Quaternary Science Reviews 209 (2019) 114e128
Contents lists available at ScienceDirect
Quaternary Science Reviews journal homepage: www.elsevier.com/locate/quascirev
Sedimentary, geochemical and hydrological history of Lake Kinneret during the past 28,000 years Lilach Lev a, b, Mordechai Stein a, c, Emi Ito d, Noa Fruchter a, c, Zvi Ben-Avraham b, Ahuva Almogi-Labin a, * a
Geological Survey of Israel, 30 Malkhe Israel St, Jerusalem, 9550161, Israel Department of Geophysical and Planetary Sciences, Tel Aviv University, P.O. Box 39040, Tel Aviv, 6997801, Israel Institute of Earth Sciences, The Hebrew University, Givat Ram, Jerusalem, Israel d Department of Earth Sciences and the Limnological Research Center, University of Minnesota, 310 Pillsbury Drive SE, Minneapolis, MN, 55455, USA b c
a r t i c l e i n f o
a b s t r a c t
Article history: Received 1 August 2018 Received in revised form 5 February 2019 Accepted 14 February 2019
The sedimentary, geochemical and hydrological history of Lake Kinneret (Sea of Galilee) is reconstructed for the past ~28 kyrs, based on three sedimentary cores drilled at the lake and a trench dug at the shore of the prehistorical OhaloeII site. During the past 28 kyrs either laminated or massive fine-grained sediments were deposited in the lake comprising primary calcites and fine-grain detritus. Sr/Ca, Mg/ Ca, 87Sr/86Sr ratios and d18O values of live and fossil ostracod shells (Cyprideis torosa) and primary calcites together with XRD, grain-size, and carbonate content analyses indicate contribution of the following types of waters to the lake: (1) Jordan River; (2) Regional runoff; and (3) Ca-chloride brines (currently comprises the Tiberias Spa brine). During the last glacial period (~28-24 ka) the lake rose to its highest stand of ~170 m below sea level (bsl), expanding over the Kinnarot Basin and converging with the southern hypersaline Lake Lisan. At that time, waters were mainly supplied to the lake by the Jordan River and regional runoff with enhanced contribution of the Dead Sea Ca-chloride brine. Primary calcites were precipitated from the lake's solution forming sequences of laminated sediments on the lake's floor. At ~24-22 ka (coinciding with Heinrich event H2 at the North Atlantic) the lake retreated below the modern level (of ~214 m bsl), depositing mainly flood-related sediments at its margins. The lake slightly rose during the Younger Dryas and subsequently declined towards the modern level with decreasing contributions of the Jordan River waters, brines and regional runoff, reflecting a continuous aridification of the region during the Holocene. © 2019 Elsevier Ltd. All rights reserved.
Keywords: Late Quaternary East Mediterranean Levant Lake Kinneret Paleo-limnology Geochemistry Sr and O isotopes Ostracods
1. Introduction Lake Kinneret (the “Sea of Galilee”) is a freshwater lake that fills the northern part of the Kinnarot Basin, one of the prominent morpho-tectonic depressions that were developed along the Dead Sea Transform (DST), e.g., Hula and the Dead Sea Basin (Fig. 1, BenAvraham et al., 1990). During the Quaternary these tectonic basins were occupied by freshwater to hypersaline lakes that derived their solutions from the Jordan River, regional runoff and springs discharging ancient brines (Stein, 2014a,b). Most of the paleohydrological and paleolimnological research so far has focused on the hypersaline last Glacial Lake Lisan and the Holocene Dead Sea that
* Corresponding author. E-mail address:
[email protected] (A. Almogi-Labin). https://doi.org/10.1016/j.quascirev.2019.02.015 0277-3791/© 2019 Elsevier Ltd. All rights reserved.
occupied during the Quaternary the Dead Sea Basin (e.g., Stein, 2001, 2014a,b and references there). The information on the hydrological-limnological history of the freshwater Lake Kinneret is more limited (e.g. Stiller, 1977; Dubowski et al., 2003; Hazan et al., 2005; Lev et al., 2014; Stein, 2014b). Lake Kinneret is mostly fed by freshwater draining Mt. Hermon, Upper Galilee Mountains and Golan Height via the Jordan and Yarmouk rivers. Mount Hermon comprises the largest and most important freshwater reservoir of the Levant, currently supplying water to Lebanon, Syria, Jordan, Israel and the Palestinian Authority. Reconstruction of the hydroclimate history of Mt. Hermon and its vicinity during the late Quaternary is crucial for models describing the hydro-climate regime of the Levant, an area that has recently suffered from prolonged periods of droughts (Cook et al., 2016). The hydro-climate history of Mt. Hermon is necessary for any study that incorporates past hydrological patterns with models of global
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
115
Fig. 1. (a) Location map. The Hula and Lake Kinneret (LK) are fresh water bodies located in the northern part of the Jordan Valley (along the Dead Sea Transform), and the Dead Sea is a hypersaline water body filling currently the Dead Sea Basin (DSB). The contour line describes the highest water level, when LK merged with the last glacial Lake Lisan that filled the DSB. The Yarmouk and the Malih sills (200 and 250 m below sea level (bsl), respectively) are also shown. (b) Cross section along the Jordan Valley showing the Yarmouk and Malih sills. (c) Sampling sites around Lake Kinneret. Black circles mark water sources and modern sediments sampling sites (described in Lev et al., 2014), black rectangles mark the trenching and coring sampling sites KIN2, SOG2 and SOG3. Bathymetry is after Ben-Avraham et al. (1990).
warming effects that predict increasing aridity in the subtropical dry regions of the Near East (e.g. Held and Soden, 2006). Here, we focus on the reconstruction of the hydrological history of Lake Kinneret, which is the largest freshwater body that receives waters from Mt. Hermon. Our reconstruction is based on mineralogical, chemical and isotope data (e.g. Sr/Ca, Mg/Ca, d18O values and 87 Sr/86Sr ratios) from ostracod shells and primary calcites that were deposited in the lake. The new data obtained in this study were used to constrain the hydro-climate history of Lake Kinneret and its water suppliers: Mt. Hermon, regional runoff and brines during the past 28 kyrs focusing mainly on the large-scale differences in lake response to glacial vs interglacial (Holocene) limnological conditions. The data are used to discuss the patterns of the Levant climate during this time interval. 2. Geological, hydrological and limnological setting 2.1. Geological setting Lake Kinneret (Sea of Galilee) is ~20 km long, ~12 km wide and its maximum depth is ~46 m, based on lake level of ~210 m below sea level (m bsl). The modern lake is located at the northern part of the Kinnarot Basin (Fig. 1), which is one of the morpho-tectonic depressions along the Dead Sea Transform (DST), with a ~4.5 km
deep sedimentary record comprising volcanics, clastic material, carbonates and salts that were accumulated during the past ~ 10 Ma (e.g. the sedimentary record recovered by the Zemah-1 drill-hole, Marcus and Slager, 1985). The modern Lake Kinneret is surrounded by Pliocene-Pleistocene basalts (in the Galilee and Golan Height), Neogene sandstones and Cretaceous and Eocene carbonate rocks (Heimann and Braun, 2000). The sedimentary sequence that filled the Kinnarot Basin during the late Pleistocene and the Holocene comprises the Kinneret Formation (Hazan et al., 2005). The formation is exposed along the southern shores of the lake displaying sequences of laminated or massive lacustrine sediments (comprising mainly primary calcite and fine detritus material) intercalated with clastic sequences of sand and pebbles (Hazan, 2003). In the center of the lake the sediments consist mostly of primary calcite and fine-detritus particles of dust origin (Stiller and Kaufman, 1985; Ganor et al., 2000; Fruchter et al., 2017). Based on the composition of sediment cores drilled near the center of the lake, Stiller and Kaufman (1985) estimated the content of primary carbonate as 70e85% of the total carbonate. Sedimentation rates in the center of the lake were estimated to be 2e7 mm/yr (Stiller, 1974). Here, we focused mainly on the sequences comprising primary calcites that were deposited from the lake's water, and are recovered from the drilled cores. The primary calcites and ostracod shells (when available) provide information on the composition of
116
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
