FAVREAU, J. 2023. Sourcing Oldowan and Acheulean Stone Tools in Eastern Africa. Aims, Methods, Challenges, and State of Knowledge

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Quaternary Science Advances 9 (2023) 100068

Contents lists available at ScienceDirect

Quaternary Science Advances


journal homepage: www.sciencedirect.com/journal/quaternary-science-advances

Review

Sourcing Oldowan and Acheulean stone tools in Eastern Africa: Aims,


methods, challenges, and state of knowledge
Julien Favreau
Department of Anthropology, McMaster University, Hamilton, ON, L8S 4L9, Canada

A R T I C L E I N F O A B S T R A C T

Keywords: Eastern Africa’s Plio-Pleistocene palaeoanthropological record has shaped our understanding of human biolog­
Oldowan ical and cultural evolution. Over the years, raw material sourcing has emerged as an important research topic in
Acheulean lithic analysis as it can allow for the identification of resource extraction points and transport distances as a
Raw material
means to infer other aspects of hominin behaviour that are of high evolutionary significance. The goal of this
Sourcing
article is to review and synthesise the aims, methods, challenges, and knowledge on raw material sourcing from
Provenance
Eastern Africa Plio-Pleistocene contexts in Eastern Africa. Beginning with a review of the Oldowan and Acheulean records, four
Plio-Pleistocene over-arching patterns are identified based on evidence from 130 localities. First, hominin toolmakers regularly
exerted selective criteria when choosing raw materials by way of opportunistic and more specialised procure­
ment strategies. Second, the fragmentation of technological activities across the palaeolandscape emerged as a
behavioural characteristic among Oldowan toolmakers before the first appearance datum of the Acheulean.
Third, hominins across Eastern Africa preferentially utilised igneous rock types followed by metamorphic and
sedimentary lithologies mirroring the regional lithostratigraphy. Finally, Acheulean toolmakers largely
mimicked their Oldowan counterparts in terms of raw material provisioning until the late Early Pleistocene,
when they began to engage in qualitatively different behaviour best evidenced by stone transport over longer
distances.
This state of archaeological knowledge serves as the basis to then review the theoretical and methodological
underpinnings of raw material sourcing followed by interpretive challenges. On a fundamental level, the iden­
tification of raw material sources is contingent on the implementation of a systematic sequence of analysis in
which the resulting data abide to the provenance postulate. This serves to preface a devoted section on the array
of analytical methods that can be successfully employed by archaeologists to source raw materials. Proven and
innovative methods are identified by bringing into dialogue key factors such as accuracy, precision, reproduc­
ibility, discriminatory power, sensitivity, destructiveness, throughput, cost, ease, and the question of spatial
scale. According to cost-benefit considerations, analytical methods that fall under the umbrella of geochemical
fingerprinting are found to generally outperform macroscopic and petrographic techniques. It is also argued that
characterising non-obsidian lithologies is best accomplished using more than one analytical method, with the
understanding that once positive baseline results are obtained subsequent testing can be narrowed down from a
methodological standpoint. Regardless of the characterisation method, it is imperative to implement effective
means to analyse the resulting data, which constitutes this article’s penultimate section. Ranging from traditional
graphical data representations to multivariate statistics and emerging branches of artificial intelligence, there are
countless means through which characteristic source signals can be identified regardless of rock type. The final
section reviews common interpretive challenges when studying raw material provisioning in deep time. While
the concepts of mobility, time-averaging, recycling, source differentiation, and selection criteria each present
unique challenges, it is argued that they can be reasonably overcome by way of multidisciplinary lines of evi­
dence. Ultimately, it is found that the future of raw material sourcing from Plio-Pleistocene contexts in Eastern
Africa is promising given what is currently known along with the subfield’s robust foundations.

E-mail address: favreaj@mcmaster.ca.

https://doi.org/10.1016/j.qsa.2022.100068
Received 16 October 2022; Received in revised form 14 December 2022; Accepted 19 December 2022
Available online 22 December 2022
2666-0334/© 2022 The Author. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-
nc-nd/4.0/).
J. Favreau Quaternary Science Advances 9 (2023) 100068

1. Introduction types can be effectively sourced, which, when combined with multi­
disciplinary evidence, will allow researchers to garner the full potential
Researchers have long attempted to identify the unique facets of of the human evolutionary record.
human evolution founded upon the analysis of hominin hard tissues,
stone artefacts, faunal remains, and trace fossils contextualised by 2. Raw material provisioning in the Earlier Stone Age
chronological and palaeoecological information. Based on the distribu­
tion of most first appearance datums in the Plio-Pleistocene palae­ The Earlier Stone Age (ESA) encompasses the Oldowan and Acheu­
oanthropological record, the Eastern African tropical zone likely lean industries, which emerged in Eastern Africa ~2.6 Ma and ~1.7 Ma,
represents the centre of endemism of Oldowan and Acheulean tool­ respectively (Asfaw et al., 1992; Braun et al., 2019; Beyene et al., 2013;
makers (Foley, 2018). However, the palaeoanthropological record is Diez-Martín et al., 2015; Lepre et al., 2011; Semaw et al., 1997, 2003).
incomplete, palimpsestic, spatially biased, and imprecisely dated, which The Acheulean emerged parallel to the late Oldowan and represents the
warrants caution as to the accuracy of the foregoing, and any other, most widespread and longest lasting cultural phenomenon on the globe,
evolutionary model (Bailey, 2007; Smith and Wood, 2017). From an persisting as late as ~160–154 ka in Eastern Africa (Kuman, 2014;
archaeological perspective, these limitations necessitate the thorough McBrearty, 2003). Many ESA site complexes are known for their vast
examination of stone tools, which outnumber any other artefact type collections of stone artefacts that were deposited near lake and river
due to their durability, using different theoretical and methodological systems and are occasionally found alongside surface modified bones
approaches to maximize their interpretive remit. One fundamental related to anthropogenic behaviours (e.g. Diez-Martín et al., 2009;
approach to lithic analysis involves raw material sourcing, a process Howell, 1961; Isaac and Isaac, 1997; Potts et al., 1999). From an arte­
typically achieved by matching stone tools with geological samples from factual perspective, Oldowan assemblages are defined by the presence of
known areas, which can allow for the identification of resource extrac­ hammerstones, cores, flakes, debris, and manuports, which may be
tion points and transport distances across palaeolandscapes to infer subdivided into several types according to the traditional typological
other aspects of hominin behaviour (e.g. Carter, 2014; Moutsiou, 2014). approach (Leakey, 1971) (Fig. 1). Acheulean assemblages also include
The probabilistic identification of raw material sources is contingent such artefact types but are primarily differentiated from the former
on analytical data that abide to the provenance postulate in which inter- based on the presence of handaxes, cleavers, knives, and picks (Fig. 1)
source variation must be greater than intra-source variation (Weigand that are collectively known as Large Cutting Tools (LCTs), which implies
et al., 1977). For this reason, archaeologists have invested considerable the ability to detach large flakes and the more pronounced imposition of
efforts to provenance obsidian stone tools because the sources of this form (Kuman, 2014).
raw material are far more homogeneous than other siliceous lithologies, From a behavioural standpoint, Acheulean sites can be differentiated
thereby allowing the establishment of characteristic fingerprints. How­ from most Oldowan sites based on their sheer size and artefact densities,
ever, there is only limited evidence for obsidian utilisation in Africa increased evidence for meat consumption, the sustained occupation of
predating the Middle Pleistocene (e.g. Gallotti and Mussi, 2015; Gowlett strategic localities in more diverse palaeoecological settings, as well as
and Crompton, 1994). In contrast, archaeologists have documented the the greater reliance on distantly sourced raw materials (de la Torre,
widespread use of quartz and quartzite by Oldowan and Acheulean 2016; Kuman, 2014). However, the comprehensive review of evidence
toolmakers during the Early and Middle Pleistocene across Eastern Af­ for Oldowan and Acheulean raw material provisioning in Eastern Africa
rica (e.g. Chavaillon, 1970; Leakey, 1971). However, far less energy has (Tables 1–2; Figs. 2 and 3) indicates more similarities than differences
been devoted to source these and other commonly exploited siliceous and four overarching patterns. First, ESA toolmakers usually selected
raw materials under the assumption that they are unsourceable at any rock types according to one or more factors, including morphometric
given spatial scale for three principal reasons: (1) elevated silica con­ characteristics, mechanical properties, functional suitability, and
centrations can muffle out unique elemental patterns; (2) metamorphic site-to-source distances, by way of opportunistic and least-effort pro­
and sedimentary rocks generally display greater elemental variation curement strategies to more specialised and energetically demanding
than volcanic lithologies; and (3) metamorphic and sedimentary out­ ones (e.g. Oldowan [Barsky et al., 2011; Braun et al., 2009a; Gold­
crops tend to cover larger areas than volcanoes and lava flows (Ebright, man-Neuman and Hovers, 2012; Gowlett et al., 2021; Harmand, 2009a;
1987; Favreau et al., 2020; Klein and Dutrow, 2008; Soto et al., 2020a). Stout et al., 2005]; Acheulean [Harmand, 2009b; Sánchez-Yustos et al.,
One can also point to the palimpsestic nature of the palae­ 2016; Merrick et al., 1994; Leakey et al., 1969; Howell et al., 1962]).
oanthropological record as a contributing factor to the lack of systematic Second, the fragmentation of technological activities along the
raw material studies. Be that as it may, the dynamic geological history of palaeolandscape manifested itself before the first appearance datum of
Eastern Africa led to the petrogenesis of various igneous, metamorphic, the Acheulean evidenced by raw materials, compositions, and attributes
and sedimentary rock types (Cahen et al., 1984), which have allowed for of some Oldowan assemblages, including, for example, Kanjera South
revelatory insights into the behaviour of early hominins regarding stone (Reeves et al., 2021), Ewass Oldupa (Cueva-Temprana et al., 2022), and
transport (e.g. Braun et al., 2008; Merrick et al., 1994). However, no DK (Reti, 2016) (Fig. 2). Simply stated, Oldowan hominins regularly
comprehensive reviews of this evidence have been undertaken to date, curated stone tools when ranging from raw material sources to occu­
which minimises our ability to better understand the adaptive signifi­ pation sites and beyond, thereby fragmenting their predominantly
cance of lithic transport in evolutionary terms. flake-centred chaîne opératoire prior to the emergence of the Acheulean,
With this preface in mind, the goal of this article is to review and when two concurrent yet divergent chaînes opératoires began to exist side
synthesise the aims, methods, challenges, and knowledge concerning by side. Importantly, the fragmentation of technological activities by
raw material sourcing from Plio-Pleistocene archaeological contexts in ESA toolmakers appears to have operated within a reduced spatial area
Eastern Africa. The first section comprises of a comprehensive review of in comparison to Middle Stone Age groups (cf. McBrearty and Brooks,
Oldowan and Acheulean raw material provisioning in Eastern Africa to 2000; Moutsiou, 2014).
identify the state of knowledge and methodological trends. The Third, early Oldowan hominins across Eastern Africa were frequently
following section reviews the theoretical and methodological un­ selective of raw material characteristics but rarely transported stone
derpinnings of raw material sourcing and delineates best practice more than a few kilometres to their final discard locations (e.g. Braun
guidelines. The third and fourth sections review analytical methods used et al., 2019; Harmand, 2009a; Stout et al., 2005). This is also supported
for raw material sourcing and ways by which multivariate data can be by evidence from the ~2-million-year-old Oldowan sites in South Afri­
analysed, respectively. The final section reviews common interpretive ca’s Cradle of Humankind World Heritage Site, including Swartkrans,
challenges when studying raw material provisioning in deep time. Sterkfontein, and Drimolen, in that raw materials were procured from
Overall, this article highlights how stone tools struck from variable rock local sources through some degree of purposeful selection (e.g. Kuman,

2
J. Favreau Quaternary Science Advances 9 (2023) 100068

Fig. 1. Oldowan and Acheulean artefact types from Oldupai Gorge (Tanzania) according to the traditional typological approach. (a) Hammerstone. (b) Anvil. (c)
Manuport. (d) Side chopper. (e) End chopper. (f) Core scraper. (g) Spheroid. (h) Polyhedron. (i) Discoid. (j–k) Awl. (l) Bifacial point. (m) Laterally trimmed flake.
(n–o) Notch. (p–q) Scraper. (r–s) Burin. (t) Punch. (u) Pièce esquillée. (v–w) Handaxe. (x–y) Cleaver. (z) Pick. Modified after Leakey (1971, 1979).

1994; Kuman and Field, 2009; Kuman et al., 2018; Stammers et al., provisioning strategies during the late Acheulean best evidenced by
2018). Over time, hominin toolmakers in Eastern Africa began to exert cosmogenic beryllium values of flint artefacts from Qesem Cave (Isreal)
stricter stone selection criteria, which occasionally entailed raw material indicating that hominins exploited underground sources (Verri et al.,
transport from more distant sources peaking at ~13 km for the Oldowan 2004, 2005). Overall, the bulk of the evidence currently indicates that
(Kanjera South) and ~60 km for the Acheulean (Isenya) (Figs. 2 and 3), early Acheulean toolmakers across Eastern Africa largely followed the
which constitute the longest verifiable distances for the ESA (Braun practices of their Oldowan counterparts and only amplified their stone
et al., 2008; Merrick et al., 1994). However, the initial shift from Old­ provisioning behaviour into something qualitatively different beginning
owan to Acheulean toolmaking was not accompanied by more complex in the late Early Pleistocene (~1 Ma). And as pointed out by others (e.g.
raw material procurement strategies despite the greater technological Carter, 2014; Féblot-Augustins, 1990), exotic raw materials from ESA
complexity inherent to the Acheulean. Instead, hominins placed more contexts are exceedingly rare. If taken further, this would suggest that
emphasis on knapping bigger blanks to accommodate the production of Oldowan and Acheulean toolmakers behaved differently than African
LCTs while continuing to manufacture smaller-sized core-flake in­ Middle Stone Age groups if lithic transport distances are understood to
dustries, which does not necessarily imply a higher degree of complexity reflect, among other things, the approximate foraging ranges and/or
nor longer transport distances as much as a shift in selective criteria. For social networks of prehistoric populations (see Brooks et al., 2018;
instance, all the early Acheulean sites in Ethiopia (Konso-Gardula, Melka McBrearty and Brooks, 2000; Moutsiou, 2014).
Wakena, Gona, Gadeb, and Melka Kunture) (see Beyene et al., 2013, The final point is the fact that both Oldowan and Acheulean tool­
2015; de la Torre, 2011; Gallotti and Mussi, 2017; Gossa and Hovers, makers across Eastern Africa preferentially utilised igneous rock types,
2022; Mussi et al., 2021; Semaw et al., 2018, 2020), Kenya (West Tur­ with an emphasis on extrusive varieties, followed by metamorphic
kana, Chesowanja, Koobi Fora, and Kilombe) (see Bishop et al., 1975; rocks, and rarely on sedimentary lithologies (Tables 1–2). This can be
Gowlett et al., 1981; Harmand, 2009b; Hoare et al., 2021; Isaac and attributed to two factors, namely the lithostratigraphic succession of
Isaac, 1997; Presnyakova et al., 2018; Texier, 2018), and Tanzania geological events that led to the surficial configuration of raw material
(Oldupai Gorge and Peninj) (see de la Torre et al., 2008; de la Torre and sources available during the Plio-Pleistocene, and more pertinently, the
Mora, 2018; Diez-Martín et al., 2014a, 2014b, 2015; Hay, 1976; Leakey, selection criteria employed by hominin toolmakers who apparently
1971; McHenry and de la Torre, 2018; Stiles, 1998) (Fig. 3) indicate that preferred not to use sedimentary rocks due to their generally inferior
raw material provisioning operated within the kilometric range of the mechanical and/or functional properties relative to igneous and meta­
Oldowan (Tables 1–2). It was only from the late Early Pleistocene (~1 morphic rock types, albeit with some notable exceptions (e.g. Braun
Ma) onwards that hominins engaged in increasingly complex raw ma­ et al., 2008, 2010; Roche et al., 1988; Stiles et al., 1974). However, not
terial procurement strategies best evidenced by the sites of Isenya every single extrusive igneous raw material was regularly used during
(>0.974 Ma), Kariandusi Upper Site (~0.970–0.780 Ma), and Gombore the ESA, as seen with the limited use of obsidian, only being recorded at
II OAM (~0.85 Ma) (Fig. 3), where maximum site-to-source distances 4/47 Oldowan sites (8.5%) (Table 1) and 22/83 Acheulean sites (26.5%)
are registered at ~60 km, ~15–30 km, and ~15–20 km, respectively (Table 2). When setting aside the exceptional case of Melka Kunture
(Gallotti and Mussi, 2017; Merrick et al., 1994). This higher degree of (Figs. 2 and 3), where ten of the obsidian-bearing sites are located, this
complexity is also potentially represented by select evidence from figure drops to only 6.4% for the Oldowan and 15.7% for the Acheulean.
DE/89B (~0.970–0.900 Ma), B11-500 (~0.615 Ma), A(W)12–10 This pattern cannot be attributed to the lack of obsidian sources avail­
(~0.609 Ma), and D14-10 (~0.499 Ma) at Olorgesailie (Isaac, 1977; able during Oldowan and Acheulean times (see Brown et al., 2013;
Potts et al., 2018), the Kariandusi Lower Site (~0.970–0.780 Ma) near Negash et al., 2020). Therefore, apart from Melka Kunture, alternative
Lake Elmenteita (Gowlett and Crompton, 1994), GqJh 1 (~0.988–0.982 plausible explanations include obsidian’s limited durability/functional
Ma) at Kilombe (Gowlett, 1978), Locality 8E (>0.7 Ma) at Gadeb (Clark suitability, or limited artefact curation, hominin mobility, and
and Kurashina, 1979), and Kaptabuya (~0.396–0.100 Ma) near Lake inter-group exchange networks.
Baringo (Merrick et al., 1994) (Fig. 3), but the relevant findings remain Overall, four important patterns are identified: (1) Oldowan and
to be definitively sourced and/or securely contextualised. Findings from Acheulean toolmakers regularly exerted selective criteria when
outside Eastern Africa equally attest to increasingly complex stone choosing raw materials; (2) the spatio-temporal fragmentation of

