Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-19T07:48:56.970Z Has data issue: false hasContentIssue false

Modern and late Quaternary clay mineral distribution in the area of the SE Mediterranean Sea

Published online by Cambridge University Press:  20 January 2017

Abstract

The present-day clay mineral distribution in the southeastern Levantine Sea and its borderlands reveals a complex pattern of different sources and distribution paths. Smectite dominates the suspended load of the Nile River and of rivers in the Near East. Illite sources are dust-bearing winds from the Sahara and southwestern Europe. Kaolinite is prevalent in rivers of the Sinai, in Egyptian wadis, and in Saharan dust. A high-resolution sediment core from the southeastern Levantine Sea spanning the last 27 ka shows that all these sources contributed during the late Quaternary and that the Nile River played a very important role in the supply of clay. Nile influence was reduced during the glacial period but was higher during the African Humid Period. In contrast to the sharp beginning and end of the African Humid Period recorded in West African records (15 and 5.5 ka), our data show a more transitional pattern and slightly lower Nile River discharge rates not starting until 4 ka. The similarity of the smectite concentrations with fluctuations in sea-surface temperatures of the tropical western Indian Ocean indicates a close relationship between the Indian Ocean climate system and the discharge of the Nile River.

Type
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

Abram, N.J., Gagan, M.K., Liu, Z., Hantoro, W.S., McCulloch, M.T., and Suwargadi, B.W. Seasonal characteristics of the Indian Ocean Dipole during the Holocene epoch. Nature 445, (2007). 299302.Google Scholar
Adamson, D.A., Gasse, F., Street, F.A., and Williams, M.A.J. Late Quaternary history of the Nile. Nature 288, (1980). 5055.Google Scholar
Aksu, A., Yasar, D., and Mudie, P.J. Paleoclimatic and paleoceanographic conditions leading to development of Sapropel layer S1 in the Aegean Sea. Palaeogeography Palaeoclimatology Palaeoecology 116, (1995). 71101.Google Scholar
Ashok, K., Guan, Z., and Yamagata, T. Impact of the Indian Ocean Dipole on the relationship between the Indian monsoon rainfall and ENSO. Geophysical Research Letters 28, (2001). 44994502.CrossRefGoogle Scholar
Ashok, K., Guan, Z., Saji, N.H., and Yamagata, T. Individual and combined influences of ENSO and the Indian Ocean Dipole on the Indian Summer Monsoon. Journal of Climate 17, (2004). 31413155.2.0.CO;2>CrossRefGoogle Scholar
Babiker, M., and Gudmundsson, A. The effects of dykes and faults on groundwater flow in an arid land: the Red Sea Hills, Sudan. Journal of Hydrology 297, (2004). 256273.CrossRefGoogle Scholar
Bard, E., Rostek, F., and Sonzogni, C. Interhemispheric synchrony of the last deglaciation inferred from alkenone palaeothermometry. Nature 385, (1997). 707710.Google Scholar
Bar-Matthews, M., Ayalon, A., and Kaufman, A. Late Quaternary paleoclimate in the Eastern Mediterranean Region from Stable Isotope Analysis of Speleothems at Soreq Cave, Israel. Quaternary Research 47, (1997). 155168.Google Scholar
Bar-Matthews, M., Ayalon, A., Kaufman, A., and Wasserburg, G.J. The Eastern Mediterranean paleoclimate as a reflection of regional events: Soreq cave, Israel. Earth and Planetary Science Letters 166, (1999). 8595.Google Scholar
Bartov, Y., Goldstein, S.L., Stein, M., and Enzel, Y. Catastrophic arid episodes in the Eastern Mediterranean linked with the North Atlantic Heinrich events. Geology 31, (2003). 439442.2.0.CO;2>CrossRefGoogle Scholar