the lake's water during the past 28,000 years. 2.2. Hydrological and limnological setting Lake Kinneret is currently a monomictic lake with a winter water temperature of ~15 C. In spring and summer, the depth of the epilimnion is 15e25 m with a maximum temperature of 29 C, and the temperature of the hypolimnion is ~15 C (Serruya, 1975). The pH of the water fluctuates from 8.0 to 9.1 and the major dissolved ions are Naþ, Caþ2, Mgþ2, Kþ and Cl (e.g., Katz and Nishri, 2013). During the last Glacial and the Holocene, Lake Kinneret coexisted with the hypersaline Lake Lisan and the Dead Sea, filling the tectonic basin of Kinnarot and the Dead Sea. Lake levels were established independently for both lakes: (1) The last glacial Lake Lisan and the Holocene Dead Sea by Bartov et al. (2002, 2003) and Torfstein et al. (2013); (2) Lake Kinneret by Hazan et al. (2004, 2005). These reconstructions show that during most of the past 70 kyrs (the last glacial period) the (paleo) Dead Sea and (paleo) Lake Kinneret were separated waterbodies and their waters did not overflow the topographic “sills” of Wadi Malih and Yarmouk River at ~ 250 and ~200 m bsl, respectively. The waters crossed the sills and the lakes converged only between ~28 and 24 ka, when lake levels reached the highest stand of ~170 m bsl (Fig. 1). During the high-stand period, the Dead Sea brine possibly reached the area of Lake Kinneret and contributed to its composition (a subject discussed below). At ~23.8 ka the surface waters in Lake Lisan and Lake Kinneret dropped to ~214 m bsl (Lev et al., 2014). The lake level decline coincided with Heinrich Event 2 (H2) in the North Atlantic (Bartov et al., 2003; Lev et al., 2014; Stein, 2014b). Bartov et al. (2003) related the lake drops during the Heinrich events to the cooling of the East Mediterranean water that caused a “shut-down” of the cyclonic “engine” producing the Levant rains. Water level in Lake Kinneret remained low possibly until ~22 ka and then rose and the lake flooded its margins. Later, during most of the post-glacial period and the Holocene Lake Kinneret level fluctuated around ~214 m bsl. 2.3. Modern hydrology of the Lake Kinneret Most of the freshwaters that currently enter Lake Kinneret are contributed by the Jordan River that receives its water mainly from Mt. Hermon (Fig. 1). Additional freshwater sources are the Yarmouk River that also drains Mt. Hermon and short rivers (Nahals) draining the basaltic Golan Height, the carbonate terrains of Galilee Mountains and their surface cover (e.g. mountain soils). Saline springs discharging at the western shore of the lake (or nearby underwater sites) constitute another important source of water to the lake (i.e., Tabgha, Fuliya and Tiberia Spa, Goldstein, 2004; KleinBenDavid et al., 2005). It was suggested that the sources of the saline water are brines whose history can be traced back to the late Neogene when marine (Mediterranean Sea) water penetrated the Jordan Valley via the Yizre'el Valley and possibly reached the Kinnarot and Dead Sea Basins (e.g. the Sedom lagoon, Starinsky, 1974; Stein, 2014a, b; Stein et al., 2000; Zak, 1967). The Ca-chloride brines and freshwater from different sources controlled the composition of the modern and ancient lakes, and their relative contributions to the lake reflect the regional hydroclimate conditions in the catchment area. 2.4. Biogenic and inorganic carbonate tracers The lacustrine sediments that were deposited in Lake Kinneret comprise mainly primary calcites and fine detritus material and contain fossil shells of fauna that inhabited the lake such as
Melanopsis (a fresh water snail), ostracods and occasionally foraminifera (e.g. Tchernov, 1975; Hazan et al., 2005; Lev, 2006; Lev et al., 2007; Mischke et al., 2010, 2014; Lev et al., 2014; Fruchter et al., 2017). The primary (authigenic) calcite, precipitates from the lake solution, from the epilimnion during spring to early summer phytoplankton blooms (due to a shift in pH conditions of lake water caused by the phytoplankton) and accumulates on the lake floor (Katz and Nishri, 2013). Ostracods, small (0.3e3 mm) bivalve crustaceans, live in most types of aquatic environments (Athersuch et al., 1989). Among them, Cyprideis torosa (Jones) is a common inhabitant of modern Lake Kinneret (Mischke et al., 2014). The modern lake is stratified during most of the year (Serruya, 1975) and ostracods, including C. torosa are absent in the modern lake below ~15 m water depth, where the environment is anoxic during most of the year (Lev et al., 2014). Sr/Ca, Mg/Ca, d18O and 87 Sr/86Sr in C. torosa shells were used by Lev et al. (2014) as a major source for tracing fluctuations in the paleo-hydrology and paleolimnology of the lake during Heinrich event H2. According to Lev et al. (2014) the d18O values of the ambient Lake water lie between 0.12 and 0.56‰ (VSMOW) and the d18O in living C. torosa shells from the same localities lie between 0.35‰ and þ0.07‰ (VPDB) indicating that the isotope values of ostracod shells (VPDB) were slightly (~0.5‰) lighter than the lake water. In an extensive empirical calibration study, Marco-Barba et al. (2012) showed that the d18O values of C. torosa accurately reflect the isotopic composition of lake waters. 3. Materials and methods 3.1. Sampling Late Holocene and Pleistocene sediments - The reconstruction of the limnological and sedimentological history of Lake Kinneret during the past ~28 kyrs is based on new data obtained from sediment cores: KIN2, SOG2 and SOG3 and the deepest part of the trench that was dug at the Ohalo-II archeological site (Ohalo-II (OH) trench, reported by Lev et al. (2014)). The elevation at the top of each core is 213, 224 and 229 m bsl respectively. Further details on the location of the cores and the trench are marked in Fig. 1 and general details are listed in Supplementary Table 1. KIN2 borehole and OH trench were drilled and dug, respectively in spring 1999 by a joint group from GFZ-Potsdam and the Hebrew University, Jerusalem at the Ohalo-II archeological site. KIN2 sediments were sampled as 3-cm thick slices at least every 20 cm. Cores SOG2 and SOG3 were drilled by Tahal (stands for “Water planning for Israel”) at 1985 (see details in Ehrlich, 1985) and are the only sediment sequences available from the margins of the lake that have higher concentrations of fossil ostracods. The two cores were sampled in 10-cm thick slices every ~10 cm” 3.2. Analytical methods XRD analysis was done using a PW 1820 diffraction spectrometer (Phillips) at the Geological Survey of Israel (GSI). Cu Ka X-radiation (l ¼ 1.5418 Å) was supplied by a copper-target tube. Scan range (2 Theta) was 2e42 and scan speed: 2 /minute. The mineral fractions in each sample were semi-quantitatively calculated. The intensity of the diffraction signal is plotted against the diffraction angle allowing the determination of the relative composition of the sample. Relative abundance of minerals is reported as: <5%, 5e25%, 25e50% and >50%. Grain-size analysis was done at the GSI on the bulk sediments by Malvern Mastersizer MS-2000 for grain-size < 2 mm. To avoid clay aggregation dry sediments samples were mixed with 25 mldistilled water with Calgon following the procedure described in
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
Crouvi et al. (2008). Carbonate content (wt% CaCO3) was determined by two methods: CO2 was analyzed at the GSI lab by loss on ignition and the CaCO3 was calculated. In addition, CaO content was determined by PerkinElmer Optima 3300 ICP-AES at the GSI. Both methods yielded consistent results. Chemical and isotope analyses of ostracods and primary calcites: The dominant ostracod species C. torosa was selected for analyses of Ca, Mg and Sr concentrations as well as d18O values and 87Sr/86Sr isotope ratios (similar treatment was applied to the samples of primary calcites). These properties provide information on the limnological conditions in the ancient lake, given that the incorporation of trace elements into the ostracod shell and the relationship between the ionic composition and the lake salinity are known (Almogi-Labin et al., 2004, Wansard, 1996a,b, Wansard et al., 1998, Marco-Barba et al., 2012). The ostracods were collected from ~30 g bulk sediment samples. The samples were washed-through a 63 mm sieve and the retained portion was air dried. The dried samples were sieved again (>250 mm) and valves of the ostracod C. torosa were identified and separated under a binocular microscope. For ostracod elemental and isotopic analysis, only adult or A-1 specimens were used. When possible, 10 (~0.3 mg) or more valves were used to measure the chemistry and 87Sr/86Sr isotope ratio. At least 4 shells were used for Sr, Mg and Ca measurements when ostracods were rare. Twentythree ostracod samples from SOG2 borehole and 11 samples from SOG3 borehole were measured for their Sr, Mg and Ca. Four of these measurements are replicates from 3 samples. The 87Sr/86Sr isotope ratios were analyzed in 22 ostracod samples from SOG2 and SOG3 while d18O values were studied in 12 samples from SOG2. Each ostracod shell or powder of primary calcite separated from the core material were washed in distilled water, cleaned ultrasonically, dissolved in 3.5 N HNO3 and analyzed for Ca, Sr concentrations by ICP-AES using Optima 3300 in the GSI. Each run included repeated determinations of the international standards SO-3. Estimated error is less than 5%. For d18O analyses, about 0.6 mg of ostracods (20 shells) or primary calcite were selected. Samples were dissolved at 25 C for 24 h in 100% phosphoric acid before measuring. Oxygen isotope ratios were measured using a 14 Gas Bench system attached to a Delta Plus mass spectrometer at the GSI. d18O values were calibrated against the international standard NBS-19 and are reported in permil (‰) relative to VPDB standard. Analytical reproducibility of duplicates is better than 0.1‰. The preparatory work for the Sr isotopic analysis included dissolution of the cleaned ostracod shells or powder of primary calcite in 1.5 ml of 3.5 N distilled HNO3 and extraction of Sr using ion exchange column filled with Sr-Spec 50e100 mesh resin (e.g. Stein et al., 1997). Strontium isotope ratios were measured with Multi-Collector Inductively Coupled Plasma Mass Spectrometer (MC-ICP-MS) NU instruments at the GSI. During the course of the study the NBS987 standard yielded 87Sr/86Sr ¼ 0.71029 ± 0.00004 (n ¼ 9, 2s). The samples were corrected to the long run laboratory value of the NBS987 standard ¼ 0.71024. Bulk carbonate chemistry - Fifty-seven samples of bulk carbonate from KIN2 were analyzed for their Ca, Sr and Mg concentrations and 20 samples were analyzed for d18O values assuming that the bulk carbonate in this core comprises primary calcite as evident from SEM analyses. About 60 mg of bulk sediment was dissolved in acetic acid, centrifuged to separate the carbonate fraction, dried on a plate, dissolved with concentrated HCl, dried again, dissolved with concentrated HNO3, dried and dissolved with 3.5N HNO3. Measurement was done by Optima 3300 ICP-AES at the GSI. Sc, Ru and Rh are added as internal standard. Precision is less than 5%.
3.3.
