Supplementary Materials for Early Levallois technology and the Lower to Middle Paleolithic...

65
www.sciencemag.org/content/345/6204/1609/suppl/DC1 Supplementary Materials for Early Levallois technology and the Lower to Middle Paleolithic transition in the Southern Caucasus D. S. Adler,* K. N. Wilkinson, S. Blockley, D. F. Mark, R. Pinhasi, B. A. Schmidt-Magee, S. Nahapetyan, C. Mallol, F. Berna, P. J. Glauberman, Y. Raczynski-Henk, N. Wales, E. Frahm, O. Jöris, A. MacLeod, V. C. Smith, V. L. Cullen, B. Gasparian *Corresponding author. E-mail: [email protected] Published 26 September 2014, Science 345, 1609 (2014) DOI: 10.1126/science.1256484 This PDF file includes: Materials and Methods Supplementary Text Figs. S1 to S16 Tables S1 to S7 References (39–192) Captions for Databases S1 and S2 Other Supplementary Materials for this manuscript include the following: (available at www.sciencemag.org/content/45/6204/1609/suppl/DC1) Databases S1 and S2

Transcript of Supplementary Materials for Early Levallois technology and the Lower to Middle Paleolithic...

www.sciencemag.org/content/345/6204/1609/suppl/DC1

Supplementary Materials for

Early Levallois technology and the Lower to Middle Paleolithic transition in the Southern Caucasus

D. S. Adler,* K. N. Wilkinson, S. Blockley, D. F. Mark, R. Pinhasi, B. A. Schmidt-Magee, S. Nahapetyan, C. Mallol, F. Berna, P. J. Glauberman, Y. Raczynski-Henk, N. Wales, E.

Frahm, O. Jöris, A. MacLeod, V. C. Smith, V. L. Cullen, B. Gasparian

*Corresponding author. E-mail: [email protected]

Published 26 September 2014, Science 345, 1609 (2014) DOI: 10.1126/science.1256484

This PDF file includes:

Materials and Methods Supplementary Text Figs. S1 to S16 Tables S1 to S7 References (39–192) Captions for Databases S1 and S2

Other Supplementary Materials for this manuscript include the following: (available at www.sciencemag.org/content/45/6204/1609/suppl/DC1)

Databases S1 and S2

Materials and Methods Excavation History of NG1 NG1 was discovered during a walkover survey of the middle Hrazdan Gorge in June 2008. The site is exposed over a 135m length in a road bulldozed by the Armenian military from their base immediately above the site. Obsidian artifacts were encountered as in situ finds in the exposed fine-grained alluvial sequence, as ex situ finds at the base of the stratigraphic section and in the road, and as slope finds that were pushed over the edge of the road when the in situ sediments were truncated by bulldozing. The quantity of artifacts, their provenance within a palaeosol and the technology used to produce them suggested that NG1 was a site worthy of detailed study. Excavations were carried out for seven weeks in June–July 2008 and 2009 in order to a) recover a statistically meaningful sample of lithic artifacts, b) to determine whether hominin activities were restricted to particular loci or spread homogenously across the exposed stratigraphic section, and c) collect a full spectrum of samples for palaeoenvironmental and chronometric investigation. The excavations were directed by D.S. Adler and B. Yeritsyan, and labor was provided by experienced graduate students and undergraduate students enrolled in the University of Connecticut’s Field School in Armenian Prehistory, directed by D.S. Adler. Geological Context The alluvial sediments containing the NG1 archaeological site are bounded at their upper and lower contacts by basaltic trachyandesites that originated as lava flows from the Gegham range, a chain of circa 100 Late Miocene, Pliocene (but see 39), and Quaternary volcanoes to the east of the site (Fig. S1) (27, 28). The upper basaltic trachyandesite (Basalt 1) is the last lava produced by the Mensakar volcano and has been mapped over a distance of 24 km in the central and lower parts of the Hrazdan Gorge. The lower lava (Basalt 7) probably also has its origins in either Mensakar or Gutanasar, but as it is buried by Basalt 1, it can only be seen in the walls of the Hrazdan Gorge over a 11-km distance. Seven further basalts underlying Basalt 1 were mapped during a geomorphological survey carried out in the central Hrazdan Gorge in 2009, all originating from volcanoes of the Gegham range. These basalts are locally interbedded with alluvial and lacustrine deposits, but so far Paleolithic artifacts have only been found in the uppermost of these sealed sediment beds at NG1. The Quaternary lava flows were constrained within the Hrazdan Gorge by Early Pleistocene lacustrine and volcano-lacustrine deposits on their western side (40, 41) and by Late Miocene-Early Pliocene andesite lavas (Kaputan Formation) and Upper Pliocene basalt to the East (26, 27). The Pleistocene-Miocene lava sequence sits, in turn on deposits of the Zangian Formation, a body of marine sands and clays with a mollusk fauna indicating an origin in the Caspian Sea and dating to the middle Miocene (42, 624–629). Prior to our study the chronology of the Gegham basaltic lavas was known as a result of 40K/40Ar and 40Ar/39Ar dating of basaltic trachyandesites from the Aknotsasar, Mensakar, Gutanasar, Hatis, Lodochnikov and Sevkatar volcanoes. Ages of between 550 and 70 ka have been reported (27, 28). Fission track (FT) dating of obsidian in rhyolite-perlite flows from the Alapars and Fantan domes, and the Djraber extrusion to the west of the Gutanasar volcano has produced results between 210–330 ka, while obsidian dikes from the Hatis volcano have been FT dated to the range 210–400 ka (43, 26, 377). Given that these obsidians are the raw material for the NG1

artifacts, the chronological data might suggest that the rhyolitic volcanism was active at the time of hominin activity. There is, however, some disagreement on the flow chronology. Obsidians from dikes in the Hatis volcano dated by both 40K/40Ar and FT have produced ages of 650 ka and 330 ka, respectively (44, 45), while FT dates from obsidians throughout the Gutanasar complex principally cluster circa 310 ± 30 ka (43, 44). In contrast Fantan obsidian was 40K/40Ar dated to 480 ± 50 ka. Thus, fundamental chronological debates remain, which we intend to address in the next phase of our work. Stratigraphy, Micromorphology, and Mineralogy Sedimentology Two columns of bulk samples were collected as continuous 5cm-thick blocks from the NG1 alluvial sequence in 2008 and 2011 and transported to the University of Winchester for laboratory study. Both sets of samples were initially air dried at 40oC and homogenized using a mortar and pestle. The 2008 samples were then each divided in two, one split being passed through a 250 µm and the other through a 2 mm mesh. Both sample fractions were used for separate dual mass specific magnetic susceptibility measurements following established procedures (46, 221–226). Organic carbon content was then determined by combusting the sample splits previously used for magnetic susceptibility measurement at 550oC for four hours and measuring the weight loss. The 2011 samples were used for grain size measurement, which was carried out using dry sieve and pipette methods (46, 86–94). Micromorphology Undisturbed blocks of sediment for micromorphology were collected from the NG1 stratigraphic sequence in Kubiena boxes during the 2009 (5 blocks) and 2013 (2 blocks) field seasons. The blocks were dried in an oven at 60˚C for 48 hours and then imbedded in a mixture of unsaturated polyester resin, styrene, and a catalyzer (MEKP) in a 7:3:0.025 ratio at the Chemistry department of University of La Laguna, Tenerife, Spain. Upon curing, they were cut into 7 x 5 x 1 cm slabs and shipped to Spectrum Petrographics Inc., Vancouver, USA for the manufacture of 18 thin sections (samples MM1–5) and to CENIEH, Burgos, Spain, for 5 thin sections (samples MM6 and 9). All thin sections are 30 µm-thick. They were observed under a polarizing Nikon Eclipse E-800 microscope at 2x, 4x, 10x, and 20x. Standardized descriptive guidelines (47) were used. Mineralogy In 2008, a column of samples was taken at 0.02 m intervals for Fourier transform infrared spectroscopy (FTIR). Representative mineralogical samples were obtained by homogenizing several grams of collected sediment. Powdered samples were analyzed by FTIR spectroscopy using a Thermo-Nicolet Nexus 470 FTIR spectrometer. A few tens of micrograms of homogenized sample or discrete particles were ground with an agate mortar and pestle. About 0.1 mg or less of the sample was mixed with about 80 mg of KBr (IR-grade). A 7 mm pellet was made using a hand press (Qwik Handi-Press, Spectra-Tech Industries Corporation) without evacuation. The spectra were collected between 4000 and 400 cm-1 at 4 cm-1 resolution. Macroscopic observations The stratigraphic sequence of NG1 comprises five lithological units (Units 5–1) of predominantly alluvial genesis that formed in two cycles (Figs. 2, S2–S5). Cycle 1 comprises

Units 5–2. Normal bedding in Units 5–3 suggests that deposition was initially on a channel to floodplain interface (Units 5 and possibly 4) but later on the floodplain, during low energy flood events (Units 4 and 3–2). Macroscopic features indicative of periodic waterlogging and incipient soil formation were observed in Units 4–2. The artifact-bearing Unit 2 has a dark gray color and a relatively high organic content (6–8%, cf. 4–6% in Units 5–3) (Table S1). Cycle 2 is separated from Cycle 1 by the unconformity at the top of the palaeosol (Unit 2) and is represented only by Unit 1. Micromorphological observations All of the stratigraphic units (5–1) exhibit a homogeneous lithological composition comprising polymictic (polygenetic) sand-sized pyroclastic shards and few quartz grains in a clayey groundmass. The pyroclastic composition is basaltic and comprises common feldspar (calcic plagioclase and sanidine), pyroxene, olivine, sphene, and vesicular and fibrous glassphenocrysts as well as few trachytic shards. These and the isolated phenocrysts show variable angular to rounded surfaces and do not exhibit strong alteration states. There are also few sand-sized, subrounded detritic rocks (quartzite and weathered limestone). In Units 4–1 the sand is unsorted, whereas the sand in the top of Unit 5 it is moderately well sorted and finer-grained (fine sand-sized and smaller) (Table S2). All of the stratigraphic units exhibit iron mottling indicative of poor drainage (Figs S6) and Units 4–1 show intersecting channels (<1 cm) filled with micritic or needle-fiber calcite. Overall, this calcitic microfabric is comparable to documented examples of pedogenic laminar groundwater calcrete (48–50). Unit 3 exhibits a granostriated b-fabric indicative of in situ clay translocation and few irregular fissures throughout the unit. Frequent massive, strongly birefringent clay infillings were observed in Units 2–1. Unit 2 contains common microscopic humified plant matter in a granular groundmass bioturbated by rootlets (Fig S7A). Microscopic fragments of obsidian flakes are present in Units 3–1 (Fig S7B). Mineralogical observations (FTIR) The FTIR results are summarized in Table S3, in which the identified mineral phases are given in decreasing absorption intensity. The analyzed sediments contain feldspars (plagioclase), quartz, carbonates, and clay minerals such as kaolinite, smectite, and/or illite. The sediments of Unit 1 and at the top of Unit 2 (Samples NG1-01 to NG1-12) show a characteristic IR absorption at 3688 cm-1 that could be assigned to hydroxyl vibration of low crystallinity kaolinite and/or of other serpentine group minerals. The carbonate concretions show absorptions of calcite sometime mixed with clay minerals. The matrix of the sediments does not show carbonate absorptions, suggesting that the calcite is mainly “confined” into infillings and laminations. Conclusions Field observations together with bulk, micromorphological, and mineralogical analyses indicate that the NG1 stratigraphic sequence represents an alluvial deposit formed by concurrent episodes of alluvial sedimentation of polygenetic pyroclastic material and pedogenesis in a poorly drained aggradational floodplain setting. The artifact-bearing Unit 2 represents the A horizon of a palaeosol of the inceptisol order, specifically an andept [an incipient soil formed on pyroclastic material and characterized by high organic matter accumulations and a weakly developed B horizon (51)]. Unit 1 represents renewed concurrence of alluvial sediment deposition and pedogenesis.

The implications of these geoarchaeological data are relevant to the context of the archaeological material and its interpretation. The alluvial, low-energy sedimentary environment associated with Units 4–1 and the existence of a vegetated surface at the top of Unit 2 prior to the deposition of Unit 1 alluvium indicates that the vast majority of the NG1 lithic assemblage was deposited by hominins on the stable surface of Unit 2 in the same stratigraphic position as it was later excavated by our team. Postdepositionally, the degree of reworking associated with bioturbation, clay translocation, and laminar calcrete formation as documented in Units 2 and 1 likely explains the minor vertical spread of the small (<2.5 cm) lithic material. These data compliment those regarding lithic conjoins discussed below and allow us to conclude that the lithic material from NG1 is largely in situ, ± a few centimeters. This degree of spatial integrity is generally uncommon among Paleolithic sites that did not experience rapid low-energy burial and is entirely absent at other LMP sites in the Caucasus. Excavation Methods An excavation/collection grid of 1x1m squares was established on the site prior to archaeological work. All artifacts found protruding from the stratigraphic section or on the roadbed immediately below the stratigraphic section were mapped in three dimensions, while artifacts found on the slope below the road were collected but not mapped (see Fig. S3–S5). Several broken artifacts found along the slope could be refitted to pieces excavated from the stratigraphic section, indicating that most if not all artifacts likely originated from the exposed fine-grained sediments. Archaeological excavations were subsequently carried out by digging into the stratigraphic section at the interface of the overlying basalt (Basalt 1) and underlying alluvial strata in nineteen 1m2 units spread along the exposure. The alluvial strata were then excavated in stratigraphic order downwards to the upper surface of the lower basalt (Basalt 7) wherever it was exposed. Excavations were positioned along the stratigraphic section in areas where the overhanging basalt was considered stable, and they penetrated into the stratigraphic section until the stability of the basalt was deemed unsafe (maximum 0.40 m). Sediment compaction and secondary carbonate formation in many excavation units required the periodic use of hammers and chisels, but otherwise WHS trowels were employed. This method resulted in a horizontally broad but shallow excavation that removed 2.358m3 of sediment along the stratigraphic section. During this procedure, the positions of all lithic artifacts (bone was not preserved on the site) were recorded in three dimensions using two Leica TCR805 total stations linked to laptop PCs running EDMWIN, an archaeologically specific GIS program (52). The total stations were also used to collect measurements along stratigraphic boundaries between units in the fresh vertical stratigraphic sections that were then imported into NewPlot, another GIS program developed by McPherron and Dibble for analysis of point-provenanced artifacts. Spot samples were collected from the cleaned stratigraphic section at the locations noted in Fig. S5 for optically stimulated luminescence (OSL) and 40Ar/39Ar dating as well as micromorphological study. Columns of samples were taken at 0.02 m intervals for FTIR measurement and for phytolith and palynology study, as continuous 0.02 m-thick blocks for tephrochronological study, and in 0.05 m-thick blocks for grain size analysis and magnetic susceptibility measurement (Fig. S5). In addition, all excavated sediment was retained with appropriate provenance, wet sieved through a 1.6 mm mesh, and dried. All artifacts and wet-sieved samples were subsequently transported to the project laboratory in Yerevan, where they were washed, labeled, bagged, photographed, analyzed, weighed, and sorted, and ultimately transported to the Institute of Archaeology and Ethnography, Armenian Academy of Science, Yerevan for curation and storage.

40Ar/39Ar Dating Samples of basalt for 40Ar/39Ar dating were collected by the simple expedient of breaking away single rock fragments of c 1 kg from the outcrop using a geological hammer (Fig. S5). The rock samples were placed in labeled zip lock bags and transported to the NERC Argon Isotope Facility, Scottish Universities Environmental Research Centre (SUERC) for argon isotope analysis. Sanidine was extracted from the tephra in the uppermost 5 cm of Unit 1 (Figs. 2 and S4) and 40Ar/39Ar measurements were made on the sanidine crystals that were recovered. Samples were prepared at the SUERC and (cryptotephra) at Royal Holloway, University of London. Basalt preparation The collected samples (Fig. S5) contained abundant phenocrysts and xenocrysts (e.g., olivine, pyroxene, quartz, and feldspar). These can often contain significant amounts of atmospheric or excess argon (e.g., 53), and as such, samples were meticulously prepared for 40Ar/39Ar dating using the methods outlined in Mark et al. (53). Briefly, visibly altered areas of the whole-rock basalt samples were removed using a slow rotating saw. The samples were crushed and sieved to a grain size of 125–250 µm, washed in de-ionized water, and then passed through a magnetic separator to remove olivine and pyroxene phenocrysts. The phenocryst-free fraction was leached in 10% HNO3 and washed (sonicated) in de-ionized water. This process was repeated six times until the water remained clear. Using a binocular microscope, small fragments of groundmass, free of phenocrysts (e.g., olivine, pyroxene) and xenocrysts (e.g., quartz, feldspar), were handpicked. Approximately 300–500 mg of groundmass for each sample was harvested, parceled into small Cu packets, and placed within Al discs for irradiation. Tephra preparation The visible tephra layer in Unit 1 was examined for the presence of both vitreous glass and sanidine crystals. Analyses of the vitreous phase followed Blockley et al., (54) and glass was identified using a petrological microscope with polarised light. Abundant glass shards were observed in the tephra layer and were then extracted and mounted for geochemical analyses. Shards were mounted onto stubs and ground and polished before analyses by electron probe microanalysis (EPMA). Single-grain major and minor element compositional analysis was carried out using a JEOL JXA-8600 electron microprobe with four wavelength spectrometers at the Research Laboratory for Archaeology and the History of Art, University of Oxford. A 15 kV beam voltage, 6nA current, and defocused (~10 micron) beam were used for analysis. The following elements were analyzed in each glass shard: Si, Ti, Al, Fe, Mn, Mg, Ca, Na, K, and P. Each element was analyzed for 10-60s on peak (10s for Na, 30s for most other elements, and 60s for low abundance elements), and the background was determined for counting for half the peak collection time on either side of the peak. Oxide concentrations for all the elements were determined by stoichiometry. Data were filtered to remove any analyses that may have hit a mineral and those with low analytical totals (<93%). Each analytical run included secondary reference glasses: ATHO-G, GOR132/5 and St/Hs6/80-g MPI-DING glasses (55). The major and minor element geochemical data are available in Database S1 along with summary secondary standard data. The tephra is rhyolitic in composition, differentiating it from Basalts 1 and 7, and the presence of fresh glass and abundant sanidine crystals indicates primary airfall.

