Hostname: page-component-848d4c4894-xfwgj Total loading time: 0 Render date: 2024-06-23T05:31:12.888Z Has data issue: false hasContentIssue false

Paleoclimate reconstruction in the Levant region from the geochemistry of a Holocene stalagmite from the Jeita cave, Lebanon

Published online by Cambridge University Press:  20 January 2017

Sophie Verheyden
Affiliation:
Université Libre de Bruxelles, Département des Sciences de la Terre et de l’Environnement (DSTE), CP 160/02 Avenue F.D. Roosevelt 50, 1050 Brussels, Belgium National Fund for Scientific Research (FNRS), Belgium
Fadi H. Nader
Affiliation:
Department of Geology, American University of Beirut, PO Box: 11-0236/2010 Beirut – Lebanon
Hai J. Cheng
Affiliation:
Department of Geology and Geophysics, University of Minnesota, 310 Pillsbury Drive, SE, Minneapolis, MN 55455, USA
Lawrence R. Edwards
Affiliation:
Department of Geology and Geophysics, University of Minnesota, 310 Pillsbury Drive, SE, Minneapolis, MN 55455, USA
Rudy Swennen
Affiliation:
Geologie, Katholieke Universiteit Leuven, Celestijnenlaan 200E, B-3001 Heverlee – Leuven, Belgium

Abstract

Dated oxygen and carbon isotopic profiles from a Holocene stalagmite (11.9–1.1 ka) from the Jeita cave, Lebanon, are compared to variations in crystallographic habit, stalagmite diameter and growth rate. The profiles show generally high δ18O and δ13C values during the late-glacial period, low values during the early Holocene, and again high values after 5.8 ka. On the basis of the good correlation between the morphological and crystallographic aspect of the stalagmite and its isotopic records, as well as the isotopic response of speleothems from central and northern Israel, we relate high δ18O and δ13C values to drier conditions. Between 6.5 and 5.8 ka an increase in isotopic values, a decrease in growth rate and stalagmite diameter suggest a transition from wet conditions in the early Holocene towards drier conditions in the mid-Holocene. The transition occurred in two steps, first a progressive change to drier conditions started at 6.5 ka but was interrupted by a short (∼ 100 years) return to wetter conditions, followed by an equally rapid (< 200 years) change to drier conditions.

