Late Holocene lake levels of the Dead Sea

18
Geological Society of America Bulletin doi: 10.1130/B25286.1 2004;116;555-571 Geological Society of America Bulletin R. Bookman (Ken-Tor), Y. Enzel, A. Agnon and M. Stein Late Holocene lake levels of the Dead Sea Email alerting services cite this article to receive free e-mail alerts when new articles www.gsapubs.org/cgi/alerts click Subscribe America Bulletin to subscribe to Geological Society of www.gsapubs.org/subscriptions/ click Permission request to contact GSA http://www.geosociety.org/pubs/copyrt.htm#gsa click viewpoint. Opinions presented in this publication do not reflect official positions of the Society. positions by scientists worldwide, regardless of their race, citizenship, gender, religion, or political article's full citation. GSA provides this and other forums for the presentation of diverse opinions and articles on their own or their organization's Web site providing the posting includes a reference to the science. This file may not be posted to any Web site, but authors may post the abstracts only of their unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education and to use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make GSA, employment. Individual scientists are hereby granted permission, without fees or further requests to Copyright not claimed on content prepared wholly by U.S. government employees within scope of their Notes Geological Society of America on August 17, 2011 gsabulletin.gsapubs.org Downloaded from

Transcript of Late Holocene lake levels of the Dead Sea

Geological Society of America Bulletin

doi: 10.1130/B25286.1 2004;116;555-571Geological Society of America Bulletin

 R. Bookman (Ken-Tor), Y. Enzel, A. Agnon and M. Stein Late Holocene lake levels of the Dead Sea  

Email alerting servicescite this article

to receive free e-mail alerts when new articleswww.gsapubs.org/cgi/alertsclick

SubscribeAmerica Bulletin

to subscribe to Geological Society ofwww.gsapubs.org/subscriptions/click

Permission request to contact GSAhttp://www.geosociety.org/pubs/copyrt.htm#gsaclick

viewpoint. Opinions presented in this publication do not reflect official positions of the Society.positions by scientists worldwide, regardless of their race, citizenship, gender, religion, or politicalarticle's full citation. GSA provides this and other forums for the presentation of diverse opinions and articles on their own or their organization's Web site providing the posting includes a reference to thescience. This file may not be posted to any Web site, but authors may post the abstracts only of their unlimited copies of items in GSA's journals for noncommercial use in classrooms to further education andto use a single figure, a single table, and/or a brief paragraph of text in subsequent works and to make

GSA,employment. Individual scientists are hereby granted permission, without fees or further requests to Copyright not claimed on content prepared wholly by U.S. government employees within scope of their

Notes

Geological Society of America

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

For permission to copy, contact [email protected] 2004 Geological Society of America 555

GSA Bulletin; May/June 2004; v. 116; no. 5/6; p. 555–571; doi: 10.1130/B25286.1; 8 figures; 3 tables.

Late Holocene lake levels of the Dead Sea

R. Bookman (Ken-Tor)†

Institute of Earth Sciences, The Hebrew University of Jerusalem, Jerusalem 91904, Israel and Geological Survey of Israel,30 Malkhe Israel Street, Jerusalem 95501, Israel

Y. EnzelInstitute of Earth Sciences, The Hebrew University of Jerusalem, Jerusalem 91904, Israel and Department of Geography,Hebrew University of Jerusalem, Jerusalem 91905, Israel

A. AgnonInstitute of Earth Sciences, The Hebrew University of Jerusalem, Jerusalem 91904, Israel

M. SteinInstitute of Earth Sciences, The Hebrew University of Jerusalem, Jerusalem 91904, Israel and Geological Survey of Israel,30 Malkhe Israel Street, Jerusalem 95501, Israel

ABSTRACT

This work presents a high-resolutionlake-level record of the late Holocene DeadSea, a hypersaline terminal lake whosedrainage basin encompasses both Mediter-ranean and hyperarid climatic zones. Thelake-level curve reflects the regional hydro-logic variations in the drainage basin,which in turn represent the Levant paleo-climates. The curve is based on 46 radio-carbon ages of organic remains from well-exposed sedimentary sequences along theDead Sea shores. These sequences recordfluvial and lacustrine depositional environ-ments. The paleolakeshores are marked byshore ridges, coarse-sand units, and ara-gonite crusts; in the modern Dead Sea, suchfeatures indicate the exact elevation of theshore.

The late Holocene Dead Sea level fluctu-ated within the range of 390 to 415 m belowsea level (mbsl). For most of the time thelake was below the topographic sill (402mbsl) separating the northern and southernbasins of the Dead Sea and was confined tothe deep northern basin. Nevertheless,short-term rises in the late Holocene DeadSea level caused the flooding of the shallowand flat southern basin.

Highstands occurred in the second andfirst centuries B.C. and the fourth centuryA.D. during the Roman and early Byzan-

†Present address: 204 Heroy Geology Laboratory,Syracuse University, Syracuse, New York 13244-1070; e-mail: [email protected].

tine periods, respectively, in the eleventhand twelfth centuries A.D. during the Cru-sader period, and at the end of the nine-teenth century A.D. The rises mark a sig-nificant change in the annual rainfall in theregion, which likely exceeded the instru-mentally measured modern average.

The curve also indicates drastic dropsthat exposed the sedimentary sequences toerosion. The oldest and probably deepestdrop in the lake level culminated during thefifteenth and fourteenth centuries B.C. aftera retreat from a higher lake stand. The lon-gest lowstand occurred after the Byzantineperiod and continued at least until theninth century A.D. This arid period coin-cided with the invasion of Moslem-Arabtribes into the area during the seventh cen-tury A.D. The dramatic fall of the Dead Sealevel during the twentieth century is pri-marily artificial and has been caused by thediversion of runoff water for the drainagebasin, but the magnitude is not consideredexceptional for the late Holocene. Althoughthe past drops in the lake never exceededthe modern artificial drop rates, they dorepresent extreme arid conditions that oc-curred frequently over the past severalthousand years.

Keywords: lake level, paleoclimate, DeadSea, Holocene, depositional environments,lacustrine deposits, radiocarbon dating.

INTRODUCTION

The Dead Sea (Fig. 1) is a terminal hyper-saline lake located in the deepest basin of the

Dead Sea rift. It is a remnant water body ofthe significantly larger Lake Lisan that filledthe Dead Sea basin and the Jordan Valley dur-ing the last glacial period (e.g., Neev and Em-ery, 1967; Begin et al., 1974; Stein, 2001, andreferences therein). These water bodies behaveessentially as ‘‘amplifier lakes’’ (see definitionby Street-Perrott and Harrison, 1985), inwhich lake levels are sensitive to the waterbalance. The changes in their lake levels andsize can be used to trace the paleohydrologicconditions in the drainage area.

The current (2003) Dead Sea level is ;415m below sea level (mbsl). The catchment areaof the lake is ;43,000 km2 (Fig. 1B), and itdrains both the Mediterranean and hyperaridclimatic areas. The Jordan River and its trib-utaries from the north are the major runoffsources. An additional water input is contrib-uted by seasonal floods and base flows fromsurrounding streams and spring seepages.

The Dead Sea basin is bounded in the eastand west by tectonic escarpments. The differ-ence in elevation between the shoulders of therift and the present shores of the lake are 300–500 m in the west and 400–1000 m in the east,whereas the southern and northern margins arealmost flat (Fig 1C). The basin is divided intotwo subbasins separated by a topographic sillat an elevation of ;402 mbsl (Fig. 1D). Themodern water body occupies the morpholog-ically deep northern basin (;300 m deep),whereas the southern shallow basin (filledwith sediments) is essentially dry (at presentit is covered by artificial evaporation ponds;Fig. 1C). This configuration plays a central

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

556 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

Figure 1. (A) The general geographic setting of the Mediterranean. (B) Average annual rainfall in millimeters over the drainage basinarea of the Dead Sea (shaded). (C) A detailed location map of the Dead Sea with contour intervals of 100 m below sea level and 200 mabove sea level. (D) North-south cross section of the Dead Sea basin. Note that the southern shallow, flat basin is flooded only abovethe sill elevation at 402 mbsl.

role in the lake-level history, because thesouthern basin acts as a large evaporation sur-face that can buffer lake-level rise.

Here, we present a lake-level curve for theDead Sea that we have reconstructed for thepast 4000 yr. The curve is based on radiocar-bon dating of sedimentary sequences that areexposed along the recently retreating westernshores of the Dead Sea. A composite levelcurve was established by using information

from two sites representing different but con-temporaneous depositional environments.Combining complementary information fromthe two sites enables the reconstruction of ahigh-resolution curve, which in turn repre-sents the hydrologic history of the region dur-ing the late Holocene. The reconstructed lev-els are then compared with the chronology ofthe Near East cultures to assess possible linksbetween climate change and cultural shifts.

SEDIMENTOLOGY AND LAKE-LEVELINTERPRETATION

The interpretation of different depositionalenvironments described in contemporaneousexposures is the key for reconstructing theDead Sea lake-level curve. Paleolimnologicaldata for accurate lake-level history is best ac-quired from a combination of synchronouslake-level indicators in the different deposi-

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 557

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

Figure 2. Schematic illustration of the depositional environments in the Dead Sea. A rise in lake level shifts the shore and buries theprevious sediments (modified after Bartov et al., 2002). (A) Clay- to silt-size detrital lacustrine deposits (L) with symmetric ripple marks(R) overlain by fluvial deposits (F). Symmetric ripple marks indicate a shallow, nearshore environment (tool for scale is 20 cm long).(B) Shore ridge composed of rounded cobbles. The cobbles are covered with aragonite crust (AC) on the side facing toward the lake.Laminated lacustrine sediments are exposed below the shore ridge (L) and a liquefied sand unit (S) above it. (C) Cobble covered withhard aragonite crusts in the shore environment. Note that the crust is composed of several laminae (coin for scale is 2 cm in diameter).(D) Lacustrine sediments composed of alternating aragonite and detrital laminae (AD) and thick, silt- to clay-size, dark detritus layers(D). The detrital laminae and layers are composed of dark, silt- to clay-size clasts brought in as suspended material during flood events(Fig. 3B). Chemical deposition of aragonite laminae occurs after massive bicarbonate input into the lake. This laminated sequence wasdeposited at the end of the nineteenth century highstand (upper laminated sequence in Fig. 5).

tional environments (Wohl and Enzel, 1995).Therefore, the information with the highestresolution is gained by physical correlation ofsystematic responses in the depositional en-vironments to changes in lake level, coupledwith a spatial perspective of the area.

The Dead Sea Holocene sediments are com-posed of fluvial and lacustrine sequences de-posited in offshore and onshore environments.The terminology used in this paper to describethe sediments is based on Manspeizer (1985),Bowman (1990), Kadan (1997), Machlus et al.