14
117
C dating
Eight samples of charcoal, 6 samples of twigs, and 5 samples of ostracod shells were dated at the radiocarbon lab at the Australian Nuclear Science and Technology Organization (ANSTO), Australia. The radiocarbon ages of the ostracod shells are corrected for reservoir age based on cases where ostracods were found together with wood or charcoal. The specific details of the reservoir age correction are given below (in the Results section). The radiocarbon ages of charcoal and wood samples and reservoir-age corrected ostracod samples were calibrated to calendar ages using OxCal 4.1. Depth-age model for each of the cores were drawn using the Bacon software for Bayesian age modeling in R (Blaauw and Christen, 2011). The age model output from Bacon is based on the probability density function for each calibrated radiocarbon age and a prior estimate of the mean sedimentation rate throughout the core (10 years/cm). 4. Results 4.1. Lithology of the cores 4.1.1. Mineralogy and grain sizes of the sediments Mineralogical data of the sediments comprising the cores are listed in Supplementary Tables 2e7. The sediments contain 60e90% calcite, 5e25% quartz and 5e25% clays. For the description of the core sediments we use the term “marls”. We distinguish between sequences comprising “laminated marls” where laminae thickness is about 0.5e1 mm, and those of “massive marls”, where layers thickness can reach a few cm. Grain size mode typically varies from 4 to 20 mm (Supplementary Fig. 4), showing distinct values for each one of the cores. 4.1.2. KIN2 core The core comprises massive (MM) and laminated marls (LM) (Fig. 2, Supplementary Fig. 1). Laminated marls appear between ~216 and ~222 m bsl. They show unimodal grain size of 4e8 mm (Supplementary Fig. 4, Supplementary Tables 2 and 5) and are composed mainly of primary calcite with some quartz and clays. Massive marls appear at the top of the core (~213e216 m bsl) and at its bottom (~222 m bsl) above pebbly sediment. Warm saline waters discharged from between the pebbles during drilling (Hazan et al., 2005). The massive marls show coarser grain size (~20e30 mm) and are composed of quartz, calcite and clays. Dolomite appears only in minor amounts. 4.1.3. SOG2 core SOG2 comprises mostly of massive marls with two short intervals of laminated marls at ~227 and ~229 m bsl (Fig. 2, Supplementary Fig. 2, Supplementary Tables 3 and 6). The carbonate content from 234.0 to 228.2 m bsl varies between 30 and 50% and is higher at the top of the core (>60%). Quartz and some clays make up 5e25% with minor amounts of gypsum at the coretop. Most of the samples show unimodal grain size distribution of ~10e20 mm (Supplementary Fig. 4). A small number of mollusks appear at the coretop. In addition, most of the sediments contain juvenile stages of C. torosa shells and adults occur only in 16 samples. The coretop is rich in freshwater diatoms with somewhat higher abundance of C. torosa. Brackish benthic diatom species occur below 226.05 m bsl, alternating with layers barren of diatoms (Ehrlich, 1985). 4.1.4. SOG3 core SOG3 comprises mostly massive marls with three short intervals
118
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
ages and use the recently published chronology by Miebach et al. (2017). The ages are reported as calibrated ages before present (1950) and are termed throughout the manuscript as “ka” (e.g. 22.5 ka). The uncertainties in the age-depth models are estimated from the linear regressions and are typically less than 15% (on the slope). 4.2.1. KIN2 core The age-depth model of KIN2 is based on five radiocarbon ages from Miebach et al. (2017), a single radiocarbon age from Hazan et al. (2005), and 6 new radiocarbon ages from this study (Fig. 3). The core covers the time interval of the last glacial period between 28 ka (bottom of the core at 222 m bsl) and 22.5 ka at the core-top (at 213.05 m bsl). This interval corresponds to Marine Isotope Stage (MIS) 2. The chronology of the top of KIN2 is consistent with that of the adjacent Ohalo trench (Lev et al., 2014), corroborating the precision of the radiocarbon ages in both sites. The average accumulation rate reaches ~150 cm/ka, higher than the rates estimated for SOG2 and SOG3. 4.2.2. SOG2 core The age-depth model of SOG2 (Fig. 3) is based on radiocarbon ages of four samples of organic debris: twigs and wood, charcoal and two ostracod shell. Samples of organic debris and ostracods from the depth of 226.1 m bsl yielded radiocarbon ages of 10.5 and 15.1 ka, respectively implying a reservoir age of ca. 5 ka for the ostracods. We use 5 ka reservoir age to correct the radiocarbon age of ostracod sample SOG2 389 from the depth of 227.9 m bsl and 800 y for ostracod sample SOG2 910 that was collected at the bottom of the core at depth of 233.1 m bsl (Table 2). This correction is based on the reservoir age calculated by Lev et al. (2007) for Melanopsis shells from the last glacial Lake Kinneret. These reservoir age corrections are consistent with different lake water compositions during the late glacial and post glacial times (see section 5.3 below). Fig. 2. Comparison of the chronology of the Ohalo (OH) trench, KIN2, SOG2 and SOG3 cores to water level curves for Lake Lisan, Lake Kinneret, and the Dead Sea during the past 30 kyrs. The curves are based on the works of Bartov et al. (2002, 2003), Bookman et al. (2004), Hazan et al. (2005), Migowski et al. (2006), Stein et al. (2010) and Torfstein et al. (2013). Four episodes of laminated sediments (LM1-4) were identified in the sedimentary sequence alternating with massive marls events during the last 28 kyr. YD - Younger Dryas, H1- Heinrich event 1, H2 - Heinrich event 2. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
of laminated marls occurring at ~231, ~234 and 235 m bsl (Fig. 2, Supplementary Fig. 3, Supplementary Tables 4 and 7). The carbonate content varies between 60 and 80% in samples below 236 m and between 10 and 50% above this depth. Quartz and some clay make up 25e50% in the entire section. A few samples show minor amounts of dolomite. The samples show grain-size distribution with modes varying between 4 and 20 mm (Supplementary Fig. 4). Small numbers of juvenile shells of C. torosa were present in most sediment samples, and adult shells were found in only 11 samples. From the bottom of the core up to 235.8 m bsl no ostracods shells were present. Diatoms were absent in large segments of the core except in its middle section from 233.65 to 235.65 m bsl where mostly brackish water species were found (Ehrlich, 1985). 4.2. Chronology of the cores The age-depth models of the cores (Fig. 3) are based on radiocarbon ages of wood and charcoal and reservoir age-corrected radiocarbon ages of ostracod shells (Table 2). In core KIN2 we combine the new data with previously determined radiocarbon
4.2.3. SOG3 core The age-depth model of SOG3 (Fig. 3) is based on five radiocarbon ages of organic debris and charcoal. The age-depth model indicates that the sediments comprising SOG 3 were deposited between ~23 and ~1.5 ka, interval that encompasses the time interval represented by KIN2 and SOG 2, and extends to the late Holocene. 4.2.4. Combined chronology The (calibrated) radiocarbon ages of the three cores combined provide an almost continuous chronology of Lake Kinneret for the past ~28 kyrs. This “combined chronology” is illustrated against the lake level curve in Fig. 2. Most of the lake level data are from Hazan et al. (2005) who identified shoreline deposits at the modern lake's margins. Here, we marked additional levels above the top of the drilled cores. The latest part of the Holocene (<2 ka) is represented in the uppermost part of core SOG3 although it was not dated. This time interval is present in dated cores retrieved from nearby locations (e.g. Stiller et al., 1988; Dubowski et al., 2003). 4.3. Oxygen isotopes Oxygen isotopes were analyzed in ostracod shells from core SOG2 and in primary calcites from KIN2 and from the lower part of Ohalo-II trench, OH7. The ostracod's and primary calcite d18O data are reported in VPDB (Supplementary Tables 8e11 and Supplementary Figs. 1e2). In the Discussion section below we explain how these d18Ocalcites were converted to d18Owater (see also Fig. 4).
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
4.4. Sr/Ca, Mg/Ca and ostracod shells
119 87
Sr/86Sr ratios in primary calcites and
KIN2 core (and OH7): Sr/Ca, Mg/Ca and 87Sr/86Sr ratios were measured in the primary calcite and the ratios range from 0.0012 to 0.0022; 0.023 to 0.088 and 0.70771 to 0.70796, respectively (Fig. 4, Supplementary Fig. 1, Supplementary Tables 8e9). SOG2 core: Sr/Ca, Mg/Ca and 87Sr/86Sr ratios in ostracod shells range from 0.0016 to 0.0039, 0.0069 to 0.01098 and 0.70771 to 0.7080, respectively (Supplementary Fig. 2, Supplementary Table 10). 87Sr/86Sr ratios in the ostracods shells show the lowest values at the top of the core and the highest values between 227 and 230 m bsl (at ~20-16 ka). SOG3 core: Sr/Ca, Mg/Ca and 87Sr/86Sr ratios in ostracod shells range from 0.004 to 0.006, 0.007 to 0.013 and 0.70765 to 0.70773, respectively (Supplementary Fig. 3, Supplementary Table 11). 5. Discussion 5.1. General In the following sections, we reconstruct the limnological, sedimentological and hydrological history of Lake Kinneret during the past 28 kyrs combining the data from the three cores and the Ohalo trench. The lithological data are used to determine the sedimentary facies and the environment of deposition (section 5.2). Combined with the available (Hazan et al., 2005) lake level curve the sedimentary facies data is used to reconstruct the limnological history of the lake (section 5.3). Then, oxygen isotopes and 87Sr/86Sr isotope ratios combined with Sr/Ca ratios in ostracods shells and primary calcites are used to reconstruct the hydrological regime, and quantify the sources of fresh and saline waters to the lake over the past 28 kyrs (sections 5.4 - 5.6). 5.2. Sedimentary facies The cores comprise mainly two sedimentary facies: massive marls (MM) and laminated marls (LM) (Fig. 2, Table 1 and Supplementary Fig. 5). The massive marls are composed mainly of quartz, calcite and clays that resemble the surface cover (e.g. mountain soils) of the Galilee Mountains (e.g. Sandler et al., 2015). The Galilee soils, in their turn evolved from desert dust material that was blown from the north Sahara deserts, reached the Levant region mainly during winter storms, settled on the land surface and underwent pedogenetic processes (e.g. Sandler et al., 2015; Palchan et al., 2018 and references there). The mineral grains comprising the surface cover were washed to the lake by seasonal floods (e.g., Palchan et al., 2018). The calcites are mainly primary mineral phases precipitated from the lake waters as is observed in the modern lake (see pictures of calcite crystals that are observed by SEM in Katz and Nishri, 2013 and Fruchter et al., 2017). This assertion is collaborated by the 87Sr/86Sr isotope ratios of the calcites (Fig. 4) that are consistent with precipitation from the lake's water.
Fig. 3. Age-depth diagrams for the KIN2 (A), SOG2 (B) and SOG3 (C) boreholes calculated by the Bacon software (Blaauw and Christen, 2011). The diagrams show 14C data of organic debris, Melanopsis and ostracod shells (Table 2). The measured
radiocarbon ages of the Melanopsis and the ostracods shells are corrected for reservoir ages (see text for further discussion). The Bayesian age-depth model output from Bacon (Blaauw and Christen, 2011), includes calibrated 14C dates (transparent blue), age-depth model (darker greys indicate more likely calendar ages; grey stippled lines show 95% confidence intervals; red curve shows single ‘best’ model based on the weighted mean age for each depth) and sedimentation rates (mm/yr) along the depth profile based on weighted mean age for each depth. Inset figures show the prior (lines) and posterior densities (area fills) for the mean accumulation rate (b; sedimentation rate) and memory (c; autocorrelation strength at 1 cm intervals), the two prior estimates in the age model. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
120
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
Table 1 Sedimentary facies. Sedimentary Facies
Mineralogy
Fine sediments and carbonate content
Biogenic constituents
Depositional environment
massive marls 1 (MM1)
Quartz (5e50%), clay (5 e50%), calcite (25e50%) Quartz (5e50%), clay (5 e50%), calcite (25e50%)
<90% fine sediments, > 50% carbonates
Yes
Marginal environment, dominated by floods
KIN2 top (213.0e214.7 m bsl)
>90% fine sediments, < 50% carbonates
No
Water depth >15 m but no laminations
SOG2 (~230e233 m bsl)
massive marls 2 (MM2)
massive marls 3 (MM3)
Quartz (5e50%), clay (5 e50%), calcite (5e75%)
<90% fine sediments, >50% carbonates
No
Water depth >15 m but no laminations
SOG3 bottom
laminated marls (LM)
Quartz (5e50%), clay (5 e50%), calcite (5e50%)
>90% fine sediments, <50% carbonates
No
High lake level environment
KIN2 SOG3
No
River/stream
KIN2 bottom
pebbles
Table 2 Radiocarbon ages. Sample
Type
Elevation (m bsl)
Age
Calendar age (ka BP)
Reference
22.5 23.1 23.6 23.5 23.6 23.8 23.9 23.9 23.9 24.8 27.3 28 12.3 11.2e12.5 16.8 21.1 20.4 20.4 20.2 24 2.2 9.2 10.5 11.1 23.1
Miebach et al., 2017 This study This study This study Miebach et al., 2017 Miebach et al., 2017 This study Hazan et al. (2005) This study This study Miebach et al., 2017 Miebach et al., 2017 This study This study This study This study This study This study This study This study This study This study This study This study This study
(14C ka BP) AMS 01 KIN2 2251 KIN2 2257 KIN2 2257 AMS 02-1 AMS 02-2 KIN2 2258 KIN2 2258 KIN2 2264 KIN2 2267 AMS 09-2 AMS 13-2 SOG2 209 SOG2 209 SOG2 389 SOG2 581 SOG2 676 SOG2 676 SOG2 753 SOG2 910 SOG3 184 SOG3 282 SOG3 384 SOG3 463 SOG3 646
wood & charcoal charcoal charcoal ostracods charcoal charcoal ostracods Melanopsis charcoal charcoal wood a u.i. snail wood ostracods ostracods wood wood wood wood ostracods charcoal charcoal charcoal charcoal wood
213.05 213.6 214.4 214.4 214.5 214.5 214.6 214.6 215.6 215.9 219.6 221.2 226.1 226.1 227.9 229.8 230.8 230.8 231.5 233.1 230.8 231.8 232.8 233.6 235.5
18.6 ± 0.1 19.4 ± 0.1 19.6 ± 0.1 19.6 ± 0.2 19.6 ± 0.1 19.7 ± 0.1 19.9 ± 0.4 20.7 ± 0.2 20.0 ± 0.1 20.8 ± 0.1 22.9 ± 0.1 25.4 ± 0.1 10.4 ± 0.1 15.1 ± 0.2 18.8 ± 0.2 17.8 ± 0.1 17.1 ± 0.1 17.2 ± 0.1 16.9 ± 0.1 20.0 ± 0.2 2.2 ± 0.1 8.2 ± 0.1 9.3 ± 0.1 9.7 ± 0.1 19.3 ± 0.1
Ages were calibrated by OXCAL 4.1. Freshwater shells were corrected for reservoir ages. See text for details. a u.i. snail e unidentified freshwater snail.