Beuning, K.R.M., Kelts, K., Russell, J., and Wolfe, B.B. Reassessment of Lake Victoria-Upper Nile River paleohydrology from oxygen isotope records of lake-sediment cellulose. Geology 30, (2002). 559562.2.0.CO;2>CrossRefGoogle Scholar
Beyth, M., Avigad, D., Wetzel, H.-U., Matthews, A., and Berhe, S.M. Crustal exhumation and indications for Snowball Earth in the East African Orogen: north Ethiopia and east Eritrea. Precambrian Research 123, (2003). 187201.Google Scholar
Biscaye, P.E. Mineralogy and sedimentation of recent deep sea clay in the Atlantic Ocean and adjacent seas and oceans. Geological Society of America Bulletin 76, (1965). 803832.Google Scholar
Bolle, M.-P., Tantawy, A.A., Pardo, A., Adatte, T., Burns, S., and Kassab, A. Climatic and environmental changes documented in the upper Paleocene to lower Eocene of Egypt. Ecologae Geologicae Helvetiae. 93, (2000). 3351.Google Scholar
Bond, G., Broecker, W.S., Johnsen, S.J., McManus, J.F., Labeyrie, L., Jouzel, J., and Bonani, G. Correlation between climate records from North Atlantic sediments and Greenland ice. Nature 365, (1993). 143147.Google Scholar
Bonnefille, R., and Chalie, F. Pollen-inferred precipitation time-series from equatorial mountains, Africa, the last 40 kyr BP. Global and Planetary Change 26, (2000). 2550.CrossRefGoogle Scholar
Bottema, S. The Younger Dryas in the Eastern Mediterranean. Quaternary Science Reviews 14, (1995). 883891.CrossRefGoogle Scholar
Cacho, I., Grimalt, J.O., Sierro, F.J., Shackleton, N., and Canals, M. Evidence for enhanced Mediterranean thermohaline circulation during rapid climatic coolings. Earth and Planetary Science Letters 183, (2000). 417429.Google Scholar
Camuffo, D. Controlling the aeolian erosion of the Great Sphinx. Studies in Conservation 38, (1993). 198205.Google Scholar
Caquineau, S., Gaudichet, A., Gomes, L., Magonthier, M.-C., and Chatenet, B. Saharan dust: clay ratio as a relevant tracer to assess the origin of soil-derived aerosols. Geophysical Research Letters 25, (1998). 983986.Google Scholar
Chester, R., Baxter, G.G., Behairy, A.K.A., Connor, K., Cross, D., Elderfield, H., and Padgham, R.C. Soil-sized eolian dust from the lower troposphere of the Eastern Mediterranean Sea. Marine Geology 24, (1977). 201217.Google Scholar
Claussen, M., Kubatzki, C., Brovkin, V., Ganopolski, A., Hoelzmann, P., and Pachur, H.-J. Simulation of an abrupt change in Saharan vegetation in the mid-Holocene. Geophysical Research Letters 26, (1999). 20372040.CrossRefGoogle Scholar
Cosca, M.A., Shimron, A., and Caby, R. Late Precambrian metamorphism and cooling in the Arabian-Nubian Shield: petrology and 40Ar/39Ar geochronology of metamorphic rocks of the Elat area (southern Israel). Precambrian Research 98, (1999). 107127.Google Scholar
Coudé-Gaussen, G., (1981). Etude détaillée d'un échantillon de poussières éolinnes prélevé au Tanezrouft, le 10 décembre 1980. Recherches Géographiques à Strasbourg 16/17, 121130.Google Scholar
Coudé-Gaussen, G., and Blanc, P. Présence de grains éolisés de palygorskite dans les poussières actuelles et les sédiments rècents d'origine désertique. Bulletin de la Société Géologique de France 8, (1985). 283292.Google Scholar
deMenocal, P., Ortiz, J., Guilderson, T., Adkins, J., Sarnthein, M., Baker, L., and Yarusinsky, M. Abrupt onset and termination of the African Humid Period: rapid climate responses to gradual insolation forcing. Quaternary Science Reviews 19, (2000). 347361.Google Scholar
deMenocal, P., Ortiz, J., Guilderson, T., and Sarnthein, M. Coherent high- and low-latitude climate variability during the Holocene warm period. Science 288, (2000). 21982202.CrossRefGoogle ScholarPubMed