3
J. Favreau Quaternary Science Advances 9 (2023) 100068

Table 1
Eastern African Oldowan sites.
Site Age (Ma) Raw Materials Maximum Transport Distance Stratigraphic Unit/s Location Country

4
J. Favreau Quaternary Science Advances 9 (2023) 100068

Table 2
Eastern African Acheulean sites.
Site Age (Ma) Raw Materials Maximum Transport Distance Stratigraphic Unit Location Country

5
J. Favreau Quaternary Science Advances 9 (2023) 100068

Site Age (Ma) Raw Materials Maximum Transport Distance Stratigraphic Unit Location Country

technological activities across the palaeolandscape emerged before the sourcing, few systematic studies have been performed such that there is
first appearance datum of the Acheulean; (3) hominin toolmakers limited information to concretely sustain interpretations of lithic
preferentially utilised igneous rock types followed by metamorphic and transport from specific sources with a high degree of probability.
sedimentary lithologies mirroring Eastern Africa’s lithostratigraphic
sequence; and (4) Acheulean toolmakers largely mimicked their Old­ 3. Theoretical and methodological framework of raw material
owan counterparts in terms of raw material provisioning until the late sourcing
Early Pleistocene, when they began to engage in qualitatively different
behaviour best evidenced by anthropogenic stone transport over Rocks used to manufacture artefacts are known as lithic raw mate­
increasingly longer distances. Another informative pattern that emerges rials, whose heterogenous spatial distribution in the form of sources
from the foregoing review is that while macroscopic, petrographic, and functioned as foci for hominin toolmakers according to the foregoing
geochemical approaches have been implemented for raw material section. From a theoretical perspective, raw material sources can be

6
J. Favreau Quaternary Science Advances 9 (2023) 100068

Fig. 2. Eastern African Oldowan sites. (1) Bokol Dora


1. (2) EG-10. (3) EG-12. (4) EG-13. (5) OGS-6a. (6)
OGS-7. (7) DAN-1. (8) DAS-7. (9) DAN-2d. (10) Bouri.
(11) A.L. 894. (12) A.L. 666. (13) FtJi 1. (14) FtJi 2.
(15) FtJi 3. (16) FtJi 5. (17) Omo 1/E. (18) Omo 57.
(19) Omo 79. (20) Omo 123. (21) FJ-1A. (22) Garba
IVE-F. (23) Lokalalei 1. (24) Lokalalei 2C. (25) Koki­
selei 5. (26) Kokiselei 6. (27) Naiyena Engol 2. (28)
Kanjera South. (29) FwJj 20. (30) FxJj 1. (31) FxJj 3.
(32) FxJj 10. (33) FxJj 82. (34) GqJh 12A. (35) GqJh
12B. (36) GqJh 12C. (37) GqJh 13A. (38) GnJi 1. (39)
GnJi 2. (40) Ewass Oldupa. (41) Trench 168. (42) DK.
(43) FLK NN. (44) FLK Zinj. (45) DS. (46) HWK E. (47)
HWK EE.

conceptualised as primary, secondary, or tertiary. Primary sources smaller than inter-source variation (Weigand et al., 1977). If this
consist of in situ outcrops such as volcanic centres and inselbergs, postulate is not supported by analytical data, then the units of analysis
whereas secondary sources refer to clusters of rocks such as gravel bars (e.g. Munsell colours, modal mineralogies, elemental profiles, and/or
and conglomerates redeposited through geomorphological processes isotopic ratios) may be described as conceptually invalid. Analytical
(Shackley, 1998, 2008). Tertiary sources or stone caches refer to focal data may also be assessed through the lens of three additional concepts,
points on palaeolandscapes that contain anthropogenic accumulations namely accuracy (quantitative validity), precision (repeatability), and
of unmodified stones or manuports (McHenry and de la Torre, 2018), reproducibility (reliability) (Frahm, 2012). Accuracy can be defined as
and their recognition in the archaeological record usually prompts the the degree to which a value is close to the truth or to an internationally
interpretation that hominins provisioned localities with knappable recognised value. Precision refers to the degree to which different values
material with the forethought of reoccupation (Potts, 1988). Surface obtained under identical conditions are in close fit. Reproducibility
artefact scatters can also function as a raw material source, whereby describes the extent to which different values obtained under different
after initial discard the artefactual by-products from any reduction stage conditions are in close fit (Frahm, 2012). Overall, researchers are
may be recycled and undergo additional use-lives (Shott, 1989). In encouraged to diligently contemplate the aforementioned concepts with
Plio-Pleistocene archaeology, stone tools are usually compared with raw regards to sources and analytical data, and execute the following
materials from primary and secondary sources based on macroscopic, sequence of analysis when designing a raw material characterisation
mineralogical, and geochemical observations with the overarching aim project (Basu et al., 2016; Beller et al., 2016; Church, 1994; Garzanti,
of identifying from where and over what distances lithics were trans­ 2016; Glascock et al., 1998; Glascock and Neff, 2003; Henderson, 2000;
ported (Fig. 4). After accounting for raw material availability, abun­ Kooyman, 2000; McHenry and de la Torre, 2018; Shackley, 1998, 2008;
dance, extractive costs, energetics, and/or workability, among other Shotton and Hendry, 1979; Sunyer et al., 2017; Weigand et al., 1977):
factors (see Browne and Wilson, 2011; Wilson, 2007), the probabilistic
identification of sources and transport distances can serve as the foun­ 1. Establish a geological reference collection for the study area by
dation to infer other facets of hominin behaviour, including, for sampling the macroscopic variability, and the vertical and lateral
example, embedded procurement, foraging ranges, social connectivity, extents of sources to document the complete intra-/inter-source
and language abilities (Carter, 2014; Frahm et al., 2016, 2020; Marwick, variation at high spatial resolution. This sampling and mapping
2003; Moutsiou, 2014). strategy cannot account for primary/secondary sources that may
Whether geological specimens and stone tools are characterised with have been exhausted by anthropogenic behaviour (e.g. quarrying) or
the naked eye, the petrographic microscope, or an instrument designed lost/buried by natural processes. Since sedimentary strata may not
to identify geochemical concentrations, raw material sourcing is based reveal the complete array of sources, stone artefacts can complement
on the provenance postulate in which intra-source variation must be the geological reference collection, and even help to identify source

7
J. Favreau Quaternary Science Advances 9 (2023) 100068

Fig. 3. Eastern African Acheulean sites. (1) KGA6-A1.


(2) KGA4-A2. (3) KGA10-A11. (4) KGA10-A6. (5)
KGA7-A1. (6) KGA7-A3. (7) KGA7-A2. (8) KGA8-A1.
(9) KGA12-A1. (10) KGA18-A1. (11) KGA20-A1.
(12) KGA20-A2. (13) MW2. (14) MW5. (15) MW1.
(16) DAN-5. (17) OGS-12. (18) BSN-12. (19) Gadeb
2A, 2B, 2C, 2E. (20) Gadeb 8A, 8D, 8E. (21) Gombore
IB. (22) Garba IVD. (23) Gombore Iγ. (24) Gombore
Iδ. (25) Gombore II OAM Test Pit C. (26) Garba XIIJ.
(27) Garba XIIIB. (28) Simbiro III. (29) Gombore II
OAM. (30) Garba IB. (31) Kokiselei 4. (32) GnJi 10.
(33) FxJj 21. (34) FxJj 37. (35) FxJj 63. (36) FxJj 65.
(37) Isenya. (38) Kariandusi Upper Site. (39) Kar­
iandusi Lower Site. (40) GqJh 15A. (41) GqJh 1. (42)
DE/89B. (43) B11-500. (44) A(W)12–10. (45) D14-10.
(46) GnJh 10 (Leakey Hominid Site). (47) GnJh 17
(Kapthurin C). (48) GnJh 02 (Leakey Handaxe Area).
(49). GnJh 03 (Leakey Handaxe Rectangle). (50) GnJh
15 (Kapthurin A). (51) GnJi 28 (Rorop Lingop). (52)
GnJh 13 (Leakey Factory Site). (53) GnJi 16 (Kapta­
buya). (54) Lewa Downs. (55) FLK West. (56) EF-HR.
(57) TK. (58) BK. (59) BK East. (60) ST Site Complex.
(61) ES2-Lepolosi (MHS-Bayasi). (62) EN1-Noolchalai
(RHS-Mugulud). (63) South Pit. (64) H21. (65) H20.
(66) G23. (67) G17 and G18. (68) H15. (69) K13 and
K14. (70) K18. (71) K19. (72) J12. (73) K10. (74) K6.
(75) Lower J6-J7 and Upper J6-J7. (76) H9-J8. (77)
KK 51. (78) Mieso 7. (79) Mieso 31. (80) MK 4. (81)
MK 91. (82) MK 101. (83) MK 123.

Fig. 4. (a–d) Graphical representation of the raw material characterisation process based on hypothetical discriminatory observations to identify lithic transport
from primary sources assuming that geological processes did not redistribute rocks across the palaeolandscape. (d) Dashed lines represent a radiating resource
provisioning strategy for heuristic purposes.

subgroups and/or undocumented sources. Before sampling, existing geomorphological processes redeposited rocks on a broader spatial
geological data can be tested to establish proof of concept. extent. In this regard, modern analogues and geospatial modelling
2. Model the spatio-temporal availability of sources because past can, to a variable degree, offer valuable insights into localised rock
geological, geomorphological, and anthropogenic processes were cycles, source-to-sink processes, and hydraulic sorting, among other
variable across time and space. For example, the earlier a primary things.
source became available the greater the probability that

8
J. Favreau Quaternary Science Advances 9 (2023) 100068

3. Characterise the geological reference collection for proof of concept,


ideally in a non-destructive manner, and report baseline results. If
successful, repeat the characterisation method for stone artefacts for
systematic archaeological testing.
4. Analyse geological samples and stone artefacts more than once,
ideally on fresh and smooth surfaces, because they can exhibit
variation due to mineralogical/elemental concentrations (i.e. nugget
effect) and surface roughness can impact analytical readings,
respectively.
5. Evaluate geological samples and stone artefacts for physical,
mineralogical, and/or geochemical alterations because they can
distort the source signal. For instance, physical processes can nega­
tively affect less durable minerals and rocks, while geochemical
processes such as authigenesis (i.e. in situ formation of minerals) and
post-depositional leaching can lead to incompatible geological
reference collections and lithic assemblages.

4. Analytical methods for raw material sourcing

While the provenance of minerals and rocks was featured among


ancient textual sources (Rapp, 2009), the concept of raw material
sourcing can be traced back to the European antiquarian William Dug­
dale (1656:778), who suggested that a Neolithic axe from southern
England was not manufactured from local flint on the basis of macro­
scopic characteristics. In the following century, William Stukeley
(1740:5) described the mineralogy of a monolith sample from Stone­ Fig. 5. Radial Venn diagram showing the factors that affect the selection,
henge (England) and ascribed its possible source. Across the nineteenth implementation, and merit of analytical methods for raw material sourcing.
century, other scholars such as Friedemann Göbel (1842), Johann Adapted from Frahm (2012, 2013), Garrison (2003), Hancock and Carter
Erasmus Wocel (1854), Augustin Damour (1866), and Otto Helm (1886) (2010), and Wadley and Kempson (2011).
expanded upon the concept of sourcing on a range of artefactual mate­
rials (Pollard et al., 2014). However, it was not until the twentieth 4.1. Macroscopic characterisation
century that empirical approaches for raw material sourcing were
established following the petrographic studies carried out by Herbert Archaeologists have commonly relied on macroscopic criteria such
Henry Thomas (1923) at Stonehenge (Shotton and Hendry, 1979), and as colour, texture, lustre, and grain-size to identify rock types and
the creation of comprehensive geological reference collections (e.g. characterise raw material sources under the premise that these may be
Cogné and Giot, 1952; Keiller et al., 1941). indexical of individual sources. In the context of African Stone Age
Across the second half of the twentieth century, the discipline of archaeology, Bond (1948) published one of the continent’s earliest
archaeology witnessed profound theoretical and methodological de­ studies on raw materials using such an approach. Over the years, var­
velopments, including the emergence of processualism that emphasised iably detailed macroscopic analyses have been performed on Oldowan
adherence to the scientific method (Reimer, 2018; Trigger, 2006). With and Acheulean raw materials (e.g. Howell et al., 1962; McHenry and de
regards to methodological advancements for artefact sourcing, the la Torre, 2018; Stollhofen et al., 2021). However, it must be noted that
pene-contemporary development of X-ray spectroscopy led to some of artefactual materials are prone to alter through a variety of geogenic
the earliest publications on obsidian geochemical sourcing (e.g. Heizer processes and may not behave as closed systems such that apparent
et al., 1965; Parks and Tieh, 1966; Weaver and Stross, 1965). In a similar source indicators can potentially act as confounders (see Price and
fashion, the refinement of neutron activation analysis in concert with Velbel, 2003; Rollinson, 1993). In turn, this could put into question
the construction of nuclear reactors across the 1950–1960s allowed for provenance interpretations based on macroscopic criteria alone, espe­
increasingly precise compositional testing (Glascock and Neff, 2003). cially considering that chemical weathering has been regularly observed
Over the past decades, several other impactful analytical techniques on materials from various archaeological contexts (e.g. Gauthier and
have been implemented to source artefactual materials (e.g. Agha-Aligol Burke, 2011; Lundblad et al., 2008; McHenry and de la Torre, 2018;
et al., 2015; Cann and Renfrew, 1964; Chataigner et al., 2003; Gale, Mussi et al., 2021; Navazo et al., 2008; Stoner and Shaulis, 2021; Zaid
1981; Kasztovszky et al., 2008; Legg et al., 2020; Merrick and Brown, et al., 2015). Therefore, it is recommended that geogenic processes be
1984; Parish et al., 2013; Redmount and Morgenstein, 1996; Sciuto accounted for at a mineralogical or geochemical level to reaffirm the
et al., 2019). conceptual validity of sourcing based on macroscopic criteria. While
Notwithstanding the sustained methodological advancements, the there remains value in the macroscopic characterisation of raw mate­
probabilistic identification of sources is contingent on analytical data rials, the method’s subjectivity makes it prone to replicability issues as
that abide to the provenance postulate in which inter-source variation well as misclassifications meaning that source attributions should be
must be greater than intra-source variation (Weigand et al., 1977). considered hypothetical at best (Shackley, 1998, 2008; e.g. Biró, 1998;
Moreover, the selection, implementation, and merit of analytical Hermes et al., 2001).
methods hinges on their accuracy (quantitative validity), precision
(repeatability), reproducibility (reliability), discriminatory power
(conceptual validity), sensitivity, destructiveness, throughput, cost, 4.2. Petrography
ease, and the question of spatial scale (Frahm, 2012, 2013; Garrison,
2003; Hancock and Carter, 2010; Wadley and Kempson, 2011) (Fig. 5). Herbert Henry Thomas (1923) conducted seminal petrographic
As such, the goals of this section are to review methods and proxies that studies at Stonehenge to identify the probable source for some of the
have been, and could be, implemented for sourcing lithic raw materials sandstone monoliths, which marked the establishment of empirical ap­
from Plio-Pleistocene archaeological contexts in Eastern Africa. proaches in the archaeological subfield of raw material sourcing