The massive marls were further divided into three different facies according to the grain size and the presence/absence of ostracods (Table 1). Facies MM1 is characterized by higher content of coarse grains (>63 mm) and the presence of ostracod shells indicating deposition in shallow water environment of less than 15 m, based on modern minimum depth of summer anoxia below which ostracods are not found today (Mischke et al., unpublished data). Facies MM2 is characterized by small grain size mode (10e20 mm, Supplementary Fig. 4) and water exceeding 15 m. Facies MM3 resembles MM1 but lacks biogenic constituents. The laminated marls (LM1-4, Fig. 2, Supplementary Figs. 1-5) are generally characterized by fine-grain sediments (4e8 mm) comprising primary calcite (up to 90%), detrital quartz, detrital calcite and clays. The laminated marls lack ostracods. They were deposited during high stand periods when the lake was stratified,
under anoxic conditions with no bioturbation activity. 5.3. Sedimentation and paleo-limnology of Lake Kinneret during the past 28 kyrs The combined age-depth models of the three cores and that of the Ohalo-II trench (Lev et al., 2014, and new data) cover the past 28 kyrs of the sedimentary and limnological history of the lake (Figs. 2 and 3). The latest Holocene is recorded at the top of SOG3 (Figs. 2 and 3, and see Supplementary Fig. 3). Late Holocene age sediments (last ~4000 years) were recovered from other cores taken near the SOG3 site (Stiller et al., 1988; Dubowski et al., 2003). The three cores show intervals of laminated marls alternating with massive marls (Fig. 2). The changes from sequences comprising laminated marls (mainly primary calcites) to sequences
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
Fig. 4. (A) 87Sr/86Sr, (B) Sr/Cawater, (C) Mg/Cawater and (D) d18OwaterVSMOW from ostracods and primary calcites in cores KIN2, SOG2, SOG3 and from the Ohalo trench. See section 5.5 for calculation of the d18Owater data in 4D. Errors on the d18Owater values are estimated as ±15% according to the errors on the clumped isotopes temperatures of Zaarur et al. (2016). The Melanopsis data are taken from Zaarur et al. (2016). For references of present-day values of Lake Kinneret, Tiberias Spa and Jordan River see Table 3. The data span the past 28 kyr. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
comprising massive marls (mostly fine detritus that comprises remobilized surface cover material) (see Supplementary Fig. 5) reflect a change in the hydrological regime. While the former indicate conditions of high-lake stands and positive net freshwater input (enriched in bi-carbonate) supporting the deposition of primary calcite, the latter likely represents shallow-water sediments affected by flood sediments that were deposited at all studied cores during low lake-stand periods. The oldest part of the record (~28-24 ka) is represented by core KIN2 and the bottom part of the Ohalo (OH) trench (Lev et al., 2014).
121
The sequences comprise laminated marls (marked in Fig. 2 as LM4) of mostly primary calcite and high sedimentation rate (~150 cm/ ka). During this time-interval the lake reached its highest stand of ~170 m bsl merging with Lake Lisan to the south (Hazan et al., 2005). The time interval between ~24e22.5 ka when the lake level dropped following the H2 event, is recorded as massive marls in all cores and in the OH trench (Fig. 2). During that period, the UpperPalaeolithic people of OhaloeII descended from their Galilee caveshelters to the exposed retreating shores of Lake Kinneret (Nadel et al., 2002). No sediment younger than 22.5 ka was recovered in KIN2. The laminated sequence LM3 begins at ~22.5 ka and terminates at ~19 ka in the deepest core SOG3. The top part of LM3 is also present in SOG2 and OH records (Fig. 2). The sedimentary record of the OH trench terminates at ~19 ka when the lake rose above the Ohalo-II site possibly forcing the Upper-Palaeolithic inhabitants to retreat back to the Galilee Mountains. Between ~19-17 ka massive marls (of MM3 type) were deposited at SOG2 and SOG3, so the lake level at this time was higher than ~230 m bsl. At ~12.5 ka, a short, laminated sequence (LM2) was deposited at SOG2 and SOG3, coinciding with the Younger Dryas (YD) cold event. Lake Kinneret rose to ~210 m bsl, contemporaneously with a significant rise at the Dead Sea (Stein et al., 2010). Between ~11-9 ka, a thick sequence of massive marls was deposited at SOG3. The age of deposition of this sequence coincides with the time of mobilization and accumulation of mountain soils in the Jordan and Beit-She'an Valley (the Fazael Formation, Stein, 2014a). This period also coincides with the interval of Sapropel S1 in the East Mediterranean when extensive floods affected the circum-Mediterranean region (Kallel et al., 1997; Almogi-Labin et al., 2009). Thus, it appears that extensive flooding and re-mobilization of soils and surface cover occurred all over the mountain backbone of Israel - the Judea, Samaria and Galilee Mountains bringing the detritus material to the Dead Sea, Jordan Valley and Lake Kinneret. The youngest interval of the laminated sequence (LM1) was deposited sometimes between 5-4 ka and could be correlated with the mid-Holocene lake level rise recorded at Tel Beit Yerah (above the Ohalo-II site) at ca. 5 ka BP (Hazan et al., 2005). LM1 is followed by massive marls indicating a lake level drop and deposition of detritus material in shallow-water environment at SOG3 (Fig. 2). Sedimentation in the southwestern margin of the lake appears to cease around ~19-18 ka as evidenced in the Ohalo trench (Lev et al., 2014) and in KIN2 core (this study). The reason for the absence of sediments is not entirely obvious. It could indicate either erosion after sediment accumulation or “starved basin” conditions where the supply of sediments was drastically decreased. At SOG2 site, located deeper in the southern part of the Kinneret Basin, sedimentation ceased after the Younger Dryas possibly indicating ongoing decrease in sediment supply to this part of the basin (Figs. 2e4). Today, there are no streams draining Galilee or Golan Heights that enter the southern part of the Kinneret Basin. Ohalo II ostracod assemblages indicate frequent fluctuations in salinity and identified 4 episodes of freshwater influx, suggesting that there may have been a stream entering the lake near the archeological site (Mischke et al., 2014). Increased salinity of the lake during low stands would not have supported the Upper Palaeolithic inhabitants occupying the Ohalo II site, so a presence of a freshwater stream nearby is perhaps to be expected. Integrating the new data with those of Hazan et al. (2005), that was mostly based on radiocarbon dating of paleo-shorelines at the exposed margins of the modern lake, allows us to refine the lake level curve (Fig. 2). The KIN2 and OH data combined indicate a high stand (~170 m bsl) and meromictic lake configuration at the peak of the last glacial MIS2 (28-24 ka BP). The lake level dropped to ~214 m bsl during H2 event (~23.8e22.5 ka) exposing the south-western
122
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
shores of the lake and allowing the Upper-Palaeolithic people to establish their settlement at the Ohalo site (Fig. 5). The lake rose again at ~19 ka, dropped at ~17 ka BP (during H1) and dropped again to ~226 m bsl between ~14-13 ka. The continuous sedimentation at SOG2 between ~22 and 13 ka implies that lake level did not fall below the depth of this core site at ~226 m bsl. This elevation is ~12 m below the low stand of the lake as marked by the Ohalo-II archeological site. These declines in Lake Kinneret levels coincide with significant lake level declines in Lake Lisan (Torfstein et al., 2013, and see Fig. 2). Thus, the two lakes appear to respond similarly to the changes in the regional hydroclimate regime. The YD event is recorded as deep water laminated sediment overlain by shallow water massive marls in SOG2 and SOG3 cores. This is the first record of the Younger Dryas time interval in the Jordan Valley.
