table 2 (Continued) Indication of Archaeological Date of Event Proposed Cause Probability of Surface Documented Date Period (Inferred) Site Description from Archaeological Source of Destruction an Earthquake Rupture

YA (∼1150 B:C:) 1150–1120 B.C. Tell Saydiyeh “The city of Stratum XII was obviously destroyed in an Fire High intense conflagration, the dense associated debris sealing a valuable corpus of finds, examination of which has established a date for this event at around 1150-1120 BC, coinciding with the withdrawal of the Egyptian empire... Some time in the last quarter of the [twelfth] century the city of Stratum XII was destroyed by fire, and at the same time the cemetery on the Lower Tell fell out of use. There is no indication as to the source of the destruction: certainly there were neither bodies nor signs of conflict amidst the ruined buildings of the Upper Tell, and it could well be that fire was the result of an accident.” (p. 86)§ Early twelfth 1150–1100 B.C. Deir ’Alla “There is little doubt that the entire complex was Earthquake century B.C. destroyed early in the 12th c. BC and that an earthquake caused the destruction.” (p. 203)∥ Late first Iron Age 1200–1100 B.C. Deir ’Alla “This period ended again with an earthquake and a Earthquake Yes victim was found completely squashed in a crack in the earth.” (p. 203)∥ Early Iron Age 1200–1100 B.C. Tell al-’Umayri “In an adjacent building to the south, destruction debris ? covered a layer of burned and broken ceramic vessels.” (p. 440)‡ ? Middle Bronze ∼1600 B:C: Tell al-’Umayri “An earthquake distorted many of the north–south Earthquake Age walls from this period at the site, including some of those in the palace.” (p. 463)† XA (∼2300 B:C:) Early Bronze IV ∼2300 B:C: Tell Abu en-Ni’aj “[...] bulldozing on the western side of Tell Abu en- Earthquake High Yes Ni'aj cleared a 26-m-long stratigraphic cross-section, revealing an earthquake slip fault. Early Bronze IV deposits capped a series of offset ash layers, indicating that the earthquake occurred during the occupation of Tell Abu en-Ni'aj.” (p. 439)‡ Early Bronze III 2650–2350 B.C. Khirbet Iskander “[ .....] section clearly revealed the tip lines of burnt Earthquake?, Fire stones, series of ash layers, mudbricks, abd detritus, clarified that these remains represent one destruction phase. [...] the destruction debris [ ...... ] included restorable pithoi and wavy-handled vessels, well- preserved wooden beams, and quantities of charred grain, lentils, and peas. The charred bones of an almost complete human arm lay on the plaster floor.” (p. 541)# (continued) 59 60 Table 2 (Continued) Indication of Archaeological Date of Event Proposed Cause Probability of Surface Documented Date Period (Inferred) Site Description from Archaeological Source of Destruction an Earthquake Rupture

WA (∼2900 B:C:) Late early ∼2900 B:C: Tell Saydiyeh “Stratum L2 was found to be associated with dense Unknown High Bronze I destruction debris (ashes, burnt mud-brick rubble and charred timber), but both Strata L2 and L3 were apparently built on the same plan. In the centrally located area, excavations revealed extremely fragmentary and disturbed Early Bronze Age architecture, the remains having been all but obliterated by the digging of graves in the thirteenth to twelfth century BC.” (p. 42)§ Early Bronze IB 3300–2850 B.C. Tell el-Fukhar “The terrace walls were so damaged, presumably from earthquakes of which we found evidence, that they were difficult to distinguish from the huge stone falls around them.” (p. 462)†

*Savage et al. (2002). †Savage et al. (2003). ‡Savage et al. (2001). §Tubb (1998). ∥Franken (1989). #Savage et al. (2005). .Fry .Mgroi .Ao aai .A-a,adL Khalil L. and Al-Taj, M. Karaki, Abou N. Meghraoui, M. Ferry, M. Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 61 completely squashed in a crack in the earth” at Tell Deir’ Alla very weak structural relays (stepovers) within the JVF,it (∼3 km east of the JVF) and Savage et al.’s (2003) previously is likely that large earthquake ruptures are characteristic in “ ” ’ M – ∼3:3 mentioned earthquake slip fault at Tell Abu en-Ni aj are length and associated with w 7.2 7.4 and a m average not considered as strong evidence for surface rupture but coseismic displacement. An estimate of the total seismic may document off-fault effects of a large earthquake. moment release for the 12 faulting events (Z–O in Fig. 9) during the last 14 ka yields an ∼3:3 mm=yr slip rate, which Summary of Paleoseismic and probably indicates a lack in the paleoseismic record (possibly Archaeoseismic Events due to the sedimentary hiatus). By contrast, using the total seismic moment for the admittedly most complete part of the We determined 3 seismic events from historical studies catalog (events Z–U) over a period of ∼3:9 ka, we obtain an (ZH–XH), 4 damaging events from archaeological studies ∼4:2 mm=yr slip rate (Fig. 10), which is comparable to the (ZA–WA), and 12 faulting events from paleoseismic investi- slip rate obtained from offset streams and paleoclimatic fluc- gations (ZT2 and YT2, ZT3, ZT4–ST4, and CT4–AT4). The con- tuations (Ferry et al., 2007). The slight difference is easily sistency in the correlation between historical, damaging, and explained by distributed deformation around the main trace faulting events constrains the occurrence of the earthquake (possibly ∼0:5 m per event). events. Table 3 presents the resulting catalog of 15 large earth- Taken as a whole, our integrated catalog of past seismic quakes (Z–O and C–A) produced by the JVF between events yields a mean recurrence interval of 1165 yr (standard A.D. 1033 and 12,060 B.C. and between 15,000 B.C. and deviation 243 yr). A close examination reveals highly vari- 18,000 B.C. (Fig. 9). It should be noted that an additional event able recurrence intervals with values ranging from 284 yr to was identified in trench T2 (WT2) for which we could not es- 2700 yr (Table 3). To some extent, we agree that very high timate an age. The fusion of the different datasets reinforces values may reflect the incompleteness of our catalog, the identification of past earthquakes. Indeed, of the 13 paleo- especially for its purely paleoseismic part (events T–O). seismic events, two are also present in the historical record and However, long intervals are also present in the assumedly one (possibly two) in the archaeological record. One event is complete historical and archaeological parts, between events present in both historical and archaeological datasets. Y and X (1508 yr) and events W and V (1150 yr), which suggests very long quiescence periods are real. In parallel, Constraints on the Holocene Behavior short interval values are clearly observed through the whole of the Jordan Valley Fault catalog: between events Y and Z (284 yr), between events W and X (391 yr), and across events O, P, and Q (190–520 yr). With major barriers (the Dead Sea and the Hula basin) to We propose that these short intervals reflect clustering and a the rupture propagation toward nearby fault segments and generally episodic behavior of the JVF over the last 14 ka.