9
J. Favreau Quaternary Science Advances 9 (2023) 100068

(Shotton and Hendry, 1979). Petrography involves the analysis of decommissioned can partly explain why this analytical method has not
~30–35 μm thin sections under transmitted light microscopy to identify been widely used in Plio-Pleistocene archaeological contexts. One may
materials based on their textural characteristics and non-opaque mineral also point out significant drawbacks, including its cost, throughput, and
composition (Perkins, 1998). This information can also reveal infor­ variably destructive nature as samples are usually powdered and
mation about petrogenesis, metamorphism, tectono-sedimentary set­ become radioactive after analysis. Overall, these limitations outweigh
tings, and weathering that can equally inform sourcing interpretations the benefits offered by NAA, particularly when considering that stone
(e.g. Zaid et al., 2015). In archaeology, the combination of macroscopic artefacts possess heritage value.
and petrographic approaches became prevalent in England starting in In recent years two forms of inductively coupled plasma mass spec­
the 1930s, when archaeologists began to establish raw material groups trometry (ICP-MS), laser ablation (LA-ICP-MS) and acid digestion (AD-
to source stone tools from Neolithic and Bronze Age contexts (e.g. Cogné ICP-MS), have largely overtaken NAA for raw material sourcing across
and Giot, 1952; Keiller et al., 1941). the discipline of archaeology primarily owing to their comparable
One of the earliest systematic accounts on raw materials in African detection capabilities and the global decommissioning of nuclear re­
Stone Age archaeology appeared in Hay’s (1976) monograph on the actors (Reimer, 2018; e.g. LeBlanc, 2004; Moreau et al., 2016). From a
geology of Oldupai Gorge in which macroscopic and petrographic data methodological standpoint, both forms of ICP-MS ionise, separate, and
were utilised for source characterisation. To this day, the combination of count a sample’s atoms according to their mass-to-charge ratios, with
macroscopic and petrographic techniques has proven useful to delineate the primary difference being whether a sample is introduced in liquid
similarities and differences among sources in the area (e.g. Favreau (AD-ICP-MS) or solid form (LA-ICP-MS). Therefore, the latter is gener­
et al., 2020; Soto et al., 2020a). Petrography has also allowed re­ ally favoured for sourcing because samples require minimal preparation
searchers to discriminate regional volcanic centres and lava flows in a and are not prone to contamination from reagents, analyses have rapid
variety of geological settings, including Oldupai Gorge, to establish the turnaround times, and laser ablation points/lines are only microscopi­
feasibility of lithic sourcing (e.g. Mollel, 2002). Equally important cally destructive (Neff, 2017). In a key study, Mollel (2002) securely
petrographic studies have been performed elsewhere in Eastern Africa, identified the sources of unmodified volcanic rocks excavated from Bed
including, for example, West Turkana (Harmand, 2009b), Melka Kun­ II sites at Oldupai Gorge using a combination of thin section petrog­
ture (Gallotti et al., 2010; Kieffer et al., 2004), Melka Wakena (Resom raphy, electron microprobe, XRF, and AD-ICP-MS despite the spatial
et al., 2018), and the Kapthurin Formation (Tryon, 2003) (Tables 1–2; proximity of volcanic centres that once drained into the palaeobasin
Figs. 2 and 3), which have enabled variably detailed insights into during the Pleistocene.
hominin raw material provisioning. But in the overall context of African Isotopes are variations of an element’s atom that have an identical
Stone Age archaeology, few petrographic studies have been conducted, number of protons but a different number of neutrons, which determines
thereby inhibiting the potential insights that may be gleaned from the whether they remain stable or become radioactive through time (Sulz­
method’s systematic application. This may be explained by key meth­ man, 2007). While isotopic analyses are commonly employed by
odological drawbacks, including destructiveness, throughput, and palaeoanthropologists to reconstruct diets, palaeoenvironmental con­
expertise-driven usage. ditions, and mobility patterns based on values obtained from fossilised
hard tissues and sediment (e.g. Lüdecke et al., 2016; Tucker et al., 2020;
4.3. Geochemical fingerprinting van der Merwe et al., 2008), they remain equally effective to source
lithic raw materials in a variety of contexts (e.g. Craig and Craig, 1972;
Geochemical fingerprinting refers to an array of analytical methods Herz, 1992). For instance, the lacustrine conditions of the Oldupai Gorge
used to identify the elemental and/or isotopic profiles of minerals and palaeobasin during the Early Pleistocene enabled the subaqueous pre­
rocks to differentiate raw material sources. Two methods that have been cipitation of bedded chert, which only became available to hominin
widely used for such purposes are energy-dispersive (ED-XRF) and toolmakers during lake lowstands (O’Neil and Hay, 1973). Considering
wavelength-dispersive (WD-XRF) X-ray fluorescence spectroscopy, that chert’s18O isotopic values are indicative of salinity levels during
which are particularly useful for identifying the concentrations of ele­ petrogenesis, Stiles et al. (1974) compared the 18O/16O ratios of chert
ments with atomic numbers greater than eleven (Rollinson, 1993). artefacts from two sites with unmodified in situ clasts to identify possible
While both techniques use X-ray photons to displace inner shell elec­ evidence of stone transport (but see Kimura, 1997). Another stable
trons of an atom’s nucleus, the detectors on WD-XRF systems have isotope that has allowed for sourcing lithic raw materials is 56Fe (e.g.
specialised crystals that diffract emitted X-rays of specific wavelengths, Mathur et al., 2020), but it remains to be applied in Eastern Africa for
which enhances accuracy at the expense of throughput (Shackley, Oldowan and Acheulean stone tools likely due to the need for compre­
2011). In fact, the most comprehensive studies on raw material sourcing hensive baseline studies, destructiveness, and the capabilities of alter­
from Plio-Pleistocene contexts in Eastern Africa have been performed nate methods.
using XRF (e.g. Braun et al., 2008; Braun et al., 2009b; Egeland et al., Radioactive isotopes are commonly utilised in palae­
2019; Finestone et al., 2020; Mercader et al., 2021; Negash et al., 2006; oanthropological research to constrain the age of fossilised hard tissues
Soto, 2020a-b). and stone artefacts (e.g. Leakey et al., 1961). One isotopic method that
Neutron activation analysis (NAA) is another important technique has been featured for raw material sourcing is fission-track dating,
through which archaeologists can identify source-specific geochemical which is based on uranium’s radioactive decay leading to the formation
fingerprints. This analytical method uses neutrons to irradiate a sample, and gradual shortening of detectable microscopic trails as a function of a
which causes radioactive nuclei to decay into daughter nuclei that emit rock’s age and thermal history (Enkelmann and Jonckheere, 2021). This
characteristic gamma rays (Glascock and Neff, 2003). NAA’s potential method has been applied to source obsidian artefacts throughout the
for sourcing artefactual materials was first realised by Robert Oppen­ Mediterranean basin and adjacent regions, East Asia, and South America
heimer and the earliest application of this method came with Sayre and (e.g. Bellot-Gurlet et al., 1999; Chataigner et al., 2003; Durrani et al.,
Dodson’s (1957) analysis of pottery from the Mediterranean region. 1971; Suzuki, 1969), but never in Eastern Africa for the ESA probably
Despite NAA’s proven potential, the technique has never been utilised to owing to the rarity of obsidian lithics, the capabilities of other analytical
source stone tools from Plio-Pleistocene contexts in Eastern Africa, and methods, costs, and destructiveness. Using a different isotopic system,
only once has it been featured for the study of artefacts from an equiv­ Vogel et al. (2006) identified the 40Ar/39Ar ratios of two obsidian
alent chrono-cultural context in continental Africa (Sampson, 2006). sources in Ethiopia in comparison to equivalent data from obsidian stone
While the sensitivity, accuracy, and precision offered by NAA is broadly tools. Since this dating technique reflects the cooling ages of rocks, it can
comparable to XRF (Hancock and Carter, 2010), the fact that nuclear prove useful when volcanic sources are tightly constrained in age and/or
reactors outfitted with NAA capabilities have been gradually exhibit similar elemental concentrations. For optimal results, the

10
J. Favreau Quaternary Science Advances 9 (2023) 100068

complete age range of volcanic centres/flows must be identified as it can In a different study, Nazaroff et al. (2013) combined macroscopic
bear important implications for the maximum and minimum ages of criteria with ED-XRF spectroscopy to distinguish chert sources and
occupation sites (Ambrose, 2012). As a case in point, Morgan et al. recognise provisioning patterns for the Neolithic site of Çatalhöyük
(2009) applied the 40Ar/39Ar method on obsidian lithics in concert with (Turkey). In another case, Gurova et al. (2016) integrated macroscopic
other means to establish the maximum ages of two Ethiopian sites, while criteria, petrographic observations, and LA-ICP-MS to characterise
electron microprobe and ED-XRF were used to probabilistically prove­ Cretaceous flint-bearing deposits and provisionally identity the sources
nance some artefacts. Other radioactive isotopic ratios that have facili­ of artefactual materials from sites of varying chrono-cultural affiliations
tated lithic raw material sourcing include 87Sr/86Sr (e.g. Brilli et al., in Bulgaria and Serbia. In a similar fashion, Skarpelis et al. (2017)
2005; Curran et al., 2001; Gale, 1981), but this method has never been detected potentially distinctive characteristics for a prominent chert
applied in ESA contexts in Eastern African likely owing to the need for outcrop on the island of Naxos (Greece) through macroscopic analysis,
comprehensive reference data, costs, destructiveness, and the potential petrography, scanning electron microscopy, XRF, and AD-ICP-MS. More
of alternate methods. recently, Brandl et al. (2018) blended visual criteria, various forms of
microscopy, and trace element analysis using LA-ICP-MS to positively
4.4. Discussion identify the source of flint-bearing limestone ballast recovered from a
shipwreck off the southern coast of Norway. In light of these case studies
The macroscopic characterisation of raw materials can prove useful and many others on a global scale (e.g. Gall and Steponaitis, 2001;
in certain contexts, but it is not recommended as a stand-alone method Harrell, 1992; Middleton and Bradley, 1989; Nash et al., 2013; Pitblado
for sourcing because rocks may be misclassified and incorrectly attrib­ et al., 2008; Prieto et al., 2022; Soto et al. 2020a-b; Stross et al., 1988), it
uted to sources (Shackley, 1998, 2008). The insights that can be gleaned can be stated that the future of sourcing non-obsidian lithologies is best
from thin section petrography make this approach useful for rock type accomplished when a multi-analytical approach is systematically
identification and source characterisation because it is conceptually applied, with the understanding that once positive baseline results are
valid, precise, reproducible, and low-cost, but it remains obtained subsequent archaeological testing can be narrowed down from
time-consuming and destructive. The implementation of XRF, NAA, and a methodological standpoint.
ICP-MS for bulk and spot geochemical fingerprinting have proven
invaluable for sourcing igneous, metamorphic, and sedimentary rocks 5. Multivariate data
from a variety of archaeological contexts because macroscopic criteria,
modal mineralogies, and petrographic textures may not necessarily Securely sourcing raw materials not only hinges on appropriate
show discriminatory differences as opposed to elemental concentrations analytical methods, but also on the implementation of effective means
and isotopic ratios. In particular, ED-XRF and LA-ICP-MS can both through which the resulting data are evaluated. While analytical data do
rapidly yield accurate, precise, and reproducible data from artefact not necessarily need to be accurate for raw material sourcing to be
(near-)surfaces with little to no destruction. conceptually valid (see Frahm, 2013), it is recommended that they are
Isotopic systems can be useful for lithic sourcing, but physical/ mutually accurate, precise, and reproducible. The latter constitutes an
chemical processes may distort source signals and isotopic techniques important facet of contemporary archaeological research (e.g. Marwick,
are partially to fully destructive (see Stephens et al., 2021). The chro­ 2017; Marwick and Birch, 2018), particularly in the context of raw
nometric dating of rocks and minerals in particular has proven to be an material studies considering their fundamental importance, the large
effective means to source stone tools (e.g. Bevins et al., 2020; Khatse­ amount of data generally produced, the costs associated with such lines
novich et al., 2020; Vogel et al., 2006). If taken further, such approaches of research, and the fact that interpretation usually rests on statistical
could offer profound insights in stone transport behaviour within and results obtained through computational means (e.g. Hermann et al.,
between subregions of Eastern Africa considering the known age con­ 2020). As the foregoing section highlighted the value of non-destructive
straints of Precambrian rocks, Cenozoic magmatism, and rifting along geochemical fingerprinting and the implementation of multi-analytical
the north-south axis of the East African Rift System, where most approaches, this section will consider how multivariate data can be
Plio-Pleistocene sites are found (see Foley, 2018; Fritz et al., 2013; transformed, explored, and statistically analysed to discriminate raw
Macgregor, 2015; Rooney, 2017, 2020a-c; Schlüter, 1997) (Figs. 2 and material sources.
3). However, this would necessitate the establishment of robust refer­
ence collections because the eastern half of the continent experienced
severe tectono-magmatic stress throughout its geological history that 5.1. Data transformation
could have overprinted and/or reset isotopic clocks. Finally, both stable
and radiometric isotope traceability indices may manifest themselves at Beginning with data handling, geochemical profiles may be stand­
overly expansive spatial scales that could be irrelevant for research ardised, normalised, or logarithmically transformed to correct for het­
projects, unless regional connections are suspected and/or occupation eroscedasticity and non-normality (Baxter, 1995). Standardisation
sites are situated along lithostratigraphic margins. As such, the spatial consists of modifying raw values by subtracting them by the mean of
scale of the research project plays an equally important role in the se­ individual elements and dividing them by the standard deviation.
lection, implementation, and merit of any analytical method through Similarly, normalisation consists of scaling the units of analysis to
which stone tools can be sourced. ensure that they are comparable (e.g. Mercader et al., 2021). Alterna­
Archaeologists working in a variety of geological contexts have held tively, geochemical data can be logarithmically transformed to stabilise
long-standing concerns to establish the sources of non-obsidian rock variance and reweigh variables to approach a normal distribution since
types through multiple means (e.g. Blomme et al., 2012; Cnudde et al., many statistical tests follow the assumption of normality (e.g. Braun
2013; de Boutray, 1981). As a case in point, Julig et al. (1998) integrated et al., 2008). When datasets do not present outliers, statistical analyses
XRF, NAA, petrography, and cathodoluminescence microscopy to of normalised and logarithmically transformed data produce compara­
characterise siliceous source material and securely provenance stone ble results (Baxter, 1995). If applicable, elements that are below the
tools from archaeological sites across parts of Michigan and Wisconsin detection limit of the analytical instrument can be assigned arbitrary
(United States) as well as Ontario (Canada). Elsewhere in North Amer­ values, zeroed, or omitted (cf. Mauran et al., 2021; Habermann et al.,
ica, ten Bruggencate et al. (2013, 2014) relied on macroscopic colour 2016; Kiehn et al., 2007; Nash et al., 2013). Overall, these trans­
identification followed by trace element geochemistry and Pb isotope formation methods are not always necessary but can help to circumvent
analysis to discriminate sources of pegmatite quartz and probabilisti­ the constant sum problem that can drive spurious correlations (Roll­
cally source artefacts in northern Manitoba and Saskatchewan (Canada). inson, 1993).

11
J. Favreau Quaternary Science Advances 9 (2023) 100068

5.2. Data exploration more broadly, regardless of rock type. In addition, there are emerging
and innovative data evaluation techniques that fall under the framework
The second step in multivariate data analysis usually consists of of data science and artificial intelligence (e.g. Bolton et al., 2020;
iterative exploration by way of histograms, boxplots, biplots, and scat­ Hazenfratz et al., 2017; Lindsay et al., 2021; Lowe et al., 2017), which
terplot matrices to establish the dimensions of the dataset, identify and hold promise when dealing with large and geographically expansive
assess outliers, ensure there are no errors or missing values, visualise and datasets.
establish preliminary relationships among variables, and ensure the
dataset meets the assumptions of any further statistical tests (Glascock 6. Interpretive challenges of raw material studies
and Neff, 2003). Analyses of variance and t-tests can be performed to
identify relationships between outcrops and macroscopic, petrographic, Despite the fundamental importance of information derived from
and/or geochemical observations (Drennan, 2009). Similarly, logistic raw material studies in Plio-Pleistocene archaeological contexts, rela­
regressions can be conducted to predict from which of two sources tively few feature the implementation of one or a combination of ap­
geological samples and artefacts originate from. Correlation matrices proaches to identify aspects such as hominin landuse, source locations,
can help to uncover relationships among variables, and correlograms and/or selection criteria at the interface with lithic technology. This
can be used to visualise positive and negative correlations as well as research gap may be attributed to a number of factors, including the
their statistical significance. Analyses of covariance can be performed if complexity of drivers conditioning raw material configurations on the
significant correlations are found among two continuous variables and palaeolandscape, site formation processes resulting in a fragmentary
one categorical variable. Alternatively, multivariate analyses of vari­ and palimpsestic archaeological record, decreased chronological reso­
ance can be conducted to determine if there are statistically significant lution with increased age, complications in differentiating among pri­
differences among sources that may not be apparent by inspecting mary, secondary, and tertiary sources, the challenges with nullifying the
graphical data representations alone. In summary, there are several possibility of lithic recycling, as well as the difficulty in identifying stone
means to evaluate and explore multivariate data that can offer valuable selection criteria. As such, the aims of this section are to review common
insights regarding correlations along with preliminary inter-/in­ interpretive challenges and corresponding solutions when studying
tra-source patterns. hominin raw material provisioning in deep time.

5.3. Data analysis 6.1. Geographic space and hominin mobility

It is often necessary to conduct statistical analyses to discriminate The spatial component of hominin behaviour that preceded and
raw material sources, particularly for non-obsidian rock types. Three drove the production and utilisation of stone tools can be challenging to
common approaches through which this can be achieved include cluster discern but may be reasonably inferred through the judicious use of
analysis, principal component analysis, and linear discriminant analysis spatial modelling and middle-range theory. Distance and travel time
(Baxter, 2001). Cluster analysis refers to a family of unsupervised ma­ represent two means by which proximity to sources can be measured
chine learning techniques that rely on different algorithms to partition and estimated. For instance, the distance between an archaeological site
sample populations into distinct groups with the goals of minimising relative to a raw material source can be calculated in two-dimensional
intra-group variance and maximising inter-group variance (Borcard Euclidean space and more precise estimates can be generated by
et al., 2018; Lowe et al., 2017; e.g. Dalpra and Pitblado, 2016; Pitblado means of palaeosurface projections (e.g. Blumenschine et al., 2008;
et al., 2013). Santonja et al., 2014). Researchers can also estimate travel time using
Principal component analysis represents a conventional approach values derived from functional anatomical studies (e.g. Molen et al.,
used for source discrimination using multivariate data (e.g. Glascock 1972; O’Neill et al., 2015) or by relying on mathematical formulas (e.g.
et al., 2007 and chapters therein). It effectively reduces the dimen­ Tobler’s hiking function, Naismith’s rule) to obtain walking speeds for
sionality of multivariate data, identifies linear combinations of variables average adult humans on flat and variable terrain. In other cases,
that explain the most variance, and expresses data in graphical space in ethnographic sources can be used to estimate prehistoric foraging radii
which variance is maximised and structure is preserved (Borcard et al., of hunter-gatherers to delineate local versus non-local raw material
2018). However, one of the weaknesses of this approach is that a priori sources (e.g. Ekshtain et al., 2014; Zaidner et al., 2021; but see Pop,
groups are unknown, meaning that geological samples from the same 2016). In any case, Pollard et al. (2014) have rightfully noted that travel
sources may not cluster in graphical space. time remains challenging to model in prehistoric contexts due to the
Another technique that is regularly used for raw material sourcing is chronological imprecision and palimpsestic nature of the archaeological
linear discriminant analysis, which falls under the framework of su­ record, which, in the absence of contextual information, can prevent
pervised machine learning (Lowe et al., 2017). This approach is useful nullifying the hypothesis that evidence of raw material transport rep­
for dimensionality reduction and the classification of artefacts into resents a time-averaged behavioural pattern.
predefined groups consisting of geological samples from known sources Be that as it may, geospatial modelling can also be an important
(e.g. Prieto et al., 2022). This approach is similar to principal component source of knowledge in the absence of detailed information on the dis­
analysis, but the primary difference is that linear combinations of vari­ tribution of raw material sources in any given study area (e.g. Clarkson
ables adhere to and enhance variance among predefined groups. For and Bellas, 2014). In the right archaeological context, one could even
accurate results, groups should be of equal size and the number of identify potential routes that facilitated the movement of material goods
predictor variables should not exceed the number of samples from the from one location to another and the associated travel costs when
smallest group (Baxter, 1994; Harbottle, 1976; Kovarovic et al., 2011). considering the anisotropic nature of past and present landscapes (e.g.
Statistical analyses can also be performed on sub-groups of geolog­ Lucero et al., 2021; Moutsiou and Agapiou, 2019). In this regard,
ical samples and artefacts defined by distinct combinations of variables modelling the geomorphology of contemporary landscapes can help to
(e.g. ten Bruggencate et al., 2013, 2014). The principal reason one may reconstruct possible configurations of raw material sources and catch­
adopt such an approach is that by increasing the size of a geological ments across palaeolandscapes at different time slices through spatial
reference collection and comparing, for example, geochemical values interpolation (e.g. Benito-Calvo et al., 2008; Flores-Prieto et al., 2015;
regardless of a sample’s macroscopic classification, it can drastically Hensel et al., 2019; Leverington et al., 2002). In turn, this could allow for
reduce the ability to probabilistically assign sources. Overall, there are better estimates of stone transport distances from primary and second­
several approaches that can allow for the effective discrimination of raw ary sources in Plio-Pleistocene archaeological contexts. While all the
material sources using geochemical fingerprints, and multivariate data foregoing concepts can serve as effective theoretical underpinnings for

