Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Earliest hominin occupation of Sulawesi, Indonesia

Abstract

Sulawesi is the largest and oldest island within Wallacea, a vast zone of oceanic islands separating continental Asia from the Pleistocene landmass of Australia and Papua (Sahul). By one million years ago an unknown hominin lineage had colonized Flores immediately to the south1, and by about 50 thousand years ago, modern humans (Homo sapiens) had crossed to Sahul2,3. On the basis of position, oceanic currents and biogeographical context, Sulawesi probably played a pivotal part in these dispersals4. Uranium-series dating of speleothem deposits associated with rock art in the limestone karst region of Maros in southwest Sulawesi has revealed that humans were living on the island at least 40 thousand years ago (ref. 5). Here we report new excavations at Talepu in the Walanae Basin northeast of Maros, where in situ stone artefacts associated with fossil remains of megafauna (Bubalus sp., Stegodon and Celebochoerus) have been recovered from stratified deposits that accumulated from before 200 thousand years ago until about 100 thousand years ago. Our findings suggest that Sulawesi, like Flores, was host to a long-established population of archaic hominins, the ancestral origins and taxonomic status of which remain elusive.

This is a preview of subscription content, access via your institution

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Sunda, Sahul and Wallacea.
Figure 2: Talepu excavations T2 and T4.
Figure 3: Finds from Talepu excavations 2 (T2) and 4/4B (T4/T4B).

Similar content being viewed by others

References

  1. Brumm, A. et al. Hominins on Flores, Indonesia, by one million years ago. Nature 464, 748–752 (2010)

    ADS  CAS  PubMed  Google Scholar 

  2. Clarkson, C. et al. The archaeology, chronology and stratigraphy of Madjedbebe (Malakunanja II): a site in northern Australia with early occupation. J. Hum. Evol. 83, 46–64 (2015)

    PubMed  Google Scholar 

  3. O’Connor, S. in Emergence and Diversity of Modern Human Behaviour in Paleolithic Asia (eds Yousuke Kaifu et al. .) 214–224 (Texas A&M Univ. Press, 2015)

  4. Morwood, M. & Van Oosterzee, P. The Discovery of the Hobbit: The Scientific Breakthrough that Changed the Face of Human History (Random House Australia, 2007)

  5. Aubert, M. et al. Pleistocene cave art from Sulawesi, Indonesia. Nature 514, 223–227 (2014)

    ADS  CAS  PubMed  Google Scholar 

  6. Van Heekeren, H. R. Early man and fossil vertebrates on the island of Celebes. Nature 163, 492 (1949)

    ADS  CAS  PubMed  Google Scholar 

  7. Bartstra, G.-J., Keates, S. G., Basoeki & Kallupa, B. On the dispersion of Homo sapiens in eastern Indonesia: the Palaeolithic of south Sulawesi. Curr. Anthropol. 32, 317–321 (1991)

    Google Scholar 

  8. van Heekeren, H. R. The Stone Age of Indonesia (Martinus Nijhoff, 1972)

  9. Hooijer, D. A. Pleistocene vertebrates from Celebes: Celebochoerus heekereni nov. gen. nov. spec. Proc. K. Ned. Akad. Wet. 51, 1024–1032 (1948)

    Google Scholar 

  10. Hooijer, D. A. Quaternary mammals west and east of Wallace’s Line. Neth. J. Zool. 25, 46–56 (1974)

    Google Scholar 

  11. van den Bergh, G. D. The Late Neogene elephantoid-bearing faunas of Indonesia and their palaeozoogeographic implications: a study of the terrestrial faunal succession of Sulawesi, Flores and Java, including evidence for early hominid dispersal east of Wallace’s Line. Scripta Geologica 117, 1–419 (1999)

    Google Scholar 

  12. Grainge, A. M. & Davies, K. G. Reef exploration in the East Sengkang Basin, Sulawesi. In Proc. 12th Annual Convention of the Indonesian Petroleum Association, V1, 207–227 (1983)