Sandine crystals were extracted for 40Ar/39Ar dating from the single tephra sample using the method outlined by Hynek et al. (56) but incorporating some minor modifications. These modifications permitted the extraction and analysis of the appropriate fraction (Sanidine feldspar) in a sample where the target material was of a finer grain size than that prepared by Hynek et al. (56). Their published methodology is considered to be highly effective with grain sizes in excess of 250 µm, but it is also applied successfully to those as small as 125 µm. However, in the NG1 Unit 1 sample, the dominant grain size was 80–125 µm with only a small component of the sample exceeding the 125 µm size fraction. Following sieving, samples were subjected to a stepped density flotation procedure using sodium polytungstate (SPT: Na6[H2W12O40]). This was used to extract grains of densities of <2.5g/cm3 and between 2.5 and 2.7g/cm3. The former was applied to clean the sample of grains of lower density than the target feldspar (c.2.5–2.6g/cm3), and the latter fraction was to concentrate potential sanidines and was retained for further processing. All fractions (<2.5 and >2.7g/cm3) were assessed to ensure that the optimum extraction density had been attained. Samples were washed in deionized water to remove SPT prior to further processing. The technique of Hynek et al., (56) involves etching the surface of prepared grains in Hydrofluoric Acid (HF) vapour prior to staining in a sodium cobaltnitrite solution. The etching ensures a clean and ‘activated’ grain surface that allows the sodium cobaltnitrite solution to differentially stain the surface of quartz, orthoclase, plagioclase, and sanidine feldspars, thus aiding their identification. Grains that stain bright yellow indicate a high-concentration of potassium, whereas grains that do not stain indicate low or absent potassium content. This, therefore, allows the high-potassium sanidine feldspars to be isolated, handpicked, and processed for further analysis. The challenge of working with finer-grained samples than Hynek et al. (56) was that published HF exposure times and proximities resulted in disintegration of the grains, rendering them unsuitable for further analysis. Consequently, a stepped exposure duration and proximity experiment was set up in the laboratories at Royal Holloway University of London. This determined the preparation conditions that ensured the highest return of grains for dating. Comparisons of the optimum conditions have been identified in Table S4. Following handpicking, grains were thoroughly washed in distilled water. Following the final wash in distilled water and methanol, the grains were parceled into Cu packets and positioned within an Al holder for irradiation. Sample irradiation and neutron fluence monitor analyses All 40Ar/39Ar measurements were made at the Natural Environmental Research Council Argon Isotope Facility, which is housed at the SUERC. International neutron fluence monitor standard Alder Creek Tuff sanidine (ACs, 1.193 ± 0.001 Ma; 57), a secondary standard referenced against a Fish Canyon sanidine (FCs) (28.02 ± 0.16 Ma) age (58), was loaded adjacent to the samples of unknown within a flame-sealed within a quartz vial. Samples were irradiated in two batches:

1. Basalt samples were Cd-shielded and irradiated for 0.08 hours in the RODEO facility at the McMaster reactor (Canada).

2. The sanidine crystals were irradiated in the Cd-lined (CLICIT) facility of the Oregon State University (USA) TRIGA reactor for 1.92 hours.

Neutron fluence monitor grains were analyzed by single crystal total fusion with a focused CO2 laser. Twenty-five crystals from the 6 different irradiation discs (one sample per disc) were analyzed (n = 150) to facilitate a J-parameter precision of c. 0.2% for each sample. J values and irradiation interference correction values are presented in a raw 40Ar/39Ar data file (Database S2). Isotope measurements were made using a MAP 215–50 noble gas mass spectrometer (as described in 59). Note: the mass spectrometers at SUERC (MAP 215-50, ARGUS and HELIX) are routinely inter-calibrated by dating of standards from the same irradiation positions and measurements of atmospheric 40Ar/36Ar and 36Ar/38Ar air ratios using a peripatetic air bottle. No systematic offset in data is observed, and there is routinely excellent inter-mass spectrometer agreement. Basalt analyses The basalt samples (300–500 mg aliquots of groundmass) were step-heated using a resistively heated double-vacuum furnace over a temperature range from 500 to 1750 °C. Isotope data were collected using a GVI ARGUS multi-collector mass spectrometer, which has a measured sensitivity of 7 × 10-14 moles/volt (60). Samples were heated for 5 minutes prior to 10 minutes cleanup. Extracted gases were cleaned using 3 GP50 SAES getters (two operated at 450 °C and one at room temperature) and a cold finger maintained at -95 ˚C using an acetone-CO2(s) slush trap. The extraction, clean up, and data collection processes were entirely automated. Experiments were conducted over 7-hour periods with hot furnace blanks (500 to 1750 ˚C) collected prior to every sample run. Average backgrounds ± standard deviations from all five blank runs (n = 30) were used to correct isotope abundances. Air calibrations were collected in batches (n = 10) immediately before and after individual experiments to monitor mass discrimination. Average 40Ar/36Ar values ± standard deviation (300.06 ± 0.17, n = 104) was used to calculate discrimination factor using the power law (61). Sanidine analyses Single grains (80-125 µm) of unknowns (NG tephra) were loaded into a Cu planchette in an ultra-high vacuum laser cell with a doubly pumped ZnSe window. Using a CO2 laser, the sanidine crystals were degassed at low temperature and subsequently fused. Radiogenic 40Ar (40Ar*) yields were thus improved; no 39Ar was liberated. All gas fractions were subjected to 180 s of purification with two SAES GP50 getters (one at room temperature the other at 450 ˚C) and a cold finger maintained at -95.5 ˚C using a mixture of dry ice (CO2[s]) and acetone. Argon isotope ratios (i.e., ion beam intensities) were measured using a MAP 215-50 mass spectrometer in peak jumping mode. The mass spectrometer has a measured sensitivity of 1.13 × 10-13 moles/volt. Backgrounds were measured after every two analyses of unknowns. Average backgrounds ± standard deviation (n = 42: 40Ar 1.02 × 10-15 moles, 39Ar 3.10 × 10-17 moles, 38Ar 1.90 × 10-17 moles, 37Ar 7.85 × 10-17 moles, 36Ar 1.38 × 10-17 moles) from the entire run sequence were used to correct raw isotope measurements of unknowns. Mass discrimination was monitored by analysis of air pipettes after every five analyses (n = 22, 40Ar/36Ar = 290.51 ± 0.61), with the power law used to calculate discrimination factors (61). Data handling and processing

The Ar isotope data were corrected for backgrounds, mass discrimination, and reactor-produced nuclides and processed using standard data reduction protocols (e.g., 62). Decay constants (63) and atmospheric argon ratios (64), the latter independently verified (65), were employed. The BGC software MassSpec was used for data regression. Data are displayed on age spectra, ideograms, and isotope correlation plots (inverse isochron plots). Plateau criteria: Uncertainty overlap at 2-sigma with confidence value of 95%, minimum amount 39Ar 60%, and minimum number of contiguous steps is three. Plateaus are mean weighted by inverse variance with associated uncertainty as standard error of the mean (SEM) for the mean square weighted deviation (MSWD) <1, and if the MSWD was >1, then calculated using SEM × MSWD0.5. Inverse isochron criteria: MSWD of contiguous steps with 0.05 probability cutoff, 60% 39Ar as minimum amount of gas and at least 3 steps. Uncertainties in age and trapped component were estimated using Monte Carlo simulation. Ideogram filtering: The SEM was determined for all samples that displayed a Gaussian (normal) distribution with a MSWD < 1. A data filter was used to screen the single crystal Ar/Ar ages for xenocrysts: any Ar/Ar age > 1.5 Median Absolute Deviations (nMADs) from the weighted mean was rejected from the age calculations. Data were not rejected on the basis of %40Ar*. Complete raw data are reported according to standard protocols (66) in Database S2. NG1 lava flows (Basalt 1 and Basalt 7) Sample NG1, Basalt 1: The step heating experiment (Fig. S8) yielded fully concordant (100% 39Ar) plateaus with an inverse isochron that defines atmospheric (64) initial trapped composition and an age that is indistinguishable from the age spectrum. Sample NG1, Basalt 7 (a, b): Both step heating experiments (Fig. S8) yielded fully concordant (100% 39Ar) plateaus with inverse isochrons that define atmospheric (64) initial trapped compositions and ages that are indistinguishable from the age spectra plots. The weighted average of all steps (i.e., both plateaus) is 441 ± 6 ka. NG1 tephra Twenty-seven sanidine crystals were fused with seventeen defining a statistically significant juvenile age population (the others deemed to be xenocrysts) (Fig. S9). These seventeen data points cast on an isotope correlation plot define an inverse isochron with atmospheric (64) 36Ar/40Ar (at 2 sigma uncertainty) and an age indistinguishable from the weighted mean population. Owing to xenocrystic contamination small populations of crystals were not step heated. The recovery of a statistically reproducible age for a significant juvenile population (63% of dated grains) supports the interpretation that the top of Unit 1 incorporates primary air fall micro-tephra originating from a single eruption. Its age makes is consistent with stratigraphic relationships to Basalt 1 and Basalt 7. The xenocrysts in the sample (1000–575 ka) all predate the basalt emplacement at NG1 (441–197 ka). 40Ar/39Ar ages relative to optimization model (67, 68)

The 40Ar/39Ar method is a relative dating technique with all ages referenced back to a standard of known age. A recent optimization model (67, 68) used constraints from 40K activity, K-Ar isotopic data, and pairs of 238U-206Pb and 40Ar/39Ar data as inputs for estimating the partial decay constants of 40K and 40Ar*/40K ratio of FCs. This calibration has reduced systematic uncertainties (i.e., improved accuracy) in the 40Ar/39Ar system and has yielded an age for FCs that is indistinguishable at the 2σ confidence level from an astronomically tuned FCs age (69). To present the most accurate 40Ar/39Ar age for the volcanic deposits from NG1, we have recalculated our age (against 58, 63) relative to the recent optimization model (67, 68). Data define ages of:

1. Basalt 1: 197 ± 7 ka 2. Combined age for Basalt 7 samples (a, b): 441 ± 6 ka 3. Unit 1 tephra: 308 ± 3 ka

The ages include full systematic uncertainties (including decay constant uncertainty) and are reported at 1 sigma (68.2% probability) confidence. Note that these ages are reported throughout the main text. Raw isotope data are presented in Database S2 to allow for ages to be recalculated relative to other standard ages and decay constants should readers see purpose to do this. NG1 Lithic Assemblage Excavations in 2008 and 2009 resulted in the recovery of 2,976 obsidian artifacts: 1,576 were excavated in situ and plotted in three dimensions, 1,163 were recovered from 122 three-dimensionally plotted water-sieved sediment samples, and 237 were collected from the road and slope in front of the stratigraphic section (Table S5, Fig. S3–S5). All sediment was collected during the excavation as plotted samples of roughly 20 liters in volume, and 134 sediment samples (51% of all sediment samples) were wet-sieved and picked to extract archaeological material (only obsidian artifacts were recovered), which was sorted into size classes, counted, and weighed (of the 134 sediment samples 12 did not contain any artifacts). This procedure resulted in the recovery of 1,163 artifacts (>2.5cm, n=0; 2.4-1.5cm, n=121; <1.5cm, n=1042), weighing 390.5 g from 1.206m3 of sediment (964.3 wet-sieved artifacts/m3). Unfortunately, the sample of unbroken specimens among the excavated and recovered artifacts is too low (see Table S5) to allow for meaningful statistical analyses of morphometric attributes, thus the following discussion focuses on summary statistics and qualitative aspects of the assemblage. Artifact context and condition All of the 2,976 artifacts recovered from NG1 are produced on obsidian with the vast majority resulting from hard-hammer percussion. Table S5 summarizes the assemblage by stratigraphic unit and technology, typology, and fragmentation, and Figure 4 and Figures S10–S14 provide representative illustrations of the assemblage. The artifacts are concentrated within Unit 2 (~20 cm thick) but appear to be distributed randomly across the exposed stratigraphic section and excavation units, and no discrete loci of intensive activity were identified. Artifact orientations along the long axis were normally horizontal, but a slight degree of post-depositional reorientation was noted on several pieces. Surface preservation of the obsidian artifacts was black to light gray in color (penetrating <1 mm), indicating post-depositional weathering.

Considerable variation was observed among artifact edges regardless of artifact mass or edge thickness. Many artifact edges were fresh, while others were weathered and abraded or unintentionally notched/denticulated. The variance in artifact edge preservation is likely due to different rates of artifact use, surface exposure, trampling on the palaeo-Hrazdan floodplain, and burial; however, high-energy movement of the artifacts is not supported by the geoarchaeological data presented above. Unfortunately, for many individual artifacts, it proved impossible to distinguish between these possible sources of edge damage. We conclude that the archaeological materials within each stratigraphic unit are geologically contemporaneous and result from repeated hominin activity on stable surfaces on this portion of the palaeo-Hrazdan floodplain. Following the classification of all artifacts according to standard techno-typological categories, no technological or typological differences were observed between the artifacts recovered from Units 4–1 (Table S5). This observation, in addition to low sample sizes from all stratigraphic units, allowed the finds to be studied as a single assemblage. Two refit groups (RGs) consisting of five artifacts were identified in the laboratory, and these strengthen the geoarchaeological arguments presented above for the post-depositional integrity of the lithic assemblage. The artifacts in RG1 (Fig. S10, top) illustrate three reduction refits and are derived from Unit 1 (artifacts 1 and 2, see Fig. 2) and the slope in front of the stratigraphic section (artifact 3), thus linking the material truncated during road construction to the in situ material excavated from the stratigraphic section (see Fig. S3). RG1 documents the bidirectional production of large blades during an early phase in the reduction sequence, followed by flake production perhaps related to the re-preparation of core convexities. Artifacts 1 and 2 were excavated in situ, and they are separated in space by 21 cm, 10 cm, and 0 cm in the x, y, z dimensions, respectively. RG2 (Fig. S10, bottom) documents an example from Unit 2 of expedient tool production and recycling, during which a large flake was truncated and both pieces (artifacts 4 and 5) were retouched along the truncation. Unfortunately, it is impossible to estimate how much time elapsed between the production of the flake and its eventual truncation and transformation into two retouched tools. Artifacts 4 and 5 from RG2 are separated in space by 11.3 cm, 1.2 cm, and 0 cm, x, y, z, respectively. The three-dimensional proximity of the two sets of artifacts from RGs 1 and 2 suggests that these and many other artifacts were deposited on stable surfaces and strengthens the argument that post-depositional processes have not greatly altered the original burial context of the NG1 assemblage. Assemblage characteristics Assemblages classified as Late Acheulian, such as those outlined in Figure 1 and Table S6 (4–10, 13, 22, 70–154), span the lengthy period between the Early Stone Age (ESA) and Middle Stone Age (MSA) in Africa, and the Lower Paleolithic (LP) and Middle Paleolithic (MP) in Eurasia. In Africa, such assemblages are typically characterized by a diversification of lithic reduction strategies and the gradual replacement of Mode 2 bifaces by Mode 3 flakes, points, and blades produced through various Levallois methods (4–6). In Eurasia, the pattern of technological evolution is broadly similar and largely contemporaneous with that identified in Africa (7–9). While in southwest Asia, Levantine sites with assemblages assigned to the Acheulo-Yabrudian (AY) document the non-Levallois manufacture of blades, broad flakes, and Quina scrapers accompanied by the gradual disappearance of bifaces (10). However, in Eurasia, biface technology remains a persistent but not dominant element of many MP industries until the end of the LMP around late OIS 6. Therefore, most researchers discuss the ESA–MSA or LP–

MP transition in terms of the general abandonment of biface technology (Mode 2) in favor of hierarchical core technologies (Mode 3), among which Levallois methods are the most prominent. In all instances, whether referring to the beginnings of the MSA or the MP, we are discussing a gradual, intermittent, often non-linear process by which specific aspects of lithic technology are lost or added, against a backdrop where technological and behavioral variability is the rule rather than the exception. In addition to a distinction based on the presence/absence of biface and/or Levallois technology, we can further differentiate among assemblages spanning the ESA–MSA and LP–MP transition by the presence/absence of steeply retouched tools (mostly unifacial scrapers with Quina or Quina-like retouch) and the presence/absence of blade production (Table S7). Following these criteria, the lithic material from NG 1 can be classified as a unique Late Acheulian transitional assemblage composed of biface and Levallois technology, Quina retouch, and blade production. The Levallois assemblage The Levallois method is consistently documented among roughly one quarter of the artifacts in all artifact classes (cores, 44.7%; tools, 24.6%; flakes, 24.5%; blades, 51.9%, débordants, 22.2%, overshot, 13.8) and accounts for 24.0% of the total debitage assemblage not including shatter or artifacts ≤2.5 cm (Table S5). The relative frequency of Levallois material in the entire assemblage and in individual artifact classes is higher at NG1 than that reported for other Late Acheulian sites (Table S6). Seventeen cores from NG1 conform to a minimum of five of Boëda’s (14) six criteria for Levallois, and these are typically produced on large flakes. More than half of these cores retain a portion of cortex on the dorsal surface, and the ventral surface is transformed into the flake release surface. Where core scar morphology could be discerned, unidirectional (n=11, 64.7%) and centripetal (n=3, 17.6%) patterns predominate. The most common Levallois cores are recurrent (n=4, 23.5%) and preferential (n=6, 35.3%) (Fig. 4). The average length (n=15; =66.7mm; σ=25mm) and width (n=15; =61.3mm; σ=22.8mm) of the cores exceed similar average measures for all categories of unbroken debitage, and the maximum last scar on the flake release surface of the cores (n=11; =63.2mm; σ=21.2mm) covers on average 80% of the average core length (n=11; =78.3mm; σ=25.8mm). Core platform edge angles are variable, but all are classified as steep since they approach but never exceed 90°. Such measurements are in keeping with Levallois core edge angle parameters as defined and supported by previous research (14, 16, 155). Débordants document the preparation and rejuvenation of dorsal convexities and flake release surfaces, as attested by their dorsal scar morphologies originating primarily from the lateral edge of the core. Overpass flakes appear to document the failed removal of preferential Levallois flakes (Fig. S11). As with the Levallois flakes discussed below, many overshot flakes exhibit dorsal scar patterns that are predominantly unidirectional but with opposing scars originating from the distal end of the core (Fig. S11.6, 11.7). Lengths for unbroken overshot flakes are highly variable (n=19; =59.6mm; σ=33.3mm), more so than for unbroken Levallois flakes (n=51; =46.8mm; σ=19.0mm). Therefore, such flaking errors occurred during all stages of reduction, perhaps due to the brittle nature of the raw material and, thus, the application of too much force during knapping was likely a perennial concern.