5.4. Freshwater and brine contribution to Lake Kinneret 87
Sr/86Sr isotope ratios combined with Sr/Ca ratios in the ostracod C. torosa and in the primary calcites retrieved from the cores were used to characterize the composition of the water that filled Lake Kinneret during the past 28 kyrs, and the contribution of the different water sources to the lake. Cyprideis torosa lives on the surface of lake's bottom sediment so it represents the lake water. The ostracod incorporates Sr into its low-Mg calcite shell in proportion to Sr/Ca ratio of the water from which it takes the cations. Wansard (1996a, b), Wansard et al. (1998) and Marco-Barba et al. (2012) monitored in the laboratory or in natural settings the distribution of Sr/Ca between C. torosa shell and water. The partition coefficient (Kd) ratio of (Sr/Ca)C. torosa/(Sr/Ca)water was determined to be 0.57 ± 0.25 by Marco-Barba et al. (2012), and to be in the range of 0.63e0.83 by Wansard (1996a, b) and by
Wansard et al. (1998). We adopted the ratio of 0.7, i.e., all measured Sr/Ca ratios of C. torosa shells are divided by 0.7 to obtain the Sr/Ca ratio of the water in which the ostracod formed its shell. For the primary (authigenic) calcites we used Kd value of 0.21 (following Katz and Nishri, 2013). Water sources that currently feed the lake are: the Jordan River (at its entrance to the lake, including also rivers and runoff that come from the basaltic Golan Height, located to the east and northeast of the Lake), runoff from the carbonatic terrains and surface cover of the Galilee mountains, located to the west and north-west of the lake, and the saline brines and springs Fuliya, Tabgha and Tiberias Spa (Fig. 6A). During the last glacial when Lake Lisan converged with Lake Kinneret the latter could be affected also by the Dead Sea brine. The modern Lake Kinneret water has Sr/Ca value of ~0.007 and 87Sr/86Sr ratios lying between 0.70747 and 0.70765, different from other freshwater sources in the vicinity (Table 3, and Lev et al., 2014): the Jordan River with low Sr/Ca and 87 Sr/86Sr values (0.0004e0.0010 and 0.7067e0.7072), water that drains the basaltic lithologies of Golan Heights (~0.003e0.005 and 0.70745e0.70750) and Galilee runoff with low Sr/Ca and high 87 Sr/86Sr values (~0.003 and ~0.7082). It is also different from values of the Lake Kinneret saline springs, with high 87Sr/86Sr and Sr/Ca ratios (e.g., Tiberias Spa 0.7078 and 0.01). It is most likely that the composition of the modern Lake Kinneret waters is a result of mixing of freshwater (Jordan River and basaltic water) with the Kinneret saline springs waters. For the modern lake, Kolodny et al. (1999) estimated ~72% of Sr contribution came from the saline springs (both gauged and ungauged springs) and ~28% from Jordan River. Recently, Fruchter et al. (2017) recalculated the relative Sr contributions to be 65% saline springs and 35% Jordan River. In this section, we describe and discuss the evolutionary trends in the composition of the lake's water. Fig. 6B shows the 87Sr/86Sr versus Sr/Cawater ratios in the lake's waters during the past 28 kyrs based
Fig. 5. Extent of Lake Kinneret and Lake Lisan in between 28 and 13 ka according to the lake level reconstructions (Fig. 2). Please note the significant shrinkage of the lake from 23.8 to 22.5 ka (H2).
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
123
Fig. 6. (A)87Sr/86Sr vs. Sr/Ca of the different water types in the drainage area of Lake Kinneret (Table 3). AB - Jordan River at Arik Bridge (entrance to the lake), YB - Jordan River at Yosef Bridge. (B) The variations in 87Sr/86Sr vs. Sr/Ca of the Lake Kinnerett water during the last ~28 kyrs. KIN2 values are inferred from primary calcites and Ohalo trench (Lev et al., 2014), SOG2 and SOG3 values are inferred from ostracods shells. Water values are highly affected from saline springs during wet periods (e.g., high-stand glacial MIS2), runoff during H2 event and low contribution of brines during lower-lake stand (e.g., Holocene) periods. Melanopsis values are from Hazan et al. (2005). (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
Table 3 d18O, Sr/Ca and
87
Sr/86Sr in freshwater and brines in Lake Kinneret and its water sources.
Source FRESHWATER Lake Kinneret Jordan River e Arik Bridge Jordan River e Yosef Bridge Kibbutzim stream Rainwater BRINES Dead Sea Tiberias Spa Tabgha springs Fuliya springs
d18O (‰)
Sr/Ca (eq.)
87
0.0 6.9 to 5.9
0.0076 0.0010 0.0004 0.0024
0.70747 0.70669 0.70723 0.70810 0.70860
1, 2 1, 2 1 3 4
0.01 0.01 0.0111 0.0041e0.0046
0.70802e0.70803 0.70785 0.70781 0.70787e0.70792
1 1, 4 3, 4 5, 6
7.3
þ4.4 to þ4.9 3.0 5.7 to 6.2 1.6
Sr/86Sr
References
1. Stein et al. (1997); 2. Gat (1970); 3. Lev (2006); 4. Gat et al. (1969); 5. Bergelson et al. (1999); 6. Klein-BenDavid et al. (2005).
on C. torosa shells and primary calcites in the cores. The lake's water show distinctly different values during the last glacial (MIS2), the H2 event, the Younger Dryas, the Holocene and modern times indicating changing contributions of source waters. Overall, the lake's waters appear to lie on and between two distinct mixing arrays: (1) Between the saline springs (Tiberias Spa and Tabgha) and the Jordan River; and (2) Between regional runoff from the Galilee Mt and the modern lake's water (Fig. 6B). In the following section, we discuss the mixing arrays for each of the relevant time intervals and describe the relative contribution of each water source to the “lake cocktail” based on the 87Sr/86Sr ratios and Sr/Ca ratios (Table 4). The calculation is based on the mixing equations (1)e(3) used by Stein et al. (1997). The high-stand of last glacial period (28-24 ka) is represented by core KIN2 that comprises mostly primary calcite. 87Sr/86Sr ratios in the precipitating waters range from 0.70771 to 0.70796 and Sr/Ca range from 0.006 to 0.01 (applying partition coefficient of 0.21, Katz and Nishri, 2013). Most of the KIN2 samples show 87Sr/86Sr ratios lying between the ratios of the saline springs (comprising Cachloride brine), freshwater of the northern Jordan River (sampled at Yosef bridge) and runoff waters from the Galilee Mountains
(Fig. 6). Another possibility is that the brine component is the Dead Sea brine that approached Lake Kinneret during MIS2 when the lakes rose to their high stand (~28-24 ka, Fig. 2). The modern saline springs of Tiberias Spa and Tabgha possibly discharge the lake solution that penetrated the Galilee aquifers during MIS2 when the lake was at its high stand (see discussion by Weber et al. (2018) on a similar process occurring at the Dead Sea). According to this view the 87Sr/86Sr and Sr/Ca ratios of the KIN2 calcites during the high stand period (~28-24 ka) reflect enhanced contribution of the Cachloride brines from the southern Lake Lisan, which expanded all over the Jordan Valley overflowing the topographic sills at Wadi Malih and Yarmouk River (Figs. 1 and 2) and converging with the rising Lake Kinneret. The mass balance calculation using the 87 Sr/86Sr and Sr/Ca ratios indicate that during this time interval the “lake cocktail” comprises 93.6% Jordan River waters, 6.3% runoff and 0.1% Dead Sea brine (Table 4). The H2 period (~23.8-22 ka BP) documented in all cores (Fig. 2) was characterized by regional aridity and significant lake level drop (Lev et al., 2014). The lake was, however, affected by floods that wash the regional surface cover. At the lake's margins, the Ohalo-II trench and core SOG2 (Fig. 1), Melanopsis shells and ostracods show
124
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
Table 4 The relative contributions of the different water sources (A) to the “lake cocktail” (B) of Lake Kinneret during different periods based on calculation follows the mixing equations (Stein et al., 1997).
87
Sr/86Sr and Sr/Ca ratios. The
(A) 87
Sr/86Sr
Sr (mg/l)
Ca (g/l)
Sr/Ca (mmol/mol)
0.70675 0.70815 0.70785 0.70800
0.100 0.375 60.0 314
0.044 0.084 3.480 16.7
1.04 2.03 7.89 8.60
Period
Age (ka)
% Jordan River
% Runoff
% Brineb
Holocene YD Post-glacial H2 Glacial
11e0 12.6e11.7 22e12.6 23.8e22 28e23.8
81.2 74.4 47.3 62.9 93.6
18.2 25.1 52.6 36.9 6.3
0.5 0.5 0.1 0.2 0.1
Source a
Jordan River Runoff Tiberias Spa Dead Sea brine (B)
a b
87
Jordan River sampled at Arik Bridge, close to the inlet to Lake Kinneret. Here, the waters comprise mixture of all type of waters from the lake's northern watershed. The brine is Tiberias Spa or Dead Sea brine (for the last glacial).
