Skip to main content Accessibility help
×
Hostname: page-component-77c89778f8-gvh9x Total loading time: 0 Render date: 2024-07-18T21:19:11.706Z Has data issue: false hasContentIssue false

References

Published online by Cambridge University Press:  14 December 2018

Martin Williams
Affiliation:
University of Adelaide
Get access

Summary

Image of the first page of this content. For PDF version, please use the ‘Save PDF’ preceeding this image.'
Type
Chapter
Information
The Nile Basin
Quaternary Geology, Geomorphology and Prehistoric Environments
, pp. 334 - 393
Publisher: Cambridge University Press
Print publication year: 2019

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

Abbate, E., Albianelli, A., Awad, A., et al. (2010). Pleistocene environments and human presence in the middle Atbara valley (Khashm El Girba), Eastern Sudan). Palaeogeography, Palaeoclimatology, Palaeoecology, 292, 1234.CrossRefGoogle Scholar
Abbate, E., Bruni, P., Ferretti, M. P., et al. (2014). The East African Oligocene intertrappean beds: Regional distribution, depositional environments and Afro/Arabian mammal dispersals. Journal of African Earth Sciences, 99, 463489.Google Scholar
Abdelsalam, M. G. (2018). The Nile’s journey through space and time: A geological perspective. Earth-Science Reviews, 177, 742773.CrossRefGoogle Scholar
Abdel Salam, Y. (1966). The ground water geology of the Gezira. Unpublished MSc thesis, University of Khartoum.Google Scholar
Abell, P. I. & Hoelzmann, P. (2000). Holocene palaeoclimates in northwestern Sudan: Stable isotope studies on molluscs. Global and Planetary Change, 26, 112.Google Scholar
Abell, P. I., Hoelzmann, P. & Pachur, H.-J. (1996). Stable isotope ratios of gastropod shells and carbonate sediments of NW Sudan as palaeoclimatic indicators. Palaeoecology of Africa, 24, 3352.Google Scholar
Abell, P. I. & Plug, I. (2000). The Pleistocene/Holocene transition in South Africa: Evidence for the Younger Dryas event. Global and Planetary Change, 26, 173179.Google Scholar
Abell, P. I. & Williams, M. A. J. (1989). Oxygen and carbon isotope ratios in gastropod shells as indicators of palaeoenvironments in the Afar region of Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology, 74, 265278.Google Scholar
Abuzied, Hussein Mohammed, Ahmed (2009). Mapping and assessment of wind erosion in central Northern State, Sudan. Unpublished PhD thesis, Desertification and Desert Cultivation Studies Institute, University of Khartoum.Google Scholar
Adamson, D. A., Clark, J. D. & Williams, M. A. J. (1974). Barbed bone points from Central Sudan and the age of the ‘Early Khartoum’ tradition. Nature, 249, 120123.CrossRefGoogle Scholar
Adamson, D. A., Clark, J. D. & Williams, M. A. J. (1987). Pottery tempered with sponge from the White Nile, Sudan. African Archaeological Review, 5, 115127.CrossRefGoogle Scholar
Adamson, D. A., Gasse, F., Street, F. A. & Williams, M. A. J. (1980). Late Quaternary history of the Nile. Nature, 287, 5055.Google Scholar
Adamson, D. A., Gillespie, R. & Williams, M. A. J. (1982). Palaeogeography of the Gezira and of the lower Blue and White Nile Valleys. In Williams, M. A. J. & Adamson, D. A. (eds.), A Land between Two Niles: Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema, pp. 165219.Google Scholar
Adamson, D., McEvedy, R. & Williams, M. A. J. (1993). Tectonic inheritance in the Nile basin and adjacent areas. Israel Journal of Earth Sciences, 41, 7585.Google Scholar
Adamson, D. & Williams, F. (1980). Structural geology, tectonics and the control of drainage in the Nile basin. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and The Nile. Rotterdam: A. A. Balkema, pp. 225252.Google Scholar
Adamson, D. A. & Williams, M. A. J. (1987). Geological setting of Pliocene rifting and deposition in the Afar Depression of Ethiopia. Journal of Human Evolution, 16, 597610.CrossRefGoogle Scholar
Addison, F. (1949). The Wellcome Excavations in the Sudan, Vol. I: Jebel Moya. Oxford: Oxford University Press, Appendices pp. 261399.Google Scholar
Addison, F. (1956). Second thoughts on Jebel Moya. Kush, 4, 41.Google Scholar
Affek, H.P., Bar-Matthews, M., Ayalon, A., Matthews, A. & Eiler, J.M. (2008). Glacial/interglacial temperature variations in Soreq cave speleothems as recorded by ‘clumped isotope’ thermometry. Geochimica et Cosmochimica Acta, 72, 53515360.CrossRefGoogle Scholar
Alimen, H., Beucher, F. & Conrad, G. (1966). Chronologie du dernier cycle Pluviale-Aride au Sahara nord-occidental. Comptes Rendus, Académie des Sciences, Paris, 263D, 58.Google Scholar
Alimen, H. & Chavaillon, J. (1963). ‘Deltas intérieurs’ d’un cours d’eau désertique. Scientia 6ème série, 57, 18.Google Scholar
Allen, J. & O’Connell, J. F. (2003). The long and the short of it: Archaeological approaches to determining when humans first colonised Australia and New Guinea. Australian Archaeology, 57, 519.CrossRefGoogle Scholar
Almásy, L. (1936). Récentes Explorations dans le Désert Libyque (1932–1936). Le Caire, Société Royale de Géographie de l’Egypte.Google Scholar
Almásy, L. E. de (1939). Unbekannte Sahara. Leipzig: Brockhaus.Google Scholar
Almeida, Manoel de (1628–1646). The history of High Ethiopia or Abassia. In Some Records of Ethiopia 1593–1646 Together with Bahrey’s History of the Galla, translated and edited by C. F. Beckingham & G. W. B. Huntingford. London: Hakluyt Society (1954).Google Scholar
Almond, D. C. (1986). Geological evolution of the Afro-Arabian dome. Tectonophysics, 131, 301332.Google Scholar
Ambrose, S. H. (1998). Late Pleistocene human population bottlenecks, volcanic winter, and differentiation of modern humans. Journal of Human Evolution, 34, 623651.Google Scholar
Amit, R., Enzel, Y., Crouvi, O., et al. (2009). The role of the Nile in initiating a massive dust influx to the Negev late in the middle Pleistocene. Geological Society of America Bulletin, 123, 873889.CrossRefGoogle Scholar
Amit, R., Enzel, Y., Grodek, T., Crouvi, O., Porat, N. & Ayalon, A. (2010). The role of rare rainstorms in the formation of calcic soil horizons on alluvial surfaces in extreme deserts. Quaternary Research, 74, 177187.Google Scholar
Amit, R., Lekach, J., Ayalon, A. & Porat, N. (2007). New insights into pedogenic processes in extremely arid environments and their paleoclimatic implications: The Negev Desert, Israel. Quaternary International, 162–163, 6175.Google Scholar
Anderson, D. M., Overpeck, J. P. & Gupta, A. K. (2002). Increase in the Asian southwest monsoon during the past four centuries. Science, 297, 596599.Google Scholar
Andrew, G. (1948). Geology of the Sudan. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 84128.Google Scholar
Andrew, G. & Karkanis, G. Y. (1945). Stratigraphical notes, Anglo-Egyptian Sudan. Sudan Notes and Records, 26, 157166.Google Scholar
Andrews, F. W. (1948). The vegetation of the Sudan. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 3261.Google Scholar
Anhuf, D., Ledru, M.-P., Behling, H., et al. (2006). Paleo-environmental change in the Amazonian and African rainforest during the LGM. Palaeogeography, Palaeoclimatology, Palaeoecology, 239, 510527.Google Scholar
Appelo, C. A. J. (1990). Ground water in the Nile delta. Nature, 346, 417418.CrossRefGoogle Scholar
Aref, M. A. M. (2003). Classification and depositional environments of Quaternary pedogenic gypsum crusts (gypcrete) from east of the Fayum Depression, Egypt. Sedimentary Geology, 155, 87108.CrossRefGoogle Scholar
Arkell, A. J. (1945). Some land and freshwater snails of the western Sudan. Sudan Notes and Records, 26, 339341.Google Scholar
Arkell, A. J. (1948). The historical background of Sudan agriculture. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 929.Google Scholar
Arkell, A. J. (1949a, reprinted 1963). The Old Stone Age in the Anglo-Egyptian Sudan. Occasional Papers No. 1. Khartoum: Sudan Antiquities Service.Google Scholar
Arkell, A. J. (1949b). Early Khartoum. London: Oxford University Press.Google Scholar
Arkell, A. J. (1953). Shaheinab. London: Oxford University Press.Google Scholar
Arkell, A. J. (1961, 2nd edn.). A History of the Sudan from the Earliest Times to 1821. London: Athlone Press.Google Scholar
Arkell, A. J. (1964). Wanyanga and an Archaeological Reconnaisaance of the South-West Libyan Desert: The British Ennedi Expedition. 1957. London: Oxford University Press.Google Scholar
Arkell, A. J. (1975). The Prehistory of the Nile Valley. Leiden: Brill.Google Scholar
Armitage, S. J., Drake, N. A., Stokes, S., et al. (2007). Multiple phases of North African humidity recorded in lacustrine sediments from the Fazzan Basin, Libyan Sahara. Quaternary Geochronology, 2, 181186.Google Scholar
Armitage, S. J., Jasim, S. A., Marks, A. E., et al. (2011). The southern route ‘Out of Africa’: Evidence for an early expansion of modern humans into Arabia. Science, 331, 453556.CrossRefGoogle ScholarPubMed
Arz, H. W., Lamy, F., Pätzold, J., Müller, P. J. & Prins, M. (2003). Mediterranean moisture source for an Early-Holocene humid period in the northern Red Sea. Science, 300, 18121.Google Scholar
Asrat, A., Barbey, B. & Gleizer, G. (2001). The Precambrian geology of Ethiopia: A review. Africa Geoscience Review, 8, 271288.Google Scholar
Atherton, M. P., Lagha, S. & Derek, F. (1978). The geochemistry and petrology of the Jabal Archenu and Jabal al ‘Awaynat alkaline ring complexes of SE Libya. In Salem, M. J. & Busweril, M. T. (eds.), The Geology of Libya, Second Symposium on the Geology of Libya, Tripoli, September 16–21. London: Academic Press, pp. 25592576.Google Scholar
Atlas of Africa. (1973). Paris: Éditions Jeune Afrique.Google Scholar
Avni, Y. (2005). Gully incision as a key factor in desertification in an arid environment, the Negev highlands, Israel. Catena, 63, 185220.CrossRefGoogle Scholar
Avni, Y. (2017). Tectonic and physiographic settings of the Levant. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 316.Google Scholar
Avni, Y., Porat, N., Plakht, J. & Avni, G. (2006). Geomorphic changes leading to natural desertification versus anthropogenic land conservation in an arid environment, the Negev Highlands, Israel. Geomorphology, 82, 177200.CrossRefGoogle Scholar
Avni, Y., Segev, A. & Ginat, H. (2012). Oligocene regional denudation of the northern Afar dome: Pre- and syn-breakup stages in the Afro-Arabian plate. Geological Society of America Bulletin, 124, 1871–1897.CrossRefGoogle Scholar
Avni, Y., Zhang, J. F., Shelach, G. & Zhou, L. P. (2010). Upper Pleistocene-Holocene geomorphic changes dictating sedimentation rates and historical land use in the valley system of the Chifeng region, Inner Mongolia, northern China. Earth Surface Processes and Landforms, 35(11), 12511268.Google Scholar
Ayliffe, D., Williams, M. A. J. & Sheldon, F. (1996). Stable carbon and oxygen isotopic composition of early-Holocene gastropods from Wadi Mansurab, north-central Sudan. The Holocene, 6, 157169.CrossRefGoogle Scholar
Ayoub, A. T. (1999). Land degradation, rainfall variability and food production in the Sahelian zone of the Sudan. Land Degradation and Development, 10, 489500.Google Scholar
Bacon, G. H. (1948). Crops of the Sudan. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 302400.Google Scholar
Bae, C. J., Douka, K. & Petraglia, M. D. (2017a). Human colonization of Asia in the Late Pleistocene: An introduction to Supplement 17. Current Anthropology, 58, (Supplement 17),S373S382.CrossRefGoogle Scholar
Bae, C. J., Douka, K. & Petraglia, M. D. (2017b). On the origin of modern humans: Asian perspectives. Science, 358: eaai9067.Google Scholar
Bagnold, R. A. (1933). A further journey to the south Libyan Desert. Geographical Journal, 82, 103129.Google Scholar
Bagnold, R. A. (1935). Libyan Sands: Travel in a Dead World. London: Immel (1987).Google Scholar
Bagnold, R. A. (1941). The Physics of Blown Sand and Desert Dunes. London: Methuen.Google Scholar
Bagnold, R. A. (1966). An approach to the sediment transport problem from general physics. United States Geological Survey Professional Paper, 422-I, 137.Google Scholar
Bagnold, R. A. (1990). Sand, Wind, and War: Memoirs of a Desert Explorer. Tucson: University of Arizona Press.Google Scholar
Bagnold, R. A., Myers, O. H., Peel, R. F. & Winkler, H. A. (1939). An expedition to the Gilf Kebir and ‘Uweinat, 1938. Geographical Journal, 93, 281313.CrossRefGoogle Scholar
Bailey, G., Alsharekh, A., Flemming, N., et al. (2007). Coastal prehistory in the southern Red Sea Basin, underwater archaeology, and the Farasan Islands. Proceedings of the Seminar for Arabian Studies, 37, 116.Google Scholar
Baker, S. (1866). The Albert N’Yanza. Great Basin of the Nile and Explorations of the Nile Sources. Volume 1. London: Sidgwick and Jackson (reprinted 1962).Google Scholar
Ball, J. (1939). Contributions to the Geography of Egypt. Cairo, Egypt: Government Press.Google Scholar
Barakat, H. N. & Hegazy, A. K. (eds.) (1997). Reviews in Ecology, Desert Conservation and Development: A Festschrift for Prof. M. Kassas on the Occasion of His 75th Birthday. Cairo: UNESCO, IDRC & South Valley University, Egypt.Google Scholar
Barboni, D., Bonnefille, R., Alexandre, A. & Meunier, J. D. (1999). Phytoliths as paleoenvironmental indicators, West Side Middle Awash, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology, 152, 87100.CrossRefGoogle Scholar
Barbour, K. M. (1961). The Republic of the Sudan: A Regional Geography. London: University of London Press.Google Scholar
Barker, P. A., Hurrell, E. R., Leng, M. J., et al. (2011). Seasonality in equatorial climate over the past 25 k.y. revealed by oxygen isotope records from Mount Kilimanjaro. Geology, 39, 11111114.Google Scholar
Baroin, C. (2003). Les Toubou du Sahara Central. Paris: Vents de Sable.Google Scholar
Barrows, T. T., Williams, M. A. J., Mills, S. C., et al. (2014). A White Nile megalake during the last interglacial period. Geology, 42, 163166.Google Scholar
Barthelme, J. W. (1984). Early evidence for animal domestication in eastern Africa. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press, pp. 200211.Google Scholar
Bar-Yosef, O. & Belfer-Cohen, A. (2001). From Africa to Eurasia: Early dispersals. Quaternary International, 75, 1928.Google Scholar
Bate, D. M. A. (1951). The mammals from Singa and Abu Hugar. In British Museum (Natural History) Fossil Mammals of Africa No 2: The Pleistocene Fauna of Two Blue Nile Sites, pp. 128.Google Scholar
Bates, R. L. & Jackson, J. A. (1987). Glossary of Geology, 3rd edn. Alexandria, VA: American Geological Institute.Google Scholar
Beadle, L. G. (1974). The Inland Waters of Tropical Africa: An Introduction to Tropical Limnology. London: Longman.Google Scholar
Beinroth, F. H. (1966). Über drei Vorkommen von Vertisols im Mittleren Sudan. Eigenschaften, Klassifikation, Enstehung und landwirtschafliche Eignung. Arbeiten aus dem Geologisch-Paläontologischen Institut der Technischen Hochschule Stuttgart, NF 49, 1115.Google Scholar
Belfer-Cohen, A. & Goring-Morris, A. N. (2017). The Upper Palaeolithic in Cisjordan. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 627637.Google Scholar
Bell, B. (1970). The oldest records of the Nile floods. Geographical Journal, 136, 569573.Google Scholar
Bell, B. (1971). The Dark Ages in ancient history. I. The First Dark Age in Egypt. American Journal of Archaeology, 75, 126.CrossRefGoogle Scholar
Bell, B. (1975). Climate and the history of Egypt: The Middle Kingdom. American Journal of Archaeology, 79, 223269.CrossRefGoogle Scholar
Benazzi, S., Douka, K., Fornai, C., et al. (2011). Early dispersals of modern humans in Europe and implications for Neanderthal behaviour. Nature, 479, 525528.Google Scholar
Benchelah, A.-C., Bouziane, H., Maka, M. & Ouahès, C. (2000). Fleurs du Sahara: Voyage ethnobotanique avec les Touaregs du Tassili. Paris: Ibis Press.Google Scholar
Bergner, A. G. N. & Trauth, M. H. (2004). Comparison of the hydrologic and hydro-chemical evolution of Lake Naivasha (Kenya) during three highstands between 175 and 60 kyr BP. Palaeogeography, Palaeoclimatology, Palaeoecology, 125, 1736.CrossRefGoogle Scholar
Bernus, E. (1974). Les Illabakan (Niger). Une tribu touarègue sahélienne et son aire de nomadisation. Paris: ORSTOM.CrossRefGoogle Scholar
Berry, L. (1962). Large scale alluvial islands in the White Nile. Revue de Géomorphogie. Dynamique, 12, 105108.Google Scholar
Beucher, F. (1971). Étude palynologique de formations néogenes et quaternaries au Sahara nord-occidental, 3 vols., Paris.Google Scholar
Beuf, S., Bijou-Duval, B., De Charpal, O., et al. (1971). Les Grés du Paléozoique inférieur au Sahara. Publication de l’Institut français du Pétrole. Paris: Technip.Google Scholar
Beuning, K. R. M., Kelts, K., Ito, E. & Johnson, T. C. (1997a). Paleohydrology of Lake Victoria, East Africa, inferred from 18O/16O ratios in sediment cellulose. Geology, 25, 10831086.2.3.CO;2>CrossRefGoogle Scholar
Beuning, K. R. M., Talbot, M. R. & Kelts, K. (1997b). A revised 30,000-year paleoclimatic record and paleohydrologic history of Lake Albert, East Africa. Palaeogeography, Palaeoclimatology, Palaeoecology, 136, 259279.CrossRefGoogle Scholar
Bird, E. (2000). Coastal Geomorphology: An Introduction. Chichester: John Wiley & Sons.Google Scholar
Birkeland, P. W. (1999). Soils and Geomorphology, 3rd edn. New York, NY: Oxford University Press.Google Scholar
Bishai, H. M. (1962). The water characteristics of the Nile in the Sudan with a note on the effect of Eichhornia crassipes on the hydrobiology of the Nile. Hydrobiologia, 19, 357382.Google Scholar
Biswas, A. K. (1970). History of Hydrology. Amsterdam: North-Holland.Google Scholar
Blanchet, C. L., Contoux, C. & Leduc, G. (2015). Runoff and precipitation dynamics in the Blue and White Nile catchments during the mid-Holocene: A data-model comparison. Quaternary Science Reviews, 130, 222230.CrossRefGoogle Scholar
Blanchet, C. L., Tjallingii, R., Frank, M., et al. (2013). High- and low-latitude forcing of the Nile River regime during the Holocene inferred from laminated sediments of the Nile deep-sea fan. Earth and Planetary Science Letters, 364, 98110.CrossRefGoogle Scholar
Blokhuis, W. A. (1993). Vertisols in the Central Clay Plain of the Sudan. Wageningen: Wageningen Agricultural University.Google Scholar
Boivin, N., Fuller, D. Q., Dennell, R., Allaby, R. & Petraglia, M. D. (2013). Human dispersal across diverse environments of Asia during the Upper Pleistocene. Quaternary International, 300, 3247.CrossRefGoogle Scholar
Bonfils, C., de Noblet-Ducoudré, N., Braconnot, P. & Joussaume, S. (2001). Hot desert albedo and climate change: Mid-Holocene monsoon in North Africa. Journal of Climate, 14, 37243737.Google Scholar
Bonne, M. (1762). Carte de L’Egypte Ancienne et Moderne. Paris: Chez Lattré, 1762.Google Scholar
Bonnefille, R. (1972). Associations polliniques actuelles et quaternaries en Éthiopie (Vallée de l’Awash et de l’Omo), 2 vols., Paris.Google Scholar
Bonnefille, R. (1976a). Végétations et climats des temps oldowayens et acheuléens à Melka-Kunturé, Éthiopie. In L’Éthiopie avant l’histoire, Vol. 1, Chavaillon, J. (ed.). Paris, CNRS, pp. 5572.Google Scholar
Bonnefille, R. (1976b). Implications of pollen from the Koobi Fora Formation, East Rudolf, Kenya. Nature, 264, 403407.Google Scholar
Bonnefille, R. (1980). Vegetation history of savanna in East Africa during the Plio-Pleistocene. IV International Palynological Conference, 3, 7889.Google Scholar
Bonnefille, R. (1983). Evidence for a cooler and drier climate in the Ethiopian uplands towards 2.4 myr ago. Nature, 303, 487491.CrossRefGoogle Scholar
Bonnefille, R., Potts, R., Chalié, F., Jolly, D. & Peyron, O. (2004). High-resolution vegetation and climate change associated with Pliocene Australopithecus afarensis. Proceedings of the National Academy of Sciences of the USA, 101, 1212512129.Google Scholar
Bonnet, C. (1992). Excavations at the Nubian royal town of Kerma: 1975–1991. Antiquity, 66, 211225.CrossRefGoogle Scholar
Borda, M. (2011). New painted shelter at Jebel Arkenu (Libya). Sahara, 22, 125129.Google Scholar
Bosworth, W., Huchon, P. & McClay, K. (2005). The Red Sea and Gulf of Aden basins. Journal of African Earth Sciences, 43, 334378.Google Scholar
Bowden, P., Van Breemen, O., Hutchison, J. & Turner, D. C. (1976). Palaeozoic and Mesozoic age trends for some ring complexes in Niger and Nigeria. Nature, 259, 297299.CrossRefGoogle Scholar
Bowen, R. & Jux, U. (1987). Afro Arabian Geology: A Kinematic View. London: Chapman & Hall.Google Scholar
Box, M. R., Krom, M. D., Cliff, R. A., et al. (2011). Response of the Nile and its catchment to millennial-scale climatic changes since the LGM from Sr isotopes and major elements of East Mediterranean sediments. Quaternary Science Reviews, 30, 431442.Google Scholar
Braconnot, P., Joussaume, S., de Noblet, N. & Ramstein, G. (2001). Mid-Holocene and Last Glacial Maximum African monsoon changes as simulated within the Paleoclimate Modelling Intercomparison Project. Global and Planetary Change, 26, 5166.Google Scholar
Bragg, M. (2003). The Adventure of English: The Biography of a Language. London: Hodder & Stoughton.Google Scholar
Brandt, S. A. (1984). New perspectives on the origins of food production in Ethiopia. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press, pp. 173190.Google Scholar
Brandt, S. A. (1986). The Upper Pleistocene and early Holocene prehistory of the Horn of Africa. African Archaeological Review, 4, 4182.CrossRefGoogle Scholar
Brass, M. (2014). The southern frontier of the Meroitic State: The view from Jebel Moya. African Archaeological Review, 31, 425445.Google Scholar
Brass, M. (2015). Interactions and pastoralism along the southern and southeastern frontiers of the Meroitic State, Sudan. Journal of World Prehistory, 28, 255288.Google Scholar
Brass, M. (2017). Early North African cattle domestication and its ecological setting: A reassessment. Journal of World Prehistory, doi.org/10.1007/s10963-017–91122-9Google Scholar
Brass, M. & Schwenniger, J.-L. (2013). Jebel Moya (Sudan): New dates from a mortuary complex at the southern Meroitic frontier. Azania: Archaeological Research in Africa, 48, 455472.CrossRefGoogle ScholarPubMed
Bräuer, G., Yokoyama, Y., Falguères, C. & Mbua, E. (1997). Modern human origins backdated. Nature, 386, 337338.CrossRefGoogle ScholarPubMed
Breed, C. S., McCauley, J. F. & Davis, P. A. (1987). Sand sheets of the eastern Sahara and ripple blankets on Mars. In Frostick, L. & Reid, I. (eds.), Desert Sediments: Ancient and Modern. Geological Society Special Publication, No. 35, pp. 337359.Google Scholar
Breeze, P. S., Groucutt, H. S., Drake, N. H., White, T. S., Jennings, R. P. & Petraglia, M. D. (2016). Palaeohydrological corridors for hominin dispersals in the Middle East ~250–70,000 years ago. Quaternary Science Reviews, 144, 155185.Google Scholar
Breuil, H. (1926). Gravures rupestres du désert libyque identique à celles des anciens Bushmen. L’Anthropologie, 36, 125127.Google Scholar
Breuil, H. (1928). Les gravures de Djebel Ouenat. Revue scientifique, Paris, 25 (February).Google Scholar
Brewer, D. J. (1989). A model for resource exploitation in the prehistoric Fayum. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Late Prehistory of the Nile Basin and the Sahara. Poznan, Poland, Poznan Archaeological Museum, pp. 127137.Google Scholar
Brewer, R. (1964). Fabric and Mineral Analysis of Soils. New York, NY: John Wiley & Sons.Google Scholar
Brookfield, M. (2010). The desertification of the Egyptian Sahara during the Holocene (the last 10,000 years) and its influence on the rise of Egyptian civilization. In Martini, I. P. & Chesworth, W. (eds.), Landscapes and Societies. Dordrecht: Springer, pp. 91108.CrossRefGoogle Scholar
Brown, D. (1980). Freshwater Snails of Africa and Their Medical Importance, 1st edn. London: Taylor and Francis.Google Scholar
Brown, D. (1994). Freshwater Snails of Africa and their Medical Importance. Revised 2nd edn. London: Taylor and Francis.CrossRefGoogle Scholar
Brown, D. S. (1965). Freshwater Gastropod Mollusca from Ethiopia. Bulletin of the British Museum (Natural History) Zoology, 12(2), 3994, and plates.Google Scholar
Brown, F. H. & Fuller, C. R. (2008). Stratigraphy and tephra of the Kibish Formation, southwestern Ethiopia. Journal of Human Evolution, 55, 366403.Google Scholar
Brown, F. H., McDougall, I. & Fleagle, J. G. (2012). Correlation of the KHS Tuff of the Kibish Formation to volcanic ash layers at other sites, and the age of early Homo sapiens (Omo I and Omo II). Journal of Human Evolution, 63, 577585.CrossRefGoogle ScholarPubMed
Brown, L. H. & Cochemé, J. (1973). A study on the Agroclimatology of the Highlands of Eastern Africa. World Meteorological Organisation, Technical Note No. 125.Google Scholar
Bruce, J. (1790). Travels to Discover the Source of the Nile. Edinburgh: Edinburgh University Press, abridged version (1964).Google Scholar
Bubenzer, O. & Riemer, H. (2007). Holocene climatic change and human settlement between the central Sahara and the Nile Valley: Archaeological and geomorphological results. Geoarchaeology, 22, 607620.CrossRefGoogle Scholar
Büdel, J. (1954). Klima-morphologische Arbeiten in Äthiopien im Frühjahr 1953. Erdkunde, 8, 139156.Google Scholar
Bunting, A. H. & Lea, J. D. (1962). The soils and vegetation of the Fung, east central Sudan. Journal of Ecology, 50, 529558.CrossRefGoogle Scholar
Buol, S. W., Hole, F. D. & McCracken, R. J. (1973). Soil Genesis and Classification. Ames, IA: Iowa State University Press.Google Scholar
Bureau de Recherches Géologiques et Minières (1981). 1:2,000,000 Geological Map of the Sudan. Orléans, France: B.R.G.M.Google Scholar
Burnett, J. R. (1948). Crop production. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 275301.Google Scholar
Burollet, P. F. (1963) Reconnaissance géologique dans le sud-est du basin de Kufra. Premier Symposium Saharien, Tripoli, 1963. Revue de l’Institut Français du Pétrole, 18(10), 219227.Google Scholar
Bussert, R., Eisawi, A. A. M., Hamed, B. & Babikir, I. A. A. (2018). Neogene palaeochannel deposits in Sudan: Remnants of a trans-Saharan river system? Journal of African Earth Sciences, 141, 921.CrossRefGoogle Scholar
Butzer, K. W. (1958). Quaternary stratigraphy and climate in the Near East. Bonner Geographische Abhandlungen, 24.Google Scholar
Butzer, K. W. (1971a). Recent History of an Ethiopian Delta: The Omo River and the level of Lake Rudolf. University of Chicago, Department of Geography Research Paper No. 136.Google Scholar
Butzer, K. W. (1971b). Environment and Archaeology, 2nd edn. London: Methuen,Google Scholar
Butzer, K. W. (1976). Early Hydraulic Civilization in Egypt: A Study in Cultural History. Prehistoric Archeology and Ecology Series. Chicago: University of Chicago Press.Google Scholar
Butzer, K. W. (1980). Pleistocene history of the Nile Valley in Egypt and Lower Nubia. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Quaternary Environments and Prehistoric Occupation in Northern Africa. Rotterdam: A. A. Balkema, pp. 253280.Google Scholar
Butzer, K. W. & Hansen, C. L. (1968). Desert and River in Nubia. Geomorphology and Prehistoric Environments at the Aswan Reservoir. Madison, WI: University of Wisconsin Press.Google Scholar
Butzer, K. W., Isaac, G. L. l., Richardson, J. L. & Washbourn-Kamau, C. (1972). Radiocarbon dating of East African lake levels. Science, 175, 10691076.Google Scholar
Buursink, J. (1971). Soils of Central Sudan. Utrecht: University of Utrecht.Google Scholar
Cahen, D. & Moeyersons, J. (1977). Subsurface movements of stone artefacts and their implications for the prehistory of Central Africa. Nature, 266, 812815.CrossRefGoogle Scholar
Camberlin, P. (1997). Rainfall anomalies in the source region of the Nile and their connection with the Indian summer monsoon. Journal of Climate, 10, 13801392.2.0.CO;2>CrossRefGoogle Scholar
Camberlin, P., Janicot, S. & Poccard, I. (2001). Seasonality and atmospheric dynamics of the teleconnection between African rainfall and tropical sea-surface temperature: Atlantic vs. ENSO. International Journal of Climatology, 21, 9731005.Google Scholar
Caneva, I. (1988). The cultural equipment of the Early Neolithic at Geili. In El Geili. The History of a Middle Nile Environment 7000BC-AD1500, Caneva, I. (ed.). Oxford, BAR International Series, 424, pp. 65147.CrossRefGoogle Scholar
Cann, R. L. (2001). Genetic clues to dispersal in human populations: Retracing the past from the present. Science, 291, 17421748.Google Scholar
Cann, R., Stoneking, M. & Wilson, A. (1987). Mitochondrial DNA and human evolution. Nature, 325, 3136.CrossRefGoogle ScholarPubMed
Caporiacco, L. & Graziosi, P. (1934). Le pitture rupestri di Ain Doua (El Auenat). Florence: Centro di Studi Coloniale.Google Scholar
Capozzi, R. & Negri, A. (2009). Role of sea-level forced sedimentary processes on the distribution of organic carbon-rich marine sediments: A review of the Late Quaternary sapropels in the Mediterranean Sea. Palaeogeography, Palaeoclimatology, Palaeoecology, 273, 249257.CrossRefGoogle Scholar
Carbonell, E., Mosquera, M., Rodriguez, X. P. & Sala, R. (1999). Out of Africa: The dispersal of the earliest technical systems reconsidered. Journal of Anthropological Archaeology, 18, 119136.CrossRefGoogle Scholar
Carrington, D. P., Gallimore, R. G. & Kutzbach, J. E. (2001). Climate sensitivity to wetlands and wetland vegetation in mid-Holocene North Africa. Climate Dynamics, 17, 151157.Google Scholar
Casini, M. (1984). Neolithic and Predynastic in the Fayum. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Origin and Early Food Development of Food-Producing Cultures in North-Eastern Africa. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 199204.Google Scholar
Castañeda, I. S., Mulitza, S., Schefuß, E., et al. (2009). Wet phases in the Sahara/Sahel region and human migration patterns in North Africa. Proceedings of the National Academy of Sciences of the USA, 106, 2015920163.Google Scholar
Caton Thompson, G. (1952). Kharga Oasis in Prehistory. London: Athlone Press.Google Scholar
Caton Thompson, G. & Gardner, E. W. (1929). Recent work on the problem of Lake Moeris. Geographical Journal, 73, 2058.CrossRefGoogle Scholar
Caton Thompson, G. & Gardner, E. W. (1932). The prehistoric geography at Kharga Oasis. Geographical Journal, 80, 369406.CrossRefGoogle Scholar
Caton Thompson, G. & Gardner, E. W. (1934). The Desert Fayum. London: The Royal Anthropological Institute of Great Britain and Ireland.Google Scholar