Levallois flakes and blades account for 74.6% of the entire Levallois assemblage, the platforms of which are typically plain (50.0% and 11.1%, respectively) or faceted (41.7% and 88.9%, respectively). Dorsal scars for unbroken Levallois flakes and blades are principally unidirectional with evidence for the lateral and distal preparation of core flake release surfaces. Average lengths for unbroken Levallois flakes (see above) and complete Levallois blades (n=8; =61.6.8mm; σ=24.9mm) highlight the variable dimensions of the Levallois cores from which they were struck and suggest that Levallois blade production may have been initiated early in the reduction sequence, when the exploitable volume of the core and core dimensions were at their greatest, and that Levallois flake production was initiated once core dimensions dropped beneath a certain threshold. While 24.6% of the retouched tools were produced on Levallois flakes or blades, artifact fragmentation, surface weathering, retouch intensity, and recycling limit the extent to which Levallois can be identified among all classes of artifacts. Whenever a technological determination was considered problematic, the conservative decision was made to count such objects in the “non-Levallois” category. Therefore, the preceding discussion and the data presented in Table S5 likely underestimate the true frequency of Levallois cores, flakes, and blades in the assemblage, particularly among the retouched tools. The non-Levallois assemblage The non-Levallois assemblage from NG1 documents a variety of hierarchical core methods all of which exhibit “…a stable relationship between fracture initiation and flake release surfaces through a succession of flake removals” (156: 165). Shea distinguishes between bifacial hierarchical core methods (e.g., Levallois cores: his Mode F1, preferential; or Mode F2, recurrent), which employ two flake release surfaces - one for fracture initiation and the other for actual flake production - and unifacial hierarchical core methods (e.g., non-Levallois: his Mode G1, platform cores; Mode G2, blade cores), which rely simply upon a “stable hierarchy of striking platform and flake release surfaces” (156: 167). The terms “hierarchical” and “non-Levallois” are used here interchangeably. The purpose is simply to recognize the stable, hierarchical relationship between the striking platform and the flake release surface of non-Levallois cores while also distinguishing them from their more complex Levallois counterparts and simpler pebble core methods, which are not represented at NG1. The non-Levallois cores at NG1 are classified as hierarchical (26.3%), core-on-flake (15.8%), blade (7.9%), and Kombewa (5.3%) (Fig. S11). All of these core types are defined as hierarchical; however, distinctions are drawn, as in the last three categories, to highlight discernable morphological differences. As a group, the non-Levallois cores vary considerably in size and morphology. The average length (n=8; =58.0mm; σ=26.4mm) and width (n=8; =45.5mm; σ=15.5mm) of unbroken hierarchical cores is similar to the average length (n=5; =60.0mm; σ=12.5mm) and width (n=5; =51.1mm; σ=11.3mm) of unbroken cores-on-flakes. The maximum last scar on the flake release surface of unbroken hierarchical cores (n=5; =41.0.8mm; σ=20.0mm) covers on average 64% of the average core length (n=5; =63.8mm; σ=25mm). Core platform edge angles are variable but are less than 90° on all specimens. Unlike the Levallois cores described above, dorsal scar patterns for the non-Levallois cores are highly variable and show no clear pattern for the lateral or distal management of core convexities. Among the non-Levallois flakes, platforms are predominately plain (52.2%) or faceted (41.3%), and the prevailing dorsal scar pattern is unidirectional, again with very limited evidence for the lateral or distal shaping of core convexities. Unbroken débordants (n=12) and overshot flakes

(n=14) exist but are difficult to assign scar morphologies. The average length ( =54.5mm; σ=23.1mm) and width ( =37.5mm; σ=14.6mm) of the débordants is very similar to the same measures for the overshot flakes ( =48.7mm; σ=21.8mm; and =31.6mm; σ=14.0mm, respectively). The assemblage contains a high frequency of small debitage (<2.5 cm) and shatter (Table S5) that cannot be assigned to a technological category. The biface assemblage The NG1 bifaces (n=7) are highly variable in size (e.g., 2,474.7–10,977.7 mm2 surface area; 46.5–384.4 g), morphology (cordiform n=3; ovate n=2; amygdaloide n=2), and degree of reduction (Fig. S12). Unlike the production of flake blanks, which is achieved through hard-hammer percussion, several of the bifaces as well as several resharpening and thinning flakes attest to the use of soft hammer percussion and the on-site modification of bifaces (Fig. S12). As a group, the bifaces display characteristics documented among specimens from both ESA and MSA/MP contexts. The large ovate and cordiform bifaces (Fig. S12.1–S12.2) are similar in morphology to Acheulian bifaces found throughout Eurasia (see 157). In contrast to the larger bifaces, the smaller cordiform, amygdaloide, and ovate bifaces (Fig. S12.4–S12.6) were manufactured on flake blanks. They are asymmetrical in edge shape and profile, and retouch is discontinuous and localized along lateral margins. The ovate biface (Fig. S12.3) is similarly asymmetrical with localized edge retouch, yet large invasive removals on one surface suggest classification as a “core-like” biface. The seventh biface (not illustrated) is an asymmetrical bifacially retouched Kombewa flake with distal thinning. The retouched tool assemblage Scrapers of various types, but in particular déjeté, transverse, and simple scrapers, dominate the retouched tool assemblage (n= 35, 53%; Figs. S13). Resharpening flakes attest to the on-site production and maintenance of the retouched tool assemblage (Fig. S14); however, conjoins linking these two artifact classes have not yet been possible. Quina scrapers were not recognized as a typological group; however, among the transverse and déjeté scrapers, six examples are produced on short, thick, wide flakes with expanding lateral edges to which has been applied Quina or demi-Quina retouch. These artifacts are similar to types produced on flint at the AY sites of Qesem Cave, Israel (158) and Hummal, Syria (159; cf. Table S6), but the Levallois concepts that characterize the NG1 assemblage are not present in these industries. Therefore, the NG1 assemblage cannot be accommodated within the definition of the AY and any of its three major industries (32). Numerous artifacts could be classified as denticulates and notched pieces; however, the majority of these likely result from post-depositional edge damage. Table S5 and Figure S14 document intentionally produced denticulates and notched pieces. Truncated-faceting, scaled, stepped, Quina, and demi-Quina retouch are the primary forms of intentional edge modification (Figs. S13–S14). Artifact sourcing by pXRF Geochemical analyses for obsidian sourcing were conducted using portable X-ray fluorescence (pXRF), an established and powerful technique for obsidian characterization worldwide (31, 160–172). Artifacts and geological specimens were analyzed for 90–120 seconds using a Niton XL2 and a Niton XL3t GOLDD+ pXRF analyzer, respectively. For recent pXRF instruments, these measurement times are sufficient, especially for the so-called “mid-Z” trace elements most important for obsidian sourcing (e.g., Rb, Sr, Zr), to attain analytical uncertainties equal to, if not

better than, those for instrumental neutron activation analysis (INAA; i.e., uncertainties ≤5-10%; 26: Fig. 6). Many artifacts were excluded from source analysis due to severely patinated surfaces resulting from 335–325 ka of hydration and weathering. All analyzed artifacts were measured at least twice in different locations as a means of quality control. Fundamental parameters (FP) correction adjusted the raw measurements for differences in composition and surface morphology as well as X-ray emission, absorption, and fluorescence. To the corrected data, as a means to ensure accuracy, we applied linear-regression calibrations based on 24 well-characterized obsidian standards analyzed by INAA and conventional laboratory-based XRF at the University of Missouri’s Research Reactor (MURR) and by electron microprobe analysis EPMA at the University of Minnesota. These obsidian standards were tested daily to monitor reproducibility. Artifacts were compared to specimens from our extensive reference collection of Southwest Asian obsidians. Figure S15 is a scatterplot of Rb, Sr, and Zr measurements for 316 NG1 artifacts and geological specimens from their four volcanic sources: Gutanasar (n=296 artifacts; the cone of which is 8 km to the NE, although several obsidian exposures occur within 2 km of the site), Pokr Arteni (n=10 artifacts), Hatis (n=9 artifacts), and Pokr Sevkar (n=1 artifacts). Figure S16 shows the geographic locations of these and other sources in Armenia, all of which have been sampled by members of our team. Spatial and temporal analysis of early Levallois technology It is our contention that the intercontinental transition from the ESA to the MSA in Africa and the LP to the MP in Eurasia occurred independently among different geographically dispersed hominin populations already adept at a variety of complex knapping procedures inherent to the Acheulian and that it is characterized by the slow replacement of bifacial technology by hierarchical flaking strategies, of which Levallois is arguably the most important method. Prehistorians increasingly accept Levallois technology as an innate property of the Acheulian that evolves from a combination of the façonnage and débitage systems discussed above (6–8, 12, 36, 86, 173, 174). A cogent argument has been made that similarities between Acheulian bifaces (Mode 2) and Levallois cores (Modes 3) are due to a common technological ancestry, in other words they are phylogenetically homologous rather than homoplasic (36). This ancestor-descendant relationship is demonstrated through formal cladistic analyses and highlights the lack of Levallois technology in regions where Acheulian traditions were not well established, for example in East Asia (36, 175). Figure 1 highlights this scenario, with the vast majority of Levallois assemblages in Eurasia overlapping in space with the ancestral Acheulian dispersal (based on: Africa, 176–178; Asia, 175, 179–183; Europe, 184), and the apparent lack of well-dated pre-OIS 6 Levallois assemblages east of 50°E (36, 185). Were the single origin and dispersal hypothesis correct, one would expect to find a spatio-temporal pattern in which the initial spread of Levallois technology tracked south to north, that is, from Africa into Eurasia, along with a truncated archaeological record where previously widespread technological behaviors (e.g., bifacial technology) were abruptly replaced by new ones developed and transported by incoming population (e.g., Levallois). Finally, one would not expect to find transitional industries such as the one

documented at NG1. None of these expectations are born out in Figure 1. Data on specific sites included in this analysis were culled only from published sources, and these can be found in Table S6. Sites selected for this analysis conform to three criteria, namely that the age of the site is known, the archaeological material came from a secure stratigraphic context, and the archaeological material was analyzed using modern protocols. For example, only sites associated with coherent and consistent absolute dates and/or careful chronostratigraphic study were used. Archaeological materials not recovered from secure stratigraphic contexts, with complex, poorly understood taphonomic records, or excavated using non-standard techniques were omitted. Archaeological assemblages not analyzed following modern protocols or not conforming to Boëda’s (14, 15) definition of Levallois technology were also excluded. Finally, previously published sites often experience reanalysis or more detailed analysis with the result being that previous claims as to antiquity, context, and technological affiliation are amended. This natural aspect of archaeological practice led to the omission from our analysis of sites that do not satisfy our three selection criteria but that appeared in previously published tables similar to our Table S6 (e.g., 7). A final consideration is terminology. The term “proto-Levallois” is typically applied to late OIS 9 and earlier assemblages (173), especially those within Western Europe where cores exhibit a volumetric core concept but either lack or exhibit minimal evidence for the predetermination of end products, one of the defining characteristics of Levallois technology (14, 15). For the purposes of Figure 1 and Table S6, we included early sites that meet the first two criteria outlined above but for which the technological diagnosis is inconclusive, thus the association in Table S6 of the term “proto-Levallois” with certain OIS 9 and earlier assemblages. Recognition of such “proto-Levallois” assemblages would appear to support the hypothesis that Levallois core reduction methods evolved directly (perhaps episodically) out of Acheulian biface production methods (e.g., 186–188). Examples of Acheulian bifaces transformed into Levallois cores are not uncommon in Late Acheulian contexts (126, 187, 189–191). However, these “proto-Levallois” assemblages, the presumed “ancestors” of later Levallois, have not been subjected to a cladistic analysis that could determine whether they are phylogenetically homologous with Levallois cores or simply an example of technological homoplasy. A case in point is the Victoria West LP industry from South Africa, the cores of which are argued to resemble Levallois cores and are termed “proto-Levallois”. These cores have been subjected to 3D geometric morphometric and cladistic analyses and, contrary to the expectations of the “proto-Levallois” hypothesis, have been shown to be the product of convergent technological evolution (188). In other words, an ancestor-descendant relationship could not be discerned between Victoria West and Levallois cores. Therefore, it has been suggested (192) that the term “proto-Levallois” be abandoned in favor of “para-Levallois” as a means of highlighting the broad morphological similarity of Victoria West cores to Levallois cores while also recognizing the distinct characteristics of the former. Until relevant material from Eurasian sites is subjected to similar analyses, the relationship between early “proto-Levallois” and later Levallois assemblages will remain unresolved. As demonstrated in Figure 1, outlined in Table S6, and argued throughout this paper, the episodic appearance (and disappearance) of Levallois technology throughout Africa and Eurasia

during the LMP suggests that inter-regional technological homoplasy, based on a shared Acheulian ancestry, is the crucial factor underwriting the eventual transition to the MP. Figure 1 demonstrates that transitional assemblages with strong links to the preceding Acheulian and a fluctuating spatio-temporal pattern of technological evolution are the norm. Additional Acknowledgements and Notes The authors wish to thank our many Armenian friends and colleagues for their support of this research, in particular Dr. P. Avetisyan (Director, Institute of Archaeology and Ethnography), Dr. B. Yeritsyan (Institute of Archaeology and Ethnography), Dr. A. Simonyan (Rector, Yerevan State University), Dr. A. Martkarov (Yerevan State University), Dr. H. Avetisyan (Yerevan State University), S. Kessedjian, H. Partevyan, E. Partevyan, and R. Partevyan. D.S.A. recognizes the generous logistical support provided by UConn’s Norian Armenian Studies Committee and Office of Study Abroad. Author responsibilities: D.S.A., director, lithic analysis, primary author; K.N.W., co-director, geoarchaeology, primary author; S.B., V.C., V.S., and A.M, tephrochronology; D.M., Ar40/Ar39 dating; R.P., co-director; B.A.S.-M., lithic analysis, data management; S.N., geomorphology; C.M., micromorphology; F.B., mineralogy; O.J., co-author; B.G., co-director; P.J.G. and Y.R-H, excavation, geology, lithic illustration; E.F., p-XRF analysis, obsidian sourcing; N.W., excavation.

Supplementary Figures S1–S16

Fig. S1. Detail of the central Hrazdan Gorge showing the location of NG1, the Paleolithic sites of Lusakert Cave, Fantan, Nurnus 1, and Jraber, and the position of volcanoes and volcanic vents.

Fig. S2. A view of NG1 from the east side of the Hrazdan Gorge illustrating the location of the site with respect to important geomorphological features. Excavations were conducted in the central and northern areas of the exposed stratigraphic section. The horizontal black line indicating the LMP floodplain exposure at NG1 runs approximately 135 meters.

Fig. S3. The central area of excavation at NG1 looking north with Basalt 1 and Units 5–1. Please reference Fig. S5 for the stratigraphic correlation.

Fig. S4. The northern area of excavation illustrating the locations of Basalts 1 and 7, the 40Ar/39Ar-dated tephra sample from Unit 1, and Units 5–1. Please reference Fig. 2 and Fig. S5 for the stratigraphic correlation.

Fig. S5. The northern end of the total NG1 stratigraphic section (top) with detail of the northern (middle, see Fig. 2) and central (bottom) areas of excavation, and the locations of all lithic artifacts, stratigraphic units, and samples. The extraction locations of the dated 40Ar/39Ar samples from Basalt 1 and Basalt 7 are indicated in the main profile (Basalt 1 and 7) and the northern section (Basalt 1) figures. Please reference Figs. S2–S4 for the photographic correlation.

Fig. S6. Micromorphology sample MM1 illustrating mottles in vughy Unit 1 sediment. Plane polarized light. See Fig. S5 for sample location, and Table S2 for description.