Figure 9. Events probability density functions for the last ∼20 ka along the JVF from historical, archaeological, and paleoseismic data. The average recurrence interval is 1165 yr (σ ˆ 243 yr) for the whole period but varies from 787 yr (σ ˆ 1212 yr) for historical and archaeological data to 1480 yr (σ ˆ 705 yr) for paleoseismic data. See text for a detailed analysis. The color version of this figure is available only in the electronic edition. 62 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

80

70

60

4.3 mm/yr 50

r /y 40 m m 11.2 mm/yr .3 8.3 mm/yr 8

30 mm/yr 4.9 11.2 mm/yr r Cumulative slip (m) Cumulative m/y 1.6 m

20 r yr m/y m/ m m 3.5 4.2 Ferry et al., 2007 10 This study

0 0 24681012 14 16 18 Age (ka BP)

Figure 10. Comparison between geomorphological data (gray curve, Ferry et al., 2007) and paleoseismological data (black curve, this study) for the last 14 ka. With an average coseismic displacement of 3.3 m (see text), inferred slip rate reaches 4:2 mm=yr for the last 5 ka, slightly higher than the minimum value and slightly lower than the average long-term value given by Ferry et al. (2007). Inferred cumulative slip from historical and archaeological events (19:8  1:8 m) satisfyingly corresponds to the first geomorphic marker (17  5 m). Furthermore, our oldest events (O, P, and Q) suggest a short-term slip rate of ∼8:3 mm=yr, comparable to the 11:2 mm=yr value from geomorphology.

This independently confirms the observation by Ferry et al. archaeoseismic events (759 B.C., 1150 B.C., 2300 B.C., and (2007) of slip rate variations along the JVF from a long-term 2900 B.C.). The oldest part of our catalog is based on paleo- value of 4:9 mm=yr to a peak value of 11 mm=yr during a 2- seismic trenching alone, which we strengthened by exploring ka-long period. any alternative interpretation. In trench T4, event VT4 is in- terpreted on the basis of increasing displacement with depth along FZ3 and related splays, suggesting that unit i1 has been Discussion affected by at least one extra event with respect to unit h. An The field investigations and collected data, their analysis, alternative explanation might be that unit h is isopach, while and results provide a remarkable succession of past earth- unit i1 is an erosive channel fill with a noticeable dip to the – quakes and related rupture parameters along a single segment west southwest (direction of flow). In this case, any horizon- of the DSF. We have identified several issues that can be sum- tal movement would produce apparent vertical displacement: marized in: (1) the complex alluvial stratigraphy (hiatus and all offsets described along FZ3 could have occurred during channeling) and related uncertainties on radiocarbon datings event ZT4 only. However, our interpretation is supported by a led to a partial identification of some paleoseismic events second observation close to the main shear zone (see labels in (e.g., A, B, C, R, and T); (2) the possibility that our paleo- Fig. 5d). Besides, the upper unit c appears to have been seismic record is short of a few events; and (3) the accuracy deposited over a preexisting scarp across the main shear zone in the identification of seismic clusters, episodes of quies- (FZ1) or to have been warped with ∼10 cm vertical displace- cence, and definition of interseismic periods. However, the ment (Fig. 5d). In the former case, the most recent displace- combination of paleoseismic trenching, historical studies, ar- ment would have occurred between A.D. 87 and A.D. 1920, chaeoseismic results, and their relationships allowed a most thus indicating either of the historical events of A.D. 749 and favorable characterization of paleoseismic events Z–V. A.D. 1033. In the latter case, warping would have taken place The analysis of paleoseismic results revealed nine fault- after the deposition of unit c and before the deposition of the ing events that we attempted to correlate (Table 3) with three wedge unit b (i.e., after A.D. 1660) at a time when no large historical events (A.D. 1033, A.D. 749, and 759 B.C.) and four earthquake is attested for the region, which is very unlikely. Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 63

Table 3 Summary of Events Identified from Historical, Archaeological, and Paleoseismic Data along the Jordan Valley Fault for the Last 18.5 ka

