Hostname: page-component-76fb5796d-5g6vh Total loading time: 0 Render date: 2024-04-26T08:53:11.108Z Has data issue: false hasContentIssue false

Problems in the Stone Age of South-east Asia Revisited

Published online by Cambridge University Press:  18 February 2014

T. E. G. Reynolds*
Affiliation:
Birkbeck College, Faculty of Continuing Education, University of London, 26 Russel Square, London, WC1. Email: te.reynolds@bbk.ac.uk

Abstract

In the 13 years since ‘Problems in the Stone Age of Southeast Asia’ was published, there has been a number of significant developments. There remains a lack of early cultural material despite the possibility that first occupation of the area may date back as far as 1.8 Myrs. It seems that the first hominins in the region were essentially ‘alithic’ in their adaptation, making the reconstruction of their behaviour extremely difficult. There is also a question as to which hominin was first ‘Out of Africa’ and into Asia and a suggestion that Homo erectus is, in fact, an Asian species that may have migrated west. This has important implications for interpretations of the significance of the so-called ‘Movius Line’. By the time stone tool use does appear regularly in the record, modern humans are present but it is still hard to identify the kinds of directional changes that are associated with the Late Pleistocene elsewhere in the world. The question of when humans were able to exploit tropical forests in the region is also one that recent work explores. The recent discoveries from Flores of stone tools that appear to pre-date the arrival of modern humans, and a possibly associated ‘dwarf’ hominin, Homo floresiensis, all require re-appraisal of the nature of human activity in the region.

Résumé

Au cours des 13 années écoulées depuis la publication de ‘Problèmes à l'âge de pierre en Asie du sud-est’, il y a eu un certain nombre de développements importants. La pénurie de matériel culturel primitive demeure malgré la possibilité que la première occupation de la région remonte à 1,8 millions d'années. Il semble que les premiers hominiens dans la région étaient essentiellement ‘alithiques’ dans leur adaptation, ce qui rend la reconstruction de leur comportement extrèmement difficile. Il reste également la question de savoir quel hominien fut le premier ‘venu d'Afrique’ à s'installer en Asie et l'hypothèse qu'Homo Erectus est, en fait, une espèce asiatique qui aurait pu émigrer vers l'ouest. Ceci a d'importantes implications pour l'interprétation de la soi-disant ‘lignée de Movius’. Au moment où apparaît régulièrement dans les archives l'utilisation des outils en pierre, la présence de l'homme moderne est avérée mais il est encore difficile d'identifier les sortes de changements de direction qu'on associe avec la fin de pléistocène ailleurs dans le monde. La question de savoir quand les humains ont été capables d'exploiter les forêts tropicales de la région est une de celles que de récents travaux explorent. Les récentes découvertes par Flores d'outils en pierre qui semblent pré-dater l'arrivée de l'homme moderne et l'existence possible d'un hominien ‘nain’, Homo Floresiensis, tout ceci nécessite une ré-évaluation de la nature des activités humaines de la région.

Résumen

En los 13 años transcurridos desde la publicación de “Problemas de la Edad de Piedra del Sureste de Asia’, se han producido un numero de desarrollos significativos. A pesar de la posibilidad de que la ocupación de la región se remonte hasta hace 1.8 millones de años, falta aún por descubrir evidencia de la cultura material temprana. Parece que los primeros homíninos de la región fueron esencialmente “alíticos” en su adaptación, lo que hace extremadamente difícil la reconstrucción de su comportamiento. Permanece también la cuestión de qué homínino fue el primero “Fuera de África” y en llegar a Asia, y la sugerencia de que Homo Erectus es, en realidad, una especie asiática que migró hacia el oeste. De ser verdad, habría importantes consecuencias para las interpretaciones de la relevancia de la ‘Línea Movius’. El hombre moderno ya está presente para cuando aparece de manera regular en el registro arqueológico el uso de instrumentos líticos, aunque es aún difícil identificar aquí los tipos de cambios direccionales asociados con el Tardo Pleistoceno en el resto del mundo. Las investigaciones recientes también han explorado si los hombres modernos pudieron explotar las selvas tropicales de la región. Los recientes descubrimientos de Flores, en los que instrumentos líticos que parecen preceder la llegada del hombre moderno, y su posible asociación con un homínino “enano”, el “Homo Floriensis”, requieren una reevaluación de la naturaleza de la actividad humana en la región.

Zusammenfassung

In den 13 Jahren seit der Veröffentlichung von ‘Problems in the Stone Age of Southeast Asia’ gab es eine Reihe bedeutender Entwicklungen. Die erste Landnahme dieser Region kann bis auf 1.8 Millionen Jahre zuröckdatiert werden, obwohl weiterhin fröhes Kulturmaterial fehlt. Die ersten Hominiden scheinen in ihrer Adaption tatsächlich, alithisch' gewesen zu sein, was die Rekonstruktion ihres Verhaltens extrem erschwert. Zudem muss geklärt werden, welcher Hominide der erste ‘Out of Africa’ in Asien war. Hier gibt es die These, dass Homo erectus tatsächlich eine asiatische Spezies ist, die eventuell nach Westen migrierte. Dies hat wichtige Implikationen för die Bedeutung der so genannten ‘Movius Linie’. Ab dem Zeitpunkt, wenn die Nutzung der Steingeräte regelmäßig in den archäologischen Quellen nachgewiesen ist, ist der moderne Mensch nachgewiesen, wobei es aber immer noch schwierig ist alle Veränderungen, die mit dem späten Pleistozän einhergehen, in anderen Teilen der Welt zu identifizieren. Die Frage, wann die Menschen fähig waren den tropischen Regenwald in dieser Region auszubeuten, wird ebenfalls untersucht. Die neuesten Entdeckungen von Steingeräten auf Flores scheinen dem Erscheinen des modernen Menschen voranzugehen. So muss der, mit diesen Funden möglicherweise assoziierte ‘Zwergen’ Hominide – Homo floresiensis – und die menschliche Verhaltensweise in dieser Region neu untersucht werden.