Fig. S7. Micromorphology sample MM2. A. Loose, bioturbated microstructure in Unit 2. Plane polarized light. B. Obsidian flakes in Unit 2. Plane polarized light. See Fig. S5 for sample location, and Table S2 for description.

Fig. S8. Age spectra and isotope correlation plots displaying inverse isochrons for Basalt 1 and Basalt 7 samples from NG1. See Fig. S5 for the extraction locations of both samples.

Fig. S9. Ideogram and isotope correlation plots displaying inverse isochrons for single sanidine crystals from the cryptotephra within Unit 1. See Fig. 2 for the extraction location of this sample.

Fig. S10. Refitting artifacts from NG1. RG1 (Unit 1): 1, bidirectional blade; 2–3, flakes. RG2 (Unit 2): 4, retouched proximal flake fragment; 5, retouched distal flake fragment. Artifacts 1–2 and 4–5 excavated in situ; artifact 3 recovered from surface. The three-dimensional coordinates (x, y, z) of artifacts 1 and 2 indicate they are separated in space by 21 cm, 10 cm, and 0 cm, respectively. Artifacts 4 and 5 are separated in space by 11.3 cm, 1.2 cm, and 0 cm, respectively.

Fig. S11. Cores and core management pieces from NG1: 1, recurrent unidirectional blade core; 2, bidirectional blade core on flake; 3, unidirectional core on Kombewa flake; 4, unidirectional blade core; 5, core trimming; 6, 7, 9, overpass flakes possibly from preferential Levallois cores; 8, débordant.

Fig. S12. Bifacial material from NG1: 1–6, bifaces and biface fragments; 7–9, biface-thinning flakes.

Fig. S13. Retouched tools from NG1: 1, 3, déjeté scrapers with Quina retouch; 2, 12, transverse scrapers; 4, raclette; 5, convergent scraper; 6, 8, 11, truncated-faceted pieces; 7, 9, déjeté scrapers; 10, truncated-faceted piece/end scraper.

Fig. S14. Retouched tools from NG1: 1–3, 5, 7, denticulates; 4, heavily recycled piece; 6, retouched Kombewa flake; 8, retouched flake; 9–13, resharpening flakes.

Fig. S15. This scatterplot of Rb, Sr, and Zr concentrations establishes the four volcanic sources of obsidian represented at NG1: Gutanasar, Pokr Arteni, Hatis, and Pokr Sevkar (see Fig. S16 for the locations of these Armenian volcanoes).

Fig. S16. The locations of NG1 (square) and known Armenian obsidian sources (circles), including those four represented at NG1: Gutanasar, Pokr Arteni, Hatis, and Pokr Sevkar. The circles for each source represent the volcanic center, approximate central location, or “type site” for which a source is named. No effort is made here to precisely represent the full primary and secondary distribution of these obsidians. Digital elevation data from SRTM3 (Shuttle Radar Topography Mission dataset version 3), and base map modified under Creative Commons terms from Wikimedia Commons.

Supplementary Tables S1–S7 Table S1. The NG1 lithostratigraphy, loss-on-ignition measurements and interpretation. Cycle Unit Top* (m) Base (m) Field description Grain size

(%Sa:Si:Cl)† Loss-on-ignition† Interpretation

1 1 0.00 0.10

Olive gray (Munsell 5 Y 5/2 [moist]) fine sandy silt/clay with rare sub-angular and sub-rounded obsidian granules scattered throughout. Particles are 'ashy' and bonded as pebble-sized, vertically orientated, prismatic blocks. Moderately sorted. Sharp boundary to:

31:30:39 (1) 3.9-4.3% (5) Floodplain alluvium

2

2 0.10 0.45

Black (10 YR 2/1) slightly humic silt with occasional medium-coarse sand-sized pumice fragments. Single wavy, horizontal, fine bed of pink (7.5 YR 7/3) tephra present at 90mm below the surface. Fine-grained matrix present as pebble-sized, blocky colloids. Moderate quantities of carbonate filaments originating from Unit 1. Diffuse boundary to:

29:38:33 (3) 4.5-8.0% (16)

A horizon in floodplain alluvium

3 0.45 0.60

Dark brown (7.5 YR 3/3) silty clay containing rare sub-angular fine pebble-sized obsidian as a single band 70mm from the unit surface. Vestiges of fine, horizontal layering towards the base. Moderately sorted. Diffuse boundary to:

22:41:37 (2) 4.7-6.3% (7) B horizon in floodplain alluvium

4 0.60 1.04

Yellowish brown (10 YR 5/4) normally bedded fine sand with rare sub-angular fine pebble-sized (waste flakes). Carbonate present as coarse, parallel, straight, continuous and discontinuous fine layers in the top 0.15m of the unit. Moderately sorted. Diffuse boundary to:

44:30:26 (3) 3.5-6.2% (16)

Floodplain alluvium/levee (pedogenically

modified)

5 1.04 1.26 Yellowish brown (10 YR 5/4) normally bedded fine sand with frequent boulder-cobble-sized sub-angular basalt clasts. Poorly sorted. Sharp boundary to Basalt 7.

33:40:27 (3) 4.1-4.9% (7) Levee/channel

* Measured from the contact between Basalt 1 and Unit 1 † Figures in parenthesis are the number of samples measured

Table S2. General micromorphological description of the NG1 stratigraphic sequence1. See Fig. S5 for sample locations.

Thin Sections Unit Lithology*† Porosity Microstructure B-fabric Postdepositional

features Lithics

(obsidian) MM1, MM6, MM9

1

Basaltic pyroclastic sand: Frequent s, vfs, ms and

cs and few gr. Subangular quartz: frequent s and vfs.

Frequent small vughs, few channels

Vughy/fine granular

Stipple-speckled

Few mottles, Few birefringent limpid clay infillings, Few micritic/needle-fibre

calcite infillings

MM2 2

Frequent small vughs, common channels, few

planes

Fine granular Stipple-speckled 16

MM3 3 Few planes, rare very small vughs Massive Granostriated 4

MM4 4 Massive Granostriated 1

MM5 5

Basaltic pyroclastic sand: Common s, vfs, and fs

and few cs. Subangular quartz: frequent s and vfs.

Frequent medium-sized

channels and few planes

Massive Granostriated Few mottles

* Frequent >20%, Commom 10–15%, Few 5–10%, Rare <5% † Silt (s), very fine sand (vfs), fine sand (fs), medium sand (ms), coarse sand (cs), gravel (gr).

Table S3. NG1sedimentological samples, context, description, and mineralogical composition inferred by FTIR spectroscopy. See Fig. S5 for sample locations.

Sample Description Mineral phases FTIR absorptions*

Sample Unit type color texture other features (decreasing intensity) 1 1 Sediment Gray sand Cl(Sm/I,K),Qz,Fel, 1a 1 Rock fragment Dark gray Qz 2 1 Sediment Gray sand Cl(Sm/I,K?),Qz,Fel, 3 1 Sediment Gray sand with white concretion Cl(Sm/I,K?),Qz,Fel, 4 2 Sediment Gray sand with white laminae Cl(Sm/I,K?),Qz,Fel, 5 2 Sediment Gray sand with white concretion Cl(Sm/I,K?),Qz,Fel, 5a 2 Rock fragment Dark gray Cl(Sm/I,K),Fel,Qz,CO3 6 2 Sediment Gray sand with white concretions Cl(Sm/I,K?),Qz,Fel 7 2 Sediment Gray sand with white concretions Cl(Sm/I,K?),Qz,Fel, 7a 2 Concretion/lamina white C 8 2 Sediment Gray sand with white concretions Cl(Sm/I,K?),Qz,Fel 9 2 Sediment Gray sand Cl(Sm/I,K?),Qz,Fel 10 2 Sediment Gray sand Cl(Sm/I,K?),Qz,Fel 11 2 Sediment Gray sand with white concretion Cl(Sm/I,K?),Qz,Fel 12 2 Sediment Brownish gray sandy-silt with white concretion Cl(Sm/I,K),Qz,Fel 13 2 Sediment Brownish gray sandy-silt with white concretion Cl(Sm/I,K),Qz,Fel 14 2 Sediment Brownish gray sandy-silt with white concretion Cl(Sm/I,K),Qz,Fel 15 2 Sediment Brownish gray silt Cl(Sm/I,K),Qz,Fel 16 2 Sediment Brownish gray silt Cl(Sm/I,K),Qz,Fel 17 2 Sediment Brownish gray silt Cl(Sm/I,K),Qz,Fel 18 2 Sediment Brownish gray silt Cl(Sm/I,K),Qz,Fel 19 2 Sediment Brownish gray silt Cl(Sm/I,K),Qz,Fel 20 2 Sediment Brownish gray silt Cl(Sm/I,K),Qz,Fel 21 2 Sediment Brownish gray silt with white concretion Cl(Sm/I,K),Qz,Fel 21a 2 Concretion/lamina off-white C,Cl(Sm/I,K) 22 2 Sediment Brown sandy-silt Cl(Sm/I,K?),Qz,Fel 23 3 Sediment Brown sandy-silt Cl(Sm/I,K?),Qz,Fel

23a 3 Rock fragment Dark gray Pl,K-Fel,Cl(Sm/I,K?) 24 3 Sediment Brown sandy-silt Cl(Sm/I,K?),Qz,Fel 25 3 Sediment Brown sandy-silt Cl(Sm/I,K?),Qz,Fel 26 3 Sediment Brown sandy-silt Cl (Sm/I,K?),Qz,Fel 27 3 Sediment Light Brown sandy-silt Cl (Sm/I,K),Qz,Fel 28 3 Sediment Light Brown sandy-silt Cl (Sm/I,K),Qz,Fel 29 3 Sediment Light Brown sandy-silt Cl (Sm/I,K),Qz,Fel 30 3 Sediment Light Brown sandy-silt Cl (Sm/I,K),Qz,Fel 31 3 Sediment Light Brown sandy-silt Cl (Sm/I,K),Qz,Fel 32 3 Sediment Light Brown sandy-silt w/ dark gray gravel Cl(Sm/I,K),Qz,Fel 32a 3 Flake Dark gray Op 33 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 34 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 35 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 36 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 37 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 38 4 Sediment Light Brown sand w/ black and white particles Td?Op?Cl(Sm/I),Fel 39 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 40 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 41 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 42 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 43 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 44 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 45 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 46 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 47 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 48 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 49 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 50 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 51 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 52 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel 53 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz 54 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Qz,Fel

55 4 Sediment Light Brown sandy-silt Cl(Sm/I,K), Qz,Fel 56 4 Sediment Light Brown sandy-silt Cl (Sm/I,K),Qz,Fel 57 4 Sediment Light Brown sandy-silt Cl(Sm/I,K),Fel,Qz

* C=calcite; Cl=clay minerals; CO3 = carbonate; Fel=Feldspar; K=kaolinite; K-Fel=K-Feldspar; Op=opal; Qz=quartz; Sm/I=smectite/illite; Td=tridymite. ?=some FTIR band position incongruities with respect to standard reference spectra.

Table S4: Preparation conditions ensuring the highest return of grains for dating. Source Grain size Proximity to HF

surface Duration of HF vapour exposure

Duration of staining

56 >125 µm Not defined c. 5 minutes 2–3 minutes

This article 80–125 µm & >125 µm 3 cm 2 minutes 2 minutes

Table S5. Techno-typological and fragmentation summary of the lithic artifacts recovered from NG1.

CLASS Unit 1 Unit 2 Unit 3 Unit 4 Slope* Total %Total-1 (N=2976)

%Total-2 (n=670)†

CORE (n=38) 4 14 0 0 20 38 1.3 5.7 Levallois 0 3 0 0 14 17 44.7

Hierarchical 2 3 0 0 5 10 26.3 Blade 2 1 0 0 0 3 7.9

Core-on-Flake 0 5 0 0 1 6 15.8 Kombewa 0 2 0 0 0 2 5.3

HAMMERSTONE (n=1) 0 0 0 0 1 1 0.03 0.1 TOOL (n=65)‡ 9 24 4 2 26 65 2.2 9.7

Biface 2 2 0 0 3 7 10.8 Single Scraper 0 6 2 0 7 15 23.1

Double Scraper 0 1 0 0 2 3 4.6 Convergent Scraper 0 0 0 0 2 2 3.1

Déjeté Scraper 0 2 1 0 0 3 4.6 Transverse Scraper 1 2 0 0 5 8 12.3

Alternate Scraper 1 0 0 0 0 1 1.5 End Scraper 1 1 0 0 1 3 4.6 Denticulate 2 4 1 1 2 10 15.4

Point 0 0 0 0 1 1 1.5 Raclette 0 1 0 0 0 1 1.5

Notch 0 1 0 0 2 3 4.6 Truncated-facetted 1 1 0 0 1 3 4.6

Retouched Piece 0 3 0 0 0 3 4.6 Burin/Spall 1 0 0 1 0 2 3.1

Technology: Levallois 4 7 0 1 4 16 24.6 Hierarchical 3 13 4 1 18 39 60.0

Kombewa 0 2 0 0 1 3 4.6 Bifacial 2 2 0 0 3 7 10.8

DEBITAGE (n=567)§ 76 248 57 13 173 567 19.0 84.6 Blade 6 11 5 2 3 27 0.9 4.8

Complete 1 4 3 0 3 11 40.7 Distal 0 3 0 1 0 4 14.8

Medial 1 1 0 1 0 3 11.1 Proximal 4 3 2 0 0 9 33.3

Technology: Levallois 1 5 3 2 3 14 51.9 Hierarchical 5 6 2 0 0 13 48.1

Flake (n=458) 62 202 46 11 137 458 15.4 80.8 Complete 17 74 19 3 70 183 40.0

Distal 14 49 8 4 31 106 23.1 Medial 16 45 14 2 16 93 20.3

Proximal 15 34 5 2 20 76 16.6 Technology: Levallois 12 61 9 1 29 112 24.5

Hierarchical 50 141 37 10 108 346 75.5 Débordant (n=27) 4 11 0 0 12 27 0.9 4.8

Complete 3 9 0 0 6 18 66.7 Distal 1 0 0 0 0 1 3.7

Medial 0 0 0 0 1 1 3.7 Proximal 0 2 0 0 5 7 25.9

Technology: Levallois 1 2 0 0 3 6 22.2 Hierarchical 3 9 0 0 9 21 77.8

Overshot (n=29) 3 14 4 0 8 29 1.0 5.1 Complete 1 8 3 0 7 19 65.5

Distal 2 6 1 0 1 10 34.5 Medial 0 0 0 0 0 0 0.0

Proximal 0 0 0 0 0 0 0.0 Technology: Levallois 0 1 0 0 4 5 17.8

Hierarchical 3 13 4 0 4 24 82.8 Kombewa (n=9) 1 3 1 0 4 9 0.3 1.6

Complete 0 0 0 0 3 3 33.3 Distal 1 2 1 0 1 5 55.6

Medial 0 1 0 0 0 1 11.1 Proximal 0 0 0 0 0 0 0.0

Resharpening (n=17) 0 7 1 0 9 17 0.6 3.0 Complete 0 5 1 0 8 14 82.4

Distal 0 0 0 0 0 0 0.0 Medial 0 2 0 0 1 3 17.6

Proximal 0 0 0 0 0 0 0.0 Type: Tool Retouch 0 3 0 0 4 7 41.2

Biface Thinning 0 4 1 0 5 10 58.8 Debitage Technology§ Unit 1 Unit 2 Unit 3 Unit 4 Slope a Total %Total

Levallois 14 68 12 3 39 136 24.0 Hierarchical 61 196 43 10 121 431 76.0

Total 75 264 55 13 160 567 100.0

Shatter (n=117) 10 86 18 1 2 117 3.9 <2.5CM (n=2189) 55 1186 924 8 16 2189 73.6

piece plotted 55 794 156 5 16 1026 46.9 bucket|| 0 392 768 3 0 1163 53.1

Unit 1 Unit 2 Unit 3 Unit 4 Slope a Total %Total TOTAL (N=2976) 154 1558 1003 24 237 2976 100.0

* Slope/Surface artifacts were recovered from immediately below the exposed stratigraphic section and the adjacent slope. † %Total-2 calculated based on all artifacts >2.5cm (n=670), excluding shatter (n=117). ‡ Quina scrapers were not recognized as a separate typological group, but Quina retouch was identified on six specimens (e.g., Fig. S13). § Debitage assigned to "Hierarchical" is the default category and likely includes many broken Levallois pieces for which a more precise determination was problematic. || 55.4% and 53% of buckets from Units 2 and 3, respectively, have been washed, picked, sorted, counted and weighted.

Table S6. Comparative data for NG1 and important sites spanning the LP–MP and ESA–MSA transition in Eurasia and Africa, respectively. Please see Fig. 1 for site locations (based on date from 4–10, 13, 22, 70–154). Map Ref. SITE (Country) Levallois Biface Classification Age [OIS] Age [ka] dating method, primary Bib.