Event Records Date* Interval† Confidence Reference Z A.D. 1033 284  1 5

ZH A.D. 1033 Ambraseys et al., 1994; Amiran et al., 1994; Ben-Menahem, 1991 ZT2 Ult:>A:D: 560–660 x1;x3; 0† This study ZT3 Ult:A:D: 500 This study Y A.D. 749 1508  2 5

YH A.D. 749 Abou Karaki, 1987; Tsafrir and Foerster, 1992; Karcz, 2004 YT1 Pen:A:D: 560–660 This study X 759 B:C:  1 391  51 4

XH 759 B:C:  1 Ben-Menahem, 1991; Nur and Ron, 1996; Ambraseys, 2005 ZA Mid-eighth century B.C. Tubb, 1988; Franken, 1989 W YA 1150 B:C:  50 1150  100 3 Franken, 1989; Nur and Cline, 2000 V 2300 B:C:  50 600  100 4

XA ∼2300 B:C: Savage et al., 2003 3YKI70falFSivtJa

*Ult., ultimate; Pen., penultimate. †Time interval from preceding event.

Over the four trench exposures, we collected and dated 28 possible that some events are still missing. The mean recur- radiocarbon samples, most of which yielded large (> 1 mg) rence interval for our paleoseismic record is 1480 yr, almost amounts of carbon (AoC) and adequate uncertainties (20– double the 787 yr mean recurrence interval obtained for the 50 yr) and appeared to be in stratigraphic order (Table 1 assumedly complete historical and archaeological periods and Fig. 5). Some samples present uncertainties that should (i.e., the last 14,000 years). This may point either to the be clarified: Samples Ka-T3-S03 (T1), Ka-T3-6N (T2), L-14 incompleteness of the long-term paleoseismic record (one (T4), and Tbc-04 (T4) present low to insufficient AoC (0.2– extra event has been identified but not dated) or to different 0.8 mg) and were not used for event determination. Although faulting behaviors over the two periods. Considering the alkali residue (AkR) fractions of samples L-02 (T4) and L-07 situation of alluvial deposits in channels and the relatively (T4) yield low AoC, their respective carbonatic and humic long sedimentary hiatus observed in trenches (Fig. 5), it is acid fractions yield similar ages that confirm the quality of likely that part of the sedimentary record of paleoseismic the results. However, because sample L-02 is older than any events has been truncated. Furthermore, plotting age versus other sample of the section by some 5 ka, including the stra- cumulative displacement for our catalog (Fig. 10) reveals that tigraphically older samples L-07 and L-06, we consider it is the paleoseismic period does not fit cumulative offsets redeposited and chose to reject it. Sample Tbc-18 presents an measured on streams. This may be only partly explained extremely low AoC and needed to be normalized to specially by distributed off-fault deformation (Griffith et al., 2010). prepared targets. Because it is used to define event R (paleo- Our catalog of seismic events allows us to identify earth- seismic event VT4) and rejuvenation is very likely, we relied quake clusters and quiescence periods and suggests episodi- on nearby samples and stratigraphy to correct its age (see city for the JVF over the last 14 ka. It appears that the description, Trench 4). We used the same approach for sam- majority of events from our catalog (9 events out of 12) ples Tbc-23 and Tbc-24, which bracket event T (paleoseis- is grouped into clusters of two or three earthquakes: events mic event XT4) and show a relatively low AoC (0.6 mg each), Y and Z (cluster YZ, interval of 284 yr), events W and X similar ages, and age inversion. Although we applied com- (WX, interval of 391 yr), events U and V (UV, interval of pensation for rejuvenation of the samples, it is possible that 600 yr), and events O, P, and Q (OPQ, mean interval of events R and T are actually slightly older. This would move 330 yr). These clusters are generally preceded and followed event T closer to event S and event R closer to events O, P, by long periods of quiescence: up to 1800 yr after cluster and Q, while extending the empty period between events R OPQ, 2335 yr before cluster UV, 1150 yr between clusters and S. UV and WX, 1508 yr between clusters WX and YZ, and at Although we combined three independent datasets into a least 977 yr after cluster YZ (time since the A.D. 1033 his- uniquely long catalog of large earthquakes for the JVF,itis torical earthquake). Although one may also consider events 64 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