    Google Scholar 

  13. Sukamto, R. Geological Map of Indonesia, Sheet VIII: Ujung Pandang . Scale 1:1,000,000. Geological Research and Development Centre, Bandung (1975)

  14. Grün, R., Eggins, S., Kinsley, L., Moseley, H. & Sambridge, M. Laser ablation U-series analysis of fossil bones and teeth. Palaeogeogr. Palaeoclimatol. Palaeoecol. 416, 150–167 (2014)

    Google Scholar 

  15. Sambridge, M., Grün, R. & Eggins, S. U-series dating of bone in an open system: the diffusion-adsorption-decay model. Quat. Geochronol. 9, 42–53 (2012)

    Google Scholar 

  16. Li, B. & Li, S.-H. Luminescence dating of K-feldspar from sediments: a protocol without anomalous fading correction. Quat. Geochronol. 6, 468–479 (2011)

    Google Scholar 

  17. Li, B., Jacobs, Z., Roberts, R. G. & Li, S.-H. Review and assessment of the potential of post-IR IRSL dating methods to circumvent the problem of anomalous fading in feldspar luminescence. Geochronometria 41, 178–201 (2014)

    Google Scholar 

  18. O’Connor, S., Ono, R. & Clarkson, C. Pelagic fishing at 42,000 years before the present and the maritime skills of modern humans. Science 334, 1117–1121 (2011)

    ADS  PubMed  Google Scholar 

  19. Summerhayes, G. R. et al. Human adaptation and plant use in highland New Guinea 49,000 to 44,000 years ago. Science 330, 78–81 (2010)

    ADS  CAS  PubMed  Google Scholar 

  20. Mellars, P. Why did modern human populations disperse from Africa ca. 60,000 years ago? A new model. Proc. Natl Acad. Sci. USA 103, 9381–9386 (2006)

    ADS  CAS  PubMed  Google Scholar 

  21. Storm, P. et al. U-series and radiocarbon analyses of human and faunal remains from Wajak, Indonesia. J. Hum. Evol. 64, 356–365 (2013)

    PubMed  Google Scholar 

  22. Barker, G. et al. The ‘human revolution’ in lowland tropical Southeast Asia: the antiquity and behavior of anatomically modern humans at Niah Cave (Sarawak, Borneo). J. Hum. Evol. 52, 243–261 (2007)

    PubMed  Google Scholar 

  23. Grün, R. et al. U-series and ESR analyses of bones and teeth relating to the human burials from Skhul. J. Hum. Evol. 49, 316–334 (2005)

    PubMed  Google Scholar 

  24. Westaway, K. E. et al. Age and biostratigraphic significance of the Punung Rainforest Fauna, East Java, Indonesia, and implications for Pongo and Homo. J. Hum. Evol. 53, 709–717 (2007)

    CAS  PubMed  Google Scholar 

  25. Westaway, K. E. et al. Establishing the time of initial human occupation of Liang Bua, western Flores, Indonesia. Quat. Geochronol. 2, 337–343 (2007)

    Google Scholar 

  26. Zaim, Y. et al. New 1.5 million-year-old Homo erectus maxilla from Sangiran (Central Java, Indonesia). J. Hum. Evol. 61, 363–376 (2011)

    MathSciNet  PubMed  Google Scholar 

  27. Indriati, E. et al. The age of the 20 meter Solo River terrace, Java, Indonesia and the survival of Homo erectus in Asia. PLoS One 6, e21562 (2011)

    ADS  CAS  PubMed  PubMed Central  Google Scholar 

  28. Cooper, A. & Stringer, C. B. Paleontology. Did the Denisovans cross Wallace’s Line? Science 342, 321–323 (2013)

    ADS  CAS  PubMed  Google Scholar 

  29. Sprintall, J. et al. The Indonesian seas and their role in the coupled ocean–climate system. Nature Geosci. 7, 487–492 (2014)