Causse, C., Conrad, G., Fontes, J.-C., et al. (1988). Le dernier ‘Humide’ pléistocène du Sahara nord-occidentale daterait de 80–100,000 ans. Compte Rendu de l’Académie des Sciences, 306, Série II, 14591464.Google Scholar
Ceram, C. W. (1967). Gods, Graves and Scholars: The Story of Archaeology, 2nd edn., translated from the German by E. B. Garside & Sophie Wilkins. Harmondsworth: Penguin Books.Google Scholar
Cerling, T. E. (1979). Paleochemistry of Plio-Pleistocene Lake Turkana, Kenya. Palaeogeography, Palaeoclimatology, Palaeoecology, 27, 247285.CrossRefGoogle Scholar
Cerling, T. E. (1996). Pore water chemistry of an alkaline lake: Lake Turkana, Kenya. In Johnson, T.C. & Odada, E. O. (eds.), The Limnology, Climatology and Paleoclimatology of the East African Rift Lakes. Amsterdam: Gordon and Breach, pp. 225240.Google Scholar
Cerling, T. E., Harris, J. M., MacFadden, B. J., et al. (1997). Global vegetation change through the Miocene/Pliocene boundary. Nature, 389, 153158.Google Scholar
Cerling, T. E., Hay, R. L. & O’Neil, J. R. (1977). Isotopic evidence for dramatic climatic changes in East Africa during the Pleistocene. Nature, 267, 137138.CrossRefGoogle ScholarPubMed
Cerling, T. E., Levin, N. E., Quade, J., et al. (2010). Comment on the paleoenvironment of Ardipithecus ramidus. Science, 328, Technical Comment 1105-d.CrossRefGoogle ScholarPubMed
Cerling, T. E., Wynn, J. G., Andanje, S. A., et al. (2011). Woody cover and hominin environments in the past 6 million years. Nature, 476, 5156.Google Scholar
Chalié, F. & Gasse, F. (2002). Late-Glacial-Holocene diatom record of water chemistry and lake-level change from the tropical East African Rift Lake Abiyata (Ethiopia). Palaeogeography, Palaeoclimatology, Palaeoecology, 187, 259283.CrossRefGoogle Scholar
Charney, J. G. (1975). Dynamics of deserts and droughts in the Sahel. Quarterly Journal of the Royal Meteorological Society, 101, 193202.Google Scholar
Charney, J. G., Stone, P. H. & Quirk, W. J. (1975). Drought in the Sahara: A biogeophysical feedback mechanism. Science, 187, 434435.Google Scholar
Chase, B. M. (2009). Evaluating the use of dune sediments as a proxy for palaeo-aridity: a southern African case study. Earth-Science Reviews, 93, 3145.Google Scholar
Chavaillon, J. (1976). Mission archéologique franco-éthiopienne de Melka-Kunturé: rapport préliminaire 1972–1975. In Chavaillon, J. (ed.), L’Éthiopie avant l’histoire, Vol. 1. Paris: CNRS, pp. 111.Google Scholar
Chialvo, J. (1975). Contribution à la géologie du confluent Atbara-Setit: Étude des sites de barrage de Rumela et Burdana (République Démocratique du Soudan). Unpublished Doctorat de 3ème Cycle thesis, Université Scientifique et Médicale de Grenoble.Google Scholar
Chorowicz, J. (2005). The East African rift system. Journal of African Earth Sciences, 43, 379410.Google Scholar
Chumakov, I. S. (1967). Pliocene and Pleistocene deposits of the Nile Valley in Nubia and Upper Egypt (in Russian). Transactions of the Geological Institute of the Academy of Sciences (USSR), 170, 111.Google Scholar
Chylek, P., Lesins, G. & Lohmann, U. (2001). Enhancement of dust source area during past glacial periods due to changes of the Hadley circulation. Journal of Geophysical Research, 106, 1847718485.CrossRefGoogle Scholar
Clark, J. D. (1971). A re-examination of the evidence for agricultural origins in the Nile Valley. Proceedings of the Prehistoric Society, 37, 3479.Google Scholar
Clark, J. D. (1973a). Preliminary report of an archaeological and geomorphological survey in the central Sudan, January to March, 1973. Unpublished report, Berkeley.Google Scholar
Clark, J. D. (1973b). Recent archaeological and geomorphological field studies in the Sudan: Some preliminary results. Nyame Akuma, 3, 5564.Google Scholar
Clark, J. D. (1975). Africa in prehistory: Peripheral or paramount? Man N.S., 10, 175198. (The 1974 Huxley Memorial Lecture).CrossRefGoogle Scholar
Clark, J. D. (1980). Human populations and cultural adaptations in the Sahara and Nile during prehistoric times. In Williams, M. A. J & Faure, H. (eds.), The Sahara and The Nile: Quaternary Environments and Prehistoric Occupation In Northern Africa. Rotterdam: A. A. Balkema, pp. 527582.Google Scholar
Clark, J. D. (1982). The cultures of the Middle Palaeolithic/Middle Stone Age. In Clark, J. D. (ed.), The Cambridge History of Africa, Vol. 1: From the Earliest Times to c. 500 BC. Cambridge: Cambridge University Press, pp. 248341.Google Scholar
Clark, J. D. (1984a). The domestication process in Northeast Africa: Ecological change and adaptive strategies. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Origin and Early Development of Food-Producing Cultures in North-Eastern Africa. Poznan, Poland, Polish Academy of Sciences and Poznan Archaeological Museum, pp. 2541.Google Scholar
Clark, J. D. (1984b). Prehistoric cultural continuity and economic change in the Central Sudan in the Early Holocene. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press, pp.113126.Google Scholar
Clark, J. D. (1987). Transitions: Homo erectus and the Acheulian: The Ethiopian sites of Gadeb and the Middle Awash. Journal of Human Evolution, 16, 809826.Google Scholar
Clark, J. D. (1989). Shabona: An Early Khartoum settlement on the White Nile. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Late Prehistory of the Nile Basin and the Sahara. Poznan, Poland: Poznan Archaeological Museum, pp. 387410.Google Scholar
Clark, J. D., Asfaw, B., Assefa, G., et al. (1984). Palaeoanthropological discoveries in the Middle Awash Valley, Ethiopia. Nature, 307, 423428.Google Scholar
Clark, J. D., Beyene, Y., WoldeGabriel, G., et al. (2003). Stratigraphic, chronological and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature, 423, 747752.CrossRefGoogle ScholarPubMed
Clark, J. D. & Brandt, S. A. (eds.) (1984). From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press.Google Scholar
Clark, J. D., Brown, K. S., Dietler, M., et al. (2008b). The Late Acheulian assemblages. In Gifford-Gonzalez, D. (ed.), Adrar Bous: Archaeology of a Central Saharan Granitic Ring Complex in Niger. Tervuren, Belgium: Royal Museum for Central Africa, pp. 5590.Google Scholar
Clark, J. D., Carter, P. L., Gifford-Gonzalez, D. & Smith, A. B. (2008c). The Adrar Bous cow and African cattle. In Gifford-Gonzalez, D. (ed.), Adrar Bous: Archaeology of a Central Saharan Granitic Ring Complex in Niger. Tervuren, Belgium: Royal Museum for Central Africa, pp. 355368.Google Scholar
Clark, J. D. & Harris, J. W. K. (1985). Fire and its role in early hominid lifeways. African Archaeological Review, 3, 327.Google Scholar
Clark, J. D. & Kurashina, H. (1979). Hominid occupation of the east-central highlands of Ethiopia in the Plio-Pleistocene. Nature, 282, 3339.CrossRefGoogle Scholar
Clark, J. D. & Schick, K. (2000). Overview and conclusion on the Middle Awash Acheulean. In de Heinzelin, J., Clark, J. D., Schick, K. D. & Gilbert, W. H. (eds.), The Acheulean and the Plio-Pleistocene Deposits of the Middle Awash Valley, Ethiopia. Tervuren, Belgium: Royal Museum of Central Africa. Annales Sciences Geologiques, 104, 193202.Google Scholar
Clark, J. D. & Stemler, A. (1975). Early domesticated sorghum from Central Sudan. Nature, 254, 588591.CrossRefGoogle Scholar
Clark, J. D. & Williams, M. A. J. (1977). Recent archaeological research in southeastern Ethiopia (1974–75): Some preliminary results. Annales d’Ethiopie, 11, 1942.CrossRefGoogle Scholar
Clark, J. D., Williams, M. A. J. & Smith, A.B. (1973). The geomorphology and archaeology of Adrar Bous, Central Sahara: A preliminary report. Quaternaria, 17, 245–97.Google Scholar
Clark, J. D. with Schultz, D. U., Kroll, E. M., et al. (2008a). The Aterian of Adrar Bous and the central Sahara. In Gifford-Gonzalez, D. (ed.), Adrar Bous: Archaeology of a Central Saharan Granitic Ring Complex in Niger. Tervuren, Belgium: Royal Museum for Central Africa, pp. 91162.Google Scholar
Clark, P.U., Archer, D., Pollard, D., Blum, J.D., Rial, J.A., Brovkin, V., Mix, A.C., Pisias, N.G., & Roy, M. (2006). The Middle Pleistocene transition: Characteristics, mechanisms, and implications for long-term changes in atmospheric pCO2. Quaternary Science Reviews, 25, 31503184.CrossRefGoogle Scholar
Claussen, M., Brovkin, V., Ganopolski, A., Kubatzki, C. & Petoukhov, V. (1998). Modelling global terrestrial vegetation-climate interaction. Philosophical Transactions of the Royal Society of London, B353, 5363.Google Scholar
Claussen, M., Kubatzki, C., Brovkin, V., et al. (1999). Simulation of an abrupt change in Saharan vegetation in the mid-Holocene. Geophysical Research Letters, 26, 20372040.CrossRefGoogle Scholar
Close, A. E. (ed.) (1987). Prehistory of Arid North Africa: Essays in Honor of Fred Wendorf. Dallas: Southern Methodist University.Google Scholar
Cockerton, H.E., Street-Perrott, F.A., Barker, P.A., Leng, M.J., Sloane, H.J., & Ficken, K.J., (2015). Orbital forcing of glacial/interglacial variations in chemical weathering and silicon cycling within the upper White Nile basin, East Africa: Stable-isotope and biomarker evidence from Lakes Victoria and Edward. Quaternary Science Reviews, 130, 5771.Google Scholar
Cohen, K. M., Finney, S. C., Gibbard, P. & Fan, J.-X. (2013). The ICS International Chronostratigraphic Chart. Episodes, 36, 199204.CrossRefGoogle Scholar
Cohen, M. N. (1977). The Food Crisis in Prehistory: Overpopulation and the Origins of Agriculture. New Haven, CT: Yale University Press.Google Scholar
Colchester, G. V. (1927). Limestone deposits of Darfur. Mimeo, pp. 15, Khartoum, May 19, 1927.Google Scholar
Collins, A. S. & Pisarevsky, S. A. (2005). Amalgamating eastern Gondwana: The evolution of the Circum-Indian Orogens. Earth Science Reviews, 71, 229270.CrossRefGoogle Scholar
Compère, P. (1984). Some algae from the Red Sea Hills in north-eastern Sudan. Hydrobiologia, 110, 6177.Google Scholar
Conrad, G. (1959). Importance et rôle des termites dans les formations pédologiques fossiles du Quaternaire de la région de Béni-Abbès. Comptes Rendus, Académie des Sciences, Paris, 249, 20892091.Google Scholar
Conrad, G. (1969). L’évolution continentale post-hercynienne du Sahara algérien (Saoura, Erg Chech, Tanezrouft, Ahnet Mouydir). Paris: CNRS , 527 pp.Google Scholar
Conway, D. (2000). The climate and hydrology of the Upper Blue Nile River. Geographical Journal, 166, 4962.Google Scholar
Conway, D. & Hulme, M. (1993). Recent fluctuations in precipitation and runoff over the Nile sub-basins and their impact on Nile discharge. Climatic Change, 25, 127151.Google Scholar
Cooke, H. B. S. (1958). Observations relating to Quaternary environments in East and Southern Africa. Alex du Toit Memorial Lecture no. 5. Geological Society of South Africa Bulletin, Annexure to Vol. 60.Google Scholar
Cooke, R., Warren, A. & Goudie, A. (1993). Desert Geomorphology. London: UCL Press.Google Scholar
Cooper, A. & Stringer, C. B. (2013). Did the Denisovans cross Wallace’s Line? Science, 342, 321323.Google Scholar
Coppens, Y., Clark Howell, F., Isaac, G. Ll. & Leakey, R. E. F. (eds.) (1976). Earliest Man and Environments in the Lake Rudolf Basin. Stratigraphy, Paleoecology, and Evolution. Chicago: University of Chicago Press.Google Scholar
Coque, R. (1962). La Tunisie présaharienne. Étude géomorphologique. Paris: Oberthür, Rennes.Google Scholar
Corti, G. (2009). Continental rift evolution: From rift initiation to incipient break-up in the Main Ethiopian Rift, East Africa. Earth-Science Reviews, 96, 153.CrossRefGoogle Scholar
Corvinus, G. (1975). Palaeolithic remains at Hadar in the Afar region. Nature, 256, 468471.CrossRefGoogle Scholar
Cosentino, D., Buchwald, R., Sampalmieri, G., et al. (2013). Refining the Mediterranean ‘Messinian gap’ with high-precision U-Pb zircon geochronology, central and northern Italy. Geology, 41, 323326.CrossRefGoogle Scholar
Costa, K., Russell, J., Konecky, B. & Lamb, H. (2014). Isotopic reconstruction of the African Humid Period and Congo Air Boundary migration at Lake Tana, Ethiopia. Quaternary Science Reviews, 83, 5867.Google Scholar
Coudé-Gaussen, G. & Rognon, P. (1983). Les poussières sahariennes. La Recherche, 147, 10501061.Google Scholar
Coudé-Gauusen, G., Rognon, P., Rapp, A. & Nihlén, T. (1987). Dating of peridesert loess in Matmata, south Tunisia, by radiocarbon and thermoluminescence methods. Zeitschrift für Geomorphologie N.F., 31, 129144.CrossRefGoogle Scholar
Coulson, D. & Campbell, A. (2001). African Rock Art: Paintings and Engravings on Stone. New York, NY: H. A. Abrams.Google Scholar
Coulthard, T. J., Ramirez, J. A., Barton, N., Rogerson, M. & Brücher, T. (2013). Were rivers flowing across the Sahara during the last interglacial? Implications for human migration through Africa. Plos ONE, 8(9), e74834. DOI:10.1371/journal.pone.0074834Google Scholar
Cramp, A. & O’Sullivan, G. (1999). Neogene sapropels in the Mediterranean: A review. Marine Geology, 153, 1128.CrossRefGoogle Scholar
Cremaschi, M., Zerboni, A., Spötl, C. & Felletti, F. (2010). The calcareous tufa in the Tadrart Acacus Mt. (SW Fezzan, Libya): An early Holocene palaeoclimate archive in the central Sahara. Palaeogeography, Palaeoclimatology, Palaeoecology, 287, 8194.Google Scholar
Crombie, M. K., Arvidson, R. E., Sturchio, N. C., El Alfy, Z. & Abu Zeid, K. (1997). Age and isotopic constraints on Pleistocene pluvial episodes in the Western desert, Egypt. Palaeogeography, Palaeoclimatology, Palaeoecology, 130, 337355.Google Scholar
Crouvi, O., Amit, R., Enzel, Y. & Gillespie, A. R. (2010). Active sand seas and the formation of desert loess. Quaternary Science Reviews, 29, 20872098.CrossRefGoogle Scholar
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. Quaternary Research, 70, 275282.Google Scholar
Crystal, D. (2004). The Stories of English. London: Allen Lane.Google Scholar
Cullen, H. M., deMenocal, P. B., Hemming, S., et al. (2000). Climate change and the collapse of the Akkadian empire: Evidence from the deep sea. Geology, 28, 379382.Google Scholar
Dafalla, H. (1975). The Nubian Exodus. London: C. Hurst & Co., and Uppsala: Scandinavian Institute of African Studies.Google Scholar
Darkoh, M. B. K. (1989). Combating Desertification in the Southern African region: An Updated Regional Assessment. Nairobi: UNEP.Google Scholar
Darkoh, M. B. K. (1999). Desertification problem: Global extent and main concepts. In Proceedings of the International Scientific Conference on Desertification and Land Degradation, Moscow, 11–15 November 1999, pp. 1132.Google Scholar
Darwin, C. (1846). An account of the fine dust which often falls on vessels in the Atlantic Ocean. Quarterly Journal of the Geological Society of London, 2, 2630.Google Scholar
Darwin, C. (1881). The Formation of Vegetable Mould through the Action of Worms, with Observations on Their Habits. London: John Murray.CrossRefGoogle Scholar
Daveau, S. (1965). Dunes ravinées et dépôts du Quaternaire récent dans le Sahel mauritanien. Revue de Géographie de l’Afrique occidentale, 1–2, 748.Google Scholar
Davies, H. R. J. & Walsh, R. P. D. (1997). Historical changes in the flood hazard at Khartoum, Sudan: Lessons and warnings for the future. Singapore Journal of Tropical Geography, 18, 123140.Google Scholar
Davies, J. L. (1969). Landforms of Cold Climates. Canberra: Australian National University Press.Google Scholar
Davis, M., Matmon, A., Rood, D. H. & Avnaim-Katav, S. (2012). Constant cosmogenic nuclide concentrations in sand supplied from the Nile River over the past 2.5 m.y. Geology, 40, 359362.CrossRefGoogle Scholar
De Deckker, P. & Williams, M. A. J. (1993). Lacustrine paleoenvironments of the Area of Bir Tarfawi-Bir Sahara reconstructed from fossil ostracods and the chemistry of their shells. In Wendorf, F., Schild, R., Close, A. E., et al. (eds.), Egypt During the Last Interglacial: The Middle Palaeolithic of Bir Tarfawi and Bir Sahara East. New York, NY: Plenum Press, pp. 115119.Google Scholar
De Heinzelin, J. (1968). Geological history of the Nile Valley in Nubia. In Wendorf, F. (ed.), The Prehistory of Nubia, Vol. 1. Dallas, TX: Southern Methodist University, 19–55.Google Scholar
De Heinzelin, J., Clark, J. D., Schick, K. D. & Gilbert, W. H. (eds.) (2000). The Acheulean and the Plio-Pleistocene Deposits of the Middle Awash Valley, Ethiopia. Musée Royal de l’Afrique Central, Tervuren, Belgium, Annales-Sciences Géologiques, 104.Google Scholar
Delibrias, G., Gasse, F. & Rognon, P. (1973). Existence de lacs importants au Pléistocène supérieur (34,000–23,000 ans BP) dans l’Afar méridional (Éthiopie). Comptes Rendus, Académie des Sciences, Paris, 277D, 26332636.Google Scholar
deMenocal, P. B. (2004). African climate change and faunal evolution during the Pliocene-Pleistocene. Earth and Planetary Science Letters, 220, 324.Google Scholar
deMenocal, P., Ortiz, J., Guilderson, T., et al. (2000). Abrupt onset and termination of the African Humid Period: Rapid responses to gradual insolation forcing. Quaternary Science Reviews, 19, 347361.Google Scholar
Dennell, R. & Roebroeks, W. (2005). An Asian perspective on early human dispersal from Africa. Nature, 438, 10991104.Google Scholar
De Putter, T., Loutre, M.-F. & Wansard, G. (1998). Decadal periodicities of Nile River historical discharge (A.D.622–1470) and climatic implications. Geophysical Research Letters, 25, 31933196.CrossRefGoogle Scholar
Derek, F., Lagha, S., Atherton, M. P. & Cliff, R. A. (1978). The rock-forming minerals of the Jabal Archenu and Jabal ‘Awaynat alkaline ring complexes of SE Libya. In Salem, M. J. & Busweril, M. T. (eds.),The Geology of Libya, Second Symposium on the Geology of Libya, Tripoli, September 16–21. London: Academic Press, pp. 25412557.Google Scholar
Derricourt, R. (2005). Getting ‘Out of Africa’: Sea crossings, land crossings and culture in the hominin migrations. Journal of World Prehistory, 19, 119132.CrossRefGoogle Scholar
De Sélincourt, A. (2015). The World of Herodotus. London: Folio.Google Scholar
De Vos, J. H. & Virgo, K. J. (1969). Soil structure in vertisols of the Blue Nile clay plains, Sudan. Journal of Soil Science, 20, 189206.Google Scholar
Dhir, R. P., Kar, A., Wadhawan, S. K., et al. (1992). Thar Desert in Rajasthan: Land, Man and Environment. Bangalore: Geological Society of India.Google Scholar
Diaz, H. F. & Bradley, R. S. (eds.) (2004). The Hadley Circulation: Present, Past and Future. Advances in Global Change Research, 21. Dordrecht: Kluwer Academic.CrossRefGoogle Scholar
Diaz, H. F. & Markgraf, V. (eds.) (1992). El Niño: Historical and Paleoclimatic Aspects of the Southern Oscillation. Cambridge: Cambridge University Press.Google Scholar
Diaz, H. F. & Markgraf, V. (eds.) (2000). El Niño and the Southern Oscillation: Multiscale Variability and Global and Regional Impacts. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Dobson, M. (1781). An account of the Harmattan, a singular African wind. Philosophical Transactions of the Royal Society of London, 71, 4657.Google Scholar
Douglas, I. (1967). Man, vegetation and the sediment yield of rivers. Nature, 215, 925928.CrossRefGoogle Scholar
Douglas, I. (1969). The efficiency of humid tropical denudation systems. Transactions of the Institute of British Geographers, 46, 116.Google Scholar
Drake, N. A., Blench, R. M., Armitage, S. J., Bristow, C. S. & White, K. H. (2011). Ancient watercourses and biogeography of the Sahara explain the peopling of the desert. Proceedings of the National Academy of Sciences of the USA, 108, 458462.Google Scholar
Drake, N. & Bristow, C. (2006). Shorelines in the Sahara: Geomorphological evidence for an enhanced monsoon from palaeolake Megachad. The Holocene, 16, 901911.CrossRefGoogle Scholar
Dresch, J. (1959). Notes sur la géomorphologie de l’Aïr. Bulletin de l’Association de Géographes français, 280–281, 220.Google Scholar
Drover, D. P. (1966). Moisture retention in some soils of the Sudan. African Soils, 11, 483496.Google Scholar
Drummond, H. (1888). On the termite as the tropical analogue of the earth-worm. Proceedings of the Royal Society of Edinburgh, 13, 137146.Google Scholar
Dubosson, J. (2011). Cattle sacrifice in the funerary rituals of the Kingdom of Kerma: The contribution of ethnoarchaeology. Kerma, Soudan, 2010–2011. Université de Neuchâtel, Documents de la mission archéologique suisse au Soudan, 2011/3, 1824.Google Scholar
Ducassou, E., Migeon, S., Mulder, T., et al. (2009). Evolution of the Nile deep-sea turbidite system during the Late Quaternary: Influence of climate change on fan sedimentation. Sedimentology, 56, 20612090.Google Scholar
Ducassou, E., Mulder, T., Migeon, S., et al. (2008). Nile floods recorded in deep Mediterranean sediments. Quaternary Research, 70, 382391.Google Scholar
Duchaufour, P. (1978). Ecological Atlas of the Soils of the World. Translated from the French by G. R. Mehuys, C. R. De Kimpe & Y. A. Martel. New York, NY: Masson.Google Scholar
Dumont, H. J. (ed.) (2009). The Nile: Origin, Environments, Limnology and Human Use. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media.CrossRefGoogle Scholar
Dumont, H. J. & Martens, K. (1984). Dragonflies (Insecta: Odonata) from the Red Sea Hills and the main Nile in Sudan. Hydrobiologia, 110, 181190.Google Scholar
Dumont, H. J., Pensaert, J. & el Moghraby, A. I. (1984). Cladocera from the Sudan: Red Sea Hills, Jebel Marra and valley of the main Nile. Hydrobiologia, 110, 163169.Google Scholar
Dutton, A. & Lambeck, K. (2012). Ice volume and sea level during the last interglacial. Science, 337, 216219.Google Scholar
Eberz, G. W., Williams, F. M. & Williams, M. A. J. (1988). Plio-Pleistocene volcanism and sedimentary facies changes at Gadeb prehistoric site, Ethiopia. Geologische Rundschau, 77, 513527.CrossRefGoogle Scholar
Edwards, D. N. (2004). The Nubian Past: An Archaeology of the Sudan. London: Routledge.Google Scholar
Ehret, C. (1984). Historical linguistic evidence for early African food production. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press.Google Scholar
Eitel, B., Blümel, W. D., Hüser, K. & Mauz, B. (2001). Dust and loessic alluvial deposits in northwestern Namibia (Damaraland, Kaokoveld): Sedimentology and palaeoclimatic evidence based on luminescence data. Quaternary International, 76, 5765.Google Scholar
Eitel, B., Kadereit, A., Blümel, W. D., Hüser, K. & Kromer, B. (2005). The Amspoort Silts, northern Namib Desert (Namibia): Formation, age and palaeoclimatic evidence of river-end deposits. Geomorphology, 64, 299314.Google Scholar
El Badri, O. (1972). Sediment transport and deposition in the Blue Nile at Khartoum, flood seasons 1967, 1968 and 1969. Unpublished M.Sc. thesis, Department of Geology, University of Khartoum, Sudan.Google Scholar
El Baz, F., Boulos, L., Breed, C., et al. (1980). Journey to the Gilf Kebir and Uweinat, Southwest Egypt, 1978. Geographical Journal, 146, 5193.Google Scholar
El Boushi, I. M. & Abdel Salam, Y. (1982). Stratigraphy and ground-water geology of the Gezira plain, central Sudan. In Williams, M. A. J. & Adamson, D. A. (eds.), A Land between Two Niles. Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema, pp. 6580.Google Scholar
El-Shabrawy, G. M. & Dumont, H. J. (2009). The Fayum Depression and its lakes. In Dumont, H. J. (ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 95124.CrossRefGoogle Scholar
El Tahir, H. A., Ballal, M. E. & Al Fadni, O. A. (2004). A Proposed Plan of Action for Research on Desertification in the Sudan: Greater Kordofan. In Proceedings of the National Forum of Scientific Research on Desertification in the Sudan, 16–18 March, 2004, Khartoum. UNESCO Chair of Desertification Studies, University of Khartoum, Sudan, 176196.Google Scholar
El Wakeel, A. S. (2004). Strategy and methodology of research on desertification in the sudan: Biodiversity sector. In Proceedings of the National Forum of Scientific Research on Desertification in the Sudan, 16–18 March, 2004, Khartoum. UNESCO Chair of Desertification Studies, University of Khartoum, Sudan, 129150.Google Scholar
Embabi, N. S. (2004). The Geomorphology of Egypt, Vol. 1: The Nile Valley and the Western Desert. The Egyptian Geographical Society.Google Scholar
Embabi, N. S. (2008). The Geomorphology of Egypt, Vol. 2: The Eastern Desert and Sinai. The Egyptian Geographical Society.Google Scholar
Emel’yanov, E. M. (1972). Principal types of recent bottom sediments in the Mediterranean Sea: their mineralogy and geochemistry. In Stanley, D. J. (ed.), The Mediterranean Sea: A Natural Sedimentation Laboratory. Dowden: Hutchinson & Ross, pp. 355386.Google Scholar
Emery, K. O., Heezen, B. C. & Allan, T. D. (1966). Bathymetry of the eastern Mediterranean Sea. Deep-Sea Research, 13, 173192.Google Scholar
Enzel, Y., Amit, R., Crouvi, O. & Porat, N. (2010). Abrasion-derived sediments under intensified winds at the latest Pleistocene leading edge of the advancing Sinai-Negev erg. Quaternary Research, 74, 121131.Google Scholar
Enzel, Y., Kushnir, Y. & Quade, J. (2015). The middle Holocene climatic records from Arabia: Reassessing lacustrine environments, shift of ITCZ in Arabian Sea, and impacts of the southwest and African monsoons. Global and Planetary Change, 129, 6991.CrossRefGoogle Scholar
Ergenzinger, P. (1968). Vorläufige Bericht uber geomorphologische Untersuchungen im Süden des Tibestigebirges. Zeitschrift für Geomorphologie N.F., 12, 98104.Google Scholar
Fadl, A. E. (1971). A mineralogical characterization of some vertisols in the Gezira and the Kenana clay plains of the Sudan. Journal of Soil Science, 22, 129135.Google Scholar
Fahim, H. M. (1973). Egyptian Nubia after resettlement. Current Anthropology, 14, 483485.Google Scholar
Fairbridge, R. W. (1962). New radiocarbon dates of Nile sediments. Nature, 196, 108110.Google Scholar
Fairbridge, R. W. (1963). Nile sedimentation above Wadi Halfa during the last 20,000 years. Kush, XI, 96107.Google Scholar
Fairbridge, R. W. (1965). Eiszeitklima in Nordafrika. Geologische Rundschau, 54, 399414.Google Scholar
Fairbridge, R. W. (1970). World paleoclimatolgy of the Quaternary. Revue de Géographie Physique et de Géologie Dynamique (2), 12(2), 97104.Google Scholar
Fantoli, A. (ed.) (1966). Contributo alla climatologie dell’Etiopia: riassunto dei risultata e tabell meteorologiche e pluviomtriche. Rome: Ministero degli Affari Esteria.Google Scholar
FAO (Food and Agriculture Organization of the United Nations). (1969). Land and Water Resources Survey of the Jebel Marra area, Republic of Sudan, Reconnaissance and Semi-detailed Soil Surveys (by L. P. White). Report LA: SF/SUD 17.Google Scholar
Faure, H. (1959). Une hypothèse sur la structure du Ténéré (Niger). Comptes Rendus, Académie des Sciences, Paris, 254D, 44854486.Google Scholar
Faure, H. (1962a). Esquisse paléogéographique du Niger oriental depuis le Crétacé. Comptes Rendus, Académie des Sciences, Paris, 277D, 19731976.Google Scholar
Faure, H. (1962b). Reconnaissance géologique des formation-post-paléozoiques du Niger oriental. Mémoire du Bureau de Recherches Géologiques et Minières (Dakar) (1966) No. 47.Google Scholar
Faure, H. (1966). Évolution des grands lacs sahariens à l’Holocène. Quaternaria, 8, 167175.Google Scholar
Faure, H. (1969). Lacs quaternaires du Sahara. Internationale Vereiningung für theoretische und angewandte Limnologie, 17, 131146.Google Scholar
Faure, H., Manguin, E. & Nydal, R. (1963). Formations lacustres du Quaternaire supérieur du Niger oriental: diatomites et âges absolus. Bulletin BRGM (Dakar), 3, 4163.Google Scholar
Feakins, S. J., deMenocal, P. B. & Eglinton, T. I. (2005). Biomarker records of late Neogene changes in northeast African vegetation. Geology, 33, 977–80.CrossRefGoogle Scholar
Feibel, C.S. (1988). Paleoenvironments of the Koobi Fora Formation, Turkana Basin, northern Kenya. Unpublished PhD dissertation, University of Utah, Salt Lake City, Utah.Google Scholar
Feibel, C. S. (1994). Freshwater stingrays from the Plio-Pleistocene of the Turkana Basin, Kenya and Ethiopia. Lethaia. 26, 359366.Google Scholar
Fielding, L., Najman, Y., Butterworth, P., et al. (2018). The initiation and evolution of the river Nile. Earth and Planetary Science Letters, 489, 166178.Google Scholar
Fielding, L., Najman, Y., Millar, I., et al. (2016). A detrital record of the Nile River and its catchment. Journal of the Geological Society, DOI: 10.1144/jgs2016-075.CrossRefGoogle Scholar
Flaux, C., Morhange, C., Marriner, N. & Rouchy, J.-M. (2011). Bilan hydrologique et biosédimentaire de la lagune du Maryût (delta du Nil, Egypte) entre 8 000 et 3 200 ans cal. B.P. Géomorphologie: Relief, Processus, Environnement, 3, 261278.Google Scholar
Fleagle, J. G., Assefa, Z., Brown, F. H. & Shea, J. J. (2008). Paleoanthropology of the Kibish Formation, southern Ethiopia: Introduction. Journal of Human Evolution, 55, 360365.Google Scholar
Fleitmann, D., Burns, S. J., Neff, U., Mangini, A. & Matter, A. (2003a). Changing moisture sources over the last 330,000 years in Northern Oman from fluid-inclusion evidence in speleothems. Quaternary Research, 60, 223232.CrossRefGoogle Scholar
Fleitmann, D., Burns, S. J., Mangini, A., et al. (2007). Holocene ITCZ and Indian monsoon dynamics recorded in stalagmites from Oman and Yemen (Socotra). Quaternary Science Reviews, 26, 170188.Google Scholar
Fleitmann, D., Burns, S. J., Mudelsee, M., et al. (2003b). Holocene forcing of the Indian monsoon recorded in a stalagmite from southern Oman. Science, 300, 17371739.CrossRefGoogle Scholar
Fleitmann, D., Burns, S. J., Pekala, M., et al. (2011). Holocene and Pleistocene pluvial periods in Yemen, southern Arabia. Quaternary Science Reviews, 30, 783787.CrossRefGoogle Scholar
Fleitmann, D. & Matter, A. (2009). The speleothem record of climate variability in Southern Arabia. Comptes Rendus, Académie des Sciences, Paris, Geoscience, 341, 633642.CrossRefGoogle Scholar
Flenley, J. (1979). The Equatorial Rain Forest: A Geological History. London: Butterworths.Google Scholar
Flint, R. F. (1959a). On the basis of Pleistocene correlation in East Africa. Geological Magazine, 96, 265284.CrossRefGoogle Scholar
Flint, R. F. (1959b). Pleistocene climates in eastern and southern Africa. Bulletin of the Geological Society of America, 70, 343374.Google Scholar
Flohn, H. (1980). The role of the elevated heat source of the Tibetan Highlands for the large-scale atmospheric circulation (with some remarks on paleoclimatic changes). In Proceedings of Symposium on Qinghai-Xizang (Tibet) Plateau (abstracts), Beijing, China, 25 May–1 June, 1980. Beijing: Academia Sinica.Google Scholar