A, B, and C as a cluster, the large uncertainty associated with shows incompleteness with a mean recurrence interval of radiocarbon dating (1500 yr) implies that the three events 1480 yr, suggesting sedimentary hiatuses and a gap in the may have occurred any time within a 3-ka period. geological record. Considering that the inland 120-km-long JVF may Taking into account that the historical and archaeologi- extend southward inside the Dead Sea basin for a further cal datasets provide a mostly complete catalog of surface- ∼20 km (Lazar and Ben-Avraham, 2002; Ben-Avraham and rupturing events, we obtain a 787-yr mean recurrence inter- Schubert, 2006) and northward in the Hula asin for ∼10 km val, ∼3:3 m of slip per event (derived from Wells and Cop- (Marco et al., 2005), this yields a maximum ∼150 km sur- persmith, 1994), and a mean 5 mm=yr slip rate for the last face-rupture length. Hence, some of our inferences rely on a 5 ka (or 4:8 mm=yr, using a regression curve; see Fig. 10), in M – maximum w 7.2 7.4 magnitude and a related average agreement with the mean value obtained by Ferry et al. ∼3:3 m coseismic left-lateral slip estimated from a maximum (2007) for the last 48 ka. This is also confirmed by the direct 120–150-km surface rupture length (Kanamori and Ander- measurement of a stream offset at Tell Saydiyeh that yields son, 1975; Wells and Coppersmith, 1994). Although the 4:9  0:3 mm=yr for the last 25 ka (see Tectonic Geomor- M 22 November 1995 w 7.3 Aqaba event is the only large phology along the Jordan Valley Fault). Finally, the last mil- modern earthquake observed along the DSF until today, its lennium of seismic quiescence along the JVF indicates up to rupture parameters may not serve as a base of comparison 5 m of slip deficit and points to either an imminent earth- for earthquakes along the JVF. Indeed, it was composed quake and/or a future earthquake cluster similar to the A.D. of two or three subevents along offshore faults (Hofstetter, 749/A.D. 1033 sequence. Neighboring segments being at a 2003), and no surface rupture could be observed so far. slightly lower, if not similar, level of tectonic loading (i.e., Unfortunately, the only coseismic slip values available were no large events since the twelfth century sequence), it is plau- modeled and may not be adequately compared to surface sible that a present-day large earthquake on the JVF may rupture values (Hofstetter, 2003). However, several coseis- trigger earthquake ruptures on nearby segments. mic or cumulative displacements were actually measured along other segments of the DSF and typically show coseis- mic values around 3 m (e.g., Klinger et al., 2000; Gomez Data and Resources et al., 2003; Meghraoui et al., 2003; Marco et al., 2005; Haynes et al., 2006), which suggests that a similar value Digital elevation model data used to produce maps are may be considered for the JVF. Furthermore, results of recent from Shuttle Radar Topography Mission (SRTM) 3 topogra- M – earthquakes, such as the 1999 w 7.3 7.4 Izmit earthquake phy from the National Aeronautics and Space Administration (Turkey) and related right-lateral strike-slip faulting, yield a and are available from Consultative Group for International comparable estimate with ∼140 km rupture length and an Agriculture Research–Consortium for Spatial Information average 3 m coseismic slip (Barka et al., 2002). CGIAR–CSI Consortium for Spatial Information at srtm.csi .cgiar.org (last accessed March 2009). Bathymetry used for Conclusions maps in Figure 1 and Figure 2 is SRTM 30‡ produced by Uni- versity of California, San Diego (http://topex.ucsd.edu/). In- The JVF is the main source of destructive earthquakes for strumental seismicity in Figure 2 was obtained from the the Jordan Valley region. Its characterization has crucial Incorporated Research Institutions for Seismology Data Man- implications for the seismic hazard assessment of large urban agement Center at www.iris.edu (last accessed April 2007). areas such as Jerusalem, Amman, and Irbid, as well as to Calibration of radiocarbon ages was performed using the numerous historical and archaeological heritage sites such OxCal software (Bronk Ramsey, 1995), available at c14.arch as Pella, Jerash, Madaba, Qumran, Jericho, and Meguiddo .ox.ac.uk/embed.php?File=oxcal.html with the IntCal04 (Fig. 6). calibration curve (Reimer et al., 2004). Through an integrated approach involving (1) the com- pilation of historical seismicity, (2) the careful reappraisal of archaeological data, and (3) detailed fault mapping, tectonic Acknowledgments geomorphology, and paleoseismic trenching, we produce an original catalog of at least 12 surface-rupturing events in the The authors are indebted to the Deanship of Scientific Research (Uni- last 14 ka and at least 16 events in the last 17 ka along the versity of Jordan), the Jordan Valley Authority, the Natural Resources Authority and Military Commandment, and Salman Al-Dhaisat (Royal Jor- Jordan Valley segment of the Dead Sea fault. The mean danian Geographic Center) for their assistance and help during our field in- recurrence interval (787 yr) indicates that the historical vestigations. We are grateful to Majdi Barjous (Natural Resources and archaeological record might be complete, while lower Authority), Hani Amoush (University of Jordan), and Zoe Shipton and and upper bounds of the extreme recurrence intervals James Kirkpatrick (Glasgow University) for their help in the field and to (284–1508 yr) imply a generally time-episodic behavior. Pieter Grootes and Marie-Josée Nadeau (Kiel University) for the radiocar- bon dating. This study was funded by the EC 5th Framework Program and The plausibility of this episodic model should be compared APAME project (Contract ICA3-CT-2002-10024). Matthieu Ferry was in to apparent episodicity of conventional renewal models (Fit- part supported by the COMPETE program (FCT Ciencia 2007 FCOMP- zenz et al., 2010). In contrast, the paleoseismological dataset 01-0124-FEDER-009326). Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 65