(2000), Bartov et al. (2002), and referencestherein. The three main depositional environ-ments are illustrated in Figure 2: the fluvialenvironment, which includes the onshore de-posits in alluvial fans, braided streams, or de-bris flows along the basin margins; the shore

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

558 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

environment, a relatively narrow zone of de-position along the lake shore; and the lacus-trine environment, which includes offshoredeposition in the main water body at differentwater depths. Each depositional environmentdominates over a particular horizontal distanceand interfingers with the one adjacent.

The procedure for reconstructing paleo–lake levels from stratigraphic levels is basedon determination of the absolute elevation ofan identified shore indicator; given the eleva-tions, physical correlation along the exposedsequences enabled the reconstruction of thedepth of deposition of a contemporaneous la-custrine deposit (Kadan, 1997; Machlus et al.,2000; Bartov et al., 2002). The depth-sensitiveenvironments are distinguished by depositswith characteristic lithology, which are direct-ly associated with lake-level elevation, andtheir lateral position relative to the lakeshorein the various sites studied. The present-dayDead Sea serves as the analogue for under-standing the depositional environments withrespect to changes in the lake level.

Fluvial Environment

Layers of fluvial sediments range from asingle pebbly layer on top of an unconformitysurface to horizontal to subhorizontal beds afew meters thick of medium- to poorly sorted,grain-supported pebbles to cobbles. Stratifi-cation is identified by changes in grain size,sorting, or upward fining. The matrix is usu-ally coarse sand. The basal contact is frequent-ly erosional. The fluvial sediments are depos-ited in alluvial fans and braided streams abovelake level (e.g., Sneh, 1979; Nemec and Steel,1984). Stratification in fluvial deposits reflectsoccasional flash floods, typical of the DeadSea area.

Debris-flow sediments usually lack stratifi-cation and consist of mud-supported thickunits of poorly sorted pebbles to boulders.Most debris-flow sediments are currently de-posited in high-angle alluvial fans along theescarpments of the Dead Sea basin above thelake level. They occur during rare heavy rain-storms (BenDavid-Novak, 1998).

Shore Environment

Sand is common in the shore environment.Sand layers are lenticular, composed of light-color, coarse to medium carbonate grains, arewell sorted, and are a few centimeters to ;1m thick. Some of the layers are finely lami-nated, whereas others appear massive. Occa-sionally, symmetric ripple marks with crest-lines roughly parallel to the modern shore arevisible (Fig. 2A). The distribution of sand lay-

ers is limited to a narrow strip oriented par-allel to the lakeshore. Frequently, thick sandunits thin out and disappear at a distance of afew tens of meters.

Coarse sand units are a common sedimen-tary feature along the recent shores of theDead Sea. Sandy units are visible at leveledshore steps that were left behind as the lakelevel retreated (Fig. 3A). The most recent stepis forming along the shore and is visible up toa few meters below lake level. Wave actionalong the shore erodes and sorts the carbonatesand grains before deposition, or, alternatively,sand is deposited at depths of a few meters asfallout from flood suspension entering the lake(e.g., Reineck and Singh, 1973; de Raff et al.,1977). Symmetric ripple marks indicate ashallow nearshore environment (de Raff et al.,1977) and a water depth lower than 5 m(Machlus et al., 2000). Thick and well-sortedsand units probably represent a relatively longperiod of wave action on a stable shoreline.

Beach ridges are identified by their distinctsedimentary structure of well-sorted, pebbly,lakeward- and/or landward-inclined ridges(foresets and backsets, respectively, e.g., Ka-dan, 1997; Adams and Wesnousky, 1998; Bar-tov et al., 2002) parallel to the lakeshore (Fig.2B). They are composed of well-rounded,clast-supported gravels commonly with platypebbles. The clasts range in size from sand tocobbles (pebbles being the majority) with amatrix of coarse sand to granules without anysilt and clay. Each ridge may be composed ofa few dipping, asymptotically parallel setswith a narrow range of grain size in each setand may extend horizontally along severalmeters. A thin sandy-silt layer covered withclay usually marks lagoon sediments deposit-ed behind the ridge (Kadan, 1997).

Beach ridges are a characteristic feature inthe modern shores of the Dead Sea (Klein,1986; Bowman, 1997). They are composed ofgravels that were transported from the fan-delta outlets along the shores built by waveaction (Kadan, 1997; Machlus et al., 2000;Bartov et al., 2002). Our field observations inthe modern shore area indicate that such ridg-es record the precise elevation of lake level atthe time they are formed and that they are as-sociated with storms, usually following thefloods that transported the coarse alluvial sed-iments into the lakeshores. We also observedthat the individual cross-bedding sets withinthe ridge represent discrete wave-energy con-ditions during storms when clasts of a partic-ular size and shape were transported (e.g., Ad-ams and Wesnousky, 1998). Backsets wereprobably built when the lake rose and pro-gressed landward, and foresets point to the re-treat of the level or to interaction of waves

and the delta front (Kadan, 1997; Bartov etal., 2002). Multiple imbricated beach ridgesbuilt one on top of the other in a horizontalarrangement indicate a stable lake stand at theperiod of their deposition, which enables theridge to accumulate and to prograde lakeward(e.g., Adams and Wesnousky, 1998). Step-likeshore ridges built on delta fronts or steep lakemargins, as are visible at the modern shore ofthe Dead Sea, indicate the abandonment andfast retreat of the water level (Orford et al.,1991; Bowman, 1997).

Aragonite crusts are composed of aragonitecrystals in a hard fabric, whereby thin indi-vidual crusts can compose packages up to afew tens of centimeters. Occasionally gypsumcrystals up to a few millimeters in size forma distinct lamina within the crusts. Aragonitecrusts are associated with all different sedi-ments, but the most pronounced appearance ofthese crusts is in shore ridges, where they cov-er the preexisting topography and gravels(Figs. 2B and 2C), sometimes cementing theminto hard beach rocks. Crusts cover fluvial de-posits in horizontal layers and finer shore andlacustrine sediments in wavy surfaces. Thecrusts appear also as fine laminae depositedwithin sand units sometimes covering ripple-mark structures. Modern and buried driftwoodon the shore is often covered with such acrust, and sometimes crusts cover onshorevegetation during later lake-level rises (Neevand Emery, 1967; Klein, 1986). Platy, 1-mm-thick to 5-cm-thick fragments of reworkedaragonite crusts form imbricate structures thatoccasionally appear in small pockets and lens-es in sand units, as matrix in pebbly shoreridges, and as thick beds forming shore ridges.

Aragonite crusts are excellent lake-level in-dicators. These crusts are the result of evap-oration along the margin of the lake and are acommon feature in hypersaline environments(e.g., Garber and Friedman, 1983). The crustsare deposited at the lake margins at depths of0 to less than 1 m (Garber, 1980; Machlus etal., 2000) and are a common feature on theretreating plains of the modern Dead Sea(Klein, 1986). Thick aragonite crusts are in-terpreted as representing a stillstand of a fewyears or longer or of repeated minor trans-gressions. The crusts are often exposed at theshore during lake-level retreat, subjectingthem to erosion that peels and crumbles themand then redeposits the fragments in the shoreenvironment.

Lacustrine Environment

Silt- to clay-size dark detritus forms mas-sive layers with thickness ranging from a fewcentimeters to a few tens of centimeters (Fig.

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 559

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

Figure 3. (A) Retreating shorelines formed since the 1970s due to the modern lake-level drop. Each strand line represents the lake levelduring its deposition. (B) Flood plume of runoff water laden with suspended detrital sediments entering the Dead Sea. These suspendedsediments will later spread and settle over the large area of the lake floor. Note the people standing on the shore for scale. (C) TheZe’elim site (318209300N, 358249300E) located at the distal part of the Ze’elim fan delta on the Dead Sea shore, photographed from thewest. The height of the Ze’elim B outcrop is ;10 m. The Ze’elim A outcrop is exposed in an incised gully ;400 m to the north. In thebackground is the eastern escarpment of the Dead Sea basin. (D) The David site (318289000N, 358249000E) located at the incised northernbank of the Nahal David fan delta, photographed from the west. The top of the outcrop extends between 390 and 402 mbsl.

2D). The dark brownish detritus beds are com-posed of quartz and calcite grains as well asmicritic limestone, dolomite clasts, and clayminerals derived from the formations exposedin the catchment area (Heim et al., 1997).

These sediments are brought by the season-al floods in suspension; after they enter thelake, they are deposited below wave-actionlevel in low-energy environments. Each layerprobably represents the sediment discharge ofeither a single flood or, where laminated, canpoint to different stages in the flood or a se-quence of floods in the same season. Theflood-suspended sediments form mud plumes(Fig. 3B) (e.g., Anderson, 2001) that are car-ried into the inner part of the lake. Conse-quently, thicker layers were deposited close tothe shore at water depths of a few meters andthinner layers at greater depths farther into thelake.

Clay-size detritus and silty sand are foundas alternating layers a few millimeters to cen-

timeters thick. Some of the silty-sand layersappear as symmetric ripple marks (Fig. 2A).Scarce aragonite laminae, mostly less than 1mm, are scattered within the detrital sedi-ments, probably because their precipitation isconcealed by the intensive detrital input (Gar-ber et al., 1987).

These fine detritus and silty sand intervalsare fallout from flood suspension entering thelake. Their grain size varies according to theflood discharge. They are deposited at a shortdistance from the shore, at a relatively shallowdepth beyond wave base. During high-energywave action (storms), ripple marks will form.

Laminated aragonite and silty-clay–size de-tritus are common deposits of Lake Lisan thatoccupied the basin during the late Pleistocene(Neev and Emery, 1967; Begin et al., 1974;Katz et al., 1977; Katz and Kolodny, 1989;Stein, 2001) and also appear in the Holocenedeposits (Yechieli et al., 1993; Heim et al.,1997; Kadan, 1997; Migowski et al., 1998).

These sediments are composed of millimeter-scale white and dark laminae, made of small(5–10 mm) aragonite needles and silty-clay–size clasts, respectively (Fig. 2D). The ara-gonite laminae differ from the aragonite crustsdescribed earlier in their soft appearance.

The alternating aragonite and detrital lami-nae of Lake Lisan were attributed by Stein etal. (1997) to the input of fresh water loadedwith bicarbonate. Stein et al. (1997) proposedthat the aragonite precipitated from the upperfreshwater body, but the process required in-put of Mg from the lower brine-type body.This model was corroborated by Barkan et al.(2001), who suggested that precipitation ofaragonite in the modern Dead Sea occurs aftermassive bicarbonate input by the winter floodsand depends on the high partial pressure ofCO2 in the lake. The detrital laminae are de-posited from the mud plumes of floods de-scribed earlier. The millimeter-scale thicknessof the detrital laminae indicates that the lam-

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

560 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

Figure 4. Schematic illustration of the Ze’elim and David fan deltas on the western shoreof the Dead Sea. The study sites are marked with thick arrows. Note the scale. (A) TheZe’elim site is located on a flat plain with a wide zone sensitive to lake-level fluctuations.This site provides a relatively continuous record of the late Holocene Dead Sea. The twolocations marked A and B are where stratigraphic sections shown in Figure 5 were mea-sured. (B) The David site is situated in the high-gradient active channel of the fan delta.It consists mostly of fluvial sediments, and only a narrow zone is influenced by lake-levelfluctuations. In this setting, the lacustrine units within the coarse-grain sequences markthe maximum lake-level elevations.

inated alternating sequences were probablydeposited below wave action and relatively farfrom the stream mouths. These sequenceswere attributed to variable lake depths (Mach-lus et al., 2000; Bartov et al., 2002), and thephysical correlation of these finely laminatedsequences with shore-environment depositsindicates that they could also have been de-posited in water depths of less than a few me-ters (Machlus et al., 2000).