Ehrmann, W., Schmiedl, G., Hamann, Y., and Kuhnt, T. Distribution of clay minerals in surface sediments of the Aegean Sea: a compilation. International Journal of Earth Sciences 96, (2007). 769780.CrossRefGoogle Scholar
Ehrmann, W., Schmiedl, G., Hamann, Y., Kuhnt, T., Hemleben, C., and Siebel, W. Clay minerals in late glacial and Holocene sediments of the northern and southern Aegean Sea. Palaeogeography Palaeoclimatology Palaeoecology. 249, (2007). 3657.Google Scholar
El-Sayed, M.M. Geochemistry and petrogenesis of the post-orogenic bimodal dyke swarms in NW Sinai, Egypt: constraints on the magmatic-tectonic processes during the late Precambrian. Chemie der Erde 66, (2006). 129141.Google Scholar
El-Sherbini, M.I., and Issa, G.I. Composition and origin of some calcrete deposits in south Western Desert of Egypt. Journal of African Earth Sciences 9, (1989). 461466.Google Scholar
Fairbanks, R.G., Mortlock, R.A., Chiu, T.-C., Cao, L., Kaplan, A., Guilderson, T.P., Fairbanks, T.W., Bloom, A.L., Grootes, P.M., and Nadeau, M.-J. Radiocarbon calibration curve spanning 0 to 50,000 years BP based on paired 230Th/234U/238U and 14C dates on pristin corals. Quaternary Science Reviews 24, (2005). 17811796.Google Scholar
Fleitmann, D., Burns, S.J., Mudelsee, M., Neff, U., Kramers, J., Mangini, A., and Matter, A. Holocene forcing of the Indian monsoon recorded in a stalagmite from Southern Oman. Science 300, (2003). 17371739.Google Scholar
Foucault, A., and Mélières, F. Palaeoclimatic cyclicity in central Mediterranean Pliocene sediments: the mineralogical signal. Palaeogeography Palaeoclimatology Palaeoecology 158, (2000). 311323.CrossRefGoogle Scholar
Giannini, A., Saravanan, R., and Chang, P. Dynamics of the boreal summer African monsoon in the NSIPP1 atmospheric model. Climate Dynamics 25, (2005). 517535.Google Scholar
Ganor, E., and Mamane, Y. Transport of Saharan dust across the Eastern Mediterranean. Atmospheric Environment 16, (1982). 581587.Google Scholar
Garzanti, E., Ando, S., Vezzoli, G., Abdel Megid, A.A., and El Kammar, A. Petrology of Nile River sands (Ethiopia and Sudan): sediment budgets and erosion patterns. Earth and Planetary Science Letters 252, (2006). 327341.Google Scholar
Gasse, F. Hydrological changes in the African tropics since the Last Glacial Maximum. Quaternary Science Reviews 19, (2000). 189211.Google Scholar
Gevrek, A.I., and Kazanci, N. A Pleistocene, pyroclastic-poor maar from central Anatolia, Turkey: influence of a local fault on a phreatomagmatic eruption. Journal of Volcanology and Geothermal Research 95, (2000). 309317.Google Scholar
Goudie, A.S., and Middleton, N.J. Saharan dust storms: nature and consequences. Earth-Science Reviews 56, (2001). 179204.CrossRefGoogle Scholar
Grindy, A.R., and Tamish, M.M.O. Some major and trace constituents of Phanerozoic Egyptian mudrocks and marls. Journal of African Earth Sciences 3, (1985). 303320.Google Scholar
Guerzoni, S., Chester, R., Dulac, F., Herut, B., Loye-Pilot, M.-D., Measures, C., Migon, C., Molinaroli, E., Moulin, C., Rossini, P., Saydam, C., Soudine, A., and Ziveri, P. The role of atmospheric deposition in the biochemistry of the Mediterranean Sea. Progress in Oceanography 44, (1999). 147190.Google Scholar
Gur, D., Steinitz, G., Kolodny, Y., Starinsky, A., and McWilliams, M. 40Ar/39Ar dating of combustion metamorphism (“Mottled Zone”, Israel). Chemical Geology 122, (1995). 171184.Google Scholar
Gvirtzman, G., and Wieder, M. 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, (2001). 18271849.Google Scholar
Hamann, Y., Ehrmann, W., Schmiedl, G., Krüger, S., Stuut, J.-B.W., and Kuhnt, T. Sedimentation processes in the Eastern Mediterranean Sea during the Late Glacial and Holocene revealed by end-member modelling of the terrigenous fraction in marine sediments. Marine Geology 248, (2008). 97114.CrossRefGoogle Scholar