Type
Research Article
Copyright
Copyright © The Prehistoric Society 2007

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

Akazawa, T., Aoki, K. & Bar-Yosef, O. (eds). 1998. Neanderthals and Modern Humans in Western Asia. London: PlenumGoogle Scholar
Akazawa, T., Aoki, K. & Kimura, T. (eds). 1992. The Evolution and Dispersal of Modern Humans in Asia, 373–378. Tokyo: Hokusen-ShaGoogle Scholar
Allen, H. 1991. A review of the Late Pleistocene/early recent stone tool assemblages of Java. Bulletin of the Indo-Pacific Prehistory Association 11, 364710.7152/bippa.v11i0.11371Google Scholar
Anderson, D.D. 1997. Cave archaeology in Southeast Asia. Geoarchaeology 12, 607–3810.1002/(SICI)1520-6548(199709)12:6<607::AID-GEA5>3.0.CO;2-23.0.CO;2-2>Google Scholar
Anshari, G., Kershaw, A.P. & Kaars, S. van der. 2001. A late Pleistocene and Holocene pollen and charcoal record from peat swamp forest, Lake Sentarum Wildlife Reserve, West Kalimantan, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 213–2810.1016/S0031-0182(01)00246-2Google Scholar
Anton, S.C. & Swisher, C.C. 2004. Early dispersals of Homo from Africa. Annual Review of Anthropology 33, 271–9610.1146/annurev.anthro.33.070203.144024Google Scholar
Argue, D., Donlon, D., Groves, C. & Wright, R. 2006. Homo floresiensis: Microcephalic, pygmoid, Australopithecus, or Homo? Journal of Human Evolution 51, 360–7410.1016/j.jhevol.2006.04.013Google Scholar
Bailey, R.C., & Headland, T.N. 1991. The tropical rainforest: is it a productive environment for human foragers? Human Ecology 19, 261–8510.1007/BF00888748Google Scholar
Bailey, R.C., Head, G., Jenike, M., Owen, B., Reichtman, R. & Zechenter, E. 1989. Hunting and gathering in tropical rainforest: it is possible? American Anthropologist 91, 598210.1525/aa.1989.91.1.02a00040Google Scholar
Barker, G., Barton, H., Beavitt, P., Bird, M., Cole, F., Daly, P., Gilbertson, D., Hunt, C., Krigbaum, J., Lampert, C., Lewis, H., Manser, J., McLaren, S., Menotti, F., Paz., V., Piper, P., Pyatt, B., Rebett, R., Reynolds, T., Stephens, M., Thompson, J., Trickett, M. & Whittaker, P. 2002. Prehistoric foragers and farmers in southeast Asia; renewed investigations at Niah Cave, Sarawak. Proceedings of the Prehistoric Society 68, 147–6410.1017/S0079497X00001481Google Scholar
Barker, G., Barton, H., Bird, M., Cole, F., Daly, P., Dykes, A.Farr, L., Gilbertson, D., Higham, T., Hunt, C., Knight, S., Kurui, E., Lewis, H., Lloyd-Smith, L., Manser, J., McLaren, S., Menotti, F., Piper, P., Pyatt, B., Rabett, R., Reynolds, T., Shimmin, J., Thompson, J. & Trickett, M. 2003. The Niah Cave Project: the fourth (2003) season of fieldwork. Sarawak Museum Journal 58(79) n.s., 45–119Google Scholar
Barker, G., Gilbertson, D. & Reynolds, T. (eds). 2005. The Human Use of Caves in Southeast Asia. Asian Perspectives 44 (1)Google Scholar
Barker, G., Gilbertson, D. & Reynolds, T. (eds) forthcoming. The Niah Cave ProjectGoogle Scholar
Barton, H. 2005. The case for rainforest foragers: the starch record at Niah Cave, Sarawak. In Barker, et al. 2005, 5672Google Scholar
Bartstra, G.J. 1985 Sangiran, the stone implements of Ngebung, and the Palaeolithic of Java. Modern Quaternary Research in Southeast Asia 9, 99113Google Scholar
Bartstra, G.J. & Basoeki, . 1989. Recent work on the Pleistocene and the Palaeolithic of Java. Current Anthropology 30, 241–4Google Scholar
Bartstra, G.J., Keates, S.G., Basoeki, & Kallupa, B. 1991. On the dispersion of Homo sapiens in eastern Indonesia: the Palaeolithic of South Sulawesi. Current Anthropology 32, 317–2110.1086/203960Google Scholar
Bartstra, G.J., Hooijer, D.A., Kallupa, B. & Anwar Akib, M. 1994. Notes on fossil vertebrates and stone tools from Sulawesi, Indonesia, and the stratigraphy of the Northern Wallanae Depression. Palaeohistoria 33/4, 118Google Scholar
Bar-Yosef, O & Pilbeam, D. (eds). 2000. The Geography of Neanderthals and Modern Humans in Europe and the Greater Mediterranean. Harvard University, Peabody Museum Bulletin 8Google Scholar
Beavitt, P. 1992. Exotic animal products and Chinese trade with Borneo. Anthropozoologica 16, 181–8Google Scholar
Bellwood, P 1979. Man’s Conquest of the Pacific. The Prehistory of Southeast Asia and Oceania. New York: Oxford University PressGoogle Scholar
Bellwood, P. 1984. Archaeological research in the Madai-Baturong Region, Sabah. Indo-Pacific Prehistory Association Bulletin 5, 385810.7152/bippa.v5i0.11223Google Scholar
Bellwood, P. 1990. Foraging towards farming: a decisive transition or a millennial blur? Review of Archaeology 11(2), 1424Google Scholar
Bellwood, P. 1996. The origins and spread of agriculture in the Indo-Pacific region: gradualism and diffusion or revolution and colonisation? In Harris, D.R. (ed.), The Origins and Spread of Agriculture and Pastoralism in Eurasia, 465–98. London: University College PressGoogle Scholar