Type
Original Articles
Copyright
University of Washington

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Alley, R.B., (2000). The Younger Dryas cold interval as viewed from central Greenland. Quaternary Science Reviews 19, 213226.Google Scholar
Alley, R.B., Mayewski, P.A., Sowers, T., Stuiver, M., Taylor, K.C., Clark, P.U., (1997). Holocene climate instability: a prominent wide-spread event 8200 yr. ago. Geology 25, (6) 483486.2.3.CO;2>CrossRefGoogle Scholar
Baker, A., Smart, P.L., (1995). Recent flowstone growth rates: field measurements in comparison to theoretical predictions. Chemical Geology 122, 121128.Google Scholar
Bar-Matthews, M., Ayalon, A., Matthews, A., Sass, E., Halicz, L., (1996). Carbon and oxygen isotope study of the active water-carbonate system in a karstic Mediterranean cave: implications for paleoclimate research in semiarid regions. Geochimica et Cosmochimica Acta 60, (2) 1996337347.Google Scholar
Bar-Matthews, M., Ayalon, A., Kaufman, A., (1997). Late Quaternary paleoclimate in the eastern Mediterranean region from stable isotope analysis of speleothems at Soreq cave, Israel. Quaternary Research 47, 155168.CrossRefGoogle Scholar
Bar-Matthews, M., Ayalon, A., Kaufman, A., Wasserburg, G.J., (1999). The Eastern Mediterranean paleoclimate as a reflection of regional events: Soreq cave Israel. Earth and Planetary Science Letters 166, 8595.CrossRefGoogle Scholar
Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A., Hawkesworth, C.J., (2003). Sea–land oxygen isotopic relationships from planctonic foraminifera and speleothems in the Eastern Mediterranean region and their implication for palaeorainfall during interglacial intervals. Geochimica et Cosmochimica Acta 67, (17) 200331813199.Google Scholar
Bjorck, S., Kromer, B., Johnsen, S., Bennike, O., Hammarlund, D., Lemdahl, G., Possnert, G., Rasmussen, T.L., Wohlfarth, B., Hammer, C.U., Spurk, M., (1996). Synchronized terrestrial–atmospheric deglacial records around the North Atlantic. Science 274, 11551160.Google Scholar
Bjorck, S., Rundgren, O., Ingolfsson, O., Funder, S., (1997). The Preboreal oscillation around the Nordic seas: terrestrial and lacustrine responses. Journal of Quaternary Science 12, 455465.Google Scholar
Burns, S.J., Fleitman, D., Matter, A., Kramers, J., Al-Subbary, A.A., (2003). Indian ocean climate and an absolute chronology over Dansgaard/Oeschger events 9 to 13. Science 301, 13651367.CrossRefGoogle Scholar
Cheng, H.J., Edwards, R.L., Ho, J., Gallup, C.D., Richards, D.A., Asmerom, Y., (2000). On the half lives of uranium-234 and thorium-230. Chemical Geology 169, 1733.CrossRefGoogle Scholar
deMenocal, P.B., (2001). Cultural response to climate change during the Late Holocene. Science 292, 667673.Google Scholar
Dubertret, L., (1975). Introduction à la carte géologique au 1/50000 du Liban. Notes et Mémoires sur le Moyen-Orient 23, 345403.Google Scholar
Dykoski, C.A., Edwards, R.L., Cheng, H., Yuan, D., Cai, Y., Zhang, M., Lin, Y., Qing, J., An, Z.S., Revenaugh, J., (2005). A high-resolution, absolute-dated Holocene and deglacial Asian monsoon record from Dongge Cave, China. Earth and Planetary Science Letters 233, 7186.CrossRefGoogle Scholar
Edgell, H.S., (1997). Karst and hydrogeology of Lebanon. Carbonates & Evaporites 12, (2) 1997220235.CrossRefGoogle Scholar
Edwards, R.L., Chen, J.H., Wasserburg, G.J., (1986). 238U–234U–230Th–232Th systematics and the precise measurement of time over the past 500.000 years. Earth and Planetary Science Letters 81, 175192.Google Scholar
Emeis, K.-C., Struck, U., Schulz, H.-M., Rosenberg, R., Bernasconi, S., Erlenkeuser, H., Sakamoto, T., Martinez-Ruiz, F., (2000). Temperature and salinity variations of the Mediterranean Sea surface waters over the last 16,000 years from records of planktonic stable oxygen isotopes and alkenone unsaturation ratios. Palaeogeography, Palaeoclimatology, Palaeoecology 158, 259280.Google Scholar
Enzel, Y., Amit, R., Dayan, U., Crouvi, O., Kahana, R., Ziv, B. and Sharon, D., in press. The climatic and physiographic controls of the eastern Mediterranean over the last Pleistocene climates in the southern Levant and its neighboring deserts.. Global and Planetary Change. doi:10.1016/j.gloplacha.2007.02.003.Google Scholar