Sr/86Sr and Sr/Ca ratios of 0.7078e0.7081 and 0.003e0.004, respectively. The values incline toward the composition of the regional runoff that drains the surface cover - the Galilee Mountains soils (e.g. Palchan et al., 2018). The mass balance calculation using the 87Sr/86Sr and Sr/Ca ratios indicate for the H2 time interval that the “lake cocktail” comprises 62.9% Jordan River waters, 36.9% runoff and 0.2% Tiberia Spa type brine (Table 4). There is a significant rise in the contribution of the runoff water to the lake as well as the brine solution. During the Younger Dryas (YD) lake level rose (Fig. 2) and deeper lake conditions are indicated by the occurrence of laminated sediments in the cores. The lake's87Sr/86Sr and Sr/Ca ratios (core SOG2) move away from the regional runoff composition towards the values of the modern lake indicating an enhanced contribution of freshwater from the Jordan sources and the occurrence of freshwater diatoms (Ehrlich, 1985). This, in turn suggests enhanced activity of the Mediterranean winter cyclones that bring rains to Mount Hermon and the southern Levant region. Similar increase in rain contribution to the region during the Younger Dryas is evident also in the lacustrine-sedimentary records of the Dead Sea (Stein et al., 2010; Liu et al., 2013). The mass balance calculation using the 87Sr/86Sr and Sr/Ca ratios indicate for the YD time interval that the “lake cocktail” comprises 74.4% Jordan River waters, 25.1% runoff and 0.5% Tiberia Spa type brine (Table 4). The (late) Holocene period is presented only in core SOG3 showing 87Sr/86Sr and Sr/Ca ratios that shift toward the modern lake composition (0.70766 and 0.007, respectively). This shift indicates also a reduced contribution of the Ca-chloride brine compared to the high-stand ~28-24 ka period. The mass balance calculation using the 87Sr/86Sr and Sr/Ca ratios indicates that at the late Holocene time interval the “lake cocktail” comprises 81.2% Jordan River waters, 18.2% runoff and 0.5% Tiberia Spa type brine. We see an enhanced flow of the Jordan River but higher dicharge of the brines. Mg/Ca ratios in the Holocene carbonates (ostracod shells) lie in the lower part of the range shown also by the last glacial primary calcite samples (Fig. 4, Supplementary Figs. 1e3). Both carbonate types show similar ratios when collected from close stratigraphic units (e.g. Fig. 4) implying no significant variation in the Kd (partition coefficient) when ostracod or primary calcite is considered. The higher Mg/Ca ratio in many of the glacial (MIS2) samples corroborate the conclusion derived from 87Sr/86Sr and Sr/Ca ratios on significant contribution of the Dead Sea brine to Lake Kinneret waters during the high stand period, yet the Mg/Ca in the water did not reach the value that allowed aragonite precipitation as the
primary calcium carbonate. Primary aragonite was not identified in the sedimentary sections of the last glacial high stand period of Lake Kinneret (Supplementary Tables 2e4). It appears that the enhanced input of freshwater from Jordan River and Galilee runoff to the lake inhibited aragonite deposition leaving the lake water to precipitate primary calcite (Hazan et al., 2005). 5.5. Oxygen isotopes and the regional hydrological regime In this section, we discuss the results of the oxygen isotope analyses of the cores and combine the d18O data with the information from the 87Sr/86Sr eSr/Ca ratios (Figs. 4, 6 and 7) to establish a model for the hydrological regime of the lake during the past 28 kyrs. Oxygen isotopes values were measured on the ostracod shells from core SOG2 and on the primary calcites in core KIN2 (Lev et al., 2014; Supplementary Figs. 1e3 and Supplementary Tables 8e11). The calcites from the lower part of core KIN2 (spanning the time of ~28-24 ka) yielded d18O ¼ 4.5‰ (VPDB) while those from the upper part of the core (~23-22 ka) show more negative values of d18O ~ 5.5‰ (VPDB). The conversion from the d18O VPDB values measured on the ostracods shells or the primary calcites (reported in VPDB) to water values reported in VSMOW uses the following equation:
d18OVSMOW ¼ 1.03086xd18OVPDBþ30.86
(1)
Then, we calculate the d18O isotope composition of the lake's water by using the following equation:
d18Ocalcite (VSMOW) -d18Owater (VSMOW) ¼ (2.78 106)/T2 e 2.89
(2)
where T is oK The T values are taken from Zaarur et al. (2016) who applied the clump isotope method to Melanopsis shells from Lake Kinneret. For the calculation, we used values of Melanopsis shells that have similar ages to the core samples. Comparison of the water composition obtained by the clumped isotopes method (on Melanopsis shells) with those calculated from the primary calcites both in the fossil and the modern shells gives similar values (Fig. 4). Zaarur et al. (2016) calculated temperatures of 18.6e22.5 C at ~ 22.5 ka and 20e27 C around 10-5 ka, with corresponding d18Owater of ~ 2.5‰ and ~0‰, respectively. The temperature corrected water-carbonate-shell fractionation is about 2‰. Assuming a similar value for the water/primary calcites or ostracods, we can plot the d18Owater along the KIN2, Ohalo and SOG3 cores (Figs. 4D
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
125
Fig. 7. Variations in d18Owater values against 87Sr/86Sr during the last 28 kyrs. The figure shows that for both the d18O and the 87Sr/86Sr isotopes values the last glacial lake's water lie between the Jordan River waters and the Dead Sea brine while the Holocene and recent lake's water deviate significantly from this mixing array. We note that during the H-events and the Holocene and the Recent the d18Owater move to more positive values (~0) that possibly reflect enhanced evaporation (see text). See Table 4 for the relative contribution of the different water sources. Note the resemblance between the composition of Tiberias spa and the lake water during the high stand that support the conclusion that Tiberias spa represents the last glacial lake water. (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.)
and 7). The figure shows that lake waters were lighter (~3‰) during the high stand period (~170 m bsl at ~26 ka) than at the Holocene or the Recent, when reaching a value of ~0‰. This trend works in an opposite direction to the global trend (as recorded in the marine foraminifera) or speleothems from nearby caves (e.g., Bar-Matthews et al., 2003 and see Fig. 8). The general pattern of d18O in the cave speleothems follows that of the Mediterranean Sea during the past ~90 kyrs (the last Glacial and the Holocene periods) (Kolodny et al., 2005; Almogi-Labin et al., 2009). Thus, in Lake Kinneret the transition from the last glacial to the Holocene was accompanied by a positive shift in the d18O values (reaching zero at present). Similar positive diversion in the lake's water d18O values during the same time interval was observed by Zaarur et al. (2016) who analyzed Melanopsis shells that live at the lake's shores (Fig. 4D). Zaarur et al. (2016) interpreted the diversion of the oxygen isotope values in Lake Kinneret from the regional and global trends as reflecting the change in the regional hydrology and precipitation pattern in the watershed of the lake. They argued that during the last glacial MIS2 the lake was predominantly fed by snow-melting waters from Mt. Hermon that were characterized by very low d18O values while during the Holocene and modern time the rivers reaching the lake are fed mainly by rain-water in the Jordan River watershed. Similar positive diversion in the d18O values in the Holocene was reported for the Yammouneh Lake in Lebanon (located at the north-west flanks of Mount Hermon). The Holocene d18O values in this lake are heavier by 3‰ during the Holocene compared to the glacial values (Fig. 8 and see Develle et al., 2010). The shift towards more positive d18O values and less radiogenic 87Sr/86Sr ratios in the Holocene and modern lake's water
Fig. 8. (A) Lake Kinneret water level curve (simplified curve, see more details in Fig. 2). (B) Temporal changes in 87Sr/86Sr in Lake Kinneret waters that reflect contributions and mixing of three water types (Jordan River, regional runoff and Dead Sea CaChloride brine). (C). d18Owater in Lake Kinneret. The lake's waters change from more d18Owater negative (and larger spread) values during the last glacial to more positive values at the Holocene and the modern Lake. The large spread in the last glacial values reflects the limnological-hydrological properties of the lake during different time intervals of the last glacial, e.g., Heinrich events or the time interval of high lake stand) (D) d18Owater of Yammouneh Lake, Lebanon (Develle et al., 2010); (E) d18O of Soreq Cave speleothems, Israel (Bar-Matthews et al., 2003). The speleothem values reflect mainly regional rains that originate from the Mediterranean Sea (Bar-Matthews et al., 2003; Kolodny et al., 2005). Note that the temporal pattern of d18Owater in Lake Kinneret reflects mainly the hydro-climate conditions in the vicinity of the lake e.g., the enhanced evaporation during the Holocene.
are consistent with diminishing contribution of the Dead Sea brine to Lake Kinneret and enhanced evaporation in the Holocene and the modern lake (see Figs. 4 and 7). 5.6. The hydro-climate of last glacial and Holocene periods at the southern Levant The limnological-geochemical data from the last glacial and Holocene Lake Kinneret reflect the hydro-climate conditions in Mount Hermon water-head and other parts of the lake catchment area: The Galilee Mountains and Golan Height (Fig. 1). All these regions are mainly affected by the winter rains associated with the Mediterranean cyclones (Dayan, 1986, Enzel et al., 2003, 2008). The
126
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
(independently) reconstructed lake level curves of both Lake Lisan and Lake Kinneret indicate surface level rise and lakes expansion during the last glacial MIS2 period (reaching lake convergence between ~ 28-24 ka). The 87Sr/86Sr isotope ratios and the Sr/Ca ratios (Fig. 6) indicate enhanced contribution of Galilee runoff water as well as Jordan River and Ca-chloride brine during the high stand period. The amounts of primary calcites and primary aragonite in Lake Kinneret and Lake Lisan, respectively and their Sr/Ca ratios attest to the enhanced supply of freshwater to the lake (see calculation for Lake Lisan in Stein et al., 1997). Higher rain precipitation during the last glacial MIS2 in the vicinity of Lake Kinneret is supported also by d18O speleothem data from the recently explored Zalmon Cave (Keinan, 2018). The cave is located within the drainage area of Nahal Zalmon that drains to the lake at the Genosar Valley close to core SOG3 location (Fig. 1). The cold weather conditions in vicinity of Lake Kinneret during the last glacial MIS2 arrived to their climax at Mount Hermon area where part of the precipitation was in the form of snow, which was transformed to ice as evident in the mountain's carbonate cave record that indicates a break in speleothem growth during MIS2 (Ayalon et al., 2013). The inputs of runoff and groundwater from the Galilee Mountains and Golan Height sources to Lake Kinneret had to surpass the waters trapped as ice in Mount Hermon. A modern example for this behavior is the severe cold and rainy winter of 1991/2 when several rain and snow storms occurred in the Lake Kinneret watershed and rain precipitation in the Judea and Galilee Mountains was a few times higher than the annual average. During the Holocene, the contribution of Galilee runoff and ground water diminished as reflected in the 87Sr/86Sr and Sr/Ca ratios of the Holocene ostracods (Fig. 6). Comparison of the last Glacial (MIS2) hydro-climate patterns in the south Levant (e.g. Israel area) and other segments of the north Mediterranean and Turkey points to an apparent contradiction. Lake Van pollen data (Pickarski and Litt, 2017) indicate cold and dry conditions, while the lake level data from Lake Lisan and Lake Kinneret as well as the indication for enhanced supply of bicarbonate and Ca to the lakes and enhanced formation of primary calcites (Lake Kinneret) and aragonite Lake Lisan, all indicate enhanced inputs of freshwater to the lakes (Stein et al., 1997; Barkan et al., 2001). It should also be noted that recent study of pollen grains collected from the KIN2 core indicate lower proportion of Mediterranean trees in the Galilee Mountains during MIS2 relative to the Holocene (Miebach et al., 2017). This picture is also consistent with the pollen data recovered from the east Mediterranean core off the Israel shores (Langgut et al., 2011). Thus, the pollen data in the east Mediterranean mountain regions of the Galilee, Greece and Turkey point to drier conditions in the conventional way of interpretation of pollen data. There are several ways to deal with this apparent contradiction. The pollen data reflect the adaptation of the vegetation to the regional hydro-climate conditions. We suggest that the regime of enhanced runoff and floods along with the heavy dust storms and strong winds and the lower temperatures over the Judea Hills and the Galilee mountains during MIS2 (e.g. Rohling, 2013; Palchan et al., 2018) could inhibit the growth of Mediterranean trees. Moreover, the Lisan Formation comprises sequences of primary aragonite and silty detritus that contain re-mobilized desert dust (composed mainly of quartz and calcite grains). Thus, the desert dust was rapidly re-mobilized to Lake Lisan as is also evident in some modern winter storms (Belmaker et al., 2011). It may be that only during the Holocene accumulation of the Terra Rossa or other mountain soils could enhance the growth of the Mediterranean trees. This matter is not concluded and requires further attention. Overall, both the Dead Sea and Lake Kinneret received during the peak of the last glacial (MIS2) large amounts of freshwaters laden
with bi-carbonate. This required enhanced precipitation in the lake's watershed both from Mount Hermon system and regional runoff. During the post-glacial period and the Holocene Lake Kinneret received less fresh and saline waters reflecting the increasing aridity in its watershed. 6. Conclusions The hydrological and limnological history of the late Quaternary (past 28 kyrs) Lake Kinneret (Sea of Galilee) was established using sedimentological, chemical (Mg/Ca and Sr/Ca), d18O and Sr-isotope data of sediment cores recovered from the lake at water depths of 213.0e230.0 m bsl. The cores comprise sequences of massive shallow-water deposits and deeper water laminated marls dominated by primary calcites. The chronology of the cores was established by radiocarbon dating of charcoal, wood and shells of the ostracod C. torosa. The Sr/Ca and 87Sr/86Sr ratios of primary calcite and C. torosa shells were combined with data of contemporaneous water sources in the lake and its vicinity to constrain their contribution to the lake during the past 28 kyrs. The main results and conclusions of this study are: 1. Laminated marls of primary calcites and fine detritus were deposited during intervals of enhanced freshwater input, high lake stands and stratified lake-configuration, during early MIS2, the Younger Dryas, and the mid-Holocene. Massive marls, comprising detritus particles (e.g. quartz, calcite, clays) were deposited during periods of low lake stands and enhanced floods mainly during the intervals of the Heinrich events, e.g. H2, H1 and the late Holocene. 2. The lithology of the cores is used as an indicator of relative water depth: Massive marls with high content of coarser grains (>63 mm) and presence of ostracod shells indicate shallow water environment. Laminated fine-grain sediments with no ostracod shells indicate higher lake level when core or trench sites were in deeper water environment. 3. Sedimentation ceased in the south-western and southern part of the lake (at water depths of 10e20 m) after ~19 and ~12 ka respectively indicating decrease in sediment supply to this part of the basin causing “starved basin” conditions. 4. Since the last glacial period the lake has received waters from four sources, each with distinct 87Sr/86SreSr/Ca ratios: (1) The Dead Sea Ca-chloride brine with 87Sr/86Sr ~ 0.7080 and Sr/Ca (eq) ~ 0.009. (2) Regional runoff (that washes the surface cover) with 87 Sr/86Sr ~ 0.7081 and Sr/Ca (eq) ~ 0.002. (3) Jordan River waters with 87Sr/86Sr ~ 0.70675 and Sr/Ca (eq) ~ 0.001. (4) Tiberias Spa brine with 87Sr/86Sr ~ 0.70785 and Sr/Ca (eq) ~ 0.008. The Tiberias Spa brine probably represents the last glacial solution of the lake that intruded the regional aquifers during the lake high stand and is currently discharging to the modern low stand lake. 5. The limnologicalehydrological history of Lake Kinneret during the past 28 kyrs: During the MIS2 high stand period (~28-24 ka, ~170 m bsl) the lake was significantly affected by the Dead Sea Ca-chloride brine and the Jordan River. During the H2 low stand period (below current elevation of ~214 m bsl), contributions of Jordan River and saline springs diminished. It appears that meteoric precipitation on Mount Hermon nearly ceased. The lake margins were affected by seasonal floods when runoff transported the soil-surface cover of Galilee Mountains to the lake.