References Bartov, Y., M. Stein, Y. Enzel, A. Agnon, and Z. E. Reches (2002). Lake levels and sequence stratigraphy of Lake Lisan, the Late Pleistocene Abed, A. M., and R. Yaghan (2000). On the paleoclimate of Jordan during precursor of the Dead Sea, Quat. Res. 57, 9–21. the last glacial maximum, Palaeogeogr. Palaeocl. 160, 23–33. Ben-Avraham, Z., and G. Schubert (2006). Deep “drop down” basin in the Abou Karaki, N. (1987). Synthése et carte sismotectonique des pays de la southern Dead Sea, Earth Planet. Sci. Lett. 251, 254–263. bordure orientale de la Méditerrannée: Sismicité du systéme de failles Ben-Avraham, Z., R. Hanel, and H. Villinger (1978). Heat flow through the du Jourdain—Mer Morte, Ph.D. Thesis, Université Louis Pasteur, Dead Sea rift, Mar. Geol. 28, 253–269. Strasbourg. Ben-Menahem, A. (1991). Four thousand years of seismicity along the Dead Akyüz, H., E. Altunel, V. Karabacak, and C. Yalciner (2006). Historical Sea rift, J. Geophys. Res. 96, no. B12, 20,195–20,216. earthquake activity of the northern part of the Dead Sea fault zone, Bronk Ramsey, C. (1995). Radiocarbon calibration and analysis of southern Turkey, Tectonophysics 426, 281–293. stratigraphy: The OxCal program, Radiocarbon 37, 425–430. Alchalbi, A., M. Daoud, F. Gomez, S. McClusky, Reilinger, R., Daëron, M., L. Benedetti, P. Tapponnier, A. Sursock, and R. C. Finkel M. A. Romeyeh, A. Alsouod, R. Yassminh, B. Ballani, R. Darawcheh, (2004). Constraints on the post ~25-ka slip rate of the Yammoûneh R. Sbeinati, Y. Radwan, R. A. Masri, M. Bayerly, R. A. Ghazzi, and fault (Lebanon) using in situ cosmogenic 36Cl dating of offset M. Barazangi (2010). Crustal deformation in northwestern Arabia limestone-clast fans, Earth Planet. Sci. Lett. 227, 105–119. from GPS measurements in Syria: Slow slip rate along the northern Daëron, M., Y. Klinger, P. Tapponnier, A. Elias, E. Jacques, and A. Sursock Dead Sea fault, Geophys. J. Int. 180, 125–135. (2007). 12,000-year-long record of 10 to 13 paleoearthquakes on the Aldersons, F., Z. Ben-Avraham, A. Hofstetter, E. Kissling, and Yammoûneh fault, Levant fault system, Lebanon, Bull. Seismol. Soc. T. Al-Yazjeen (2003). Lower-crustal strength under the Dead Sea basin Am. 97, 749–771. from local earthquake data and rheological modeling, Earth Planet. El-Isa, Z. H., and H. Mustafa (1986). Earthquake deformations in the Lisan Sci. Lett. 214, 129–142. deposits and seismotectonic implications, Geophys. J. R. Astr. Soc. 86, Al-Taj, M. (2000). Active faulting along the Jordan Valley segment of the 413–424. Jordan—Dead Sea Transform, Ph.D. Thesis, University of Jordan, Ellenblum, R., S. Marco, A. Agnon, T. K. Rockwell, and A. Boas (1998). Amman. Crusader castle torn apart by earthquake at dawn, 20 May 1202, Al-Tarazi, E., E. Sandvol, and F. Gomez (2006). The February 11, 2004 Geology 26, 303–306. M ˆ 5:2 Dead Sea earthquake L in Jordan and its tectonic implication, Ferry, M., and M. Meghraoui (2008). Reply to comment of Dr M. Klein on: Tectonophysics 422, 149–158. “A 48-kyr-long slip rate history for the Jordan Valley segment of the Altunel, E., M. Meghraoui, V. Karabacak, S. Akyüz, M. Ferry, Ç. Yalçiner, Dead Sea fault”, Earth Planet. Sci. Lett. 268, 241–242. and M. Munschy (2009). Archaeological sites (Tell and Road) offset Ferry, M., M. Meghraoui, N. Abou Karaki, M. Al-Taj, H. Amoush, by the Dead Sea fault in the Amik Basin, Southern Turkey, Geophys. S. Al-Dhaisat, and M. Barjous (2007). A 48-kyr-long slip rate history J. Int. 179, 1313–1329. for the Jordan Valley segment of the Dead Sea fault, Earth Planet. Sci. Ambraseys, N. N. (2004). The 12th century seismic paroxysm in the Middle Lett. 260, 394–406. East: A historical perspective, in Investigating the Records of Past Fitzenz, D. D., M. A. Ferry, and A. Jalobeanu (2010). Long-term slip history Earthquakes, 21st course of the International School of Geophysics discriminates among occurrence models for seismic hazard assess- Bologna, Italy, 733–758. ment, Geophys. Res. Lett. 37, L20307, doi 10.1029/2010GL044071. Ambraseys, N. (2005). Historical earthquakes in Jerusalem—A methodolo- Franken, H. J. (1989). Deir' Alla (Tell) in Archaeology of Jordan, Vol. II1.