12
J. Favreau Quaternary Science Advances 9 (2023) 100068

archaeological research, any investigation of hominin mobility at the challenging to discern whether artefacts were manufactured from a
interface with geographic space according to stone tools and raw ma­ primary or secondary raw material source. A proven approach is to study
terials is contingent on assumptions, analogies, and high-quality data. an artefact’s size, morphology, cortex, neocortex, and abrasion state
along with site formation processes (e.g. Fernandes, 2012). However,
6.2. Stone transport and time-averaging this approach’s interpretive remit is contingent on the preservation of
key lithic attributes, the accessibility/visibility of secondary sources (e.
In Plio-Pleistocene archaeology, site-to-source transport distances g. Goldman-Neuman and Hovers, 2012; Harmand, 2009a, 2009b; Roche
are usually interpreted as singular and unidirectional behavioural epi­ et al., 2018; Stout et al., 2005), and the nature of the archaeological site.
sodes. However, this may not always be a sound interpretation when The importance of the latter cannot be understated in the context of
considering the palimpsest nature of the archaeological record (Bailey, Plio-Pleistocene archaeology since revisionist studies have demon­
2007; Dibble et al., 2017; Stern, 1994). For instance, Blumenschine et al. strated that many lithic and faunal assemblages are, to varying extents,
(2008) tested a distance-decay model for quartzite artefacts from Old­ palimpsest accumulations (cf. Domínguez-Rodrigo et al., 2007; Leakey,
owan assemblages relative to a primary raw material source according 1971).
to macroscopic similarities at Oldupai Gorge (Fig. 2). They identified a From a historical perspective, the recognition that lithic and faunal
correlation between site-to-source distance and artefact weight, while assemblages were not pristine anthropogenic accumulations resulted in
artefact sizes and reduction attributes were not correlated to distance. a shift from site-specific investigations towards a greater emphasis on
Therefore, the authors suggested that other variables must have affected the palaeolandscape (e.g. Isaac, 1981). Within the context of raw ma­
the reduction, utilisation, and discard patterns of quartzite tools along terial studies, this led to the development of one iconic theory known as
the palaeolandscape. While this remains plausible, it must be acknowl­ the stone cache model, which posits that hominins provisioned focal
edged that one of the study’s underlying assumptions was that each points with unmodified stones or manuports with the forethought of
assemblage represented a contemporary analytical unit. reoccupation (Potts, 1988). While the establishment of tertiary sources
In a comparable study, Luncz et al. (2016) tested a distance-decay remains plausible (e.g. Howell et al., 1962; Semaw et al., 2020), it must
model for the distribution of chimpanzee hammerstones relative to be acknowledged that large clasts can naturally accumulate alongside
raw material sources in the Taï National Park (Côte d’Ivoire) (cf. Boesch artefacts in low-energy lacustrine environments and fluvial settings.
and Boesch, 1984). Hammerstone weights and use-wear patterns abided Often, this stands as a simpler and occasionally better explanation for
to the model’s predictions, whereby weight decreased and use-wear the co-occurrence of unmodified stones and lithics at some Oldowan
increased with growing site-to-source distance. However, the authors sites in particular (cf. de la Torre and Mora, 2005; Leakey, 1971). If it is
recognised that chimpanzees predominantly carried stones over short possible to rule out the exploitation of secondary and tertiary sources,
distances. Therefore, the distribution and utilisation of hammerstones then it becomes interpretatively sound to begin the identification pro­
reflects a time-averaged behavioural pattern (Luncz et al., 2016). This cess of primary sources assuming they are accessible/visible, there are
middle-range model highlights the need for archaeological in­ discriminatory differences among them, and that artefactual materials
terpretations to be tightly integrated with an understanding of site for­ plot within the range of variation at the intra-source scale.
mation processes and techno-typological data because inferred
site-to-source transport distances could represent cumulative events 6.5. Selection criteria
rather than singular behavioural episodes.
Another challenge facing raw material studies is the range of possible
6.3. Recycling selection criteria imposed by hominin toolmakers (e.g. site-to-source
distance, fracture predictability, angularity, edge durability, hardness,
Another challenge facing raw material studies is identifying whether colour, etc.), which partly explains the frequent focus on functional and
hominins followed a linear artefact production process or recycled economic characteristics as they are more readily discernible. For
pieces from surface scatters. As a case in point, Marder et al. (2006) instance, Key et al. (2020) conducted experimental cutting tests using
studied attributes of Acheulean bifaces from two Middle Pleistocene raw materials that were exploited by Pleistocene hominins at Oldupai
levels at Revadim Quarry (Israel). The older bifaces were primarily Gorge (Figs. 1 and 2) under the premise that edge sharpness and dura­
manufactured from flint, which may have been procured from fluvial bility guided stone selection to produce flakes and LCTs (Kuman, 2014;
deposits ~3–5 km away, and were characterised as thick, intensively Toth, 1985). More specifically, they found that quartzite flakes have
retouched, and symmetrical. The younger bifaces were manufactured sharper but less durable edges than chert, while basalt flake edges have a
from comparable rock types, but they exhibited double-patinated flaked lower initial sharpness but are more durable with sustained use.
surfaces and were visibly shorter and thinner. Therefore, it was argued Therefore, Key et al. (2020) suggested that both Oldowan and Acheulean
that the younger bifaces were recycled pieces that may have functioned toolmakers primarily selected raw materials to manufacture function­
as cores (Marder et al., 2006). ally optimised cutting tools to maximize energy return rates.
The foregoing case study, among others from the Levant (e.g. Agam In contrast to the foregoing study, Stout et al. (2005) visually char­
et al., 2015; Shimelmitz, 2015; Wojtczak, 2015), illustrates the fact that acterised Oldowan lithics and clasts sampled from coeval conglomerates
hominins occasionally recycled lithics, which calls for careful attribute at Gona (Table 1; Fig. 2) to identify qualitative variables that may have
analysis to demonstrate whether raw materials were procured from been relevant to early toolmakers. The middle-range hypothesis is based
geological sources or surface scatters. However, the extent to which on the observation that modern flintknappers usually spend consider­
Oldowan and Acheulean toolmakers across Eastern Africa repurposed able time and effort identifying suitable raw materials as part of the tool
artefacts from surface scatters is unaccounted for and probably under­ manufacturing process. It was demonstrated that Oldowan hominins
estimated (see Amick, 2015) because the primary means through which preferentially selected glassy, smooth, and fine-grained igneous lithol­
it can be identified (i.e. double patination) is characteristic of arid re­ ogies with small phenocrysts, which may have facilitated tool produc­
gions (Macholdt et al., 2017) and is commonly reported on rock types (e. tion (Stout et al., 2005). Following a similar methodology,
g. McDonald, 1991 [chert]) that were rarely utilised by early hominins Goldman-Neuman and Hovers (2012) found differences in the selection
in Eastern Africa (Tables 1–2). strategies for two Oldowan sites at Hadar (Table 1; Fig. 2). For one site, it
was argued that hominins purposefully sought out high-quality mate­
6.4. Primary, secondary, and tertiary sources rials with fine-grained textures and high fracture predictability, whereas
at the other toolmakers apparently selected more homogenous but
Even when lithic recycling is nullified or indeterminate, it can be lower-quality rocks (Goldman-Neuman and Hovers, 2012). In another

13
J. Favreau Quaternary Science Advances 9 (2023) 100068

comparative study of two Oldowan sites at West Turkana (Table 1; review the theoretical and methodological underpinnings of raw mate­
Fig. 2), Harmand (2009a) found that hominins at one site preferentially rial sourcing followed by common interpretive challenges.
selected angular medium-grained phonolites over rounded clasts likely On a fundamental level, the identification of raw material sources is
due to their natural striking platforms and better fracture predictability. contingent on the implementation of a systematic, comprehensive, and
For the other locality, it was argued that the stone selection criteria were multifactorial sequence of analysis. Moreover, sourcing is a probabilistic
less strict evidenced by an abundance of rounded clasts and lower exercise because archaeological interpretations hinge on equally accu­
technical elaboration (Harmand, 2009a). rate, precise, and reproducible data that abide to the provenance
Another means by which researchers can identify the selection postulate. From a methodological standpoint, several factors affect the
criteria of hominin toolmakers is through the comparison of available selection, implementation, and merit of analytical methods through
lithologies at sources in contrast to the assemblage’s raw material di­ which stone artefacts, or any other material, can be confidently prove­
versity. For instance, studies on the earliest known Oldowan site of nanced. These include accuracy, precision, reproducibility, discrimina­
Bokol Dora 1 (Table 1; Fig. 2) found that hominins preferentially utilised tory power, sensitivity, destructiveness, throughput, cost, ease, and the
fine-grained rhyolite from nearby conglomerate beds and that the question of spatial scale (Frahm, 2012, 2013; Garrison, 2003; Hancock
assemblage was over-represented by rock types relative to their natural and Carter, 2010; Wadley and Kempson, 2011). Based on a series of
frequencies at sources (Braun et al., 2019). Similarly, raw material cost-benefit considerations, the array of analytical methods that fall
studies at Omo (Table 1; Fig. 2) have shown that hominins purposefully under the umbrella of geochemical fingerprinting generally outperforms
selected quartz from a limited array of naturally available rock types macroscopic and petrographic approaches since they can yield accurate,
relative to their frequencies at secondary sources (Delagnes et al., 2011). precise, and reproducible data with little to no destructive effects on
Using a combination of approaches, Braun et al. (2009b) characterised artefactual materials with rapid turnaround times. There also exist
the fracture predictability and durability of rock types represented in the several other approaches dealing with isotopic systems that have yet to
Oldowan assemblage from Kanjera South (Table 1; Fig. 2). It was shown be applied in the context of Plio-Pleistocene archaeology in Eastern
that more durable lithologies were selected over others despite their less Africa, but which hold promise for large-scale projects considering the
frequent occurrence in conglomerates, and that rebound hardness or well-documented and spatio-temporally constrained pulses of geological
fracture predictability did not factor as heavily in stone selection (Braun activity across the eastern half of continental Africa (e.g. Fritz et al.,
et al., 2009b). Stated otherwise, hominins regularly transported 2013; Rooney, 2017, 2020a-c). Regardless of the method of choice,
abrasion-resistant raw materials such as granites and quartzites at a characterising non-obsidian lithologies is best accomplished using a
higher frequency than their natural abundance on the palaeolandscape, multi-analytical approach in light of the growing body of
and they also invested more effort in the reduction and technical elab­ proof-of-concept studies and systematic archaeological testing (e.g.
oration of materials from more distant sources (Reeves et al., 2021). Brandl et al., 2018; Gurova et al., 2016; Julig et al., 1998; Nazaroff et al.,
Altogether, the foregoing case studies on the identification of selection 2013; ten Bruggencate et al., 2014). Once positive baseline results are
criteria by ESA toolmakers reveal the range of possibilities through obtained through multiple means, subsequent archaeological testing can
which this can be achieved, which can serve as a baseline to investigate be narrowed down methodologically.
other factors that could have equally affected hominin behaviour (e.g. Since raw material characterisation studies usually results in the
Browne and Wilson, 2011; Wilson, 2007). production of multivariate data, this warrants thorough evaluation and
reporting that are best performed following the principles of open and
7. Conclusions reproducible science. While data transformations are not always
necessary, iterative data exploration and statistical analyses are often
Eastern Africa’s Plio-Pleistocene palaeoanthropological record has required to discriminate non-obsidian sources and probabilistically
shaped our understanding of human biological and cultural evolution. source stone tools with a high degree of certainty. Ranging from tradi­
Despite being incomplete, palimpsestic, spatially biased, and impre­ tional graphical data representations to multivariate statistics and
cisely dated, this record offers an exceptional window into the human emerging branches of artificial intelligence (e.g. Glascock et al., 2007;
evolutionary past in the form of stone tools. Built upon generations of Lowe et al., 2017), there are countless means through which charac­
research, raw material studies represent an important component of teristic source signals can be identified regardless of rock type. Moving
contemporary lithic analysis through which it is possible to identify raw from data analysis to archaeological interpretation, the case studies
material sources and stone transport distances across palaeolandscapes reviewed herein on mobility, time-averaging, recycling, source differ­
to then reconstruct other behavioural aspects that are of high evolu­ entiation, and selection criteria pose challenges that are particularly
tionary significance. Several key behavioural trends were identified in relevant to the study of hominin raw material provisioning in deep time.
reviewing evidence of raw material provisioning during the ESA These interpretive challenges can also entail problems of replicability
(Tables 1–2). More specifically, it was established that both Oldowan when considering the assumptions, geological backdrop, methodolog­
and Acheulean toolmakers regularly exerted selective criteria when ical contingencies, and/or variables employed in any given study.
choosing raw materials, fragmented their activities along the palae­ Altogether, these challenges help to explain the limited efforts that have
olandscape, and preferentially utilised igneous rock types followed by been historically devoted in sourcing lithic raw materials of
metamorphic and sedimentary lithologies (e.g. Goldman-Neuman and Plio-Pleistocene age. With this being said, there are a multiplicity of
Hovers, 2012; Harmand, 2009b; Howell et al., 1962; Stout et al., 2005). means to securely infer aspects of hominin behaviour beyond resource
Finally, it was found that Acheulean toolmakers largely mimicked their extraction points and transport distances when raw material analyses
Oldowan counterparts in terms of raw material provisioning until the are tightly integrated with an understanding of site formation processes,
late Early Pleistocene, when they began to engage in qualitatively chronology, and information derived from lithic techno-typological
different behaviour best evidenced by the purposeful transport of stone studies. Overall, this calls for a structured multidisciplinary frame­
over increasingly longer distances (cf. Braun et al., 2008; Egeland et al., work that together with continual methodological advances will entail
2019; Gallotti and Mussi, 2017; Harmand, 2009a, 2009b; Merrick et al., the ability of future researchers to best surmount these common inter­
1994). Another informative pattern that emerges from this compre­ pretive challenges. Ultimately, the future of sourcing raw materials from
hensive review is that while macroscopic, petrographic, and geochem­ Eastern Africa’s Plio-Pleistocene palaeoanthropological record is bright
ical approaches have been implemented for raw material sourcing, few considering what is currently known along with the robust theoretical
systematic studies have been performed such that there is limited in­ and methodological foundations that are in place so as to garner the full
formation to concretely sustain interpretations of lithic transport with a potential of a non-renewable record.
high degree of probability. This trend served as a stepping-stone to