Fairchild, I.J., Smith, C.L., Baker, A., Fuller, L., Spötl, C., Mattey, D., McDermott, F., E.I., M.F., (2006). Modification and preservation of environmental signals in speleothems. Earth Science Reviews 75, (1–4) 2006105153.CrossRefGoogle Scholar
Fleitmann, D., Burns, S.J., Mangini, A., Mudelsee, M., Kramers, J., Villa, I., Neff, U., Al-Subbary, A.A., Buettner, A., Hippler, D., Matter, A., (2007). Holocene ITCZ and Indian monsoon dynamics recorded in stalagmites from Oman and Yemen (Socotra). Quaternary Science Reviews 26, 170188.CrossRefGoogle Scholar
Fleitmann, D., Burns, S.J., Neff, U., Mangini, A., Matter, A., (2003). Changing moisture sources over the last 330,000 years in Northern Oman from fluid inclusion evidence in speleothems. Quaternary Research 60, 223232.CrossRefGoogle Scholar
Frumkin, A., Carmi, I., Zak, I., Magaritz, M., (1994). Middle Holocene environmental change determined from the salt Caves of Mount Sodom, Israel. Bar-Yosef, O., Kra, R.S. Late Quaternary Chronology and Plaeoclimates of the Eastern Mediterranean. Radiocarbon 315332.Google Scholar
Frumkin, A., Carmi, I., Gopher, A., Ford, D.C., Schwarcz, H.P., Tsuk, T., (1999a). A Holocene millenial-scale climatic cycle from a speleothem in Nahal Qanah Cave, Israel. The Holocene 9, (6) 1999677682.CrossRefGoogle Scholar
Frumkin, A., Ford, D.C., Schwarcz, H.P., (1999b). Continental oxygen isotopic record of the last 170,000 years in Jerusalem. Quaternary Research 51, 317327.CrossRefGoogle Scholar
Frumkin, A., Ford, D.C., Schwarcz, H.P., (2000). Paleoclimate and vegetation of the last glacial cycles in Jerusalem from a speleothem record. Global Biogeochemical cycles 14, (3) 2000863870.Google Scholar
Genty, D., Baker, A., Vokal, B., (2001). Intra- and inter-annual growth rate of modern stalagmites. Chemical Geology 176, 191212.CrossRefGoogle Scholar
Genty, D., Blamart, D., Ghaleb, B., Plagnes, V., Causse, C.h., Bakalowicz, M., Zouari, K., Chkir, N., Hellstrom, J., Wainer, K., Bourges, F., (2006). Timing and dynamics of the last deglaciation from European and North African d13C stalagmite profiles–comparison with Chinese and South Hemisphere stalagmites. Quaternary Science Reviews 25, 21182142.Google Scholar
Genty, D., Blamart, D., Ouahdi, R., Gilmour, M., Baker, A., Jouzel, J., Van-Exter, S., (2003). Precise dating of Dansgaard–Oeschger climate oscillations in western Europe from stalagmite data. Nature 421, 833837.Google Scholar
Goodfriend, G.A., (1999). Terrestrial stable isotope records of Late Quaternary paleoclimates in the Eastern Mediterranean region. Quaternary Science Reviews 18, 501513.Google Scholar
Gvirtzman, G., Wieder, M., (2001). Climate of the last 53,000 years in the Eastern Mediterranean, based on soil-sequence stratigraphy in the coastal plain of Israel. Quaternary Science Reviews 20, 18271849.Google Scholar
Hendy, C.H., (1971). The isotopic geochemistry of speleothems: I. The calculation of the effects of different modes of for indicators. Geochimica et Cosmochimica Acta 35, 801824.CrossRefGoogle Scholar
Hill, C., and Forti, P. (1997). Cave Minerals of the World. 2nd Edition 463, National Speleological Society, Huntsville, AL, USA., 109110.Google Scholar
Johnson, S., Clausen, H.B., Dansgaard, W., Gundestrup, N.S., Hansson, M., Johnsson, P., Steffensen, P., Sveinbjornsdottir, A.E., (1992). A deep ice core from east Greenland. Meddelelser om Gronland. Geoscience 29, 22 pp.Google Scholar
Kaufmann, G., Dreybrodt, W., (2004). Stalagmite growth and palaeo-climate: an inverse approach. Earth and Planetary Science Letters 224, 529545.Google Scholar
Karkabi, S. (1990). Al Ouat'Ouate N5: Special Jeita–Cinquantenaire de la spéléologie libanaise. Spéléo-Club du Liban, Lebanon.136.Google Scholar
Kolodny, Y., Stein, M., Machlus, M., (2005). Sea-rain-lake relation in the Last Glacial East Mediterranean revealed by d18O–d13C in lake Lisan aragonites. Geochimica et Cosmochimica Acta 69, 40454060.Google Scholar
McDermott, (2004). Palaeo-climate reconstruction from stable isotope variations in speleothems: a review. Quaternary Science Reviews 23, 901918.Google Scholar