Bellwood, P. 2001. Early agriculturalist population diasporas? Farming, languages, and genes. Annual Review of Anthropology 30, 18120710.1146/annurev.anthro.30.1.181Google Scholar
Bergh, G.D. van den, De Vos, J., Sondaar, P.Y. & Aziz, F. 1996a. Pleistocene of zoogeographic evolution of Java (Indonesia) and glacio-eustatic sea level fluctuations: a background for the presence of Homo. Bulletin of the Indo-Pacific Prehistory Association 14, 721Google Scholar
Bergh, G.D. van den, Mubroto, B., Aziz, F., Sondaar, P.Y, & De Vos, J. 1996b. Did Homo erectus reach the island of Flores? Bulletin of the Indo-Pacific Prehistory Association 14, 2736Google Scholar
Bergh, G.D. van den, De Vos, J. & Sondaar, P.Y. 2001. The Late Quaternary palaeogeography of mammal evolution in the Indonesian Archipelago. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 38540810.1016/S0031-0182(01)00255-3Google Scholar
Bettis, E.A. III, Zaim, Y., Larick, R., Ciochon, R.L., Suminto, R., Rizal, Y., Reagan, M. & Heizler, M. 2004. Landscape development preceding Homo erectus immigration into Central Java, Indonesia: the Sangiran Formation Lower Lahar. Palaeogeography, Palaeoclimatology and Palaeoecology 206, 115–3110.1016/j.palaeo.2004.01.016Google Scholar
Bird, M., Prentice, C., Taylor, D. & Hunt, C. 2005. A savannah corridor in Sundaland? Quaternary Science Reviews 24, 2228–4210.1016/j.quascirev.2005.04.004Google Scholar
Blench, R. 2004. Fruits and arboriculture in the Indo-Pacific Region. Bulletin of the Indo-Pacific Prehistory Association 24, 3150 (Indo-Pacific Prehistory: the Taipei Papers 2)Google Scholar
Bowdler, S. 1992. Homo sapiens in Southeast Asia and the Antipodes: archaeological versus biological interpretations. In Akazawa, et al. (eds) 1992, 559–89Google Scholar
Bowdler, S. 1996. The human colonisation of Sunda and Sahul: cultural and behavioural considerations. Bulletin of the Indo-Pacific Prehistory Association 14, 374210.7152/bippa.v14i0.11586Google Scholar
Bowler, J.M., Johnson, H., Olley, J.M., Prescott, J.R., Roberts, R.G., Shawcross, W. & Spooner, N.A. 2003. New ages for human occupation and climatic changes at Lake Mungo, Australia. Nature 421, 837–40Google Scholar
Brauer, G. 1992. The origins of modern Asians: by regional evolution or by replacement? In Akazawa, et al. (eds) 1992, 401–13Google Scholar
Brosius, P.J. 1991. Foraging in tropical rain forests: the case of the Penan of Sarawak, East Malaysia (Borneo). Human Ecology 19, 123–5010.1007/BF00888743Google Scholar
Brown, P, Sutkikna, T., Morwood, M.J., Soejono, R.P., Jatmiko, ,Wayhu Saptomo, E. & Due, Rocus Awe. 2004. A new small-bodied hominin from the Late Pleistocene of Flores, Indonesia. Nature 431, 1055–61Google Scholar
Brumm, A., Aziz, F., Bergh, G.D. van den, Morwood, M.J., Moore, M.W., Kurniawan, I., Hobbs, D.R. & Fullagar, R. 2006. Early stone technology on Flores and its implications for Homo floresiensis. Nature 441, 624–8Google Scholar
Bulbeck, F.D. 2002. Hunter-Gatherer occupation of the Malay Peninsula from the Ice Age to the Iron Age. In Mercader, J (ed.) 2002b, 119–60Google Scholar
Chappell, J. 2000. Pleistocene seedbeds of western Pacific maritime cultures and the importance of chronology. Modern Quaternary Research in Southeast Asia 16, 7798Google Scholar
Choi, K. & Driwantoro, D. 2007. Shell tool use by early members of Homo erectus in Sangiran, central Java, Indonesia: cut mark evidence. Journal of Archaeological Science 34, 485810.1016/j.jas.2006.03.013Google Scholar
Ciochon, R.L., Long, V.T., Larida, R., Gonzalez, L., Grun, R., De Vos, J., Yonge, C., Taylor, L., Yoshida, H. & Reagan, M. 1996. Dated co-occurrence of Homo erectus and Gigantopithecus from Tham Khuyen Cave, Vietnam. Proceedings of the National Academy of Science, USA 93. 3016–2010.1073/pnas.93.7.3016Google Scholar
Coles, B. 1998. Doggerland: a speculative survey. Proceedings of the Prehistoric Society 64, 458110.1017/S0079497X00002176Google Scholar
Colinvaux, P.A. & Bush, M.B. 1991. The rain-forest ecosystem as a resource for hunting and gathering. American Anthropologist 93, 153–6210.1525/aa.1991.93.1.02a00100Google Scholar
Cranbrook, , Earl of. 2000. Northern Borneo environments of the past 40,000 years; Archaeozoological evidence. Sarawak Museum Journal 55, 61110Google Scholar
Culot, E. 2006. How the Hobbit shrugged: tiny hominid's story takes new turn. Review of American Paleoanthropology Society meeting, 24–26 April 2006, San Juan, Puerto Rico. Science 312, 983–4Google Scholar
Davidson, I. & Noble, W. 1992. Why the first colonisation of the Australian region is the earliest evidence of modern human behaviour. Archaeology in Oceania 27(3), 135–4210.1002/j.1834-4453.1992.tb00297.xGoogle Scholar