Fontes, J.-C., Gasse, F., Callot, Y., et al. (1985). Freshwater to marine-like environments from Holocene lakes in Northern Sahara. Nature, 317, 608610.CrossRefGoogle Scholar
Fontes, J.-C., Moussié, C., Pouchan, P. & Weidmann, M. (1973). Phases humides au Pléistocène supérieur et à l’Holocène dans le sud de l’Afar (T.F.A.I.). Comptes Rendus, Académie des Sciences, Paris, 277D, 19731976.Google Scholar
Fontes, J.-C. & Pouchan, P. (1975). Les cheminées du Lac Abhé (TFAI): stations hydroclimatiques de l’Holocène. Comptes Rendus, Académie des Sciences, Paris, 280, 383386.Google Scholar
Foucault, A. & Stanley, D. J. (1989). Late Quaternary palaeoclimatic oscillations in East Africa recorded by heavy minerals in the Nile delta. Nature, 339, 4446.Google Scholar
Fraedrich, K., Jiang, J., Gerstengarbe, F.-W. & Werner, P. C. (1997). Multiscale detection of abrupt climate changes: Application to River Nile flood levels. International Journal of Climatology, 17, 13011315.Google Scholar
Francis, P. & Oppenheimer, C. (2004). Volcanoes. Oxford: Oxford University Press.Google Scholar
Francis, P. W., Thorpe, R. S. & Ahmed, F. (1973). Setting and significance of Tertiary-Recent volcanism in the Darfur province of Western Sudan. Nature Physical Science, 243, 3032.CrossRefGoogle Scholar
Freydier, R., Michard, A., De Lange, G. &Thomson, J. (2001). Nd isotopic compositions of Eastern Mediterranean sediments: Tracers of the Nile influence during sapropel S1 formation. Marine Geology, 177, 4562.CrossRefGoogle Scholar
Fritz, H., Abdelsalam, M., Ali, K. A., et al. (2013). Orogen styles in the East African Orogen: A review of the Neoproterozoic to Cambrian tectonic evolution. Journal of African Earth Sciences, 86, 65106.CrossRefGoogle ScholarPubMed
Frumkin, A. (2009). Stable isotopes of a subfossil Tamarix tree from the Dead Sea region, Israel, and their implications for the Intermediate Bronze Age environmental crisis. Quaternary Research, 71, 319328.CrossRefGoogle Scholar
Fuller, D. (1998). Palaeoecology of the Wadi Muqadam: A preliminary report on plant and animal remains. Sudan & Nubia, 2, 5260.Google Scholar
Fuller, D. & Smith, L. (2004). The prehistory of the Bayuda: New evidence from the Wadi Muqadam. In Kendall, T. (ed.), Nubian Studies 1998: Proceedings of the Ninth Conference of the International Society of Nubian Studies. Boston: Northwestern University, pp. 265281.Google Scholar
Furon, R. (1963). The Geology of Africa. Translated by A. Hallam & L.A. Stevens. London: Oliver & Boyd.Google Scholar
Gabriel, B. (1986). Die östliche Libysche Wüste im Jungquartär. Berliner geographische Studien, 19, 1216.Google Scholar
Gaitskell, A. (1959). Gezira: A Story of Development in the Sudan. London: Faber & Faber.Google Scholar
Gallagher, J. P. (1976). Ethno-archaeology in south-central Ethiopia. In Abebe, B., Chavaillon, J. & Sutton, J. E. G. (eds.), Proceedings of the Panafrican Congress of Prehistory and Quaternary Studies VIIth session, 1971, Addis Ababa, pp. 325331.Google Scholar
Galloway, R. W. (1965a). A note on world precipitation during the last glaciation. Eiszeitalter und Gegenwart, 16, 7677.Google Scholar
Galloway, R. W. (1965b). Late Quaternary climates in Australia. Journal of Geology, 73, 603618.CrossRefGoogle Scholar
Galloway, R. W. (1970). The full-glacial climate in the southwestern United States. Annals of the Association of American Geographers, 60, 245256.Google Scholar
Gani, N. D. S., Gani, M. R. & Abdelsalam, M. G. (2007). Blue Nile incision on the Ethiopian Plateau: Pulsed plateau growth, Pliocene uplift, and hominin evolution. GSA Today, 17, 411.Google Scholar
Garcin, Y. (2006). Interactions entre l’érosion, l’hydrologie lacustre et la végétation en zone tropicale: application au basin de Masoko (Tanzanie) durant les derniers 45 000 ans. Unpublished doctoral thesis, Université Paul Cézanne, Faculté des Sciences et Techniques, 158 pp.Google Scholar
Garcin, Y., Vincens, A., Williamson, D., Buchet, G. & Guiot, J. (2007). Abrupt resumption of the African Monsoon at the Younger Dryas-Holocene climatic transition. Quaternary Science Reviews, 26, 690704.CrossRefGoogle Scholar
Garcin, Y., Vincens, A., Williamson, D., Guiot, J. & Buchet, J. (2006a). Wet phases in tropical Africa during the last glacial period. Geophysical Research Letters, 33 (L07703), 14.Google Scholar
Garcin, Y., Williamson, D., Taieb, M., et al. (2006b). Centennial to millennial changes in maar-lake deposition during the last 45, 000 years in tropical Southern Africa (Lake Masoko, Tanzania). Palaeogeography, Palaeoclimatology, Palaeoecology, 239, 334354.CrossRefGoogle Scholar
Garzanti, E., Andò, S., Padoan, M., Vezzoli, G. & El Kammar, A. (2015). The modern Nile sediment system: Processes and products. Quaternary Science Reviews, 130, 956.Google Scholar
Garzanti, E., Andò, S., Vezzoli, G., Megid, A. A. & Kammar, A. (2006). Petrology of Nile River sands (Ethiopia and Sudan): Sediment budgets and erosion patterns. Earth and Planetary Science Letters, 252, 327341.Google Scholar
Gasse, F. (1975). L’évolution des lacs de l’Afar Central (Éthiopie et TFAI) du Plio-Pléistocène à l’Actuel. D.Sc. thesis, 3 vols. University of Paris VI.Google Scholar
Gasse, F. (1976). Intérêt de l’étude des Diatomées pour la reconstitution des paléoenvironnements lacustres. Exemple des lacs d’âge Holocène de l’Afar (Éthiopie et TFAI). Revue de Géographie Physique et de Géologie Dynamique (2) 18(2–3), 199216.Google Scholar
Gasse, F. (1977). Evolution of Lake Abhé (Ethiopia and TFAI) from 70 000 B.P. Nature, 265, 4245.Google Scholar
Gasse, F. (1978). Les diatomées holocènes d’une tourbière (4 040 m) d’une montagne éthiopienne: le Mont Badda. Revue Algologique NS, 13(2), 105149.Google Scholar
Gasse, F. (1980). Les diatomées lacustres plio-pléistocènes du Gadeb (Ethiopie): Systématique, paléoécologie, biostratigraphie. Revue Algologique, 3, 1249.Google Scholar
Gasse, F. (1990). Tectonic and climatic controls on lake distribution and environments in Afar from Miocene to Present. In Katz, B. J. (ed.), Lacustrine Basin Exploration: Case Studies and Modern Analogs. American Association of Petroleum Geologists Special Volumes, M50, 1941.Google Scholar
Gasse, F. (2000a). Water resources variability in tropical and subtropical Africa in the past. In Water Resources Variability in Africa during the XXth Century. International Association of Hydrological Sciences Publication, No. 252, pp. 97105.Google Scholar
Gasse, F. (2000b). Hydrological changes in the African tropics since the Last Glacial Maximum. Quaternary Science Reviews, 19, 189211.Google Scholar
Gasse, F. (2002a). Kilimanjaro’s secrets revealed. Science, 298, 548549.Google Scholar
Gasse, F. (2002b). Diatom-inferred salinity and carbonate oxygen isotopes in Holocene waterbodies of the western Sahara and Sahel (Africa). Quaternary Science Reviews, 21, 737767.CrossRefGoogle Scholar
Gasse, F., Barker, P., Gell, P. A., Fritz, S. C. & Chalié, F. (1997). Diatom-inferred salinity in palaeolakes: An indirect tracer of climate change. Quaternary Science Reviews, 16, 547563.CrossRefGoogle Scholar
Gasse, F., Chalié, F., Vincens, A., Williams, M. A. J. & Williamson, D. (2008). Climatic patterns in equatorial and southern Africa from 30, 000 to 10, 000 years ago reconstructed from terrestrial and near-shore proxy data. Quaternary Science Reviews, 27, 23162340.Google Scholar
Gasse, F. & Fontes, J.-C. (1989). Palaeoenvironments and palaeohydrology of a tropical closed lake (Lake Asal, Djibouti) since 10,000 yr B.P. Palaeogeography, Palaeoclimatology, Palaeoecology, 69, 67102.CrossRefGoogle Scholar
Gasse, F., Juggins, S. & BenKhelifa, L. (1995). Diatom-based transfer functions for inferring past hydrochemical characteristics of African lakes. Palaeogeography, Palaeoclimatology, Palaeoecology, 117, 3154.CrossRefGoogle Scholar
Gasse, F., Richard, O., Robbe, D., Rognon, P. & Williams, M. A. J. (1980b). Évolution tectonique et climatique de l’Afar central d’après les sédiments plio-pléistocènes. Bulletin de la Société Géologique de France, (7) 22, 9871001.Google Scholar
Gasse, F. & Roberts, C. N. (2004). Late Quaternary hydrologic changes in the arid and semiarid belt of northern Africa. In Diaz, H. F. & Bradley, R. S. (eds.), The Hadley Circulation: Present, Past and Future. Dordrecht: Kluwer Academic, pp. 313345.CrossRefGoogle Scholar
Gasse, F., Rognon, P. & Street, F. A. (1980a). Quaternary history of the Afar and Ethiopian Rift lakes. In Williams, M. A. J. & Faur, H.e (eds.), The Sahara and The Nile. Rotterdam: A. A. Balkema, pp. 361400.Google Scholar
Gasse, F. & Street, F. A. (1978). Late Quaternary lake-level fluctuations and environments of the northern Rift Valley and Afar region (Ethiopia and Djibouti). Palaeogeography, Palaeoclimatology, Palaeoecology, 24, 279325.Google Scholar
Gast, M. (2000). Moissons du désert. Utilisation des ressources naturelles au Sahara central. Paris: Ibis Press.Google Scholar
Gatto, M. C. & Zerboni, A. (2015). Holocene supra-regional environmental changes as trigger for major socio-cultural processes in Northeastern Africa and the Sahara. African Archaeological Review, 32, 301333.Google Scholar
Gautier, A. (1993). The Middle Paleolithic: Archaeofaunas from Bir Tarfawi (Western Desert, Egypt). In Wendorf, F., Schild, R., Close, A. E., et al. (eds.), Egypt During the Last Interglacial: The Middle Palaeolithic of Bir Tarfawi and Bir Sahara East. New York, NY: Plenum Press, pp. 121143.CrossRefGoogle Scholar
Geological Research Authority of the Sudan. (1995). 1: 2,000,000 Geological Map of the Sudan. Khartoum, Sudan: Ministry of Mining and Energy.Google Scholar
Geological Research Authority of the Sudan & Robertson Research International Limited. (1995). Accompanying Geological Notes to the 1: 1,000,000 scale Geological Atlas of the Republic of the Sudan. Bulletin No. 40, compiled by staff of the Geological Research Authority of the Sudan and Robertson Research International Ltd. Ministry of Mining and Energy, Geological Research Authority of the Sudan, Khartoum, The Republic of the Sudan.Google Scholar
Geyh, M. & Thiedig, F. (2008). The Middle Pleistocene Al Mahrúqah Formation in the Murzuq Basin, northern Sahara, Libya: evidence for orbitally-forced humid episodes during the last 500,000 years. Palaeogeography, Palaeoclimatology, Palaeoecology, 257, 121.Google Scholar
Gibbard, P. & Cohen, K. M. (2008). Global chronostratigraphical correlation table for the last 2.7 million years. Episodes, 31, 243247.Google Scholar
Gibbard, P. L., Head, M. J., Walker, M. J. C. & the Subcommission on Quaternary Stratigraphy (2010). Formal ratification of the Quaternary System/Period and the Pleistocene Series/Epoch with a base at 2.58 Ma. Journal of Quaternary Science, 25, 96102.Google Scholar
Gibbs, R. J. (1967). The geochemistry of the Amazon River system. Part 1. The factors that control the salinity and the composition and concentration of the suspended deposits. Geological Society of America Bulletin, 78, 12031232.Google Scholar
Gifford-Gonzalez, D. (ed.) (2008). Adrar Bous: Archaeology of a Central Saharan Granitic Ring Complex in Niger. Tervuren, Belgium: Royal Museum for Central Africa.Google Scholar
Gilbert, G. K. (1914). The transportation of debris by running water. United States Geological Survey Professional Paper, 86.CrossRefGoogle Scholar
Gillan, J. A. (1918). Jebel Marra and the Deriba lakes. Sudan Notes and Records, 1, 263267.Google Scholar
Ginat, H., Zilberman, E. & Saragusti, I. (2003). Early Pleistocene lake deposits and Lower Paleolithic finds at Nahal (wadi) Zihor, Southern Negev desert, Israel. Quaternary Research, 59, 445458.Google Scholar
Glasby, P., O’Flaherty, A. & Williams, M. A. J. (2010). A geospatial visualisation of a late Pleistocene fluvial wetland surface in the Flinders Ranges, South Australia. Geomorphology, 118, 130151.Google Scholar
Glasby, P., Williams, M. A. J., McKirdy, D., Symonds, R. & Chivas, A. R. (2007). Late Pleistocene environments in the Flinders Ranges, Australia: Preliminary evidence from microfossils and stable isotopes. Quaternary Australasia, 24(2), 1928.Google Scholar
Glennie, K. W. (1970). Desert Sedimentary Environments. Developments in Sedimentology, 14. Amsterdam: Elsevier.Google Scholar
Goda, S. E., Musnad, H. A. & El Siddig, E. A. (2004). Strategy and Methodology of Research on Desertification in the Sudan: Forestry Sector. In Proceedings of the National Forum of Scientific Research on Desertification in the Sudan, 16–18 March, 2004, Khartoum. UNESCO Chair of Desertification Studies, University of Khartoum, Sudan, 7888.Google Scholar
Gopher, A., Lev-Yadun, S. & Abbo, S. (2017). Domesticating plants in the Near East. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 737742.CrossRefGoogle Scholar
Goren-Inbar, N., Alperson, N., Kislev, M. E., et al. (2004). Evidence of hominin control of fire at Gesher Benot Ya’aqov, Israel. Science, 304, 725727.Google Scholar
Goring-Morris, A. N. & Belfer-Cohen, A. (2017). The Early and Middle Epipalaeolithic of Cisjordan. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 639649.Google Scholar
Goudie, A. S. (1985). The drainage of Africa since the Cretaceous. Geomorphology, 67, 437456.Google Scholar
Goudie, A. S. (2013). Arid and Semi-Arid Geomorphology. Cambridge: Cambridge University Press.Google Scholar
Goudie, A., Atkinson, B. W., Gregory, K. J., et al. (eds.) (1985). The Encyclopaedic Dictionary of Physical Geography. Oxford: Blackwell.Google Scholar
Gowlett, J. A. J. (1984). Ascent to Civilization: The Archaeology of Early Man. London: Collins.Google Scholar
Grabham, G. W. (1917). Notes on the Geology of the Southern Gezira. Geological Survey Report 55/62, Khartoum (unpublished).Google Scholar
Grabham, G. W. (1934). Water Supplies in the Anglo-Egyptian Sudan. Sudan Geological Survey Bulletin, No. 2, 142. (Khartoum).Google Scholar
Grainger, A. (1990). The Threatening Desert: Controlling Desertification. London: Earthscan.Google Scholar
Greene, H. (1948). Soils of the Anglo-Egyptian Sudan. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 144175.Google Scholar
Greigert, J. & Pougnet, R. (1967). Essai de description des formations géologiques de la République du Niger. Mémoires du Bureau de Recherches Géologiques et Minières (Dakar) No. 48, 1236.Google Scholar
Griffin, D. L. (1999). The late Miocene climate of northeastern Africa: Unravelling the signals in the sedimentary succession. Journal of the Geological Society of London, 156, 817826.Google Scholar
Griffin, D. L. (2002). Aridity and humidity: Two aspects of the late Miocene climate of North Africa and the Mediterranean. Palaeogeography, Palaeoclimatology, Palaeoecology, 182, 6591.Google Scholar
Griffin, D. L. (2006). The late Neogene Sahabi rivers of the Sahara and their climatic and environmental implications for the Chad Basin. Journal of the Geological Society of London, 163, 905921.CrossRefGoogle Scholar
Griffin, D. L. (2011). The late Neogene Sahabi rivers of the Sahara and the hamadas of the eastern Libya-Chad border area. Palaeogeography, Palaeoclimatology, Palaeoecology, 309, 176185.Google Scholar
Griffiths, J. G. (1966). Hecataeus and Herodotus on ‘A gift of the river’. Journal of Near Eastern Studies, 25, 5761.Google Scholar
Griffiths, J. F. (1972a). Wet and dry tropics. In Griffiths, J. F. (ed.), Climates of Africa. World Survey of Climatology, Vol. 10. Amsterdam: Elsevier, pp. 221257.Google Scholar
Griffiths, J. F. (1972b). The Mediterranean Zone. In Griffiths, J. F. (ed.), Climates of Africa. World Survey of Climatology, Vol. 10. Amsterdam: Elsevier, pp. 3774.Google Scholar
Griffiths, J. F. (1972c). Ethiopian Highlands. In Griffiths, J. F. (ed.), Climates of Africa. World Survey of Climatology, Vol. 10. Amsterdam: Elsevier, pp. 369388.Google Scholar
Griffiths, J. F. & Soliman, K. H. (1972). The Northern Desert (Sahara). In Griffiths, J. F. (ed.), Climates of Africa. World Survey of Climatology, Vol. 10. Amsterdam: Elsevier, pp. 75131.Google Scholar
Grosman, L. & Munro, N. D. (2017). The Natufian Culture: The harbinger of food producing societies. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant, Cambridge: Cambridge University Press, pp. 699707.Google Scholar
Groucutt, H. S. et al. (2015). Rethinking the dispersal of Homo sapiens out of Africa. Evolutionary Anthropology, 24, 149164.Google Scholar
Grove, A. T. (1958). The ancient erg of Hausaland and similar formations on the south side of the Sahara. Geographical Journal, 124, 528533.Google Scholar
Grove, A. T. (1980). Geomorphic evolution of the Sahara and the Nile. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 716.Google Scholar
Grove, A. T. (1993). Africa’s climate in the Holocene. In Shaw, T., Sinclair, P., Andah, B. & Okpoko, A. (eds.), The Archaeology of Africa: Food, Metals and Towns. London: Routledge, pp. 3242.Google Scholar
Grove, A. T. & Pullan, R. A. (1963). Some aspects of the Pleistocene paleogeography of the Chad Basin. In Howell, F. Clark & Bourlière, F. (eds.), African Ecology and Human Evolution. Chicago: Aldine, pp. 230245.Google Scholar
Grove, A. T. & Warren, A. (1968). Quaternary landforms and climate on the south side of the Sahara. Geographical Journal, 134, 194208.Google Scholar
Grove, J. (2004). Little Ice Ages: Ancient and Modern, Vols. I and II. London: Routledge.Google Scholar
Grove, J. M. (1988). The Little Ice Age. London: Methuen.Google Scholar
Gunn, R. H. (1982). The plains of the central and southern Sudan. In Williams, M. A. J. & Adamson, D. A. (eds.), A Land between Two Niles: Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema, pp. 81109.Google Scholar
Guo, Z., Petit-Maire, N. & Kröpelin, S. (2000). Holocene non-orbital climatic events in present-day arid areas of northern Africa and China. Global and Planetary Change, 26, 97103.Google Scholar
Haaland, R. (1981). Migratory Herdsmen and Cultivating Women: The Structure of Neolithic Seasonal Adaptation in the Khartoum Nile Environment. Bergen, Mimeo, 254 pp.Google Scholar
Haaland, R. (1984). Continuity and Discontinuity: How to account for a two thousand years gap in the cultural history of the Khartoum Nile environment. Norwegian Archaeological Review, 17, 3951.Google Scholar
Haberlah, D., Glasby, P., Williams, M. A. J., et al. (2010a). ‘Of droughts and flooding rains’: an alluvial loess record from central South Australia spanning the last glacial cycle. In Bishop, P. & . Pillans, B (eds.), Australian Landscapes. Geological Society, London, Special Publications, 346, pp. 185223.Google Scholar
Haberlah, D., Williams, M. A. J., Halverson, G., et al. (2010 b). Loess and floods: high-resolution multi-proxy data of Last Glacial Maximum (LGM) slackwater deposition in the Flinders Ranges, semi-arid South Australia. Quaternary Science Reviews, 29, 26732693.CrossRefGoogle Scholar
Habicht, J. K. A. (1979). Paleoclimate, Paleomagnetism, and Continental Drift. AAPG Studies in Geology No. 9. Tulsa, OK: American Association of Petroleum Geologists.Google Scholar
Hagedorn, H. & Jäkel, D. (1969). Bemerkungen zur Quartäran Entwicklung des Reliefs im Tibesti-Gebirge (Tchad). Bulletin de l’ASEQUA, 23, 2542.Google Scholar
Hamdan, M. A. & Brook, G. A. (2015). Timing and characteristics of Late Pleistocene and Holocene wetter periods in the Eastern Desert and Sinai of Egypt based on 14C dating and stable isotope analysis of spring tufa deposits. Quaternary Science Reviews, 130, 168188.CrossRefGoogle Scholar
Hamdan, M. A., Ibrahim, M. I. A., Shiha, M. A., et al. (2016). An exploratory Early and Middle Holocene sedimentary record with palynoforms and diatoms from Faiyum Lake, Egypt. Quaternary International, http://dx.doi.org/10.1016/j.quaint.2015.12.049Google Scholar
Hamilton, A. C. (1977). An upper Pleistocene pollen diagram from highland Ethiopia. Abstracts, Xth INQUA Congress, Birmingham, UK, p. 193.Google Scholar
Hamilton, A. C. (1982). Environmental History of East Africa: A Study of the Quaternary. London: Academic Press.Google Scholar
Hammer, M. F., Karafet, T. M., Redd, A. J., et al. (2001). Hierarchical patterns of global human Y-chromosome diversity. Molecular Biology and Evolution, 18, 11891203.Google Scholar
Hammerton, D. (1966). Hot springs and fumaroles at Jebel Marra, Sudan. Nature, 209, 1290.Google Scholar
Hammerton, D. (1968). Recent discoveries in the caldera of Jebel Marra. Sudan Notes and Records, 49, 136148.Google Scholar
Hamroush, H. A. & Stanley, D. J. (1990). Paleoclimatic oscillations in East Africa interpreted by analysis of trace elements in Nile delta sediments. Episodes, 13, 264269.Google Scholar
Hamza, W. (2009). The Nile Delta. In Dumont, H. J. (ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 7594.CrossRefGoogle Scholar
Harlan, J. R. (1975). Crops & Man. Madison, WI: American Society of Agronomy, Crop Science Society of America.Google Scholar
Harrington, J. W. (1967). The first, first principles of geology. American Journal of Science, 265, 449461.Google Scholar
Harrison, J. C. (1955). An interpretation of gravity anomalies in the eastern Mediterranean. Philosophical Transactions of the Royal Society of London, A 248, 283325.Google Scholar
Harrison, M. N. & Jackson, J. K. (1958). Ecological Classification of the Vegetation of the Sudan. Forests Department Bulletin 2 (New Series), Ministry of Agriculture, Republic of Sudan.Google Scholar
Harvey, P. & Grove, A. T. (1982). A prehistoric source of the Nile. Geographical Journal, 148, 327336.Google Scholar
Hassan, F. A. (1976). Heavy minerals and the evolution of the modern Nile. Quaternary Research, 6, 425444.CrossRefGoogle Scholar
Hassan, F. A. (1980). Prehistoric settlements along the Main Nile. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and The Nile, Rotterdam: A. A. Balkema, pp. 421450.Google Scholar
Hassan, F. A. (1981). Historical Nile floods and their implications for climatic change. Science, 212, 11421145.CrossRefGoogle ScholarPubMed
Hassan, F. A. (1986a). Chronology of the Khartoum ‘Mesolithic’ and ‘Neolithic’ and related sites in the Sudan: statistical analysis and comparisons with Egypt. African Archaeological Review, 4, 83102.Google Scholar
Hassan, F. A. (1986b) Holocene lakes and prehistoric settlements of the Western Faiyum, Egypt. Journal of Archaeological Science, 13, 483501.CrossRefGoogle Scholar
Hassan, F. A. (1997). Holocene palaeoclimates of Africa. African Archaeological Review, 14, 213230.Google Scholar
Hassan, F. A., Hamdan, M. A., Flower, R. J. & Keatings, K. (2012). Oxygen and carbon isotopic records in Holocene freshwater mollusc shells from the Faiyum palaeolakes, Egypt: Palaeoenvironmental and palaeoclimatic implications. Quaternary International, 266, 175187.Google Scholar
Hassanein Bey, A. M. (1924). Through Kufra to Darfur. Geographical Journal, 64(4), 273291; 64(5), 253366.Google Scholar
Hassanein Bey, A. M. (1925). The Lost Oasis. London: Butterworths.Google Scholar
Hastenrath, S. (1974). Glaziale und periglaziale Formbildung in Hoch-Semyen, Nord-Äthiopien. Erdkunde, 28, 176186.CrossRefGoogle Scholar
Hastenrath, S. (1977). Pleistocene mountain glaciation in Ethiopia. Journal of Glaciology, 18, 309313.Google Scholar
Hawks, J. D. & Wolpoff, M. H. (2001). The four faces of Eve: Hypothesis compatibility and human origins. Quaternary International, 75, 4150.Google Scholar
Haynes, C. V. (1980). Geological evidence of pluvial climates in the Nabta area of the Western Desert of Egypt. In Wendorf, F. & Schild, R. (eds.), Prehistory of the Eastern Sahara. New York, NY: Academic Press, pp. 353–37f1.Google Scholar
Haynes, C. V., Jr. (1987). Holocene migration rates of the Sudano-Sahelian wetting front, Arba’in Desert, Eastern Sahara. In Close, A. E. (ed.), Prehistory of Arid North Africa: Essays in Honour of Fred Wendorf. Dallas, TX: Southern Methodist University, pp. 6984.Google Scholar
Haynes, C. V., Jr. (2001). Geochronology and climate change of the Pleistocene-Holocene transition in the Darb el Arba’in Desert, Eastern Sahara. Geoarchaeology, 16, 119141.3.0.CO;2-V>CrossRefGoogle Scholar
Haynes, C. V., Eyles, C. H., Pavlish, L. A., Ritchie, J. C. & Rybak, M. (1989). Holocene palaeoecology of the eastern Sahara: Selima Oasis. Quaternary Science Reviews, 8, 109136.CrossRefGoogle Scholar
Haynes, C. V. & Mead, A. B. (1987). Radiocarbon dating and paleoclimatic significance of subfossil Limicolaria in northwestern Sudan. Quaternary Research, 28, 8699.Google Scholar
Heine, K. & Heine, J. T. (2002). A paleohydrologic interpretation of the Homeb Silts, Kuiseb River, central namib Desert (Namibia) and paleoclimatic implications. Catena, 48, 107130.Google Scholar
Heinrich, H. (1988). Origin and consequences of cyclic ice-rafting in the northeast Atlantic Ocean during the last 130, 000 years. Quaternary Research, 29, 142152.CrossRefGoogle Scholar
Heiser, C. B. (1973). Seed to Civilization: The Story of Man’s Food. San Francisco, CA: W. H. Freeman.Google Scholar
Hennekam, R., Donders, T. H., Zwiep, K. & de Lange, G. J. (2015). Integral view of Holocene precipitation and vegetation changes in the Nile catchment area as inferred from its delta sediments. Quaternary Science Reviews, 130, 189199.Google Scholar
Hennekam, R., Jilbert, T., Schnetger, B. & de Lange, G. J. (2014). Solar forcing of Nile discharge and sapropel S1 formation in the early- to mid-Holocene. Paleoceanography, 29, 2013PA002553.CrossRefGoogle Scholar
Herodotus, (1954, reprinted 1960). The Histories. Translated by Aubrey de Sélincourt. Middlesex: Penguin.Google Scholar
Hershovitz, I., Weber, G. W., Quam, R., et al. (2018). The earliest modern humans outside Africa. Science, 359, 456459.Google Scholar
Hewitt, G. (2000). The genetic legacy of the Quaternary ice ages. Nature, 405, 907913.Google Scholar
Higgs, N. C., Thomson, J., Wilson, T. R. S. & Croudace, I. W. (1994). Modification and complete removal of eastern Mediterranean sapropels by postdepositional oxidation. Geology, 22, 423426.Google Scholar
Higham, T., et al. (2011). The earliest evidence for anatomically modern humans in northwestern Europe. Nature, 479, 521524.Google Scholar
Hobbs, H. F. C. (1918). Notes on Jebel Marra, Darfur. Geographical Journal, 52, 357363.Google Scholar
Hoelzmann, P. (1993a). Holozäne Limnite im NW-Sudan. PhD thesis, Freie Universität, Berlin, pp. 1191.Google Scholar
Hoelzmann, P. (1993b). Palaeoecology of Holocene lacustrine sediments in Western Nubia, SE Sahara. In Thorweihe, U. & Schandelmeier, H. (eds.), Geoscientific Research in Northeast Africa. Rotterdam: A. A. Balkema, pp. 569574.Google Scholar
Hoelzmann, P., Gasse, F., Dupont, L. M., et al. (2004). Palaeoenvironmental changes in the arid and subarid belt (Sahara-Sahel-Arabian Peninsula) from 150 kyr to Present. In Battarbee, R. W., Gasse, F. & Stickley, C. E. (eds.), Past Climate Variability through Europe and Africa. Dordrecht: Springer Science+Business Media, pp. 219256.Google Scholar
Hoelzmann, P., Jolly, D., Harrison, S. P., et al. (1998). Mid-Holocene land-surface conditions in northern Africa and the Arabian Peninsula: A data set for the analysis of biogeophysical feedbacks in the climate system. Global Biogeochemical Cycles, 12, 3551.Google Scholar
Hoelzmann, P., Keding, B., Berke, H., Kröpelin, S. & Kruse, H.-J. (2001). Environmental change and archaeology: lake evolution and human occupation in the Eastern Sahara during the Holocene. Palaeogeography, Palaeoclimatology, Palaeoecology, 169, 193217.CrossRefGoogle Scholar
Hoelzmann, P., Kruse, H.-J. & Rottinger, F. (2000). Precipitation estimates for the eastern Saharan palaeomonsoon based on a water balance model of the West Nubian Palaeolake Basin. Global and Planetary Change, 26, 105120.Google Scholar
Hofmann, C., Courtillot, V., Férard, G., et al. (1997). Timing of the Ethiopian flood basalt event and implications for plume birth and global change. Nature, 389, 838841.CrossRefGoogle Scholar
Holdaway, S. J., Phillipps, R., Koopman, A., Linseele, V. & Wendrich, W. (2017c). The L Basin archaeological record. In Holdaway, S. J. & Wendrich, W. (eds.),The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press, pp. 5197.Google Scholar
Holdaway, S. J., Phillipps, R. & Wendrich, W. (2017a). The Desert Fayum reinvestigated: The Fayum in context. In Holdaway, S. J. & Wendrich, W. (eds.),The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press, pp. 233241.Google Scholar
Holdaway, S. J. & Wendrich, W. (eds.) (2017). The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Monumenta Archaeologica, 39. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press.Google Scholar
Holdaway, S. J., Wendrich, W. & Phillipps, R. (2017b). The Desert Fayum reinvestigated: The evidence considered. In Holdaway, S. J. & Wendrich, W. (eds.),The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press, pp. 213231.Google Scholar
Holmes, J. & Hoelzmann, P. (2017). The Late Pleistocene-Holocene African Humid period as evident in lakes. Oxford Research Encyclopedia of Climate Science, 44 pp. DOI:10.1093/acrefore/9780190228620.013.531.CrossRefGoogle Scholar
Hövermann, J. (1954). Über die Höhenlage der Schneegrenze in Äthiopien und ihre Schwankungen in historischer Zeit. Nachrichten der Akademie Wissenschaften in Göttingen. Mathematisch-physikalische Klasse IIa, No. 6, 111137.Google Scholar
Hsü, K. J. (1983. The Mediterranean Was a Desert: A Voyage of the Glomar Challenger. Princeton, NJ: Princeton University Press.Google Scholar
Hsü, K. J., Montadert, L., Bernoulli, D., et al. (1977). History of the Mediterranean salinity crisis. Nature, 267, 399403.CrossRefGoogle Scholar
Hublin, J.-J., et al. (2017). New fossils from Jebel Irhoud, Morocco and the pan-African origin of Homo sapiens. Nature, 546, 289292.Google Scholar
Huckriede, R. & Venzlaff, H. (1962). Über eine pluvialzeitliche Molluskenfauna aus Kordofan (Sudan). Paläontologische Zeitschrift. Stuttgart: H. Schmidt Festband, pp. 93109.Google Scholar
Hugot, G. (1977). Un secteur du quaternaire lacustre mauritanien: Tichitt (Aouker). Éléments pour servir à une étude géomorphologique. Vol. 1. Mémoires de l’Institut mauritanien de la Recherche Scientifique (Section préhistoire), pp. 1190.Google Scholar
Hugot, H. J. (1962). Missions Berliet Ténéré-Tchad. Paris: Arts et Métiers Graphiques.Google Scholar
Hull, E. (1896). Observations on the geology of the Nile Valley and on the evidence of the greater volume of the river at a former period. Quarterly Journal of the Geological Society of London, 52, 308319.Google Scholar