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
During the Younger Dryas the lake level rose, depositing laminated marls. The lake received larger freshwater input from the Jordan River as compared to the H2 period. During the late Holocene and modern times the lake has received diminishing contributions of runoff waters and saline springs and more of the Jordan River. 6. The MIS2 high-stand period is characterized by enhanced contributions of Jordan River that is a mixture comprising all fresh water types from the lake watershed. This picture is consistent with the enhanced activity of the Mediterranean cyclones and increased contribution of winter rains to the southern Levant and to Mount Hermon (the largest “water reservoir” of the southern Levant). During the Holocene and modern times and particularly during the H2 period the contribution of freshwater from Mount Hermon via the Jordan River diminished indicating the weakening of the Mediterranean cyclones and the aridification of the southern Levant region. Acknowledgments We wish to thank N. Tepelyakov, I. Segal, O. Yoffe, O. Berlin, T. Zilberman, N. Teutsch and D. Stiber (Geochemistry lab) and M. Kitin (Micropaleontology lab), all from the Geological Survey of Israel. Dr. I. Stefanova from the Department of Earth Sciences, University of Minnesota, USA is warmly thanked for her help in applying the BACON software. We thank the anonymous reviewers for very thoughtful and constructive reviews, and the editorial handling of Ana Moreno, all of which greatly improved this manuscript. This research was supported by USAeIsrael Binational Science Foundation BSF grant 2010347(to AA and EI) and Israel Science Foundation ISF grant 1663/16(to MS and AA). Appendix A. Supplementary data Supplementary data to this article can be found online at https://doi.org/10.1016/j.quascirev.2019.02.015. References Almogi-Labin, A., Bar-Matthews, M., Shriki, D., Kolosvsky, E., Paterne, M., Schilman, B., Ayalon, A., Aizenshtat, Z., Matthews, A., 2009. Climatic variability during the last ~90 ka of the southern and northen Levantine Basin as evident from marine records and speleothems. Quat. Sci. Rev. 28, 2882e2896. Almogi-Labin, A., Schilman, B., Flako-Zaritsky, S., 2004. Micro-faunal ecosystem of the Timsah springs: environmental and stable isotopes characterization. In: Israel Geological Survey Report GSI/29/2004, 46 pp. Athersuch, J., Home, D.J., Whittaker, J.E., 1989. Marine and brackish water ostracods (superfamilies cypridacea and cytheracea). In: Synopses of the British Fauna (New Series) No. 43. Linnean Society of London and Estuarine and Coastal Sciences Association. E. J. Brill, Leiden, p. 343. Ayalon, A., Bar-Matthews, M., Frumkin, A., Matthews, A., 2013. Last Glacial warm events on Mount Hermon: the southern extension of the Alpine karst range of the east Mediterranean. Quat. Sci. Rev. 59, 43e56. Barkan, E., Luz, B., Lazar, B., 2001. Dynamics of carbon dioxide system at the Dead Sea. Geochem. Cosmochim. Acta 65, 355e368. Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A., Hawkesworth, C.J., 2003. Seaeland oxygen isotopic relationships from planktonic foraminifera and speleothems in the Eastern Mediterranean region and their implication for paleorainfall during interglacial intervals. Geochem. Cosmochim. Acta 67, 3181e3199. Bartov, Y., Goldstein, S.L., Stein, M., Enzel, Y., 2003. Catastrophic arid episodes in the eastern Mediterranean linked with the North Atlantic Heinrich events. Geology 31, 439e442. Bartov, Y., Stein, M., Enzel, Y., Agnon, A., Reches, Z., 2002. Lake levels and sequence stratigraphy of Lake Lisan, the late Pleistocene precursor of the Dead Sea. Quat. Res. 57, 9e21. Belmaker, R., Lazar, B., Stein, M., Beer, J., 2011. Short residence time and fast transport of fine detritus in the Judean Desert: clues from 7Be in settled dust. Geophys. Res. Lett. 38, L16714. https://doi.org/10.1029/2011GL048672. Ben-Avraham, Z., Amit, G., Golan, A., Begin, Z.B., 1990. The bathymetry of Lake Kinneret and its structural significance. Isr. J. Earth Sci. 39, 77e84.
127
Bergelson, G., Nativ, R., Bein, A., 1999. Salinization and dilution history of ground water discharging into the Sea of Galilee, the Dead Sea Transform, Israel. Appl. Geochem. 14, 91e118. Blaauw, M., Christen, A., 2011. Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Anal. 6, 457e474. Bookman, R., Enzel, Y., Agnon, A., Stein, M., 2004. Late Holocene Lake levels of the Dead Sea. Geol. Soc. Am. Bull. 116, 555e571. Cook, B.I., Anchukaitis, K.J., Touchan, R., Meko, D.M., Cook, E.R., 2016. Spatiotemporal drought variability in the Mediterranean over the last 900 years. J. Geophys. Res.: Atmosphere 121, 2060e2074. Crouvi, O., Amit, R., Enzel, Y., Porat, N., Sandler, A., 2008. Sand dunes as a major proximal dust source for late Pleistocene loess in the Negev Desert, Israel. Quat. Res. 70, 275e282. Dayan, U., 1986. Climatology of Back trajectories from Israel based on synoptic analysis. J. Clim. Appl. Meteorol. 25, 591e595. Develle, A.-L., Herreros, J., Vidal, L., Sursock, A., Gasse, F., 2010. Controlling factors on a paleo-lake oxygen isotope record (Yammoûneh, Lebanon) since the Last Glacial Maximum. Quat. Sci. Rev. 29, 865e886. Dubowski, Y., Erez, J., Stiller, M., 2003. Isotopic paleolimnology of Lake Kinneret. Limnol. Oceanogr. 48, 68e78. Ehrlich, A., 1985. The eco-biostratigraphic significance of the fossil diatoms of Lake Kinneret. GSI Curr. Res. 5, 24e30. Enzel, Y., Bookman, R., Sharon, D., Gvirtzman, H., Dayan, U., Ziv, B., Stein, M., 2003. Late Holocene climates of the Near East deduced from Dead Sea level variations and modern regional winter rainfall. Quat. Res. 60, 263e273. Enzel, Y., Amit, R., Dayan, U., Crouvi, O., Kahana, R., Ziv, B., Sharon, D., 2008. The climatic and physiographic controls of the eastern Mediterranean over the late Pleistocene climates in the southern Levant and its neighboring deserts. Glob. Planet. Chang. 60, 165e192. Fruchter, N., Lazar, B., Nishri, A., Almogi-Labin, A., Eisenhauer, A., Beeri-Shlevin, Y., Stein, M., 2017. 88Sr/86Sr fractionation and calcite accumulation rate in the Sea of Galilee. Geochem. Cosmochim. Acta 215, 17e32. Ganor, E., Deutsch, Y., Foner, H.A., 2000. Mineralogical composition and sources of airborne settling particles on Lake Kinneret (the Sea of Galilee), Israel. Water, Air, Soil Pollut. 118, 245e262. Gat, J.R., 1970. Environmental isotope balance of Lake Kinneret, isotope hydrology. In: Symposium on the Use of Isotopes in Hydrology. IAEA, pp. 109e127. Gat, J.R., Mazor, E., Tzur, Y., 1969. The stable isotope composition of mineral waters in the Jordan Rift Valley. Isr. J. Earth Sci. 7, 334e352. Goldstein, T., 2004. Lake Flood Craters in the Northwestern Part of Lake Kinneret Structure, Evolution and Relation to Water Flow (Hebrew). M.Sc. Thesis. Tel Aviv University, Israel. Hazan, N., 2003. Reconstruction of Kinneret Lake Levels in the Last 40,000 Years. M.Sc Thesis. The Hebrew University of Jerusalem, Israel. Hazan, N., Stein, M., Agnon, A., Marco, S., Nadel, D., Negendank, J.F.W., Schwab, M.J., Neev, D., 2005. The late Quaternary limnological history of Lake Kinneret (Sea of Galilee), Israel. Quat. Res. 63, 60e77. Hazan, N., Stein, M., Marco, S., 2004. Lake Kinneret levels and active faulting in the Tiberias area. Isr. J. Earth Sci. 53, 199e205. Heimann, A., Braun, D., 2000. Quaternary stratigraphy of the Kinnarot Basin, Dead Sea Transform, northeastern Israel. Isr. J. Earth Sci. 49, 31e44. Held, I.M., Soden, B.J., 2006. Robust responses of the hydrological cycle to global warming. J. Clim. 19, 5686e5699. Kallel, N., Paterne, M., Duplessy, J.C., Vergnaud-Grazzini, C., Pujol, C., Labeyrie, L., Arnold, M., Fontugne, M., Pierre, C., 1997. Enhanced rainfall in the Mediterranean region during the last sapropel event. Oceanol. Acta 20, 697e712. Katz, A., Nishri, A., 2013. Calcium, magnesium and strontium cycling in stratified, hardwater lakes: Lake Kinneret (Sea of Galilee), Israel. Geochem. Cosmochim. Acta 105, 372e394. Keinan, J., 2018. Paleo-environment of the Northern Jordan Rift Based on Speleothems from the Zalmon Cave, Israel. MSc Thesis. The Hebrew University of Jerusalem, Israel. Klein-BenDavid, O., Gvirtzman, H., Katz, A., 2005. Geochemical identification of fresh water sources in brackish groundwater mixtures; the example of Lake Kinneret (Sea of Galilee), Israel. Chem. Geol. 214, 45e59. Kolodny, Y., Katz, A., Starinsky, A., Moise, T., Simon, E., 1999. Chemical tracing of salinity sources in Lake Kinneret (Sea of Galilee), Israel. Limnol. Oceanogr. 