- gical discussion, J. Seismol. 9, 329–340. II2. Field Reports, Surveys and Sites, D. Homès-Fredericq and Ambraseys, N. N. (2006). Earthquakes and Archeology, J. Archaeol. Sci. 33, J. Hennessy (eds), Peeters, Leuven, Belgium. 1008–1016. Garfunkel, Z., I. Zak, and R. Freund (1981). Active faulting in the Dead Sea Ambraseys, N. N. (2009). Earthquakes in the Mediterranean and Middle rift, in The Dead Sea Rift: Selected Papers of the International East, A Multidisciplinary Study of Seismicity up to 1900, Cambridge Symposium on the Dead Sea Rift Amsterdam, Netherlands, 1–26. University Press, Cambridge, 947 pp. Ginat, H., Y. Enzel, and Y. Avni (1998). Translocated Plio-Pleistocene Ambraseys, N. N., and J. A. Jackson (1998). Faulting associated with drainage systems along the Arava fault of the Dead Sea Transform, historical and Recent earthquakes in the eastern Mediterranean region, Tectonophysics 284, 151–160. Geophys. J. Int. 133, 390–406. Ginzburg, A., and Z. Ben-Avraham (1997). A seismic refraction study Ambraseys, N. N., C. P. Melville, and R. D. Adams (1994). The Seismicity of of the north basin of the Dead Sea, Israel, Geophys. Res. Lett. 24, Egypt, Arabia and the Red Sea: A Historical Review, Cambridge 2063–2066. University Press, Cambridge, 181 pp. Griffith, W. A., S. Nielsen, G. Di Toro, and S. A. F. Smith Rough faults, Amiran, D. H. K., E. Arieh, and T. Turcotte (1994). Earthquakes in Israel and distributed weakening, and off-fault deformation, J. Geophys. Res., adjacent areas: Macroseismic observations since 100 B.C.E, Isr. 115, no. B08409, doi 10.1029/2009JB006925. Explor. J. 44, 260–305. Gomez, F., M. Meghraoui, A. Darkal, F. Hijazi, M. Mouty, Y. Sulaiman, Arieh, E., Y. Rotstein, and U. Peled (1982). The Dead Sea earthquake of 23 R. Sbeinati, R. Darawcheh, R. Al-Ghazzi, and M. Barazangi (2003). April 1979, Bull. Seismol. Soc. Am. 72, 1627–1634. Holocene faulting and earthquake recurrence along the Serghaya Avni, R., D. D. Bowman, A. Shapira, and A. Nur (2002). Erroneous branch of the Dead Sea fault system in Syria and Lebanon, Geophys. interpretation of historical documents related to the epicenter J. Int. 153, 658–674. of the 1927 Jericho earthquake in the Holy Land, J. Seismol. 6, Gomez, F., T. Nemer, C. Tabet, M. Khawlie, M. Meghraoui, and 469–476. M. Barazangi (2007). Strain partitioning of active transpression within Barka, A., H. S. Akyüz, G. Sunal, Z. Çakir, A. Dikbas, B. Yerli, E. Altunel, the Lebanese restraining bend of the Dead Sea fault (Lebanon and R. Armijo, B. Meyer, J.-B. de Chabalier, T. K. Rockwell, J. R. Dolan, SW Syria), Geological Society, London, Special Publications 290, R. Hartleb, T. Dawson, S. Christofferson, A. Tucker, T. Fumal, 285–303. R. Langridge, H. Stenner, W. Lettis, J. Bachhuber, and W. Page Guidoboni, E. (1989). I terremoti prima del Mille in Italia e nell’area (2002). The surface rupture and slip distribution of the 17 August mediterranea, Storia-Geofisica-Ambiente, Bologna, 768. 1999 Izmit earthquake (M 7.4), North Anatolian fault, Bull. Seismol. Guidoboni, E., A. Comastri, and G. Traina (1994). Catalog of Ancient Earth- Soc. Am. 92, 43–60. quakes in the Mediterranean Area up to the 10th Century, ING-SGA, Bartov, Y., A. Agnon, Y. Enzel, and M. Stein (2006). Late Quaternary Bologna, 504 pp. faulting and subsidence in the central Dead Sea basin, Isr. J. Earth Haynes, J., T. Niemi, and M. Atallah (2006). Evidence for ground-rupturing Sci. 55, 18–31. earthquakes on the Northern Wadi Araba fault at the archaeological 66 M. Ferry, M. Meghraoui, N. Abou Karaki, M. Al-Taj, and L. Khalil