    ADS  CAS  Google Scholar 

  30. Mijares, A. S. et al. New evidence for a 67,000-year-old human presence at Callao Cave, Luzon, Philippines. J. Hum. Evol. 59, 123–132 (2010)

    PubMed  Google Scholar 

  31. Moore, M. W., Sutikna, T., Jatmiko, Morwood, M. J. & Brumm, A. Continuities in stone flaking technology at Liang Bua, Flores, Indonesia. J. Hum. Evol. 57, 503–526 (2009)

    CAS  PubMed  Google Scholar 

  32. Eggins, S. M. et al. In situ U-series dating by laser-ablation multi-collector ICPMS: new prospects for Quaternary geochronology. Quat. Sci. Rev. 24, 2523–2538 (2005)

    ADS  Google Scholar 

  33. Grün, R. et al. ESR and U-series analyses of teeth from the palaeoanthropological site of Hexian, Anhui Province, China. J. Hum. Evol. 34, 555–564 (1998)

    PubMed  Google Scholar 

  34. Huntley, D. J., Godfrey-Smith, D. I. & Thewalt, M. L. Optical dating of sediments. Nature 313, 105–107 (1985)

    ADS  Google Scholar 

  35. Aitken, M. J. An Introduction to Optical Dating: The Dating of Quaternary sediments by the Use of Photon-Stimulated Luminescence (Oxford Univ. Press, 1998)

  36. Lian, O. B. & Roberts, R. G. Dating the Quaternary: progress in luminescence dating of sediments. Quat. Sci. Rev. 25, 2449–2468 (2006)

    ADS  Google Scholar 

  37. Roberts, R. G. et al. Optical dating in archaeology: thirty years in retrospect and grand challenges for the future. J. Archaeol. Sci. 56, 41–60 (2015)

    Google Scholar 

  38. Wintle, A. G. Anomalous fading of thermo-luminescence in mineral samples. Nature 245, 143–144 (1973)

    ADS  CAS  Google Scholar 

  39. Huntley, D. J. & Lamothe, M. Ubiquity of anomalous fading in K-feldspars and the measurement and correction for it in optical dating. Can. J. Earth Sci. 38, 1093–1106 (2001)

    ADS  CAS  Google Scholar 

  40. Thomsen, K. J., Murray, A., Jain, M. & Bøtter-Jensen, L. Laboratory fading rates of various luminescence signals from feldspar-rich sediment extracts. Radiat. Meas. 43, 1474–1486 (2008)

    CAS  Google Scholar 

  41. Buylaert, J.-P., Murray, A., Thomsen, K. J. & Jain, M. Testing the potential of an elevated temperature IRSL signal from K-feldspar. Radiat. Meas. 44, 560–565 (2009)

    CAS  Google Scholar 

  42. Thiel, C. et al. Luminescence dating of the Stratzing loess profile (Austria)—testing the potential of an elevated temperature post-IR IRSL protocol. Quat. Int. 234, 23–31 (2011)

    Google Scholar 

  43. Li, B. & Li, S.-H. Luminescence dating of Chinese loess beyond 130 ka using the non-fading signal from K-feldspar. Quat. Geochronol. 10, 24–31 (2012)

    Google Scholar 

  44. Hutton, J. & Prescott, J. Field and laboratory measurements of low-level thorium, uranium and potassium. Nucl. Tracks Radiat. Meas. 20, 367–370 (1992)

    CAS  Google Scholar 

  45. Bøtter-Jensen, L. & Mejdahl, V. Assessment of beta dose-rate using a GM multi-counter system. Nucl. Tracks Radiat. Meas. 14, 187–191 (1988)

    Google Scholar 

  46. Jacobs, Z. & Roberts, R. G. An improved single grain OSL chronology for the sedimentary deposits from Diepkloof Rockshelter, Western Cape, South Africa. J. Archaeol. Sci. 63, 175–192 (2015)