Ref. dating method, secondary

NG1 Nor Geghi 1 (AM) + + Late Acheulian OIS 9, early (9e) ~335–325 Ar/Ar this

work geology/(bio)stratigraphy EUROPE

pre-OIS 9

EU-1 St. Acheul, Rue Marcellin Betholot (FR)

rare + Acheulian OIS 14?

geology/(bio)stratigraphy 70, 71 Fréille terrace level

EU-2 Cagny-la-Garenne (FR)

Proto + Acheulian OIS 12?

geology/(bio)stratigraphy 70, 71 Cagny terrace level

EU-3 Swanscombe (GB) single

specimen + Acheulian OIS 11

geology/(bio)stratigraphy 72, 73 Rickson's Pit

EU-4 Atapuerca (ES)

rare + Acheulian OIS 10, late

~345±26 wm (4 ESR / U-series)

74–76 Gran Dolina, level TD10.1 (transition OIS 10–9?) geology/(bio)stratigraphy

EUROPE

OIS 9

EU-5 Ambrona (ES)

+ + Acheulian OIS 9, early (9e) ~336±36 wm (2 ESR / U-series)

77 upper and lower members geology/(bio)stratigraphy

EU-6 Petit Bost (FR)

+ + Acheulian OIS 9, early (9e) ~325±30 wm (2 TL)

78 level 2 geology/(bio)stratigraphy

EU-7 Domeny (ES) + + Acheulian OIS 9, early (9e–c) >317±49 Ar/Ar

79 geology/(bio)stratigraphy

EU-8 Cagny l'Epinette (FR) Proto + Acheulian OIS 9, early? (9e?)

geology/(bio)stratigraphy 80

EU-9 Aridos 1 (ES) + + Acheulian OIS 9, early? (9e?)

geology/(bio)stratigraphy 81

EU-10 Orgnac 3 (FR)

Proto + Acheulian OIS 9, early? (9e?) >302.9 Ar/Ar / U-Th

82 level 5b geology/(bio)stratigraphy

EU-11 La Micoque (FR)

Proto + Late Acheulian OIS 9 ~350–288 ESR / U-series

7, 83 level L 2/3 geology/(bio)stratigraphy

EU-12 Gentelles base (FR) + + Late Acheulian OIS 9

geology/(bio)stratigraphy 8

EU-13 Torre in Pietra (IT)

+ + Late Acheulian OIS 9?

geology/(bio)stratigraphy 84 layer "m"

EU-14 Torralba (ES) + + Late Acheulian OIS 9? >243±18 U-Th / U-series

85 geology/(bio)stratigraphy

EU-15 Cave dall'Olio (IT) + + Late Acheulian OIS 9?

geology/(bio)stratigraphy 8

EU-16 Purfleet (GB) Proto + Early MP

OIS 9, late (9c–a) ~324 TL

7, 86 (Late Acheulian) geology/(bio)stratigraphy

EU-17 Puig d'en Roca III (ES) + + Late Acheulian OIS 9, late (9c–a) <317±49 Ar/Ar

79 geology/(bio)stratigraphy

EU-18 Achenheim (FR)

+ no Early MP OIS 9, late (9c/a) ~258±23 geology/(bio)stratigraphy

87, 88 layer 20a wm (2 TL)

EU-19 Solent River (GB)

Proto + Late Acheulian OIS 9, late (~9b)

geology/(bio)stratigraphy 89 river terraces

EU-10 Orgnac 3 (FR)

+ + Late Acheulian OIS 9, late >302.9 Ar/Ar / U-Th

82 levels 4–3 geology/(bio)stratigraphy

EUROPE

OIS 8

EU-20 Mesvin IV (BE) + + Late Acheulian transition OIS 9–8 ~283±30 wm (2 U-Th)

90 geology/(bio)stratigraphy

EU-21 Les Bosses (FR)

+ + Late Acheulian transition OIS 9–8 ~274±12 wm (6 TL)

91 level 1 geology/(bio)stratigraphy

EU-10 Orgnac 3 (FR)

+ + Late Acheulian OIS 8, early <302.9 Ar/Ar / U-Th

82 levels 2–1 geology/(bio)stratigraphy

EU-22 Kesselt - Op de Schanz (BE) + no Early MP OIS 8, early

geology/(bio)stratigraphy 92

EU-23 Markkleeberg (DE)

+ + Late Acheulian OIS 8, early

geology/(bio)stratigraphy 93 assemblage FC-1–FC-2

EU-18 Achenheim (FR)

+ no Early MP OIS 8, early

geology/(bio)stratigraphy 87, 88 layers 20''', 20'', 20'

EU-24 Rheindahlen (DE)

+ no Early MP OIS 8?, early

geology/(bio)stratigraphy 94 level B5

EU-25 Raspide 2 (FR) + + Late Acheulian OIS 8

geology/(bio)stratigraphy 95

EU-26 Argoeuves (FR) + + Late Acheulian OIS 8 geology/(bio)stratigraphy 96

Lower terrace of the Somme

EU-11 La Micoque (FR)

Proto + Late Acheulian OIS 8 >241–288 ESR / U-series

7, 83 levels 3-4 geology/(bio)stratigraphy

EU-27 Baume Bonne (FR)

+ + Late Acheulian OIS 8, late

geology/(bio)stratigraphy 97, 98 unit III

EU-18 Achenheim (FR)

+ no Early MP OIS 8, late

geology/(bio)stratigraphy 87, 88 layers 20, 19, 18

EUROPE

OIS 7

EU-28 Thames Valley Sites (GB)

+ + Late Acheulian transition OIS 8–7

geology/(bio)stratigraphy 99 Terraces

EU-29 Le Pucheuil (FR)

+ + Late Acheulian transition OIS 8–7

geology/(bio)stratigraphy 100 assemblage C

EU-30 Galeria Pesada (PT)

+ + Late Acheulian OIS 7, early (7e) 241±22 ESR / U-series

101 level B2 geology/(bio)stratigraphy

EU-4 Atapuerca (ES)

+ + Early MP OIS 7, early (7e) 240±44 TL / IRSL

76, 102 Gran Dolina, level TD11 geology/(bio)stratigraphy

EU-31 Bonneval (FR) + +? Late Acheulian OIS 7, early (7e) 240

103

EU-32 Barbas (FR)

+ + Late Acheulian OIS 7, early (7e) 239±44 TL

104 level 4 geology/(bio)stratigraphy

EU-33 La Cotte Saint-Brelade (GB)

+ no Early MP OIS 7, early (7e) 238±35 TL

105 units C-D geology/(bio)stratigraphy

EU-34 Maastricht-Belvédère (NL) + no Early MP OIS 7, early (7e) ~258±19 wm (1 TL / 1 ESR)

106 geology/(bio)stratigraphy

EU-35 Biache - Saint-Vaast (FR)

+ no Early MP OIS 7, early? (7e?) ~230±18 wm (1 ESR / 2 ESR / U-series / 1 TL) 107–

109 level IIa geology/(bio)stratigraphy

EU-24 Rheindahlen (DE)

+ no Early MP OIS 7, early? (7e?)

geology/(bio)stratigraphy 94 level B3

EU-36 Cantalouette I (FR)

+ no Early MP OIS 7, early (7d?) 223±20 TL

110 unit 4 geology/(bio)stratigraphy

EU-37 Korolevo (UA)

+ no Early MP OIS 7, early (7d?) 220±35 OSL

111 level Vb geology/(bio)stratigraphy

EU-38 Weimar-Ehringsdorf (DE) + + Early MP OIS 7 ~230 (find layer) U-Th / U-series 112,

243±6.2 (lower travertine) geology/(bio)stratigraphy 113

EU-39 Le Rissori (BE)

+ no? Early MP OIS 7

geology/(bio)stratigraphy 114 Level Iia

EU-40 Bečov I (CZ)

+ no? Early MP OIS 7

geology/(bio)stratigraphy 115 level A-III-6

EU-41 Hundisburg (DE) + + Late Acheulian OIS 7

geology/(bio)stratigraphy 116

EU-42 Nové Mesto nad Váhom (SK) + no Early MP OIS 7

geology/(bio)stratigraphy 117

EU-43 Salouel (FR) + +? Early MP OIS 7 >200±57 ESR / U-series

118 geology/(bio)stratigraphy

EU-44 Biśnik Cave (PL)

+ no Early MP OIS 7 230±51 (TL Flint) / 204±30

(wm 2 TL Sed.) TL / U-Th 119

layer 18 125±10 (U/Th) geology/(bio)stratigraphy

EU-44 Biśnik Cave (PL)

+ no Early MP OIS 7, late (7c–a?) 279±97 (TL) TL / U-Th

119 layer 15 216±25 (U/Th) geology/(bio)stratigraphy

EU-45 Abri Vaufrey (FR)

+ no? Early MP OIS 7, late (7c–a) 208±8 U-series

120 level IX geology/(bio)stratigraphy

EU-46 Campsas (FR) + +?

OIS 7, late (7b)

geology/(bio)stratigraphy 121

EU-47 Gran Rois (FR) + +

OIS 7, late (7a)

geology/(bio)stratigraphy 9

EU-48 Racibórz Studienna 2 (PL) + no Early MP OIS 7, late – OIS 6, early

geology/(bio)stratigraphy 117

EU-29 Le Pucheuil (FR)

+ no Early MP OIS 7, late – OIS 6, early

geology/(bio)stratigraphy 100

assemblage B

EU-49 Bapaume Les Osiers (FR) + + OIS 7, late – OIS 6,

early ~195 IRSL

122 geology/(bio)stratigraphy

EU-50 San Bernardino (IT)

+ no Early MP transition OIS 7–6 ~184±6 wm (8 ESR)

9 level VIII-VII geology/(bio)stratigraphy

EU-51 Therdonne (FR) + no Early MP transition OIS 7–6 178±11 TL

123 geology/(bio)stratigraphy

EU-52 Dzierzyslaw 1 (PL) + no Early MP transition OIS 7–6

geology/(bio)stratigraphy 117

ASIA

pre-OIS 7

AS-1 Revadim (IL) rare + Acheulian OIS 13–9? > ~500–300? U-Th (carbonate-coated lithics) 124–

126 geology/(bio)stratigraphy

AS-2 Hummal (SY)

no + AY OIS 11e –

422±55 until 243±40 TL 127,

128 El Kowm Basin transition OIS 8–7 geology/(bio)stratigraphy

AS-3 Qesem (IL)

no + AY OIS 11, early (11e)? ~397±20 wm (1 TL / 1 ESR) 10,129,

130 first occupation based on oldest TL / ESR geology/(bio)stratigraphy

AS-4 Tabun (IL)

rare + AY OIS 11, late – OIS 8 ~390±50 until 267±22 ESR / U-series + TL 131–

133 E (Units XI-X) geology/(bio)stratigraphy

AS-5 Azych (GE)

rare + Late Acheulian OIS 10 360–330 geology/(bio)stratigraphy 13 Layers VI-V

AS-3 Qesem (IL)

no + AY OIS 9–7 326+20/-18 (TL) U-series / TL / ESR 10,129,

130 based on dates of Low. + Upp. sequence until ~218±15 (TL/U-series) geology/(bio)stratigraphy ASIA

OIS 7

AS-4 Tabun (IL)

+ no Early MP (Tabun-D-type)

transition OIS 8–7 – OIS 7?

256±26 until 244±28 (TL) TL / ESR 132–135 D (Units IX-VII) ~200 (ESR) geology/(bio)stratigraphy

AS-6 Oumm Qoubeiba (SY)

no + AY transition OIS 8–7 <245±16 U/Th

136 El Kowm Basin geology/(bio)stratigraphy

AS-7 Misliya Cave (IL)

no + AY transition OIS 8–7 244±30 wm (9 TL)

22 Lower Terrace, sq. Q28, Q29 geology/(bio)stratigraphy

AS-8 Karain Cave E (TR)

+ no Early MP OIS 7 250–200 TL / ESR 137,

138 Layer F geology/(bio)stratigraphy

AS-9 Jamal Cave (IL) no + AY OIS 7 >227±35/-27 U/Th

139 geology/(bio)stratigraphy

AS-10 Yabrud I (SY)

no + AY OIS 7 ~225±11 wm (3 ESR-CU / TL)

136 levels 19-18 geology/(bio)stratigraphy

AS-11 Hayonim (IL)

+ no AY OIS 7 > ~220 TL / ESR

140 layer G geology/(bio)stratigraphy

AS-11 Hayonim (IL)

+ no Early MP (Tabun-D-type) OIS 7, late ~220

TL / ESR 140

layer F geology/(bio)stratigraphy

AS-12 Jebel Qattar JQ-1 (SA)

+ no Early MP OIS 7, late 211±16 OSL

141 lower geology/(bio)stratigraphy

AS-2 Hummal (SY) + no Hummalian OIS 7, late – OIS 6, ~220–150 TL 127,

El Kowm Basin (Tabun-D-like) early geology/(bio)stratigraphy 128

AS-7 Misliya Cave (IL)

+ no Early MP OIS 7, late – OIS 6, early 212±27 until 166±23

TL (21) 22

Upper Terrace, sq. L15, L10, J15, N12 geology / (bio)stratigraphy AFRICA

pre-OIS 9

AF-1 Kathu Pan 1 (ZA)

+ + Fauresmith OIS 13–12? ~472±45 wm (1 ESR / U-series / 1 OSL)

4 layer 4a geology/(bio)stratigraphy

AF-2 Kharga Oasis 10 (KO10) (EG) rare + Acheulian OIS 11–9? ~400–300 U-series 142, 143

AF-3 Grotte des Rhinoceros (MA) rare + Acheulian transition OIS 10–9 ~350–300 ESR 144, 145

AF-4 Cap Chatelier (MA) rare + Acheulian transition OIS 10–9 ~350–300 OSL 140, 145

AFRICA

OIS 9

AF-5 GnJh-03, GnJh-17 (KE)

rare + ESA–MSA

<OIS 13 <509±9 Ar/Ar

146 Kapthurin Formation K3 transition geology/(bio)stratigraphy

AF-6 Rorop Lingop (KE)

rare + Early MSA OIS 9? <509±9 Ar/Ar

5 Kapthurin Formation K4 Sangoan/Fauresmith (or older, but <OIS 13) >284±12 geology/(bio)stratigraphy

AF-7 Leakey Handaxe Site (LHA) (KE)

rare + Acheulian OIS 9? <509±9 Ar/Ar 146,

147 Kapthurin Formation K4 (or older, but <OIS 13) >284±12 geology/(bio)stratigraphy

AF-8 Factory Site (FS) (KE)

rare + Acheulian OIS 9? <509±9 Ar/Ar 146,

147 Kapthurin Formation K4 (or older, but <OIS 13) >284±12 geology/(bio)stratigraphy

AF-1 Kathu Pan 1 (ZA)

rare no Early MSA OIS 9, late (OIS 9a?) 291±45 OSL

4 layer 3 geology/(bio)stratigraphy

AF-9 Bundu Farm (ZA) prepared

cores single

specimen

terminal Acheulian OIS 9–7? ~340–190

ESR 5

layer G6 (ESA–MSA transition) geology/(bio)stratigraphy

AFRICA

OIS 8

AF-10 ETH72-8B (ET)

+ no? Early MSA transition OIS 9–8 >276±4 Ar/Ar

6 Gademotta Formation G3, Unit 9 geology/(bio)stratigraphy

AF-11 Kulkuletti (ET)

+ no? Early MSA transition OIS 9–8? ~280±8 Ar/Ar

5 Gademotta Formation G3, Unit 10 >183±10 geology/(bio)stratigraphy

AF-12 Florisbed (ZA)

+ +? Early MSA OIS 8, late ~268±26 wm (1 ESR / 2 OSL) 148,

149 oldest MSA layers geology/(bio)stratigraphy

AF-13 Sterkfontein (ZA) + +? Early MSA OIS 8, late 252±42 ESR

5 geology/(bio)stratigraphy

AF-14 Sterkfontein: Lincoln Cave (ZA)

+ +? Early MSA OIS 8, late ~252.6±35.6 U-series

5, 150 CoHK geology/(bio)stratigraphy

AF-15 Gademotta (ET)

+ no Early MSA OIS 8–7 <280±8 Ar/Ar

5 Gademotta Formation G3, Unit 11 >183±10 geology/(bio)stratigraphy

AFRICA

OIS 7

AF-16 Border Cave (ZA)

+ no Early MSA OIS 7, early ~238–217 ESR

5 6BS geology/(bio)stratigraphy

AF-17 Koimilot, Loc. 1 (KE)

rare no Early MSA OIS 7 ~200 (~200–250) tephrostratigraphic correlation 6, 147,

151 Kapthurin Formation geology/(bio)stratigraphy

AF-17 Koimilot, Loc. 2 (KE)

rare no Early MSA OIS 7 ~200 (~200–250) tephrostratigraphic correlation 6, 147,

151 Kapthurin Formation geology/(bio)stratigraphy

AF-18 Kharga Oasis, Site REF-4 (EG) + no Early MSA OIS 7 220±20 U-series 152

AF-19 Omo-Kibish, AHS (ET)

+ + Early MSA OIS 7, late

6 Kibish Formation, Member I

AF-20 Omo-Kibish, KHS (ET)

+ + Early MSA OIS 7, late

6 Kibish Formation, Member I

AF-21 ETH72-7B, ETH72-1 (ET)

+ + Early MSA OIS 7, late – OIS 6, early ~195

6

Gademotta Formation

AF-22 Sai Island 8-B-11 (SD)

+ + MSA OIS 7, late – OIS 6,

early <223±19

OSL 153, 154 BLG/TLG/UG (Sangoan) >152±10

Abbreviations and notes: Numbers in the far left column refer to individual sites in Fig. 1, for example, Nor Geghi 1 (NG1); Europe (EU 1–52); Asia (AS 1–12); and Africa (AF 1–22). Country codes provided by the International Organization for Standardization (ISO); + indicates presence; Proto-Levallois (Proto); Early Stone Age (ESA); Middle Stone Age (MSA); Acheulo-Yabrudian (AY); Middle Paleolithic (MP); Oxygen Isotope Stage (OIS); weighted mean (wm), see references for raw data; age estimates (ka) in italics diverge significantly from the geochronological framework established for the site; Argon/Argon (Ar/Ar); Electron Spin Resonance (ESR); Infrared Stimulated Luminescence (IRSL); Optically Stimulated Luminescence (OSL); Uranium Series (U-Series); Uranium-Thorium (U-Th).