SITE DESCRIPTION

The drastic (mostly artificially induced)lake-level fall of ;80 cm/yr during the secondhalf of the twentieth century exposed depositsof late Holocene age around the Dead Sea.Gullies incised into the newly exposed plainsand fan deltas along the shores of the lakereveal a very detailed depositional history. We

examined and described two of the most well-exposed sites along the western Dead Seashore—the Nahal David fan delta near EinGedi (termed the David site herein) and theZe’elim plain east of the archaeological site ofMassada (termed the Ze’elim site herein) (Fig.1C).

The Ze’elim site is located on the south-western shore of the present-day Dead Sea ona flat plain at an elevation of ;400 mbsl (Fig.3C). In the Ze’elim site, mainly lacustrine de-posits are exposed, indicating that it was cov-ered by water throughout most of the past fewthousand years and was relatively preservedfrom erosion. The flat plain is sensitive tometer-scale lake-level fluctuations that affect awide zone of deposition (Fig. 4A). Therefore,this site provides a relatively continuous sed-imentological and paleolimnological record ofthe late Holocene Dead Sea. Lacustrine sedi-

ments recognized in the Ze’elim exposuresrepresent highstands, but their absolute ele-vation cannot be determined. The site, how-ever, provides information on lake-level dropsthat exposed the lacustrine deposits to shoreand fluvial processes.

The David site is located at an elevation of390 to 402 mbsl in the Nahal David Holocenefan delta and consists mostly of fluvial sedi-ments (Fig. 3D). The fan delta is situatedwithin 1 km of the escarpment (Fig. 4B) form-ing a relatively steep topography, which re-sults in a narrow, high-gradient zone of de-position. In contrast to the Ze’elim site, theDavid site was subjected to erosion, so de-positional hiatuses appear in the section. How-ever, in such a setting, tongues of lacustrineand shore deposits within the fluvial sequenc-es of the fan delta mark the maximum lake-level elevations during highstands.

The two sites are complementary to eachother in that they recorded in their sedimentsboth highstands and lowstands. The exposureswere documented in detail in the field, includ-ing stratigraphic and physical correlationalong and between several exposures at thesame site. A precise topographic survey pro-vided the absolute elevation of each sedimen-tological unit and sample. The sequences weredivided into stratigraphic units according totheir lithology and the unconformities divid-ing them (Fig. 5). In the David site, two off-shore fine-grained units were recognized (Fig.6) and integrated within the composite lake-level record (Fig. 7). The Dead Sea lake-levelelevations presented in this work follow thedatum used by the Survey of Israel and are2.57 m higher than the Dead Sea Works datumsometimes used in other studies.

Ze’elim Site

Two gullies in the Ze’elim site were select-ed for the study (Figs. 3C and 4A), and theirdetailed and relatively continuous depositionalrecord is described and discussed. The north-ern exposure, Ze’elim A, is more lacustrine incharacter, although fluvial and shore depositsalso exist. This exposure, presented in Figure5, is composed of more than 9 m of well-stratified sediments.

The oldest exposed unit, termed unit I,comprises eastward-dipping laminated fine-grained detrital sediments that are truncatedand overlain by pebbles and by a shore ridge(unconformity A) (Fig. 5). The shore ridgeconsists of well-sorted imbricated crumblyaragonite-crust fragments layered in foresetsdipping toward the east, extending horizontal-ly a few meters, and having a thickness of upto 1 m. Unit II consists mostly of dark clay-

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 561

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

Figure 5. The Ze’elim site sections. On the left, the sedimentary description and radiocarbon chronology. Stratigraphic units are markedby Roman numerals (I–V) and major unconformities by capital letters (A–D). Arrows mark the radiocarbon sample location. On theright, the lake-level curve is reconstructed from the sedimentological information. Dashed line indicates unconformities. A break in thecurve means sedimentological information was (1) lost by erosion during lowstands, or (2) uninterpretable because level indicators werenot deposited during highstands. Calibrated radiocarbon dates (1s) are marked on the curve at the absolute elevation of the samples.Vertical arrows indicate that the deposition of the sample was below lake level in a lacustrine environment. Sample numbers on thecalibrated ranges are referenced from Tables 1 and 2.

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

562 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

Figure 6. The David site. The 300 m exposure is presented in a schematic illustration on the bottom of the figure. Note the lacustrineunits and shore ridges that represent lake-level indicators. A, B, and C are enlarged segments of the outcrop from which informationfor the lake-level curve was gathered. Radiocarbon dates are presented in years before present. For calibration ranges, see Table 3, andfor the David site reconstructed lake-level curve, see Figure 7.

to silt-size and sand detrital units, with a fewsymmetric ripple marks and aragonite crusts.A thin sand unit with an aragonite crust rep-resents unconformity B, which separates thelacustrine units. Alternating laminae of ara-

gonite and clay- to silt-size detritus are com-mon in unit III, in addition to the massiveclay- to silt-size layers and a sand layer a fewcentimeters thick with aragonite crusts at 404mbsl. A 10-cm-thick coarse-sand unit covered

by aragonite crust at the elevation of ;403mbsl represents the period of exposure anderosion marked as unconformity C. Unit IVoverlies the unconformity and is similar in li-thology to unit III. A 20-cm-thick coarse-sand

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 563

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

Figure 7. Unification of the lake-level infor-mation from the Ze’elim sites A and B andDavid site. At the bottom of the figure, thecomposite curve of the Dead Sea levels. Thecolor frames represent the different strati-graphic units that were determined fromthe sections (Figs. 5 and 7). Dashed linesmark unconformities. An incompletely col-ored frame means that the informationfrom the site is not full. The elevation of 402mbsl is in bold for reference between thecurves.N

layer with aragonite crusts that crops out at;402 mbsl marks the unconformity D surface.The uppermost and youngest sequence, unit V,consists of fine-silt–size to clay-size layersand a distinct alternating aragonite and detritallaminae sequence (upper laminated sequence,Fig. 5).

The Ze’elim B exposure, south of theZe’elim A outcrop (Figs. 3C and 4A), com-prises .10 m of stratified sediments exposedalong 300 m of an incised stream on theZe’elim alluvial fan. The exposure wasmapped in detail and is summarized in threecolumnar sections that represent the lateralstratigraphic relationships (Fig. 5).

The easternmost section at the Ze’elim Boutcrop exposes the eastward-dipping, lami-nated, fine-grained sediments of unit I identi-fied at the Ze’elim A exposure. An erosionalunconformity marked by pebbles and arago-nite crusts was correlated physically in thefield to unconformity A and is covered by silt-to clay-size lacustrine sediments of unit II.Unconformity B is marked by an erosionalchannel filled with pebbles between 407 and406 mbsl in the Ze’elim B outcrop. Aragonitecrusts cover the bottom and top of thechannel.

Above this fluvial channel the depositionalsequence is fully exposed along the outcrop,and the physical correlation is simple (Fig. 5).Laminations of detrital silt to clay alternatingwith aragonite laminae comprise unit III (406to 402 mbsl). Unconformity C cuts unit IIIand is marked by aragonite crusts coveringpebbles and coarse sand. Above unconformityC, sandy units are more frequent, but lacus-trine sediments dominate the eastern part ofthe Ze’elim B outcrop (units IV and V). Apebbly shore ridge (401–402 mbsl) coveredwith aragonite crust terminates toward the eastand transforms into unconformity D, which isalso covered with aragonite crusts and abun-dant organic debris, including a buried treetrunk. The uppermost alternating aragoniteand detrital laminae sequence (in unit V) is

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

564 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

TABLE 1. AMS RADIOCARBON DATES FROM THE ZE’ELIM A SITE

Sampleno.

Laboratoryno.

Unit 14C age(yr B.P.)

Calculated distribution†

(calendar years)Modeled distribution‡

(calendar years)

1 KIA3213 Base of V 600 6 40 s: A.D. 1300–1370 (54.9%) s: A.D. 1381–1410A.D. 1385–1400 (13.3%) 2s: A.D. 1345–14352s: A.D. 1290–1420 (95.4%) Agreement—87%

2 KIA3214A IV 780 6 30 s: A.D. 1220–1235 (8.7%)A.D. 1240–1280 (59.5%) Poor agreement§2s: A.D. 1210–1290

3 KIA3215 660 6 30 s: A.D. 1290–1315 (23.4%) s: A.D. 1290–1315 (23.4%)A.D. 1350–1390 (44.8%) A.D. 1350–1390 (44.8%)2s: A.D. 1280–1330 (41.3%) 2s: A.D. 1280–1330 (41.3%)A.D. 1340–1400 (54.1%) A.D. 1340–1400 (54.1%)

Agreement—101.6%4 KIA3216 680 6 30 s: A.D. 1280–1310 (40.2%) s: A.D. 1295–1305 (7.8%)

A.D. 1360–1390 (28%) A.D. 1350–1390 (60.4%)2s: A.D. 1270–1330 (52.7%) 2s: A.D. 1290–1330 (23.6%)A.D. 1340–1400 (42.7%) A.D. 1340–1400 (71.8%)

Agreement—86.4%5 KIA3217 IV 690 6 30 s: A.D. 1280–1305 (49.8%) s: A.D. 1282–1299

A.D. 1370–1385 (18.4%) 2s: A.D. 1265–13252s: A.D. 1270–1330 (61.6%) Agreement—118.2%A.D. 1350–1390 (33.8%)

6 KIA3218 700 6 30 s: A.D. 1270–1310 (57.5%) s: A.D. 1282–1298A.D. 1370–1380 (10.7%) 2s: A.D. 1264–13082s: A.D. 1260–1320 (71.6%) Agreement—119.5%A.D. 1350–1390 (23.8%)

7 KIA3219 760 6 30 s: A.D. 1244–1285 s: A.D. 1267–12882s: A.D. 1220–1295 2s: A.D. 1254–1297

Agreement—111%8 AA43159 IV 722 6 36 s: A.D. 1263–1298 s: A.D. 1240–1280

2s: A.D. 1220–1310 (86%) 2s: A.D. 1220–1285A.D. 1360–1390 (9.4%) Agreement—52.2%

9 AA43158 814 6 37 s: A.D. 1190–1200 (2.9%) s: A.D. 1220–1263A.D. 1210–1275 (65.3%) 2s: A.D. 1190–12802s: A.D. 1160–1290 Agreement—116.3%