14
J. Favreau Quaternary Science Advances 9 (2023) 100068

Author’s note Baxter, M.J., 1995. Standardization and transformation in principal component analysis,
with applications to archaeometry. Applied Statistics 44, 513–527. https://doi.org/
10.2307/2986142.
A preprint of this article is available with the Open Science Frame­ Baxter, M.J., 2001. Multivariate analysis in archaeology. In: Brothwell, D.R., Collard, A.
work (https://osf.io/nf27g). M. (Eds.), Handbook of Archaeological Sciences. Wiley, London, pp. 685–694.
Beller, J.A., Greenfield, H.J., Fayek, M., Shai, I., Maeir, A.M., 2016. Provenance and
exchange of basalt grinding stones of EB III Tell es-Safi/Gath, Israel. J. Archaeol. Sci.:
Declaration of competing interest Reports 9, 226–237. https://doi.org/10.1016/j.jasrep.2016.07.025.
Bellot-Gurlet, L., Bigazzi, G., Dorighel, O., Oddone, M., Poupeau, G., Yegingil, Z., 1999.
The fission-track analysis: an alternative technique for provenance studies of
The authors declare that they have no known competing financial prehistoric obsidian artefacts. Radiat. Meas. 31, 639–644. https://doi.org/10.1016/
interests or personal relationships that could have appeared to influence S1350-4487(99)00157-2.
the work reported in this paper. Benito-Calvo, A., Pérez-González, A., Parés, J.M., 2008. Quantitative reconstruction of
late cenozoic landscapes: a case study in the sierra de Atapuerca (burgos, Spain).
Earth Surf. Process. Landforms 33, 196–208. https://doi.org/10.1002/esp.1534.
Data availability Bevins, R.E., Pirrie, D., Ixer, R.A., O’Brien, H., Pearson, M.P., Power, M.R., Shail, R.K.,
2020. Constraining the provenance of the Stonehenge ‘Altar Stone’: evidence from
automated mineralogy and U–Pb zircon age dating. J. Archaeol. Sci. 120, 105188
All data are included in the manuscript. https://doi.org/10.1016/j.jas.2020.105188.
Beyene, Y., Katoh, S., WoldeGabriel, G., Hart, W.K., Uto, K., Sudo, M., Kondo, M.,
Acknowledgements Hyodo, M., Renne, P.R., Suwa, G., Asfaw, B., 2013. The characteristics and
chronology of the earliest Acheulean at Konso, Ethiopia. Proc. Natl. Acad. Sci. USA
110, 1584–1591. https://doi.org/10.1073/pnas.1221285110.
First and foremost, I wish to acknowledge the terrific level of support Beyene, Y., Asfaw, B., Sano, K., Suwa, G., 2015. Konso-Gardula Research Project. In:
and mentorship I received from my doctoral supervisory committee, Archaeological Collections: Background and the Early Acheulean Assemblages, ume
2. The University of Tokyo, Tokyo. http://umdb.um.u-tokyo.ac.jp/DKankoub/Bullet
including Drs. Tristan Carter, Eduard Reinhardt, Christian Tryon, and in/no48/.
Julio Mercader. Their thoughtful comments, constructive criticism, and Biró, K.T., 1998. Lithic Implements and the Circulation of Raw Materials in the Great
help with references greatly improved the quality and scope of this Hungarian Plain during the Late Neolithic Period. Hungarian National Museum,
Budapest.
article. I would also like to thank Dr. María Soto who channelled my Bishop, W.W., Pickford, M., Hill, A., 1975. New evidence regarding the Quaternary
interests in raw material sourcing through years of mentorship and geology, archaeology and hominids of Chesowanja, Kenya. Nature 258, 204–208.
collaboration. I am grateful for the funding provided by the Social Sci­ https://doi.org/10.1038/258204a0.
Blomme, A., Degryse, P., Van Peer, P., Elsen, J., 2012. The characterization of
ences and Humanities Research Council of Canada for my Joseph-
sedimentary quartzite artefacts from Mesolithic sites, Belgium. Geol. Belg. 15,
Armand Bombardier Canada Graduate Scholarship (767-2019-2655) 193–199. https://popups.uliege.be/1374-8505/index.php?id=3690.
and the Partnership Grant Program (895-2016-1017) for the Stone Blumenschine, R.J., Masao, F.T., Tactikos, J.C., Ebert, J.I., 2008. Effects of distance from
stone source on landscape-scale variation in Oldowan artifact assemblages in the
Tools, Diet, and Sociality at the Dawn of Humanity project led by Dr.
Paleo-Olduvai Basin, Tanzania. J. Archaeol. Sci. 35, 76–86. https://doi.org/
Julio Mercader, as well as the Government of Ontario, The Leakey 10.1016/j.jas.2007.02.009.
Foundation (41660), and the Department of Anthropology at McMaster Boesch, C., Boesch, H., 1984. Mental map in wild chimpanzees: an analysis of hammer
University. I express my gratitude to the Tanzanian Commission for transports for nut cracking. Primates 25, 160–170. https://doi.org/10.1007/
BF02382388.
Science and Technology (2021-648-NA-2017/035), the Ministry of Bolton, M.S.M., Jensen, B.J., Wallace, K., Praet, N., Fortin, D., Kaufman, D., De Batist, M.,
Natural Resources and Tourism, the Antiquities Division (EA.150/297/ 2020. Machine learning classifiers for attributing tephra to source volcanoes: an
01:5/2018/2019), the Ngorongoro Conservation Area (BE.504/620/ evaluation of methods for Alaska tephras. J. Quat. Sci. 35, 81–92. https://doi.org/
10.1002/jqs.3170.
01/53), and the Executive Secretary of the Mining Commission Bond, G., 1948. Rhodesian stone age man and his raw materials. S. Afr. Archaeol. Bull. 3,
(00001258) for authorising my fieldwork at Oldupai Gorge and labo­ 55–60. https://doi.org/10.2307/3886949.
ratory analyses in Canada. I wish to also recognise the Maasai commu­ Borcard, D., Gillet, F., Legendre, P., 2018. Numerical Ecology with R. Springer, Cham.
https://doi.org/10.1007/978-1-4419-7976-6.
nity at Oldupai Gorge, without whom my doctoral research would not Brandl, M., Martinez, M.M., Hauzenberger, C., Filzmoser, P., Nymoen, P., Mehler, N.,
have been possible. Finally, I wish to thank the editor and anonymous 2018. A multi-technique analytical approach to sourcing Scandinavian flint:
reviewer for their suggestions and constructive comments on an earlier provenance of ballast flint from the shipwreck “Leirvigen 1,”. Norway. PLOS ONE 13,
e0200647. https://doi.org/10.1371/journal.pone.0200647.
draft of this article.
Braun, D.R., Plummer, T., Ditchfield, P., Ferraro, J.V., Maina, D., Bishop, L.C., Potts, R.,
2008. Oldowan behavior and raw material transport: perspectives from the Kanjera
References Formation. J. Archaeol. Sci. 35, 2329–2345. https://doi.org/10.1016/j.
jas.2008.03.004.
Braun, D.R., Plummer, T., Ferraro, J.V., Ditchfield, P., Bishop, L.C., 2009a. Raw material
Agam, A., Marder, O., Barkai, R., 2015. Small flake production and lithic recycling at
quality and Oldowan hominin toolstone preferences: evidence from Kanjera South,
Late Acheulian Revadim, Israel. Quat. Int. 361, 46–60. https://doi.org/10.1016/j.
Kenya. J. Archaeol. Sci. 36, 1605–1614. https://doi.org/10.1016/j.jas.2009.03.025.
quaint.2014.06.070.
Braun, D.R., Harris, J.W.K., Maina, D.N., 2009b. Oldowan raw material procurement and
Agha-Aligol, D., Lamehi-Rachti, M., Oliaiy, P., Shokouhi, F., Farahani, M.F., Moradi, M.,
use: evidence from the Koobi Fora Formation. Archaeometry 51, 26–42. https://doi.
Jalali, F.F., 2015. Characterization of Iranian obsidian artifacts by PIXE and
org/10.1111/j.1475-4754.2008.00393.x.
multivariate statistical analysis. Geoarchaeology 30, 261–270. https://doi.org/
Braun, D.R., Harris, J.W.K., Levin, N.E., McCoy, J.T., Herries, A.I.R., Bamford, M.K.,
10.1002/gea.21509.
Bishop, L.C., Richmond, B.G., Kibunjia, M., 2010. Early hominin diet included
Ambrose, S.H., 2012. Obsidian dating and source exploitation studies in Africa:
diverse terrestrial and aquatic animals 1.95 Ma in East Turkana, Kenya. Proc. Natl.
implications for the evolution of human behaviour. In: Liritzis, I., Stevenson, C.M.
Acad. Sci. USA 107, 10002–10007. https://doi.org/10.1073/pnas.1002181107.
(Eds.), Obsidian and Ancient Manufactured Glasses. University of New Mexico Press,
Braun, D.R., Aldeias, V., Archer, W., Arrowsmith, J.R., Baraki, N., Campisano, C.J.,
Albuquerque, pp. 56–72.
Deino, A.L., DiMaggio, E.N., Dupont-Nivet, G., Engda, B., Feary, D.A., 2019. Earliest
Amick, D.S., 2015. The recycling of material culture today and during the Paleolithic.
known Oldowan artifacts at >2.58 Ma from Ledi-Geraru, Ethiopia, highlight early
Quat. Int. 361, 4–20. https://doi.org/10.1016/j.quaint.2014.08.059.
technological diversity. Proc. Natl. Acad. Sci. USA 116, 11712–11717. https://doi.
Asfaw, B., Beyene, Y., Suwa, G., Walter, R.C., White, T.D., WoldeGabriel, G., Yemane, T.,
org/10.1073/pnas.1820177116.
1992. The earliest acheulean from konso-gardula. Nature 360, 732–735. https://doi.
Brilli, M., Cavazzini, G., Turi, B., 2005. New data of 87Sr/86Sr ratio in classical marble: an
org/10.1038/360732a0.
initial database for marble provenance determination. J. Archaeol. Sci. 32,
Bailey, G., 2007. Time perspectives, palimpsests and the archaeology of time.
1543–1551. https://doi.org/10.1016/j.jas.2005.04.007.
J. Anthropol. Archaeol. 26, 198–223. https://doi.org/10.1016/j.jaa.2006.08.002.
Brooks, A.S., Yellen, J.E., Potts, R., Behrensmeyer, A.K., Deino, A.L., Leslie, D.E.,
Barsky, D., Chapon-Sao, C., Bahain, J.J., Beyene, Y., Cauche, D., Celiberti, V.,
Ambrose, S.H., Ferguson, J.R., d’Errico, F., Zipkin, A.M., Whittaker, S., Post, J.,
Desclaux, E., de Lumley, H., de Lumley, M.A., Marchal, F., Moullé, P.E., 2011. The
Veatch, E.G., Foecke, K., Clark, J.B., 2018. Long-distance stone transport and
early oldowan stone-tool assemblage from fejej FJ-1A, Ethiopia. J. Afr. Archaeol. 9,
pigment use in the earliest Middle Stone Age. Science 360, 90–94. https://doi.org/
207–224. https://doi.org/10.3213/2191-5784-10196.
10.1126/science.aao2646.
Basu, A., Bickford, M.E., Deasy, R., 2016. Inferring tectonic provenance of siliciclastic
Brown, F.H., Nash, B.P., Fernandez, D.P., Merrick, H.V., Thomas, R.J., 2013.
rocks from their chemical compositions: a dissent. Sediment. Geol. 336, 26–35.
Geochemical composition of source obsidians from Kenya. J. Archaeol. Sci. 40,
https://doi.org/10.1016/j.sedgeo.2015.11.013.
3233–3251. https://doi.org/10.1016/j.jas.2013.03.011.
Baxter, M.J., 1994. Stepwise discriminant analysis in archaeometry: a critique.
J. Archaeol. Sci. 21, 659–666. https://doi.org/10.1006/jasc.1994.1065.

15
J. Favreau Quaternary Science Advances 9 (2023) 100068

Browne, C.L., Wilson, L., 2011. Resource selection of lithic raw materials in the Middle noolchalai (ancient RHS-mugulud) in Peninj (lake natron, Tanzania). Quat. Int.
Palaeolithic in southern France. J. Hum. Evol. 61, 597–608. https://doi.org/ 322–323, 237–263. https://doi.org/10.1016/j.quaint.2013.10.011.
10.1016/j.jhevol.2011.08.004. Diez-Martín, F., Sánchez Yustos, P., Uribelarrea, D., Baquedano, E., Mark, D.F.,
Cahen, L., Snelling, N.J., Delhal, J., Vail, J.R., Bonhomme, M., Ledent, D., 1984. The Mabulla, A., Fraile, C., Duque, J., Díaz, I., Pérez-González, A., Yravedra, J.,
Geochronology and Evolution of Africa. Clarendon Press, Oxford. Egeland, C.P., Organista, E., Domínguez-Rodrigo, M., 2015. The origin of the
Cann, J.R., Renfrew, C., 1964. The characterization of obsidian and its application to the acheulean: the 1.7 million-year-old site of FLK West, olduvai Gorge (Tanzania). Sci.
mediterranean region. Proc. Prehist. Soc. 30, 111–133. https://doi.org/10.1017/ Rep. 5, 17839 https://doi.org/10.1038/srep17839.
S0079497X00015097. Domínguez-Rodrigo, M., Barba, R., Egeland, C.P., 2007. Deconstructing Olduvai: A
Carter, T., 2014. The contribution of obsidian characterization studies to early Taphonomic Study of the Bed I Sites. Springer, New York. https://doi.org/10.1007/
prehistoric archaeology. In: Yamada, M., Ono, A. (Eds.), Lithic Raw Material 978-1-4020-6152-3.
Exploitation and Circulation in Prehistory: a Comparative Perspective in Diverse Drennan, R.D., 2009. Statistics for Archaeologists: A Common Sense Approach, second
Palaeoenvironments. ERAUL, Liège, pp. 23–33. ed. Springer, Dordrecht. https://doi.org/10.1007/978-1-4419-0413-3.
Chataigner, C., Badalian, R., Bigazzi, G., Cauvin, M.-C., Jrbashian, R., Karapetian, S.G., Dugdale, W., 1656. The Antiquities of Warwickshire Illustrated: from Records, Leiger-
Norelli, P., Oddone, M., Poidevin, J.-L., 2003. Provenance studies of obsidian Books, Manuscripts, Charters, Evidences, Tombes, and Armes. Warren, London.
artefacts from Armenian archaeological sites using the fission-track dating method. Durrani, S.A., Khan, H.A., Taj, M., Renfrew, C., 1971. Obsidian source identification by
J. Non-Cryst. Solids 323, 167–171. https://doi.org/10.1016/S0022-3093(03)00300- fission track analysis. Nature 233, 242–245. https://doi.org/10.1038/233242a0.
4. Ebright, C.A., 1987. Quartzite petrography and its implications for prehistoric use
Chavaillon, J., 1970. Découverte d’un niveau oldowayen dans la basse vallée de l’Omo archeological analysis. Archaeol. E. N. Am. 15, 29–45. https://www.jstor.org/
(Ethiopie). Bull. Soc. Prehist. Fr. 67, 7–11. https://doi.org/10.3406/ stable/40914353.
bspf.1970.10428. Egeland, C.P., Fadem, C.M., Byerly, R.M., Henderson, C., Fitzgerald, C., Mabulla, A.Z.,
Church, T., 1994. Lithic Resource Studies: A Sourcebook for Archaeologists. University of Baquedano, E., Gidna, A., 2019. Geochemical and physical characterization of lithic
Tulsa, Oklahoma. raw materials in the Olduvai Basin, Tanzania. Quat. Int. 526, 99–115. https://doi.
Clark, J.D., Kurashina, H., 1979. Hominid occupation of the East-central highlands of org/10.1016/j.quaint.2019.09.036.
Ethiopia in the plio-pleistocene. Nature 282, 33–39. https://doi.org/10.1038/ Ekshtain, R., Malinsky-Buller, A., Ilani, S., Segal, I., Hovers, E., 2014. Raw material
282033a0. exploitation around the middle paleolithic site of ‘ein qashish. Quat. Int. 331,
Clarkson, C., Bellas, A., 2014. Mapping stone: using GIS spatial modelling to predict lithic 248–266. https://doi.org/10.1016/j.quaint.2013.07.025.
source zones. J. Archaeol. Sci. 46, 324–333. https://doi.org/10.1016/j. Enkelmann, E., Jonckheere, R., 2021. Fission track dating. In: Alderton, D., Elias, S.A.
jas.2014.03.035. (Eds.), Encyclopedia of Geology, second ed. Academic Press, Cambridge,
Cnudde, V., Dewanckele, J., De Kock, T., Boone, M., Baele, J.-M., Crombé, P., pp. 116–131. https://doi.org/10.1016/B978-0-12-409548-9.11991-8.
Robinson, E., 2013. Preliminary structural and chemical study of two quartzite Favreau, J., Soto, M., Nair, R., Bushozi, P.M., Clarke, S., Durkin, P.R., Hubbard, S.M.,
varieties from the same geological formation: a first step in the sourcing of quartzites Inwood, J., Itambu, M., Larter, F., Lee, P., Mwambwiga, A., Patalano, R., Tucker, L.,
utilized during the Mesolithic in northwest Europe. Geol. Belg. 16, 27–34. https://po Mercader, J., 2020. Petrographic characterization of raw material sources at Oldupai
pups.uliege.be/1374-8505/index.php?id=3981. Gorge, Tanzania. Front. Earth Sci. 8, 158. https://doi.org/10.3389/
Cogné, J.E., Giot, P.-R., 1952. Étude pétrographique des haches polies de Bretagne. Bull. feart.2020.00158.
Soc. Prehist. Fr. 49, 388–395. https://doi.org/10.3406/bspf.1952.5080. Féblot-Augustins, J., 1990. Exploitation des matières premières dans l’Acheuléen
Craig, H., Craig, V., 1972. Greek marbles: determination of provenance by isotopic d’Afrique : perspectives comportementales. Paléo 2, 27–42. https://doi.org/
analysis. Science 176, 401–403. https://doi.org/10.1126/science.176.4033.401. 10.3406/pal.1990.987.
Cueva-Temprana, A., Lombao, D., Soto, M., Itambu, M., Bushozi, P., Boivin, N., Fernandes, P., 2012. Itinéraires et transformations du silex: une pétroarchéologie
Petraglia, M., Mercader, J., 2022. Oldowan technology amid shifting Environments~ refondée, application au Paléolithique moyen. PhD Dissertation, Université de
2.03–1.83 million years ago. Frontiers in Ecology and Evolution 10, 788101. https:// Bordeaux.
doi.org/10.3389/fevo.2022.788101. Finestone, E., Braun, D.R., Plummer, T.W., Bartilol, S., Kiprono, N., 2020. Building ED-
Curran, J.M., Meighan, I.G., Simpson, D.D.A., Rogers, G., Fallick, A.E., 2001. 87Sr/86Sr: a XRF datasets for sourcing rhyolite and quartzite artifacts: a case study on the Homa
new discriminant for provenancing neolithic porcellanite artifacts from Ireland. Peninsula, Kenya. J. Archaeol. Sci.: Report 33, 102510. https://doi.org/10.1016/j.
J. Archaeol. Sci. 28, 713–720. https://doi.org/10.1006/jasc.2000.0582. jasrep.2020.102510.
Dalpra, C.L., Pitblado, B.L., 2016. Discriminating quartzite sources petrographically in Flores-Prieto, E., Quénéhervé, G., Bachofer, F., Shahzad, F., Maerker, M., 2015.
the upper gunnison basin, Colorado: implications for paleoamerican lithic- Morphotectonic interpretation of the Makuyuni catchment in Northern Tanzania
procurement studies. PaleoAmerica 2, 22–31. https://doi.org/10.1080/ using DEM and SAR data. Geomorphology 248, 427–439. https://doi.org/10.1016/j.
20555563.2015.1137684. geomorph.2015.07.049.
Damour, A., 1866. Sur la composition des haches en pierre trouvées dans les monuments Foley, R.A., 2018. Evolutionary geography and the afrotropical model of hominin
celtiques et chez les tribus sauvages. Rev. Archéol. 13, 190–207. https://www.jstor. evolution. Bull. Mem. Soc. Anthropol. Paris 30, 17–31. https://doi.org/10.3166/
org/stable/41746734. bmsap-2018-0001.
de Boutray, B., 1981. Étude pétrographique comparative de quartzites enfumés utilisés Frahm, E., 2012. Evaluation of archaeological sourcing techniques: reconsidering and
par les Paléoesquimaux de l’Artique québécois. Quatrième Colloque sur le Re-deriving hughes’ four-fold assessment scheme. Geoarchaeology 27, 166–174.
Quaternaire du Québec 35, 29–40. https://doi.org/10.7202/1000375ar. https://doi.org/10.1002/gea.21399.
de la Torre, I., 2011. The early stone age lithic assemblages of Gadeb (Ethiopia) and the Frahm, E., 2013. Is obsidian sourcing about geochemistry or archaeology? A reply to
developed oldowan/early acheulean in East Africa. J. Hum. Evol. 60, 768–812. Speakman and Shackley. J. Archaeol. Sci. 40, 1444–1448. https://doi.org/10.1016/
https://doi.org/10.1016/j.jhevol.2011.01.009. j.jas.2012.10.001.
de la Torre, I., 2016. The origins of the Acheulean: past and present perspectives on a Frahm, E., Feinberg, J.M., Schmidt-Magee, B.A., Wilkinson, K.N., Gasparyan, B.,
major transition in human evolution. Phil. Trans. Biol. Sci. 371, 20150245 https:// Yeritsyan, B., Adler, D.S., 2016. Middle Palaeolithic toolstone procurement
doi.org/10.1098/rstb.2015.0245. behaviors at Lusakert cave 1, Hrazdan valley, Armenia. J. Hum. Evol. 91, 73–92.
de la Torre, I., Mora, R., 2005. Unmodified lithic material at Olduvai Bed I: manuports or https://doi.org/10.1016/j.jhevol.2015.10.008.
ecofacts? J. Archaeol. Sci. 32, 273–285. https://doi.org/10.1016/j.jas.2004.09.010. Frahm, E., Jones, C.O., Corolla, M., Wilkinson, K.N., Sherriff, J.E., Gasparyan, B.,
de la Torre, I., Mora, R., 2018. Technological behaviour in the early acheulean of EF-HR Adler, D.S., 2020. Comparing lower and middle palaeolithic lithic procurement
(olduvai Gorge, Tanzania). J. Hum. Evol. 120, 329–377. https://doi.org/10.1016/j. behaviors within the hrazdan basin of central Armenia. J. Archaeol. Sci.: Report 32,
jhevol.2018.01.003. 102389. https://doi.org/10.1016/j.jasrep.2020.102389.
de la Torre, I., Mora, R., Martínez-Morena, J., 2008. The early acheulean in Peninj (lake Fritz, H., Abdelsalam, M., Ali, K.A., Bingen, B., Collins, A.S., Fowler, A.R., Ghebreab, W.,
natron, Tanzania). J. Anthropol. Archaeol. 27, 244–264. https://doi.org/10.1016/j. Hauzenberger, C.A., Johnson, P.R., Kusky, T.M., Macey, P., 2013. Orogen styles in
jaa.2007.12.001. the East African orogen: a review of the neoproterozoic to cambrian tectonic
Delagnes, A., Boisserie, J.-R., Beyene, Y., Chuniaud, K., Guillemot, C., Schuster, M., 2011. evolution. J. Afr. Earth Sci. 86, 65–106. https://doi.org/10.1016/j.
Archaeological investigations in the lower Omo valley (shungura formation, jafrearsci.2013.06.004.
Ethiopia): new data and perspectives. J. Hum. Evol. 61, 215–222. https://doi.org/ Gale, N.H., 1981. Mediterranean obsidian source characterisation by strontium isotope
10.1016/j.jhevol.2011.03.008. analysis. Archaeometry 23, 41–51. https://doi.org/10.1111/j.1475-4754.1981.
Dibble, H.L., Holdaway, S.J., Lin, S.C., Braun, D.R., Douglass, M.J., Iovita, R., tb00953.x.
McPherron, S.P., Olszewski, D.I., Sandgathe, D., 2017. Major fallacies surrounding Gall, D.G., Steponaitis, V.P., 2001. Composition and provenance of greenstone artifacts
stone artifacts and assemblages. J. Archaeol. Method Theor 24, 813–851. https:// from Moundville. SE. Archaeol. 20, 99–117. https://www.jstor.org/stable/
doi.org/10.1007/s10816-016-9297-8. 40713210.
Diez-Martín, F., Sánchez, P., Domínguez-Rodrigo, M., Mabulla, A., Barba, R., 2009. Were Gallotti, R., Mussi, M., 2015. The unknown Oldowan: ~1.7-million-year-old
Olduvai hominins making butchering tools or battering tools? Analysis of a recently standardized obsidian small tools from Garba IV, Melka Kunture, Ethiopia. PLoS One
excavated lithic assemblage from BK (Bed II, Olduvai Gorge, Tanzania). 10, e0145101. https://doi.org/10.1371/journal.pone.0145101.
J. Anthropol. Archaeol. 28, 274–289. https://doi.org/10.1016/j.jaa.2009.03.001. Gallotti, R., Mussi, M., 2017. Two acheuleans, two humankinds: from 1.5 to 0.85 Ma at
Diez-Martín, F., Sánchez Yustos, P., de la Rúa, D.G., Gómez González, J.Á., de Luque, L., Melka kunture (upper awash, Ethiopian highlands). Journal of Anthropological
Barba, R., 2014a. Early acheulean technology at Es2-lepolosi (ancient MHS-bayasi) Sciences 95, 137–181. https://doi.org/10.4436/jass.95001.
in Peninj (lake natron, Tanzania). Quat. Int. 322–323, 209–236. https://doi.org/ Gallotti, R., Collina, C., Raynal, J.-P., Kieffer, G., Geraads, D., Piperno, M., 2010. The
10.1016/j.quaint.2013.08.053. early middle Pleistocene site of Gombore II (Melka kunture, upper awash, Ethiopia)
Diez-Martín, F., Sánchez Yustos, P., Gómez González, J.Á., Luque, L., de la Rúa, D.G., and the issue of acheulean bifacial shaping strategies. Afr. Archaeol. Rev. 27,
Domínguez-Rodrigo, M., 2014b. Reassessment of the early acheulean at EN1- 291–322. https://doi.org/10.1007/s10437-010-9083-z.