Demeter, F., Bacon, A-M., Nguyen, K.T., Long, V.T.Durringer, P., Rousse, S., Coppens, Y., Matsumura, H., Dodo, Y., Nguyen, M.H. & Tomako, A. 2005. Discovery of a second human molar and maxilla fragment in the late Middle to Late Pleistocene cave of Ma U'Oi (northern Vietnam). Journal of Human Evolution 49, 39340210.1016/j.jhevol.2004.12.004Google Scholar
Denham, T. 2005. Envisaging early agriculture in the Highlands of New Guinea: landscapes, plants and practices. World Archaeology 37, 29030610.1080/00438240500095447Google Scholar
Dennell, R. & Roebroeks, W. 2005. An Asian perspective on early human dispersal from Africa. Nature 438, 1099–10410.1038/nature04259Google Scholar
De Vos, J.Sondaar, P. & Swisher, C.C. 1994. Dating hominid sites in Indonesia. Science 266, 1722–710.1126/science.7992059Google Scholar
Endicott, K. & Bellwood, P. 1991. The possibility of independent foraging in the rainforest of peninsular Malaysia. Human Ecology 19, 151–8510.1007/BF00888744Google Scholar
Everett, A.H., Evans, J. & Busk, G. 18791980. Report on the exploration of the caves of Borneo: and introductory remarks; and a note on the bones collected. Proceedings of the Royal Society of London 30, 310–19Google Scholar
Falk, D., Hildebolt, C., Smith, K., Morwood, M.J., Sutikna, T., Brown, P., Jatmiko, ,Wayu Saptomo, E., Brunsden, B. & Prior, F. 2005. The brain of LB1, Homo floresiensis. Science 308, 242–5Google Scholar
Fifield, L.K., Bird, M.I., Turney, C.S.M., Hausladen, P.A., Santos, G.M. & Tada, M.L. di. 2001. Radiocarbon dating of the human occupation of Australia prior to 40 ka [SC]BP[D] – successes and pitfalls. Radiocarbon 43, 1139–4510.1017/S0033822200041795Google Scholar
Flenley, J.R. 1998. Tropical forests under the climates of the last 30,000 years. Climatic Change 39, 177–97Google Scholar
Gabunia, L, Vekua, A., Lordkipanidze, D., Swisher, C.C., Ferring, R. et al. 2000. Earliest Pleistocene hominid cranial remains from Dmanisi, Republic of Georgia: taxonomy, geological setting, and age. Science 288, 1019–2510.1126/science.288.5468.1019Google Scholar
Gamble, C. 1993. Timewalkers: the prehistory of global colonisation. London: PenguinGoogle Scholar
Gillespie, R. 2002. Dating the first Australians. Radiocarbon 44, 455–7210.1017/S0033822200031830Google Scholar
Glover, I.C. 1978. Report on a visit to archaeological sites near Medan, Sumatra Utara, July 1975. Bulletin of the Indo-Pacific Prehistory Association 1, 5660Google Scholar
Glover, I.C. 1981. Leang Burung 2: an Upper Palaeolithic rockshelter in South Sulawesi, Indonesia. Modern Quaternary Research in Southeast Asia 6, 138Google Scholar
Gosden, C. 1992. Production systems and the colonization of the Western Pacific. World Archaeology 24, 556910.1080/00438243.1992.9980193Google Scholar
Gosden, C. 1995. Arboriculture and agriculture in coastal Papua New Guinea. Antiquity 69, 807–1710.1017/S0003598X00082351Google Scholar
Groube, L. 1989. The taming of the rainforest: a model for Late Pleistocene forest exploitation in New Guinea. In Harris, D. & Hillman, G.C. (eds), Foraging and Farming: the evolution of plant exploitation, 292317. London: Unwin HymanGoogle Scholar
Groves, C. 1996. Hovering on the brink: nearly but not quite getting to Australia. In Rousham, E. & Freedman, L. (eds), Perspectives in Human Biology 2, 83–7. Nedlands: University of Western Australia, Centre for Human Biology 10.1142/9789812819727_0006Google Scholar
Groves, C. 2004. Homo floresiensis: some initial reactions to the publication and discovery and replies from Brown and Morwood. Before Farming (Online Version) 2004/4, article 1:2Google Scholar
Hanebuth, T.J.J., Stattegger, K. & Grootes, P.M. 2000. Rapid flooding of the Sunda Shelf: A late-glacial sea-level record. Science 288, 1033–5Google Scholar
Harris, D.R. 1977. Subsistence strategies across the Torres Straits. In Allen, J., Golson, J. & Jones, R. (eds), Sunda and Sahul, 421–63. London: AcademicGoogle Scholar
Harrisson, T. 1958. The caves of Niah: a history of prehistory. Sarawak Museum Journal 8, 549–95Google Scholar
Harrisson, T. 1959. New archaeological and ethnological results from Niah Cave, Sarawak. Man 59, 1810.2307/2796008Google Scholar
Hayden, B. 1979. Palaeolithic Reflections. Lithic Technology and Ethnographic Excavations Among Australian Aboringines. Canberra: Australian Institute of Aboriginal StudiesGoogle Scholar
Heaney, L.R. 1991. A synopsis of climatic and vegetational change in southeast Asia. Climatic Change 19, 5361Google Scholar
Henneberg, M. & Thorne, A. 2004. Homo floresiensis: some initial reactions to the publication and discovery and replies from Brown and Morwood. Before Farming (Online Version) 2004/4, article 1:2–4Google Scholar
Hou, Y., Potts, R., Yuan, B., Guo, Z., Deino, A., Wei, W., Clark, J., Zie, G. & Huang, W. 2000. Mid-Pleistocene Acheulean-like stone technology from the Bose Basin, South China. Science 287, 1622–6Google Scholar