10 AA43157 IV 798 6 36 s: A.D. 1215–1275 s: A.D. 1160–1175 (8.1%)2s: A.D. 1160–1290 A.D. 1185–1235 (60.1%)

2s: A.D. 1120–1140 (1.5%)A.D. 1150–1260 (93.9%)Agreement—63.1%

11 KIA3220 III 1630 6 40 s: A.D. 360–370 (1.6%) s: A.D. 340–470)A.D. 380–470 (47.9%) 2s: A.D. 260–280 (1.4%)A.D. 480–540 (18.8%) A.D. 320–540 (94%)2s: A.D. 260–280 (1.7%) Agreement—101.7%A.D. 340–540 (93.7%)

12 KIA3221 III 1760 6 40 s: A.D. 230–350 (66.2%) s: A.D. 230–350A.D. 370–380 (2%) 2s: A.D. 210–4002s: A.D. 130–390 Agreement—108.4%

13 KIA3222 1800 6 50 s: A.D. 130–260 (57%) s: A.D. 130–260A.D. 280–290 (2.3%) 2s: A.D. 120–330A.D. 300–330 (8.9%) Agreement—109.9%2s: A.D. 80–390

14 KIA3223 III 1950 6 60 s: 40–30 B.C. (2.1%) s: A.D. 60–140 (65%)20–10 B.C. (3.1%) A.D. 160–170 (3.2%)A.D. 1–A.D. 130 (63%) 2s: A.D. 40–2202s: 100–70 B.C. (1.6%) Agreement—88.5%60 B.C.–A.D. 230 (93.8%)

15 KIA3224 III 1940 6 40 s: A.D. 20–90 (51.5%) s: A.D. 20–90A.D. 100–130 (16.7%) 2s: A.D. 1–1302s: 50 B.C.– A.D. 140 Agreement—111.3%

16 KIA3225 1930 6 50 s: A.D. 20–130 s: A.D. 20–95 (64.1%)2s: 50 B.C.– A.D. 220 A.D. 100–110 (4.1%)

2s: 20 B.C.–A.D. 140Agreement—116.9%

17 KIA3227A III 1990 6 40 s: 50 B.C.– A.D. 60 s: 45 B.C.–A.D. 302s: 100–70 B.C. (2.5%) 2s: 60 B.C.–A.D. 7060 B.C.–A.D. 90 (89.6%) Agreement—108.3%A.D. 100–130 (3.3%)

18 KIA3228 III 2120 6 40 s: 200–90 B.C. (62%) s: 115–40 B.C.80–60 B.C. (6.2%) 2s: 150 B.C.–A.D. 12s: 360–290 B.C. (9.5%) Agreement—88.1%240–40 B.C. (85.9%)

19 KIA3232 III 2050 6 40 s: 12 B.C.–A.D. 10 s: 170–125 B.C. (39.6%)2s: 170 B.C.–A.D. 30 (93.7%) 120–75 B.C. (28.6%)A.D. 40–50 (1.7%) 2s: 200–50 B.C.

Agreement—80.3%20 KIA3233 2120 6 30 s: 200–90 B.C. (68.2%) s: 200–185 B.C. (5.8%)

2s: 350–320 B.C. (4.3%) 180–110 B.C. (62.4%)210–40 B.C. (91.1%) 2s: 210–60 B.C.

Agreement—110.6%

(Continued )

eroded at the western part of the exposure.Fluvial units compose the top of the exposedsequence, increasing in thickness westwardinto the Ze’elim alluvial fan.

Nahal David Site

Two fine-grained units interfinger withshore ridges and fluvial deposits in the Davidsite exposure (Fig. 6). The lower unit is com-posed of alternating aragonite and clay-sizedetrital laminae and wedges out from east (1m thick) to west, where the aragonite laminaeretain their nature but the detritus is sandier.This unit disappears completely and interfin-gers with shore deposits at ;394.5 mbsl in thewesternmost exposure in the site (Fig. 6).

Bedded fluvial deposits that are covered bytwo horizons of thick aragonite crusts overliethe lower unit (Figs. 6A and 6B). These fluvialdeposits change laterally into a buried shoreridge at 401 mbsl. The shore ridge is com-posed of foresets, each with well-sorted grav-els of different grain size covered and ce-mented by thick aragonite crusts. The ridgeslopes toward the east and wedges out at 404mbsl (Fig. 6C).

The 20–30-cm-thick upper unit is com-posed of aragonite and clay-size detrital lam-inae and was deposited on the sloping surfaceof the underlying shore ridge. The unit inter-fingers to the west with a higher shore ridgeat 398.5 mbsl (Figs. 6A and 6B). This shoreridge is also composed of well-sorted foresetsand is covered with massive aragonite crusts.

Fluvial sediments dominate the rest of theupper part of the exposure. The modern retreateroded the evidence of younger lacustrineunits that are still preserved at the Ze’elimsite.

RADIOCARBON DATING

The chronology of the sections described atthe David and Ze’elim sites is essential for thereconstruction of the Dead Sea lake-levelcurve. The exposures are rich in organic de-bris of vegetation growing along streams andsprings around the Dead Sea (Ken-Tor et al.,2001a, b). By examination of the samples andtheir associated sediments, we were able toexclude contamination by the sparse recentvegetation found on top of the sections in thishyperarid saline environment.

We collected 46 samples from the Ze’elimA and B and David sites (Figs. 5 and 6, re-spectively); these samples are listed in Tables1, 2, and 3, respectively, which present theAMS (accelerator mass spectrometry) datingresults. The Ze’elim A exposure is composedmostly of lacustrine deposits, hence providing

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 565

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

TABLE 1. (Continued )

Sampleno.

Laboratoryno.

Unit 14C age(yr B.P.)

Calculated distribution†

(calendar years)Modeled distribution‡

(calendar years)

21 KIA3234 III 2230 6 30 s: 380–350 B.C. (9.3%) s: 370–350 B.C. (7.9%)320–230 B.C. (50.6%) 320–200 B.C. (60.3%)220–200 B.C. (8.3%) 2s: 390–200 B.C.2s: 390–200 B.C. Agreement—99.7%

22 KIA3235 2190 6 30 s: 360–280 B.C. (42.7%) s: 360–280 B.C. (45.5%)260–170 B.C. (25.5%) 260–190 B.C. (22.7%)2s: 380–160 B.C. 2s: 370–170 B.C.

Agreement—101.6%23 AA43156 Base of III 2485 6 40 s: 770–520 B.C. s: 600–500 B.C. (46.6%)

2s: 790–480 B.C. (86.7%) 490–480 B.C. (2.1%)470–410 B.C. (8.7%) 470–410 B.C. (19.5%)

2s: 670–400 B.C.Agreement—92.6%

24 AA42803 II 2518 6 35 s: 790–750 B.C. (13.4%) s: 690–660 B.C. (14.9%)690–660 B.C. (9.8%) 650–550 B.C. (53.3%)650–540 B.C. (45%) 2s: 740–520 B.C.2s: 800–510 B.C. Agreement—105.4%

25 AA42804 II 2445 6 34 s: 760–690 B.C. (22.4%) s: 760–680 B.C.550–410 B.C. (45.8%) 2s: 770–570 B.C.2s: 770–610 B.C. (38.3%) Agreement—96.2%600–400 B.C. (57.1%)

26 AA42805 II 2715 6 42 s: 900–825 B.C. s: 900–820 B.C.2s: 970–950 B.C. (2.3%) 2s: 920–800 B.C.930–800 B.C. (93.1%) Agreement—104.5%

27 AA42807 II 2775 6 37 s: 980–950 B.C. (10.9%) s: 1000–890 B.C.940–890 B.C. (34.1%) 2s: 1010–840 B.C.880–830 B.C. (23.2%) Agreement—101.2%2s: 1000–830 B.C.

28 AA42809 Base of II 3130 6 44 s: 1490–1480 B.C. (2.6%) s: 1440–1370 B.C. (53.9%)1450–1370 B.C. (52.5%) 1340–1310 B.C. (14.3%)1340–1310 B.C. (13.1%) 2s: 1500–1290 B.C. (92.9%)2s: 1520–1290 B.C. (93.9%) 1280–1260 B.C. (2.5%)1280–1260 B.C. (1.5%) Agreement—104.6%

29 AA42806 I 3220 6 36 s: 1520–1440 B.C. s: 1520–1445 B.C.2s: 1600–1560 B.C. (6.3%) 2s: 1600–1560 B.C. (6.7%)1530–1410 B.C. (89.1%) 1540–1410 B.C. (88.7%)

Agreement—100.4%30 AA42810 I 3703 6 37 s: 2140–2030 B.C. s: 2140–2020 B.C. (58%)

2s: 2200–1970 B.C. 2000–1970 B.C. (10.2%)2s: 2200–2160 B.C. (4.8%)2150–1940 B.C. (90.6%)Agreement—96.1%

†Calibration by the OxCal program v. 3.5 (Bronk Ramsey, 1995) of single dates with s and 2s probability(68.2% and 95.4%, respectively). The percent of probability distribution of each range is in parentheses.

‡Calibration of overlapping calendar date ranges using the stratigraphic order of the samples. The calibrationis calculated by using sequences that are defined as a group of events, which are known to follow one anotherwithout possible overlap. Agreement index represents the overlap between the calculated and modeleddistributions. Samples without overlapping age ranges below and above could not be refined by using the model.

§Sample excluded from the model analysis because of poor statistic agreement probably due to recycling. Thissample was not used in the reconstruction of the lake-level curve.

the most continuous chronology. With only afew exceptions, the analyzed radiocarbon agesare consistent with their stratigraphic order.The oldest age is 3703 6 37 radiocarbon yrB.P. and was recovered from the Ze’elim Asite. The youngest age is 93 6 36 yr B.P. andwas recovered from the upper part of theZe’elim B outcrop. Since the late 1960s, thetops of the sequences have been exposedabove lake-level.

The radiocarbon age of an organic samplecollected from a sedimentary section dates thetermination of carbon dioxide exchange of theplant with the atmosphere. The organic debrisis then transported into the lake by the sea-sonal floods; it can be recycled along thestreams and shores, and finally deposited andburied within the sediments that compose thesection. Yechieli (1993) and Ken-Tor et al.

(2001b) indicated that the time interval be-tween the beginning of the radiocarbon decayand the deposition of the organic debris isshort in the Dead Sea arid environment, fromless than a few decades to no more than twocenturies; i.e., the radiocarbon age of the or-ganic debris approximately dates the age ofdeposition in the section.