16
J. Favreau Quaternary Science Advances 9 (2023) 100068

Garrison, E.G., 2003. Techniques in Archaeological Geology. Springer, New York. material sites – a case study on understanding prehistoric human settlement activity
https://doi.org/10.1007/978-3-319-30232-4. in the southwestern Ethiopian Highlands. E&G Quaternary Science Journal 68,
Garzanti, E., 2016. From static to dynamic provenance analysis—sedimentary petrology 201–213. https://doi.org/10.5194/egqsj-68-201-2019.
upgraded. Sediment. Geol. 336, 3–13. https://doi.org/10.1016/j. Hermann, A., Forkel, R., McAlister, A., Cruickshank, A., Golitko, M., Kneebone, B.,
sedgeo.2015.07.010. McCoy, M., Reepmeyer, C., Sheppard, P., Sinton, J., Weisler, M., 2020. Pofatu, a
Gauthier, G., Burke, A.L., 2011. The effects of surface weathering on the geochemical curated and open-access database for geochemical sourcing of archaeological
analysis of archaeological lithic samples using non-destructive polarized energy materials. Sci. Data 7, 141. https://doi.org/10.1038/s41597-020-0485-8.
dispersive XRF. Geoarchaeology 26, 269–291. https://doi.org/10.1002/gea.20346. Hermes, O.D., Luedtke, B.E., Ritchie, D., 2001. Melrose green rhyolite: its geologic
Glascock, M.D., Neff, H., 2003. Neutron activation analysis and provenance research in setting and petrographic and geochemical characteristics. J. Archaeol. Sci. 28,
archaeology. Meas. Sci. Technol. 14, 1516–1526. https://doi.org/10.1088/0957- 913–928. https://doi.org/10.1006/jasc.2000.0605.
0233/14/9/304. Herz, N., 1992. Provenance determination of Neolithic to classical Mediterranean
Glascock, M.D., Braswell, G., Cobean, R.H., 1998. A systematic approach to obsidian marbles by stable isotopes. Archaeometry 34, 185–194. https://doi.org/10.1111/
source characterization. In: Shackley, M.S. (Ed.), Archaeological Obsidian Studies: j.1475-4754.1992.tb00491.x.
Method and Theory. Springer, New York, pp. 15–66. https://doi.org/10.1007/978- Hoare, S., Brink, J.S., Herries, A.I.R., Mark, D.F., Morgan, L.E., Onjala, I., Rucina, S.M.,
1-4757-9276-8_2. Stanistreet, I.G., Stollhofen, H., Gowlett, J.A.J., 2021. Geochronology of a long
Glascock, M.D., Speakman, R.J., Popelka-Filcoff, R.S. (Eds.), 2007. Archaeological Pleistocene sequence at Kilombe volcano, Kenya: from the oldowan to middle stone
Chemistry: Analytical Techniques and Archaeological Interpretations. American age. J. Archaeol. Sci. 125, 105273 https://doi.org/10.1016/j.jas.2020.105273.
Chemical Society, Washington. https://doi.org/10.1021/bk-2007-0968. Howell, F.C., 1961. Isimila: a paleolithic site in Africa. Sci. Am. 204, 118–131. https://
Göbel, F., 1842. Ueber den Einfluß der Chemie auf die Ermittelung der Völker der Vorzeit www.jstor.org/stable/24937108.
oder Resultate der chemischen Untersuchung metallischer Alterthümer: Howell, F.C., Cole, G.H., Kleindienst, M.R., 1962. Isimila: an acheulian occupation site in
insbesondere der in den Ostseegouvernements vorkommenden, behufs der the iringa highlands, southern highlands province, tanganyika. In: Mortelmans, G.,
Ermittelung der Völker, von welchen sie abstammen. Enke Verlag, Erlangen. http Nenquin, J. (Eds.), Actes du IVe Congrès Panafricain de Préhistoire et de l’Étude du
://hdl.handle.net/10062/39514. Quaternaire. Musée Royal de l’Afrique Centrale, Tervuren, pp. 43–80.
Goldman-Neuman, T., Hovers, E., 2012. Raw material selectivity in late pliocene Isaac, G.L., 1977. Olorgesailie: Archaeological Studies of a Middle Pleistocene Lake Basin
oldowan sites in the makaamitalu basin, hadar, Ethiopia. J. Hum. Evol. 62, 353–366. in Kenya. University of Chicago Press, Chicago.
https://doi.org/10.1016/j.jhevol.2011.05.006. Isaac, G., 1981. Stone Age visiting cards: approaches to the study of early land-use
Gossa, T., Hovers, E., 2022. Continuity and change in lithic techno-economy of the early patterns. In: Hodder, I., Isaac, G., Hammond, N. (Eds.), Pattern of the Past: Studies in
Acheulian on the Ethiopian highland: a case study from locality MW2; the Melka the Honour of David Clarke. Cambridge University Press, Cambridge, pp. 131–155.
Wakena site-complex. PLoS One 17, e0277029. https://doi.org/10.1371/journal. Isaac, G.L., Isaac, B., 1997. In: Koobi Fora Research Project Volume 5: Plio-Pleistocene
pone.0277029. Archaeology, ume 5. Clarendon Press, Oxford.
Gowlett, J.A.J., 1978. Kilombe – an acheulean site complex in Kenya. In: Bishop, W.W. Julig, P.J., Long, D.G.F., Hancock, R.G.V., 1998. Cathodoluminescence and petrographic
(Ed.), Geological Background to Fossil Man. Scottish Academic Press, Edinburgh, techniques for positive identification of quartz-rich artifacts from Late Paleo-Indian
pp. 337–360. https://www.lyellcollection.org/doi/10.1144/gsl.sp.1978.006.01.24. sites in the Great Lakes Region. The Wisconsin Archeologist 79, 68–88.
Gowlett, J.A.J., Crompton, R.G., 1994. Kariandusi: Acheulean morphology and the Kasztovszky, Z.S., Biró, K.T., Markó, A., Dobosi, V., 2008. Cold neutron prompt gamma
question of allometry. Afr. Archaeol. Rev. 12, 3–42. https://doi.org/10.1007/ activation analysis – a non destructive method for characterisation of high silica
BF01953037. content chipped stone tools and raw materials. Archaeometry 50, 12–29. https://doi.
Gowlett, J.A.J., Harris, J.W.K., Walton, D., Wood, B.A., 1981. Early archaeological sites, org/10.1111/j.1475-4754.2007.00348.x.
hominid remains and traces of fire from Chesowanja, Kenya. Nature 294, 125–129. Keiller, A., Piggott, S., Wallis, F.S., 1941. First report of the sub-committee of the south-
https://doi.org/10.1038/294125a0. western group of museums and art galleries on the petrological identification of
Gowlett, J.A.J., Stanistreet, I.G., Albert, R.M., Blackbird, S.J., Herries, A.I.R., Hoare, S., stone axes. Proc. Prehist. Soc. 7, 50–72. https://doi.org/10.1017/
Kogai, P., Komboh, C.K., Mark, D.F., Muriuki, R.M., Murphy, H., Rucina, S.M., S0079497X00020272.
Stollhofen, H., 2021. New oldowan localities at high level within Kilombe caldera, Key, A., Proffitt, T., de la Torre, I., 2020. Raw material optimization and stone tool
Kenya. L’Anthropologie 126, 102976. https://doi.org/10.1016/j. engineering in the early stone age of olduvai Gorge (Tanzania). J. R. Soc. Interface
anthro.2021.102976. 17, 20190377. https://doi.org/10.1098/rsif.2019.0377.
Gurova, M., Andreeva, P., Stefanova, E., Stefanov, Y., Kočić, M., Borić, D., 2016. Flint Khatsenovich, A.M., Shelepaev, R.A., Rybin, E.P., Shelepov, Y.Y., Marchenko, D.V.,
raw material transfers in the prehistoric Lower Danube Basin: an integrated Odsuren, D., Gunchinsuren, B., Olsen, J.W., 2020. Long distance transport and use of
analytical approach. J. Archaeol. Sci.: Report 5, 422–441. https://doi.org/10.1016/ mica in the initial upper paleolithic of central Asia: an example from the kharganyn
j.jasrep.2015.12.014. gol 5 site (northern Mongolia). J. Archaeol. Sci.: Report 31, 102307. https://doi.org/
Habermann, J.M., McHenry, L.J., Stollhofen, H., Tolosana-Delgado, R., Stanistreet, I.G., 10.1016/j.jasrep.2020.102307.
Deino, A.L., 2016. Discrimination, correlation, and provenance of Bed I Kieffer, G., Raynal, J.-P., Bardin, G., 2004. Volcanic markers in coarse alluvium at Melka
tephrostratigraphic markers, Olduvai Gorge, Tanzania, based on multivariate kunture (upper awash, Ethiopia). In: Chavaillon, J., Piperno, M. (Eds.), Studies on
analyses of phenocryst compositions. Sediment. Geol. 339, 115–133. https://doi. the Early Paleolithic Site of Melka Kunture, Ethiopia. Istituto Italiano di Preistoria e
org/10.1016/j.sedgeo.2016.03.026. Protostoria, Florence, pp. 93–101.
Hancock, R.G.V., Carter, T., 2010. How reliable are our published archaeometric Kiehn, A.V., Brook, G.A., Glascock, M.D., Dake, J.Z., Robbins, L.H., Campbell, A.C.,
analyses? Effects of analytical techniques through time on the elemental analysis of Murphy, M.L., 2007. Fingerprinting specular hematite from mines in Botswana,
obsidians. J. Archaeol. Sci. 37, 243–250. https://doi.org/10.1016/j.jas.2009.10.004. southern Africa. In: Glascock, M.D., Speakman, R.J., Popelka-Filcoff, R.S. (Eds.),
Harbottle, G., 1976. Activation analysis in archaeology. In: Newton, G.W.A. (Ed.), Archaeological Chemistry: Analytical Techniques and Archaeological Interpretation.
Radiochemistry: a Specialist Periodical Report. The Chemical Society, London, American Chemical Society, Washington, pp. 460–479. https://doi.org/10.1021/bk-
pp. 33–72. 2007-0968.ch025.
Harmand, S., 2009a. Variability in raw material selectivity at the late pliocene sites of Kimura, Y., 1997. The MNK chert factory site: the chert-using strategy by early hominids
Lokalalei, West Turkana, Kenya. In: Hovers, E., Braun, D.R. (Eds.), Interdisciplinary at olduvai Gorge, Tanzania. Afr. Stud. Monogr. 18, 1–28. https://jambo.africa.
Approaches to the Oldowan. Springer, New York, pp. 85–97. https://doi.org/ kyoto-u.ac.jp/kiroku/asm_normal/abstracts/pdf/ASM%20%20Vol.18%20No.1%
10.1007/978-1-4020-9060-8_8. 201997/Yuki%20KIMURA.pdf.
Harmand, S., 2009b. Raw materials and techno-economic behaviors at oldowan and Klein, C., Dutrow, B., 2008. Mineral Science, 23rd Edition. John Wiley & Sons, New
acheulean sites in the West Turkana region, Kenya. In: Adams, B., Blades, B.S. (Eds.), Jersey.
Lithic Materials and Paleolithic Societies. Wiley-Blackwell, West Sussex, pp. 3–14. Kooyman, B.P., 2000. Understanding Stone Tools and Archaeological Sites. University of
https://doi.org/10.1002/9781444311976.ch1. Calgary Press, Calgary. https://doi.org/10.2307/j.ctv6gqs13.
Harrell, J.A., 1992. Ancient Egyptian limestone quarries: a petrological survey. Kovarovic, K., Aiello, L.C., Cardini, A., Lockwood, C.A., 2011. Discriminant function
Archaeometry 34, 195–211. https://doi.org/10.1111/j.1475-4754.1992.tb00492.x. analyses in archaeology: are classification rates too good to be true? J. Archaeol. Sci.
Hay, R.L., 1976. Geology of the Olduvai Gorge: A Study of Sedimentation in a Semiarid 38, 3006–3018. https://doi.org/10.1016/j.jas.2011.06.028.
Basin. University of California Press, Berkeley. Kuman, K., 1994. The archaeology of Sterkfontein – past and present. J. Hum. Evol. 27,
Hazenfratz, R., Munita, C.S., Neves, E.G., 2017. Neural networks (SOM) applied to INAA 471–495. https://doi.org/10.1006/jhev.1994.1065.
data of chemical elements in archaeological ceramics from central amazon. Science Kuman, K., 2014. Acheulean industrial complex. In: Smith, C. (Ed.), Encyclopedia of
& Technology of Archaeological Research 3, 334–340. https://doi.org/10.1080/ Global Archaeology. Springer, New York, pp. 7–18. https://doi.org/10.1007/978-1-
20548923.2018.1470218. 4419-0465-2_653.
Heizer, R.F., Williams, H., Graham, J.A., 1965. Notes on Mesoamerican Obsidians and Kuman, K., Field, A.S., 2009. The oldowan industry from Sterkfontein caves, South
Their Significance in Archaeological Studies, vol. 1. Contributions of the University Africa. In: Schick, K., Toth, N. (Eds.), The Cutting Edge: New Approaches to the
of California Archaeological Research Facility, pp. 94–103. Archaeology of Human Origins. Stone Age Institute Press, Gosport, pp. 151–169.
Helm, O., 1886. Mycenean amber imported from the Baltic. In: Schliemann, H. (Ed.), Kuman, K., Sutton, M.B., Pickering, T.R., Heaton, J.L., 2018. The oldowan industry from
Tiryns: the Prehistoric Palace of the Kings of Tiryns. The Results of the Latest Swartkrans cave, South Africa, and its relevance for the african oldowan. J. Hum.
Excavations. John Murray, London, pp. 369–372. https://doi.org/10.1017/ Evol. 123, 52–69. https://doi.org/10.1016/j.jhevol.2018.06.004.
CBO9780511740299.009. Leakey, M.D., 1971. Olduvai Gorge Volume 3: Excavations in Beds I and II, 1960-1963.
Henderson, J., 2000. The Science and Archaeology of Materials: an Investigation of Cambridge University Press, Cambridge.
Inorganic Materials. Routledge, London. https://doi.org/10.4324/9780203630143. Leakey, M.D., 1979. Olduvai Gorge: My Search for Early Man. Collins, London.
Hensel, E.A., Bödeker, O., Bubenzer, O., Vogelsang, R., 2019. Combining Leakey, L.S.B., Evernden, J.F., Curtis, G.H., 1961. Age of bed I, olduvai Gorge,
geomorphological–hydrological analyses and the location of settlement and raw tanganyika. Nature 191, 478–479. https://doi.org/10.1038/191478a0.