Haug, G.H., Hughen, K.A., Sigman, D.M., Peterson, L.C., and Röhl, U. Southward migration of the Intertropical Convergence Zone through the Holocene. Science 293, (2001). 13041308.CrossRefGoogle ScholarPubMed
Heller-Kallai, L., and Kalman, Z.H. Some naturally occurring illite-smectite interstratifications. Clay Clay Minerals 20, (1972). 165168.Google Scholar
Heller-Kallai, L., Nathan, Y., and Zak, I. Clay mineralogy of Triassic sediments in southern Israel and Sinai. Sedimentology 20, (1973). 513521.Google Scholar
Hemleben, C., (2002). Short Cruise Report, R.V. Meteor Cruise 51. Leg 3 Valletta-Malta to Istanbul-Turkey 14.11.-10.12.2001.Google Scholar
Holeman, J.N. The sediment yield of major rivers of the world. Water Resources Research 4, (1968). 737747.CrossRefGoogle Scholar
Hurrell, J.W., Kushnir, Y., Ottersen, G., and Visbeck, M. An overview of the North Atlantic Oscillation. Geophysical Monograph 134, (2003). 135.Google Scholar
Ibrahim, K.M., Tarawneh, K., and Rabba, I. Phases of activity and geochemistry of basaltic dike systems in northeast Jordan parallel to the Red Sea. Journal of Asian Earth Sciences 21, (2003). 467472.Google Scholar
Innocenti, F., Agostini, S., Di Vincenzo, G., Doglioni, C., Manetti, P., Savascin, M.Y., and Tonarini, S. Neogene and Quaternary volcanism in Western Anatolia: magma sources and geodynamic evolution. Marine Geology 221, (2005). 397421.CrossRefGoogle Scholar
Janicot, S., Harzallah, A., Fontaine, B., and Moron, V. West African monsoon dynamics and eastern equatorial Atlantic and Pacific SST Anomalies (1970–88). Journal of Climate 11, (1997). 18741882.Google Scholar
Joseph, J.H., Manes, A., and Ashbel, D. Desert aerosols transported by Khamsinic depressions and their climatic effects. Journal of Applied Meterology 12, (1973). 792797.2.0.CO;2>CrossRefGoogle Scholar
Karakas, Z., and Kadir, S. Devitrification of volcanic glasses in Konya volcanic unit, Turkey. Turkish Journal of Earth Sciences 9, (2000). 3946.Google Scholar
Katz, O., Beyth, M., Miller, N., Stern, R., Avigad, D., Basu, A., and Anbar, A. A Late Neoproterozoic (∼630 Ma) high-magnesium andesite suite from southern Israel: implications for the consolidation of Gondwanaland. Earth and Planetary Science Letters 218, (2004). 475490.CrossRefGoogle Scholar
Klein, B., Roether, W., Manca, B.B., Bregant, D., Beitzel, V., Kovacevic, V., and Luchetta, A. The large deep water transient in the Eastern Mediterranean. Deep-Sea Research I 46, (1999). 371414.Google Scholar
Kohfeld, K.E., and Harrison, S.P. DIRTMAP: the geological record of dust. Earth-Science Reviews 54, (2001). 81114.CrossRefGoogle Scholar
Krom, M.D., Cliff, R.A., Eijsink, L.M., Herut, B., and Chester, R. The characterisation of Saharan dusts and Nile particulate matter in surface sediments from the Levantine basin using Sr isotopes. Marine Geology 155, (1999). 319330.Google Scholar
Kröpelin, S., Verschuren, D., Lezine, A.M., Eggermont, H., Cocquyt, C., Francus, P., Cazet, J.P., Fagot, M., Rumes, B., Russell, J.M., Darius, F., Conley, D.J., Schuster, M., von Suchodoletz, H., and Engstrom, D.R. Climate-driven ecosystem succession in the Sahara: the past 6000 years. Science 320, (2008). 765768.Google Scholar
Kuhlmann, H., Meggers, H., Freudenthal, T., and Wefer, G. The transition of the monsoonal and N Atlantic climate system off NW Africa during the Holocene. Geophysical Research Letters 31, (2004). 14.CrossRefGoogle Scholar