The radiocarbon ages were calibrated to cal-endar dates by the OxCal program (BronkRamsey, 1995, 2001) and are presented as amean and associated 1s and 2s standard er-rors (calculated distribution values in calendaryears are listed in Tables 1 to 3). The associ-ation of several calendar ranges with one ra-diocarbon age is a consequence of the ‘‘wig-gles’’ in the calibration curve. Because thestratigraphic order of the samples in each se-quence is known, we incorporate this con-

straint into the analysis of the radiocarbonages. Such prior information taken into ac-count in the calibration analysis results in amore precise chronology and reduced uncer-tainty of stratigraphically ordered overlappingcalendar ranges (modeled distribution in cal-endar years, listed in Tables 1–3). In using thistype of analysis, there is a risk of imposinginconsistent dates into the chronology (BronkRamsey, 1995); therefore, each radiocarbon-modeled distribution is presented with anagreement index (Table 1 to 3). An unaltereddistribution (modeled distribution overlapswith calculated distribution) will have anagreement of 100%, but it is possible for theagreement index to rise above 100%, if themodeled distribution overlaps the highestprobability of the prior distribution. Two sam-ples show poor agreement, probably becauseof recycling (780 6 30 and 279 6 20 yr B.P.;Tables 1 and 2, respectively). These sampleswere excluded from the calibration analysisand not used in the lake-level curvereconstruction.

RECONSTRUCTION OF LAKE-LEVELCURVE FOR THE PAST 4000 YR

The reconstruction of the late Holocenelake-level curve of the Dead Sea is based onlake-level indicators from the Ze’elim and Da-vid sites. For each columnar section describedpreviously, an independent lake-level curvewas interpreted on the basis of the informationabout depositional environments, lake-levelindicators, and radiocarbon ages (Figs. 5 and6). There were no substantial contradictionsbetween the individual curves (Fig. 7); thus,by combining the various curves, a compositelake-level curve for the Dead Sea was pro-duced (Fig. 8). The following is a detailed de-scription of the curve reconstruction.

The oldest sediments described (unit I, Fig.5) correspond to a lake level higher than 411mbsl dated to 2140–1445 B.C. (3703 6 37and 3220 6 36 radiocarbon yr B.P., Table 1).At that time the lake level was falling from anearlier highstand (prior to the fifteenth centuryB.C.), but no indicator for the absolute lake-level elevation was found at our sites. How-ever, unit I may correspond to a distinct shoreridge identified in a western location in theNahal Darga fan delta (Fig. 1C) at ;370 mbsl,where its age was estimated at ;3000–4000yr B.P. (Kadan, 1997). Unit I is eroded by un-conformity A. Organic debris within the ara-gonite crusts on the unconformity were datedbetween 1440 and 1120 B.C. (3130 6 44,2977 6 48, and 2985 6 39 yr B.P., Tables 1and 2). According to the data from the Ze’elimA outcrop, the lake dropped below 410.5 mbsl

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

566 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

TABLE 2. AMS RADIOCARBON DATES FROM THE ZE’ELIM B SITE

Sampleno.

Laboratoryno.

Unit 14C age(yr B.P.)

Calculated distribution†

(calendar years)Modeled distribution‡

(calendar years)

1 KIA1651 V 93 6 36 s: A.D. 1690–1730 (18.8%) s: A.D. 1690–1730 (15.5%)A.D. 1810–1850 (20%) A.D. 1810–1900 (52.7%)A.D. 1860–1920 (29.4%) 2s: A.D. 1680–1770 (27.6%)2s: A.D. 1670–1740 (27.1%) A.D. 1800–1930 (67.8%)A.D. 1800–1960 (68.3%) Agreement—100.9%

2 KIA8260 V 279 6 20 s: A.D. 1530–1550 (24%)A.D. 1630–1660 (44.2%) Poor agreement§2s: A.D. 1520–1580 (43.5%)A.D. 1620–1670 (51.9%)

3 KIA8261 V 135 6 31 s: A.D. 1670–1710 (12%) s: A.D. 1670–1740 (40.2%)A.D. 1720–1760 (13.3%) A.D. 1750–1760 (2.3%)A.D. 1800–1820 (5.7%) A.D. 1800–1860 (25.7%)A.D. 1830–1890 (24.7%) 2s: A.D. 1670–1780 (57.9%)A.D. 1910–1950 (12.5%) A.D. 1790–1890 (37.5%)2s: A.D. 1670–1780 (41.1%) Agreement—99.3%A.D. 1790–1960 (54.3%)

4 AA43163 V 236 6 39 s: A.D. 1630–1680 (38.3%) s: A.D. 1641–1670A.D. 1770–1810 (23.9%) 2s: A.D. 1520–1560 (2.4%)A.D. 1930–1950 (5.9%) A.D. 1620–1690 (82.1%)2s: A.D. 1520–1590 (10.2%) A.D. 1750–1810 (10.9%)A.D. 1620–1690 (41.5%) Agreement—131.1%A.D. 1730–1810 (36%)A.D. 1920–1950 (7.8%)

5 KIA8259 V 260 6 24 s: A.D. 1530–1540 (3.4%) s: A.D. 1641–1661A.D. 1630–1670 (64.8%) 2s: A.D. 1530–1550 (1.4%)2s: A.D. 1520–1570 (16.4%) A.D. 1630–1680 (86.1%)A.D. 1620–1670 (70.7%) A.D. 1780–1800 (8%)A.D. 1780–1800 (8.3%) Agreement—113.4%

6 KIA8258 Base of IV 909 6 23 s: A.D. 1040–1100 (44.2%)A.D. 1110–1210 (24%)2s: A.D. 1030–1210

7 AA43162 895 6 36 s: A.D. 1040–1090 (29.2%)A.D. 1120–1140 (12.5%)A.D. 1150–1210 (26.5%)2s: A.D. 1030–1220

8 AA43160 Base of II 2977 6 48 s: 1310–1120 B.C.2s: 1380–1330 B.C. (6.2%)1320–1040 B.C. (89.2%)

9 AA43161 2985 6 39 s: 1300–1120 B.C.2s: 1380–1330 B.C. (6.2%)1320–1050 B.C. (89.2%)

†Calibration by the OxCal program v. 3.5 (Bronk Ramsey, 1995) of single dates with s and 2s probability(68.2% and 95.4%, respectively). The percent of probability distribution of each range is in parentheses.

‡Calibration of overlapping calendar date ranges using the stratigraphic order of the samples. The calibrationis calculated by using sequences that are defined as a group of events, which are known to follow one anotherwithout possible overlap. Agreement index represents the overlap between the calculated and modeleddistributions. Samples without overlapping age ranges below and above could not be refined by using the model.

§Sample excluded from the model analysis because of poor statistic agreement probably due to recycling. Thissample was not used in the reconstruction of the lake-level curve.

before rising again. In the Nahal Darga fandelta, this unconformity was identified at 414mbsl (Kadan, 1997); in the Sedom diapir (Fig.1C), narrow salt-cave passages, which repre-sent arid conditions (Frumkin et al., 1991),correlate to this drop in lake level (Fig. 8).

Shore indicators at elevations between 409and 407 mbsl at the Ze’elim A outcrop indi-cate that lacustrine unit II (Fig. 5) was depos-ited under conditions of relatively low lakelevels that fluctuated between 1000 and 550B.C. (2775 6 37 to 2518 6 35 yr B.P., Table1). At the Sedom diapir, this period was theweakest of the wet periods (Frumkin et al.,1991).

Unconformity B (Fig. 5) marks a drop inlake level at least to 407 mbsl. This drop oc-curred between 690 and 410 B.C. (2518 6 35and 2485 6 40 yr B.P., Table 1).

Unit III was deposited in the Ze’elim site(Fig. 5) under relatively high lake-level con-ditions. Radiocarbon ages indicate that at thesame time the lake reached 394.5 mbsl in theDavid site, where the lacustrine lower unitthins out and interfingers with shore deposits(Fig. 6). Radiocarbon samples from theseunits cover the time range between 370 B.C.and A.D. 350 (2230 6 30 to 1760 6 40 yrB.P., Table 1; 2128 6 27 to 1771 6 30 yrB.P., Table 3). The rising trend began duringthe third and second century B.C. and reachedits highest stand in the first century B.C. His-torical and archaeological evidence strengthenthis determination. Rujm el Bahr was an ac-tive mooring place for boats in the northernshore of the Dead Sea (Fig. 1C) dated to thesecond and first century B.C., according to theconstruction stones (Bar-Adon, 1989). An-

choring stones were found at the highstand el-evation recorded at the David site. Klein(1986) reported on a very rainy period duringthe first century B.C. that ended just beforecommencement of the Christian calendar,when the lake level started to drop. It was pro-posed that this highstand reached higher ele-vations (330 mbsl; Klein, 1986), but this sug-gestion later proved to be an overestimate(Frumkin et al., 1991; Neev and Emery,1995). Other works in the Dead Sea area sug-gest a highstand during that period. Wide salt-cave passages in the Sedom diapir (Frumkinet al., 1991), lacustrine sediments in exposuresin the Nahal Darga fan delta (Kadan, 1997;Enzel et al., 2000), and a laminated aragoniteand fine-grained detritus sequence in a saltcore drilled in the northern basin (Heim et al.,1997) all represent enhanced input of runoffthat raised the lake level.

A second highstand is recorded in the upperpart of unit III in the Ze’elim A exposure (Fig.5). Limited information from the Ze’elim Asediments indicates that after a relatively fastdrop to 404 mbsl, represented by a thin sandand aragonite crust unit, the lake level roseagain. The rising of the level lasted for a shortperiod, and the organic sample collected fromthe middle of the lacustrine layer marking thishighstand was dated to A.D. 340–470 (16306 40, Table 1). Evidence for this highstandelevation was recovered by using archaeolog-ical information. An excavated structure inEin Gedi (Fig. 1C) at an elevation of ;390mbsl was dated at between A.D. 70 and 130(Mazar and Dunayevsky, 1964). The occupa-tion level in the structure, which is a bath-house, is covered by shore deposits that lacklate Roman and Byzantine artifacts, althoughthe village was occupied at this period. Thearchaeologically sterile shore deposits attestthat the lake level reached the structure ele-vation at the latest by A.D. 300 and coveredit, preventing human use (Fig. 8; Y. Hirsch-feld, 2003, personal commun.).

After the continuous deposition of unit IIIover a few centuries, the development of un-conformity C in the Ze’elim site marked anerosion period during a drop in lake level (Fig.5). The level elevation was lower than 404mbsl, according to the shore ridges in the Da-vid site (Fig. 6C). This drop in lake level be-gan after the fifth century A.D., and the firstsediments deposited on top of unconformity Care dated to the eleventh century A.D. and lat-er times (Fig. 5) (909 6 23 and 895 6 36 yrB.P., Table 2). This is the longest hiatus in thechronology. Because a sedimentary sequenceof a few hundred years was eroded, there is apossibility that information on lake-level fluc-tuation was lost. Historic documentation on

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 567

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

TABLE 3. AMS radiocarbon dates from the David site

Sampleno.

Laboratoryno.

Unit 14C age(yr B.P.)