Huffman, O.F., Shipman, P., Hertler, C., De Vos, J. & Aziz, F. 2005. Historical evidence of the 1936 Mojokerto skull discovery, East Java. Journal of Human Evolution 48, 321–63Google Scholar
Huffman, O.F., Zaim, Y., Kappelman, J., Ruez Jnr, D.R., De Vos, J., Rizal, Y., Aziz, F. & Hertler, C. 2006. Relocation of the 1936 Mojokerto skull discovery site near Perning, East Java. Journal of Human Evolution 50, 431–51Google Scholar
Hyodo, M. 2001. The Sangiran geomagnetic excursion and its chronological contribution to the Quaternary geology of Java. In Simanjutak, T. et al. (eds), Sangiran: man, culture, and environment in Pleistocene times, 320–36. Jakarta: Yayosan Obor IndonesiaGoogle Scholar
Hyodo, M., Nakaya, H., Urabe, A., Saegusa, H., Shunrong, X., Jiyun, Y. & Xuepin, J. 2002. Paleomagnetic dates of hominid remains from Yuanmou, China and other Asian sites. Journal of Human Evolution 43, 2741Google Scholar
Kaifu, Y., Baba, H., Aziz, F., Indriati, E., Schrenk, F. & Jacob, T. 2005. Taxonomic affinities and evolutionary history of the early Pleistocene hominids of Java: Dentognathic evidence. American Journal of Physical Anthropology 128, 709–26Google Scholar
Kaifu, Y., Y., ,Arif, J., Yokoyama, K., Baba, H., Suparka, E. & Haji, G. 2007. A new Homo erectus molar from Sangiran. Journal of Human Evolution 52, 222–6Google Scholar
Ke, Y., Su, B., Song, X., Lu, D., Chen, L., Li, H., Qi, C., Marzuki, S., Deka, R., Underhill, P., Xiao, C., Shriver, M., Lell, J., Wallace, D., Spencer Wells, R., Seielstad, M., Oefner, P., Zhu, J., Huang, W., Chakraborty, R., Chen, Z. & Jin, L. 2001. African origin of modern humans in East Asia: a tale of 12,000 Y chromosomes. Science 292, 1151–310.1126/science.1060011Google Scholar
Keates, S.G. 2002. The Movius Line: fact or fiction? Bulletin of the Indo-Pacific Prehistory Association 22, 1724Google Scholar
Keates, S.G. & Bartstra, G.J. 199. Island migration of early modern Homo sapiens in Southeast Asia: The artefacts from the Walanae Depression, Sulawesi, Indonesia. Palaeohistoria 33/34, 1930Google Scholar
Kershaw, A.P., Penny, D., Kaars, S. van der, Anshari, G. & Thanotherampillai, A. 2001. Vegetation and climate in lowland southeast Asia at the last glacial maximum. In Metcalfe, I. et al. (eds), Faunal and Floral Migrations and Evolution in SE Asia-Australasia, 227–36. Rotterdam: BalkemaGoogle Scholar
Kidder, J.H. & Durband, A.C. 2004. A re-evaluation of the metric diversity within Homo erectus. Journal of Human Evolution 46, 29731310.1016/j.jhevol.2003.12.003Google Scholar
Kramer, A. 1993. Human taxonomic diversity in the Pleistocene: does Homo erectus represent multiple hominid species? American Journal of Physical Anthropology 91, 161–7110.1002/ajpa.1330910203Google Scholar
Lahr, M.M. & Foley, R. 2004. Human evolution writ small. Nature 431, 1043–4Google Scholar
Langbroek, M. & Roebroeks, W. 2000. Extraterrestrial evidence on the age of the hominids from Java. Journal of Human Evolution 38, 59560010.1006/jhev.1999.0394Google Scholar
Larick, R., Ciochon, R.L., Zaim, Y., Sudijono, Suminto,Rizal, Y., Aziz, F., Reagan, M. & Meizler, M. 2001. Early Pleistocene 40Ar/39Ar ages for Bapeng formation hominins, Central Java, Indonesia. Proceedings of the National Academy of Science 98, 4866–71Google Scholar
Latinis, D.K. 2000. The development of subsistence system models for Island Southeast Asia and Near Oceania: the nature and role of arboriculture and arboreal-based economies. World Archaeology 32, 4167Google Scholar
Leinder, J.J.M., Aziz, F., Sondaar, P. & De Vos, J. 1985. The age of the hominid-bearing deposits of Java: state of the art. Geologie en Mijnbouw 64, 167–73Google Scholar
Lieberman, D.E. & Shea, J. 1994. Behavioral differences between archaic and modern humans in the Levantine Mousterian. American Anthropologist 96(2), 300–32Google Scholar
Long, V.T., De Vos, J. & Ciochon, R.L. 1996. The fossil mammalian fauna of the Lang Trang caves, Vietnam, compared with Southeast Asian fossil and recent mammal faunas: the geographical implications. Bulletin of the Indo-pacific Prehistory Association 15, 119–28Google Scholar
Lourandos, H. 1997. Continent of Hunter-Gatherers: new perspectives in Australian prehistory. Cambridge: University PressGoogle Scholar
Loy, T.H., Spriggs, M. & Wickler, S. 1992. Direct evidence for human use of plants 28,000 years ago: starch residues on stone artifacts from the northern Solomon Islands. Antiquity 66, 89891210.1017/S0003598X00044811Google Scholar
Martin, R.D. 2006. Technical Comment in Science online,.Google Scholar
Maw, B. 1993. The first discovery of an early man's fossilized maxillar bone fragment in Myanmar paleoanthropology. The East Asian Tertiary/Quaternary Newsletter 16, 72Google Scholar