Kuhnt, T., Schmiedl, G., Ehrmann, W., Hamann, Y., and Anderson, A. Stable isotopic composition of Holocene benthic foraminifers from the Eastern Mediterranean Sea: past changes in productivity and deep water oxygenation. Palaeogeography Palaeoclimatology Palaeoecology 268, (2008). 106115.Google Scholar
Kusky, T., and El-Baz, F. Neotectonics and fluvial geomorphology of the northern Sinai Peninsula. Journal of African Earth Sciences 31, (2000). 213235.Google Scholar
Lado, M., and Ben-Hur, M. Soil mineralogy effects on seal formation, runoff and soil loss. Applied Clay Science 248, (2004). 209224.Google Scholar
Lamb, H.F., Bates, C.R., Coombes, P.V., Marshall, M.H., Umer, M., Davies, S.J., and Dejen, E. Late Pleistocene desiccation of Lake Tana, source of the Blue Nile. Quaternary Science Reviews 26, (2007). 287299.Google Scholar
Landmann, G., Abu Qudaira, G.H., Shawabkeh, K., Wrede, V., and Kempe, S. Geochemistry of the Lisan and Damya Formation in Jordan, and implications for palaeoclimate. Quaternary International 89, (2002). 4557.Google Scholar
Malanotte-Rizzoli, P., and Hecht, A. Large-scale properties of the Eastern Mediterranean: a review. Oceanologica Acta 11, (1988). 323335.Google Scholar
Maldonado, A., and Stanley, D.J. Clay mineral distribution patterns as influenced by depositional processes in the Southeastern Levantine Sea. Sedimentology 28, (1981). 2132.CrossRefGoogle Scholar
Mayewski, P.A., Rohling, E.J., Stager, J.C., Karlén, W., Maasch, K.A., Meeker, L.D., Meyerson, E.A., Gasse, F., van Kreveld, S., Holmgren, K., Lee-Thorp, J., Rosqvist, G., Rack, F., Staubwasser, M., Schneider, R.R., and Steig, E.J. Holocene climate variability. Quaternary Research 62, (2004). 243255.Google Scholar
Milliman, J.D., and Syvitski, J.P.M. Geomorphic/tectonic control of sediment discharge to the ocean: the importance of small mountainous rivers. Journal of Geology 100, (1992). 525544.Google Scholar
Moreno, A., Cacho, I., Canals, M., Prins, M.A., Sanchez-Goni, M.-F., Grimalt, J.O., and Weltje, G.J. Saharan dust transport and high-latitude glacial climatic variability: the Alboran Sea record. Quaternary Research 58, (2002). 318328.Google Scholar
Neev, D., and Emery, K.O. The Destruction of Sodom, Gomorrah, and Jericho: Geological, Climatological, and Archaeological Background. (1995). Oxford University Press, New York. 175 ppGoogle Scholar
Paquet, H., Coude-Gaussen, G., and Rognon, P. Etude mineralogique de poussieres le long d'un itineraire entre 19° et 35° de latitude nord. Revue de Geologie Dynamique et de Geographie Physique 25, (1984). 257265.Google Scholar
Pickard, G.L., and Emery, W.J. Descriptive Physical Oceanography — An Introduction. (1982). Pergamon Press, San Diego. 249 ppGoogle Scholar
Pinardi, N., and Masetti, E. Variability of the large scale general circulation of the Mediterranean Sea from observations and modelling: a review. Palaeogeography Palaeoclimatology Palaeoecology 158, (2000). 153174.Google Scholar
Pye, K. Aeolian dust transport and deposition over Crete and adjacent parts of the Mediterranean Sea. Earth Surface Processes and Landforms 17, (1992). 271288.Google Scholar
Refaat, A.A., and Imam, M.M. The Tayiba Red Beds: transitional marine-continental deposits in the precursor Suez Rift, Sinai, Egypt. Journal of African Earth Sciences 28, (1999). 487506.Google Scholar
Renssen, H., Brovkin, V., Fichefet, T., and Goosse, H. Simulation of the Holocene climate evolution in Northern Africa: the termination of the African Humid Period. Quaternary International 150, (2006). 95102.Google Scholar
Rimmelé, G., Oberhänsli, R., Goffé, B., Jolivet, L., Candan, O., and Cetinkaplan, M. First evidence of high-pressure metamorphism in the “Cover Series” of the southern Menderes Massif. Tectonic and metamorphic implications for the evolution of SW Turkey. Lithos 71, (2003). 1946.Google Scholar