Calculated distribution†

(calendar years)Modeled distribution‡

(calendar years)

1 KIA11646 Modern 245 6 28 s: A.D. 1640–1670 (51.8%)A.D. 1780–1800 (16.4%)2s: A.D. 1520–1560 (6.8%)A.D. 1630–1680 (60.9%)A.D. 1770–1810 (23.9%)A.D. 1930–1950 (3.8%)

2 AA43397 Upper 833 6 36 s: A.D. 1165–1170 (2.2%)A.D. 1185–1260 (66%)2s: A.D. 1060–1090 (2.2%)A.D. 1120–1140 (2.5%)A.D. 1150–1280 (90.6%)

3 KIA11645 Upper 968 6 26 s: A.D. 1020–1050 (26%)A.D. 1090–1120 (25.3%)A.D. 1140–1160 (16.9%)2s: A.D. 1000–1160

4 KIA11647 Lower 1771 6 30 s: A.D. 230–340 s: A.D. 130–150 (5.2%)2s: A.D. 130–350 (94.1%) A.D. 210–330 (63%)A.D. 370–380 (1.3%) 2s: A.D. 130–350

Agreement—92.8%5 KIA11649 Lower 2163 6 27 s: 360–310 B.C. (28%) s: 205–150 B.C. (42.2%)

230–160 B.C. (40.2%) 140–110 B.C. (26%)2s: 360–280 B.C. (38.4%) 2s: 340–290 B.C. (2.8%)260–110 B.C. (57%) 240–50 B.C. (92.6%)

Agreement—68.5%6 KIA11648 Lower 2128 6 27 s: 200–105 B.C. s: 350–320 B.C. (12.6%)

2s: 350–320 B.C. (7%) 210–150 B.C. (55.6%)210–50 B.C. (88.4%) 2s: 360–310 B.C. (18%)

240–100 B.C. (77.4%)Agreement—94.7%

†Calibration by the OxCal program v. 3.5 (Bronk Ramsey, 1995) of single dates with s and 2s probability(68.2% and 95.4%, respectively). The percent of probability distribution of each range is in parentheses.

‡Calibration of overlapping calendar date ranges using the stratigraphic order of the samples. The calibrationis calculated by using sequences that are defined as a group of events, which are known to follow one anotherwithout possible overlap. Agreement index represents the overlap between the calculated and modeleddistributions. Samples without overlapping age ranges below and above could not be refined by using the model.

harsh drought years, hunger, and plagues arereported during the sixth to ninth centuriesA.D. (Klein, 1986). Salt deposition in thenorthern basin (Heim et al., 1997) and narrowsalt-cave passages in the Sedom diapir (Frum-kin et al., 1991) all indicate that at least sincethe sixth century A.D., the lake level was rel-atively low, and the rising trend started onlyafter the ninth century A.D. (Fig. 8).

In the David site, the upper unit (Fig. 6)marks a rapid, short lake-level rise to 398.5mbsl between A.D. 1020 and 1260 (968 6 26and 833 6 36 yr B.P., Table 3). This age issimilar to the ages of samples collected justabove unconformity C from the Ze’elim B site(;402 mbsl, Fig. 5), also indicating a fast lev-el rise. At the same time, wide salt-cave pas-sages were created in the Sedom diapir (Fig.8) (Frumkin et al., 1991), and historical doc-uments from the tenth and eleventh centuriesA.D. indicate a rainy period and prosperity inthe region (Klein, 1986) that may have coin-cided with the initiation of this lake rise. La-custrine conditions continued at least untilA.D. 1390 (Fig. 5) (798 6 36 to 660 6 30 yrB.P., Table 1).

After the thirteenth century A.D. and aboveunit IV, most of the information comes fromthe Ze’elim sites that were not eroded duringthe drastic modern retreat of the Dead Sea, as

occurred in the David site. At A.D. 1381–1410 (600 6 40 yr B.P., Table 1), the lakelevel dropped to ;402 mbsl, according to theinformation from the Ze’elim A site (uncon-formity D, Fig. 5). Because the time intervalbetween the dates below and above the un-conformity in the Ze’elim A exposure is short,the lake level probably did not drop signifi-cantly below this level, and only minimal ero-sion occurred. In the Ze’elim B exposure, thesamples collected from unconformity D at thesame elevation were much younger, dating toA.D. 1641–1670 (260 6 24 and 236 6 39 yrB.P., Table 2). The difference in dates mayrepresent an exposure period between thefourteenth to the seventeenth centuries A.D.,when lake level was relatively stable at ;402mbsl.

Unit V was deposited above the unconfor-mity D surface (Fig. 5), marking a rise in lakelevel. Despite the difficulty in establishing anaccurate chronology for the past four centuriesbecause of wide ranges of calibrated radiocar-bon ages, correlation to documented historicevents established a reliable time frame (Ken-Tor et al., 2001a, b). Shore environment re-corded in unit V at 400–401 mbsl (Fig. 5) isdated to A.D. 1670–1860 (135 6 31 yr B.P.,Table 2). The dated sample was collected froma sand layer, which was liquefied by an earth-

quake in A.D. 1834 (Ken-Tor et al., 2001a, b).According to documented records on the his-toric Dead Sea during that year, the level wasat ;395 mbsl (Klein, 1961, 1986), implyinga water depth of ;5 m at Ze’elim.

The uppermost laminated sequence in unitV (Fig. 5) was dated to A.D. 1690–1900 (936 36 yr B.P., Table 2) and marks a distinctrise in lake level. Despite the wide calendar-age range, this lacustrine sequence is correlat-ed with the end of the nineteenth-century lakerise, which was measured instrumentally (Fig.8; Klein, 1986; Underhill, 1967). The lacus-trine laminated sequence provides a minimallevel of 400–402 mbsl, and the known histor-ic lake-level maximum elevation is ;390mbsl, meaning the lacustrine sediments weredeposited at about 10 m depth on the lakefloor. The base of the highest fine-sand unit isat 401 mbsl, and it was liquefied by the laststrong earthquake in the Dead Sea area, the1927 earthquake (Fig. 5; Ken-Tor et al.,2001a, b; Shapira et al., 1993). The historiclake level during that year was still high, at;390 mbsl (Klein, 1986).

DISCUSSION

The reconstructed lake-level curve of thelate Holocene Dead Sea (Fig. 8) shows thatthe lake level fluctuated within 25 m duringthe past 4000 yr. For most of this period thelevel did not reach above the sill (402 mbsl)separating the northern and southern basins ofthe Dead Sea. In the following, we discuss thepattern revealed by the curve and consider theimplications of the sedimentological recordregarding the limnological and hydrologicconditions. We also examine the relationshipbetween regional climatic conditions as in-ferred from the lake levels and cultural shiftsin the region.

Nonclimatic factors could have affected thereconstructed lake-level curve. The major fac-tors are the tectonic subsidence of the basinfloor that creates accommodation space andthe rate of accumulation of sediment that fillsit. The question is, what was the relative con-tribution of the tectonic subsidence vs. that ofthe sediment accumulation and has tectonicsubsidence affected the presented curve? Weaddress this question by first considering therole of sedimentation, and we use today’s sillelevation as a datum.

When the lake level exceeded the sill andflooded the southern basin (Fig. 1D), sedi-ments were deposited over the sill and in-creased its elevation. The cumulative increasein elevation is estimated as ;10 m (412–402mbsl) from the line drawn on the basis of theprogressive increase of the shore indicators’

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

568 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al. on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 569

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

Figure 8. (A) The composite lake-level curve of the Dead Sea during the past 4000 yr. Calibrated radiocarbon dates with 1s ranges aremarked on the curve at the absolute elevation of the samples. Sample numbers marked by A, B, and D are referenced from Table 1(Ze’elim A site), Table 2 (Ze’elim B site), and Table 3 (David site), respectively. The thick line at the bottom of the curve emphasizesthe sediment accumulation at the Ze’elim site (see text). The inset at the top shows the measured historic lake-level curve since A.D.1800 (modified from Klein, 1986, and extended to the year 2000 by using the Hydrological Service data). (B) The Dead Sea lake-levelcurve reconstructed according to Mount Sedom caves (Fig. 1C) and assuming a diapir rising rate of 6 mm/yr (Frumkin, 1997; Frumkinet al., 2001). The radiocarbon dates adjusted to the calendar range are marked with horizontal bars and their sample number (Table11.1 in Frumkin, 1992). (C) Core DS 7–1 SC drilled offshore Ein Gedi in the deep northern basin (Fig. 1C). The horizontal extensionof the core represents its depth in meters. Calibrated radiocarbon dates are marked with horizontal bars and the sample number (Heimet al., 1997). Their locations in the core are marked with arrows, and they are adjusted to the calendar range of the Dead Sea curve.N

elevation at the Ze’elim site (Fig. 8), locatedin close proximity to the sill (Fig. 1C). Duringmost of the period, the lake level was adjustedto this buffering elevation as expected; how-ever, the deviation from this datum line andflooding of the southern basin implies a sig-nificant hydrologic change.

During flooding of the southern basin, en-hanced evaporation from the large additionalarea of shallow waters (;35%, Fig. 1D; Hall,1997; Bartov et al., 2002) had buffered therising lake level. Furthermore, the depositionof laminated aragonite during highstandspoints to stratified conditions in the lake.These conditions imply that a low-salinity up-per water layer developed and improved theevaporation from the surface waters becauseof their high activity. During negative fresh-water input, lake level had fallen below thesill elevation in the northern deep basin. As aresult, the concentration of the lake brine hadincreased, the activity of the saline water haddecreased, and the falling of the lake level wasbuffered and stabilized. Therefore, each timethe Dead Sea overtops the topographic thresh-old and rises to a higher elevation or dropsbelow, it indicates a significant change infreshwater input.

The long-term tectonic subsidence rate canbe estimated from normal faults that were ac-tive during the Lake Lisan period at the Mas-sada plain, ;5 km to the southwest of theZe’elim site (Marco and Agnon, 1995). Thesefaults show vertical displacements of a fewmeters, yielding an average subsidence rate of,0.1 mm/yr, which is smaller by an order ofmagnitude than the calculated sedimentationrate of ;1 mm/yr for the Lisan Formation(Stein, 2001). Post-Lisan and Holocene dis-placements were recognized on the margins ofthe Massada plain and yield subsidence ratesin the range of 0.3–0.6 mm/yr (Bartov andSagy, 2003), which are lower than the sedi-mentation rates calculated in the Ze’elim sitenearby (3–13 mm/yr, Ken-Tor et al., 2001a).This result implies that the tectonic subsidencehas been compensated by the sediment accu-mulation at the margins of the lake where the

field work was done, at least during the LateHolocene.

The climate influence on the lake level ofthe Dead Sea is a continuous process that canovercome the effects of temporary earthquakeevents that can change the level; i.e., a changein the hypsometric curve of the lake due to anearthquake throw is not expected to be main-tained for a long period. In addition, no dis-crepancy appears between paleoshore eleva-tions in the contemporaneous sites, and nodisturbance appears in the datum line recon-structed from the shore indicators at theZe’elim site.