Hulme, M. & Trilsbach, A. (1989). The August 1988 storm over Khartoum. Weather, 44, 8290.Google Scholar
Hunting Technical Services Limited. (1958). Jebel Marra Investigations. Report on phase I studies. Report to Ministry of Irrigation and Hydro-Electric Power and Ministry of Agriculture, Government of Republic of Sudan.Google Scholar
Hunting Technical Services Limited. (1963). Report No 1. Gezira Extension Area soil survey and land classification. Khartoum, Republic of the Sudan, Ministry of Agriculture.Google Scholar
Hunting Technical Services Limited. (1964). The White Nile east bank, Rabak to Khartoum: Soils and engineering reconnaissance. Roseires Soil Survey Report No. 6, pp. 1–128. Khartoum, Republic of the Sudan, Ministry of Agriculture.Google Scholar
Hunting Technical Services Limited. (1965). Report No 11. The White Nile east bank, Melut to Rabak: Exploratory soils and engineering survey. Khartoum, Republic of the Sudan, Ministry of Agriculture.Google Scholar
Hunting Technical Services Limited. (1974). Land use planning survey in southern Darfur. Annex 1. Soil and vegetation resources. Government of Sudan.Google Scholar
Hunting Technical Services Limited. (1976). Savanna development project phase II. Part One. Soils and geomorphology. Republic of the Sudan and FAO, Rome.Google Scholar
Huntley, D. J., Godfrey-Smith, D. I. & Thewalt, M. L. W. (1985). Optical dating of sediments. Nature, 313, 105107.Google Scholar
Hurni, H. (1982). Hochgebirge von Semien – Äthiopien. Klima und Dynamik der Höhenstufung von der letzten Kaltzeit bis zur Gegenwart. Geographica Bernensia G 13, Beiheft 7 zum Jahrbuch der Geographischen Gesellschaft von Bern.Google Scholar
Hurni, H. (1999). Sustainable management of natural resources in African and Asian mountains. Ambio, 28, 382389.Google Scholar
Hurst, H. E. (1950). The hydrology of the Sobat and White Nile and the topography of the Blue Nile and Atbara. In The Nile Basin, Vol. 8. Physical Department Paper No. 55, 195. Cairo, Egypt: Government Press.Google Scholar
Hurst, H. E. (1952). The Nile: A General Account of the River and the Utilisation of its waters, 2nd edn. London: Constable, pp. 1326.Google Scholar
Hurst, H. E. & Phillips, P. (1931). General description of the basin. In The Nile Basin, Vol. 1. Physical Department Paper. Cairo, Egypt: Government Press,Google Scholar
Hurst, H. E. & Phillips, P. (1938). The hydrology of the Lake Plateau and the Bahr el Jebel. In The Nile Basin, Vol. 5. Ministry of Public Works, Physical Department Paper No. 35. Cairo, Egypt: Schindler’s Press.Google Scholar
Ingman, M., Kaessmann, H., Pääbo, S. & Gyllensten, U. (2000). Mitochondrial genome variation and the origin of modern humans. Nature, 408, 708713.Google Scholar
Isaac, G. Ll. (1982). The earliest archaeological traces. In Clark, J. D. (ed.), The Cambridge History of Africa, Vol. I: From the Earliest Times to c. 500 BC. Cambridge: Cambridge University Press, pp. 157224.Google Scholar
Ismail, E. H. & Abdelsalam, M. G. (2012). Morpho-tectonic analysis of the Tekeze River and the Blue Nile drainage systems on the Northwestern Plateau, Ethiopia. Journal of African Earth Sciences, 69, 1447.Google Scholar
Issar, A. & Eckstein, Y. (1969). The lacustrine beds of Wadi Feiran, Sinai: their origin and significance. Israel Journal of Earth Science, 18, 2128.Google Scholar
Issawi, B. (1976). An introduction to the physiography of the Nile Valley. In Wendorf, F. & Schild, R. (eds.), Prehistory of the Nile Valley. New York, NY: Academic Press, pp. 322.Google Scholar
Issawi, B. (1981). General geology of Bir Sahara and adjacent areas. In Schild, R. & Wendorf, F. (eds.), The Prehistory of an Egyptian Oasis. Warsaw: Polish Academy of Science, pp. 1924.Google Scholar
Jackson, J. K. (1957). Changes in the climate and vegetation of the Sudan. Sudan Notes and Records, 38, 4766.Google Scholar
Jarosz, E. (1997). Tidal Dynamics in the Bab El Mandab Strait. PhD dissertation, Louisiana State University, Department of Oceanography & Coastal Sciences.Google Scholar
Jenny, H. (1941). Factors of Soil Formation. New York, NY: McGraw-Hill.Google Scholar
Jewitt, T. N. (1950). Goz soils of the Anglo-Egyptian Sudan. In Transactions 4th International Congress of Soil Science, Amsterdam, 1, pp. 285288.Google Scholar
Jewitt, T. N. (1955). Gezira Soil. Republic of Sudan Ministry of Agriculture Bulletin No. 12, Khartoum.Google Scholar
Jewitt, T. N. & Manton, J. S. (1954). Soil exhaustion in the Goz sands of the Sudan. In Transactions 5th International Congress of Soil Science, Leopoldville, pp. 413416.Google Scholar
Joffe, J. S. (1949). Pedology, 2nd edn. New Brunswick, NJ: Pedology Publications.Google Scholar
Johnson, D. L. (1993). Dynamic denudation evolution of tropical, subtropical and temperate landscapes with three-tiered soils: Toward a general theory of landscape evolution. Quaternary International, 17, 6778.CrossRefGoogle Scholar
Johnson, M. O., Mudd, S. M., Pillans, B., et al. (2014). Quantifying the rate and depth dependence of bioturbation based on optically stimulated luminescence (OSL) dates and meteoric 10Be. Earth Surface Processes and Landforms, 39, 11891198.CrossRefGoogle Scholar
Johnson, T. C. (1996). Sedimentary processes and signals of past climatic change in the large lakes of the East African Rift Valley. In Johnson, T. C. & Odada, E.O. (eds.), The Limnology, Climatology and Paleoclimatology of the East African Rift Lakes. Amsterdam: Gordon and Breach, pp. 367412.Google Scholar
Johnson, T. C., Kelts, K. & Odada, E. (2000). The Holocene history of Lake Victoria. Ambio, 29, 211.Google Scholar
Johnson, T. C. & Malala, J. O. (2009). Lake Turkana and its link to the Nile. In Dumont, H. J. (ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 287304.CrossRefGoogle Scholar
Johnson, T. C., Scholz, C. A., Talbot, M. R., et al. (1996). Late Pleistocene desiccation of Lake Victoria and rapid evolution of cichlid fishes. Science, 273, 10911093.Google Scholar
Jonglei Investigation Team. (1955). Equatorial Nile Project: The Jonglei Report. Khartoum: Sudan Government.Google Scholar
Joseph, A., Frangi, J. P. & Aranyossy, J. F. (1992). Isotope characteristics of meteoric water and groundwater in the Sahelo-Sudanian zone. Journal of Geophysical Research, 97, (No. D7), 75437551.Google Scholar
Jouzel, J., Masson-Delmotte, V., Cattani, O., et al. (2007). Orbital and millennial Antarctic climate variability over the past 800,000 years. Science, 317, 793796.CrossRefGoogle ScholarPubMed
Kalb, J. E. (1995). Fossil elephantids, Awash paleolake basins, and the Afar triple junction, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology, 114, 357368.Google Scholar
Kappelmann, J., Rasmussen, D. T., Sanders, W. J., et al. (2003). Oligocene mammals from Ethiopia and faunal exchange between Afro-Arabia and Eurasia. Nature, 426, 549552.Google Scholar
Kassas, M. (1972). Ecological consequences of water development projects. In Polunin, N. (ed.), The Environmental Future. London: Macmillan, pp. 215246.Google Scholar
Kassas, M. (1995). Desertification: A general review. Journal of Arid Environments, 30, 115128.CrossRefGoogle Scholar
Kassas, M. & Batanouny, K. H. (1984). Plant ecology. In Cloudsley-Thompson, J. L. (ed.), Key Environments: Sahara Desert. Oxford: Pergamon, pp. 7790.Google Scholar
Ke, Y. & 22 others (2001). African origin of modern humans in East Africa: A tale of 12,000 Y chromosomes. Science, 292, 11511153.Google Scholar
Kearey, P. & Vine, F. J. (1996). Global Tectonics, 2nd edn. Oxford: Blackwell.Google Scholar
Keding, B. (1993). Leiterband sites in the Wadi Howar, North Sudan. In Krzyzaniak, L., Kobusiewicz, M. & Alexander, J. (eds.), Environmental Change and Human Culture in the Nile Basin and Northern Africa until the Second Millennium B.C. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 371380.Google Scholar
Keding, B. (1998–2002). Two seasons in the Wadi Howar region (1996–1998): A preliminary report. Kush, 28, 8996.Google Scholar
el Din Hussein, Kemal (1928). L’exploration du désert de Libye. La Géographie, 40, 171183; 320336.Google Scholar
Kendrew, W. G. (1957). Climatology, Treated Mainly in Relation to Distribution in Time and Place, 2nd edn. Oxford: Oxford University Press.Google Scholar
Kendrew, W. G. (1961). Climates of the Continents. Oxford: Oxford University Press.Google Scholar
Kieffer, B., Arndt, N., Lapierre, H., et al. (2004). Flood and shield basalts from Ethiopian magmas from the African superswell. Journal of Petrology, 45, 793834.Google Scholar
Kieniewicz, J. M. & Smith, J. R. (2007). Hydrologic and climatic implications of stable isotope and minor element analyses of authigenic calcite silts and gastropod shells from a mid-Pleistocene pluvial lake, Western Desert, Egypt. Quaternary Research, 68, 431444.Google Scholar
Kimbel, W. H., Walter, R. C., Johanson, D. C., et al. (1996). Late Pliocene Homo and Oldowan tools from the Hadar formation (Kada Hadar Member), Ethiopia. Journal of Human Evolution, 31, 549561.CrossRefGoogle Scholar
Kirkby, M. J. & Chorley, R. J. (1967). Throughflow, overland flow and erosion. Bulletin of the International Association of Scientific Hydrology, 12, 521.Google Scholar
Kislev, M. E. & Simchoni, O. (2017). Early agriculture in the Southern Levant. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 733736.CrossRefGoogle Scholar
Klein, R. G. (1989). The Human Career: Human Biological and Cultural Origins. Chicago: University of Chicago Press.Google Scholar
Kozlowski, J. K. & Ginter, B. (1989). The Fayum Neolithic in the light of new discoveries. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Late Prehistory of the Nile Basin and the Sahara. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 157179.Google Scholar
Kozlowski, J. K. & Ginter, B. (1993). Holocene changes in the Fayum: Lake Moeris and the evolution of climate in Northeastern Africa. In Krzyzaniak, L., Kobusiewicz, M. & Alexander, J. (eds.), Environmental Change and Human Culture in the Nile Basin and Northern Africa until the Second Millennium B.C. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 327336.Google Scholar
Krom, M. D., Stanley, D., Cliff, R. A. & Woodward, J. C. (2002). River Nile sediment fluctuations over the past 7000 yr and their key role in sapropel development. Geology, 30, 7174.Google Scholar
Kroon, D., Little, A. M., Lourens, L. J., et al. (1998). Oxygen isotope and sapropel stratigraphy in the Eastern Mediterranean during the last 3.3 million years. In Robertson, A.H. F., Emeis, K.-C., Richter, C. & Camerlengui, A. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results, 160,181189.Google Scholar
Kröpelin, S. (1993a). Zur Rekonstruktion der spätquartären Umwelt am Unteren Wadi Howar (Südöstliche Sahara/NW Sudan). Berliner Geographische Abhandlungen, 54, 1193.Google Scholar
Kröpelin, S. (1993b). The Gilf Kebir and Lower Wadi Howar: Contrasting Early and Mid-Holocene environments in the Eastern Sahara. In Krzyzaniak, L., Kobusiewicz, M. & Alexander, J. (eds.), Environmental Change and Human Culture in the Nile Basin and Northern Africa until the Second Millennium B.C. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 249258.Google Scholar
Kröpelin, S., Verschuren, D., Lézine, A.-M., et al. (2008). Climate-driven ecosystem succession in the Sahara: The past 6000 years. Science, 320, 765768.Google Scholar
Krzyzaniak, L. (1991). Early farming in the Middle Nile Basin: Recent discoveries at Kadero (Central Sudan). Antiquity, 65, 515532.Google Scholar
Kubiëna, W. (1950). Bestimmungsbuch und Systematik der Böden Europas. Stuttgart: Ferdinand Enke Verlag.Google Scholar
Kuper, R. (1989). The Eastern Sahara from North to South: Data and dates from the B.O.S. Project. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Later Prehistroy of the Nile Basin and the Sahara. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 197203.Google Scholar
Kuper, R. & Kröpelin, S. (2006). Climate-controlled Holocene occupation in the Sahara: Motor of Africa’s evolution. Science, 313, 803807.Google Scholar
Kuper, R. & Riemer, H. (2010). Archaeological survey at western Jebel Ouenat, SE Libya. Libya Antiqua, 1, 4555.Google Scholar
Lacaille, A. D. (1951). The Stone Industry of Singa and Abu Hugar. In British Museum (Natural History) Fossil Mammals of Africa No 2, The Pleistocene Fauna of Two Blue Nile Sites, pp. 4350.Google Scholar
Lærdal, T., Talbot, M. R. & Russell, J. M. (2002). Late Quaternary sedimentation and climate in the Lakes Edward and George area, Uganda-Congo. In Odada, E. O. & Olago, D. O. (eds.), The East African Great Lakes: Limnology, Palaeolimnology and Biodiversity. Dordrecht: Kluwer Academic, pp. 429470.Google Scholar
Lamb, H. F., Bates, C. R., Coombes, P. V., et al. (2007). Late Pleistocene desiccation of Lake Tana, source of the Blue Nile. Quaternary Science Reviews, 26(3–4), 287299.Google Scholar
Lamb, H. H. (1970). Volcanic dust in the atmosphere; with a chronology and assessment of its meteorological significance. Philosophical Transactions of the Royal Society of London A, 266(1178), 426533.Google Scholar
Lamb, H. H. (1972). Climate: Present, Past and Future. Vol. 1: Fundamentals and Climate Now. London: Methuen.Google Scholar
Lamb, H. H. (1977). Climate: Present, Past and Future. Vol. 2. Climatic History and the Future. London: Methuen.Google Scholar
Lambeck, K., Purcell, A., Flemming, N. C., et al. (2011). Sea level and shoreline reconstructions for the Red Sea: Isostatic and tectonic considerations and implications for hominin migration out of Africa. Quaternary Science Reviews, 30, 35423574.Google Scholar
Lancaster, N., Kocurek, G., Singhvi, A., et al. (2002). Late Pleistocene and Holocene dune activity and wind regimes in the western Sahara Desert of Mauritania. Geology, 30, 991994.Google Scholar
Laronne, J. B. & Reid, I. (1993). Very high rates of bedload sediment transport by ephemeral desert rivers. Nature, 366, 148150.CrossRefGoogle Scholar
Larrasoaña, J. C., Roberts, A. P., Rohling, E. J., Winklhofer, M. & Wehausen, R. (2003). Three million years of monsoon variability over the northern Sahara. Climate Dynamics, 21, 689698.Google Scholar
Lawson, A. C. (1927). The Valley of the Nile. University of California Chronicle, 29.Google Scholar
Lazar, B. & Stein, M. (2011). Freshwater on the route of hominids out of Africa revealed by U-Th in Red Sea corals. Geology, 39, 10671070.Google Scholar
Leach, M. & Mearns, R. (eds.) (1996). The Lie of the Land: Challenging Received Wisdom on the African Environment. Oxford: James Currey.Google Scholar
Lebon, J. H. G. & Robertson, V. C. (1961). The Jebel Marra, Darfur, and its region. Geographical Journal, 127, 3049.CrossRefGoogle Scholar
Lee, K. E. & Wood, T. G. (1971). Termites and Soils. London: Academic Press.Google Scholar
Leier, A. L., DeCelles, P. G. & Pelletier, J. D. (2005). Mountains, monsoons and megafans. Geology, 33, 289292.CrossRefGoogle Scholar
Lemma, A. (1973). Schistosomiasis: The social challenge of controlling a man-made disease. Impact of Science on Society, 13, 133142.Google Scholar
Leopold, L. B., Wolman, M. G. & Miller, J. P. (1964). Fluvial Processes in Geomorphology. San Francisco: W. H. Freeman.Google Scholar
Le Quellec, J.-L. (2009). Les images rupestres du Jebel el-‘Uweynât. Archéo-Nil, 19, 1326.Google Scholar
Le Quellec, J.-L. (2011). Un chemin dans l’Uweynât. Sahara, 22, 149154.Google Scholar
Leroy, S. A. G. & Dupont, L. (1994). Development of vegetation and continental aridity in northwestern Africa during the Late Pliocene: The pollen record of ODP Site 658. Palaeogeography, Palaeoclimatology, Palaeoecology, 109, 295316.CrossRefGoogle Scholar
Leroy, S. A. G. & Dupont, L. M. (1997). Marine palynology of ODP Site 658 (NW Africa) and its contribution to the stratigraphy of Late Pliocene. Geobios, 30, 351359.Google Scholar
Lewis-Williams, J. D. (1981). Believing and Seeing: Symbolic Meanings in Southern San Rock Paintings. London: Academic Press.Google Scholar
Lewis-Williams, J. D. (1990). Documentation, analysis and interpretation: Dilemmas in rock art research. South African Archaeological Bulletin, 45, 126136.CrossRefGoogle Scholar
Lewis-Williams, J. D. (1991). Upper Palaeolithic art in the 1990s: A southern African perspective. South African Journal of Science, 87, 422430.Google Scholar
Lézine, A.-M. & Casanova, J. (1989). Pollen and hydrological evidence for the interpretation of past climates in tropical West Africa during the Holocene. Quaternary Science Reviews, 8, 4555.Google Scholar
Lézine, A.-M., Casanova, J. & Hillaire-Marcel, C. (1990). Across an early Holocene humid phase in western Sahara: Pollen and isotope stratigraphy. Geology, 18, 264267.Google Scholar
Lézine, A.-M., Hély, C., Grenier, C., Braconnot, P. & Krinner, G. (2011). Sahara and Sahel vulnerability to climate changes, lessons from Holocene hydrological data. Quaternary Science Reviews, 30, 30013012.Google Scholar
Linné, S. (1965). The ethnologist and the American Indian potter. In Matson, F. R. (ed.), Ceramics and Man. Viking Fund Publications in Anthropology No. 41, Wenner-Gren Foundation for Anthropological Research.Google Scholar
Linseele, V., Van Neer, W., Thys, S., et al. (2014). New archaeozoological data from the Fayum ‘Neolithic’ with a critical assessment of the evidence for early stock keeping in Egypt. PLoS ONE, DOI: 10.1371/journal.pone.0108517.Google Scholar
Linstädter, J. & Kröpelin, S. (2004). Wadi Bakht revisited: Holocene climate change and prehistoric occupation in the Gilf Kebir region of the Eastern Sahara, SW Egypt. Geoarchaeology, 19, 753778.Google Scholar
Lisiecki, L. E. & Raymo, M. E. (2005). A Pliocene-Pleistocene stack of 57 globally distributed benthic ∂18O records. Paleoceanography, 20, PA1003.Google Scholar
Lisiecki, L. E. & Raymo, M. E. (2007). Plio-Pleistocene climate evolution: Trends and transitions in glacial cycle dynamics. Quaternary Science Reviews, 26, 5669.Google Scholar
Liu, B. & Coulthard, T. J. (2014). Mapping the interactions between rivers and sand dunes: Implications for fluvial and aeolian geomorphology. Geomorphology, 231, 246257.Google Scholar
Livingstone, D. A. (1962). Age of deglaciation in the Ruwenzori range, Uganda. Nature, 194, 859860.Google Scholar
Livingstone, D. A. (1967). Postglacial vegetation of the Ruwenzori Mountains in Equatorial Africa. Ecological Monographs, 37, 2552.Google Scholar
Livingstone, D. A. (1975). Late Quaternary climatic change in Africa. Annual Review of Ecology and Systematics, 6, 249280.Google Scholar
Livingstone, D. A. (1980). Environmental changes in the Nile headwaters. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Quaternary Environments and Prehistoric Occupation in Northern Africa. Rotterdam: A. A.Balkema, pp. 339359.Google Scholar
Lombardini, E. (1865a). Saggio idrolico sul Nilo. Milan, Italy.Google Scholar
Lombardini, E. (1865b). Essai sur l’hydrologie du Nil. Paris, France.Google Scholar
Lourens, L. J., Antonarakou, A., Hilgen, F. J., et al. (1996). Evaluation of the Plio-Pleistocene astronomical timescale. Paleoceanography, 11, 391413.Google Scholar
Low, J. (1811). Egypt from the best authorities. Map published in the New and Complete American Encyclopedia or Universal Dictionary of Arts and Sciences, by E. Low, successor to John Low, New York, 1811.Google Scholar
Lynes, H. (1921). Notes on the natural history of Jebel Marra. Sudan Notes and Records, 4, 119137.Google Scholar
Lynes, H. & Campbell Smith, W. (1921). Preliminary note on the rocks of Darfur. Geological Magazine, 58, 206215.Google Scholar
Lyons, H. G. (1906). The Physiography of the River Nile and its Basin. Cairo, Egypt: National Printing Department.Google Scholar
MacEachern, S. (2000). Genes, tribes and African history. Current Anthropology, 41, 357384.CrossRefGoogle ScholarPubMed
Macgregor, D. S. (2012). The development of the Nile drainage system: Integration of onshore and offshore evidence. Petroleum Geoscience, 18, 417431.Google Scholar
Macklin, M. G., Lewin, J. & Woodward, J. C. (2012). The fluvial record of climate change. In Vita-Finzi, C. (ed.), River History. Philosophical Transactions of the Royal Society of London, Series A, 370, 21432172.Google Scholar
Macklin, M. G., Toonen, W. H. J., Woodward, J. C., et al. (2015). A new model of river dynamics, hydroclimatic change and human settlement in the Nile Valley derived from meta-analysis of the Holocene fluvial archive. Quaternary Science Reviews, 130, 109123.Google Scholar
Macklin, M. G., Woodward, J. C., Welsby, D. A., et al. (2013). Reach-scale river dynamics moderate the impact of rapid Holocene climate change on floodwater farming in the desert Nile. Geology, 41, 695698.CrossRefGoogle Scholar
Mahaney, W. C. (1990). Ice on the Equator: Quaternary Geology of Mount Kenya, East Africa. Sister Bay, WI: Wm Caxton.Google Scholar
Mainguet, M. (1994). Desertification: Natural Background and Human Mismanagement. 2nd edn. Berlin: Springer-Verlag.Google Scholar
Mainguet, M. & Canon, L. (1976). Vents et paléovents du Sahara. Tentative d’approche paléoclimatique. Revue de Géographie Physique et de Géologie Dynamique (2), 18(2–3), 241250.Google Scholar
Mainguet, M., Canon, L. & Chemin, M. C. (1980). Le Sahara: géomorphologie et paléogéomorphologie éoliennes. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 1735.Google Scholar
Makled, W. A. & Mandur, M. M. M. (2016). Nannoplankton calendar: Application of nannoplankton biochronology in sequence stratigraphy and basin analysis in the subsurface offshore Nile Delta, Egypt. Marine and Petroleum Geology, 72, 374392.Google Scholar
Maley, J. (1980). Les changements climatiques de la fin du Tertiaire en Afrique: leur conséquence sur l’apparition du Sahara et de sa végétation. In Williams, M. A. J. & Faure, H. (eds.),The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 6386.Google Scholar
Maley, J. (1981). Études palynologiques dans le bassin du Tchad et paléoclimatologie de l’Afrique Nord-tropicale de 30 000 ans à l’époque actuelle. DSc thesis, University of Montpellier.Google Scholar
Maley, J. (1982). Dust, clouds, rain types, and climatic variations in tropical north Africa. Quaternary Research, 18, 116.CrossRefGoogle Scholar
Maley, J. (1996). The African rain forest: Main characteristics of changes in vegetation and climate from the Upper Cretaceous to the Quaternary. Proceedings of the Royal Society of Edinburgh, 104B, 3173.Google Scholar
Maley, J. (2000). Last Glacial Maximum lacustrine and fluviatile formations in the Tibesti and other Saharan mountains, and large-scale climatic teleconnections linked to the activity of the Subtropical Jet Stream. Global and Planetary Change, 26, 121136.Google Scholar
Maley, J. (2004). Le bassin du Tchad au Quaternaire récent: formations sédimentaires, paléoenvironnements et préhistoire. La question des Paléotchads. In Renault-Miskovsky, J. & Semah, A.-M. (eds.), Guide de la Préhistoire mondiale. Paris: Artcom, pp. 179217.Google Scholar
Maley, J. (2010). Climate and palaeoenvironment evolution in north tropical Africa from the end of the Tertiary to the Upper Quaternary. Palaeoecology of Africa, 30, 227278.Google Scholar
Maley, J., Cohen, J., Faure, H., Rognon, P. & Vincent, P. M. (1970). Quelques formations lacustres et fluviatiles associées a différentes phases du volcanisme au Tibesti (nord du Tchad). Cahiers ORSTOM série Géologie, 2, 127152.Google Scholar
Mallinson, M. D. S. (1997). Preliminary report: SARS Survey from Omdurman to Gabolab 1997. Sudan & Nubia, 1, 2527.Google Scholar
Mallinson, M. D. S. (1998). The SARS Survey from Omdurman to Gabolab 1997, The Survey. Sudan & Nubia, 2, 4245.Google Scholar
Marinova, E., Linseele, V. & Vermeersch, P. (2008). Holocene environments and subsistence patterns near the Tree Shelter, Red Sea Mountains, Egypt. Quaternary Research, 70, 392397.CrossRefGoogle Scholar
Marinova, M. M., Meckler, A. N. & McKay, C. P. (2014). Holocene freshwater carbonate structures in the hyper-arid Gebel Uweinat region of the Sahara Desert (Southwestern Egypt). Journal of African Earth Sciences, 89, 5055.Google Scholar
Maritan, L., Iacumin, P., Zerboni, A., et al. (2018). Fish and salt: The successful recipe of White Nile hunter-gatherer-fishers. Journal of Archaeological Science, 92, 4862.Google Scholar
Mark, B. G. & Osmaston, H. A. (2008). Quaternary glaciation in Africa: Key chronologies and climatic implications. Journal of Quaternary Science, 23, 589608.Google Scholar
Marks, A. E. (1989). The later prehistory of the Central Nile Valley: A view from its Eastern Hinterlands. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Late Prehistory of the Nile Basin and the Sahara. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 443450.Google Scholar
Marks, A. E. (1993). Climatic and cultural changes in the Southern Atbai, Sudan, from the fifth through the third millennium B.C. In Krzyzaniak, L., Kobusiewicz, M. & Alexander, J. (eds.), Environmental Change and Human Culture in the Nile Basin and Northern Africa until the Second Millennium B.C. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 431438.Google Scholar
Marks, A. E. & Fattovich, R. (1989). The later prehistory of the Eastern Sudan: A preliminary view. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Late Prehistroy of the Nile Basin and the Sahara. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 451458.Google Scholar
Marriner, N., Flaux, C., Kaniewski, D., et al. (2012b). ITCZ and ENSO-like modulation of Nile delta hydro-geomorphology during the Holocene. Quaternary Science Reviews, 45, 7384.Google Scholar
Marriner, N., Flaux, C., Morhange, C. & Kaniewski, D. (2012a). Nile Delta’s sinking past: Quantifiable links with Holocene compaction and climate-driven changes in sediment supply? Geology, 40, 10831086.CrossRefGoogle Scholar
Marshall, M. H., Lamb, H. F., Huws, D., et al. (2011). Late Pleistocene and Holocene drought events at Lake Tana, the source of the Blue Nile. Global and Planetary Change, 78, 147161.CrossRefGoogle Scholar
Martens, K. (1984). On the freshwater ostracods (Crustacea, Ostracoda) of the Sudan, with special reference to the Red Sea Hills, including a description of a new species. Hydrobiologia, 110, 137161.Google Scholar
Matthews, A., Ayalon, A. & Bar-Matthews, M. (2000). D/H ratios of fluid inclusions of Soreq cave (Israel) speleothems as a guide to the Eastern Mediterranean Meteoric Line relationships in the last 120 ky. Chemical Geology, 166, 183191.CrossRefGoogle Scholar
Mawson, R. & Williams, M. A. J. (1984). A wetter climate in eastern Sudan 2, 000 years ago? Nature, 309, 4951.CrossRefGoogle Scholar
Maxwell, T. A., Issawi, B. & Haynes, C. V., Jr. (2010). Evidence for Pleistocene lakes in the Tushka region, south Egypt. Geology, 38, 11351138.CrossRefGoogle Scholar
May, J.-H., Wells, S. G., Cohen, T. J., et al. (2015). A soil chronosequence on Lake Mega-Frome beach ridges and its implications for late Quaternary pedogenesis and paleoenvironmental conditions in the drylands of southern Australia. Quaternary Research, 83, 150165.Google Scholar
Mayewski, P. A., Rohling, E. E., Stager, J. C., et al. (2004). Holocene climate variability. Quaternary Research, 62, 243255.Google Scholar
McBrearty, S. (1990). Consider the humble termite: Termites as agents of post-depositional disturbance at African archaeological sites. Journal of Archaeological Science, 17, 111143.Google Scholar
McBurney, C. B. M. & Hey, R. W. (1955). Prehistory and Pleistocene Geology in Cyrenaican Libya. Cambridge: Cambridge University Press.Google Scholar
McCann, J. C. (1999). Green Land, Brown Land, Black Land: An Environmental History of Africa, 1800–1990. New Haven, CT: Heinemann and Oxford: James Currey.Google Scholar
McCauley, J. F., Schaber, G. G., Breed, C. S., et al. (1982). Subsurface valleys and geoarchaeology of the eastern Sahara revealed by Shuttle Radar. Science, 218, 10041020.Google Scholar
McCauley, J. F., Schaber, G. G., McHugh, W. P., et al. (1986). Paleodrainages of the eastern Sahara: The radar rivers revisited (SIR-A/B implications for a mid-Tertiary trans-African drainage system. Institute of Electrical and Electronics Engineers Special Volume GE-24, 624648.Google Scholar
McDougall, I., Brown, F. H. & Fleagle, J. G. (2005). Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature, 433, 733736.Google Scholar
McDougall, I., Brown, F. H. & Fleagle, J. G. (2008). Sapropels and the age of hominins in Omo I and II, Kibish, Ethiopia. Journal of Human Evolution, 55, 409420.CrossRefGoogle ScholarPubMed
McDougall, I., Morton, W. H. & Williams, M. A. J. (1975). Age and rates of denudation of Trap Series basalts at Blue Nile gorge, Ethiopia. Nature, 254, 207209.Google Scholar
McEvedy, R. (1992). Late Cainozoic environments in central Sudan: A study in images and their interpretation. Unpublished MSc thesis, Macquarie University, Sydney.Google Scholar
McGarry, S., Bar-Matthews, M., Matthews, A., et al. (2004). Constraints on hydrological and paleotemperature variations in the Eastern Mediterranean region in the last 140 ka given by the ∂D values of speleothem fluid inclusions. Quaternary Science Reviews, 23, 919934.Google Scholar
McGee, D., Broecker, W. S. & Winckler, G. (2010). Gustiness: The driver of glacial dustiness? Quaternary Science Reviews, 29, 23402350.CrossRefGoogle Scholar
McHugh, W. P. (1980). Archaeological sites of the Gilf Kebir. In Baz, El et al., Journey to the Gilf Kebir and ‘Uweinat, Southwest Egypt, 1978. Geographical Journal, 146, 6468.Google Scholar
McHugh, W. P., Breed, C. S., Schaber, G. G., McCauley, J. F. & Szabo, B. J. (1988). Acheulian sites along the ‘radar rivers’, southern Egyptian Sahara. Journal of Field Archaeology, 15, 361379.Google Scholar
McHugh, W. P., Schaber, G. G., Breed, C. S. & McCauley, J. F. (1989). Neolithic adaptation and the Holocene functioning of Tertiary palaeodrainages in southern Egypt and northern Sudan. Antiquity, 63, 320336.Google Scholar
McKenzie, J. A. (1993). Pluvial conditions in the eastern Sahara following the penultimate deglaciation: Implications for changes in atmospheric circulations with global warming. Palaeogeography, Palaeoclimatology, Palaeoecology, 103, 95105.Google Scholar
McPherron, S. P., Alemseged, Z., Marean, C. W., et al. (2010). Evidence for stone-tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature, 466, 857860.Google Scholar
McTainsh, G. H. (1980). Harmattan dust deposition in northern Nigeria. Nature, 286, 587588.CrossRefGoogle Scholar