Table S7. Technological features used to classify different industries during the LP–MP and ESA–MSA transition in Eurasia and Africa, respectively.

Classification* Mode Biface Technology†

Levallois Technology

Non-Levallois Technology

Quina retouch Blades

MSA/MP 3 +/- dominant + +/- + AY 2/3 + - dominant + +

NG1 2/3 + + + + + ESA/LP 2 dominant (+) + - (+)

* Early Stone Age (ESA); Lower Paleolithic (LP); Nor Geghi 1 (NG1); Acheulo-Yabrudian (AY); Middle Stone Age (MSA); Middle Paleolithic (MP). † + present; - absent; +/- infrequent; (+) at some Late Acheulian sites only.

Database S1 (separate excel file): NG1 tephra glass compositions.

Database S2 (separate excel file): Experimental and isotopic data for NG1 Basalts 1 and 7 and the Unit 1 tephra.

References 1. I. McDougall, F. H. Brown, J. G. Fleagle, Stratigraphic placement and age of modern humans

from Kibish, Ethiopia. Nature 433, 733–736 (2005). Medline doi:10.1038/nature03258

2. J. J. Hublin, The origin of Neandertals. Proc. Natl. Acad. Sci. U.S.A. 106, 16022–16027 (2009). Medline doi:10.1073/pnas.0904119106

3. G. Clark, World Prehistory: A New Outline. (Cambridge Univ. Press, London, 1969).

4. N. Porat, M. Chazan, R. Grün, M. Aubert, V. Eisenmann, L. K. Horwitz, New radiometric ages for the Fauresmith industry from Kathu Pan, southern Africa: Implications for the Earlier to Middle Stone Age transition. J. Archaeol. Sci. 37, 269–283 (2010). doi:10.1016/j.jas.2009.09.038

5. A. I. R. Herries, A chronological perspective on the Acheulian and its transition to the Middle Stone Age in southern Africa: The question of the Fauresmith. Int. J. Evol. Biol. 2011, 1–25 (2011). doi:10.4061/2011/961401

6. C. A. Tryon, J. T. Faith, Variability in the Middle Stone Age of Eastern Africa. Curr. Anthropol. 54 (S8), S234–S254 (2013). doi:10.1086/673752

7. M. White, N. Ashton, Lower Palaeolithic core technology and the origins of the Levallois method in north‐western Europe. Curr. Anthropol. 44, 598–609 (2003). doi:10.1086/377653

8. F. Fontana, M.-H. Moncel, G. Nenzioni, G. Onorevoli, C. Peretto, J. Combier, Widespread diffusion of technical innovations around 300,000 years ago in Europe as a reflection of anthropological and social transformations? New comparative data from the western Mediterranean sites of Orgnac (France) and Cave dall’Olio (Italy). J. Anthropol. Archaeol. 32, 478–498 (2013). doi:10.1016/j.jaa.2013.08.003

9. A. Picin, M. Peresani, C. Falguères, G. Gruppioni, J.-J. Bahain, San Bernardino Cave (Italy) and the appearance of Levallois technology in Europe: Results of a radiometric and technological reassessment. PLOS ONE 8, e76182 (2013). Medline doi:10.1371/journal.pone.0076182

10. R. Barkai, A. Gopher, S. E. Lauritzen, A. Frumkin, Uranium series dates from Qesem Cave, Israel, and the end of the Lower Palaeolithic. Nature 423, 977–979 (2003). Medline doi:10.1038/nature01718

11. R. Shimelmitz, R. Barkai, A. Gopher, Systematic blade production at late Lower Paleolithic (400-200 kyr) Qesem Cave, Israel. J. Hum. Evol. 61, 458–479 (2011). Medline doi:10.1016/j.jhevol.2011.06.003

12. S. L. Kuhn, Roots of the Middle Paleolithic in Eurasia. Curr. Anthropol. 54, S255–S268 (2013). doi:10.1086/673529

13. V. B. Doronichev, PaleoAnthropol. 2008, 107 (2008).

14. E. Boëda, in The Definition and Interpretation of Levallois Technology, H. L. Dibble, O. Bar-Yosef, Eds. (Prehistory Press, Madison, WI, 1995), pp. 41–69.

15. E. Boëda, Le Concept Levallois: Variabilité des Méthodes (CNRS Éditions, Paris, 1994).

16. P. J. Brantingham, S. L. Kuhn, Constraints on Levallois core technology: A mathematical model. J. Archaeol. Sci. 28, 747–761 (2001). doi:10.1006/jasc.2000.0594

17. S. J. Lycett, M. I. Eren, Levallois economics: An examination of ‘waste’ production in experimentally produced Levallois reduction sequences. J. Archaeol. Sci. 40, 2384–2392 (2013). doi:10.1016/j.jas.2013.01.016

18. S. J. Lycett, M. I. Eren, Levallois lessons: The challenge of integrating mathematical models, quantitative experiments and the archaeological record. World Archaeol. 45, 519–538 (2013). doi:10.1080/00438243.2013.821670

19. M. I. Eren, S. J. Lycett, Why Levallois? A morphometric comparison of experimental ‘preferential’ Levallois flakes versus debitage flakes. PLOS ONE 7, e29273 (2012). Medline doi:10.1371/journal.pone.0029273

20. R. Foley, M. M. Lahr, Mode 3 technologies and the evolution of modern humans. Camb. Archaeol. J. 7, 3 (1997). doi:10.1017/S0959774300001451

21. S. J. Armitage, S. A. Jasim, A. E. Marks, A. G. Parker, V. I. Usik, H. P. Uerpmann, The southern route “out of Africa”: Evidence for an early expansion of modern humans into Arabia. Science 331, 453–456 (2011). Medline doi:10.1126/science.1199113

22. H. Valladas, N. Mercier, I. Hershkovitz, Y. Zaidner, A. Tsatskin, R. Yeshurun, L. Vialettes, J. L. Joron, J. L. Reyss, M. Weinstein-Evron, Dating the Lower to Middle Paleolithic transition in the Levant: A view from Misliya Cave, Mount Carmel, Israel. J. Hum. Evol. 65, 585–593 (2013). Medline doi:10.1016/j.jhevol.2013.07.005

23. M. Meyer, Q. Fu, A. Aximu-Petri, I. Glocke, B. Nickel, J. L. Arsuaga, I. Martínez, A. Gracia, J. M. de Castro, E. Carbonell, S. Pääbo, A mitochondrial genome sequence of a hominin from Sima de los Huesos. Nature 505, 403–406 (2014). Medline doi:10.1038/nature12788

24. K. Prüfer, F. Racimo, N. Patterson, F. Jay, S. Sankararaman, S. Sawyer, A. Heinze, G. Renaud, P. H. Sudmant, C. de Filippo, H. Li, S. Mallick, M. Dannemann, Q. Fu, M. Kircher, M. Kuhlwilm, M. Lachmann, M. Meyer, M. Ongyerth, M. Siebauer, C. Theunert, A. Tandon, P. Moorjani, J. Pickrell, J. C. Mullikin, S. H. Vohr, R. E. Green, I. Hellmann, P. L. Johnson, H. Blanche, H. Cann, J. O. Kitzman, J. Shendure, E. E. Eichler, E. S. Lein, T. E. Bakken, L. V. Golovanova, V. B. Doronichev, M. V. Shunkov, A. P. Derevianko, B. Viola, M. Slatkin, D. Reich, J. Kelso, S. Pääbo, The complete genome sequence of a Neanderthal from the Altai Mountains. Nature 505, 43–49 (2014). Medline doi:10.1038/nature12886

25. Materials and methods are available as supporting material on Science Online.

26. R. Badalian, G. Bigazzi, M.-C. Cauvin, C. Chataigner, R. Jrbashyan, S. G. Karapetyan, M. Oddone, J.-L. Poidevin, An international research project on Armenian archaeological sites: Fission-track dating of obsidians. Radiat. Meas. 34, 373–378 (2001). doi:10.1016/S1350-4487(01)00189-5

27. E. V. Arutyunyan, V. A. Lebedev, I. V. Chernyshev, A. K. Sagatelyan, Geochronology of Neogene-Quaternary volcanism of the Geghama Highland (Lesser Caucasus, Armenia). Dokl. Earth Sci. 416, 1042–1046 (2007). doi:10.1134/S1028334X07070136

28. V. A. Lebedev, I. V. Chernyshev, K. N. Shatagin, S. N. Bubnov, A. I. Yakushev, The quaternary volcanic rocks of the Geghama highland, Lesser Caucasus, Armenia: Geochronology, isotopic Sr-Nd characteristics, and origin. J. Volcanol. Seismol. 7, 204–229 (2013). doi:10.1134/S0742046313030044

29. S. Joannin, J.-J. Cornée, P. Münch, M. Fornari, I. Vasiliev, W. Krijgsman, S. Nahapetyan, I. Gabrielyan, V. Ollivier, P. Roiron, C. Chataigner, Early Pleistocene climate cycles in continental deposits of the Lesser Caucasus of Armenia inferred from palynology, magnetostratigraphy, and 40Ar/39Ar dating. Earth Planet. Sci. Lett. 291, 149–158 (2010). doi:10.1016/j.epsl.2010.01.007

30. V. Ollivier, S. Nahapetyan, P. Roiron, I. Gabrielyan, B. Gasparyan, C. Chataigner, S. Joannin, J.-J. Cornée, H. Guillou, S. Scaillet, P. Munch, W. Krijgsman, Quaternary volcano-lacustrine patterns and palaeobotanical data in southern Armenia. Quat. Int. 223-224, 312–326 (2010). doi:10.1016/j.quaint.2010.02.008

31. E. Frahm, B. A. Schmidt, B. Gasparyan, B. Yeritsyan, S. Karapetian, K. Meliksetian, D. S. Adler, Ten seconds in the field: Rapid Armenian obsidian sourcing with portable XRF to inform excavations and surveys. J. Archaeol. Sci. 41, 333–348 (2014). doi:10.1016/j.jas.2013.08.012

32. A. J. Jelinek, The Tabun Cave and Paleolithic Man in the Levant. Science 216, 1369–1375 (1982). Medline doi:10.1126/science.216.4553.1369

33. S. McBrearty, J. Archaeol. Res. 69, 7 (2013).

34. J. J. Shea, Homo sapiens Is as Homo sapiens Was. Curr. Anthropol. 52, 1–35 (2011). doi:10.1086/658067

35. T. Hopkinson, A. Nowell, M. White, PaleoAnthropology 2013, 61 (2013).

36. S. Lycett, Why is there a lack of Mode 3 Levallois technologies in East Asia? A phylogenetic test of the Movius–Schick hypothesis. J. Anthropol. Archaeol. 26, 541–575 (2007). doi:10.1016/j.jaa.2007.07.003

37. R. Stöckli, E. Vermote, N. Saleous, R. Simmon, D. Herring, The Blue Marble Next Generation - A true color earth dataset including seasonal dynamics from MODIS (NASA Earth Observatory, 2005); http://earthobservatory.nasa.gov/Features/BlueMarble/bmng.pdf.

38. L. Augustin, C. Barbante, P. R. Barnes, J. M. Barnola, M. Bigler, E. Castellano, O. Cattani, J. Chappellaz, D. Dahl-Jensen, B. Delmonte, G. Dreyfus, G. Durand, S. Falourd, H. Fischer, J. Flückiger, M. E. Hansson, P. Huybrechts, G. Jugie, S. J. Johnsen, J. Jouzel, P. Kaufmann, J. Kipfstuhl, F. Lambert, V. Y. Lipenkov, G. C. Littot, A. Longinelli, R. Lorrain, V. Maggi, V. Masson-Delmotte, H. Miller, R. Mulvaney, J. Oerlemans, H. Oerter, G. Orombelli, F. Parrenin, D. A. Peel, J. R. Petit, D. Raynaud, C. Ritz, U. Ruth, J. Schwander, U. Siegenthaler, R. Souchez, B. Stauffer, J. P. Steffensen, B. Stenni, T. F. Stocker, I. E. Tabacco, R. Udisti, R. S. Van De Wal, M. Van Den Broeke, J. Weiss, F. Wilhelms, J. G. Winther, E. W. Wolff, M. Zucchelli; EPICA community members, Eight glacial cycles from an Antarctic ice core. Nature 429, 623–628 (2004). Medline doi:10.1038/nature02599

39. J. Mitchell, R. Westaway, Chronology of Neogene and Quaternary uplift and magmatism in the Caucasus: Constraints from K–Ar dating of volcanism in Armenia. Tectonophysics 304, 157–186 (1999). doi:10.1016/S0040-1951(99)00027-X

40. D. Kharazyan, Geological Map of the Republic of Armenia: Scale 1/500,000 (Ministry of Nature Protection of the Republic of Armenia, Geological Agency, Yerevan, Armenia, 2005).

41. V. A. Lebedev, I. V. Chernyshev, A. I. Yakushev, Dokl. Earth Sci. 437, 808 (2011).

42. D. V. Nalivkin, Geology of the USSR (Oliver and Boyd, Edinburgh, UK, 1973).

43. M. Oddone, G. Bigazzi, Y. Keheyan, S. Meloni, J. Radioanal. Nucl. Chem. 243, 673–682 (2000). doi:10.1023/A:1010666118830

44. A. N. Komarov, N. V. Skovorodkin, S. G. Karapetian, Geochemica 6, 693 (1972).

45. S. G. Karapetyan, Structural and Compositional Features of Young Rhyolitic Volcanoes in the Armenian SSR (National Academy of Sciences of the Armenian SSR, Yerevan, Armenia, 1972).

46. S. J. Gale, P. G. Hoare, Quaternary Sediments: Petrographic Methods for the Study of Unlithified Rocks (Belhaven Press, London, 1991).

47. G. Stoops, Guidelines for Analysis And Description of Soil and Regolith Thin Sections (Soil Science Society of America, Madison, WI, 2003).

48. L. Bal, Neth. J. Agric. Sci. 23, 18 (1975).

49. R. G. C. Bathurst, Carbonate Sediments and Their Diagenesis (Elsevier, Amsterdam, 1971).

50. N. P. James, J. Sediment. Petrol. 42, 817 (1972).

51. L. P. Wilding, N. E. Smeck, G. F. Hall, Pedogenesis and Soil Taxonomy: The Soil Orders: The Soil Orders. Vol. 11 (Elsevier, Amsterdam, 1983).

52. S. P. McPherron, H. L. Dibble, http://oldstoneage.com.

53. D. F. Mark, S. Gonzalez, D. Huddart, H. Böhnel, Dating of the Valsequillo volcanic deposits: Resolution of an ongoing archaeological controversy in Central Mexico. J. Hum. Evol. 58, 441–445 (2010). Medline doi:10.1016/j.jhevol.2010.02.009

54. S. Blockley, S. D. F. Pyne-O’Donnell, J. J. Lowe, I. P. Matthews, A. Stone, A. M. Pollard, C. S. M. Turney, E. G. Molyneux, A new and less destructive laboratory procedure for the physical separation of distal glass tephra shards from sediments. Quat. Sci. Rev. 24, 1952–1960 (2005). doi:10.1016/j.quascirev.2004.12.008

55. K. P. Jochum, B. Stoll, K. Herwig, M. Willbold, A. W. Hofmann, M. Amini, S. Aarburg, W. Abouchami, E. Hellebrand, B. Mocek, I. Raczek, A. Stracke, O. Alard, C. Bouman, S. Becker, M. Dücking, H. Brätz, R. Klemd, D. de Bruin, D. Canil, D. Cornell, C.-J. de Hoog, C. Dalpé, L. Danyushevsky, A. Eisenhauer, Y. Gao, J. E. Snow, N. Groschopf, D. Günther, C. Latkoczy, M. Guillong, E. H. Hauri, H. E. Höfer, Y. Lahaye, K. Horz, D. E. Jacob, S. A. Kasemann, A. J. R. Kent, T. Ludwig, T. Zack, P. R. D. Mason, A. Meixner, M. Rosner, K. Misawa, B. P. Nash, J. Pfänder, W. R. Premo, W. D. Sun, M. Tiepolo, R. Vannucci, T. Vennemann, D. Wayne, J. D. Woodhead, MPI-DING reference glasses for

in situ microanalysis: New reference values for element concentrations and isotope ratios. Geochem. Geophys. Geosyst. 7, 1525 (2006). doi:10.1029/2005GC001060

56. S. A. Hynek, F. H. Brown, D. P. Fernandez, A rapid method for hand picking potassium-rich feldspar from silicic tephra. Quat. Geochronol. 6, 285–288 (2011). doi:10.1016/j.quageo.2010.11.004

57. S. Nomade, P. R. Renne, N. Vogel, A. L. Deino, W. D. Sharp, T. A. Becker, A. R. Jaouni, R. Mundil, Alder Creek sanidine (ACs-2): A Quaternary 40Ar/39Ar dating standard tied to the Cobb Mountain geomagnetic event. Chem. Geol. 218, 315–338 (2005). doi:10.1016/j.chemgeo.2005.01.005

58. P. R. Renne, C. C. Swisher, A. L. Deino, D. B. Karner, T. L. Owens, D. J. DePaolo, Intercalibration of standards, absolute ages and uncertainties in 40Ar/39Ar dating. Chem. Geol. 145, 117–152 (1998). doi:10.1016/S0009-2541(97)00159-9