17
J. Favreau Quaternary Science Advances 9 (2023) 100068

Leakey, M., Tobias, P.V., Martyn, J.E., Leakey, R.E.F., 1969. An acheulean industry with exploited unstable environments ~ 2 million years ago. Nat. Commun. 12, 3.
prepared core technique and the discovery of a contemporary hominid mandible at https://doi.org/10.1038/s41467-020-20176-2.
Lake Baringo, Kenya. Proc. Prehist. Soc. 35, 48–76. https://doi.org/10.1017/ Merrick, H.V., Brown, F.H., 1984. Obsidian sources and patterns of source utilization in
S0079497X00013402. Kenya and northern Tanzania: some initial findings. Afr. Archaeol. Rev. 2, 129–152.
LeBlanc, D., 2004. Caractérisation géochimique de matières premières lithiques: Analyse https://doi.org/10.1007/BF01117229.
de la quartzite de Mistassini (colline Blanche, rivière Témiscamie) et de la calcédonie Merrick, H.V., Brown, F.H., Nash, W.P., 1994. Use and movement of obsidian in the early
du lac-Saint-Jean (île aux Couleuvres, lac Saint-Jean). MSc Thesis, Université du and middle stone ages of Kenya and northern Tanzania. In: Childs, S.T. (Ed.),
Québec à Chicoutimi. Society, Culture, and Technology in Africa, MASCA Research Papers in Science and
Legg, R., Nielson, J., Demel, S., 2020. Novel use of cathodoluminescence to identify Archaeology, vol. 11, pp. 29–44 (Philadelphia).
differences in source rocks for Late Paleoindian quartzite tools. Archaeometry 62, Middleton, A.P., Bradley, S.M., 1989. Provenancing of Egyptian limestone sculpture.
875–887. https://doi.org/10.1111/arcm.12568. J. Archaeol. Sci. 16, 475–488. https://doi.org/10.1016/0305-4403(89)90069-1.
Lepre, C.J., Roche, H., Kent, D.V., Harmand, S., Quinn, R.L., Brugal, J.-P., Texier, P.-J., Molen, N.H., Rozendal, R.H., Boon, W., 1972. Fundamental characteristics of human gait
Lenoble, A., Feibel, C.S., 2011. An earlier origin for the Acheulian. Nature 477, in relation to sex and location. Proc. K. Ned. Akad. Wet. Ser. C Biol. Med. Sci. 45,
82–85. https://doi.org/10.1038/nature10372. 215–223.
Leverington, D.W., Teller, J.T., Mann, J.D., 2002. A GIS method for reconstruction of late Mollel, G.F., 2002. Petrology and Geochemistry of the Southeastern Ngorongoro
Quaternary landscapes from isobase data and modern topography. Comput. Geosci. Volcanic Highland; and Contribution to “Sourcing” of Stone Tools at Olduvai Gorge,
28, 631–639. https://doi.org/10.1016/S0098-3004(01)00097-8. Tanzania. MSc Thesis, Rutgers University.
Lindsay, J.J., Hughes, H.S., Yeomans, C.M., Andersen, J.C., McDonald, I., 2021. Moreau, L., Brandl, M., Filzmoser, P., Hauzenberger, C., Goemaere, É., Jadin, I.,
A machine learning approach for regional geochemical data: platinum-group Collet, H., Hauzeur, A., Schmitz, R.W., 2016. Geochemical sourcing of flint artifacts
element geochemistry vs geodynamic settings of the North Atlantic Igneous from western Belgium and the German rhineland: testing hypotheses on gravettian
Province. Geosci. Front. 12, 101098 https://doi.org/10.1016/j.gsf.2020.10.005. period mobility and raw material economy. Geoarchaeology 31, 229–243. https://
Lowe, D.J., Pearce, N.J., Jorgensen, M.A., Kuehn, S.C., Tryon, C.A., Hayward, C.L., 2017. doi.org/10.1002/gea.21564.
Correlating tephras and cryptotephras using glass compositional analyses and Morgan, L.E., Renne, P.R., Taylor, R.E., WoldeGabriel, G., 2009. Archaeological age
numerical and statistical methods: review and evaluation. Quat. Sci. Rev. 175, 1–44. constraints from extrusion ages of obsidian: examples from the Middle Awash,
https://doi.org/10.1016/j.quascirev.2017.08.003. Ethiopia. Quat. Geochronol. 4, 193–203. https://doi.org/10.1016/j.
Lucero, G.F., Castro, S.C., Cortegoso, V., 2021. GIS modeling of lithic procurement in quageo.2009.01.001.
highlands: archaeological and actualistic approach in the Andes. J. Archaeol. Sci.: Moutsiou, T., 2014. The Obsidian Evidence for the Scale of Social Life during the
Report 38, 103026. https://doi.org/10.1016/j.jasrep.2021.103026. Palaeolithic. BAR International Series 2613. Archaeopress, Oxford.
Lüdecke, T., Schrenk, F., Thiemeyer, H., Kullmer, O., Bromage, T.G., Sandrock, O., Moutsiou, T., Agapiou, A., 2019. Least cost pathway analysis of obsidian circulation in
Fiebig, J., Mulch, A., 2016. Persistent C3 vegetation accompanied plio-pleistocene early holocene–early middle holocene Cyprus. J. Archaeol. Sci.: Report 26, 101881.
hominin evolution in the Malawi Rift (chiwondo beds, Malawi). J. Hum. Evol. 90, https://doi.org/10.1016/j.jasrep.2019.101881.
163–175. https://doi.org/10.1016/j.jhevol.2015.10.014. Mussi, M., Altamura, F., Di Bianco, L., Bonnefille, R., Gaudzinski-Windheuser, S.,
Luncz, L.V., Proffitt, T., Kulik, L., Haslam, M., Wittig, R.W., 2016. Distance-decay effect Geraads, D., Melis, R.T., Panera, J., Piarulli, F., Pioli, L., Ruta, G., Sánchez-Dehesa
in stone tool transport by wild chimpanzees. Philosophical Transactions of the Royal Galán, S., Méndez-Quintas, E., 2021. After the emergence of the acheulean at Melka
Society B 283, 20161607. https://doi.org/10.1098/rspb.2016.1607. kunture (upper awash, Ethiopia): from Gombore IB (1.6 Ma) to Gombore Iγ (1.4 Ma),
Lundblad, S.P., Mills, P.R., Hon, K., 2008. Analysing archaeological basalt using non- Gombore Iδ (1.3 Ma) and Gombore II OAM test Pit C (1.2 Ma). Quat. Int. https://doi.
destructive energy-dispersive X-ray fluorescence (EDXRF): effects of post- org/10.1016/j.quaint.2021.02.031.
depositional chemical weathering and sample size on analytical precision. Nash, D.J., Coulson, S., Staurset, S., Ullyott, J.S., Babutsi, M., Hopkinson, L., Smith, M.P.,
Archaeometry 50, 1–11. https://doi.org/10.1111/j.1475-4754.2007.00345.x. 2013. Provenancing of silcrete raw materials indicates long-distance transport to
Macgregor, D., 2015. History of the development of the East African Rift System: a series Tsodilo Hills, Botswana, during the Middle Stone Age. J. Hum. Evol. 64, 280–288.
of interpreted maps through time. J. Afr. Earth Sci. 101, 232–252. https://doi.org/ https://doi.org/10.1016/j.jhevol.2013.01.010.
10.1016/j.jafrearsci.2014.09.016. Navazo, M., Colina, A., Domínguez-Bella, S., Benito-Calvo, A., 2008. Raw stone material
Macholdt, D.S., Jochum, K.P., Pöhlker, C., Arangio, A., Förster, J.-D., Stoll, B., Weis, U., supply for Upper Pleistocene settlements in Sierra de Atapuerca (Burgos, Spain): flint
Weber, B., Müller, M., Kappl, M., Shiraiwa, M., Kilcoyne, A.L.D., Weigand, M., characterization using petrographic and geochemical techniques. J. Archaeol. Sci.
Scholz, D., Haug, D., Al-Amri, A., Andreae, M.O., 2017. Characterization and 35, 1961–1973. https://doi.org/10.1016/j.jas.2007.12.009.
differentiation of rock varnish types from different environments by microanalytical Nazaroff, A.J., Baysal, A., Çiftçi, Y., 2013. The importance of chert in central Anatolia:
techniques. Chem. Geol. 459, 91–118. https://doi.org/10.1016/j. lessons from the Neolithic assemblage at Çatalhöyük, Turkey. Geoarchaeology 28,
chemgeo.2017.04.009. 340–362. https://doi.org/10.1002/gea.21446.
Marder, O., Milevski, I., Matskevich, Z., 2006. The handaxes of Revadim Quarry: typo- Neff, H., 2017. Inductively coupled plasma-mass spectrometry (ICP-ms). In: Gilbert, A.S.
technological considerations and aspects of intra-site variability. In: Goren-Inbar, N., (Ed.), Encyclopedia of Geoarchaeology. Springer, Dordrecht, pp. 433–449. https://
Sharon, G. (Eds.), Axe Age: Acheulian Toolmaking from Quarry to Discard. Equinox, doi.org/10.1007/978-1-4020-4409-0_19.
London, pp. 223–242. Negash, A., Shackley, M.S., Alene, M., 2006. Source provenance of obsidian artifacts
Marwick, B., 2003. Pleistocene exchange networks as evidence for the evolution of from the early stone age (ESA) site of Melka konture, Ethiopia. J. Archaeol. Sci. 33,
language. Camb. Archaeol. J. 13, 67–81. https://doi.org/10.1017/ 1647–1650. https://doi.org/10.1016/j.jas.2005.11.001.
S0959774303000040. Negash, A., Nash, B.P., Brown, F.H., 2020. An initial survey of the composition of
Marwick, B., 2017. Computational reproducibility in archaeological research: basic Ethiopian obsidian. J. Afr. Earth Sci. 172, 103977 https://doi.org/10.1016/j.
principles and a case study of their implementation. J. Archaeol. Method Theor 24, jafrearsci.2020.103977.
424–450. https://doi.org/10.1007/s10816-015-9272-9. O’Neil, J.R., Hay, R.L., 1973. 18O/16O ratios in cherts associated with the saline lake
Marwick, B., Birch, S.E.P., 2018. A standard for the scholarly citation of archaeological deposits of East Africa. Earth Planet Sci. Lett. 19, 257–266. https://doi.org/10.1016/
data as an incentive to data sharing. Advances in Archaeological Practice 6, 0012-821X(73)90126-X.
125–143. https://doi.org/10.1017/aap.2018.3. O’Neill, M.C., Lee, L.-F., Demes, B., Thompson, N.E., Larson, S.G., Stern Jr., J.T.,
Mathur, R., Burns, J., Powell, W., Boryk, R., Sheetz, B., D’Amico, P., Harney, P., 2020. Umberger, B.R., 2015. Three-dimensional kinematics of the pelvis and hind limbs in
Evaluation of Fe isotope values as a provenance tool for chert artefacts from the chimpanzee (Pan troglodytes) and human bipedal walking. J. Hum. Evol. 86, 32–42.
north-eastern United States. Archaeometry 62, 156–168. https://doi.org/10.1111/ https://doi.org/10.1016/j.jhevol.2015.05.012.
arcm.12572. Parish, R.M., Swihart, G.M., Li, Y.S., 2013. Evaluating fourier transform infrared
Mauran, G., Caron, B., Détroit, F., Nankela, A., Bahain, J.-J., Pleurdeau, D., Lebon, M., spectroscopy as a non-destructive chert sourcing technique. Geoarchaeology 28,
2021. Data pretreatment and multivariate analyses for ochre sourcing: application to 289–307. https://doi.org/10.1002/gea.21437.
Leopard Cave (Erongo, Namibia). J. Archaeol. Sci.: Report 35, 102757. https://doi. Parks, G.A., Tieh, T.T., 1966. Identifying the geographical source of artefact obsidian.
org/10.1016/j.jasrep.2020.102757. Nature 211, 289–290. https://doi.org/10.1038/211289a0.
McBrearty, S., 2003. Patterns of technological change at the origin of Homo sapiens. Perkins, D., 1998. Mineralogy. Prentice Hall, New Jersey.
Before Farming 3, 1–6. https://doi.org/10.3828/bfarm.2003.3.9. Pitblado, B.L., Dehler, C., Neff, H., Nelson, S.T., 2008. Pilot study experiments sourcing
McBrearty, S., Brooks, A.S., 2000. The revolution that wasn’t: a new interpretation of the quartzite, gunnison basin, Colorado. Geoarchaeology 23, 742–778. https://doi.org/
origin of modern human behavior. J. Hum. Evol. 39, 453–463. https://doi.org/ 10.1002/gea.20240.
10.1006/jhev.2000.0435. Pitblado, B.L., Cannon, M.B., Neff, H., Dehler, C.M., Nelson, S.T., 2013. LA-ICP-MS
McDonald, M.M.A., 1991. Systematic reworking of lithics from earlier cultures in the analysis of quartzite from the upper gunnison basin, Colorado. J. Archaeol. Sci. 40,
early holocene of dakhleh oasis, Egypt. J. Field Archaeol. 18, 269–273. https://doi. 2196–2216. https://doi.org/10.1016/j.jas.2012.11.016.
org/10.1179/009346991792208281. Pollard, A.M., Bray, P.J., Gosden, C., 2014. Is there something missing in scientific
McHenry, L.J., de la Torre, I., 2018. Hominin raw material procurement in the Oldowan- provenance studies of prehistoric artefacts? Antiquity 88, 625–631. https://doi.org/
Acheulean transition at Olduvai Gorge. J. Hum. Evol. 120, 378–401. https://doi.org/ 10.1017/S0003598X00101255.
10.1016/j.jhevol.2017.11.010. Pop, C.M., 2016. Simulating lithic raw material variability in archaeological contexts: a
Mercader, J., Akuku, P., Boivin, N., Bugumba, R., Bushozi, P., Camacho, A., Carter, T., Re-evaluation and revision of brantingham’s neutral model. J. Archaeol. Method
Clarke, S., Cueva-Temprana, A., Durkin, P., Favreau, J., Fella, K., Haberle, S., Theor 23, 1127–1161. https://doi.org/10.1007/s10816-015-9262-y.
Hubbard, S., Inwood, J., Itambu, M., Koromo, S., Lee, P., Mohammed, A., Potts, R., 1988. Early Hominid Activities at Olduvai. Aldine de Gruyter, New York.
Mwambwiga, A., Olesilau, L., Patalano, R., Roberts, P., Rule, S., Saladie, P., https://doi.org/10.4324/9780203700969.
Siljedal, G., Soto, M., Umbsaar, J., Petraglia, M., 2021. Earliest Olduvai hominins