The lake-level curve, which is based onsedimentological interpretation, can be com-pared to the pattern of the instrumental lakelevels measured during the nineteenth andtwentieth centuries (Fig. 8; Klein, 1961, 1986;Underhill, 1967). The high-resolution instru-mental record displays annual saw-tooth–likefluctuations that are superimposed on a longertrend that reaches a maximum at the end ofthe nineteenth century and then declines. Thereconstructed curve exhibits a similar generalpattern. Although the small annual fluctua-tions are smoothed out in the sedimentologicalrecord, the major changes in the hydrologicconditions such as the rise at the end of thenineteenth century are revealed in the recon-structed curve. The overall long trends of ris-ing or dropping levels reflect the accumulationand subtraction of rainy and dry years fromthe Dead Sea water balance, respectively, re-cording the history of the drainage area in thescale of decades to centuries.

Lake-Level Rises

The Dead Sea level clearly rose above thesill four times, flooding the southern basin forperiods of a few decades to a few centuries.Highstands occurred during the Roman andByzantine periods in the second and first cen-turies B.C. and the fourth century A.D., re-spectively; during the Crusader period in theeleventh and twelfth centuries A.D., and from

the end of the nineteenth century A.D. to thebeginning of the twentieth century.

The lake-level curve shows good correla-tion with the curve reconstructed from eleva-tions of relict cave passages dated by radio-carbon on organic samples in the Sedom saltdiapir at the western shore of the southern ba-sin (Fig. 1C) (Frumkin et al., 1991; Frumkin,1997). Overall, the records are synchronized,yet some of the highstands’ correspondence intime is not perfect (Fig. 8). The rises in lakelevel at stages 5 and 9 in the Sedom recordprecede the timing of lake-level rises in ourcurve. This discrepancy can be explained bythe instantaneous response of the salt rock tothe change in discharge in the caves, whereasthe rise in lake level may take many years(Frumkin et al., 1991). Stage 7 correlates intime to the two sequential highstands aroundthe beginning of the Christian era that are re-corded by unit III in the sedimentary record;however, the high solubility of the salt prob-ably could not allow the cave data to be sen-sitive enough, so the two events were recordedas one. In addition, the maximum highstandelevations in this study are lower than the Se-dom record highstands (Fig. 8). In the Sedomdiapir record, the elevations were determinedby taking into account the rising rate of thediapir (6 mm/yr in Fig. 8), which may varytemporally and spatially. When applying high-er rates, this discrepancy disappears (Frumkin,1997).

Further information on the limnologicalconditions during highstands comes fromcomparing the Ze’elim and David sedimentarysequences on the lake margins to a core fromthe deeper environment of the lake offshoreEin Gedi (core DS 7–1 SC drilled at 203 mbsl;Fig. 1C; Heim et al., 1997). This core revealsan evaporitic sequence of several salt layersintercalated with laminated aragonite and fine-grained detritus. The laminated sequences arecorrelative to the high levels seen in the lake-level curve (Fig. 8). The lower laminated se-quence in the core correlates to the highstanddated to the second and first centuries B.C.The upper laminated sequence correlates to

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

570 Geological Society of America Bulletin, May/June 2004

BOOKMAN (KEN-TOR) et al.

the highstand at the end of nineteenth century.A few other undated thin detrital layers withinthe salt could be correlated to the highstandsduring the fourth century and the eleventh andtwelfth centuries A.D. This agreement be-tween the lake-level chronology and the coredsediments demonstrates that during enhancedfreshwater input and higher lake levels, bicar-bonate and suspended detrital materialbrought by the freshwater floods entered thedeeper part of the lake. This supply of freshwater and sediments interrupts the salt depo-sition from the supersaturated Dead Sea brineand causes the formation of a density-layeredlake structure that enables the precipitation ofaragonite laminae from the upper water layer(Barkan et al., 2001; Stein et al., 1997).

Lake-Level Falls

Five unconformities representing significantlowstands—i.e., when the level dropped to be-low 402 mbsl—are recorded in the curve (Fig.8). These unconformities are caused by thefall of the lake level to below the elevationrecorded by the respective hiatuses. Most ofthe unconformities represent sedimentary hi-atuses, but only the relatively dense radiocar-bon dating enables the distinction between un-conformities that were formed by long orshorter lowstands and the subsequent levelrise. The data from the Ze’elim and David ex-posures do not allow us to determine the min-imum stand of the lake during the past 4000yr. However, the core DS En that was drilledat the Ein Gedi shore at an elevation of ;410mbsl (Fig. 1C) reveals continuous sedimenta-tion and no unconformities in its top 4.5 m,which were dated to the past 3000 yr (Mi-gowski et al., 1998). This result indicates thatlake level did not drop below 415 mbsl duringthis time interval.

For the past few decades, the Dead Sea lev-el has been declining faster (;0.8 m/yr) thanany time in the resolved geologic record. Thelevel fell from a highstand of 392 mbsl (1895–1930, Klein, 1986) to the present (2003) ele-vation of 417 mbsl. This dramatic fall hasbeen primarily artificially induced by the di-version of runoff water for human needs. Al-though the lake-level curve indicates signifi-cant lowstands in the late Holocene, thechanges in level never exceeded the modernrate. At the same time, however, the lake-levelcurve indicates that extreme long-termdroughts occurred frequently over the pastseveral thousand years.

Climate Implications and CulturalChanges

The study of lake-level changes in the lateHolocene Dead Sea suggests significant fluc-

tuations in the regional hydrologic conditions.Transferring lake-level variations into morespecific regional climatology requires quanti-fication and correlation with the modern ana-logues. The analogues are the maintained rainrecord from Jerusalem since the mid nine-teenth century (e.g., Amiran, 1995) and mea-sured lake levels of the Dead Sea previous tothe significant interference in its water balancein the 1960s (Klein, 1986; Underhill, 1967).Correlation and analysis of these modern rec-ords enable Enzel et al. (2003) to determinethat episodes of lake-level rises and falls arecharacterized by an average of 650 mm/yr and445 mm/yr, respectively, in Jerusalem. Histor-ical reports of individual drought years werealways associated with ,500 mm of rainfallin Jerusalem (Klein, 1986). As Jerusalem rain-fall is highly correlated with many rain sta-tions in northern Israel, the headwaters of theDead Sea, the climatic implications of the lakefluctuations could be extended to larger areasin the Levant that are under similar atmo-spheric conditions (Enzel et al., 2003). Thischange in annual rainfall that was responsiblefor a shortage or an excess of water supplycould have influenced drastically the arid andsemiarid environments of the area.

The question to address is whether suchchanges in the amount of rainfall affected thehistorical inhabitants of the region. The lakerises of the second and first centuries B.C. andthe fourth century A.D. coincided with theprosperity of the Roman and Byzantine peri-ods proposed by Issar et al. (1992). Duringthese periods there was an increase in popu-lation in the region, and desert margins weresettled. The next significant rise in the DeadSea level occurred during the eleventh andtwelfth centuries A.D., in the Crusader period.Comprehensive archaeological evidence fromthis period suggests widespread agriculturaldevelopment in the region (Ellenblum, 1991),which could reflect favorite hydrologic con-ditions, as indicated by the lake curve. Theenhanced rainfall during these periods did notautomatically lead to cultural flourishing, butit certainly could have reinforced and stabi-lized these cultures, which were highly depen-dent on water resources and local agriculture.

The longest lake-level drop can also be con-nected to cultural changes in the region. Thehiatus in the curve that started in the fifth cen-tury A.D. and continued until the tenth cen-tury A.D. coincides with the termination ofthe Byzantine regime in the Near East and theseventh century A.D. invasion of Moslem-Arab tribes into the area (Issar et al., 1992).Although other historical and political factorscould have influenced this cultural change, the

coincidence with long arid conditions isnoteworthy.

CONCLUSIONS

1. A new high-resolution lake-level curvewas reconstructed for the late Holocene DeadSea from dense radiocarbon dating of detailedsedimentary sections and level indicators inthe Ze’elim plain and the Nahal David fandelta.

2. The levels of the Dead Sea fluctuatedwithin the range of 390 to 415 mbsl duringthe past 4000 yr. Most of the time, the lakelevel was below 402 mbsl, and the Dead Seawas confined to its deep northern basin. Nev-ertheless, short-term rises in the Dead Sea lev-el caused the flooding of the southern shallowbasin.

3. The significant high- and lowstands arecorrelated with cultural shifts in the region.The highstands of the second and first centu-ries B.C., the fourth century A.D., and theeleventh and twelfth centuries A.D. coincidewith the Roman, Byzantine, and Crusader pe-riods, respectively. The lake drop that startedafter the fifth century A.D. and prevailed atleast until the ninth century A.D. coincidedwith the decline of the Byzantine regime andinvasion of the Arab tribes into the area.

4. The drop from the late nineteenth centuryhighstand of 390 mbsl to the present elevationof 417 mbsl is mainly due to artificial waterdiversion. The modern rate of lake-level fallsignificantly exceeds all natural drops in theDead Sea history. However, drops of similarmagnitude are not unusual in the late Holo-cene record, which yields evidence of long pe-riods of droughts in the region.

ACKNOWLEDGMENTS

We thank Shmulik Marco, Nicolas Waldmann,Claudia Migowski, Yuval Bartov, Amos Frumkin,Galit Kadan, and the Geological Survey of Israeltechnicians for productive discussions and assis-tance in the field work. We also thank George Burrand the National Science Foundation–Arizona AMSLaboratory for their hospitality, and Tamar Sofer forthe draft of the Dead Sea topographic map. Themanuscript benefited from the review of NicolePetit-Maire and an anonymous referee. This re-search was supported by the Israel Science Foun-dation (grant 504.99–13), the German-Israel Foun-dation for Research and Development (G-642–113.8/1999), and the Ring Family Foundation forAtmospheric Global Studies.

REFERENCES CITED

Adams, K.D., and Wesnousky, S.G., 1998, Shoreline pro-cesses and age of the Lake Lahontan high-stand in theJessup embayment, Nevada: Geological Society ofAmerica Bulletin, v. 110, p. 1318–1332.

Amiran, D.H.K., 1995, Climatic data from Jerusalem: Je-

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from

Geological Society of America Bulletin, May/June 2004 571

LATE HOLOCENE LAKE LEVELS OF THE DEAD SEA

rusalem, Jerusalem Institute for Israel Studies, 141 p.(in Hebrew, introduction in English).

Anderson, R.Y., 2001, Rapid changes in late Pleistoceneand stream discharge determined from medium- andcoarse-grained sediment in saline lakes: Global andPlanetary Change, v. 28, p. 73–83.

Bar-Adon, P., 1989, Excavations in the Judean Desert, inZusman, A., and Straws, D., eds., Antiquities: Jeru-salem, The Antiquities Authorities, p. 3–14 (inHebrew).

Barkan, E., Luz, B., and Lazar, B., 2001, Dynamics of thecarbon dioxide system in the Dead Sea: Geochimicaet Cosmochimica Acta, v. 65, p. 355–368.