    Google Scholar 

  47. Li, B., Roberts, R. G., Jacobs, Z. & Li, S.-H. A single-aliquot luminescence dating procedure for K-feldspar based on the dose-dependent MET-pIRIR signal sensitivity. Quat. Geochronol. 20, 51–64 (2014)

    CAS  Google Scholar 

  48. Galbraith, R. Graphical display of estimates having differing standard errors. Technometrics 30, 271–281 (1988)

    Google Scholar 

  49. Galbraith, R. F., Roberts, R. G., Laslett, G. M., Yoshida, H. & Olley, J. M. Optical dating of single and multiple grains of quartz from Jinmium rock shelter, northern Australia: part I, experimental design and statistical models. Archaeometry 41, 339–364 (1999)

    Google Scholar 

  50. Galbraith, R. & Roberts, R. G. Statistical aspects of equivalent dose and error calculation and display in OSL dating: an overview and some recommendations. Quat. Geochronol. 11, 1–27 (2012)

    Google Scholar 

  51. Li, B., Roberts, R. G., Jacobs, Z., Li, S.-H. & Guo, Y.-J. Construction of a ‘global standardised growth curve’(gSGC) for infrared stimulated luminescence dating of K-feldspar. Quat. Geochronol. 27, 119–130 (2015)

    Google Scholar 

  52. Li, B., Roberts, R. G. & Jacobs, Z. On the dose dependency of the bleachable and non-bleachable components of IRSL from K-feldspar: improved procedures for luminescence dating of Quaternary sediments. Quat. Geochronol. 17, 1–13 (2013)

    Google Scholar 

  53. Auclair, M., Lamothe, M. & Huot, S. Measurement of anomalous fading for feldspar IRSL using SAR. Radiat. Meas. 37, 487–492 (2003)

    CAS  Google Scholar 

  54. Huntley, D. J. & Lian, O. B. Some observations on tunnelling of trapped electrons in feldspar and their implications for optical dating. Quat. Sci. Rev. 25, 2503–2512 (2006)

    ADS  Google Scholar 

  55. Kars, R. H., Wallinga, J. & Cohen, K. M. A new approach towards anomalous fading correction for feldspar IRSL dating—tests on samples in field saturation. Radiat. Meas. 43, 786–790 (2008)

    CAS  Google Scholar 

  56. Li, B. & Li, S.-H. Investigations of the dose-dependent anomalous fading rate of feldspar from sediments. J. Phys. D 41, 225502 (2008)

    ADS  Google Scholar 

  57. Kirschvink, J. L. The least-squares line and plane and the analysis of paleomagnetic data. Geophys. J. R. Astron. Soc. 62, 699–718 (1980)

    ADS  Google Scholar 

  58. Lurcock, P. C. & Wilson, G. S. PuffinPlot: a versatile, user-friendly program for paleomagnetic analysis. Geochem. Geophys. Geosyst. 13, Q06Z45 (2012)

    Google Scholar 

  59. Torsvik, T., Briden, J. & Smethurst, M. Super-IAPD interactive analysis of palaeomagnetic data. http://www.geodynamics.no/resources.html (2000)

  60. Butler, R. F. Paleomagnetism: Magnetic Domains to Geologic Terranes Ch. 6, 103–120 (Blackwell Scientific, 1992)

  61. Rivera, T. A., Storey, M., Schmitz, M. D. & Crowley, J. L. Age intercalibration of 40Ar/39Ar sanidine and chemically distinct U/Pb zircon populations from the Alder Creek Rhyolite Quaternary geochronology standard. Chem. Geol. 345, 87–98 (2013)

    ADS  CAS  Google Scholar 

  62. Storey, M., Roberts, R. G. & Saidin, M. Astronomically calibrated 40Ar/39Ar age for the Toba supereruption and global synchronization of late Quaternary records. Proc. Natl Acad. Sci. USA 109, 18684–18688 (2012)