McTainsh, G. H. (1984). The nature and origin of the aeolian mantles of central northern Nigeria. Geoderma, 33, 1337.CrossRefGoogle Scholar
McTainsh, G. (1987). Desert loess in northern Nigeria. Zeitschrift für Geomorphologie N.F., 31, 145165.CrossRefGoogle Scholar
McTainsh, G. H. & Walker, P. H. (1982). Nature and distribution of Harmattan dust. Zeitschrift für Geomorphologie N.F., 26, 417436.Google Scholar
Meehan, B. (1991). Wetland hunters: Some reflections. In Haynes, C. D., Ridpath, M. G. & Williams, M. A. J. (eds.), Monsoonal Australia: Landscape, Ecology and Man in the Northern Lowlands. Rotterdam: A. A. Balkema, pp. 197206.Google Scholar
Mellars, P. (2006). Going East: New genetic and archaeological perspectives on the modern human colonization of Eurasia. Science, 313, 796800.Google Scholar
Menardi Noguera, A. & Zboray, A. (2011). Rock art in the landscape setting of the western Jebel Uweinat (Libya). Sahara, 22, 85116.Google Scholar
Mercier, N., Valladas, H., Bar-Yosef, O., Vandermeersch, B. & Joron, J.-L. (1993). Thermoluminescene date for the Mousterian burial site of Es-Skhul, Mt. Carmel. Journal of Archaeological Science, 20, 169174.CrossRefGoogle Scholar
Mercone, D., Thomson, J., Abu-Zied, R. H., Croudace, I. W. & Rohling, E. J. (2001). High-resolution geochemical and micropalaeontological profiling of the most recent eastern Mediterranean sapropel. Marine Geology, 177, 2544.Google Scholar
Merla, G. (1963). Missione geologica nell’etiopia meridionale del Consiglio Nazionale delle Richerche 1959–1960. Notizie geomorfologiche e geologiche. Annales Museo Geologica Bologna, 31, 156.Google Scholar
Messerli, B. (1972). Formen und Formungsprozesse in der Hochgebirgsregion des Tibesti. Hochgebirgsforschung, 2, 2886.Google Scholar
Messerli, B. & Winiger, M. (1992). Climate, environmental change, and resources of the African mountains from the Mediterranean to the Equator. Mountain Research and Development, 12, 315336.Google Scholar
Messerli, B., Winiger, M. & Rognon, P. (1980). The Saharan and East African uplands during the Quaternary. In Williams, M. A. J. & Faure, H. (eds.),The Sahara and the Nile. Quaternary Environments and Prehistoric Occupation in Northern Africa. Rotterdam: A. A. Balkema, pp. 87132.Google Scholar
Meyer, A., Montero, C. M. & Spreinat, A. (1996). Molecular phylogenetic inferences about the evolutionary history of East African cichlid fish radiations. In Johnson, T.C. & Odada, E. O. (eds.), The Limnology, Climatology and Paleoclimatology of the East African Rift Lakes. Amsterdam: Gordon and Breach, pp. 303323.Google Scholar
Milliman, J. D. (1997). Fluvial sediment discharge to the sea and the importance of regional tectonics. In Ruddiman, W. F. (ed.), Tectonic Uplift and Climate Change. New York, NY: Plenum Press, pp. 239257.Google Scholar
Milliman, J. D. & Meade, R. H. (1983). World-wide delivery of river sediment to the oceans. Journal of Geology, 91, 121.Google Scholar
Milne, G. (1935). Some suggested units of classification and mapping, particularly for East African soils. Soil Research, 4, 183198.Google Scholar
Milne, G. (1936). Normal erosion as a factor in soil profile development. Nature, 138, 548549.Google Scholar
Minucci, E. (1938). Richerche geologiche nella regione del Semien. In Missione di studio al Lago Tana. Reale Accademia d’Italia, 1, 3746.Google Scholar
Miserey, Y. (2008). Comment le Sahara est devenu un désert. Le Figaro, 14 May 2008, p. 11.Google Scholar
Mitchell, C. W. (1984). Soils. In Cloudsley-Thompson, J. L. (ed.), Key Environments: Sahara Desert. Oxford: Pergamon, pp. 4155.Google Scholar
Mix, A. C., Bard, E. & Schneider, R. (2001). Environmental processes of the ice age: Land, oceans, glaciers (EPILOG). Quaternary Science Reviews, 20, 627657.Google Scholar
Mo, K., Bell, G. D. & Thiaw, W. M. (2001). Impact of sea surface temperature anomalies on the Atlantic tropical storm activity and West African rainfall. Journal of the Atmospheric Sciences, 58, 34773496.2.0.CO;2>CrossRefGoogle Scholar
Moernaut, J., Verschuren, D., Charlet, F., Kristen, I., Fagot, M. & De Batist, M. (2010). The seismic-stratigraphic record of lake-level fluctuations in Lake Challa: Hydrological stability and change in equatorial Africa over the last 140 kyr. Earth and Planetary Science Letters, 290, 214223.Google Scholar
Moeyersons, J. (1978). The behaviour of stones and stone implements buried in consolidating and creeping Kalahari Sands. Earth Surface Processes, 3, 115128.Google Scholar
Moeyersons, J., Vermeersch, P. M., Beeckman, H. & Van Peer, P. (1999). Holocene environmental changes in the Gebel Umm Hammad, Eastern Desert, Egypt. Geomorphology, 26, 297312.Google Scholar
Mohamed, Y. A., El Sammani, M. O. & Shadad, M. Z. (1982). North Kordofan Rural Water Supply Baseline Survey. Sudan: Institute of Environmental Studies, University of Khartoum.Google Scholar
Mohamed, Y. A., Van den Hurk, B. J. J. M., Savenije, H. H. G. & Bastiaanssen, W. G. M. (2005). Hydroclimatology of the Nile: Results from a Regional Climate Model. Hydrology and Earth System Sciences Discussions, 2(1), 319364.Google Scholar
Mohammed-Ali, A. & Marks, A. E. (1984). The prehistory of Shaqadud in the Western Butana, Central Sudan: A preliminary report. Norwegian Archaeological Review, 17, 5259.CrossRefGoogle Scholar
Mohr, E. C. J. & van Baren, F. A. (1959). Tropical Soils: A Critical Study of Soil Genesis as Related to Climate, Rock and Vegetation. London: Interscience Publishers.Google Scholar
Mohr, E. C. J., van Baren, F. A. & van Schuylenborgh, J. (1972). Tropical Soils: A Comprehensive Study of Their Genesis. The Hague: Mouton.Google Scholar
Mohr, P. (1968). The Cenozoic volcanic succession in Ethiopia. Bulletin Volcanologique, 32, 514.Google Scholar
Mohr, P. (1974). Mapping of the major structures of the African rift system from ERTS-1 imagery. Smithsonian Astrophysical Observatory Special Report, 361, 170.Google Scholar
Monod, T. (1963). The Late Tertiary and Pleistocene in the Sahara. In Howell, F. Clark & Bourlière, F. (eds.), African Ecology and Human Evolution. Chicago: Aldine, pp. 116229.Google Scholar
Moorehead, A. (1971). The White Nile. London: Hamish Hamilton.Google Scholar
Moorehead, A. (1972). The Blue Nile. London: Harper & Row.Google Scholar
Morales, C. (ed.) (1979). Saharan Dust: Mobilization, Transport, Deposition. Chichester, John Wiley & Sons.Google Scholar
Morcos, S. A. & Messieh, S. N. (1973). Change in the current regime in the Suez Canal after construction of the Aswan High Dam. Nature, 242, 3839.Google Scholar
Moreau, R. E. (1963). The distribution of tropical African birds as an indicator of past climatic changes. In Howell, F.Clark & Bourlière, F. (eds.), African Ecology and Human Evolution. Chicago: Aldine, pp. 2842.Google Scholar
Moussa, A., et al. (2016). Lake Chad sedimentation and environments during the late Miocene and Pliocene: New evidence from mineralogy and chemistry of the Bol core sediments. Journal of African Earth Sciences, 118, 192204.Google Scholar
Moy, C. M., Seltzer, G. O., Rodbell, D. T. & Anderson, D. M. (2002). Variability of El Niño/ Southern Oscillation activity at millennial timescales during the Holocene epoch. Nature, 420, 162165.Google Scholar
Muchane, M. W. (1996). Comparison of the isotope record in micrite, Lake Turkana, with the historical weather record over the last century. In Johnson, T.C. & Odada, E. O. (eds.),The Limnology, Climatology and Paleoclimatology of the East African Rift Lakes. Amsterdam: Gordon and Breach, pp. 431441.Google Scholar
Mukherjee, R., Rao, C. R. & Trevor, J. C. (1955). The Ancient Inhabitants of Jebel Moya (Sudan). Oxford: Oxford University Press.Google Scholar
Muriel, C. E. (1901). Report on the Forests of the Sudan. Cairo, Egypt: Government Printer.Google Scholar
Murphy, J. O. & Whetton, P. H. (1989). A re-analysis of a tree ring chronology from Java. Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen, Series B, 92(3), 241257.Google Scholar
Museen der Stadt Köln (1978). Sahara: 10 000 Jahre zwischen Weide und Wüste. Museen der Stadt Köln.Google Scholar
Mussi, M., Caneva, I. & Zarattini, A. (1984). More on the terminal Palaeolithic of the Fayum Depression. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Origin and Early Food Development of Food-Producing Cultures in North-Eastern Africa. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 185191.Google Scholar
Mustafa, M. A. & Mahdi, A. A. (2004). Strategy and plan of action of scientific research on desertification. In Proceedings of the National Forum of Scientific Research on Desertification in the Sudan, 16–18 March, 2004, Khartoum. UNESCO Chair of Desertification Studies, University of Khartoum, Sudan.Google Scholar
Mustafa, M. A. & Saeed, A. B. (2004). Strategy and methodology of research on desertification in the Sudan: Soil and water conservation sector. In Proceedings of the National Forum of Scientific Research on Desertification in the Sudan, 16–18 March, 2004, Khartoum. UNESCO Chair of Desertification Studies, University of Khartoum, Sudan, pp. 36–58.Google Scholar
Muzzolini, A. (1995). Les images rupestres du Sahara. Toulouse: Muzzolini.Google Scholar
Nachtigal, G. (1889). Sahara and Sudan. IV. Wadai and Darfur (1889). Translated from the original German by Allan G. B. Fisher and Humphrey J. Fisher. London: C. Hurst & Co.Google Scholar
National Research Council, Board on Science and Technology for International Development. (1996). Lost Crops of Africa, Vol. 1: Grains. Washington, DC: National Academies Press.Google Scholar
Newbold, D. (1924). A desert odyssey of a thousand miles. Sudan Notes and Records, 7, 4392, 104107.Google Scholar
Newbold, D. (1928). Rock pictures and archaeology in the Libyan Desert. Antiquity, 2, 261291.Google Scholar
Newbold, D. & Shaw, W. B. K. (1928). An exploration of the South Libyan Desert. Sudan Notes and Records, 11, 103194.Google Scholar
Ngomanda, A., Neumann, K., Schweizer, A. & Maley, J. (2009). Seasonality change and the third millennium BP rainforest crisis in southern Cameroon (Central Africa). Quaternary Research, 71, 307318.Google Scholar
Nichol, J. E. (1991). The extent of desert dunes in northern Nigeria as shown by image enhancement. Geographical Journal, 157, 1324.Google Scholar
Nichol, J. E. (1999). Geomorphological evidence and Pleistocene refugia in Africa. Geographical Journal, 165, 7989.Google Scholar
Nicholson, S. E. (1996). A review of climate dynamics and climate variability in eastern Africa. In Johnson, T. C., & Odada, E. O. (eds.),The Limnology, Climatology and Paleoclimatology of the East Afrfican Lakes. Amsterdam, Gordon and Breach, pp. 2556.Google Scholar
Nicholson, S. E. (2011). Dryland Climatology. Cambridge: Cambridge University Press.CrossRefGoogle Scholar
Nicholson, S. E. & Selato, J. C. (2000). The influence of La Niña on African rainfall. International Journal of Climatology, 20, 17611776.Google Scholar
Nicoll, K. & Sallam, E. S. (2016). Paleospring tufa deposition in the Kurkur Oasis region and implications for tributary integration with the River Nile in southern Egypt. Journal of African Earth Sciences, DOI: 10.1016/jafrearsci.2016.10.014Google Scholar
Nielsen, E. (1973). Coastal erosion in the Nile delta. Nature and Resources, 11, 1418.Google Scholar
Nile Basin Initiative (2012). State of the River Nile Basin 2012. Nile Basin Initiative, Entebbe, Uganda.Google Scholar
Nilsson, E. (1931). Quaternary glaciations and pluvial lakes in British East Africa. Geografiska Annaler, A13, 249349.Google Scholar
Nilsson, E. (1935). Traces of ancient changes of climate in East Africa. Geografiska Annaler, 1/2, 121.Google Scholar
Nilsson, E. (1940). Ancient changes of climate in British East Africa and Abyssinia: A study of ancient lakes and glaciers. Geografiska Annaler, A 22, 179.Google Scholar
Nilsson, E. (1949). The pluvials of East Africa: An attempt to correlate Pleistocene changes of climate. Geografiska Annaler, A 31, 204211.Google Scholar
Nye, P. H. (1954). Some soil-forming processes in the humid tropics. I. A field study of a catena in the West African forest. Journal of Soil Science, 5, 721.CrossRefGoogle Scholar
Nye, P. H. (1955). Some soil-forming processes in the humid tropics. IV. The action of the soil fauna. Journal of Soil Science, 6, 7383.Google Scholar
Nyssen, J., Poesen, J., Moeyersons, J., Deckers, J., Haile, M. & Lang, A. (2004). Human impact on the environment in the Ethiopian and Eritrean highlands: A state of the art. Earth-Science Reviews, 64, 273320.Google Scholar
Obeid, M. M., Bari, E. A., Wickens, G. E. & Williams, M. A. J. (1982). The vegetation of the central Sudan. In Williams, M. A. J. & Adamson, D. A. (eds.), A Land between Two Niles: Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema, pp. 143164.Google Scholar
O’Connell, J. F. & Allen, J. (2015). The process, biotic impact, and global implications of the human colonization of Sahul about 47,000 years ago. Journal of Archaeological Science, 56, 7384.Google Scholar
Olivieri, A., Achilli, A., Pala, M., et al. (2006). The mtDNA legacy of the Levantine early Upper Palaeolithic in Africa. Science, 315, 17671770.Google Scholar
Ortlieb, L. (2004). Historical chronology of ENSO and the Nile flood record. In Battarbee, R. W., Gasse, F. & Stickley, C. E. (eds.), Past Climate Variability through Europe and Africa. Dordrecht: Springer Science+Business Media, pp. 257278.Google Scholar
Osborne, A. H., Vance, D., Rohling, E. J., Barton, N., Rogerson, M. & Fello, N. (2008). A humid corridor across the Sahara for the migration of early modern humans out of Africa 120,000 years ago. Proceedings of the National Academy of Sciences of the USA, 105, 1644416447.Google Scholar
Osman, O. E. & Hastenrath, S. L. (1969). On the synoptic climatology of summer rainfall over central Sudan. Archiv fur Meteorologie, Geophysik und Bioklimatologie, Series B, 17, 297324.Google Scholar
Osman, Y. Z. & Shamseldin, A. Y. (2002). Qualitative rainfall prediction models for central and southern Sudan using El Niño-Southern Oscillation and Indian Ocean sea-surface temperature indices. International Journal of Climatology, 22, 1861–1878.Google Scholar
Osmaston, H. A., Mitchell, W. A. & Osmaston, J. A. N. (2005). Quaternary glaciation of the Bale Mountains, Ethiopia. Journal of Quaternary Science, 20, 593606.Google Scholar
Osmond, J. K. & Dabous, A. A. (2004). Timing and intensity of groundwater movement during Egyptian Sahara pluvial periods by U-series analysis of secondary U in ores and carbonates. Quaternary Research, 61, 8594.Google Scholar
Otero, O., Lecuyer, C., Fourel, F., Martineau, F., Mackaye, H. T., Vignaud, P. & Brunet, M. (2011). Freshwater fish ∂18O indicates a Messinian change of the precipitation regime in Central Africa. Geology, 39, 435438.Google Scholar
Otto-Bliesner, B. L., Russell, J. M., Clark, P. U., et al. (2014). Coherent changes of southeastern equatorial and northern African rainfall during the last deglaciation. Science, 346, 12231227.Google Scholar
Owen, H. G. (1983). Atlas of Continental Displacement, 200 million Years to the Present. Cambridge: Cambridge University Press.Google Scholar
Owen, R. B., Potts, R., Behrensmeyer, A. K. & Ditchfield, P. (2008). Diatomaceous sediments and environmental change in the Pleistocene Olorgesailie Formation, southern Kenya Rift Valley. Palaeogeography, Palaeoeclimatology, Palaeoecology, 269, 1737.CrossRefGoogle Scholar
Özdoğan, M. (2017). The archaeology of early farming in Southeast Turkey. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 723731.Google Scholar
Ozenda, P. (1977). Flore du Sahara. 2ème édition. Paris, Editions du Centre National de la Recherche Scientifique, 622 pp.Google Scholar
Pachur, H.-J. (1997). Der Ptolemäus-See in Westnubien als Paläoklimaindikator. Petermanns Geographische Mitteilungen, 141, 227250.Google Scholar
Pachur, H.-J. & Altmann, N. (2006). Die Ostsahara im Spätquartär: Ökosystemwandel im gröten hyperariden Raum der Erde. Berlin: Springer.Google Scholar
Pachur, H.-J. & Hoelzmann, P. (1991). Paleoclimatic implications of late Quaternary lacustrine sediments in Western Nubia, Sudan. Quaternary Research, 36, 257276.CrossRefGoogle Scholar
Pachur, H.-J. & Hoelzmann, P. (2000). Late Quaternary palaeoecology and palaeoclimates of the Eastern Sahara. Journal of African Earth Sciences, 30, 929939.Google Scholar
Pachur, H.-J. & Kröpelin, S. (1987). Wadi Howar: Paleoclimatic evidence from an extinct river system in the southeastern Sahara. Science, 237, 298300.CrossRefGoogle ScholarPubMed
Pachur, H.-J., Kröpelin, S., Hoelzmann, P., Goschin, M. & Altmann, N. (1990). Late Quaternary fluvio-lacustrine environments of western Nubia. Berliner Geowissentschaftliche Abhandlungen, A120(1), 203260.Google Scholar
Pachur, H.-J. & Rottinger, F. (1997). Evidence for a large extended paleolake in the Eastern Sahara as revealed by spaceborne radar lab images. Remote Sensing of Environment, 61, 437440.Google Scholar
Padoan, M., Garzanti, E., Harlavan, Y. & Villa, I. M. (2011). Tracing Nile sediment sources by Sr and Nd isotope signatures (Uganda, Ethiopia, Sudan). Geochimica et Cosmochimica Acta, 75, 36273644.Google Scholar
Parkin, D. W. (1974). Trade-winds during the glacial cycles. Proceedings of the Royal Society of London, A 337, 73100.Google Scholar
Parkin, D. W. & Shackleton, N. (1973). Trade-winds and temperature correlations down a deep-sea core off the Saharan coast. Nature, 245, 455457.CrossRefGoogle Scholar
Parmenter, C. & Folger, D. W. (1974). Eolian biogenic detritus in deep sea sediments: A possible index of equatorial Ice Age aridity. Science, 185, 695698.CrossRefGoogle ScholarPubMed
Parry, D. E. & Wickens, G. E. (1981). The Qozes of Sorthern Darfur, Sudan Republic. Geographical Journal, 147, 307320.Google Scholar
Paterson, R. T. (1948). Darfur Province. In Tothill, J. D. (ed.), Agriculture in the Sudan, Oxford: Oxford University Press, pp. 851874.Google Scholar
Paton, T. R. (1978). The Formation of Soil Material. London: Allen & Unwin.Google Scholar
Paton, T. R. (1986). Perspectives on a Dynamic Earth. London: Allen & Unwin.CrossRefGoogle Scholar
Paton, T. R., Humphreys, G. S. & Mitchell, P. B. (1995). Soils: A New Global View. London: UCL Press.Google Scholar
Paton, T. R. & Williams, M. A. J. (1972). The concept of laterite. Annals of the Association of American Geographers, 62, 4256.Google Scholar
Peel, R. F. (1939). An expedition to the Gilf Kebir and ‘Uweinat. 4. The Gilf Kebir. Geographical Journal, 93, 295307.Google Scholar
Peel, R. F. (1941). Denudational landforms of the central Libyan desert. Journal of Geomorphology, 4, 323.Google Scholar
Peel, R. F. (1966). The landscape in aridity. Transactions of the Institute of British Geographers, 38, 123.Google Scholar
Peel, R. F. & Bagnold, R. A. (1939). An expedition to the Gilf Kebir and ‘Uweinat. 3. Archaeology. Geographical Journal, 93, 291295.Google Scholar
Pennington, B. T., Sturt, F., Wilson, P., Rowland, J. & Brown, A. G. (2017). The fluvial evolution of the Holocene Nile Delta. Quaternary Science Reviews, 170, 212231.Google Scholar
Penny, J. T. & Racek, A. A. (1968). Comprehensive Revision of a World-Wide Collection of Freshwater Sponges of Spongillidae. United States National Museum Bulletin 272, Washington, DC: Smithsonian Institution Press.Google Scholar
Pentecost, A. (2005). Travertine. Berlin: Springer.Google Scholar
Pesce, A. (1968). Gemini Space Photographs of Libya and Tibesti: A Geological and Geographical Analysis. Petroleum Exploration Society of Libya, Tripoli.Google Scholar
Petit-Maire, N. (ed.) (1982). Le Shati. Lac pléistocène du Fezzan (Libye). Paris: C.N.R.S.Google Scholar
Petit-Maire, N. (ed.) (1991). Paléoenvironnements du Sahara. Lacs holocènes à Taoudenni (Mali). Paris: C.N.R.S.Google Scholar
Petit-Maire, N. & Riser, J. (eds.) (1983). Sahara ou Sahel? Quaternaire récent du basin de Taoudenni (Mali). Paris: C.N.R.S.Google Scholar
Petraglia, M., Korisettar, R., Boivin, N., et al. (2007). Middle Pleistocene assemblages from the Indian subcontinent before and after the Toba super-eruption. Science, 317, 114116.Google Scholar
Philibert, A., Tibby, J. & Williams, M. (2010). A Middle Pleistocene diatomite from the western piedmont of Jebel Marra, Darfur, western Sudan, and its hydrological and climatic significance. Quaternary International, 216, 145150.Google Scholar
Phillips, J. S. (1987). Sinai during the Paleolithic: The early periods. In Close, A. E. (ed.), Prehistory of Arid North Africa: Essays in Honour of Fred Wendorf. Dallas: Southern Methodist University, pp. 105121.Google Scholar
Phillipps, R., Holdaway, S., Ramsay, R., Emmitt, J., Wendrich, W. & Linseele, V. (2016). Lake level changes, lake edge basins and the paleoenvironment of the Fayum north shore, Egypt, during the Early to Mid-Holocene. Open Quaternary, 2: 2, 112.Google Scholar
Phillipps, R., Holdaway, S. J. & Wendrich, W. (2017). The Fayum in the context of Northeast Africa. In Holdaway, S. J. & Wendrich, W. (eds.),The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press, pp. 916.Google Scholar
Phillipps, R., Holdaway, S., Wendrich, W. & Cappers, R. (2012). Mid-Holocene occupation of Egypt and global climatic change. Quaternary International, 251, 6476.Google Scholar
Pik, R., Marty, B., Carignan, J. & Lave, J. (2003). Stability of Upper Nile drainage network (Ethiopia) deduces from (U – Th)/He thermochronometry: Implication of uplift and erosion of the Afar plume dome. Earth and Planetary Science Letters, 215, 7388.Google Scholar
Pik, R., Marty, B., Carignan, J., Yirgu, G. & Ayalew, T. (2008). Timing of East African Rift development in southern Ethiopia: Implication for mantle plume activity and evolution of topography. Geology, 36, 167170.Google Scholar
Pillans, B., Spooner, N. & Chappell, J. (1997). The dynamics of soils in North Queensland: Rates of mixing by termites determined by single grain luminescence dating. In Roach, I. C. (ed.), Regolith and Landscapes in Eastern Australia. Canberra: CRC LEME, pp. 100101.Google Scholar
Pons, A. & Quézel, P. (1957). Première étude palynologique de quelques paléosols sahariens. Transactions de l’Institut de Recherches Sahariennes (Alger), 15, 1540.Google Scholar
Pons, A. & Quézel, P. (1958). Premières remarques sur l’étude palynologique d’un guano fossile du Hoggar. Comptes Rendus, Académie des Sciences, Paris, 246D, 22902292.Google Scholar
Potter, E. C. (1976). Pleistocene glaciation in Ethiopia: New evidence. Journal of Glaciology, 17, 148150.Google Scholar
Prasad, G. (1971). A Quaternary diatomite from the western Sudan. Palaeogeography, Palaeoclimatology, Palaeoecology, 10, 287292.CrossRefGoogle Scholar
Prasad, S., Negendank, J. F. W. & Stein, M. (2009). Varve counting reveals high resolution radiocarbon reservoir age variations in palaeolake Lisan. Journal of Quaternary Science, 24, 690696.Google Scholar
Prave, A. R., Bates, C. R., Donaldson, C. H., Toland, H., Condon, D. J. & Mark, D. (2016). Geology and geochronology of the Tana Basin, Ethiopia: LP volcanism, super eruptions and Eocene-Oligocene environmental change. Earth and Planetary Letters, 443, 18.Google Scholar
Prenni, A. J., Petters, M. D., Kreidenweis, S. M., et al. (2009). Relative roles of biogenic emissions and Saharan dust as ice nuclei in the Amazon basin. Nature Geosciences, 2, 402405.Google Scholar
Puchol, N., Blard, P.-H., Pik, R., Tibari, B. & Lavé, J. (2016). Variability of magmatic and cosmogenic 3He in Ethiopian river sands of detrital pyroxenes: Impact on denudation rate determinations. Chemical Geology, DOI:10.1016/j.chemgeo.2016.10.033.Google Scholar
Pye, K. (1987). Aeolian Dust and Dust Deposits. London: Academic Press.Google Scholar
Quade, J., Levin, N., Semaw, S., et al. (2004). Paleoenvironments of the earliest stone toolmakers, Gona, Ethiopia. Bulletin of the Geological Society of America, 116, 15291544.Google Scholar
Quézel, P. (1962). A propos de l’olivier de Lapérrine de l’Adrar Greboun. In Hugot, H. J. (ed.), Missions Berliet Ténéré-Tchad. Paris: Arts et Métiers Graphiques, pp. 329332.Google Scholar
Quézel, P. (1997). High mountains of the Central Sahara: Dispersal, origin and conservation of the flora. In Barakat, H. N. & Hegazy, A. K. (eds.), Reviews in Ecology, Desert Conservation and Development: A Festschrift for Prof. M. Kassas on the Occasion of His 75th Birthday. Cairo: UNESCO, IDRC & South Valley University, Egypt, pp. 159175.Google Scholar
Quézel, P. & Martinez, C. (1958–59). Le dernier interpluvial au Sahara central. Libyca, 6–6, 211225.Google Scholar
Quézel, P. & Martinez, C. (1962). Premiers résultats de l’analyse palynologique de sédiments recueillis au Sahara méridional à l’occasion de la mission Berliet-Tchad. In Hugot, H. J. (ed.), Missions Berliet Ténéré-Tchad. Paris: Arts et Métiers Graphiques, pp. 313327.Google Scholar
Railsback, L. B., Gibbard, P. L., Head, M. J., Voarintsoa, N. R. G. & Toucanne, S. (2015). An optimized scheme of lettered marine isotope substages for the last 1.0 million years, and the climatostratigraphic nature of isotope stages and substages. Quaternary Science Reviews, 111, 94106.Google Scholar
Rampino, M. R. & Ambrose, S. H. (2000). Volcanic winter in the Garden of Eden: The Toba supereruption and the late Pleistocene human population crash. In McCoy, F. W. & Heiken, G. (eds.), Volcanic Hazards and Disasters in Human Antiquity. Geological Society of America Special Paper, 345, 7182.Google Scholar
Randell, J. R. (1961). A rainfall map of the Sudan Gezira. Sudan Notes and Records, 42, 2936.Google Scholar
Rasmussen, M., Guo, X., Wang, Y., et al. (2011). An Aboriginal Australian genome reveals separate human dispersals into Asia. Science, 334, 9498.Google Scholar
Rattray, J. M. (1960). The Grass Cover of Africa. FAO Agricultural Studies, No. 49.Google Scholar
Retallack, G. J. (2001). Soils of the Past: An Introduction to Paleopedology, 2nd edn. New York, NY; Blackwell Science.Google Scholar
Retallack, G. J. & Huang, C. (2010). Depth to gypsic horizon as a proxy for paleoprecipitation in paleosols of sedimentary environments. Geology, 38, 403406.Google Scholar
Revel, M., Ducassou, E., Grousset, F. E., et al. (2010). 100,000 years of African monsoon variability recorded in sediments of the Nile margin. Quaternary Science Reviews, 29, 13421362.CrossRefGoogle Scholar
Revel, M., Ducassou, E., Skonieczny, C., et al. (2015). 20, 000 years of Nile River dynamics and environmental changes in the Nile catchment area as inferred from Nile upper continental slope sediments. Quaternary Science Reviews, 130, 200221.Google Scholar
Richter, D., Grün, R., Joannes-Boyau, R., et al. (2017). The age of the hominin fossils from Jebel Irhoud, Morocco, and the origins of the Middle Stone Age. Nature, 546, 293296.Google Scholar
Richter, J. (1989). Neolithic sites in the Wadi Howar (Western Sudan). In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Later Prehistroy of the Nile Basin and the Sahara. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 431442.Google Scholar
Richter, T. (2017). Natufian and Early Neolithic in the Black Desert, Eastern Jordan. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 715722.Google Scholar
Ritchie, J. C., Eyles, C. H. & Haynes, C. V. (1985). Sediment and pollen evidence for an early to mid-Holocene humid period in the eastern Sudan. Nature, 314, 352355.Google Scholar
Ritchie, J. C. & Haynes, C. V. (1987). Holocene vegetation zonation in the eastern Sahara. Nature, 330, 645647.Google Scholar
Robbins, L. H. (1984). Late prehistoric and pastoral adaptations west of Lake Turkana, Kenya. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press, pp. 206211.Google Scholar
Roberts, N., Taieb, M., Barker, P., Damnatti, B., Icole, M. & Williamson, D. (1993). Timing of the Younger Dryas event in East Africa from lake-level changes. Nature, 366, 146148.Google Scholar
Robertson, A. C. (1950). The Hafir. What-Why-Where-How. Ministry of Agriculture, Sudan Government, Khartoum, Bulletin No. 1.Google Scholar
Robertson Research & Geological Research Authority of the Sudan (1988a). Sudan Geology, 1:1,000,000 Dongola and Berber Sheet.Google Scholar
Robertson Research & Geological Research Authority of the Sudan (1988b). Sudan Geology, 1:1,000,000 Khartoum Sheet.Google Scholar
Roche, H. (1980). Premiers Outils Taillés d’Afrique. Paris, Société d’Ethnographie, 114 pp.Google Scholar
Roche, H., Delagnes, A., Brugal, J.-P., et al. (1999). Early hominid stone tool production skill 2.34 Myr ago in West Turkana, Kenya. Nature, 399, 5760.CrossRefGoogle ScholarPubMed
Rodd, F. R. (1926). People of the Veil. London: Macmillan.Google Scholar
Roden, J., Abdelsalam, M. G., Atekwana, E., El-Qady, G. & Tarabees, E. A. (2011). Structural influence of the pre-Eonile drainage system of southern Egypt: Insights from magnetotelluric and gravity data. Journal of African Earth Sciences, 61, 358368.Google Scholar
Rodrigues, D., Abell, P. I. & Kröpelin, S. (2000). Seasonality in the early Holocene climate of Northwest Sudan: Interpretation of Etheria elliptica shell isotopic data. Global and Planetary Change, 26, 181187.Google Scholar
Rögner, K., Knabe, K., Roscher, B., Smykatz-Kloss, W. & Zöller, L. (2004). Alluvial loess in the Central Sinai: Occurrence, origin, and palaeoclimatological consideration. Paleoecology of Quaternary Drylands, Lecture Notes in Earth Sciences, 102, 7999.Google Scholar
Rognon, P. (1967). Le Massif de l’Atakor et ses bordures (Sahara centrale). Étude géomorphologique. Paris: CNRS.Google Scholar
Rognon, P. (1976a). Essai d’interprétation des variations climatiques au Sahara depuis 40,000 ans. Revue de Géographie Physique et de Géologie Dynamique (2), 18(2–3), 337340.Google Scholar
Rognon, P. (1976b). Constructions alluviales holocènes et oscillations climatiques du Sahara méridional. Bulletin de l’Association Géographique de France, Paris, No. 433, 7783.Google Scholar
Rognon, P. (1980). Les phases d’aridité du Pléistocène supérieur et de l’Holocène au Sahara: Arguments sédimentologiques. Palaeoecology of Africa, 18, 111133.Google Scholar
Rognon, P. (1989). Biographie d’un Désert. Paris: Plon.Google Scholar
Rognon, P. & Gasse, F. (1973). Dépôts lacustres quaternaries de la basse vallée de l’Awash (Afar, Éthiopie); leurs rapports avec la tectonique et le volcanisme sous-aquatique. Revue de Géographie Physique et de Géologie Dynamique (2), 15(3), 295316.Google Scholar
Rognon, P. & Williams, M. A. J. (1977). Late Quaternary climatic changes in Australia and North Africa: A preliminary interpretation. Palaeogeography, Palaeoclimatology, Palaeoecology, 21, 285327.CrossRefGoogle Scholar