18
J. Favreau Quaternary Science Advances 9 (2023) 100068

Potts, R., Behrensmeyer, A.K., Ditchfield, P., 1999. Paleolandscape variation and early Semaw, S., Rogers, M.J., Cáceres, I., Stout, D., Leiss, A.C., 2018. The early acheulean
Pleistocene hominid activities: members 1 and 7, Olorgesailie formation, Kenya. ~1.6–1.2 Ma from Gona, Ethiopia: issues related to the emergence of the acheulean
J. Hum. Evol. 37, 747–788. https://doi.org/10.1006/jhev.1999.0344. in Africa. In: Gallotti, R., Mussi, M. (Eds.), The Emergence of the Acheulean in East
Potts, R., Behrensmeyer, A.K., Faith, J.T., Tryon, C.A., Brooks, A.S., Yellen, J.E., Deino, A. Africa and beyond: Contributions in Honor of Jean Chavaillon. Springer, Cham,
L., Kinyanjui, R., Clark, J.B., Haradon, C., Levin, N.E., Meijer, H.J.M., Veatch, E.G., pp. 115–128. https://doi.org/10.1007/978-3-319-75985-2_6.
Owen, R.B., Renaut, R.W., 2018. Environmental dynamics during the onset of the Semaw, S., Rogers, M.J., Simpson, S.W., Levin, N.E., Quade, J., Dunbar, N., McIntosh, W.
middle stone age in eastern Africa. Science 360, 86–90. https://doi.org/10.1126/ C., Cáceres, I., Stinchcomb, G.E., Holloway, R.L., Brown, F.H., Butler, R.F., Stout, D.,
science.aao2200. Everett, M., 2020. Co-occurrence of Acheulian and Oldowan artifacts with Homo
Presnyakova, D., Braun, D.R., Conard, N.J., Feibel, C., Harris, J.W., Pop, C.M., erectus cranial fossils from Gona, Afar, Ethiopia. Sci. Adv. 6, eaaw4694 https://doi.
Schlager, S., Archer, W., 2018. Site fragmentation, hominin mobility and LCT org/10.1126/sciadv.aaw4694.
variability reflected in the early Acheulean record of the Okote Member, at Koobi Shackley, M.S., 1998. Gamma rays, X-rays and stone tools: some recent advances in
Fora, Kenya. J. Hum. Evol. 125, 159–180. https://doi.org/10.1016/j. archaeological geochemistry. J. Archaeol. Sci. 25, 259–270. https://doi.org/
jhevol.2018.07.008. 10.1006/jasc.1997.0247.
Price, J.R., Velbel, M.A., 2003. Chemical weathering indices applied to weathering Shackley, M.S., 2008. Archaeological petrology and the archaeometry of lithic materials.
profiles developed on heterogeneous felsic metamorphic parent rocks. Chem. Geol. Archaeometry 50, 194–215. https://doi.org/10.1111/j.1475-4754.2008.00390.x.
202, 397–416. https://doi.org/10.1016/j.chemgeo.2002.11.001. Shackley, M.S. (Ed.), 2011. X-Ray Fluorescence Spectrometry (XRF) in Geoarchaeology.
Prieto, A., Aldea-Moreira, X., Arzarello, M., Berruti, G.L.F., Caracausi, S., Daffara, S., de Springer, New York. https://doi.org/10.1007/978-1-4419-6886-9.
la Peña, P., Favreau, J., García-Rojas, M., Huysecom, E., Janardhana, B., Jha, D.K., Shimelmitz, R., 2015. The recycling of flint throughout the lower and middle paleolithic
Lahari, L., Molefyane, T.R., Pruvost, C., Rodríguez-Álvarez, X.P., Thomas, M., sequence of tabun cave, Israel. Quat. Int. 361, 34–45. https://doi.org/10.1016/j.
Vaishnav, H.K., Villeneuve, Q., de Lombera-Hermida, A., 2022. How to deal with an quaint.2014.08.033.
elephant in the room? Understanding “non-flint” raw materials: characterisation and Shott, M.J., 1989. On tool-class use lives and the formation of archaeological
technological organisation. Revista ArkeoGazte Aldizkaria 12, 73–98. https://arkeog assemblages. Am. Antiq. 54, 9–30. https://doi.org/10.2307/281329.
azte.org/wp-content/uploads/2022/07/3-MON-Prieto-et-al..pdf. Shotton, F.W., Hendry, G.L., 1979. The developing Field of petrology in archaeology.
Rapp, G., 2009. Archaeomineralogy, second ed. Springer, Berlin. https://doi.org/ J. Archaeol. Sci. 6, 75–84. https://doi.org/10.1016/0305-4403(79)90034-7.
10.1007/978-3-540-78594-1. Skarpelis, N., Carter, T., Contreras, D.A., Mihailović, D.D., 2017. Characterization of the
Redmount, C.A., Morgenstein, M.E., 1996. Major and trace element analysis of modern siliceous rocks at Stélida, an early prehistoric lithic quarry (Northwest Naxos,
Egyptian pottery. J. Archaeol. Sci. 23, 741–762. https://doi.org/10.1006/ Greece), by petrography and geochemistry: a first step towards chert sourcing.
jasc.1996.0070. J. Archaeol. Sci.: Report 12, 819–833. https://doi.org/10.1016/j.
Reeves, J.S., Braun, D.R., Finestone, E.M., Plummer, T.W., 2021. Ecological perspectives jasrep.2016.11.015.
on technological diversity at Kanjera South. J. Hum. Evol. 158, 103029 https://doi. Smith, R.J., Wood, B., 2017. The principles and practice of human evolution research:
org/10.1016/j.jhevol.2021.103029. are we asking questions that can be answered? Comptes Rendus Palevol 16,
Reimer, R., 2018. Lithic sourcing in Canada. Can. J. Archaeol. 42, 137–143. https:// 670–679. https://doi.org/10.1016/j.crpv.2016.11.005.
www.jstor.org/stable/44878257. Soto, M., Favreau, J., Campeau, K., Carter, T., Abtosway, M., Bushozi, P.M., Clarke, S.,
Resom, A., Asrat, A., Gossa, T., Hovers, E., 2018. Petrogenesis and depositional history of Durkin, P.R., Hubbard, S.M., Inwood, J., Itambu, M., Koromo, S., Larter, F., Lee, P.,
felsic pyroclastic rocks from the Melka Wakena archaeological site-complex in South Mwambwiga, A., Nair, R., Olesilau, L., Patalano, R., Tucker, L., Mercader, J., 2020a.
central Ethiopia. J. Afr. Earth Sci. 142, 93–111. https://doi.org/10.1016/j. Fingerprinting of quartzitic outcrops at Oldupai Gorge, Tanzania. J. Archaeol. Sci.:
jafrearsci.2018.03.003. Report 29, 102010. https://doi.org/10.1016/j.jasrep.2019.102010.
Reti, J.S., 2016. Quantifying oldowan stone tool production at olduvai Gorge, Tanzania. Soto, M., Favreau, J., Campeau, K., Carter, T., Durkin, P.R., Hubbard, S.M., Nair, R.,
PLoS One 11, e0147352. https://doi.org/10.1371/journal.pone.0147352. Bushozi, P.M., Mercader, J., 2020b. Systematic sampling of quartzites in sourcing
Roche, H., Brugal, J.-P., Lefevre, D., Ploux, S., Texier, P.-J., 1988. Isenya: état des analysis: intra-outcrop variability at Naibor Soit, Tanzania (Part I). Archaeological
recherches sur un nouveau site acheuléen d’Afrique orientale. Afr. Archaeol. Rev. 6, and Anthropological Sciences 12, 100. https://doi.org/10.1007/s12520-020-01054-
27–55. https://doi.org/10.1007/BF01117111. w.
Roche, H., de la Torre, I., Arroyo, A., Brugal, J.-P., Harmand, S., 2018. Naiyena Engol 2 Stammers, R.C., Caruana, M.V., Herries, A.I.R., 2018. The first bone tools from
(West Turkana, Kenya): a case study on variability in the oldowan. Afr. Archaeol. Kromdraai and stone tools from Drimolen, and the place of bone tools in the South
Rev. 35, 57–85. https://doi.org/10.1007/s10437-018-9283-5. African Earlier Stone Age. Quat. Int. 495, 87–101. https://doi.org/10.1016/j.
Rollinson, H.R., 1993. Using Geochemical Data: Evaluation, Presentation, Interpretation. quaint.2018.04.026.
Routledge, London. Stephens, J.A., Ducea, M.N., Killick, D.J., Ruiz, J., 2021. Use of non-traditional heavy
Rooney, T.O., 2017. The Cenozoic magmatism of East-Africa: Part I — flood basalts and stable isotopes in archaeological research. J. Archaeol. Sci. 127, 105334 https://doi.
pulsed magmatism. Lithos 286–287, 264–301. https://doi.org/10.1016/j. org/10.1016/j.jas.2021.105334.
lithos.2017.05.014. Stern, N., 1994. The implications of time-averaging for reconstructing the land-use
Rooney, T.O., 2020a. The cenozoic magmatism of East Africa: Part II – rifting of the patterns of early tool-using hominids. J. Hum. Evol. 27, 89–105. https://doi.org/
mobile belt. Lithos 360–361, 105291. https://doi.org/10.1016/j. 10.1006/jhev.1994.1037.
lithos.2019.105291. Stiles, D., 1998. Raw material as evidence for human behaviour in the Lower Pleistocene:
Rooney, T.O., 2020b. The cenozoic magmatism of East Africa: Part III–rifting of the the Olduvai case. In: Petraglia, M.D., Korisettar, R. (Eds.), Early Human Behaviour in
craton. Lithos 360–361, 105390. https://doi.org/10.1016/j.lithos.2020.105390. Global Context: the Rise and Diversity of the Lower Palaeolithic Record. Routledge,
Rooney, T.O., 2020c. The cenozoic magmatism of East Africa: Part V – magma sources London, pp. 130–145. https://doi.org/10.4324/9780203203279.
and processes in the East African Rift. Lithos 360–361, 105296. https://doi.org/ Stiles, D.N., Hay, R.L., O’Neil, J.R., 1974. The MNK chert factory site, olduvai Gorge,
10.1016/j.lithos.2019.105296. Tanzania. World Archaeol. 5, 285–308. https://doi.org/10.1080/
Sampson, C.G., 2006. Acheulian quarries at hornfels outcrops in the Upper Karoo region 00438243.1974.9979575.
of South Africa. In: Goren-Inbar, N., Sharon, G. (Eds.), Axe Age: Acheulian Stollhofen, H., Stanistreet, I.G., Toth, N., Schick, K.D., Rodríguez-Cintas, A., Albert, R.M.,
Toolmaking from Quarry to Discard. Equinox, London, pp. 75–107. Farrugia, P., Njau, J.K., Pante, M.C., Herrmann, E.W., Ruck, L., Bamford, M.K.,
Sánchez-Yustos, P., Diez-Martín, F., Domínguez-Rodrigo, M., Fraile, C., Duque, J., Blumenschine, R.J., Masao, F.T., 2021. Olduvai’s oldest Oldowan. J. Hum. Evol. 150,
Uribelarrea, D., Mabulla, A., Baquedano, E., 2016. Techno-economic human 102910 https://doi.org/10.1016/j.jhevol.2020.102910.
behavior in a context of recurrent megafaunal exploitation at 1.3 Ma. Evidence from Stoner, W.D., Shaulis, B.J., 2021. Chemical mapping to evaluate post-depositional
BK4b (Upper Bed II, Olduvai Gorge, Tanzania). J. Archaeol. Sci.: Report 9, 386–404. diagenesis among the earliest ceramics in the Teotihuacan Valley, Mexico. Minerals
https://doi.org/10.1016/j.jasrep.2016.08.019. 11, 384. https://doi.org/10.3390/min11040384.
Santonja, M., Panera, J., Rubio-Jara, S., Pérez-González, A., Uribelarrea, D., Domínguez- Stout, D., Quade, J., Semaw, S., Rogers, M.J., Levin, N.E., 2005. Raw material selectivity
Rodrigo, M., Mabulla, A.Z.P., Bunn, H.T., Baquedano, E., 2014. Technological of the earliest stone toolmakers at Gona, Afar, Ethiopia. J. Hum. Evol. 48, 365–380.
strategies and the economy of raw materials in the TK (Thiongo Korongo) lower https://doi.org/10.1016/j.jhevol.2004.10.006.
occupation, Bed II, Olduvai Gorge, Tanzania. Quat. Int. 322–323, 181–208. https:// Stross, F.H., Hay, R.L., Asaro, F., Bowman, H.R., Michel, H.V., 1988. Sources of the
doi.org/10.1016/j.quaint.2013.10.069. quartzite of some ancient Egyptian sculptures. Archaeometry 30, 109–119. https://
Sayre, E.V., Dodson, R.W., 1957. Neutron activation study of mediterranean potsherds. doi.org/10.1111/j.1475-4754.1988.tb00439.x.
Am. J. Archaeol. 61, 35–41. https://doi.org/10.2307/501078. Stukeley, W., 1740. Stonehenge, a Temple Restor’d to the British Druids. W. Innys & R.
Schlüter, T., 1997. Geology of East Africa. Gebrüder Borntraeger, Berlin. Manby, London.
Sciuto, C., Geladi, P., La Rosa, L., Linderholm, J., Thyrel, M., 2019. Hyperspectral Sulzman, E.W., 2007. Stable isotope chemistry and measurement: a primer. In:
imaging for characterization of lithic raw materials: the case of a mesolithic dwelling Michener, R., Lajtha, K. (Eds.), Stable Isotopes in Ecology and Environmental
in northern Sweden. Lithic Technol. 44, 22–35. https://doi.org/10.1080/ Science, second ed. Blackwell Publishing, Malden, pp. 1–21. https://doi.org/
01977261.2018.1543105. 10.1002/9780470691854.ch1.
Semaw, S., Renne, P., Harris, J.W.K., Feibel, C.S., Bernor, R.L., Fesseha, N., Mowbray, K., Sunyer, M.R., Torcal, R.M., Figueroa, J.P., Martínez-Moreno, J., Benito-Calvo, A., 2017.
1997. 2.5-million-year-old stone tools from Gona, Ethiopia. Nature 385, 333–336. Quartzite selection in fluvial deposits: the N12 level of Roca dels Bous (Middle
https://doi.org/10.1038/385333a0. Palaeolithic, southeastern Pyrenees). Quat. Int. 435, 49–60. https://doi.org/
Semaw, S., Rogers, M.J., Renne, P.R., Butler, R.F., Dominguez-Rodrigo, M., Stout, D., 10.1016/j.quaint.2015.09.010.
Hart, W.S., Pickering, T., Simpson, S.W., 2003. 2.6-Million-year-old stone tools and Suzuki, M., 1969. Fission track dating and uranium contents of obsidian (I). Quat. Res. 8,
associated bones from OGS-6 and OGS-7, Gona, Afar, Ethiopia. J. Hum. Evol. 45, 123–130. https://doi.org/10.4116/jaqua.8.123.
169–177. https://doi.org/10.1016/S0047-2484(03)00093-9. ten Bruggencate, R.E., Fayek, M., Brownlee, K., Milne, S.B., Hamilton, S., 2013.
A combined visual-geochemical approach to establishing provenance for pegmatite

19
J. Favreau Quaternary Science Advances 9 (2023) 100068

quartz artifacts. J. Archaeol. Sci. 40, 2702–2712. https://doi.org/10.1016/j. Vogel, N., Nomade, S., Negash, A., Renne, P.R., 2006. Forensic 40Ar/39Ar dating: a
jas.2013.02.008. provenance study of Middle Stone Age obsidian artifacts from Ethiopia. J. Archaeol.
ten Bruggencate, R.E., Fayek, M., Milne, B., Brownlee, K., 2014. Characterizing quartz Sci. 33, 1749–1765. https://doi.org/10.1016/j.jas.2006.03.008.
artefacts: a case study from Manitoba’s northern Boreal forest. Archaeometry 56, Wadley, L., Kempson, H., 2011. A review of rock studies for archaeologists, and an
913–926. https://doi.org/10.1111/arcm.12092. analysis of dolerite and hornfels from the Sibudu area, KwaZulu-Natal. South. Afr.
Texier, P.-J., 2018. Technological assets for the emergence of the acheulean? Reflections Humanit. 23, 87–107. https://www.sahumanities.org/index.php/sah/article/view
on the Kokiselei 4 lithic assemblage and its place in the archaeological context of /329.
West Turkana, Kenya. In: Gallotti, R., Mussi, M. (Eds.), The Emergence of the Weaver, J., Stross, F., 1965. Analysis by X-Ray Fluorescence of Some American
Acheulean in East Africa and beyond: Contributions in Honor of Jean Chavaillon. Obsidians, vol. 1. Contributions of the University of California Archaeological
Springer, Cham, pp. 33–52. https://doi.org/10.1007/978-3-319-75985-2_3. Research Facility, pp. 89–93.
Thomas, H.H., 1923. The source of the stones of Stonehenge. Antiq. J. 3, 239–260. Weigand, P.C., Harbottle, G., Sayre, E.V., 1977. Turquoise sources and source analysis:
https://doi.org/10.1017/S0003581500005096. mesoamerica and the Southwestern USA. In: Earle, T.K., Ericson, J.E. (Eds.),
Toth, N., 1985. The oldowan reassessed: a close look at early stone artifacts. J. Archaeol. Exchange Systems in Prehistory. Academic Press, New York, pp. 15–34. https://doi.
Sci. 12, 101–120. https://doi.org/10.1016/0305-4403(85)90056-1. org/10.1016/B978-0-12-227650-7.50008-0.
Trigger, B.G., 2006. A History of Archaeological Thought, second ed. Cambridge Wilson, L., 2007. Understanding prehistoric lithic raw material selection: application of a
University Press, Cambridge. https://doi.org/10.1017/CBO9780511813016. gravity model. J. Archaeol. Method Theor 14, 388–411. https://doi.org/10.1007/
Tryon, C.A., 2003. The Acheulian to MSA Transition: Tephrostratigraphic Context for s10816-007-9042-4.
Archaeological Change in the Kapthurin Formation. University of Connecticut. PhD Wocel, J., 1854. Archäologische parallelen. Sitzungsberichte der Kaiserlichen. Akademie
Dissertation. der Wissenschafien. Philosophisch-Historische Classe (Wien) 11, 716–761.
Tucker, L., Favreau, J., Itambu, M., Larter, F., Mollel, N., Mwambwiga, A., Patalano, R., Wojtczak, D., 2015. Cores on flakes and bladelet production, a question of recycling? The
Roberts, P., Soto, M., Mercader, J., 2020. Initial assessment of bioavailable strontium perspective from the Hummalian industry of Hummal, Central Syria. Quat. Int. 361,
at Oldupai Gorge, Tanzania: potential for early mobility studies. J. Archaeol. Sci. 155–177. https://doi.org/10.1016/j.quaint.2014.10.021.
114, 105066 https://doi.org/10.1016/j.jas.2019.105066. Zaid, S.M., Elbadry, O., Ramadan, F., Mohamed, M., 2015. Petrography and
van der Merwe, N.J., Masao, F.T., Bamford, M.K., 2008. Isotopic evidence for contrasting geochemistry of Pharaonic sandstone monuments in Tall San Al Hagr, Al Sharqiya
diets of early hominins Homo habilis and Australopithecus boisei of Tanzania. South Governorate, Egypt: implications for provenance and tectonic setting. Turk. J. Earth
Afr. J. Sci. 104, 153–155. http://hdl.handle.net/11427/27727. Sci. 24, 344–364. https://doi.org/10.3906/yer-1407-20.
Verri, G., Barkai, R., Bordeanu, C., Gopher, A., Hass, M., Kaufman, A., Kubik, P., Zaidner, Y., Centi, L., Prévost, M., Mercier, N., Falguères, C., Guérin, G., Valladas, H.,
Montanari, E., Paul, M., Ronen, A., Weiner, S., Boaretto, E., 2004. Flint mining in Richard, M., Galy, A., Pécheyran, C., Tombret, O., Pons-Branchu, E., Porat, N.,
prehistory recorded by in situ produced cosmogenic 10Be. Proc. Natl. Acad. Sci. USA Shahack-Gross, R., Friesem, D.E., Yeshurun, R., Turgeman-Yaffe, Z., Frumkin, A.,
101, 7880–7884. https://doi.org/10.1073/pnas.0402302101. Herzlinger, G., Ekshtain, R., Shemer, M., Varoner, O., Sarig, R., May, H.,
Verri, G., Barkai, R., Gopher, A., Hass, M., Kubik, P.W., Paul, M., Ronen, A., Weiner, S., Hershkovitz, I., 2021. Middle Pleistocene Homo behavior and culture at 140,000 to
Boaretto, E., 2005. Flint procurement strategies in the late lower palaeolithic 120,000 years ago and interactions with Homo sapiens. Science 372, 1429–1433.
recorded by in situ produced cosmogenic 10Be in tabun and Qesem caves (Israel). https://doi.org/10.1126/science.abh3020.
J. Archaeol. Sci. 32, 207–213. https://doi.org/10.1016/j.jas.2004.09.001.

20

You might also like