Robinson, S.A., Black, S., Sellwood, B.W., and Valdes, P.J. A review of palaeoclimates and palaeoenvironments in the Levant and Eastern Mediterranean from 25,000 to 5000 years BP: setting the environmental background for the evolution of human civilisation. Quaternary Science Reviews 25, (2006). 15171541.Google Scholar
Rosenthal, E., Jones, B.F., and Weinberger, G. The chemical evolution of Kurnub Group paleowater in the Sinai-Negev province—a mass balance approach. Applied Geochemistry 13, (1998). 553569.Google Scholar
Rossignol-Strick, M. Sea–land correlation of pollen records in the Eastern Mediterranean for the glacial–interglacial transition: biostratigraphy versus radiometric time-scale. Quaternary Science Reviews 14, (1995). 893915.CrossRefGoogle Scholar
Rossignol-Strick, M. The Holocene climatic optimum and pollen records of sapropel 1 in the eastern Mediterranean, 9000–6000 BP. Quaternary Science Reviews 18, (1999). 515530.Google Scholar
Rossignol-Strick, M., Nesteroff, W., Olive, P., and Vergnaud-Grazzini, C. After the deluge: Mediterranean stagnation and sapropel formation. Nature 295, (1982). 105110.Google Scholar
Saaroni, H., Ziv, B., Bitan, A., and Alpert, P. Easterly wind storms over Israel. Theoretical and Applied Climatology 59, (1998). 6177.Google Scholar
Saji, N.H., and Yamagata, T. Indian ocean dipole mode events and African rainfall variability. CLIVAR Exchanges 27, (2003). 14.Google Scholar
Saji, N.H., Goswami, B.N., Vinayachandran, P.N., and Yamagata, T. A dipole mode in the tropical Indian Ocean. Nature 401, (1999). 360363.Google Scholar
Sandler, A., and Herut, B. Composition of clays along the continental shelf off Israel: contribution of the Nile versus local sources. Marine Geology 167, (2000). 339354.Google Scholar
Schilman, B., Almogi-Labin, A., Bar-Matthews, M., Labeyrie, L., Paterne, M., and Luz, B. Long- and short-term carbon fluctuations in the Eastern Mediterranean during the late Holocene. Geology 29, (2001). 10991102.Google Scholar
Schott, F.A., McCreary, J.P. Jr. The monsoon circulation of the Indian Ocean. Progress in Oceanography 51, (2001). 1123.Google Scholar
Segev, A., Weissbrod, T., and Lang, B. 40Ar/39Ar dating of the Aptian-Albian igneous rocks in Makhtesh Ramon (Negev, Israel) and its stratigraphic implications. Cretaceous Research 26, (2005). 633656.Google Scholar
Shaw, J.E., Baker, J.A., Menzies, M.A., Thirlwall, M.F., and Ibrahim, K.M. Petrogenesis of the largest intraplate volcanic field on the Arabian Plate (Jordan): a mixed lithosphere–asthenosphere source activated by lithospheric extension. Journal of Petrology 44, (2003). 16571679.CrossRefGoogle Scholar
Siani, G., Paterne, M., Michel, E., Sulpizio, R., Sbrana, A., Arnold, M., and Haddad, G. Mediterranean Sea surface radiocarbon reservoir age changes since the last glacial maximum. Science 294, (2001). 19171920.Google Scholar
Singer, A. Illite in aridic soils, desert dusts and desert loess. Sedimentary Geology 59, (1988). 251259.Google Scholar
Sirocko, F., Garbe-Schönberg, D., McIntyre, A., and Molfino, B. Teleconnections between the subtropical monsoons and high-latitude climates during the last deglaciation. Science 272, (1996). 526529.Google Scholar
Stanley, D.J., and Galili, E. Sediment dispersal along northern Israel coast during the early Holocene: geological and archaeological evidence. Marine Geology 130, (1996). 1117.Google Scholar
Stanley, D.J., and Wingerath, J.G. Clay mineral distributions to interpret Nile cell provenance and dispersal: I. Lower River Nile to delta sector. Journal of Coastal Research 12, (1996). 911929.Google Scholar