Rohling, E., Abu-Zied, R., Casford, J., Hayes, A., & Hoogakker, B. (2009). The marine environment: Present and past. In Woodward, J. C. (ed.), The Physical Geography of the Mediterranean. Oxford: Oxford University Press, pp. 3367.Google Scholar
Rohling, E. J., Cane, T. R., Cooke, S., et al. (2002). African monsoon variability during the previous interglacial maximum. Earth and Planetary Science Letters, 202, 6175.Google Scholar
Rohling, E. J., Grant, K. M., Roberts, A. P. & Larrasoaña, J.-C. (2013). Paleoclimate variability in the Mediterranean and Red Sea regions during the last 500,000 years. Current Anthropology, 54,S183S201.Google Scholar
Ron, H. & Levi, S. (2001). When did hominids first leave Africa? New high-resolution magnetostratigraphy from the Erk-el-Ahmar Formation, Israel. Geology, 29, 887890.Google Scholar
Rosenberg, T. M., Preusser, E., Fleitmann, D., et al. (2011). Humid periods in southern Arabia: Windows of opportunity for modern human dispersal. Geology, 39, 11151118.CrossRefGoogle Scholar
Roskin, J., Porat, N., Tsoar, H., Blumberg, D. G. & Zander, A. M. (2011a). Age, origin and climatic controls on vegetated linear dunes in the northwestern Negev Desert (Israel). Quaternary Science Reviews, 30, 16491674.Google Scholar
Roskin, J. & Tsoar, H. (2017). Late Quaternary chronologies of the northern Sinai/northwestern Negev dunefield and their palaeoclimatic and palaeoenvironmental implications. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge, Cambridge University Press, pp. 505520.Google Scholar
Roskin, J., Tsoar, H., Porat, N. & Blumberg, D. G. (2011b). Palaeoclimate interpretations of Late Pleistocene vegetated linear dune mobilization episodes: Evidence from the northwestern Negev dunefield, Israel. Quaternary Science Reviews, 30, 33643380.CrossRefGoogle Scholar
Rossignol, M. (1961). Analyses polliniques de sédiments marins quaternaires en Israel. I. Sédiments récents. Pollens et Spores, 3, 303324.Google Scholar
Rossignol, M. (1962). Analyses polliniques de sédiments marins quaternaires en Israel. II. Sédiments pléistocènes. Pollens et Spores, 4, 121148.Google Scholar
Rossignol-Strick, M. (1985). Mediterranean Quaternary sapropels, an immediate response of the African monsoon to variations in insolation. Palaeogeography, Palaeoclimatology, Palaeoecology, 49, 237263.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
Rossignol-Strick, M., Nesterhoff, W., Olive, P. & Vergnaud-Grazzini, C. (1982). After the deluge: Mediterranean stagnation and sapropel formation. Nature, 295, 105110.Google Scholar
Russell, E. W. (ed.) (1962). The Natural Resources of East Africa. Nairobi, Kenya: D. A. Hawkins, Ltd. and East African Literature Bureau.Google Scholar
Ruxton, B. P. (1958). Weathering and subsurface erosion at the piedmont angle, Balos, Sudan. Geological Magazine, 95, 353377.Google Scholar
Ruxton, B. P. & Berry, L. (1978). Clay plains and geomorphic history of the central Sudan: A review. Catena, 5, 251283.Google Scholar
Rzóska, J. (1961). Observations on tropical rainpools and general remarks on temporary waters. Hydrobiologia, 17, 265286.Google Scholar
Rzóska, J. (1976) (ed.). The Nile: Biology of an Ancient River. The Hague: Junk.CrossRefGoogle Scholar
Said, R. (1981). The Geological Evolution of the River Nile. New York, NY: Springer-Verlag.CrossRefGoogle Scholar
Said, R. (1993). The River Nile: Geology, Hydrology and Utilization. Oxford: Pergamon.Google Scholar
Salama, R. B. (1987). The evolution of the River Nile: The buried saline rift lakes in Sudan. Journal of African Earth Sciences, 6, 899913.Google Scholar
Salama, R. B. (1994). The Sudanese buried saline lakes. In Rosen, M. R. (ed.), Paleoclimate and Basin Evolution of Playa Systems. Geological Society of America, Boulder, Colorado, Special Paper 289, pp. 3348.Google Scholar
Salvatori, S. (2012). Disclosing archaeological complexity of the Khartoum Mesolithic. New data at the site and regional level. African Archaeological Review, 29, 399472.CrossRefGoogle Scholar
Salvatori, S., Usai, D., Faroug, M. A., et al. (2014). Archaeology at Al Khiday: New insight on the prehistory and history of Central Sudan. In Anderson, J.R. & Welsby, D. A. (eds.), The Fourth Cataract and Beyond. Proceedings of the 12th International Conference of Nubian Studies. British Museum Publications on Egypt and Sudan 1. Leuven: Peeters, pp. 243247.Google Scholar
Salvatori, S., Usai, D. & Zerboni, A. (2011). Mesolithic site formation and palaeoenvironment along the White Nile (Central Sudan). African Archaeological Review, 28, 177211.Google Scholar
Sandford, K. S. (1933a). Geology and geomorphology in the southern Libyan Desert. Geographical Journal, 82, 213219.Google Scholar
Sandford, K. S. (1933b). Volcanic craters in the Libyan Desert. Nature, 131, 4649.CrossRefGoogle Scholar
Sandford, K. S. (1933c). Past climate and early Man in the Southern Libyan Desert. Geographical Journal, 82, 219222.Google Scholar
Sandford, K. S. (1934). Paleolithic Man and the Nile Valley in Middle and Upper Egypt. The University of Chicago Oriental Institute Publications, 18, 1131.Google Scholar
Sandford, K. S. (1935). Geological observations on the north-western frontiers of the Anglo-Egyptian Sudan and the adjoining part of the southern Libyan Desert. Quarterly Journal of the Geological Society of London, 91, 323381.Google Scholar
Sandford, K. S. (1936). Observations on the distribution of land and freshwater mollusca in the S. Libyan desert. Quarterly Journal of the Geological Society of London, 92, 207220.CrossRefGoogle Scholar
Sandford, K. S. (1953). Notes on sand dunes in Egypt and Sudan. Geographical Journal, 119, 363366.Google Scholar
Sandford, K. S. & Arkell, W. J. (1929a). Palaeolithic Man and the Nile-Faiyum Divide: A Study of the Region during Pliocene and Pleistocene Times. The University of Chicago Oriental Institute Publications, 10, Prehistoric Survey of Egypt and Western Asia, 1, pp. 177.Google Scholar
Sandford, K. S. & Arkell, W. J. (1929b). Palaeolithic Man and the Nile Valley in Lower Egypt. The University of Chicago Oriental Institute Publications, 46, 1105.Google Scholar
Sandford, K. S. & Arkell, W. J. (1933). Palaeolithic Man and the Nile Valley in Nubia and Upper Egypt. The University of Chicago Oriental Institute Publications, 17, 192.Google Scholar
Sarnthein, M. (1978). Sand deserts during glacial maximum and climatic optimum. Nature, 272, 4345.CrossRefGoogle Scholar
Sarnthein, M., Tetzlaff, G., Koopmann, B., Wolter, K. & Pflaumann, U. (1981). Glacial and interglacial wind regimes over the eastern subtropical Atlantic and north-west Africa. Nature, 293, 193196.Google Scholar
Sarnthein, M., Thiede, J., Pflaumann, U., et al. (1982). Atmospheric and oceanic circulation patterns off Northwest Africa during the past 25 million years. In von Rad, U., Hinz, K., Sarnthein, M. & Seibold, E. (eds.), Geology of the Northwest African Continental Margin. Berlin: Springer-Verlag, pp. 545604.Google Scholar
Schaefer, J. M., Denton, G. H., Barrell, D. J. A., et al. (2006). Near-synchronous interhemispheric termination of the Last Glacial Maximum in mid-latitudes. Science, 312, 15101513.Google Scholar
Schaller, K. F. & Kuls, W. (1972). Äthiopien-Ethiopia. Berlin:Springer-Verlag (In German and English).Google Scholar
Schild, R. & Wendorf, F. (1981). The Prehistory of an Egyptian Oasis. Wroclaw, Poland: Ossolineum.Google Scholar
Schild, R. & Wendorf, F. (1993). Middle Palaeolithic lakes in the Southwestern Desert of Egypt. In Wendorf, F., Schild, R., Close, A. E., et al., Egypt During the Last Interglacial: The Middle Palaeolithic of Bir Tarfawi and Bir Sahara East. New York: Plenum Press, pp. 1565.Google Scholar
Schild, R. & Wendorf, F. (2001). Geoarchaeology of the Holocene Climatic Optimum at Nabta Playa, Southwestern Desert, Egypt. Geoarchaeology, 16, 728.Google Scholar
Schlüter, T. (1997). Geology of East Africa. Berlin: Gebrüder Borntraeger.Google Scholar
Schuck, W. (1989). From lake to well: 5, 000 years of settlement in Wadi Shaw. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Later Prehistoroy of the Nile Basin and the Sahara. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 421429.Google Scholar
Schulz, E. (1994). The southern limit of the Mediterranean vegetation in the Sahara during the Holocene. Historical Biology, 9, 137156.Google Scholar
Schumm, S. A. (1977). The Fluvial System. New York, NY: Wiley-Interscience.Google Scholar
Schumm, S. A. & Parker, R. S. (1973). Implications of complex responses of drainage systems for Quaternary alluvial stratigraphy. Nature Physical Science, 243, 99100.Google Scholar
Schuster, M., Duringer, P., Ghienne, J.-F., et al. (2006). The age of the Sahara desert. Science, 311, 821.Google Scholar
Schütz, L., Jaenicke, R. & Pietrek, H. (1981). Saharan dust transport over the North Atlantic Ocean. In Péwé, T. L. (ed.), Desert Dust: Origin, Characteristics, and Effect on Man. Geological Society of America Special Paper, 186, 87100.Google Scholar
Schwarcz, H. P., Grün, R., Vandermeersch, B., Bar-Yosef, O., Vallas, H. & Tchernov, E. (1988). ESR dates for the hominid burial site of Qafzeh in Israel. Journal of Human Evolution, 17, 733737.CrossRefGoogle Scholar
Scott, H. (1958). Biogeographical research in High Simien (Northern Ethiopia) 1952–53. Proceedings of the Linnean Society of London, 170, 191.Google Scholar
Scrivner, A. E., Vance, D. & Rohling, E. J. (2004). New neodymium isotope data quantify Nile involvement in Mediterranean anoxic episodes. Geology, 32, 565568.CrossRefGoogle Scholar
Semaw, S., Renne, P., Harris, J. W. K., Feibel, C.S., Bernor, R.L., Fesseka, N. & Mowbray, K. (1997). 2.5-million-year-old stone tools from Gona, Ethiopia. Nature, 385, 333336.Google Scholar
Sembroni, A., Molin, P., Pazzaglia, F. J., Faccenna, C. & Abebe, B. (2016). Evolution of continental-scale drainage in response to mantle dynamics and surface processes: An example from the Ethiopian Highlands. Geomorphology, 261, 1229.Google Scholar
Sepulchre, P., Ramstein, G., Fluteau, F., Schuster, M., Tiercelin, J.-J. & Brunet, M. (2006). Tectonic uplift and Eastern African aridification. Science, 313,1419–23.Google Scholar
Servant, M. (1973). Séquences continentales et variations climatiques: évolution du Bassin du Tchad au Cénozoique Supérieur. Paris: ORSTOM.Google Scholar
Servant, M., Ergenzinger, P. & Coppens, Y. (1969). Datations absolues sur un delta lacustre quaternaire au Sud du Tibesti (Angamma). Comptes Rendus Sommaire de la Société Géologique de France, 1, 313314.Google Scholar
Servant, M. & Servant-Vildary, S. (1980). L’environnement quaternaire du basin du Tchad. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 133162.Google Scholar
Shahin, M. (1985). Hydrology of the Nile Basin. Amsterdam: Elsevier.Google Scholar
Sharon, G. (2017). The Acheulian of the Levant. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 539547.Google Scholar
Shaw, W. B. K. (1933). Neolithic and later times. Geographical Journal, 82, 222224.Google Scholar
Shaw, W. B. K. (1936). An expedition to the southern Libyan Desert. Geographical Journal, 87, 193221.Google Scholar
Shiner, J. L. (1971). The prehistory and geology of Northern Sudan. Report to the NSF. Parts I and II.Google Scholar
Shinnie, P. L. (1967). Meroe: A Civilization of the Sudan. London: Thames and Hudson.Google Scholar
Shriäinen, A. (1984). Two Southern Sudanese pottery traditions in a historical perspective. Norwegian Archaeological Review, 17, 1118.Google Scholar
Shukri, N. M. (1949). The mineralogy of some Nile sediments. Quarterly Journal of the Geological Society, 105, 511534.Google Scholar
Singhvi, A. K., Williams, M. A. J., Rajaguru, S. N., et al. (2010). A ~200 ka record of climatic change and dune activity in the Thar Desert, India. Quaternary Science Reviews, 29, 30953105.Google Scholar
Sir Gibb, Alexander & Partners (1954). Estimation of irrigable areas in the Sudan 1951–3. Report to the Sudan Government. London: Metcalf and Cooper Ltd.Google Scholar
Smith, A. B. (1980). The Neolithic tradition in the Sahara. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 451465.Google Scholar
Smith, A. B. (1884). The origins of food production in northeast Africa. Palaeoecology of Africa, 16, 317324.Google Scholar
Smith, A. B. (1992). Pastoralism in Africa: Origins and Development Ecology. London: Hurst & Company.Google Scholar
Smith, A. G., Hurley, A. M. & Briden, J. C. (1981). Phanerozoic Paleocontinetal World Maps. Cambridge: Cambridge University Press.Google Scholar
Smith, B. D. (1995). The Emergence of Agriculture. New York, NY: Scientific American Library.Google Scholar
Smith, J. (1949). Distribution of tree species in the Sudan in relation to rainfall and soil texture. Sudan Government Ministry of Agriculture, Bulletin 4.Google Scholar
Smith, J. R., Giegengack, R., Schwarcz, H. P., et al. (2004). A reconstruction of Quaternary pluvial environments and human occupations using stratigraphy and geochronology of fossil-spring tufas, Kharga Oasis, Egypt. Geoarchaeology, 19, 407439.Google Scholar
Smith, L. (1998–2002). The SARS Survey along the Omdurman-Gabolab Road 1997: Interim Report on the pottery and small finds. Kush, 28, 131156.Google Scholar
Snow, O. W. (1948). Animal foodstuffs. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 668687.Google Scholar
Soil Survey Staff. (1960). Soil Classification: A Comprehensive System (7th Approximation). Washington, DC: USDA Soil Conservation Service.Google Scholar
Soil Survey Staff. (2010). Keys to Soil Taxonomy, 11th edn. Washington, DC: USDA-Natural Resources Conservation Service.Google Scholar
Sparks, B. W. (1972). Geomorphology. London: Longman.Google Scholar
Spate, O. H. K. (1952). Toynbee and Huntington: A study in determinism. Geographical Journal, 118, 406424.CrossRefGoogle Scholar
Stager, J. C., Cumming, B. & Meeker, L. (1997). A high-resolution 11,400-yr diatom record from Lake Victoria, East Africa. Quaternary Research, 47, 8189.Google Scholar
Stager, J. C., Cumming, B. F. & Meeker, L. D. (2003). A 10, 000-year high-resolution diatom record from Pilkington Bay, Lake Victoria, East Africa. Quaternary Research, 59, 172181.Google Scholar
Stager, J. C. & Johnson, T. C. (2000). A 12, 400 14C yr offshore diatom record from east central Lake Victoria, East Africa. Journal of Paleolimnology, 23, 373383.Google Scholar
Stager, J. C. & Johnson, T. C. (2008). The Late Pleistocene desiccation of Lake Victoria and the origin of its endemic biota. Hydrobiologia, 596, 516.Google Scholar
Stager, J. C., Mayewski, P. A. & Meeker, L. D. (2002). Cooling cycles, Heinrich event 1, and the desiccation of Lake Victoria. Palaeogeography, Palaeoclimatology Palaeoecology, 183, 169178.CrossRefGoogle Scholar
Stager, J. C., Reinthal, P. N. & Livingstone, D. A. (1986). A 25,000-year history for Lake Victoria, East Africa, and some comments on its significance for the evolution of cichlid fishes. Freshwater Biology, 16, 1519.Google Scholar
Stahl, A. B. (1984). A history and critique of investigations into early African agriculture. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press, pp. 921.Google Scholar
Stanley, D. J., Krom, M. D., Cliff, R. A. & Woodward, J. A. (2003). Nile flow failure at the end of the Old Kingdom, Egypt: Strontium isotopic and petrologic evidence. Geoarchaeology, 18, 395402.Google Scholar
Stanley, D. J. & Warne, A. G. (1993a). Nile Delta: Recent geological evolution and human impact. Science, 260, 628634.CrossRefGoogle ScholarPubMed
Stanley, D. J. & Warne, A. G. (1993b). Sea level and initiation of Predynastic culture in the Nile Delta. Nature, 363, 435438.Google Scholar
Stanley, D. J. & Warne, A. G. (1998). Nile Delta in its destructive phase. Journal of Coastal Research, 14, 794825.Google Scholar
Stemler, A. B. L. (1980). Origins of plant domestication in the Sahara and Nile valley. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Quaternary Environments and Prehistoric Occupation in Northern Africa. Rotterdam: A. A. Balkema, pp. 503526.Google Scholar
Stokes, S., Maxwell, T. A., Haynes, C. V. & Horrocks, J. (1998). Latest Pleistocene and Holocene sand-sheet construction in the Selima Sand Sea, Eastern Sahara. In Alsharhan, A. S., Glennie, K. W., Whittle, G. L. & Kendall, C. G. St. C. (eds.),Quaternary Deserts and Climatic. Rotterdam: A. A.Balkema, pp. 175195.Google Scholar
Stone, A. E. C. & Thomas, D. S. G. (2008). Linear dune accumulation chronologies from the southwest Kalahari, Namibia: Challenges of reconstructing late Quaternary palaeoenvironments from aeolian landforms. Quaternary Science Reviews, 27, 16671681.Google Scholar
Street, F. A. (1979a). Late Quaternary lakes in the Ziway-Shala basin, southern Ethiopia. Unpublished PhD thesis, University of Cambridge.Google Scholar
Street, F. A. (1979b). Late Quaternary precipitation estimates for the Ziway-Shala basin, southern Ethiopia. Palaeoecology of Africa, 11, 135143.Google Scholar
Street, F. A. (1980). The relative importance of climate and local hydrogeological factors in influencing lake-level fluctuations. Palaeoecology of Africa, 12, 137158.Google Scholar
Stringer, C. (2000). Coasting out of Africa. Nature, 405, 2427.Google Scholar
Stringer, C. B., Grün, R., Schwarcz, H. P. & Goldberg, P. (1989). ESR dates for the hominin burial site of Es Skhul in Israel. Nature, 338, 756758.CrossRefGoogle ScholarPubMed
Stringer, C. & McKie, R. (1996). African Exodus: The Origins of Modern Humanity. London: Jonathan Cape.Google Scholar
Sudan Meteorological Service. (no date). Climatological normals (provisional) 1931–1960, Khartoum.Google Scholar
Sukova, L. (2011). The ‘Venus’ of Jebel Uweinat (SE Libya). Sahara, 22, 117124.Google Scholar
Sultan, M., Sturchio, N., Hassan, F. A., et al. (1997). Precipitation source inferred from stable isotopic composition of Pleistocene groundwater and carbonate deposits in the Western Desert of Egypt. Quaternary Research, 48, 2937.Google Scholar
Sutcliffe, J. V. (2009). The hydrology of the Nile Basin. In Dumont, H. J. (ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 335364.Google Scholar
Sutcliffe, J. V. & Parks, Y. P. (1999). The Hydrology of the Nile. International Association of Hydrological Sciences Special Publication, 5, 1179.Google Scholar
Sutton, J. E. G. (1974). The aquatic civilization of middle Africa. Journal of African History, 15, 527546.Google Scholar
Sutton, J. E. G. (1977). The African Aqualithic. Antiquity, 51, 2534.Google Scholar
Swain, A. (1997). Ethiopia, the Sudan and Egypt: The Nile River dispute. Journal of Modern African Studies, 35, 674694.Google Scholar
Swap, R., Garstang, M., Greco, S., Talbot, R. & Kallberg, P. (1992). Saharan dust in the Amazon Basin. Tellus, 44B, 133149.CrossRefGoogle Scholar
Swezey, C. (2001). Eolian sediment response to late Quaternary climate changes: Temporal and spatial patterns in the Sahara. Palaeogeography, Palaeoclimatology, Palaeoecology, 167, 119155.CrossRefGoogle Scholar
Swezey, C. (2003). The role of climate in the creation and destruction of continental stratigraphic records: An example from the northern margin of the Sahara Desert. SEPM Special Publication, No. 77, 207225.Google Scholar
Szabo, B. J., Haynes, C. V. & Maxwell, T. A. (1995). Ages of Quaternary pluvial episodes determined by uranium-series and radiocarbon dating of lacustrine deposits of Eastern Sahara. Palaeogeography, Palaeoeclimatology, Palaeoecology, 113, 227242.CrossRefGoogle Scholar
Taïeb, M. (1974). Évolution quaternaire du Bassin de l’Awash (Rift éthiopien et Afar). D. ès Sc. thesis, Paris, 2 vols.Google Scholar
Talbot, M. R. (1980). Environmental responses to climatic change in the West African Sahel over the past 20 000 years. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, Rotterdam, pp. 3762.Google Scholar
Talbot, M. R. (1985). Major bounding surfaces in aeolian sandstones: A climatic model. Sedimentology, 32, 257265.Google Scholar
Talbot, M. R. & Williams, M. A. J. (1978). Erosion of fixed dunes in the Sahel, Central Niger. Earth Surface Processes, 3, 107113.Google Scholar
Talbot, M. R. & Williams, M. A. J. (1979) Cyclic alluvial fan sedimentation on the flanks of fixed dunes, Janjari, central Niger. Catena, 6, 4362.Google Scholar
Talbot, M. R. & Williams, M. A. J. (2009). Cenozoic evolution of the Nile basin. In Dumont, H. J. (ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 3760.Google Scholar
Talbot, M. R., Williams, M. A. J. & Adamson, D. A. (2000). Strontium isotope evidence for late Pleistocene reestablishment of an integrated Nile drainage network. Geology, 28, 343346.Google Scholar
Talling, J. F. (1957). The longitudinal succession of water characteristics in the White Nile. Hydrobiologia, 11, 7389.Google Scholar
Talling, J. F. (1976). Water characteristics. In Rzóska, J. (ed.), The Nile: Biology of an Ancient River. The Hague: Junk, pp. 357384.Google Scholar
Tchakerian, V. P. (ed.) (1995). Desert Aeolian Processes. London: Chapman & Hall.Google Scholar
Tchakerian, V. P. (2009). Palaeoclimatic interpretations from desert dunes and sediments. In Parsons, A. J. & Abrahams, A. D. (eds.), Geomorphology of Desert Environments, 2nd edn. Berlin: Springer, pp. 757772.CrossRefGoogle Scholar
Teller, J. T., McGinn, R. A., Rajapara, H. M., Shukla, A. D. & Singhvi, A. K. (2018). Optically stimulated luminescence ages from the Lake Agassiz basin in Manitoba. Quaternary Research, DOI: 10.1017/qua.2017.107Google Scholar
Templeton, A. R. (2002). Out of Africa again and again. Nature, 416, 4551.Google Scholar
Thomas, D. S. G. & Middleton, N. J. (1994). Desertification: Exploding the Myth. Chichester: John Wiley & Sons.Google Scholar
Thompson, K. (1976). Swamp development in the headwaters of the White Nile. In Rzóska, J. (ed.), The Nile: Biology of an Ancient River. The Hague: Junk, pp. 177196.Google Scholar
Thompson, L. G., Mosley-Thompson, E., Davis, M. E., et al. (2002). Kilimanjaro ice core records: Evidence of Holocene climate change in tropical Africa. Science, 298, 589593.Google Scholar
Thomson, J., Mercone, D., De Lange, G. J. & Van Santvoort, P. J. M. (1999). Review of recent advances in the interpretation of eastern Mediterranean sapropel S1 from geochemical evidence. Marine Geology, 153, 7789.Google Scholar
Thorne, A. G. & Wolpoff, M. H. (1981). Regional continuity in Australasian Pleistocene hominid evolution. American Journal of Physical Anthropology, 55, 337349.Google Scholar
Thorp, M. B. (1969). Some aspects of the geomorphology of the Aïr Mountains, southern Sahara. Transactions of the Institute of British Geographers, 47, 2546.Google Scholar
Thurmond, A. K., Stern, R. J., Abdelsalam, M. G., Nielsen, K. C., Abdeen, M. M. & Hinz, E. (2004). The Nubian Swell. Journal of African Earth Sciences, 39, 401407.CrossRefGoogle Scholar
Tiercelin, J. J. (1981). Rifts continentaux, tectonique, climats, sédiments. Exemples: la sédimentation dans le Nord du Rift Gregory (Kenya) et dans le Rift de l’Afar (Éthiopie) depuis le Miocène. Unpublished DSc. thesis, Universite d’Aix-Marseille II. Marseille.Google Scholar
Tiercelin, J. J. (1986). The Pliocene Hadar Formation, Afar depression of Ethiopia. In Frostick, L. E. & Reid, I. (eds.), Desert Sediments: Ancient and Modern. Geological Society Special Publication, No. 35. Oxford: Blackwell Scientific, pp. 221240.Google Scholar
Tierney, J. E. & deMenocal, P. B. (2013). Abrupt shifts in Horn of Africa hydroclimate since the Last Glacial Maximum. Science, 342, 843846.Google Scholar
Tillet, T. (ed.) (1997). Sahara: Paléomilieux et Peuplement Préhistorique au Pléistocène Supérieur. Paris: l’Harmattan.Google Scholar
Tobler, R., Rohrlach, A., Soubrier, J., et al. (2017). Aboriginal mitogenomes reveal 50,000 years of regionalism in Australia. Nature, 544, 180184.Google Scholar
Toonen, W., Graham, A., Pennington, B., Hunter, M., Strutt, K., Barker, D., Masson-Berghoff, A. & Emery, V. (2017). Holocene fluvial history of the Nile’s west bank floodplain at ancient Thebes (Luxor, Egypt) and its relation with cultural dynamics and basin-wide hydroclimatic variability. Geoarchaeology, DOI: 10.1002/gea.21631.CrossRefGoogle Scholar
Tothill, J. D. (1946a). The origin of the Sudan Gezira clay plain. Sudan Notes and Records, 27, 153183.Google Scholar
Tothill, J. D. (1946b). The fossil-bearing clay at Erkowit. Sudan Notes and Records, 27, 229232.Google Scholar
Tothill, J. D. (1948). A note on the origins of the soils of the Sudan from the point of view of the man in the field. In Tothill, J. D. (ed.), Agriculture in the Sudan. Oxford: Oxford University Press, pp. 129143.Google Scholar
Trauth, M. H., Maslin, M. A., Deino, A. L., Strecker, M. R., Bergner, A. G. N. & Dühnforth, M. (2007). High- and low-latitude forcing of Plio-Pleistocene East African climate and human evolution. Journal of Human Evolution, 53, 475486.CrossRefGoogle ScholarPubMed
Trauth, M.H., Maslin, M.A., Deino, A.L., Junninger, A., Lesoloyia, M., Odada, E.O., Olago, D.O., Olaka, L.A., Strecker, M.R., & Tiedemann, R. (2010). Human evolution in a variable environment: The amplifier lakes of Eastern Africa. Quaternary Science Reviews, 29, 29812988.Google Scholar
Treble, P. C., Baker, A., Ayliffe, L. K., et al. (2017). Hydroclimate of the Last Glacial Maximum and deglaciation in southern Australia’s arid margin interpreted from speleothem records (23–15 ka). Climate of the Past, 13, 667687.Google Scholar
Trigger, B. G. (1982). The rise of civilization in Egypt. In Clark, J. D. (ed.), The Cambridge History of Africa, Vol. I: From the Earliest Times to c. 500 BC. Cambridge: Cambridge University Press, pp. 478547.Google Scholar
Tryon, C. A., Faith, J. T., Peppe, D. J., et al. (2015). The Pleistocene prehistory of the Lake Victoria basin. Quaternary International, 404, 100114.Google Scholar
Tsoar, H. & Pye, K. (1987). Dust transport and the question of desert loess formation. Sedimentology, 34, 139153.Google Scholar
Turney, C. S. M., Kershaw, A. P., Clemens, S. C., Branch, N., Moss, P. T. & Fifield, L. K. (2004). Millennial and orbital variations of El Niño/Southern Oscillation and high latitude climate in the last glacial period. Nature, 428, 306310.Google Scholar
Umer, M., Lamb, H. F., Bonnefille, R., et al. (2007). Late Pleistocene and Holocene vegetation history of the Bale Mountains, Ethiopia. Quaternary Science Reviews, 26, 22292246.Google Scholar
Underhill, P. A., Passarino, G., Lin, A. A., et al. (2001). The phylogeography of Y chromozone binary haplotypes and the origins of modern human populations. Annals of Human Genetics, 65, 4362.CrossRefGoogle Scholar
UNEP. (1992a). Status of desertification and implementation of the United Nations Plan of Action to Combat Desertification. Report of the Executive Director. Nairobi: United Nations Environment Programme,Google Scholar
UNEP. (1992b). Middleton, N. & Thomas, D. S. G. (eds.), World Atlas of Desertification, 1st edn. London: Arnold.Google Scholar
UNEP. (1997). Middleton, N. & Thomas, D. S. G. (eds.), World Atlas of Desertification, 2nd edn. London: Arnold.Google Scholar
UNESCO. (2007). International Sediment Initiative Conference, 12–15 November, 2006, Khartoum. UNESCO Chair in Water Resources, Khartoum.Google Scholar
United Nations, Ethiopia. (1971). Report on the geology, geochemistry and hydrology of hot springs of the eastern rift system. Rome: United Nations Development Programme.Google Scholar
United States Department of Agriculture. (1951). Soil Survey Manual. Agriculture Handbook No. 18.Google Scholar
UNOCHA (UN Office for the Coordination of Humanitarian Affairs). 1999a. OCHA Situation Report #2. Retrieved from: www.reliefweb.int (accessed 19 December 2002).Google Scholar
UNOCHA (UN Office for the Coordination of Humanitarian Affairs). 1999b. OCHA Situation Report #3. Retrieved from: www.reliefweb.int (accessed 19 December 2002).Google Scholar
Usai, D. & Salvatori, S. (2007). The oldest representation of a Nile boat. Antiquity, 81, (314) (Antiquity Gallery). Retrieved from: www.antiquity.ac.uk/ProjGall/usaiGoogle Scholar
Vail, J. R. (1972a). Jebel Marra, a dormant volcano in Darfur Province, western Sudan. Bulletin Volcanologique, 36, 251265.Google Scholar
Vail, J. R. (1972b). Geological reconnaissance in the Zalingei and Jebel Marra areas of western Darfur Province, Sudan. Bulletin of the Geological Survey, Sudan, 19, 150.Google Scholar
Vail, J. R. (1974). Geological map of the Democratic Republic of Sudan and adjacent areas, at scale 1: 2 000 000. Ministry of Overseas Development (British Government) Publication DOS 1203B.Google Scholar
Vail, J. R. (1976). Outline of the geochronology and tectonic units of the Basement complex of north-east Africa. Proceedings of the Royal Society of London, 350A, 127141.Google Scholar
Vail, J. R. (1978). Outline of the Geology and Mineral Deposits of the Democratic Republic of the Sudan and Adjacent Areas. Overseas Geology and Mineral Resources No. 49. London: Her Majesty’s Stationary Office.Google Scholar
Vail, J. R. (1983). Pan-African crustal accretion in Northeast Africa. Journal of African Earth Sciences, 1, 285294.Google Scholar
Vail, J. R. (1985). Pan-African (Late Precambrian) tectonic terrains and the reconstruction of the Arabian-Nubian Shield. Geology, 13, 839842.Google Scholar
Vaks, A., Bar-Matthews, M., Ayalon, A., et al. (2003). Paleoclimate 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
Vaks, A., Bar-Matthews, M., Ayalon, A., et al. (2006). Paleoclimate and location of the border between Mediterranean climate region and the Saharo-Arabian Desert as revealed by speleothems from the northern Negev Desert, Israel. Earth and Planetary Science Letters, 249, 384399.Google Scholar
Vaks, A., Bar-Matthews, M., Ayalon, A., Matthews, A., Halicz, L. & Frumkin, A. (2007). Desert speleothems reveal climatic window for African exodus of early modern humans. Geology, 35, 831834.Google Scholar
Vaks, A., Bar-Matthews, M., Matthews, A., Ayalon, A. & Frumkin, A. (2010). Middle-Late Quaternary paleoclimate of northern margins of the Saharan-Arabian Desert: reconstruction from speleothems of Negev Desert, Israel. Quaternary Science Reviews, 29, 26472662.Google Scholar
Van Damme, D. (1984). The Freshwater Mollusca of Northern Africa: Distribution, Biogeography and Palaeoecology. Dordrecht: Kluwer Academic.Google Scholar