Maxwell, A.L. & Liu, K-B. 2002. Late Quaternary pollen and associated records from the monsoonal areas of continental South and SE Asia. In Kershaw, A.P et al. (eds), Bridging Wallace's line, Advances in Geology 34, 189228. Reiskirchen: CatenaGoogle Scholar
Mellars, P. 2006. A new radiocarbon resolution and the dispersal of modern humans in Eurasia. Nature 439, 931–510.1038/nature04521Google Scholar
Mercader, J. 2002a. Introduction. The Paleolithic settlement of rain forests. In Mercader, (ed.) 2002b, 131Google Scholar
Mercader, J. (ed.). 2002b. Under The Canopy. The Archaeology of Tropical Rainforests. New Brunswick, New Jersey: Rutgers University PressGoogle Scholar
Mijares, A.S.B. 2001. An expedient lithic technology in northern Luzon (Philippines). Lithic Technology 26, 138–5210.1080/01977261.2001.11720983Google Scholar
Molengraaff, G.A.F. 1921. Modern deep-sea research in the east Indian archipelago. Geographical Journal 57, 95121Google Scholar
Moore, M.W. & Brumm, A. 2007. Stone artefacts and hominins in island Southeast Asia: new insights from Flores, eastern Indonesia. Journal of Human Evolution 52, 8510210.1016/j.jhevol.2006.08.002Google Scholar
Morwood, M., O'Sullivan, P.B., Aziz, F. & Raza, A. 1998. Fission-track ages of stone tools and fossils on the east Indonesian island of Flores. Nature 392, 173–6Google Scholar
Morwood, M., Aziz, F., O'Sullivan, P., Nasruddin, ,Hobbs, D.R. & Raza, A. 1999. Archaeological and palaeontological research in central Flores, east Indonesia: results of fieldwork, 1997–98. Antiquity 73, 273–8610.1017/S0003598X00088244Google Scholar
Morwood, M.J., Soejono, R.P., Roberts, R.G., Sutikna, T., Turney, C.S.M., Westaway, K.E., Rink, W.J., Zhao, J., Bergh, G.D. van den, Awe Due, R., Hobbs, D.R., Moore, M.W., Bird, M.I. & Fifield, L.K. 2004. Archaeology and age of a new hominin from Flores in eastern Indonesia. Nature 431, 1087–91Google Scholar
Morwood, M.J., Brown, P., Jatmiko, ,Sutikna, T., Wahyu Saptono, E., Westaway, K.E., Awe Due, R., Roberts, R.G., Maeda, T., Wasisto, S. & Djubianto, T. 2005. Further evidence for the small-bodied hominins from the Late Pleistocene of Flores, Indonesia. Nature 437, 1012–1710.1038/nature04022Google Scholar
Cuong, Nguyen Lang. 1992. A reconsideration of hominid fossils in Vietnam. In Akazawa, et al. (eds) 1992, 321–35Google Scholar
O'Connor, S. & Veth, P. 2000. The world's first mariners: Savannah dwellers in an island continent. Modern Quaternary Research in Southeast Asia 16, 99137Google Scholar
O'Connor, S. & Veth, P. 2005. Early Holocene shell fish hooks from Lene Hara Cave, East Timor establish complex fishing technology was in use in Island Southeast Asia five thousand years before Austronesian settlement. Antiquity 79, 249–5610.1017/S0003598X0011405XGoogle Scholar
O'Connor, S., Spriggs, M. & Veth, P. 2002. Excavation at Lene Hara Cave establishes occuption in East Timor at least 30,000–35,000 years ago. Antiquity 76, 4550Google Scholar
O'Connor, S., Aplin, K., Spriggs, M., Veth, P. & Ayliffe, L. 2002. From savannah to rain forest: changing environments and human occupation at Liang Lemdubu, the Aru Islands, Makulu, Indonesia. In Kershaw, P. & David, B. (eds), Bridging Wallace's Line: environmental and human history and dynamics of the Southeast Asian-Australian region, 279306. Reiskirchen: Catena. Advances in GeoecologyGoogle Scholar
Pannell, S. & O'Connor, S. 2005. Towards a cultural Topography of Cave Use in East Timor: A Preliminary Study, In Barker, G, Gilbertson, D. & Reynolds, T. (eds.) The Human Use of Caves in Southeast Asia. Asian Perspectives 44 (1): 193206Google Scholar
Pasveer, J.M. 2004. 26, 000 Years of Rainforest Exploitation on the Bird's Head of Papua, Indonesia. Modern Quaternary Research in Southeast Asia 17Google Scholar
Pavlides, C & Gosden, C. 1994. 35,000–year–old sites in the rainforests of West New Britain, Papua New Guinea. Antiquity 68, 604–10Google Scholar
Pawlik, A.F. 2004. The Palaeolithic Site of Arubo 1 in Central Luzon, Philippines. Bulletin of the Indo-Pacific Prehistory Association 24 (Indo-Pacific Prehistory: The Taipei Papers (Volume 2), 312Google Scholar
Pawlik, A.F. and Ronquillo, W.P. 2003. The Palaeolithic in the Philippines. Lithic Technology 28. 7993Google Scholar
Paz, V. 2005. Rock Shelters, Caves, and Archaeobotany in Island Southeast Asia. In Barker, G., Gilbertson, D., Reynolds, T. (eds.) The Human Use of Caves in Southeast Asia. Asian Perspectives 44 (1), 107118Google Scholar
Puech., P.F. 1983. Tooth wear, diet, and the artefacts of Java man. Current Anthropology 24. 381–2Google Scholar
Pookajorn, S. 1987. The Technological and Functional Morphology Analyses of the Lithic Tools from the Hoabinhian Excavation at the Ban Kao Area; Kanchenaburi Province, Thailand. Bankok: Silpakorn UniversityGoogle Scholar
Pope, G. G. 1988. Recent Advances in Far Eastern paleoanthropology. Annual Review of Anthropology 17: 4377.Google Scholar