site of Qasr Tilah, Dead Sea Transform fault system, Jordan, J. Seis- Migowski, C., A. Agnon, R. Bookman, J. F. W. Negendank, and M. Stein mol. 10, 415–430. (2004). Recurrence pattern of Holocene earthquakes along the Dead Hofstetter, A. (2003). Seismic observations of the 22/11/1995 Gulf of Aqaba Sea Transform revealed by varve-counting and radiocarbon dating earthquake sequence, Tectonophysics 369, 21–36. of lacustrine sediments, Earth Planet. Sci. Lett. 222, 301–314. Jordan Seismological Observatory (2004). The Dead Sea earthquake, Nemer, T., and M. Meghraoui (2006). Evidence of coseismic ruptures along ML = 4.9—a preliminary report, Natural Resources Authority, The the Roum fault (Lebanon): A possible source for the AD 1837 earth- Hashemite Kingdom of Jordan, 6 pp. quake, J. Struct. Geol. 28, 1483–1495. Kanamori, H., and D. G. Anderson (1975). Theoretical basis of some Nemer, T., M. Meghraoui, and K. Khair (2008). The Rachaya-Serghaya fault empirical relations in seismology, Bull. Seismol. Soc. Am. 65, system (Lebanon): Evidence of coseismic ruptures, and the AD 1759 1073–1095. earthquake sequence, J. Geophys. Res. 113, no. B05312, doi 10.1029/ Karabacak, V., E. Altunel, M. Meghraoui, and H. S. Akyüz (2010). Field 2007JB005090. evidences from northern Dead Sea fault zone (south Turkey): New Niemi, T. M., and N. Mansoor (2002). Nearly a millennium of seismic quies- findings for the initiation age and slip rate, Tectonophysics 480, cence in Aquaba, Jordan along the southern Dead Sea Transform, 172–182. paper no. 227, Annual Meeting of the Geological Society of America, Karcz, I. (2004). Implications of some early Jewish sources for estimates Denver, Colorado, 27–30 October 2002, 227–210. of earthquake hazard in the Holy Land, in Investigating the Records Niemi, T. M., H. Zhang, M. Atallah, and J. B. J. Harrison (2001). Late of Past Earthquakes, 21st course of the International School of Pleistocene and Holocene slip rate of the northern Wadi Araba fault, Geophysics Bologna, Italy, 759–792. Dead Sea Transform, Jordan, J. Seismol. 5, 449–474. Katz, O., and O. Crouvi (2007). The geotechnical effects of long human Nur, A., and E. H. Cline (2000). Poseidon’s Horses: Plate tectonics and habitation (2000 < years): Earthquake induced landslide hazard in earthquake storms in the Late Bronze Age Aegean and eastern the city of Zefat, northern Israel, Eng. Geol. 95, 57–78. Mediterranean, J. Archaeol. Sci. 27, 43–63. Ken-Tor, R., A. Agnon, Y. Enzel, M. Stein, S. Marco, and J. Negendank Nur, A., and H. Ron (1996). And the walls came tumbling down: Earthquake (2001). High-resolution geological record of historic earthquakes in history in the Holy Land, in Archaeoseismology, S. Stiros and R. Jones the Dead Sea basin, J. Geophys. Res. 106, 2221–2234. (eds), British School at Athens, Athens, Greece, 75–85. King, G., G. Bailey, and D. Sturdy (1994). Active tectonics and human Petrunin, A., and S. V. Sobolev (2006). What controls thickness of sediments survival strategies, J. Geophys. Res. 99, 20,063–20,078. and lithospheric deformation at a pull-apart basin?, Geology 34, Klein, M. (2008). A comment on: “A 48-kyr-long slip rate history for 389–392. the Jordan Valley segment of the Dead Sea fault” EPSL 260 Poirier, J.-P., and M. A. Taher (1980). Historical seismicity in the near and (2007). 394–406, Earth Planet. Sci. Lett. 268, 239–240. Middle East, North Africa, and Spain from Arabic documents (VIIth- Klinger, Y., J.-P. Avouac, N. Abou Karaki, L. Dorbath, D. Bourles, and XVIIIth century), Bull. Seismol. Soc. Am. 70, 2185–2201. J. L. Reyss (2000). Slip rate on the Dead Sea Transform in northern Quennell, A. M. (1959). Tectonics of the Dead Sea rift, in 20th Araba Valley (Jordan), Geophys. J. Int. 142, 755–768. International Geological Congress, Association of African Geological Lazar, M., and Z. Ben-Avraham (2002). First images from the bottom of the Surveys, 385–405. Dead Sea—Indications of recent tectonic activity, in Honoring Eytan Quennell, A. M (1984). The Western Arabia Rift System, Geological Society, Sass and Amitai Katz on the occasion of their 70th birthdays, Jerusa- London, Special Publications 17, 775–788. lem, Israel, 211–218. Reches, Z., and D. F. Hoexter (1981). Holocene seismic and tectonic activity Le Beon, M., Y. Klinger, A. Q. Amrat, A. Agnon, L. Dorbath, G. Baer, in the Dead Sea area, in The Dead Sea Rift: Selected Papers of the J.-C. Ruegg, O. Charade, and O. Mayyas (2008). Slip rate and locking International Symposium on the Dead Sea Rift Amsterdam, depth from GPS profiles across the southern Dead Sea Transform, Netherlands, 235–254. J. Geophys. Res. 113, no. B11403, doi 10.1029/2007JB005280. Reilinger, R., S. McClusky, P. Vernant, and S. Lawrence (2006). Malkawi, A. I. H., and A. S. Alawneh (2000). Paleoearthquake features as GPS constraints on continental deformation in the Africa-Arabia- indicators of potential earthquake activities in the Karameh Dam site, Eurasia continental collision zone and implications for the dynamics Nat. Hazards 22, 1–16. of plate interactions, J. Geophys. Res. 111, doi 10.1029/ Marco, S., M. Hartal, N. Hazan, L. Lev, and M. Stein (2003). Archaeology, 2005JB004051. history, and geology of the A.D. 749 earthquake, Dead Sea Transform, Reimer, P.J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. Beck, P.G. Blackwell, Geology 31, 665–668. C. E. Buck, G. S. Burr, K. B. Cutler, P. E. Damon, R. L. Edwards, Marco, S., T. K. Rockwell, A. Heimann, U. Frieslander, and A. Agnon R. G. Fairbanks, M. Friedrich, T. P.Guilderson, C. Herring, K. A. Hugh- (2005). Late Holocene activity of the Dead Sea Transform revealed en, B. Kromer, F. G. McCormac, S. W. Manning, C. B. Ramsey, in 3D paleoseismic trenches on the Jordan Gorge segment, Earth R. W. Reimer, S. Remmele, J. R. Southon, M. Stuiver, S. Talamo, Planet. Sci. Lett. 234, 189–205. F. W. Taylor, J. van der Plicht, and C. E. Weyhenmeyer (2004). IntCal04 Marco, S., M. Stein, A. Agnon, and H. Ron (1996). Long-term earthquake Terrestrial radiocarbon age calibration, 0–26 cal kyr BP, Radiocarbon clustering: A 50,000-year paleoseismic record in the Dead Sea graben, 43, 1029–1058. J. Geophys. Res. 101, 6179–6192. Ryberg, T., M. Weber, Z. Garfunkel, and Y. Bartov (2007). The shallow McCalpin, J. P. (1998). Paleoseismology, Academic Press, New York, velocity structure across the Dead Sea Transform fault, Arava Valley, 588 pp. from seismic data, J. Geophys. Res. 112, no. B0830, doi 10.1029/ McClusky, S., R. Reilinger, S. Mahmoud, S. D. Ben, and A. Tealeb (2003). 2006JB004563. GPS constraints on Africa (Nubia) and Arabia plate motions, Geophys. Salamon, A., A. Hofstetter, Z. Garfunkel, and H. Ron (2003). Seismotec- J. Int. 155, 126–138. tonics of the Sinai subplate—The eastern Mediterranean region, Meghraoui, M., F. Gomez, R. Sbeinati, J. Van der Woerd, M. Mouty, Geophys. J. Int. 155, 149-–173. A. Darkal, Y. Radwan, I. Layyous, H. M. Najjar, R. Darawcheh, Savage, S., K. Zamora, and D. Keller (2001). Archaeology in Jordan, Am. J. F. Hijazi, R. Al-Ghazzi, and M. Barazangi (2003). Evidence for Archaeol. 105, 427–461. 830 years of seismic quiescence from paleoseismology, archeoseis- Savage, S., K. Zamora, and D. Keller (2002). Archaeology in Jordan, 2001 mology and historical seismicity along the Dead Sea fault in Syria, Season, Am. J. Archaeol. 106, 435–458. Earth Planet. Sci. Lett. 210, 35–52. Savage, S., K. Zamora, and D. Keller (2003). Archaeology in Jordan, 2002 Meyers, E., A. Kraabel, and J. Strange (1976). Ancient synagogue excava- Season, Am. J. Archaeol. 107, 449–475. tions at Khirbet Shema’, Upper Galilee, Israel 1970–1972, Annu. Am. Savage, S., K. Zamora, and D. Keller (2005). Archaeology in Jordan, 2004 Sch. Orient. Res. 42. Season, Am. J. Archaeol. 109, 527–555. Episodic Behavior of the Jordan Valley Section of the Dead Sea Fault 67