    ADS  CAS  PubMed  Google Scholar 

  63. Elburg, M. A. & Foden, J. Geochemical response to varying tectonic settings: an example from southern Sulawesi (Indonesia). Geochim. Cosmochim. Acta 63, 1155–1172 (1999)

    ADS  CAS  Google Scholar 

  64. Wilson, M. E. J., Bosence, D. W. J. & Limbong, A. Tertiary syntectonic carbonate platform development in Indonesia. Sedimentology 47, 395–419 (2000)

    ADS  CAS  Google Scholar 

Download references

Acknowledgements

This research was supported by grants from the Australian Research Council (ARC) to M.J.M. (DP0770234/DP1093342) and G.D.v.d.B. (FT100100384), and funds from the Geological Survey Institute of Indonesia. L.B. and R.G.R. are supported by ARC Fellowships FT140100384 and FL130100116, respectively. A.B.’s involvement was supported by ARC fellowship DE130101560. The stone tool analysis was supported by ARC Fellowship DP1093342 to M.W.M., and M.S. was funded by the Villum Foundation. The fieldwork was authorized by the directors of the Geological Survey Institute of Indonesia, A. Djumarma Wirakusumah and Y. Kusumahbrata. We further acknowledge the Indonesian State Ministry of Research and Technology and the National Centre for Archaeology in Jakarta (ARKENAS), for facilitating the research. Field assistants included H. Oktaviana, Dadang, S. Sudjarwadi, Ngaliman, T. Suryana and U. P. Wibowo. During the excavations we were assisted by the landowner, Wahe, and 11 other local labourers. Y. Jafari prepared the sediment samples for optical dating and L. Kinsley assisted with the uranium-series analysis. We thank K. Westaway, D. Granger, B. Pillans and B. Jones for additional field support, and S. van der Mije for allowing access to the vertebrate collection at Naturalis. S. Hayes is thanked for providing feedback on the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

M.J.M. and G.D.v.d.B. conceived the study with F.A., as part of a wider project led by M.J.M., in collaboration with A.B., I.K., S. and E.S. Samples for optical dating were collected and analysed by B.L. and R.G.R. R.G. conducted the uranium-series dating and M.S. analysed samples for 40Ar/39Ar dating. A.B. and M.W.M. identified and analysed the stone artefacts. G.D.v.d.B. and I.K. analysed the fossil specimens. G.D.v.d.B. and R.S. recorded the site stratigraphy. D.Y. collected and analysed samples for palaeomagnetism. S. conducted a regional geological survey supervised by G.D.v.d.B. and M.J.M. G.D.v.d.B. and A.B. wrote the manuscript, with contributions from the other authors.

Corresponding authors

Correspondence to Gerrit D. van den Bergh or Bo Li.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Extended data figures and tables

Extended Data Figure 1 Research location and geology of southwest Sulawesi and Talepu.