59. V. C. Smith, D. F. Mark, R. A. Staff, S. P. E. Blockley, C. B. Ramsey, C. L. Bryant, T. Nakagawa, K. K. Han, A. Weh, K. Takemura, T. Danhara; Suigetsu 2006 Project members, Toward establishing precise 40Ar/39Ar chronologies for Late Pleistocene palaeoclimate archives: An example from the Lake Suigetsu (Japan) sedimentary record. Quat. Sci. Rev. 30, 2845–2850 (2011). doi:10.1016/j.quascirev.2011.06.020

60. D. F. Mark, D. Barfod, F. M. Stuart, J. Imlach, The ARGUS multicollector noble gas mass spectrometer: Performance for 40Ar/39Ar geochronology. Geophys. Geochem. Geosyst 10, 1 (2009). doi:10.1029/2009GC002643

61. P. R. Renne, W. S. Cassata, L. E. Morgan, The isotopic composition of atmospheric argon and 40Ar/39Ar geochronology: Time for a change? Quat. Geochronol. 4, 288–298 (2009). doi:10.1016/j.quageo.2009.02.015

62. D. F. Mark, C. M. Rice, A. E. Fallick, N. H. Trewin, M. R. Lee, A. Boyce, J. K. W. Lee, 40Ar/39Ar dating of hydrothermal activity, biota and gold mineralization in the Rhynie hot-spring system, Aberdeenshire, Scotland. Geochim. Cosmochim. Acta 75, 555–569 (2011). doi:10.1016/j.gca.2010.10.014

63. R. H. Steiger, E. Jäger, Subcommission on geochronology: Convention on the use of decay constants in geo- and cosmochronology. Earth Planet. Sci. Lett. 36, 359–362 (1977). doi:10.1016/0012-821X(77)90060-7

64. J.-L. Lee, K. Marti, J. P. Severinghaus, K. Kawamura, H.-S. Yoo, J. B. Lee, J. S. Kim, A redetermination of the isotopic abundances of atmospheric Ar. Geochim. Cosmochim. Acta 70, 4507–4512 (2006). doi:10.1016/j.gca.2006.06.1563

65. D. F. Mark, F. M. Stuart, M. De Podesta, New high-precision measurements of the isotopic composition of atmospheric argon. Geochim. Cosmochim. Acta 75, 7494–7501 (2011). doi:10.1016/j.gca.2011.09.042

66. P. R. Renne, A. L. Deino, W. E. Hames, M. T. Heizler, S. R. Hemming, K. V. Hodges, A. A. P. Koppers, D. F. Mark, L. E. Morgan, D. Phillips, Data reporting norms for 40Ar/39Ar geochronology. Quat. Geochronol. 4, 346–352 (2009). doi:10.1016/j.quageo.2009.06.005

67. P. R. Renne, R. Mundil, G. Balco, K. Min, K. R. Ludwig, Joint determination of 40K decay constants and 40Ar∗/40K for the Fish Canyon sanidine standard, and improved accuracy

for 40Ar/39Ar geochronology. Geochim. Cosmochim. Acta 74, 5349–5367 (2010). doi:10.1016/j.gca.2010.06.017

68. P. R. Renne, G. Balco, K. R. Ludwig, R. Mundil, K. Min, Response to the comment by W. H. Schwarz et al. on “Joint determination of 40K decay constants and 40Ar∗/40K for the Fish Canyon sanidine standard, and improved accuracy for 40Ar/39Ar geochronology” by P. R. Renne et al. (2010). Geochim. Cosmochim. Acta 75, 5097–5100 (2011). doi:10.1016/j.gca.2011.06.021

69. K. F. Kuiper, A. Deino, F. J. Hilgen, W. Krijgsman, P. R. Renne, J. R. Wijbrans, Synchronizing rock clocks of Earth history. Science 320, 500–504 (2008). Medline doi:10.1126/science.1154339

70. A. Tuffreau, in The Definition and Interpretation of Levallois Technology, H. L. Dibble, O. Bar-Yosef, Eds. (Monographs in World Archaeology 23, Prehistory Press, Madison, WI, 1995), pp. 413–427.

71. A. Tuffreau, P. Antoine, in The Earliest Occupation of Europe, W. Roebroeks, T. van Kolfschoten, Eds. (Leiden Univ. Press, Leiden, Netherlands, 1995), pp. 147–164.

72. D. A. Roe, The Lower and Middle Palaeolithic Periods in Britain (Routledge and Kegan Paul, London, 1981).

73. D. R. Bridgland, The Quaternary of the Thames (Chapman and Hall, London, 1994).

74. M. Terradillos-Bernal, Les gisements de Galería, Gran Dolina TD10 et Ambrona (le Complexe Inférieur) : Trois modèles technologiques dans le deuxième tiers du Pléistocène moyen. Anthropologie 117, 494–514 (2013). doi:10.1016/j.anthro.2013.10.004

75. C. Falguères, J. J. Bahain, Y. Yokoyama, J. L. Arsuaga, J. M. Bermudez de Castro, E. Carbonell, J. L. Bischoff, J. M. Dolo, Earliest humans in Europe: The age of TD6 Gran Dolina, Atapuerca, Spain. J. Hum. Evol. 37, 343–352 (1999). Medline doi:10.1006/jhev.1999.0326

76. E. Carbonell, M. Mosquera, A. Ollé, X. P. Rodríguez, M. Sahnouni, R. Sala, J. M. Vergès, Structure morphotechnique de l’industrie lithique du Pléistocène inférieur et moyen d’Atapuerca (Burgos, Espagne). Anthropologie 105, 259–280 (2001). doi:10.1016/S0003-5521(01)80016-9

77. C. Falguères, J.-J. Bahain, A. Pérez-González, N. Mercier, M. Santonja, J.-M. Dolo, The Lower Acheulian site of Ambrona, Soria (Spain): Ages derived from a combined ESR/U-series model. J. Archaeol. Sci. 33, 149–157 (2006). doi:10.1016/j.jas.2005.07.006

78. L. Bourguignon, H. Djema, P. Bertran, C. Lahaye, P. Guibert, in Les sociétés du Paléolithique dans un grand Sud-Ouest de la France: nouveaux gisements, nouveaux résultats, nouvelles méthodes, J. Jaubert, J.-G. Bordes, Eds. (Société préhistorique française, Paris, 2008), pp. 41–55.

79. E. Carbonell, X. P. Rodríguez, Veleia 24–25, 331 (2007–2008).

80. A. Lamotte, A. Tuffreau, in L’Acheuléen dans la vallée de la Somme et Paléolithique moyen dans le nord de la France: données récentes, A. Tuffreau, Ed. (Centre d'Etudes et Recherches Préhistoriques, Lille, France, 2001), pp. 149–153.

81. M. Santonja, P. Villa, The Lower Paleolithic of Spain and Portugal. J. World Prehist. 4, 45–94 (1990). doi:10.1007/BF00974819

82. V. Michel, G. Shen, C. C. Shen, C. C. Wu, C. Vérati, S. Gallet, M. H. Moncel, J. Combier, S. Khatib, M. Manetti, Application of U/Th and 40Ar/39Ar dating to Orgnac 3, a Late Acheulean and Early Middle Palaeolithic site in Ardèche, France. PLOS ONE 8, e82394 (2013). Medline doi:10.1371/journal.pone.0082394

83. C. Falguères, J.-J. Bahain, H. Saleki, U-Series and ESR dating of teeth from Acheulian and Mousterian levels at La Micoque (Dordogne, France). J. Archaeol. Sci. 24, 537–545 (1997). doi:10.1006/jasc.1996.0137

84. S. Grimaldi, Paléo 10, 109 (1998).

85. M. Santonja, P. Villa, in Axe Age. Acheulian Toolmaking from Quarry to Discard, N. Goren-Inbar, G. Sharon, Eds. (Approaches in Anthropological Archaeology, Equinox, London, 2006), pp. 429–478.

86. D. R. Bridgland, P. Harding, P. Allen, I. Candy, C. Cherry, W. George, D. J. Horne, D. H. Keen, K. E. H. Penkman, R. C. Preece, E. J. Rhodes, R. Scaife, D. C. Schreve, J.-L. Schwenninger, I. Slipper, G. R. Ward, M. J. White, T. S. White, J. E. Whittaker, An enhanced record of MIS 9 environments, geochronology and geoarchaeology: Data from construction of the High Speed 1 (London–Channel Tunnel) rail-link and other recent investigations at Purfleet, Essex, UK. Proc. Geol. Assoc. 124, 417–476 (2013). doi:10.1016/j.pgeola.2012.03.006

87. J. Buraczynski, J. Butrym, in Loess and Environment, M. Pesci, Ed. (Catena Supplement, 1987), pp. 81–94.

88. J. Junkmanns, Archäologisches Korrespondenzblatt 21, 1 (1991).

89. R. Westaway, D. Bridgland, M. White, The Quaternary uplift history of central southern England: Evidence from the terraces of the Solent River system and nearby raised beaches. Quat. Sci. Rev. 25, 2212–2250 (2006). doi:10.1016/j.quascirev.2005.06.005

90. C. Ryssaert, Anthropologica Praehistorica 117, 13 (2006).

91. M. Jarry, D. Colonge, L.-A. Lelouvier, V. Mourre, Les Bosses (La Magdelaine, Lot, France): Un gisement paléolithique moyen antérieur à l'avant-dernier Interglaciaire sur la moyenne terrasse du Lot (Société Préhistorique Française, Paris, 2007).

92. A. van Baelen, E. P. M. Meijs, P. Van Peer, J.-P. De Warrimont, M. De Bie, Notae Praehistoricae 27, 19 (2007).

93. J. Schäfer, T. Laurat, J. Kegler, Praehistoria Thuringica 10, 141 (2004).

94. J. Thissen, Die Paläolithischen Freilandstationen von Rheindahlen im Löss Zwischen Maas und Niederrhein (Rheinische Ausgrabungen 59, Verlag Philipp von Zabern, Mainz, Germany, 2006).

95. D. Cologne et al., De la transition paléolithique inférieur-moyen dans la vallée de la Garonne: l’Acheuléen supérieur de Raspide 2 (Blagnac, Haute-Garonne) (Société préhistorique française, Paris, 2010).

96. A. Tuffreau, in The Transition from the Lower to Middle Palaeolithic and the Origin of Modern Man, A. Ronen, Ed. (British Archaeological Reports, International Series 151, Oxford, 1982), pp. 137–149.

97. J. Gagnepain, C. Gaillard, in Les premiers peuplements en Europe. Données récentes sur les modalités de peuplement et sur le cadre chronostratigraphique, géologique et paléogéographique des industries du Paléolithique ancien et moyen en Europe (Rennes, 22–25 septembre 2003), N. Molines, M. H. Moncel, Eds. (British Archaeological Reports, International Series 1364, Oxford, 2005), pp. 73–86.

98. J. Gagnepain, C. Gaillard, La grotte-abri de la Baume Bonne: une séquence chrono-stratigraphique de 300,000 ans (Alpes-de-Haute-Provence, Quinson) (Rapport de fouilles programmées, Service régional d’archéologie, Ed., Aix-en-Provence, France, 1996).

99. B. Scott, N. Ashton, S. Lewis, S. Parfitt, M. White, in The Ancient Human Occupation of Britain, N. Ashton, S. G. Lewis, C. B. Stringer, Eds. (Elsevier, Amsterdam, 2011), pp. 67–89.

100. A. Delagnes, A. Ropars, Eds., Paléolithique moyen en pays de Caux (Haute-Normandie). Le Pucheuil, Ettoutteville: Deux gisements de plein air en milieu loessique (Documents d'archéologique Francaise 56, Éditions de la Maison des Sciences de l’Homme, Paris, 1996).

101. A. E. Marks, J.-Ph. Brugal, V. P. Chabai, K. Monigal, P. Goldberg, B. Hockett, E. Peman, M. Elorza, C. Mallol, Paléo 14, 77 (2002).

102. G. W. Berger, A. Pérez-González, E. Carbonell, J. L. Arsuaga, J. M. Bermúdez de Castro, T. L. Ku, Luminescence chronology of cave sediments at the Atapuerca paleoanthropological site, Spain. J. Hum. Evol. 55, 300–311 (2008). Medline doi:10.1016/j.jhevol.2008.02.012

103. X. Sun, N. Mercier, C. Falgueres, J.-J. Bahain, J. Despriee, G. Bayle, H. Lu, Recuperated optically stimulated luminescence dating of middle-size quartz grains from the Palaeolithic site of Bonneval (Eure-et-Loir, France). Quat. Geochronol. 5, 342–347 (2010). doi:10.1016/j.quageo.2009.05.004

104. E. Boëda, B. Kervazo, N. Mercier, H. Valladas, in Proceedings of the International Round Table: Reduction Processes (Chaînes Opératoires) for the European Mousterian, A. Bietti, S. Grimaldi, Eds. (Rome, 1996), pp. 465–504.

105. P. Callow, J. M. Cornford, La Cotte de St. Brelade 1961-1978: Excavation by C. B. M. McBurney (Kluwer Academic, Norwich, UK, 1986).

106. J. Vandenberghe, W. Roebroeks, T. van Kolfschoten, Maastricht-Belvédère: Stratigraphy. Palaeoenvironment and Archaeology of the Middle and Late Pleistocene Deposits. Part II (Mededelingen Rijks Geologische Dienst, Leiden, Netherlands, 1993).

107. J. J. Bahain, La méthode de datation par résonance de spin électronique (ESR) au Muséum national d’histoire naturelle. Vingt ans de recherches méthodologiques et d’applications géochronologiques (thesis, Université Michel de Montaigne Bordeaux 3, Bordeaux, France, 2007).

108. Y. Yokoyama, in Hominidae, G. Giacobini, Ed. (Proceedings of the 2nd International Congress of Human Paleontology, Turin, Italy, 1989), pp. 387–390.

109. J. Sommé, A. Tuffreau, M. J. Aikten, P. Auguste, J. Chaline, J. P. Colbeaux, N. Cunat-Boge, R. Geeaerts, J. Hus, J. Huxtable, E. Juvigne, A. V. Munaut, S. Occhietti, P. Pichet, J. J. Puissegur, D. D. Rousseau, B. Van Vliet-Lanoe, in Le gisement paléolithique moyen de Biache-Saint-Vaast (Pas-de-Calais) Stratigraphie, environnement études archéologique, A. Tuffreau, J. Sommé, Eds. (Mémoires de la Société Préhistorique Française, Paris, 1988), pp. 115–119.

110. M. Brenet, M. Folgado, P. Bertran, A. Lenoble, P. Guibert, E. Vieillevigne, in Les sociétés du Paléolithique dans un grand Sud-Ouest de la France: Nouveaux gisements, nouveaux résultats, nouvelles méthodes, J. Jaubert, J.-G. Bordes, I. Ortega, Eds. (Société Préhistorique Française, Paris, 2008), pp. 57–81.

111. P. Haesaerts, L. V. Koulakovskaya, in The European Middle Paleolithic, L. V. Koulakovskaya, Ed. (Shlyakh, Kiev, Ukraine, 2006), pp. 21–38.

112. R. Mallick, N. Frank, A new technique for precise uranium-series dating of travertine micro-samples. Geochim. Cosmochim. Acta 66, 4261–4272 (2002). doi:10.1016/S0016-7037(02)00999-7

113. B. Blackwell, H. P. Schwarcz, U-Series analyses of the lower travertine at Ehringsdorf, DDR. Quat. Res. 25, 215–222 (1986). doi:10.1016/0033-5894(86)90058-X

114. A. Adam, Les pointes pseudo-Levallois du gisement moustérien Le Rissori, à Masnuy-Saint-Jean (Hainaut, Belgique). Anthropologie 106, 695–730 (2002). doi:10.1016/S0003-5521(02)01143-3

115. A. Wisniewski, J. Fridrich, in Middle Palaeolithic Activity and Palaeoecology, J. M. Burdukiewicz, A. Wisniewski, Eds. (Studia Archeologiczne 41, Biblioteka Cyfrowa, Wydawnictwa Uniwersytetu Wroclawskiego, Wroclaw, Poland 2010), pp. 217–243.

116. S. Ertmer, Jahresschr. Mus. Landkreises Börde 51, 13 (2011).

117. A. Wisniewski, Quat. Int. 326–327, 364 (2013).

118. N. Ameloot-van der Heijden, C. Dupuis, N. Limondin, A. V. Munaut, J. J. Puissegur, Anthropologie 100, 555 (1996).

119. K. Cyrek, M. Sudol, L. Czyzewski, G. Osipowicz, M. Grelowska, Middle Palaeolithic cultural levels from Middle and Late Pleistocene sediments of Biśnik Cave, Poland. Quat. Int. 326-327, 20–63 (2014). doi:10.1016/j.quaint.2013.12.014

120. B. Blackwell, H. P. Schwarcz, in La Grotte Vaufrey: Paléoenvironnement, chronologie, activités humaines, J. P. Rigaud, Ed. (Mémoires de la Société Préhistorique Française, Paris, 1988), pp. 365–380.

121. J. Jaubert, C. Servelle, in L’Acheuléen dans l’Ouest de l’Europe, A. Tuffreau, Ed. (University of Sciences & Technology, Centre d'Etudes et Recherches Préhistoriques, Lille, France, 1996), pp. 77–108.