Stanley, D.J., Mart, Y., and Nir, Y. Clay mineral distributions to interpret Nile cell provenance and dispersal: II. Coastal plain from Nile delta to Northern Israel. Journal of Coastal Research 13, (1997). 506533.Google Scholar
Stuut, J.-B.W., and Lamy, F. Climate variability at the southern boundaries of the Namib (southwestern Africa) and Atacama (northern Chile) coastal deserts during the last 120,000 yr. Quaternary Research 62, (2004). 301309.Google Scholar
Talbot, M.R., and Laerdal, T. The Late Pleistocene–Holocene palaeolimnology of Lake Victoria, East Africa, based upon elemental and isotopic analyses of sedimentary organic matter. Journal of Paleolimnology 23, (2000). 141164.Google Scholar
Tierney, J.E., Russell, J., Huang, Y., Sinninghe Damste, J.S., Hopmans, E.C., and Cohan, A.S. Northern Hemisphere controls on tropical southeast African climate during the past 60,000 years. Science 322, (2008). 252255.Google Scholar
Tütken, T., Eisenhauer, A., Wiegand, B., and Hansen, B.T. Glacial–interglacial cycles in Sr and Nd isotopic composition of Arctic marine sediments triggered by the Svalbard/Barents Sea ice sheet. Marine Geology 182, (2002). 351372.Google Scholar
Ukstins, I.A., Renne, P.R., Wolfenden, E., Baker, J.A., Ayalew, D., and Menzeis, M. Matching conjugate volcanic rifted margins: 40Ar/39Ar chrono-stratigraphy of pre- and syn-rift bimodal flood volcanism in Ethiopia and Yemen. Earth and Planetary Science Letters 198, (2002). 289306.Google Scholar
Venkatarathnam, K., and Ryan, W.B.F. Dispersal patterns of clay minerals in the sediments of the eastern Mediterranean Sea. Marine Geology 11, (1971). 261282.Google Scholar
Vidal, L., Labeyrie, L., Cortijo, E., Arnold, M., Duplessy, J.C., Michel, E., Becqué, S., and van Weering, T.C.E. Evidence for changes in the North Atlantic Deep Water linked to meltwater surges during the Heinrich events. Earth and Planetary Science Letters 146, (1997). 1327.Google Scholar
Wanas, H.A., and Soliman, H.E. Allogenic and authigenic clays of the Lower Palaeozoic sandstones of the Naqus Formation at Gebel Gunna, central Sinai, Egypt: their recognition and geological significance. Journal of African Earth Sciences 32, (2001). 4760.Google Scholar
Webster, P.J., Moore, A.M., Loschnigg, J.P., and Leben, R.R. Coupled ocean–atmosphere dynamics in the Indian Ocean during 1997–98. Nature 401, (1999). 356360.Google Scholar
Weldeab, S., Emeis, K.-C., Hemleben, C., and Siebel, W. Provenance of lithogenic surface sediments and pathways of riverine suspended matter in the Eastern Mediterranean Sea: evidence from 143Nd/144Nd and 87Sr/96Sr ratios. Chemical Geology 186, (2002). 139149.Google Scholar
Weldeab, S., Emeis, K.-C., Hemleben, C., Vennemann, T.W., and Schulz, H. Sr and Nd isotope composition of Late Pleistocene sapropels and nonsapropelic sediments from the Eastern Mediterranean Sea: implications for detrital influx and climatic conditions in the source areas. Geochimica et Cosmochimica Acta 66, (2002). 35853598.Google Scholar
Williams, M., Talbot, M.R., Aharon, P., Salaam, Y.A., Williams, F., and Brendeland, N.I. Abrupt return of the summer monsoon 15,000 years ago: new supporting evidence from the lower White Nile valley and Lake Albert. Quaternary Science Reviews 25, (2006). 26512665.Google Scholar
Wüst, G. Die Tiefenzirkulation des Mittelländischen Meeres in den Kernschichten des Zwischen- und des Tiefenwassers. Deutsche Hydrographische Zeitschrift 13, (1960). 105131.Google Scholar
Wüst, G. On the vertical circulation of the Mediterranean Sea. Journal of Geophysical Research 66, (1961). 32613271.Google Scholar