Sbeinati, M. R., R. Darawcheh, and M. Mouty (2005). The historical earth- Zilberman, E., R. Amit, N. Porat, Y. Enzel, and U. Avner (2005). Surface quakes of Syria: An analysis of large and moderate earthquakes from ruptures induced by the devastating 1068 AD earthquake in the south- 1365 B.C. to 1900 A.D, Ann. Geophys. 48, 347–435. ern Arava valley, Dead Sea rift, Israel, Tectonophysics 408, 79–99. Sbeinati, M. R., M. Meghraoui, G. Suleyman, P. Grootes, M.-J. Nadeau, H. Al Najjar, F. Gomez, and R. Al-Ghazzi (2010). Timing of earth- Centro de Geofísica de Évora quake ruptures at the Al Harif Roman aqueduct (Dead Sea fault, Syria) Universidade de Évora from archaeoseismology and palaeoseismology, special paper 471 in Rua Romão Ramalho, 59 Ancient Earthquakes (special volume), M. Sintubin, I. Stewart, 7002-554 Évora, Portugal T. Niemi, and E. Altunel (Editors), Geological Society of America [email protected] Bulletin, 243–267. (M.F.) Shapira, A., R. Avni, and A. Nur (1993). A new estimate for the epicenter of the Jericho earthquake of 11 July 1927, Isr. J. Earth Sci. 42, 93–96. Sieh, K. (1996). The repetition of large-earthquake ruptures, Proc. Natl. Acad. Sci. Unit. States Am. 93, 3764–3771. Institut de Physique du Globe Stein, R. S., A. A. Barka, and J. H. Dieterich (1997). Progressive failure on UMR 7516, 5 rue René Descartes 67084 Strasbourg, France the North Anatolian Fault since 1939 by earthquake stress triggering, (M.M.) Geophys. J. Int. 128, 594–604. ten Brink, U. S., Z. Ben-Avraham, R. E. Bell, M. Hassouneh, D. F. Coleman, G. Andreasen, G. Tibor, and B. J. Coakley (1993). Structure of the Department of Environmental and Applied Geology Dead Sea pull-apart basin from gravity analyses, J. Geophys. Res. University of Jordan 98, 21,877–21,894. Amman 11942, Jordan Tsafrir, Y., and G. Foerster (1992). The dating of the ’Earthquake of the (N.A.) Sabbatical Year’ of 749 C.E. in Palestine, Bull. Sch. Orient. Afr. Stud. 55, 231–235. Department of Earth and Environmental Sciences Tubb, J. N. (1988). Tell es-Sa’idiyeh: Preliminary report on the first three – The Hashemite University seasons of renewed excavations, Levant 20, 23 89. P.O. Box 150459 Tubb, J. N. (1998). Canaanites, The British Museum Press, London, 160. Zarqa 13115, Jordan Wdowinski, S., Y. Bock, G. Baer, L. Prawirodirdjo, N. Bechor, S. Naaman, (M.A.) R. Knafo, Y. Forrai, and Y. Melzer (2004). GPS measurements of current crustal movements along the Dead Sea fault, J. Geophys. Res. 109, no. B05403, doi 10.1029/2003JB002640. Department of Archeology Wells, D. L., and K. J. Coppersmith (1994). New empirical relationships University of Jordan among magnitude, rupture length, rupture width, rupture area, and Amman 11942, Jordan (L.K.) surface displacement, Bull. Seismol. Soc. Am. 84, 974–1002. Willis, B. (1928). Earthquakes in the Holy Land, Bull. Seismol. Soc. Am. 18, 73–103. Manuscript received 13 April 2010