44, 1035e1044. Kolodny, Y., Stein, M., Machlus, M., 2005. Seaerainelake relation in the last glacial east Mediterranean revealed by d18Oed13C in lake lisan aragonites. Geochem. Cosmochim. Acta 69, 4045e4060. Langgut, D., Almogi-Labin, A., Bar-Matthews, M., Weinstein-Evron, M., 2011. Vegetation and Climate changes in the South-Eastern Mediterranean during the Last Glacial-Interglacial cycle (86 ka): new marine pollen record. Quat. Sci. Rev. 30, 3960e3972. Lev, L., 2006. Melanopsis Shells as Chronometers and Paleo-Environmental Tracers of Lake Kinneret and its Vicinity, Israel. M.Sc. Thesis. Tel Aviv University, Tel Aviv. Lev, L., Almogi-Labin, A., Mischke, S., Ito, E., Ben-Avraham, Z., Stein, M., 2014. Paleohydrology of Lake Kinneret during the Heinrich event H2. Palaeogeogr. Palaeoclimatol. Palaeoecol. 396, 183e193. Lev, L., Boaretto, E., Heller, J., Marco, S., Stein, M., 2007. The feasibility of using Melanopsis shells as radiocarbon chronometers, Lake Kinneret, Israel. Radiocarbon 49, 1003e1015. Liu, T., Broecker, W.S., Stein, M., 2013. Rock varnish evidence for a younger Dryas
128
L. Lev et al. / Quaternary Science Reviews 209 (2019) 114e128
wet period in the Dead Sea basin of the Near East. Geophys. Res. Lett. 40, 1e7. Marco-Barba, J., Ito, E., Carbonell, E., Mesquita-Joanes, F., 2012. Empirical calibration of shell chemistry of Cyprideis torosa (Jones, 1850) (Crustacea: Ostracoda). Geochem. Cosmochim. Acta 93, 143e163. Marcus, E., Slager, J., 1985. The sedimentary-magmatic sequence of Zemah 1 well (Jordan-Dead Sea Rift, Israel) and its emplacement in time and space. Isr. J. Earth Sci. 34, 1e10. Miebach, A., Chen, C., Schwab, M.J., Stein, M., Litt, T., 2017. Vegetation and climate during the Last Glacial high stand (ca. 28-22 ka BP) of the Sea of Galilee, Northern Israel. Quat. Sci. Rev. 156, 47e56. Migowski, C., Stein, M., Prasad, S., Negendank, J.F.W., Agnon, A., 2006. Holocene climate variability and cultural evolution in the Near East from the Dead Sea sedimentary record. Quat. Res. 66, 421e431. Mischke, S., Almogi-Labin, A., Al-Saqarat, B., Rosenfeld, A., Elyashiv, H., Boomer, I., Stein, M., Lev, L., Ito, E., 2014. An expanded ostracod-based conductivity transfer function for climate reconstruction in the Levant. Quat. Sci. Rev. 93, 91e105. Mischke, S., Almogi-Labin, A., Ortal, R., Rosenfeld, A., Schwab, M.J., Boomer, I., 2010. Quantitative reconstruction of lake conductivity in the Quaternary of the Near East (Israel) using ostracods. J. Paleolimnol. 43, 667e688. Nadel, D., Tsatskin, A., Bar-Yosef Mayer, D.E., Belmaker, M., Boaretto, E., Kislev, M.E., Hershkovitz, I., Rabinovich, R., Simmons, T., Weiss, E., Zohar, I., Asfur, O., Emmer, G., Ghraieb, T., Grinberg, E., Halabi, H., Weissbrod, L., Zaidner, Y., 2002. The Ohalo II 1999e 2000 seasons of excavation: a preliminary report. In: Mitekufat Haeven, Journal of the Israel Prehistoric Society, vol. 32, pp. 17e48. Palchan, D., Stein, M., Goldstein, S.L., Almogi-Labin, A., Tirosh, T., Erel, Y., 2018. Synoptic conditions of fine-particle transport to the last interglacial Red SeaDead Sea from Nd-Sr compositions of sediment cores. Quat. Sci. Rev. 179, 123e136. Pickarski, N., Litt, T., 2017. A new high-resolution pollen sequence at Lake Van, Turkey: insights into penultimate interglacialeglacial climate change on vegetation history. Clim. Past 13, 689e710. Rohling, E.J., 2013. Quantitative assessment of glacial fluctuations in the level of Lake Lisan, Dead Sea rift. Quat. Sci. Rev. 70, 63e72, 2013. Sandler, A., Meunier, A., Velde, B., 2015. Mineralogical and chemical variability of mountain red/brown Mediterranean soils. Geoderma 239e240, 156e167. Serruya, S., 1975. Wind, water temperature and motions in Lake Kinneret: general pattern. Verh. Internat. Verein. Limnol. 19, 73e87. Starinsky, A., 1974. Relationship between Ca-Chloride Brines and Sedimentary Rocks in Israel (In Hebrew). Ph.D. Thesis. The Hebrew University of Jerusalem, Jerusalem. Stein, M., 2001. The history of neogene-quaternary water bodies in the Dead Sea basin. J. Paleolimnol. 26, 271e282. Stein, M., 2014a. Late Quaternary limnological history of lake Kinneret. In: Zohary, T., Sukenik, A., Berman, T., Nishri, A. (Eds.), Lake Kinneret. Springer Netherlands, pp. 39e58.
Stein, M., 2014b. The evolution of Neogene-Quaternary water-bodies in the Dead Sea rift valley. In: Modern Approaches in Solid Earth Sciences. Springer, Netherlands, pp. 279e316. Stein, M., Starinsky, A., Agnon, A., Katz, A., Raab, M., Spiro, B., Zak, I., 2000. The impact of brine-rock interaction during marine evaporite formation on the isotopic Sr record in the oceans: evidence from Mt. Sedom, Israel. Geochem. Cosmochim. Acta 64, 2039e2053. Stein, M., Starinsky, A., Katz, S., Goldstein, S.L., Machlus, M., Schramm, A., 1997. Strontium isotopic, chemical and sedimentological evidence for the evolution of Lake Lisan and the Dead Sea. Geochem. Cosmochim. Acta 61, 3975e3992. Stein, M., Torfstein, A., Gavrieli, I., Yechieli, Y., 2010. Abrupt aridities and salt deposition in the post-glacial Dead Sea and their North Atlantic connection. Quat. Sci. Rev. 29, 567e575. Stiller, M., 1974. Rate of Transport and Sedimentation in Lake Kinneret. Ph.D Thesis. Weizmann Institute of Science, Rehovot. Stiller, M., 1977. Origin of sedimentation components in Lake Kinneret, traced by their isotopic composition. In: Golterman, H.L. (Ed.), Interaction between Sediments and Freshwater. Junk, pp. 57e64. Stiller, M., Carmi, I., Kaufman, A., 1988. Organic and inorganic 14C concentrations in the sediments of Lake Kinneret and the Dead Sea (Israel) and the factors which control them. Chem. Geol. 73, 63e78. Stiller, M., Kaufman, A., 1985. Paleoclimatic trends revealed by the isotopic composition of carbonates in Lake Kinneret. In: Zeitschrift fur Gletscherkunde und Glazialgeologie, vol. 21, pp. 79e87. Tchernov, E., 1975. The molluscs of the Sea of Galilee. Malacologia 15, 147e184. Torfstein, A., Goldstein, S.L., Enzel, Y., Stein, M., 2013. Impacts of abrupt climate changes in the Levant from last glacial Dead Sea levels. Quat. Sci. Rev. 69, 1e7. Wansard, G., De Deckker, P., Julia, R., 1998. Variability in ostracod partition coefficients D(Sr) and D(Mg) implications for lacustrine palaeoenvironmental reconstructions. Chem. Geol. 146, 39e54. Wansard, G., 1996a. Quantification of paleotemperature changes during Isotopic Stage 2 in the La Draga continental sequence (NE Spain) based on the Mg/Ca ratio of freshwater ostracods. Quat. Sci. Rev. 15, 237e245. Wansard, G., 1996b. Nouvelle approche de la quantification des paleotemperatures a partir du rapport (Mg/Ca) des valves d'ostracods lacustres. In: C.R. Acad. Sci. Paris, Ser IIa 323, pp. 493e500. Weber, N., Yechieli, Y., Stein, M., Yokochi, R., Gavrieli, I., Zappala, J., Mueller, P., Lazar, B., 2018. The circulation of the Dead Sea brine in the regional aquifer. Earth Planet. Sci. Lett. 493, 242e261. Zaarur, S., Affek, H.P., Stein, M., 2016. Last Glacial-Holocene temperatures and hydrology of the Sea of Galilee and Hula valley from clumped isotopes in Melanopsis shells. Geochem. Cosmochim. Acta 179, 142e155. Zak, I., 1967. Geology of Mount Sedom (In Hebrew). Ph.D. Thesis. the Hebrew University of Jerusalem, Jerusalem.