a, Digital elevation model of southwest Sulawesi. 1: Talepu; 2: Maros karst area. Area enclosed by rectangle shown in d; map data: copyright USGS/NASA SRTM (2007). b, Geological map of southwest Sulawesi (data from refs 11, 63, 64). The Walanae Depression (WD) is an elongate fault-bounded basin (also known as the West Sengkang Basin) separated from the Bone Mountains to the east by a major fault, the East Walanae fault (EWF), which formed in response to east–west compression and strike-slip movements along the Walanae fault zone. To the west the basin is bordered by the Western Dividing Range, consisting of uplifted Miocene volcanics deposited in a shallow marine environment13. The Walanae Depression basin infill consists of a several-kilometre-thick regressive sequence, named the Walanae Formation. In the southern part of the Walanae Depression, the Walanae Formation is folded and deformed by Pleistocene compression, whereas to the north near Lake Tempe deposition continues to the present day. 1: Talepu; 2: Paroto (alluvial terrace of the Walanae River); 3: Beru8; 4: Tanrung River (palaeontological site: coastal terrace deposits11); 5: Sompe8,9,10; 6: Celeko8,9,10; 7: Maros karst archaeological sites5. c, Schematic stratigraphic scheme for the northern Sengkang Basin at the latitude of the Talepu site (green dotted line in b). The Walanae Formation basin fill represents a regressive sequence that was strongly influenced by tectonic movements along the Walanae fault zone. The youngest unit of the Walanae Formation is the Beru member (deltaic sands, clays and gravels), which contains fossil vertebrate remains of the Walanae Fauna11. The lower part of the Beru member (unit A) is characterized by sedimentary structures indicative of shallow marine/estuarine/fluvial depositional environments. The upper part of the Beru member (unit B) consists of fully terrestrial fluvio-lacustrine deposits, which merge into the modern floodplain along the depocentral axis of the Walanae Depression. The coarser-grained unit B of the Beru member was not deposited in the Sengkang anticline, which started to rise during the Middle Pleistocene, or in the southern portion of the Walanae Depression south of Talepu. East of the Walanae fault Zone, in the East Sengkang Basin, uplift and folding during the Pliocene caused a depositional hiatus. Here Late Miocene deformed marine deposits of the Walanae Formation are unconformably overlaid by a conglomerate up to 5 m thick, the Tanrung Formation, which contains a distinct fossil vertebrate fauna11. During the Middle and Late Pleistocene, uplift of the Western Dividing Range generated the formation of alluvial fans and influxes of coarse-grained boulder conglomerates into the Walanae Depression. d, Geological map of Talepu area with sub-horizontal layering, fault-bounded to the east by steeply west-dipping strata of the Sengkang anticline. 1: Modern alluvium; 2: Late Pleistocene alluvial terrace; 3–7: lithological sub-units of the Walanae Formation: 3: fluvio-lacustrine facies of the upper part of the Beru member; 4: fluvio-estuarine facies of the lower part of the Beru member; 5: shallow marine facies of the Samaoling member; 6: deep marine facies of the Burecing member; 7: coral reef facies of the Tacipi member; 8: strike and dip; 9: sub-horizontal layering; 10: major fault; 11: sites with surface-collected stone artefacts; 12: sites with in situ stone artefacts; 13: fossil vertebrate localities.

Extended Data Figure 2 Talepu site and excavation images.

a, View to the east of the north–south trending ridge. Talepu is located behind the palm trees on the left. b, View towards the East Baulk of Talepu excavation 2 in 2009. In 2010 the 1 m × 2 m excavation was extended to a 2 m × 2 m excavation. c, The 12 m deep Talepu excavation 2 in 2012. d, Talepu excavation 4 in 2010. View towards the North Baulk. e, Talepu excavation 2 in 2012. Photograph shows the North Baulk at 4–5 m depth, with, in the upper part, the base of gravel unit A7 and the three holes left by sampling for optical dating.

Extended Data Figure 3 Composition of gravels and distribution of stone artefacts in T2, and size distribution of recent anoa molars.

a, Gravel compositions based on pebble counts (200 pebbles per level). Overall, the composition of the gravel is dominated by volcanic pebbles, which become more abundant with increasing depth, probably as a result of less intense weathering further down (near the surface, the volcanic clasts are frequently weathered to a crumbly clayey ‘ghost’). Note the increase both in weathering-resistant silicified rock pebbles and in heavily weathered indeterminable clasts towards the top of the sequence. b, Total number of artefacts per 10 cm spit (black triangles), total amount of gravel clasts per spit (red graph) and the maximum clast diameter per spit (blue graph: values represent the mean maximum clast diameter of the ten largest clasts). Note the higher concentration of both gravel and artefacts in the topsoil: pebbles and artefacts are concentrated by winnowing of sand and clay by sheetwash processes. c, Detail of the topsoil as exposed in the west baulk of excavation T2. d, Detail of the basal sequence as exposed in the north baulk of excavation T2. Note the cross-bedded foresets of the pebbly sand of sub-unit D2, with inter-bedded mud laminae, indicative of tidal activity. Diameter of the round sample hole (for optical dating) is ~10 cm. e, Histogram of the transverse diameter measurements (in millimetres) of recent lowland anoa (Bubalus [Anoa] depressicornis) lower molars measured in the collections of the Naturalis Biodiversity Centre, Leiden, the Netherlands (n = 32). The lowland anoa is the largest living anoa, bigger in body size than the mountain anoa, Bubalus [Anoa] quarlesi. The lower molar fragment from unit A of the Talepu-2 excavation (Fig. 3t) has a preserved basal transverse diameter of 14.4 mm and an estimated basal transverse diameter of 15.5 mm, slightly above the size range for extant lowland anoa.