Bartov, Y., and Sagy, A., 2003, Recent faulting in the DeadSea—Pull-apart structure in the Lisan Formation: Is-rael Geological Society, field trip guidebook: Israel,Ein Boqeq, p. 1–16.

Bartov, Y., Stein, M., Enzel, Y., Agnon, A., and Reches,Z., 2002, Lake levels and sequence stratigraphy ofLake Lisan, the late Pleistocene precursor of the DeadSea: Quaternary Research, v. 57, p. 9–21.

Begin, B.Z., Ehrlich, A., and Nathan, Y., 1974, LakeLisan—The Pleistocene precursor of the Dead Sea:Geological Survey of Israel Bulletin, v. 63, 30 p.

BenDavid-Novak, H., 1998, Modern and Holocene debrisflows along the western escarpment of the Dead Sea[M.S. thesis]: Jerusalem, Hebrew University, 129 p.(in Hebrew, English abstract).

Bowman, D., 1990, Climatically triggered Gilbert-type la-custrine fan deltas, the Dead Sea area, Israel, in Co-lella, A., and Prior, D.B., eds., Coarse-grained deltas:International Association of Sedimentologists SpecialPublication 10, p. 273–280.

Bowman, D., 1997, Geomorphology of the Dead Sea west-ern margin, in Niemi, T.M., et al., eds., The DeadSea—The lake and its setting: Oxford, Oxford Uni-versity Press, p. 217–225.

Bronk Ramsey, C., 1995, Radiocarbon calibration and anal-ysis of stratigraphy: The OxCal program: Radiocar-bon, v. 37, p. 425–430.

Bronk Ramsey, C., 2001, Development of the radiocarbonprogram OxCal: Radiocarbon, v. 43, p. 355–363.

de Raff, J.F.M., Boersma, J.R., and van-Gelder, A., 1977,Wave-generated structures and sequences from a shal-low marine succession, Lower Carboniferous: Sedi-mentology, v. 24, p. 451–483.

Ellenblum, R., 1991, Frankish rural settlement in crusaderPalestine [Ph.D. thesis]: Jerusalem, Hebrew Univer-sity, 383 p. (in Hebrew).

Enzel, Y., Kadan, G., and Eyal, Y., 2000, Holocene earth-quakes inferred from a fan-delta sequence in the DeadSea graben: Quaternary Research, v. 53, p. 137–144.

Enzel, Y., Bookman (Ken-Tor), R., Sharon, D., Gvirtzman,H., Dayan, U., Ziv, B., and Stein, M., 2003, Late Ho-locene climates of the Near East deduced from DeadSea level variations and regional winter rainfall: Qua-ternary Research, v. 60, p. 263–273.

Frumkin, A., 1992, The karst system of the Mount Sedomdiapir [Ph.D. thesis]: Jerusalem, Hebrew University,210 p. (in Hebrew, English abstract).

Frumkin, A., 1997, The Holocene history of the Dead Sealevels, in Niemi, T.M., et al., eds., The Dead Sea—The lake and its setting: Oxford, Oxford UniversityPress, p. 237–248.

Frumkin, A., Magaritz, M., Carmi, I., and Zak, I., 1991,The Holocene climatic record of the salt caves of

Mount Sedom, Israel: The Holocene, v. 1,p. 191–200.

Frumkin, A., Kadan, G., Enzel, Y., and Eyal, Y., 2001, Ra-diocarbon chronology of the Holocene Dead Sea: At-tempting a regional correlation: Radiocarbon, v. 43,p. 1179–1190.

Garber, R.A., 1980, The sedimentology of the Dead Sea[Ph.D. thesis]: Troy, New York, Rensselaer Polytech-nic Institute, 169 p.

Garber, R.A., and Friedman, G.M., 1983, Coated grainsalong the Dead Sea shore in Peryt Tadeusz, M., ed.,Coated grains: Berlin, Springer-Verlag, p. 163–168.

Garber, R.A., Levy, Y., and Friedman, G.M., 1987, Thesedimentology of the Dead Sea: Carbonates and Evap-orites, v. 2, p. 43–57.

Hall, J.K., 1997, Topography and bathymetry of the DeadSea depression, in Niemi, T.M., et al., eds., The DeadSea—The lake and its setting: Oxford, Oxford Uni-versity Press, p. 11–21.

Heim, C., Nowaczyk, N.R., Negendank, J.F.W., Leroy,S.A.G., and Ben-Avraham, Z., 1997, Near East de-sertification: Evidence from the Dead Sea: Naturwis-senschaften, v. 84, p. 398–401.

Issar, A.S., Govrin, Y., Geyh, M.A., Wakshal, E., and Wolf,M., 1992, Climate change during the upper Holocenein Israel: Israel Journal of Earth-Sciences, v. 40,p. 219–223.

Kadan, G., 1997, Evidence for Dead Sea lake-level fluc-tuations and recent tectonism from the Holocene fan-delta of Nahal Darga [M.S. thesis]: Beer-Sheva, Ben-Gurion University of the Negev, 54 p. (in Hebrew,English abstract).

Katz, A., and Kolodny, N., 1989, Hypersaline brine dia-genesis and evolution in the Dead Sea–Lake Lisansystem (Israel): Geochimica et Cosmochimica Acta,v. 53, p. 59–67.

Katz, A., Kolodny, Y., and Nissenbaum, A., 1977, The geo-chemical evolution of the Pleistocene Lake Lisan–Dead Sea system: Geochimica et Cosmochimica Acta,v. 41, p. 1609–1626.

Ken-Tor, R., Agnon, A., Enzel, Y., Stein, M., Marco, S.,and Negendank, J.F.W., 2001a, High-resolution geo-logical record of historic earthquakes in the Dead Seabasin: Journal of Geophysical Research, v. 106,p. 2221–2234.

Ken-Tor, R., Stein, M., Enzel, Y., Agnon, A., Marco, S.,and Negendank, J.F.W., 2001b, Precision of calibratedradiocarbon ages of historic earthquakes in the DeadSea basin: Radiocarbon, v. 43, p. 1371–1382.

Klein, C., 1961, On the fluctuations of the level of the DeadSea since the beginning of the 19th century: Jerusalem,Hydrological Service, Hydrological Paper 7, revisededition, 83 p.

Klein, C., 1986, Fluctuations of the level of the Dead Seaand climatic fluctuations in Israel during historicaltimes [Ph.D. thesis]: Jerusalem, Hebrew University,(in Hebrew, English abstract), 208 p.

Machlus, M., Enzel, Y., Goldstein, S.L., Marco, S., andStein, M., 2000, Reconstructing low levels of LakeLisan by correlating fan-delta and lacustrine deposits:Quaternary International, v. 73/74, p. 137–144.

Manspeizer, W., 1985, The Dead Sea rift: Impact of climateand tectonism on Pleistocene and Holocene sedimen-tation, in Biddle, K.T., and Christie-Blick, N., eds.,Strike-slip deformation, basin formation and sedimen-tation: Society of Economic Paleontologists and Min-eralogists Special Publication 37, p. 143–158.

Marco, S., and Agnon, A., 1995, Prehistoric earthquake de-formations near Masada, Dead Sea graben: Geology,v. 23, p. 695–698.

Mazar, B., and Dunayevsky, I., 1964, En-Gedi—Third sea-son of excavations, preliminary report: Israel Explo-ration Journal, v. 14, p. 121–130.

Migowski, C., Ken-Tor, R., Negendank, J.F.W., Stein, M.,and Mingram, J., 1998, Post-Lisan record documentedin sediment sequences from the western shore areaand the central basin of the Dead Sea: Terra Nostra98/6: 3rd ELDP Workshop, Ptolemais: Bonn, Alfred-Wegener-Stiftung, p. 100–102.

Neev, D., and Emery, K.O., 1967, The Dead Sea, deposi-tional processes and environments of evaporates: Geo-logical Survey of Israel Bulletin, v. 41, 147 p.

Neev, D., and Emery, K.O., 1995, The destruction of Sod-om, Gomorrah, and Jericho: Oxford, Oxford Univer-sity Press, 175 p.

Nemec, W., and Steel, R.J., 1984, Alluvial and coastal con-glomerates: Their significant features and commentson gravelly mass-flow deposits, in Koster, E.H., andSteel, R.J., eds., Sedimentology of gravels and con-glomerates: Canadian Society of Petroleum GeologistsMemoir 10, p. 1–31.

Orford, J.D., Carter, R.W.G., and Jennings, S.C., 1991,Coarse clastic barrier environments: Evolution andimplications for Quaternary sea level interpretation:Quaternary International, v. 9, p. 87–104.

Reineck, H.-E., and Singh, I.B., 1973, Depositional sedi-mentary environments, with reference to terrigenousclastics: Berlin, Springer-Verlag, 439 p.

Shapira, A., Avni, R., and Nur, A., 1993, A new estimatefor the epicenter of the Jericho earthquake of 11 July1927: Israel Journal of Earth-Sciences, v. 42,p. 93–96.

Sneh, A., 1979, Late Pleistocene fan-deltas along the DeadSea rift: Journal of Sedimentary Petrology, v. 49,p. 541–552.

Stein, M., 2001, The sedimentary and geochemical recordof Neogene–Quaternary water bodies in the Dead Seabasin—Inferences for the regional paleoclimate his-tory: Journal of Paleolimnology, v. 21, p. 271–282.

Stein, M., Starinsky, A., Katz, A., Goldstein, S.L., Machlus,M., and Schramm, A., 1997, Strontium isotopic,chemical, and sedimentological evidence for the evo-lution of Lake Lisan and the Dead Sea: Geochimicaet Cosmochimica Acta, v. 61, p. 3975–3992.

Street-Perrott, F.A., and Harrison, S.P., 1985, Lake-level andclimate reconstruction, in Hecht, A.D., ed., Paleocli-mate analysis and modeling: New York, Wiley,p. 291–340.

Underhill, H.W., 1967, Dead Sea levels and the P.E.F. mark:Palestine Exploration Quarterly, p. 45–53.

Wohl, E.E., and Enzel, Y., 1995, Data for paleohydrology,in Gregory, K.J., et al., eds., Global continental pa-leohydrology: Chichester, England, John Wiley andSons, p. 23–59.

Yechieli, Y., 1993, The effects of water level changes inclosed lakes (Dead Sea) on the surrounding ground-water and country rocks [Ph.D. thesis]: Rehovot, Is-rael, Weizman Institute.

Yechieli, Y., Magaritz, M., Levy, Y., Weber, U., Kafri, U.,Woelfli, W., and Bonani, G., 1993, Late Quaternarygeological history of the Dead Sea area, Israel: Qua-ternary Research, v. 39, p. 59–67.

MANUSCRIPT RECEIVED BY THE SOCIETY 26 OCTOBER 2002REVISED MANUSCRIPT RECEIVED 17 JULY 2003MANUSCRIPT ACCEPTED 23 AUGUST 2003

Printed in the USA

on August 17, 2011gsabulletin.gsapubs.orgDownloaded from