Pope, G. G. 1989. Bamboo and Human Evolution. Natural History 10, 4956Google Scholar
Pope, G. G. 1992. Replacement versus Regionally Continuous Models: The Paleobehavioural and Fossil Evidence from East Asia. In Akazawa, et al. (eds) 1992, 314Google Scholar
Pope, G. G., Frayer, D.W., Liangcharoen, M., Kulasing, P., & Nakabanlang, N. 1981. Palaeothropological investigations of the Thai-American expedition in northern Thailand (1978–1980). Asian Perspectives 21, 147–63Google Scholar
Pope, G. G., Barr, S., Macdonald, A., & Nakabaanlang, S. 1986. Earliest radiometrically dated artefacts from Southeast Asia. Current Anthropology 27. 275–9Google Scholar
Rabett, R. 2005. The early exploitation of Southeast Asian mangroves: bone technology from caves and open sites. In Barker, et al. (eds) 2005, 154–79Google Scholar
Raghavan, P., Groves, C. & Pathmanathan, G. 2003. Homo erectus – was our stone age mechanical engineer an indigenous species of Asia? Yes! He was a product of ‘Evolution in isolation’. Journal of the Anatomical Society of India 52, 1619Google Scholar
Reynolds, T.E.G. 1993. Problems in the Stone Age of Southeast Asia. Proceedings of the Prehistoric Society 59, 115Google Scholar
Reynolds, T.E.G. forthcoming. The stone artefacts. In Barker, et al. (eds), forthcoming, Volume 2Google Scholar
Rightmire, G.P. 1990. The Evolution of Homo erectus. Comparative Anatomical Studies of an Extinct Human Species. Cambridge: University PressGoogle Scholar
Rolland, N. 2002. The initial hominid colonisation of Asia: a survey of anthropic evidence from biogeographic and ecological perspectives. Bulletin of the Indo-Pacific Prehistory Association 22/Melaka Paper 6, 315Google Scholar
Santa Luca, A. 1980. The Ngandong Fossil hominids. A Comparative Study of a Far Eastern Homo Erectus Group. New Haven: Yale University, Department of AnthropologyGoogle Scholar
Schepartz, L.A., Miller-Antonio, S. & Bakken, D.A. 2000. Upland resources and the early Palaeolithic occupation of Southern China, Vietnam, Laos, Thailand and Burma. World Archaeology 32, 11310.1080/004382400409862Google Scholar
Semah, F., Semah, A-M., Djubiantono, T. & Simanjutak, T. 1992. Did they also make stone tools? Journal of Human Evolution 23, 439–6610.1016/0047-2484(92)90092-NGoogle Scholar
Semah, F., Saleki, H. & Falgueres, C. 2000. Did early man reach Java during the late Pliocene? Journal of Archaeological Science 27, 763–9Google Scholar
Semah, F., Semah, A-M. & Simanjuntak, T. 2002. More than a million years of human occupation in insular Southeast Asia. The early archaeology of eastern and central Java. In Mercader, (ed.) 2002, 161–90Google Scholar
Shipman, P. 2001. The Man Who Found the Missing link: the extraordinary Life of Eugene Dubois. London: Weidenfeld & NicholsonGoogle Scholar
Simanjuntak, T. 2001. New insights on the tools of Pithecanthropus. In Simanjuntak, T., Prasetyo, B. & Handini, R. (eds), Sangiran: man, culture and environment in Pleistocene times, 154–70. Jakarta: Yayosan Obor IndonesiaGoogle Scholar
Simanjuntak, T. & Nurani Asikin, I. 2004. Early Holocene human settlement in eastern Java. Bulletin of the Indo-Pacific Prehistory Association 24/Taipei Paper 2, 1319Google Scholar
Simanjuntak, T. & Semah, F. 1996. A new insight into the Sangiran flake industry. Bulletin of the Indo-Pacific Prehistory Association 14, 22–610.7152/bippa.v14i0.11584Google Scholar
Sondaar, P., Bergh, G.D. van den, Mubroto, F., Aziz, J., De Vos, J. & Batu, U.L. 1994. Middle Pleistocene faunal turnover and colonisation of Flores (Indonesia) by Homo erectus. Comptes Rendues de l'Academie des Sciences de Paris 319(SII), 1255–62Google Scholar
Sorenson, P. 1976. Preliminary note on the relative and absolute chronology of two early Palaeolithic sites from north Thailand. In Ghosh, A.K. (ed.), Le Paleolithique inferieur et moyen en Inde, en Asie centrale, en Chine et dans le Sud-Est Asiatique, 237–51. Nice: IXe Congres, Union International de Sciences Prehistorique et Protohistorique, Colloque VIIGoogle Scholar
Spriggs, M. 2000a. Out of Asia: the spread of Southeast Asian Pleistocene and Neolithic maritime cultures in Island Southeast Asia and the western Pacific. Modern Quaternary Research in Southeast Asia 16, 5176Google Scholar
Spriggs, M. 2000b. Can hunter-gatherers live in tropical rainforests? The Pleistocene Melanesian evidence. In Schweitzer, P.P., Biesele, M. & Hitchcock, R.K. (eds), Hunters and Gatherers in the Modern World: conflict resistance and self-determination, 287304. New York: BerghahnGoogle Scholar
Spriggs, M., O'Connor, S. & Veth, P. 200. Vestiges of early pre-agricultural economy in the landscape of East Timor. In Karlstrom, A. & Kallen, A. (eds), Fishbones and Glittering Emblems: Southeast Asian archaeology 2002, 4958. Stockholm: Museum of Far Eastern AntiquitiesGoogle Scholar
Stone, R. 2006. Java Man's first tools. Review of the Indo-Pacific Prehistory Association Congress, 20–26 March 2006, Manila. Science 312, 361Google Scholar
Storm, P. 2001. The evolution of humans in Australasia from an environmental perspective. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 363–83Google Scholar
Storm, P., Aziz, F., De Vos, J., Kosaasih, D., Baskoro, S., Ngaliman, , & Hoek Ostende, L.W. van den. 2005. Late Pleistocene Homo sapiens in a tropical rainforest fauna in East Java. Journal of Human Evolution 49, 36545Google Scholar
Sun, X., Li, X., Luo, Y. & Chen, X. 2000. The vegetation and climate at the last glaciation on the emerged continental shelf of the South China Sea. Palaeogeography, Palaeoclimatology, Palaeoecology 160, 301–16Google Scholar
Swisher, C.C., Curtis, G.H., Jacob, T., Getty, A.G., Suprijo, A., & Widiasmoro, , 1994. Age of the earliest known hominids in Java, Indonesia. Science 263, 1119–2110.1126/science.8108729Google Scholar
Swisher, C.C., Rink, W.J., Anton, S.C., Schwarcz, H.P., Curtis, G.H., Suprijo, A. & Widiasmoro, . 1996. Latest Homo erectus of Java: potential contemporaneity with Homo sapiens in Southeast Asia. Science 274, 1870–4Google Scholar
Tan, Ha Van. 1995. The Ngoum Rockshelter and the Paleolithic flake industries in mainland Southeast Asia. In Chin, S.T. (ed.), Papers on Archaeology in Southeast Asia, 171–9. Hong Kong: University Museum & Art Gallery, University of Hong KongGoogle Scholar
Tan, Ha Van. 1997. The Hoabinhian and Before. Bulletin of the Indo-Pacific Prehistory Association 16, 3541Google Scholar
Taylor, D., Yen, O.H., Sanderson, P.G. & Dodson, J. 2001. Late Quaternary peat formation and vegetation dynamics in a lowland tropical swamp; Nee Soon, Singapore. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 269–87Google Scholar
Theunissen, B., De Vos, J., Sondaar, P. & Aziz, F. 1990. The establishment of a chronological framework for the hominid-bearing deposits of Java: a historical survey. Geological Society of America Special Paper 242, 395410.1130/SPE242-p39Google Scholar
Tougard, C. & Montuire, S. 2006. Pleistocene paleoenvironmental reconstructions and mammalian evolution in South-East Asia: focus on fossil faunas from Thailand. Quaternary Science Reviews 25, 126–4110.1016/j.quascirev.2005.04.010Google Scholar
Townsend, P.K. 1990. On the possibility/impossibility of tropical forest hunting and gathering. American Anthropologist 92, 745–7Google Scholar
Trinkaus, E. 2005. Early modern humans. Annual Review of Anthropology 34, 207–30Google Scholar
Van Andel, T.H. & Davies, W. (eds). 2003. Neanderthals and Modern Humans in the European Landscape During the Last Glaciation. Cambridge: McDonald Insititute MonographGoogle Scholar
Vekua, A., Lordkipanidze, D., Rightmire, G.P., Agusti, J., Ferring, R., Maisuradze, G., Moustkhelishivili, A., Nioradze, M., Ponce de Leon, M.S., Tappen, M., Tvalchrelidze, M. & Zollikofer, L.P.E. 2002. A new skull of early Homo from Dmanisi, Georgia. Science 297, 85–910.1126/science.1072953Google Scholar
Verhoeven, Th. 1968. Pleistozane Funde auf Flores, Timor and Sumba. In Anthropica. Gedenkschrift zum 100. Geburtstag von P.W. Schmidt, 393403. Studia Instituti Anthropos 21Google Scholar
Veth, P., Spriggs, M. & O'Connor, S. 2005. The continuity of cave use in the Tropics: examples from East Timor and the Aru Islands, Maluku. In Barker, et al. (eds), 180–92Google Scholar
Villmoare, B. 2005. Metric and non-metric randomisation methods, geographic variation, and the single-species hypothesis for Asian and African Homo erectus. Journal of Human Evolution 49, 680701Google Scholar
Voris, H.K. 2000. Maps of Pleistocene sea levels in Southeast Asia: shorelines, river systems and time durations. Journal of Biogeography 27, 1153–67Google Scholar
Widianto, H. 2005. The oldest Homo erectus stone tools in Java: from the Lower Pucangan Formation in Sangiran, Abstract. (accessed 30/11/2005)Google Scholar
Wolpoff, M.H., Mannheim, B., Mann, A., Hawks, J., Caspari, K.R., Frayer, D.W., Gill, G.W. & Clark., G. 2004. Why not the Neandertals? World Archaeology 36, 527–46Google Scholar
Wu, X. 1992. The origin and dispersal of anatomically modern humans in East and Southeast Asia. In Akazawa, et al. (eds) 1992, 373–8Google Scholar
Yamei, H., Potts, R., Yuan, B., Guo, Z., Deino, A., Wang, W., Clark, J., Xie, G. & Huang, W. 2000. Mid-Pleistocene Acheulean-like stone technology of the Bose Basin, South China. Science 287, 122Google Scholar
Majid, Zuraina. 1982. The West Mouth, Niah, in the prehistory of Southeast Asia. Sarawak Museum Journal 31 (special monograph)Google Scholar
Majid, Zuraina. 1998. Radiocarbon dates and culture sequence in the Lenggong Valley and beyond. Malaysian Museums Journal 34, 241–9Google Scholar