Van Damme, D. & Pickford, M. (2003). The late Cenozoic Thiaridae (Mollusca, Gastropoda, Cerithioidea) of the Albertine Rift Valley (Uganda-Congo) and their bearing on the origin and evolution of the Tanganyikan thalassoid malacofauna. Hydrobiologia, 498, 183.Google Scholar
Van Damme, D. & Van Bocxlaer, B. (2009). Freshwater molluscs of the Nile Basin, past and present. In Dumont, H. J. (ed.), The Nile. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 585629.Google Scholar
Van Neer, W. (1993). Fish remains from the Last Interglacial at Bir Tarfawi (Eastern Sahara, Egypt). In Wendorf, F., Schild, R., Close, A. E., et al. (eds.), Egypt During the Last Interglacial: The Middle Palaeolithic of Bir Tarfawi and Bir Sahara East, Plenum Press: New York, pp. 144155.Google Scholar
Van Neer, W. (2004). Evolution of prehistoric fishing in the Egyptian Nile Valley. Journal of African Archaeology, 2, 251269.Google Scholar
Van Peer, P. (1998). The Nile corridor and the Out-of-Africa model. Current Anthropology, 39, 115140.CrossRefGoogle Scholar
Van Zinderen Bakker, E. M. (1978). Late Mesozoic and Tertiary palaeoenvironments of the Sahara region. In van Zinderen Bakker, E. M. (ed.), Antarctic Glacial History and World Palaeoenvironments. Rotterdam: A. A.Balkema, pp. 129135.Google Scholar
Vekua, A., Lordkipanidze, D., Rightmire, G. P., et al. (2002). A new skull of early Homo from Dmanisi, Georgia. Science, 297, 8589.Google Scholar
Vermeersch, P. M. & Van Neer, W. (2015). Nile behaviour and Late Palaeolithic humans in Upper Egypt during the Late Pleistocene. Quaternary Science Reviews, 130, 155167.Google Scholar
Vernet, R. (1995). Climats Anciens du Nord de l’Afrique. Paris: l’Harmattan.Google Scholar
Véron, A. J., Flaux, C., Marriner, N., et al. (2013). A 6000-year geochemical record of human activities from Alexandria (Egypt). Quaternary Science Reviews, 81, 138147.Google Scholar
Verschuren, D., Jaap, S. Sinninghe Damsté, J. S., Moernaut, J., Kristen, I., Blaauw, M., Fagot, M., Haug, G.H. & CHALLACEA project members. (2009). Half-precessional dynamics of monsoon rainfall near the East African equator. Nature, 462, 637641.Google Scholar
Verschuren, D., Laird, K. R. & Cumming, B. F. (2000). Rainfall and drought in equatorial east Africa during the past 1,100 years. Nature, 403, 410414.CrossRefGoogle ScholarPubMed
Vigne, J.-D., Gourichon, L., Helmer, D., Martin, L. & Peters, J. (2017). The beginning of animal domestication and husbandry in Southwest Asia. In Enzel, Y. & Bar-Yosef, O. (eds.), Quaternary of the Levant. Cambridge: Cambridge University Press, pp. 753759.Google Scholar
Vinnicombe, P. (1976). People of the Eland. Pietermaritzburg: Natal University Press.Google Scholar
Vita-Finzi, C. (1969a). The Mediterranean Valleys: Geological Changes in Historical Times. Cambridge: Cambridge University Press.Google Scholar
Vita-Finzi, C. (1969b). Geological opportunism. In Ucko, P. J. & Dimbleby, G. W. (eds.), The Domestication and Exploitation of Plants and Animals. London: Aldine, pp. 3134.Google Scholar
Vita-Finzi, C. (1978). Archaeological Sites in their Setting. London: Thames and Hudson.Google Scholar
Vita-Finzi, C. (2008). The Sun: A User’s Manual. New York: Springer Science+Business Media.Google Scholar
Vittimberga, P. & Cardello, R. (1963). Sédimentologie et pétrographie du Bassin du Paléozoïque de Kufra. Premier Symposium Saharien, Tripoli, 1963. Revue de l’Institut Français du Pétrole, 18(10), 228240.Google Scholar
Vizy, E. K. & Cook, K. H. (2003). Connections between the summer East African and Indian rainfall regimes. Journal of Geophysical Research, 108(D16), 4510, 129.Google Scholar
Vrba, E. S., Denton, G. H., Partridge, T. C. & Burckle, L. H. (eds.) (1995). Paleoclimate and Evolution, with Emphasis on Human Origins. New Haven, CT: Yale University Press.Google Scholar
Wagner, B., Wennrich, V., Viehberg, F., et al. (2018). Holocene rainfall runoff in the central Ethiopian highlands and evolution of the River Nile drainage system as revealed from a sediment record from Lake Dendi. Global and Planetary Change, 163, 2943.Google Scholar
Walker, G. T. (1924). Correlations in seasonal variations of weather. IX: A further study of world weather. Memoirs of the Indian Meteorological Department, 24, 275332.Google Scholar
Walker, M. (2005). Quaternary Dating Methods. Chichester: John Wiley & Sons.Google Scholar
Walker, M., Johnsen, S., Rasmussen, S. O., et al. (2009). Formal definition and dating of the GSSP (Global Stratotype Section and Point) for the base of the Holocene using the Greenland NGRIP ice core, and selected auxiliary records. Journal of Quaternary Science, 24(1), 317.Google Scholar
Walker, M. J. C., Berkelhammer, M., Bjorck, S., et al. (2012). Formal subdivision of the Holocene Series/Epoch: A Discussion Paper by a Working Group of INTIMATE (Integration of ice-core, marine and terrestrial records) and the Subcommission on Quaternary Stratigraphy (International Commission on Stratigraphy). Journal of Quaternary Science, 27, 649659.Google Scholar
Walter, R. C., Buffler, R. T., Briggemann, J. H., et al. (2000). Early human occupation of the Red Sea coast of Eritrea during the last interglacial. Nature, 405, 6569.Google Scholar
Warren, A. (1970). Dune trends and their implications in the central Sudan. Zeitschrift für Geomorphologie, Neue Folge, Supplementband, 10, 154180.Google Scholar
Warren, A. (2013). Dunes: Dynamics, Morphology, History. Royal Geographical Society with IBG. Oxford: Wiley-Blackwell.Google Scholar
Warren, A. & Khogali, M. (1992). An Assessment of Desertification and Drought in the Sudano-Sahelian region 1985–1991. New York: UNSO.Google Scholar
Washbourn-Kamau, C. K. (1971). Late Quaternary lakes in the Nakuru-Elmenteita Basin, Kenya. Geographical Journal, 137, 522534.Google Scholar
Washburn, A. L. (1973). Periglacial Processes and Environments. London: Edward Arnold.Google Scholar
Washburn, A. L. (1979). Geocryology: A Survey of Periglacial Processes and Environments. London: Edward Arnold.Google Scholar
Washington, R., Todd, M. C., Lizcano, G., et al. (2006). Links between topography, wind, deflation, lakes and dust: The case of the Bodele Depression, Chad. Geophysical Research Letters, 33(9), L09401.Google Scholar
Wasson, R. J., Rajaguru, S. N., Misra, V. N., et al. (1983). Geomorphology, late Quaternary stratigraphy and palaeoclimatology of the Thar dunefield. Zeitschrift für Geomorphologie N.F., 45, 117151.Google Scholar
Waters, C. N. and 23 others (2016). The Anthropocene is functionally and stratigraphically distinct from the Holocene. Science, 351, 137.Google Scholar
Watson, E., Forster, P., Richards, M. & Bandelt, H.-J. (1997). Mitochondrial footprints of human expansions in Africa. American Journal of Human Genetics, 61, 691704.Google Scholar
Watson, J. P. (1961). Some observations on soil horizons and insect activity in granite soils. In Proceedings of the 1st Federal Science Congress, 1960, Salisbury, pp. 16.Google Scholar
Watson, J. P. (1962). The soil below a termite mound. Journal of Soil Science, 13, 4651.Google Scholar
Watson, J. P. (1964). A soil catena on granite in Southern Rhodesia. I. Field observations. Journal of Soil Science, 15, 238257.Google Scholar
Webster, P. J. (2004a). The elementary Hadley circulation. In Diaz, H. F. & Bradley, R. S. (eds.), The Hadley Circulation: Present, Past and Future. Dordrecht: Kluwer Academic, pp. 960.Google Scholar
Webster, P. J. (2004b). The coupled monsoon system. In Wang, B. (ed.), The Asian Monsoon. Berlin: Springer, pp. 366.Google Scholar
Webster, R. (1965). A catena of soils on the Northern Rhodesian Plateau. Journal of Soil Science, 16, 3143.Google Scholar
Wehausen, R. & Brumsack, H.-J. (1998). The formation of Pliocene Mediterranean sapropels: Constraints from high-resolution major and minor element studies. In Robertson, A. H. F., Emeis, K.-C., Richter, C. & Camerlengui, A. (eds.), Proceedings of the Ocean Drilling Program, Scientific Results,160, 207217.Google Scholar
Weiss, H. (2000). Beyond the Younger Dryas: Collapse as adaptation to abrupt climate change in ancient West Asia and the Eastern Mediterranean. In Bawden, G. & Reycraft, R. M. (eds.), Environmental Disaster and the Archaeology of Human Response, Maxwell Museum of Anthropology, University of New Mexico, Albuquerque. Anthropological Papers No. 7, pp. 7598.Google Scholar
Wells, J. (1921). Road construction across the cotton soils of the southern Sudan. Sudan Notes and Records, 4, 108111.Google Scholar
Welsby, D. A., Macklin, M. G. & Woodward, J. C. (2002). Human responses to Holocene environmental changes in the northern Dongola reach of the Nile, Sudan. In Friedman, R. (ed.), Egypt and Nubia: Gifts of the Desert. London: British Museum Press, pp. 2838.Google Scholar
Welsby, D. A. & Phillipson, D. W. (2008). Empires of the Nile. London: Folio Society.Google Scholar
Wendorf, F. (ed.) (1968). The Prehistory of Nubia. Dallas, TX: Southern Methodist University.Google Scholar
Wendorf, F., Close, A. E. & Schild, R. (1987). Early domestic cattle in the eastern Sahara. Palaeolecology of Africa, 18, 441448.Google Scholar
Wendorf, F. & Schild, R. (1974). A Middle Stone Age Sequence from the Central Rift Valley, Ethiopia. Wroclaw, Poland: Ossolineum.Google Scholar
Wendorf, F. & Schild, R. (1976). Prehistory of the Nile Valley. New York: Academic Press.Google Scholar
Wendorf, F. & Schild, R. (1980). Prehistory of the Eastern Sahara. New York: Academic Press.Google Scholar
Wendorf, F. & Schild, R. (1984a). Introduction. In Wendorf, F., Schild, R. & Close, A. E. (eds.), Cattle Keepers of the Eastern Sahara: The Neolithic of Bir Kiseiba. Dallas, TX: Department of Anthropology, Southern Methodist University, pp. 18.Google Scholar
Wendorf, F. & Schild, R. (1984b). The emergence of food production in the Egyptian Sahara. In Clark, J. D. & Brandt, S. A. (eds.), From Hunters to Farmers: The Causes and Consequences of Food Production in Africa. Berkeley, CA: University of California Press, pp. 93101.Google Scholar
Wendorf, F., Schild, R. & Close, A. E. (eds.) (1984a). Cattle Keepers of the Eastern Sahara: The Neolithic of Bir Kiseiba. Dallas, TX: Department of Anthropology, Southern Methodist University.Google Scholar
Wendorf, F., Schild, R., Close, A. E., et al. (1984b). New radiocarbon dates on the cereals from Wadi Kubbaniya. Science, 225, 645646.CrossRefGoogle ScholarPubMed
Wendorf, F., Schild, R. & Close, A. (eds.) (1993a). Egypt During the Last Interglacial: The Middle Paleolithic of Bir Tarfawi and Bir Sahara East. New York: Plenum Press.Google Scholar
Wendorf, F., Schild, R. & Close, A. (1993b). Summary and conclusions. In Wendorf, F., Schild, R., Close, A. E., et al. (eds.), Egypt During the Last Interglacial: The Middle Palaeolithic of Bir Tarfawi and Bir Sahara East. New York: Plenum Press, pp. 552573.Google Scholar
Wendorf, F., Schild, R., El Hadidi, N., et al. (1979). The use of barley in the Egyptian late Palaeolithic. Science, 205, 13411347.Google Scholar
Wendrich, W., Holdaway, S. J., Phillipps, R. & Emmitt, J. J. (2017a). The K Basin archaeological record. In Holdaway, S. J. & Wendrich, W. (eds.), The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press, pp. 99164.Google Scholar
Wendrich, W., Phillipps, R., Holdaway, S. J., Linseele, V., Emmitt, J. J. & Marston, J. M. (2017b). Kom K. In Holdaway, S. J. & Wendrich, W. (eds.), The Desert Fayum Reinvestigated: The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Los Angeles, CA: UCLA Cotsen Institute of Archaeology Press, pp. 165212.Google Scholar
Wenke, R. J. (1984). Early agriculture in the Southern Fayum Depression: Some test survey results and research implications. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Origin and Early Food Development of Food-Producing Cultures in North-Eastern Africa. Poznan, Poland: Polish Academy of Sciences and Poznan Archaeological Museum, pp. 193198.Google Scholar
Wenke, R. J. (1991). The evolution of early Egyptian civilization: Issues and evidence. Journal of World Prehistory, 5, 279329.Google Scholar
Wenke, R. J. & Casini, M. (1989). The Epipalaeolithic-Neolithic transition in Egypt’s Fayum Depression. In Krzyzaniak, L. & Kobusiewicz, M. (eds.), Late Prehistory of the Nile Basin and the Sahara. Poznan, Poland, Poznan Archaeological Museum, pp. 139155.Google Scholar
Werdecker, J. (1967). Map of Hoch Semyen (Äthiopien). 1:50,000 scale. Erdkunde, 12/1, Beilage II, printed in Bayerisches Landesvermessungsamt, Munich.Google Scholar
Wescott, W. A., Morley, C. K. & Karanja, F. M. (1996). Tectonic controls on the development of Rift-Basin lakes and their sedimentary character: Examples from the East African Rift System. In Johnson, T.C. & Odada, E. O. (eds.), The Limnology, Climatology and Paleoclimatology of the East African Rift Lakes. Amsterdam: Gordon and Breach, pp. 321.Google Scholar
West, R. G. (1977). Pleistocene Geology and Biology, with Especial Reference to the British Isles, 3rd edn. London: Longman.Google Scholar
Western, D. (1997). In the Dust of Kilimanjaro. Washington, DC: Island Press.Google Scholar
Whetton, P., Adamson, D. & Williams, M. (1990). Rainfall and river flow variability in Africa, Australia and East Asia linked to El Niño – Southern Oscillation events. In Bishop, P. (ed.), Lessons for Human Survival: Nature’s Record from The Quaternary. Geological Society of Australia Symposium Proceedings, 1, 7182.Google Scholar
Whetton, P., Allan, R. & Rutherfurd, I. (1996). Historical ENSO teleconnections in the Eastern Hemisphere: Comparison with the latest El Niño series of Quinn. Climatic Change, 32, 103109.Google Scholar
Whetton, P. H. & Rutherfurd, I. (1994). Historical ENSO teleconnections in the Eastern Hemisphere. Climatic Change, 28, 221253.Google Scholar
White, T. D., Asfaw, B., DeGusta, D., et al. (2003). Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature, 423, 742747.Google Scholar
Whiteman, A. J. (1971). The Geology of the Sudan Republic. London: Clarendon Press.Google Scholar
Whyte, R. O. (1951). Management and conservation in the Anglo-Egyptian Sudan. Commonwealth Bureau of Pastures and Field Crops Bulletin, 41, 6979.Google Scholar
Wickens, G. E. (1975a). Quaternary plant fossils from the Jebel Marra volcanic complex and their palaeoclimatic interpretation. Palaeogeography, Palaeoeclimatology, Palaeoecology, 17, 109122.Google Scholar
Wickens, G. E. (1975b). Changes in the climate and vegetation of the Sudan since 20 000 B.P. Boissiera, 24, 4365.Google Scholar
Wickens, G. E. (1976a). The Flora of Jebel Marra (Sudan Republic) and Its Geographical Affinities. Kew Bulletin Additional Series V, HMSO.Google Scholar
Wickens, G. E. (1976b). Speculations on long distance dispersal and the flora of Jebel Marra, Sudan. Kew Bulletin, 31(1), 105150.Google Scholar
Wickens, G. E. (1982). Palaeobotanical speculations and Quaternary environments in the Sudan. In Williams, M. A. J. & Adamson, D. A. (eds.), A Land between Two Niles: Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema, pp. 2350.Google Scholar
Wickens, G. E. (1984). Flora. In Cloudsley-Thompson, J. L. (ed.), Key Environments: Sahara Desert. Oxford: Pergamon, pp. 6775.Google Scholar
Willcocks, W. (1904). The Nile in 1904. London: E. & F.N. Spon.Google Scholar
Williams, F. M. (2016). Understanding Ethiopia: Geology and Scenery. Dordrecht: Springer Science+Business Media.Google Scholar
Williams, F. M., Williams, M. A. J. & Aumento, F. (2004). Tensional fissures and crustal extension rates in the northern part of the Main Ethiopian Rift. Journal of African Earth Sciences, 38, 183197.Google Scholar
Williams, M. (2012a). River sediments. In Vita-Finzi, C. (ed.), River History. Philosophical Transactions of the Royal Society of London, Series A, 370, 20932122.Google Scholar
Williams, M. (2012b). Geomorphology and Quaternary geology of the region between Wadi el Arab and Kerma. Kerma, Soudan 2011–2012, Université de Neuchâtel, Documents de la mission archéologique suisse au Soudan, 2012/4, 1015.Google Scholar
Williams, M. (2014). Climate Change in Deserts: Past, Present and Future. New York: Cambridge University Press.Google Scholar
Williams, M. (2015). Interactions between fluvial and eolian geomorphic systems and processes: Examples from the Sahara and Australia. Catena, 134, 413.Google Scholar
Williams, M. & Jacobsen, G. E. (2011). A wetter climate in the desert of northern Sudan 9900–7600 years ago. Sahara, 22, 714.Google Scholar
Williams, M., McCarthy, M. & Pickup, G. (1995). Desertification, drought and landcare: Australia’s role in an international convention to combat desertification. Australian Geographer, 26, 2332.Google Scholar
Williams, M., Nitschke, N. & Chor, C. (2006). Complex geomorphic response to late Pleistocene climatic changes in the arid Flinders Ranges of South Australia. Géomorphologie: Relief, processus, environnement, 4, 249258.Google Scholar
Williams, M. & Nottage, J. (2006). Impact of extreme rainfall in the central Sudan during 1999 as a partial analogue for reconstructing early Holocene prehistoric environments. Quaternary International, 150(1), 8294.Google Scholar
Williams, M., Prescott, J. R., Chappell, J., Adamson, D., Cock, B., Walker, K. & Gell, P. (2001). The enigma of a late Pleistocene wetland in the Flinders Ranges, South Australia. Quaternary International, 83–85, 129144.Google Scholar
Williams, M., Talbot, M., Aharon, P., Abdl Salaam, Y., Williams, F. & Brendeland, K. I. (2006). 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, 26512665.Google Scholar
Williams, M. A. J. (1966). Age of alluvial clays in the western Gezira, Republic of the Sudan. Nature, 211, 270271.Google Scholar
Williams, M. A. J. (1968a). Termites and soil development near Brocks Creek, Northern Territory. Australian Journal of Science, 31, 153154.Google Scholar
Williams, M. A. J. (1968b). Soil salinity in the west central Gezira, Republic of the Sudan. Soil Science, 105, 451464.Google Scholar
Williams, M. A. J. (1968c). A dune catena on the clay plains of the west central Gezira, Republic of the Sudan. Journal of Soil Science, 19, 367378.Google Scholar
Williams, M. A. J. (1969). Prediction of rainsplash erosion in the seasonally wet tropics. Nature, 222, 763765.Google Scholar
Williams, M. A. J. (1975). Late Pleistocene tropical aridity synchronous in both hemispheres? Nature, 253, 617618.Google Scholar
Williams, M. A. J. (1978). Termites, soils and landscape equilibrium in the Northern Territory of Australia. In Davies, J. L. & Williams, M. A. J. (eds.), Landform Evolution in Australasia. Canberra: Australian National University Press, pp. 128141.Google Scholar
Williams, M. A. J. (1984). Geology. In Key Environments: Sahara Desert, Cloudsley-Thompson, J. L. (ed.). Oxford, Pergamon, pp. 3139.Google Scholar
Williams, M. A. J. (2008). Geology, geomorphology and prehistoric environments. In Gifford-Gonzalez, D. (ed.), Adrar Bous: Archaeology of a Central Saharan Granitic Ring Complex in Niger. Royal Museum for Central Africa. Tervuren, Belgium, pp. 2554.Google Scholar
Williams, M. A. J. (2009a). Late Pleistocene and Holocene environments in the Nile basin. Global and Planetary Change, 69, 115.Google Scholar
Williams, M. A. J. (2009b). Human impact on the Nile Basin: Past, Present, Future. In Dumont, H. J. (ed.), The Nile: Origin, Environments, Limnology and Human Use. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 771779.Google Scholar
Williams, M. A. J. (2013). Quaternary Geology and Environmental History of the Nile in the Sudan. Bulletin No. 42, 1–77, Geological Research Authority of the Sudan, Ministry of Energy & Mining, Khartoum, Sudan.Google Scholar
Williams, M. A. J., Abell, P. I. & Sparks, B. W. (1987). Quaternary landforms, sediments, depositional environments and gastropod isotope ratios at Adrar Bous, Tenere Desert of Niger, south-central Sahara. In Frostick, L. & Reid, I. (eds.), Desert Sediments: Ancient and Modern. Geological Society Special Publication No. 35, pp. 105125.Google Scholar
Williams, M. A. J. & Adamson, D. A. (1973). The physiography of the central Sudan. Geographical Journal, 139, 498508.Google Scholar
Williams, M. A. J. & Adamson, D. A. (1974). Late Pleistocene desiccation along the White Nile. Nature, 248, 584586.Google Scholar
Williams, M. A. J. & Adamson, D. A. (1980). Late Quaternary depositional history of the Blue and White Nile Rivers in central Sudan. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 281304.Google Scholar
Williams, M. A. J. & Adamson, D. A. (eds.) (1982). A Land between Two Niles: Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema.Google Scholar
Williams, M. A. J. & Adamson, D. A. (2008). A biophysical model for the formation of late Pleistocene valley-fills in the arid Flinders Ranges of South Australia. South Australian Geographical Journal, 107, 114.Google Scholar
Williams, M. A. J., Adamson, D. A. & Abdulla, H. H. (1982). Landforms and soils of the Gezira: A Quaternary legacy of the Blue and White Nile rivers. In Williams, M. A. J. & Adamson, D. A. (eds.), A Land between Two Niles: Quaternary Geology and Biology of the Central Sudan. Rotterdam: A. A. Balkema, pp. 111142.Google Scholar
Williams, M. A. J., Adamson, D., Cock, B. & McEvedy, R. (2000). Late Quaternary environments in the White Nile region, Sudan. Global and Planetary Change, 26, 305316.Google Scholar
Williams, M. A. J., Adamson, D., Prescott, J. R. & Williams, F. M. (2003). New light on the age of the White Nile. Geology, 31, 10011004.Google Scholar
Williams, M. A. J., Adamson, D. A., Williams, F. M., Morton, W. H. & Parry, D. E. (1980). Jebel Marra volcano: A link between the Nile Valley, the Sahara and Central Africa. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 305337.Google Scholar
Williams, M. A. J., Ambrose, S. H., van der Kaars, S., et al. (2009a). Environmental impact of the 73 ka Toba super-eruption in South Asia. Palaeogeography, Palaeoclimatology, Palaeoecology, 284, 295314.Google Scholar
Williams, M. A. J., Assefa, G. & Adamson, D. A. (1986). Depositional context of Plio-Pleistocene hominid-bearing formations in the Middle Awash Valley, southern Afar Rift, Ethiopia. In Frostick, L., Renaut, R., Reid, I. & Tiercelin, J. J. (eds.), Sedimentation in the African Rifts. Geological Society Special Publication No. 25. Oxford: Blackwell Scientific, pp. 233243.Google Scholar
Williams, M. A. J. & Balling, R. C. Jr. (1996). Interactions of Desertification and Climate. London: Arnold, with WMO and UNEP.Google Scholar
Williams, M. A. J., Bishop, P. M., Dakin, F. M. & Gillespie, R. (1977). Late Quaternary lake levels in southern Afar and the adjacent Ethiopian Rift. Nature, 267, 690693.Google Scholar
Williams, M. A. J., Clark, J. D., Adamson, D. A. and Gillespie, R. (1975). Recent Quaternary research in central Sudan. Bulletin de l’ASEQUA, 46, 7586.Google Scholar
Williams, M. A. J., Duller, G. A. T., Williams, F. M., et al. (2015a). Causal links between Nile floods and eastern Mediterranean sapropel formation during the past 125 kyr confirmed by OSL and radiocarbon dating of Blue and White Nile sediments. Quaternary Science Reviews, 130, 89108.Google Scholar
Williams, M., Dunkerley, D., De Deckker, P., Kershaw, P. & Chappell, J. (1998). Quaternary Environments, 2nd edn. London: Arnold.Google Scholar
Williams, M. A. J., Dunkerley, D. L., De Deckker, P., Kershaw, A. P. and Stokes, T. (1993). Quaternary Environments, 1st edn. London: Edward Arnold.Google Scholar
Williams, M. A. J. & Faure, H. (eds.) (1980). The Sahara and the Nile: Quaternary environments and prehistoric occupation in northern Africa. Rotterdam: A. A. Balkema.Google Scholar
Williams, M. A. J. & Hall, D. N. (1965). Recent expeditions to Libya from the Royal Military Academy, Sandhurst. Geographical Journal, 131, 482501.Google Scholar
Williams, M. A. J., Medani, A. H., Talent, J. A. & Mawson, R. (1974). A note on upper Quaternary mollusca west of Jebel Aulia. Sudan Notes and Records, 54, 168172.Google Scholar
Williams, M. A. J., Street, F. A. & Dakin, F. M. (1978). Fossil periglacial deposits in the Semien Highlands, Ethiopia. Erdkunde, 32, 4046.Google Scholar
Williams, M. A. J. & Talbot, M. R. (2009). Late Quaternary environments in the Nile Basin. In Dumont, H. J. (ed.), The Nile: Origin, Environments, Limnology and Human Use. Monographiae Biologicae, 89. Dordrecht: Springer Science+Business Media, pp. 6172.Google Scholar
Williams, M. A. J., Usai, D., Salvatori, S., et al. (2015b). Late Quaternary environments and prehistoric occupation in the lower White Nile valley, central Sudan. Quaternary Science Reviews, 130, 7288.Google Scholar
Williams, M. A. J. & Williams, F. M. (1980). Evolution of the Nile Basin. In Williams, M. A. J. & Faure, H. (eds.), The Sahara and the Nile. Rotterdam: A. A. Balkema, pp. 207224.Google Scholar
Williams, M. A. J., Williams, F. M. & Bishop, P. (1981). Late Quaternary history of Lake Besaka, Ethiopia. Palaeoecology of Africa, 13, 93104.Google Scholar
Williams, M. A. J., Williams, F. M., Duller, G. A. T., et al. (2010). Late Quaternary floods and droughts in the Nile Valley, Sudan: New evidence from optically stimulated luminescence and AMS radiocarbon dating. Quaternary Science Reviews, 29, 11161137.Google Scholar
Williams, M. A. J., Williams, F. M., Gasse, F., Curtis, G. H. & Adamson, D. A. (1979). Plio-Pleistocene environments at Gadeb prehistoric site, Ethiopia. Nature, 282, 2933.Google Scholar
Williamson, D., Jackson, M., Banerjee, S. K. & Petit-Maire, N. (2004). The magnetism of a glacial aeolianite from Lanzarote (Canary Islands): Coupling between luvic calcisol formation and Saharan dust trapping processes during wet deposition events off northwestern Sahara. Geophysical Journal International, 157, 10901104.Google Scholar
Williamson, P. G. (1982). Molluscan biostratigraphy of the Koobi Fora hominid-bearing deposits. Nature, 295, 140142.Google Scholar
Wilson, I. G. (1971). Desert sandflow basins and a model for the development of ergs. Geographical Journal, 137, 180199.Google Scholar
Winchell, F., Stevens, C. J., Murphy, C., Champion, L. & Fuller, D. Q. (2017). Evidence for sorghum domestication in fourth millennium BC eastern Sudan. Current Anthropology, https://doi.org/10.1086/693898Google Scholar
Winkler, H. A. (1939a). Archaeological Survey of Egypt, Rock Drawings of Southern Upper Egypt, II. London: The Sir Robert Mond Expedition.Google Scholar
Winkler, H. A. (1939b). An Expedition to the Gilf Kebir and ‘Uweinat, 1938. 5. Rock-pictures at ‘Uweinat. Geographical Journal, 93, 307310.Google Scholar
Winton, T. (2015). Island Home: A Landscape Memoir. NSW, Australia: Penguin Random House.Google Scholar
WoldeGabriel, G., Ambrose, S. H., Barboni, D., Bonnefille, R., Bremond, L., Currie, B., DeGusta, D., Hart, W.K., Murray, A., Renne, P., Jolly-Saad, M.C., Stewart, K.M., & White, T.D. (2009). The geological, isotopic, botanical, invertebrate and lower vertebrate surroundings of Ardipithecus ramidus. Science, 65, 65e165e5.Google Scholar
WoldeGabriel, G., Haile-Lelassie, Y., Renne, P. R., et al. (2001). Geology and palaeontology of the Late Miocene Middle Awash valley, Afar rift, Ethiopia. Nature, 412, 175178.Google Scholar
WoldeGabriel, G., White, T. D., Suwa, G., et al. (1994). Ecological and temporal placement of early Pliocene hominids at Aramis, Ethiopia. Nature, 371,330333.Google Scholar
Wong, H. K. & Zarudski, E. F. K. (1969). Thickness of unconsolidated sediments in the Eastern Mediterranean Sea. Bulletin of the Geological Society of America, 80, 26112614.Google Scholar
Woodward, J., Macklin, M., Fielding, L., et al. (2015a). Shifting sediment sources in the world’s longest river: A strontium isotope record for the Holocene Nile. Quaternary Science Reviews, 130, 124140.Google Scholar
Woodward, J. C., Macklin, M. G., Krom, M. D. & Williams, M. A. J. (2007). The Nile: Evolution, Quaternary river environments and material fluxes. In Gupta, A. (ed.), Large Rivers: Geomorphology and Management. Chichester: John Wiley & Sons, pp. 261292.Google Scholar
Woodward, J. C., Macklin, M. G. & Welsby, D. (2001). The Holocene fluvial sedimentary record and alluvial geoarchaeology in the Nile valley of northern Sudan. In Maddy, D. R., Macklin, M. G. & Woodward, J. C. (eds.), River Basin Sediment Systems: Archives of Environmental Change. Rotterdam: A. A. Balkema, pp. 327355.Google Scholar
Woodward, J. C., Williams, M. A. J., Garzanti, E., Macklin, M. G. and Marriner, N. (2015b). From source to sink: Exploring the Quaternary history of the Nile. Quaternary Science Reviews, 130, 38.Google Scholar
Wright, D. K. (2018). East and southern African Neolithic: Geography and overview. In Smith, C. (ed.), Encyclopedia of Archaeology. New York: Springer Science+Business Media, DOI: 10.1007/978–3-319–51726-1_1888–2.Google Scholar
Wynn, J. G., Alemseged, Z., Bobe, R., Geraads, D., Reed, D. & Roman, D. C. (2006). Geological and palaeontological context of a Pliocene juvenile hominin at Dikika, Ethiopia. Nature, 443, 332336.Google Scholar
Yemane, K., Bonnefille, R. & Faure, H. (1985). Palaeoclimatic and tectonic implications of Neogene microflora from the north-western Ethiopian highlands. Nature, 318, 653656.Google Scholar
Zaki, A. S. & Giegengack, R. (2016). Inverted topography in the southeastern part of the Western Desert of Egypt. Journal of African Earth Sciences, 121, 5661.Google Scholar
Zaki, A. S., Pain, C. F., Edgett, K. S. & Giegengack, R. (2018). Inverted stream channels in the Western Desert of Egypt: Synergistic remote, field observations and laboratory analysis on Earth with applications to Mars. Icarus, 18, doi.org/10.1016/j.icarus.2018.03.001Google Scholar
Zboray, A. (2009). Rock art of the Libyan Desert, 2nd edn. (DVD). Newbury: Fliegel Jezernicky Expeditions Ltd.Google Scholar
Zerboni, A. (2005). Cambiamenti climatici olocenici nel Sahara central: nuovi archive paleoambientali. Unpublished doctoral research thesis, University of Milan.Google Scholar
Zerboni, A. (2008). Holocene rock varnish on the Messak plateau (Libyan Sahara): Chronology of weathering processes. Geomorphology, 102, 640651.Google Scholar
Zhao, Y., Colin, C., Liu, Z., Paterne, M., Siani, G. & Xie, X. (2012). Reconstructing precipitation changes in northeastern Africa during the Quaternary by clay mineralogical and geochemical investigations of Nile deep-sea fan sediments. Quaternary Science Reviews, 57, 5870.Google Scholar
Zhao, Y., Liu, Z., Colin, C., et al. (2011). Variations of the Nile suspended discharges during the last 1.75 Myr. Palaeogeography, Palaeoclimatology, Palaeoecology, 311, 230241.Google Scholar