122. S. Balescu, A. Tuffreau, Archaeol. Almanac 16, 5 (2004).

123. J.-L. Locht, P. Antonie, D. Hérisson, G. Gadebois, N. Debenham, Gallia Préhistorie 52, 1 (2010).

124. R. Rabinovich, O. Ackermann, E. Aladjem, R. Barkai, R. Biton, I. Milevski, N. Solodenko, O. Marder, Elephants at the Middle Pleistocene Acheulian open-air site of Revadim Quarry, Israel. Quat. Int. 276–277, 183–197 (2012). doi:10.1016/j.quaint.2012.05.009

125. O. Marder, Y. Milevski, R. Rabinovich, Revadim Quarry, Israel Antiquities Authority, Ed. (Excavations and Surveys in Israel 119, 2007); http://www.antiquities.org.il/home_heb.asp..

126. G. Gvirtzman, M. Wieder, O. Marder, H. Khalaily, R. Rabinovich, H. Ron, Geological and pedological aspects of an Early-Paleolithic site: Revadim, Central Coastal Plain, Israel. Geoarchaeol. 14, 101–126 (1999). doi:10.1002/(SICI)1520-6548(199902)14:2<101::AID-GEA1>3.0.CO;2-1

127. J.-M. L. Tensorer, R. Jagher, P. Rentzel, T. Hauck, K. Ismail-Meyer, C. Pümpin, D. Wojtczak, Long-term site formation processes at the natural springs Nadaouiyeh and Hummal in the El Kowm Oasis, Central Syria. Geoarchaeology 22, 621–640 (2007). doi:10.1002/gea.20177

128. J.-M. Le Tensorer, D. Wojtczak, P. Schmid, K. Meyer, A.- S. Martineau, D. Schuhmann, T. Hauck, Hummal et Nadaouiyeh (El Kowm, Syrie centrale), Resultats de la campagne 2007; http://elkowm.unibas.ch/Bilder/Publikationen/2007/RAP%20Fouille%2007.pdf. (2007).

129. N. Mercier, H. Valladas, C. Falguères, Q. Shao, A. Gopher, R. Barkai, J.-J. Bahain, L. Vialettes, J.-L. Joron, J.-L. Reyss, New datings of Amudian layers at Qesem Cave (Israel): Results of TL applied to burnt flints and ESR/U-series to teeth. J. Archaeol. Sci. 40, 3011–3020 (2013). doi:10.1016/j.jas.2013.03.002

130. A. Gopher, A. Ayalon, M. Bar-Matthews, R. Barkai, A. Frumkin, P. Karkanas, R. Shahack-Gross, The chronology of the late Lower Paleolithic in the Levant based on U–Th ages of speleothems from Qesem Cave, Israel. Quat. Geochronol. 5, 644–656 (2010). doi:10.1016/j.quageo.2010.03.003

131. W. J. Rink, H. P. Schwarcz, A. Ronen, A. Tsatskin, Confirmation of a near 400 ka age for the Yabrudian industry at Tabun Cave, Israel. J. Archaeol. Sci. 31, 15–20 (2004). doi:10.1016/S0305-4403(03)00080-3

132. N. Mercier, H. Valladas, Reassessment of TL age estimates of burnt flints from the Paleolithic site of Tabun Cave, Israel. J. Hum. Evol. 45, 401–409 (2003). Medline doi:10.1016/j.jhevol.2003.09.004

133. N. Mercier, H. Valladas, G. Valladas, J.-L. Reyss, A. Jelinek, L. Meignen, J.-L. Joron, TL dates of burnt flints from Jelinek’s excavations at Tabun and their implications. J. Archaeol. Sci. 22, 495–509 (1995). doi:10.1006/jasc.1995.0048

134. R. Grün, C. Stringer, Tabun revisited: Revised ESR chronology and new ESR and U-series analyses of dental material from Tabun C1. J. Hum. Evol. 39, 601–612 (2000). Medline doi:10.1006/jhev.2000.0443

135. R. Shimelmitz, S. L. Kuhn, Paleoanthropology 2013, 1 (2013).

136. N. Porat, M. Chazan, H. Schwarcz, L. K. Horwitz, Timing of the Lower to Middle Paleolithic boundary: New dates from the Levant. J. Hum. Evol. 43, 107–122 (2002). Medline doi:10.1006/jhev.2002.0562

137. M. Otte, I. Yalçinkaya, J. Kozlowski, O. Bar-Yosef, I. Lopez Bayón, H. Taskiran, Long-term technical evolution and human remains in the Anatolian palaeolithic. J. Hum. Evol. 34, 413–431 (1998). Medline doi:10.1006/jhev.1997.0199

138. M. Otte et al., Anthropologie 99, 529 (1995).

139. M. Weinstein-Evron, A. Tsatskin, N. Porat, J. Kronfeld, S. Afr. J. Sci. 95, 186 (1999).

140. N. Mercier, H. Valladas, L. Froget, J.-L. Joron, J.-L. Reyss, S. Weiner, P. Goldberg, L. Meignen, O. Bar-Yosef, A. Belfer-Cohen, M. Chech, S. L. Kuhn, M. C. Stiner, A.-M. Tillier, B. Arensburg, B. Vandermeersch, Hayonim Cave: A TL-based chronology for this Levantine Mousterian sequence. J. Archaeol. Sci. 34, 1064–1077 (2007). doi:10.1016/j.jas.2006.09.021

141. M. D. Petraglia, A. Alsharekh, P. Breeze, C. Clarkson, R. Crassard, N. A. Drake, H. S. Groucutt, R. Jennings, A. G. Parker, A. Parton, R. G. Roberts, C. Shipton, C. Matheson, A. Al-Omari, M. A. Veall, Hominin dispersal into the Nefud Desert and Middle palaeolithic settlement along the Jubbah Palaeolake, Northern Arabia. PLOS ONE 7, e49840 (2012). Medline doi:10.1371/journal.pone.0049840

142. C. S. Churcher, M. R. Kleindienst, H. P. Schwarcz, Faunal remains from a Middle Pleistocene lacustrine marl in Dakhleh Oasis, Egypt: Palaeoenvironmental reconstructions. Palaeogeogr. Palaeoclimatol. Palaeoecol. 154, 301–312 (1999). doi:10.1016/S0031-0182(99)00104-2

143. J. R. Smith, R. Giegengack, H. P. Schwarcz, M. M. A. McDonald, M. R. Kleindienst, A. L. Hawkins, C. S. Churcher, A reconstruction of Quaternary pluvial environments and human occupations using stratigraphy and geochronology of fossil-spring tufas, Kharga Oasis, Egypt. Geoarchaeol. 19, 407–439 (2004). doi:10.1002/gea.20004

144. J. P. Raynal, F. Z. Sbihi Alaoui, D. Geraads, L. Magoga, A. Mohi, The earliest occupation of North-Africa: The Moroccan perspective. Quat. Int. 75, 65–75 (2001). doi:10.1016/S1040-6182(00)00078-1

145. E. J. Rhodes, J. P. Raynal, D. Geraads, F. Z. Sbihialaoui, C. R. Acad. Sci. Paris 319, 1109 (1994).

146. A. L. Deino, S. McBrearty, 40Ar/39Ar dating of the Kapthurin Formation, Baringo, Kenya. J. Hum. Evol. 42, 185–210 (2002). Medline doi:10.1006/jhev.2001.0517

147. C. A. Tryon, S. McBrearty, P.-J. Texier, Levallois lithic technology from the Kapthurin Formation, Kenya: Acheulian origin and Middle Stone Age diversity. Afr. Archaeol. Rev. 22, 199–229 (2006). doi:10.1007/s10437-006-9002-5

148. R. Grün, J. S. Brink, N. A. Spooner, L. Taylor, C. B. Stringer, R. G. Franciscus, A. S. Murray, Direct dating of Florisbad hominid. Nature 382, 500–501 (1996). Medline doi:10.1038/382500a0

149. R. Grün, Direct dating of human fossils. Yearb. Phys. Anthropol. 131, 2–48 (2006). doi:10.1002/ajpa.20516

150. S. C. Reynolds, R. J. Clarke, K. A. Kuman, The view from the Lincoln Cave: Mid- to late Pleistocene fossil deposits from Sterkfontein hominid site, South Africa. J. Hum. Evol. 53, 260–271 (2007). Medline doi:10.1016/j.jhevol.2007.02.004

151. S. McBrearty, C. A. Tryon, in Transitions Before the Transition: Evolution and Stability in the Middle Paleolithic and Middle Stone Age, E. Hovers, S. L. Kuhn, Eds. (Springer, New York, 2006), pp. 257–276.

152. A. L. Hawkins et al., Nyame Akuma 55, 8 (2001).

153. P. Van Peer, R. Fullagar, S. Stokes, R. M. Bailey, J. Moeyersons, F. Steenhoudt, A. Geerts, T. Vanderbeken, M. De Dapper, F. Geus, The Early to Middle Stone Age transition and the emergence of modern human behaviour at site 8-B-11, Sai Island, Sudan. J. Hum. Evol. 45, 187–193 (2003). Medline doi:10.1016/S0047-2484(03)00103-9

154. P. van Peer, V. Rots, J. M. Vroomans, Before Farming 3, article 1 (2004).

155. P. Van Peer, The Levallois Reduction Strategy (Prehistory Press, Madison, WI, 1992).

156. J. J. Shea, Lithic Modes A–I: A new framework for describing global-scale variation in stone tool technology illustrated with evidence from the East Mediterranean Levant. J. Archaeol. Method Theory 20, 151–186 (2013). doi:10.1007/s10816-012-9128-5

157. N. Goren-Inbar, G. Sharon, 2006. Axe Age: Acheulian Tool-making from Quarry to Discard (Equinox, London, 2006).

158. R. Barkai, A. Gopher, in The Lower and Middle Palaeolithic in the Middle East and Neighbouring Regions, J.-M. Tensorer, R. Jagher, M. Otte, Eds. (ERAUL 126, Lèige, France, 2011), pp. 49–57.

159. J.-M. Le Tensorer, V. Von Falkenstein, H. Le Tensorer, S. Muhesen, in The Lower and Middle Palaeolithic in the Middle East and Neighbouring Regions, J.-M. Tensorer, R. Jagher, M. Otte, Eds. (ERAUL 126, Lèige, Belgium, 2011), pp. 235–248.

160. E. Frahm, Validity of “off-the-shelf” handheld portable XRF for sourcing Near Eastern obsidian chip debris. J. Archaeol. Sci. 40, 1080–1092 (2013). doi:10.1016/j.jas.2012.06.038

161. S. Phillips, R. J. Speakman, Initial source evaluation of archaeological obsidian from the Kuril Islands of the Russian Far East using portable XRF. J. Archaeol. Sci. 36, 1256–1263 (2009). doi:10.1016/j.jas.2009.01.014

162. N. Craig, R. J. Speakman, R. S. Popelka-Filcoff, M. Aldenderfer, L. Flores Blanco, M. B. Vega, M. D. Glascock, C. Stanish, Macusani obsidian from southern Peru: A characterization of its elemental composition with a demonstration of its ancient use. J. Archaeol. Sci. 37, 569–576 (2010). doi:10.1016/j.jas.2009.10.021

163. P. Jia, T. Doelman, C. Chen, H. Zhao, S. Lin, R. Torrence, M. D. Glascock, Moving sources: A preliminary study of volcanic glass artifact distributions in northeast China using PXRF. J. Archaeol. Sci. 37, 1670–1677 (2010). doi:10.1016/j.jas.2010.01.027

164. A. Nazaroff, K. Prufer, B. Drake, Assessing the applicability of portable X-ray fluorescence spectrometry for obsidian provenance research in the Maya lowlands. J. Archaeol. Sci. 37, 885–895 (2010). doi:10.1016/j.jas.2009.11.019

165. D. V. Burley, P. J. Sheppard, M. Simonin, Tongan and Samoan volcanic glass: pXRF analysis and implications for constructs of ancestral Polynesian society. J. Archaeol. Sci. 38, 2625–2632 (2011). doi:10.1016/j.jas.2011.05.016

166. M. McCoy, P. R. Mills, S. Lundblad, T. Rieth, J. G. Kahn, R. Gard, A cost surface model of volcanic glass quarrying and exchange in Hawai‘i. J. Archaeol. Sci. 38, 2547–2560 (2011). doi:10.1016/j.jas.2011.04.017

167. P. Sheppard, G. Irwin, S. Lin, C. McCaffrey, Characterization of New Zealand obsidian using PXRF. J. Archaeol. Sci. 38, 45–56 (2011). doi:10.1016/j.jas.2010.08.007

168. R. H. Tykot, L. Lai, C. Tozzi, in Proceedings of the 37th International Symposium on Archaeometry, I. Turbanti-Memmi Ed. (Springer, New York, 2011), pp. 321–328.

169. N. Forster, P. Grave, Non-destructive PXRF analysis of museum-curated obsidian from the Near East. J. Archaeol. Sci. 39, 728–736 (2012). doi:10.1016/j.jas.2011.11.004

170. N. Goodale, D. G. Bailey, G. T. Jones, C. Prescott, E. Scholz, N. Stagliano, C. Lewis, pXRF: A study of inter-instrument performance. J. Archaeol. Sci. 39, 875–883 (2012). doi:10.1016/j.jas.2011.10.014

171. E. Frahm, R. Doonan, V. Kilikoglou, Handheld portable x-ray fluorescence of Aegean obsidians. Archaeometry 56, 228–260 (2014). doi:10.1111/arcm.12012

172. E. Frahm, Characterizing obsidian sources with portable XRF: Accuracy, reproducibility, and field relationships in a case study from Armenia. J. Arch Sci 49, 105–125 (2014). doi:10.1016/j.jas.2014.05.003

173. M. White, N. Ashton, B. Scott, Dev. Quat. Sci. 14, 53 (2011).

174. C. Shipton, C. Clarkson, J. N. Pal, S. C. Jones R. G. Roberts, C Harris, M. C. Gupta, P. W. Ditchfield, M. D. Petraglia, Generativity, hierarchical action and recursion in the technology of the Acheulean to Middle Palaeolithic transition: A perspective from Patpara, the Son Valley, India. J. Hum. Evol. 65, 93 (2013).

175. S. J. Lycett, C. J. Norton, A demographic model for Palaeolithic technological evolution: The case of East Asia and the Movius Line. Quat. Int 211, 55–65 (2010). doi:10.1016/j.quaint.2008.12.001

176. R. R. Klein, The Human Career. Human Biological and Cultural Origins (Univ. of Chicago Press, Chicago, IL, 1999).

177. M. M. Lahr, in Out of Africa I: The First Hominin Colonization of Eurasia, J. G. Fleagle, . J. Shea, F. E. Grine, A. L. Baden, R. E. Leakey, Eds. (Vertebrate Paleobiology and Paleoanthropology 27, Springer, Dordrecht, Netherlands, 2010), pp. 27–46.

178. K. Kuman, in Breathing Life Into Fossils: Taphonomic Studies in Honor of C. K. (Bob) Brain, T. R. Pickering, K. Schick, N. Toth, Eds. (Stone Age Institute Publication Series 2, Stone Age Institute Press, Gosport, UK, 2007), pp. 181–198.

179. N. Galanidou, J. Cole, G. Iliopoulos, J. McNabb, Antiquity 87 (2013); http://antiquity.ac.uk/projgall/galanidou336/.

180. R. Dennell, The Palaeolithic Settlement of Asia (Cambridge Univ. Press, Cambridge, 2009).

181. C. Gaillard, M. Singh, K. K. Rishi, Man and Environment XXXIII, 1 (2008).

182. S. Pappu, Y. Gunnell, K. Akhilesh, R. Braucher, M. Taieb, F. Demory, N. Thouveny, Early Pleistocene presence of Acheulian hominins in South India. Science 331, 1596–1599 (2011). Medline doi:10.1126/science.1200183

183. A. Brumm, M. W. Moore, Aust. Archaeol. 74, 32 (2012).

184. O. Jöris, in Cambridge World Prehistory, C. Renfrew, P. Bahn, Eds. (Cambridge Univ. Press, Cambridge, 2014), pp. 1703–1746.

185. P. Biagi, Archaeology of Early Northeastern Africa/Studies in African Archaeology 9, 1005 (2006).

186. N. Rolland, in The Definition and Interpretation of Levallois Technology, H. L. Dibble, O. Bar-Yosef, Eds. (Prehistory Press, Madison, WI, 1995), pp. 333–359.

187. H. DeBono, N. Goren-Inbar, J. Israel Preh. Soc. 31, 23 (2001).

188. S. J. Lycett, Are Victoria West cores “proto-Levallois”? A phylogenetic assessment. J. Hum. Evol. 56, 175–191 (2009). Medline doi:10.1016/j.jhevol.2008.10.001

189. O. Marder, H. Khalaily, R. Rabinovich, G. Gvirtzman, M. Wieder, N. Porat, H. Ron, R. Bankirer, I. Saragusti, J. Israel Prehist. Soc 28, 21 (1998).

190. O. Marder, I. Milevski, Z. Matskevich, in Axe Age: Acheulian Tool-making from Quarry to Discard, N. Goren-Inbar, G. Sharon, Eds. (Equinox, London, 2006), pp. 223–242.

191. Y. Zaidner, D. Druck, M. Weinstein-Evron, in Axe Age: Acheulian Tool-making from Quarry to Discard, N. Goren-Inbar, G. Sharon, Eds. (Equinox, London, 2006), pp. 243–266.

192. S. J. Lycett, N. von Cramon-Taubadel, J. A. J. Gowlett, A comparative 3D geometric morphometric analysis of Victoria West cores: Implications for the origins of Levallois technology. J. Archaeol. Sci. 37, 1110–1117 (2010). doi:10.1016/j.jas.2009.12.011