McLaren, S.J., Gilbertson, D.D., Gratten, J.P., Hunt, C.O., Duller, G.A.T., Barker, G.A., (2004). Quaternary palaeogeomorphologic evolution of the Wadi Faynan area, southern Jordan. Palaeogeography, palaeoclimatology, palaeoecology 205, 131154.CrossRefGoogle Scholar
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. Quaternary Research 66, 421431.CrossRefGoogle Scholar
Nader, F.H., (2004). The Jiita Cave (Lebanon). Gunn, J. Encyclopedia of Caves and Karst Science Fitzroy Dearborn, New York–London.463464.Google Scholar
Nader, F.H., Swennen, R., Ellam, R., (2004). Stratabound dolomite versus volcanism-associated dolomite: an example from Jurassic platform carbonates in Lebanon. Sedimentology 51, (2) 339 360.Google Scholar
Nader, F.H., Swennen, R., Ottenburgs, R., (2003). Karst-meteoric dedolomitization in Jurassic carbonates, Lebanon. Geologica Belgica 6/1–2, 323.Google Scholar
Neff, U., Burns, S.J., Mangini, A., Mudelsee, M., Fleitmann, D., Matter, A., (2001). Strog coherence between solar variability and the monsoon in Oman between 9 and 6 kyr ago. Nature 411, 290293.CrossRefGoogle Scholar
Robinson, S.A., Black, S., Sellwood, B.W., Valdes, P.J., (2006). A review of palaeoclimates and palaeoenvironments in the Levant and Eastern Mediterranean from 25,000 to 5,000 years BP: setting the environmental background for the evolution of human civilization. Quaternary Science Reviews 25, 15171541.Google Scholar
Rossignol-Strick, M., (1995). Sea–land correlation of pollen records in the Eastern Mediterranean for the glacial–interglacial transition: biostratigraphy versus radiometric time-scale. Quaternary Science Reviews 14, 893915.Google Scholar
Rossignol-Strick, M., (1999). The Holocene climatic optimum and pollen records of sapropel 1 in the Eastern Mediterranean, 9000–6000 BP. Quaternary Science Reviews 18, 515530.Google Scholar
Shakun, J.D., Burns, S.J., Fleitman, D., Kramers, J., Matter, A., Al-Subary, A., (2007). A high-resolution, absolute-dated deglacial speleothem record of Indian Ocean climate from Socotra Island, Yemen. Earth and Planetary Science Letters 259, 442456.Google Scholar
Shen, C.-C., Edwards, R.L., Cheng, H., Dorale, J.A., Thomas, R.B., Moran, S.B., Weinstein, S., Edmonds, H.N., (2002). Uranium and thorium isotopic and concentration measurements by magnetic sector inductively coupled plasma mass spectrometry. Chemical Geology 185, (3–4) 165178.CrossRefGoogle Scholar
Staubwasser, M., Weiss, H., (2006). Holocene climate and cultural evolution in late prehistoric–early historic West Asia. Quaternary Research 66, 372387.CrossRefGoogle Scholar
United Nations Development Program (UNDP) (1970). Liban Etude Des Eaux Souterraines. Government of Lebanon, Beirut, Lebanon.Google Scholar
Vaks, A., Bar-Matthews, M., Ayalon, A., Schilman, B., Gilmour, M., Wawkesworth, C.J., Frumkin, A., Kaufman, A., Matthews, A., (2003). Palaeoclimate reconstruction based on the timing of speleothem growth and oxygen and carbon isotope composition in a cave located in the rain shadow in Israel. Quaternary Research 59, 182193.CrossRefGoogle Scholar
Verheyden, S., (2001). Speleothems as palaeoclimatic archives. Isotopic and geochemical study of the cave environment and its Late Quaternary records.. Unpublished Ph. D. Thesis, Vrije Universiteit Brussel, Belgium., 132p.Google Scholar
Verheyden, S., Baele, J.M., Keppens, E., Genty, D., Cattani, O., Cheng, H., Lawrence, E., Zhang, H., Van Strijdonck, M., Quinif, Y., (2006). The Proserpine stalagmite (Han-Sur-Lesse Cave, Belgium): preliminary environmental interpretation of the last 1000 years as recorded in a layered speleothem. Geologica Belgica 9, (3-4) 245256.Google Scholar
Wachter, E., Hayes, J.M., (1985). Exchange of oxygen isotopes in carbon-dioxide–phosphoric acid systems. Chemical Geology 52, 365374.Google Scholar
Walley, C.D., (2001). The Lebanon passive margin and the evolution of the Levantine Neotethys. Ziegler, P.A., Cavazza, W., Robertson, A.H.F., Crasquin-Soleau, S. Peri-Tethys Memoir 6: Peri-Tethyan Rift/Wrench Basins and Passive Margins, Mémoire du Muséum national d'Histoire naturelle, Paris 86, 407439.Google Scholar
Wang, Y.J., Cheng, H., Edwards, R.L., An, Z.S., Wu, J.Y., Shen, C.C., Dorale, J.A., (2001). A high-resolution absolute-dated Late Pleistocene monsoon record from Hulu cave, China. Science 294, 23452348.CrossRefGoogle ScholarPubMed