Save book to Kindle

To save this book to your Kindle, first ensure coreplatform@cambridge.org is added to your Approved Personal Document E-mail List under your Personal Document Settings on the Manage Your Content and Devices page of your Amazon account. Then enter the ‘name’ part of your Kindle email address below. Find out more about saving to your Kindle.

Note you can select to save to either the @free.kindle.com or @kindle.com variations. ‘@free.kindle.com’ emails are free but can only be saved to your device when it is connected to wi-fi. ‘@kindle.com’ emails can be delivered even when you are not connected to wi-fi, but note that service fees apply.

Find out more about the Kindle Personal Document Service.

  • References
  • Martin Williams, University of Adelaide
  • Book: The Nile Basin
  • Online publication: 14 December 2018
  • Chapter DOI: https://doi.org/10.1017/9781316831885.025
Available formats
×

Save book to Dropbox

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Dropbox.

  • References
  • Martin Williams, University of Adelaide
  • Book: The Nile Basin
  • Online publication: 14 December 2018
  • Chapter DOI: https://doi.org/10.1017/9781316831885.025
Available formats
×

Save book to Google Drive

To save content items to your account, please confirm that you agree to abide by our usage policies. If this is the first time you use this feature, you will be asked to authorise Cambridge Core to connect with your account. Find out more about saving content to Google Drive.

  • References
  • Martin Williams, University of Adelaide
  • Book: The Nile Basin
  • Online publication: 14 December 2018
  • Chapter DOI: https://doi.org/10.1017/9781316831885.025
Available formats
×