Extended Data Figure 4 Fossil samples used for uranium-series dating and demagnetization results of representative palaeomagnetic samples.

Faunal remains from Talepu excavation 4, used for uranium-series laser ablation dating (ai) and representative NRM intensity plots of progressive demagnetization (j). ai, Close-ups of the surface sections for each fossil with the laser spot profiles. All fossils originate from excavation T4, sub-unit E2. Scale bars next to fossils are 2 cm; white scale bars in close-ups are 1 mm. Numbers between brackets are Australian National University laboratory numbers. a, Specimen TLP10-F8 (ANU-2946), Celebochoerus upper left first incisor; laser ablation transect on cut section of root. b, c, Two sections measured on different transects of the same specimen, TLP10-F1 (ANU-2947 and ANU-2948), a Celebochoerus lower left canine. d, Specimen TLP10-F7 (ANU-2951), rib fragment of Celebochoerus; e, Specimen TLP10-F4 (ANU-2954), Celebochoerus upper left third molar; laser ablation transect on cut section of root; f, Specimen TLP10-F3 (ANU-2956), Celebochoerus upper right third molar (same individual as previous); laser ablation transect on cut section of root. g, Specimen TLP10-F9 (ANU-2955), Celebochoerus upper molar fragment; laser ablation transect on cut section of enamel and dentine. h, Specimen TLP10-F6 (ANU-2949), bone fragment. i, TLP10-F2 (ANU-2942), Celebochoerus upper left fourth premolar; laser ablation transect on cut section of root. j, NRM intensity plot of progressive demagnetization (upper left), equal area projections (middle left) and vector end-point demagnetization orthogonal plots (bottom left) for two Talepu palaeomagnetic samples (T2-510-1, T4-180-4). To the right the demagnetization curve for an additional sample (T2-320-4) is given. The inset shows the zoomed-out trajectory endpoints of sample T2-320-4. Open squares on the equal area projection diagrams indicate an upper hemisphere magnetic direction.

Extended Data Figure 5 Lithological and magnetic properties against depth for the composite stratigraphic column at Talepu.

a, Columns from left to right show lithology, sand/silt/clay ratios, NRM magnetic intensities before and after demagnetization, and magnetic declination and inclination directions. The intensities before and after demagnetization represent averages for each sampled level. Declination and inclination values are the averages of the higher coercivity stable magnetization (ChRM) with their 95% confidence ranges. b, Equal-area projections of NRM and ChRM directions for all sampled levels, and the mean direction (circles with crosses; the mean is of all sampled levels except the two levels with deviating inclinations: n = 22) and present-day magnetic direction in the area (red crosses).

Extended Data Figure 6 Decay curves and dose–response curves of the IRLS signals for K-feldspars from Talepu.

a, Representative IRSL and MET-pIRIR decay curves for a single aliquot of sample TUT-OSL2, stimulated at different temperatures (shown above each curve). b, Dose–response curves for the IRSL (50 °C) and MET-pIRIR (100–250 °C) signals for the same aliquot. The natural signals are shown on the x axis using the same symbols as the regenerative signals. The data points were fitted using a single saturating-exponential function. The best-fit curves are shown as solid lines and the characteristic saturation dose (D0) values are indicated. c, Radial plot showing the D0 values from 38 aliquots for different samples. The grey band shows the mean of the D0 values.

Extended Data Figure 7 Results from residual dose, dose recovery and anomalous fading tests.

a, Residual doses measured for bleached aliquots of the four samples from the upper trench, plotted against stimulation temperature. Each data point represents the mean and standard error for four aliquots. b, Results of the dose recovery test conducted on sample TUT-OSL1. The measured/given dose ratios are shown for the IRSL and MET-pIRIR signals at the different stimulation temperatures. Each data point represents the mean and standard error for four aliquots. The data shown in red squares were obtained using a hot IR bleach at the end of each SAR cycle, as per the conventional MET-pIRIR procedure. The data shown in black circles were obtained with the modified MET-pIRIR procedure (Supplementary Table 4), using a solar simulator bleach instead of a hot IR bleach. The dashed line denotes a ratio of unity, and the solid lines indicate ratios 10% larger and smaller than unity. The data (circles) obtained using the modified MET-pIRIR procedure fall within the latter band. c, Decay of the sensitivity-corrected IRSL and MET-pIRIR signals with their standard errors of six aliquots from TUT-OSL3, plotted against log(t/tc) where t is the delayed period for each measurement and tc is the time for the first measurement (tc = 720, 870, 1,040, 1,240 and 1,480 s for the signals measured at 50, 100, 150, 200 and 250 °C, respectively). The sensitivity-corrected signals were normalized to the first measurements. d, Anomalous fading rates (g values) and their standard errors for the IRSL and MET-p IRIR signals of TUT-OSL3 obtained using the data sets in c, plotted against stimulation temperature. All the g values have been normalized to a delay time of 2 days.

Extended Data Figure 8 Radial plots of single-aliquot De values for the TUT samples.

a, TUT-OSL1. b, TUT-OSL2. c, TUT-OSL3. d, TUT-OSL9. The grey band in each plot shows the weighted mean of the measured De values estimated using the central age model. The De estimate and the overdispersion (OD) value for each De distribution based on the central age model are also shown in each plot.

Extended Data Figure 9 De versus temperature plots for the TUT samples and the dose–response curve for sample TLT-OSL6.

a, Plots of the weighted mean De against stimulation temperature for the TUT samples. The dashed line in each plot shows the plateau range of De values. Each data point represents the mean and standard error for 8 (TUT-OSL1), 10 (TUT-OSL2), 12 (TUT-OSL3) and 13 (TUT-OSL9) aliquots. b, Dose–response curve for the sensitivity-corrected MET-pIRIR 250 °C signal from an aliquot of sample TLT-OSL6. The regenerative-dose data points and their standard errors were fitted using a single saturating-exponential function, and the best-fit curve is shown as a full line. The natural signal of this aliquot (red circle on the y axis) falls in the saturated region of the curve (see dashed line), so only a minimum De can be estimated.

Extended Data Figure 10 40Ar/39Ar fusion ages of single sanidine crystals from sample TAL-10-01 (T4, sub-unit E2, 2.5 m below surface).

Top: the individual ages ± 1σ. Bottom: histogram based on over 100 analyses. Excluding the two youngest ages, the crystal ages span a period from 5 to 11 million years ago, but with the main age population at about 9.4 million years ago. The ages correspond with the Late Miocene collision phase, when potassium-rich volcanics of the Camba Formation were formed along the Western Dividing Range63.

Supplementary information

Supplementary Tables

This file contains Supplementary Tables 1-9. (PDF 353 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

van den Bergh, G., Li, B., Brumm, A. et al. Earliest hominin occupation of Sulawesi, Indonesia. Nature 529, 208–211 (2016). https://doi.org/10.1038/nature16448

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature16448

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing