[go: up one dir, main page]

Academia.eduAcademia.edu
The Zooarchaeology of Pleistocene Africa Jessica C. Thompson , Alex Bertacchi Emily Y. Hallett , and Briana Pobiner , Hannah M. Keller , Introduction Zooarchaeology, otherwise called archaeozoology, is the study of animal bones from archaeological sites, with an emphasis on the interactions between modern humans or extinct human relatives (collectively “hominins”) and other animals. Africa has the longest archaeological record, potentially extending into the Pliocene (Harmand et al., 2015) and spanning the Pleistocene. However, most Pleistocene archaeological sites do not preserve vertebrate fossils, which represent rich sources of paleoenvironmental and behavioral data. This chapter begins with a summary of how these remains, when they are available, are studied by zooarchaeologists. These analyses typically include some measure of taxonomic and skeletal part abundance and, in some cases, an assessment of the taphonomic history of an assemblage – the processes it underwent on its way to becoming a collection of fossils. We then offer a region-by-region summary of J. C. Thompson (*) Department of Anthropology, Yale University, New Haven, CT, USA Division of Anthropology, Yale Peabody Museum, New Haven, CT, USA e-mail: jessica.thompson@yale.edu A. Bertacchi · H. M. Keller Department of Anthropology, Yale University, New Haven, CT, USA E. Y. Hallett Department of Anthropology, Loyola University Chicago, Chicago, IL, USA Pan-African Evolution Research Group, Max Planck Institute for the Science of Human History, Jena, Germany B. Pobiner Department of Anthropology, National Museum of Natural History, Smithsonian Institution, Washington, DC, USA © The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 A. Beyin et al. (eds.), Handbook of Pleistocene Archaeology of Africa, https://doi.org/10.1007/978-3-031-20290-2_126 1955 1956 J. C. Thompson et al. the current knowledge of Pleistocene zooarchaeological assemblages in Africa, by compiling data from published records in books and journal articles available as of the end of 2020. We close with a discussion of these trends and the emergence of key debates and theoretical/methodological advances within African zooarchaeology. Zooarchaeological Methods of Data Collection and Analysis Taxonomic and Skeletal Part Representation The most fundamental type of data from a faunal assemblage is a list of identified taxa, which draws from a list of identified skeletal elements and portions of elements recovered from sites. Zooarchaeology usually relies on paleontological methods to identify fossils based on their morphology, and, then, these fossils are quantified according to a variety of measures to know how many specimens, skeletal parts, and animals they represent (Lyman, 2008). More recently, zooarchaeologists have begun to employ molecular methods to taxonomically identify fossils and check whether they preserve ancient DNA or proteins (Horsburgh et al., 2016; Prendergast et al., 2017). Zooarchaeological assemblages may contain remains of large or small mammals, reptiles, birds, amphibians, fish, shellfish, or even insects. The most diverse group of large-bodied mammals (>5 kg), the ungulates, dominate most Pleistocene African zooarchaeological reports and are our focus here. Their dominance is for five primary reasons: (1) Africa houses substantial ungulate diversity by virtue of much of its landmass being in the tropical belt where ideal habitats for the survival of these mammals existed for a long time; (2) these taxa preserve well and are recovered through most archaeological recovery strategies because they possess dental and sometimes bone elements that can survive longer after the animals are dead; (3) they are readily identifiable at least at the family level, often more specifically through their differentiated dental and horn core elements but also because they have extant representatives; (4) they have a range of habitat preferences that can derive from a range of past ecological conditions; and (5) they are inferred to have been the likely targets of ancient hominin subsistence behavior. The most basic unit of quantification in zooarchaeology is the NSP or number of specimens. These key terms in bold are the most common measures of zooarchaeological abundances. The NISP, or number of identified specimens, may equal the NSP or may be a subset that has been further identified (Lyman, 2008). For example, the NISP for large mammals in an assemblage could reference all specimens from large mammals, all specimens from large mammals that could be identified to a skeletal part, or all specimens from large mammals that could be identified to a particular taxonomic level. This level of specificity is not always clear in all reports, but knowing the basis for the NISP is an essential first step in comparing assemblages. Because elements tend to become less identifiable as they become more fragmented (and therefore lose more of their The Zooarchaeology of Pleistocene Africa 1957 morphological features), the ratio of NSP to NISP is an expedient way to measure fragmentation. Other units of faunal abundance are not counts of specimens but rather are derived measures. The MNE (minimum number of elements), is the minimum number of a given skeletal part that is represented in an assemblage. This can be just the element (e.g., the MNE for radius) or a more specific category (e.g., the MNE for the left radius). As with the NISP, there is no standard protocol for calculating the MNE (Marean et al., 2001). Some analysts count the number of bone portions (e.g., three complete portions of the proximal end of a right radius can only be an MNE of three right radii because each one has only one proximal portion). Other researchers look more carefully to see which portions overlap, even if they are not complete, because any area of overlap must represent two different animals (e.g., three incomplete portions of the proximal end of a right radius could represent one, two, or three animals at minimum, irrespective of whether the portions that are present overlap). Where overlaps do not exist, it may still be possible to count as separate animals the same elements if they come from individuals of different body sizes, ages at death, or distinct archaeological contexts/layers. The MNI, or minimum number of individuals, is obtained by dividing the MNE of the most frequent element for a certain taxon, also taking into account its side and age, by the number of times that the element occurs in a complete skeleton (Binford, 1978). MNEs of eight right radii, three right femora, and four left tibiae all from the same type of animal, body size, age, and context would mean that at least eight individuals are present – an MNI of eight. A range of other quantitative analyses can provide more information about how the nutrients associated with various skeletal parts influenced the ways in which animals were butchered and potentially how decisions were made to transport those parts back to a central location (Lupo, 2001). The minimum animal unit (MAU) for a given element is the MNE (without taking into account its side – right or left), divided by the number of times that the element occurs in a complete skeleton. For example, the MAU derived from an MNE of six for all femora (right and left) would be three because the femur occurs twice in the body. The value of %MAU is calculated by dividing the MAU for an element by the MAU for the most frequent element and then multiplying it by 100 (Binford, 1978). An analogous measure is the normed NISP, which starts with NISP rather than MNE and therefore skips the analytical steps and assumptions that underlie MNE calculations (Reynard et al., 2016). Because different skeletal parts are associated with different nutritional quantities in a living animal, a relative representation of skeletal elements converts dry fossils into more meaningful assessments of their roles in ancient subsistence systems (Faith & Gordon, 2007; Morin, 2007). In contrast to these in-depth measures that often include many fragments, a simple presence or absence in a taxonomic list only requires that at least one specimen from each listed taxon be present in an assemblage. This is the most basic type of zooarchaeological data, and, because different species have different habitat preferences, the presence of particular taxa in an assemblage can reveal past 1958 J. C. Thompson et al. environmental or ecological conditions (Faith & Lyman, 2019). For example, the presence of hippopotamus (Hippopotamus spp.) indicates proximity to water, whereas the presence of wildebeest (Connochaetes spp.) indicates grassland. Taxonomic identifications may also inform about hominin diet, if it can be demonstrated that hominins were responsible for bringing remains to the site as part of their subsistence activities. However, interpreting taxonomic lists beyond these basic observations can quickly become complex. A high diversity of taxa may be reflective of many things: a more diverse mosaic of habitats, a severely time-averaged assemblage that samples diachronic habitat changes, a larger assemblage in which rarer taxa are statistically more likely to be sampled, a more diverse set of bone accumulators (different carnivores, scavengers, hominins, etc.), a change in prey selection by the same primary predator, or even a change in recovery/analysis methods (for example, sieve gap size or analytical size cutoffs) by archaeologists so that certain taxa are better represented. For these reasons, quantification of faunal assemblages lies at the nexus between theory and method, and higher-order interpretations should always take place within a framework that is informed by an understanding of the depositional history of a site. Taphonomy Zooarchaeology emerged from paleontology, in which more emphasis is placed on the taxonomic identification of fossils (Gifford-Gonzalez, 2018). However, at Pleistocene sites where mobile groups of hominins often accumulated faunal remains at the same locations as other predators, scavengers, and even abiotic agents (such as streams), untangling the depositional history of an assemblage is challenging. For these reasons, Pleistocene African zooarchaeology has played an outsized role in the development of zooarchaeological methods (Thompson, 2020). This has especially been the case with respect to taphonomy, the study of the transformative processes affecting animal remains during and after deposition. Taphonomic processes can break, disassociate, dissolve, or otherwise modify an animal carcass from the time of death to the time of recovery as a fossil or a subfossil (FernandezJalvo & Andrews, 2016). Thus, they may bias reconstructions of the past by deleting information, but they may also add information about both the depositional and post-depositional contexts of the remains. Many taphonomic processes leave traces of the history of a bone assemblage in the form of fragmentation, burning patterns, or other bone surface modifications (BSMs; e.g., cut marks from stone tools, percussion marks from hammerstones, tooth marks from carnivores or hominins, dissolution from stomach acid or sediments, etchings from roots or fungi). However, they may also result in the systematic removal of certain bones or other hard parts (teeth, shells, etc.) from the assemblage or they may simply reduce existing fragments to a less identifiable state (Merritt & Davis, 2017). Some analysts spend considerable effort recording data The Zooarchaeology of Pleistocene Africa 1959 about the general skeletal parts, breakage patterns, and surface modifications on less identifiable fragments, while others choose to focus on more identifiable specimens. The level of understanding about how an assemblage formed influences the kinds of questions zooarchaeologists can ask about past behaviors. Different components of an assemblage can provide different information about taphonomic history, which can, in turn, lead to divergent interpretations of site formation and ancient behavior (Yravedra & Domínguez-Rodrigo, 2009). Specieslevel identifications are more common on teeth and horn cores and – at a higher taxonomic level – foot elements (Marean et al., 1992). These elements are associated with the least nutritive parts of the body (Pickering et al., 2003) and so may be the last ones to be butchered, gnawed, or transported by either humans or carnivores. However, they are also some of the densest parts of the skeleton (Lam et al., 2003) and may therefore have some of the highest chances of preservation. This relationship between “identifiability,” resistance to taphonomic degradation, and distribution of nutrients on an animal skeleton means that taphonomic processes and analytical methods can affect which skeletal parts are reported from an assemblage and in what frequency. Importantly, while analysts have no control over past taphonomic processes, they have substantial control over the methods they use. Variation in how analysts have reported assemblages across Africa makes it impossible to compare details of collections that may have been subjected to different taphonomic histories. However, because taxonomic identifications at the genus or species levels are dominated by dental and horn core elements, it is possible to make general comparisons of taxonomic changes in archaeofaunal assemblages over time. Taxonomic Patterns Archaeofaunal assemblages complement paleontological data by providing additional reference points for the geographic and chronological distribution of taxa, thus enabling a better understanding of past faunal community structure. The modern zoogeographic separation of African faunas into the Palearctic region (north of the Sahara) and the African region (south of the Sahara) can be traced back to the Middle Pleistocene or terminal Early Pleistocene (Geraads, 2010; O’Regan et al., 2005). Prior to this, there is evidence for deeper connections between these regions as well as between Africa and Eurasia (Werdelin & Sanders, 2010). The emergence of most modern taxonomic groups took place gradually over the last ~1 Ma (Faith et al., 2019). Early Pleistocene larger mammal communities generally included extinct taxa with no modern analogues, many of which persisted longer in subSaharan Africa than in northern Africa (O’Regan et al., 2005), such as Dinofelis, Homotherium, Smilodon (saber-toothed cats), Deinotherium (giant elephant relatives), Hipparion (three-toed horses), Sivatherium (short-necked giraffes), and Ancylotherium (clawed ungulates related to horses and rhinoceroses). The Early 1960 J. C. Thompson et al. Pleistocene also saw the introduction of certain genera into Africa, including Equus, which became ubiquitous components of African faunal assemblages before becoming restricted to sub-Saharan Africa today (Bernor et al., 2010). Thus, there were both taxonomic groups and community structures – along with the role of hominins within them – with no modern analogues (Faith et al., 2019). In northern Africa, Ennouchi (1962) described faunal remains from Jebel Irhoud in Morocco as typical of the Early (Lower) Pleistocene: Rhinoceros sp., Equus mauritanicus, Asinus africanus, Canis anthus, Gazella atlantica, Gazella cuvieri, Gazella dorcas, Alcelaphus bubalis, Connochaetes taurinus prognu, Bos taurus ibericus, and Bos primigenius. Subsequent work revealing the deposits to be Middle Pleistocene, ~300 ka, attests to the problematic chronologies available for earlier work (Richter et al., 2017). Early Pleistocene faunas in northern Africa are now known to include genera with no modern representatives, such as Parmularius, Rabaticeras, Numidocapra, Sivatherium, and Hipparion (Geraads, 2010). In general, species from the Late Pleistocene, which became locally extirpated in the Holocene, tended to be larger in body size than their modern descendants. For example, in northern Africa, Panthera leo leo (Atlas lion) is found throughout Pleistocene archaeological sites and is the largest subspecies of lion, roughly double the size of modern sub-Saharan African lions (Hallett, 2018). Following migration from Eurasia to North Africa, Ursus arctos crowtheri (Atlas bear) and Megaceroides algericus (extinct North African deer) both survived in the region into the Holocene (Fernandez et al., 2015; Merzoug & Sari, 2008). Time-averaging, especially of older deposits, makes it challenging to understand the pace and environmental context of major changes, but in eastern Africa, faunal communities comprising mostly modern genera had emerged by the end of the Middle Pleistocene (Potts et al., 2018). By the Late Pleistocene, extinct species were often present within modern genera. In eastern Africa, extinct taxa show how grasslands changed in character toward the end of the Pleistocene and into the Holocene. For example, in the Lake Victoria Basin, extinct species belonging to the extant genera of Aepyceros (Faith et al., 2014), Damaliscus (Faith et al., 2012), and Syncerus as well as the extinct genera Rusingoryx and Megalotragus (Faith et al., 2011, 2020) all show adaptations to grasslands, which, in many cases, were more arid than today. In terms of hominin exploitation of these faunas, there is a trend across the Middle Pleistocene for archaeofaunas to contain greater proportions of small-bodied ungulates, although this may be attributable to the increasing representation of rock shelter and cave sites later in time, where these fragile remains may be better preserved and more likely to sample adjacent rocky or forested habitats (Smith et al., 2019). Early Pleistocene faunas are poorly described from southern-central Africa and are mainly from paleontological sites. These show an abundance of woody vegetation and faunal associations, suggesting that the region was a long-term corridor between eastern and southern Africa (Lüdecke et al., 2016). Middle and Late Pleistocene large mammal communities in this area continue to offer information The Zooarchaeology of Pleistocene Africa 1961 about past biogeographic processes between the better studied eastern and southern regions. Extinct taxa include Equus capensis, Metridiochoerus sp., Syncerus antiquus, Megalotragus priscus, and Antidorcas bondi. Two teeth identified as the giant warthog Metridiochoerus sp. are reported from the “Tshangula” deposit at Redcliff Cave in Zimbabwe, which is dated to 25,650 ± 1,800 BP (26,355–33,822 cal yr BP; Cruz-Uribe, 1983). If confirmed, this would be the last known occurrence of the genus (see also Faith [2014]). The giant alcelaphine Megalotragus priscus is also attested at Redcliff, while the only other possible occurrence in the region is reported from an early excavation at Leopard’s Hill in Zambia for which there is no information about the stratigraphic context (Cooke, 1950). The gazelle Eudorcas thomsonii is an eastern African species that today does not occur south of central Tanzania but is attested in the Late Pleistocene deposits at Leopard’s Hill and Kalemba in Zambia. On the other hand, the blesbok Damaliscus lunatus is a southern African species that is not attested north of the Limpopo River today, but its occurrence is documented at Redcliff and Twin Rivers. The mountain reedbuck Redunca fulvorufula is absent from southerncentral Africa today, with its distribution broken in two main populations in eastern and southern Africa. Its abundant fossils throughout the Late Pleistocene sequence at Redcliff suggest that habitat fragmentation for this species might have been a recent phenomenon (Coetzee & van Zinderen Bakker, 1970; Happold & Happold, 1989). In southern Africa, based on a study of 25 Plio-Pleistocene deposits from South Africa, there were multiple iterations of large mammal community formation in the Early Pleistocene, including a large mammal turnover between 3.0 and 2.0 million years ago (Ma), more specifically between 2.61 and 2.4 Ma (Hanon et al., 2019). The genera Tragelaphus and Antidorcas (bovids), Papio (baboon), and Equus (zebra) were particularly diverse and represented by multiple species between 2.5 and 0.5 Ma. Paleoecological reconstructions from specific sites during the Early Pleistocene include forests or closed woodlands, shrublands/bushlands, edaphic grasslands, wooded grasslands, and more dry and open environments. The taxonomy of bovids in particular, as well as suids, indicates a shift from more wet-closed to more open-dry (grassland) environments at about 2.5 Ma. Taxonomic patterns reveal continued expansions and contractions of grassland ecosystems in the Middle and Late Pleistocene. On the southern coast of Africa, cave sites preserve fossils from large gregarious ungulates in a region that today supports mainly small, territorial taxa. In this case, sea-level changes in the Middle to Late Pleistocene offered access to the Paleo-Agulhas Plain, an extinct grazing ecosystem on an expanse of continental shelf that is now submerged under sea water (Marean et al., 2020). The impact of major climatic and environmental changes is recorded in faunal assemblages near the end of the Pleistocene across southern Africa, based on the reduced frequency of largebodied grazing bovids and the increase of small-bodied browsers at Buffelskloof, 1962 J. C. Thompson et al. Nelson Bay Cave, and Equus Cave (Faith, 2011b; Klein, 1972, 1978a; Klein et al., 1991). At the end of the Late Pleistocene, habitat loss caused some large-bodied grazers to become locally extirpated (e.g., Connochaetes gnou) or extinct (e.g., Equus capensis and Syncerus antiquus). The blue antelope (Hippotragus leucophaeus) is an interesting exception that survived in small refugia before going extinct in historic times (Faith & Thompson, 2013). Another example of the loss of productive grazing ecosystems comes from the now arid interior, where Faith (2013) found a taxonomically rich landscape with ample hunting opportunities near Boomplaas Cave during the Last Glacial Maximum, in a region that lost much of its large grazer biodiversity in the Holocene. As researchers continue to fill in details about faunal turnover and exchange, it has become clear that it is difficult to generalize about environmental conditions and their effects on hominin–faunal relationships even within smaller subregions. Some highly productive ecosystems no longer exist nor do they have close modern analogues. Faunal turnovers may be masked or exaggerated by time-averaged assemblages, and conditions may have changed over short geographic distances (Faith et al., 2021). Future work that resolves more of this detail will be essential for understanding the diversity of faunal species that promoted hominin survival in Pleistocene Africa and their emergent subsistence behavior. Patterns in Pleistocene Zooarchaeological Research Overview Pleistocene zooarchaeological assemblages in Africa are widely distributed but concentrated in the corridors where the most intensive archaeological research has occurred: the Mediterranean coast, the Nile River corridor, the Rift Valley of eastern Africa, and the cone of southern Africa (Fig. 1). All site locations are derived from original published sources, including maps or public repositories, according to the criteria set out in the following section. Regions are purely geographically defined, and, although they do follow modern political boundaries for ease of presentation, they do not reflect historical, linguistic, or other cultural groupings. We use the current International Commission on Stratigraphy start point for the Lower (Early) Pleistocene at 2.58 Ma, the start of the Middle Pleistocene at 781 thousand years ago (ka), and the start and end points of the Upper (Late) Pleistocene at 129 ka and 11.7 ka, respectively (Cohen et al., 2013). Sites are only included in our analysis if they contain vertebrate remains from the Pleistocene portions of the deposits, in association with other indicators of human behavior (such as stone tools or butchery marks), and if they have come from excavated contexts. This excludes localities that are purely paleontological, multicomponent sites where archaeofaunal assemblages are present but are not Pleistocene in age, The Zooarchaeology of Pleistocene Africa 1963 Fig. 1 A map of the African mainland showing the regions of Africa as defined in this chapter, with locations of Pleistocene archaeofaunas that meet our search criteria, published as of the end of 2020 and surface collections in which there is no clear association between Pleistocene archaeological materials and faunal remains. Similarly, we exclude dissertations, theses, unpublished reports, or other “gray literature” as the primary sources of information, in order to facilitate evaluation of trends in the most easily available zooarchaeological data. We make an exception when a dissertation is the only available source of detail about a faunal assemblage, but the presence of that assemblage is known from the literature that otherwise satisfies the earlier criteria – for example, where a faunal assemblage is mentioned in a journal article but the assemblage size is only available in a dissertation. This analysis differs from other compilations that are more region-specific, include paleontological sites, or trade a broad chronological span of the Pleistocene for more in-depth discussions 1964 J. C. Thompson et al. of key sites (e.g., Jousse 2017; Plug and Badenhorst 2001; Smith et al. 2019; Van Neer, 1989). Table 1 provides the details of each site, including an estimate of the precision for each location. Locations on the individual regional maps are numerically matched with the sites in this Table. Table 1 also includes the year(s) in which the site was excavated and whether the assemblage is associated with stone tools of general “Earlier” (ESA), “Middle” (MSA), or “Later” Stone (LSA) Age technologies. Under these headings, we include the northern African equivalents of “Lower/ Early,” “Middle” (or Mousterian), and “Late” Paleolithic, which have a different naming system for historical reasons. We also include specific age constraints when they are available, and the general methods used to obtain those dates. Dating methods are simplistically categorized to offer general information about the empirical basis of assemblage ages. “Chronometric” refers to any dating method that is based on physical or chemical phenomena and which provides numerical age estimations in calendar years before present (e.g., argon–argon, potassium–argon, uranium–thorium, uranium–lead, radiocarbon, optically stimulated luminescence, thermoluminescence, electron spin resonance, obsidian hydration, etc.) (Richter & Wagner, 2015). We report these numerical ages as ranges and include ages based on correlation with chronometrically dated materials (such as the tephra correlation, in which, for example, a volcanic deposit may be directly dated at Locality A, then found again at Locality B, but is not directly dated again at Locality B). The other three approaches we list are commonly used relative methods that should be treated as having increasing levels of uncertainty in this order: “biochronology,” “paleomagnetic,” and “typology.” Typology is inferred as a method for all sites, because they all have associated artifacts, and, when it is specifically listed, it is because that is the only dating method that was used. Sites that are “undated” have no numerical age range estimates provided in the literature. Figure 2 shows the distribution of ages of Pleistocene archaeofaunas in different African regions. We also offer data on which assemblages are associated with Pleistocene hominin remains (in an excavated context or surface-collected context adjacent to an excavated faunal assemblage and likely part of it) and what classes of vertebrate faunas have been reported. Additional information includes a general assessment of assemblage size as an order of magnitude, augmented in the comments with any direct quotes from the relevant reference. Finally, we note where there has been a formal taphonomic analysis (specifically designed to understand assemblage formation), which assemblages have reported butchery marks, and which have reported bone tools. These are important distinctions because some assemblages may have reports of cut marks but still lack a formal taphonomic analysis that reconstructs assemblage formation histories more broadly, such as fragmentation, disarticulation, transport, carnivore and hominin involvement, and post-depositional processes. References are not comprehensive for a site but rather only relate to the information provided in Table 1. In compiling Table 1, it became clear that the concept of an “assemblage” is not always straightforward. For example, an assemblage may include all remains from a site, a single layer within a site, or even a subset of specific kinds of animal Table 1 All Pleistocene archaeofaunal assemblages in Africa published as of the end of 2020. Locations are rounded to nearest three decimals. Other information is keyed at the end of the table Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments No No Yes LPH 12.5– 12.6 Chron No V 1000s Yes Yes No Campmas et al. (2016)* and Chibane (2016) OA late Yes nineteenth century, 1931–1947, 1998–present No No EP 1900– 2400 Chron, BChron No V 100s Yes Yes No Cáceres et al. (2015a), Clark (1967)*, and Sahnouni et al. (2011, 2018) 5.656 OA 1947; Yes 1990s–2000s No No EP 1780– 1950 BChron, No PMag, Type V 1000s Yes Yes No Cáceres et al. (2015a), and Sahnouni et al. (2002, 2010, 2011, 2013)* 36.202 5.656 OA 2000s Yes No No EP 1770– 1950 BChron, No PMag, Type M, B 1000s Yes Yes No Sahnouni et al. (2010, 2013)* Algeria 25.267 6.533 OA 1948, 1950s, Yes 1960s, 1970s No No EPMP Undated BChron, No Type V Unknown No NA No Hocine (2016)* and Thomas (1977) Es Sayar Algeria 35.170 4.133 OA 1970s No No Yes LP 13.1– 16.5 No V Unknown No NA No Amara (1977), Close (1980)*, and Sari (2020) 7 Gisement Aterien des Phacocheres Algeria 36.757 3.030 C 1960s, 1980s– present No Yes No LP Undated Type No V Unknown Yes Yes No Campmas (2017)* and Hadjouis (1985, 1994) 8 Grotte Madeleine, Taza Cave Algeria 36.713 5.543 C 1920s, 1986–1990s No Yes Yes LP 12 - >39 Chron Yes V Unknown No NA No Close (1980)* and Meier et al. (2003) 9 Lac Karar Algeria 35.250 −0.917 OA 1900s Yes Yes No EPMP Undated BChron No V Unknown No NA No Boule (1900) and Clark (1967)* Source site Country Lat Long 1 Abri Alain Algeria 35.700 −0.700 RS 1927–1930; 1934 2 Ain Boucherit Algeria 36.167 5.667 3 Ain Hanech Algeria 36.202 4 El Kherba Algeria 5 Erg Tihodaine 6 Chron References (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age 10 Mansourah Algeria 36.350 6.600 OA nineteenth century; mid twentieth century; early twentieth century Yes No No PLEP 11 Oued Bousmane Algeria 35.575 8.217 RS 2014 No Yes Yes 12 Oued Djebbana (Bir el-Ater) Algeria 34.800 8.060 OA 1920s No Yes 13 Tamar Hat Algeria 36.650 5.430 RS 1930s; 1950s; 1960s; 1973 No 14 Tighennif (form. Ternifine) Algeria 35.415 0.329 OA 1882, 1954–1956, 1981–1983 15 Bir Sahara East BS-14 Egypt 22.626 28.717 OA 1988 Dates (ka) Hominin remains Vert Fauna Size Undated BChron No V Unknown No NA No LPH Undated Type No U 100s No NA No No LP > 35 Chron No V Unknown No Yes No No Yes LP 12–25 Chron No V 1000s Yes Yes Yes Close (1980)*, Hogue and Barton (2016), Merzoug (2005, 2012), Merzoug and Sari (2008), and Sari and Kim (2017) Yes No No MP 800 PMag Yes V Unknown No No No Geraads et al. (1986)*, Geraads (2012), and Tong (1986) No Yes No LP 101–105 Chron No V 10s NA No Dating Bone Taph Butch Tools Comments No Oldowan only References Bayle (1854), Chaid-Saoudi et al. (2006)*, Joleaud (1918), Sahnouni et al. (2011), and Thomas (1884) Bahra et al. (2020)* Invertebrate taphonomy only Poorly preserved Close (1980), Dibble et al. (2013), Morel (1974a*, b), Steele et al. (2019), and Vanhaeren et al. (2006) Gautier (1993), Issawi (1993)*, and Nicoll (2018) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 16 Bir Sahara East E-88-1 Egypt 22.626 28.717 OA 1986 No Yes No MP 175 Chron No M, B 10s No NA No 17 Bir Tarfawi BT 86-2 Egypt 22.920 28.750 OA 1973–1974; 1985–1988 No Yes No LP 85 Chron No V 10s Yes No No Gautier (1993), Nicoll (2018), and Schild and Wendorf (1993)* 18 Bir Tarfawi BT 87-2 Egypt 22.920 28.750 OA 1973–1974; 1985–1988 No Yes No LP 102.2 Chron No V 10s Yes No No Gautier (1993), Nicoll (2018), and Schild and Wendorf (1993)* 19 Bir Tarfawi BT E-87-5 Egypt 22.920 28.750 OA 1973–1974; 1985–1988 No Yes No MP 139–154 Chron No V 10s Yes No No Gautier (1993) 20 Bir Tarfawi BT-14 Egypt 22.920 28.750 OA 1973–1974; 1985–1988 No Yes No LP 85–135 Chron No V 10,000s Yes Yes No Bluszcz (1993); Gautier (1980, 1993), Kowalski et al. (1993), Nicoll (2018), and Van Neer (1993) 21 Bir Tarfawi E-87-1 Egypt 22.920 28.750 OA 1973–1974; 1985–1988 No Yes No MP 140–162 Chron No V 10s Yes No No Gautier (1993), Schild and Wendorf (1993)*, and Schwarcz and Grün (1993) 22 Bir Tarfawi E-87-4 Egypt 22.920 28.750 OA 1973–1974; 1985–1988 No Yes No MP 148 Chron No V 10s Yes No No Gautier (1993), Schild and Wendorf (1993)*, and Schwarcz and Grün (1993) 23 Dakhleh Oasis 348 Egypt 25.510 29.120 OA 1977–present Yes Yes Yes MP- 200–400 Chron, H Type No V Unknown No No No Churcher et al. (2008) and Kato et al. (2014) Few References Close and Wendorf (1993) and Issawi (1993)* (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated Hominin remains Vert Fauna Size 24 Dakhleh Oasis 357 Egypt 25.510 29.120 OA MP- 200–400 Chron, H Type No V Unknown No No No 25 Dishna E6104 Egypt 26.100 32.500 Yes LPH Undated Type No V 10s No NA No Not many Gautier (1976), Paulissen and Vermeersch (1987)*, andWendorf and Schild (1976) 26 Dishna E61M10 Egypt 26.100 No Yes LPH Undated Type No V 10s No NA No Not many Gautier (1976), Paulissen and Vermeersch (1987)*, and Wendorf and Schild (1976) 27 Dishna E61M9 Egypt No late 1960s–1980s No Yes LPH Undated Type No V 10s No NA No Not many Gautier (1976), Paulissen and Vermeersch (1987)*, and Wendorf and Schild (1976) 28 Esna E71K9 OA 1970s No No Yes LP 10– 21.59 Chron No M, F 100s No NA No Greenwood and Todd (1976), Vermeersch (2008)*, and Wendorf and Schild (1976) 29 OA 1962–1965 No No Yes LP 14.5– 16.5, 18–19 Chron No V 100s Yes Yes No Churcher and Smith (1972), Vermeersch and Van Neer (2015), and Yeshurun (2018)* ESA MSA LSA Age 1977–present Yes Yes Yes OA late No 1960s–1980s No 32.500 OA late No 1960s–1980s 26.100 32.500 OA Egypt 25.400 32.500 Gebel Silsila 2B, Egypt Kom Ombo 24.642 32.938 Dates (ka) Dating Bone Taph Butch Tools Comments References Churcher et al. (2008) and Kato et al. (2014) Source site Country Lat Long Year(s) Type excavated MSA LSA Age Dates (ka) 30 Kharga Oasis KO10C Egypt 24.700 30.708 OA 31 Nazlet Khater 4 Egypt 31.378 26.778 32 Sodmein Cave Egypt 26.240 33 Taramsa Egypt 34 Wadi Abu Noshra sites 35 Wadi Kubbaniya E-78-3 (all trenches) Hominin remains Vert Fauna Size 1930–1933, Yes 1976, 1982–1983, 1983, 2000–present No No MP Undated Type No V Unknown No NA No CatonThompson and Gardner (1932)*, CatonThompson (1952), Churcher et al. (1999), Dachy et al. (2018), Gardner (1932, 1935), and Leigh and Butzer (1968) OA 1980s No No Yes LP 33 Chron Yes V Unknown No NA No Crevecoeur (2012) and Vermeersch et al. (1984)* 33.970 C 1970s, No 1980s, 1990s, 2000s Yes Yes LPH 6 - >115 Chron No V Unknown No Yes No Moeyersons et al. (2002)* and Vermeersch et al. (2015) 26.120 32.685 OA 1990s No Yes No LP 18.48 Chron Yes V Unknown No NA No Van Peer et al. (2010) and Vermeersch et al. (1998)* Egypt 28.700 33.950 OA 1980s No Yes Yes LP 26–48 Chron No M, B 100s No NA Yes Phillips (1988*, 1994) Egypt 24.224 32.841 OA 1978–1980s No No Yes LP 17–18.5 Chron No V 10,000s No Yes No Gautier and Van Neer (1989), Vermeersch and Van Neer (2015)*, Wendorf and Said (1967), and Wendorf et al. (1988) ESA Dating Bone Taph Butch Tools Comments References (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 36 Wadi Kubbaniya E-78-4 (all trenches) Egypt 24.220 32.841 OA 1978–1980s No No Yes LP 17–18.5 Chron No V 10,000s No Yes No Gautier and Van Neer (1989), Vermeersch and Van Neer (2015)*, Wendorf and Said (1967), and Wendorf et al. (1988) 37 Wadi Kubbaniya E-78-5 Egypt 24.220 32.840 OA 1978–1980s No No Yes LP Undated Type No V 100s No Yes No Banks et al. (2015)* and Gautier and Van Neer (1989) 38 Wadi Kubbaniya E-78-5f Egypt 24.222 32.856 OA 1978–1980s No No Yes LP 15.83 Chron No V 100s No Yes No Gautier and Van Neer (1989), Pazdur et al. (1994), and Vermeersch and Van Neer (2015)* 39 Wadi Kubbaniya E-78-9 Egypt 24.222 32.841 OA 1978–1980s No No Yes LP 18.2 Chron No V 1000s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* 40 Wadi Kubbaniya E-81-1 (all trenches) Egypt 24.223 32.841 OA 1978–1980s No No Yes LP 17–18.5 Chron No V 10,000s No Yes No Gautier and Van Neer (1989), Wendorf and Said (1967), Wendorf et al. (1988), and Vermeersch and Van Neer (2015)* References Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age 41 Wadi Kubbaniya E-81-3 Egypt 24.223 32.852 OA 1978–1980s No No Yes LP 42 Wadi Kubbaniya E-81-4 Egypt 24.224 32.852 OA 1978–1980s No No Yes 43 Wadi Kubbaniya E-81-5 Egypt 24.228 32.845 OA 1978–1980s No No 44 Wadi Kubbaniya E-81-6 Egypt 24.224 32.930 OA 1978–1980s No 45 Wadi Kubbaniya E-81-7 Egypt 24.220 32.840 OA 1978–1980s 46 Wadi Kubbaniya E-82-3 Egypt 24.224 32.841 OA 47 Wadi Kubbaniya E-83-1 Egypt 24.220 32.840 48 Wadi Kubbaniya E-83-2 Egypt 24.220 32.840 Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 18.1– 18.4 Chron No V 1000s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* LP 18.4– 20.7 Chron No V 10,000s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* Yes LP 12.5 Chron No V 100s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* No Yes LP 17.4– 19.4 Chron No V 1000s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* No No Yes LP Undated Type No V 100s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* 1978–1980s No No Yes LP 19.8 No V 1000s No Yes No Gautier and Van Neer (1989) and Vermeersch and Van Neer (2015)* OA 1978–1980s No No Yes LP Undated Type No V 100s No Yes No Banks et al. (2015)* and Gautier and Van Neer (1989) OA 1978–1980s No No Yes LP Undated Type No V 100s No Yes No Banks et al. (2015)* and Gautier and Van Neer (1989) Chron References (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 49 Wadi Kubbaniya E-83-4 Egypt 24.220 32.840 OA 1978–1980s No No Yes LP 50 Wadi Kubbaniya E-84-1 Egypt 24.220 32.840 OA 1978–1980s No No Yes 51 Wadi Kubbaniya E-84-2 Egypt 24.220 32.840 OA 1978–1980s No No 52 Wadi Kubbaniya WK26 Egypt 24.220 32.840 OA 2014 No 53 Hagfet ed Dabba Libya 32.680 21.560 C 1947–1948, 2008 54 Haua Fteah 32.900 22.051 C 1950s; 2007–2012 Libya Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Undated Type No V 100s No Yes No Banks et al. (2015)* and Gautier and Van Neer (1989) LP Undated Type No V 100s No Yes No Banks et al. (2015)* and Gautier and Van Neer (1989) Yes LP Undated Type No V 100s No Yes No Banks et al. (2015)* and Gautier and Van Neer (1989) No Yes LP 12–19 Chron No V Unknown No NA No Banks et al. (2015)* and Gautier and Van Neer (1989) No No Yes LP 39–40 Chron No M Unknown No Yes No Barker et al. (2008), Bate (1955), Derricourt (1971), Klein and Scott (1986), McBurney and Hey (1955), and Reade et al. (2016)* No Yes Yes LPH 6.9–73 Chron Yes V 10,000s Yes No Douka et al. (2014)*, Klein and Scott (1986), McBurney (1967), and Tobias (1967) Dating Yes References Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size 55 Uan Tabu Libya 24.860 10.528 RS 1960–1963; 1990–1993 No Yes Yes 56 Aïn Bahya Morocco 33.700 −7.200 OA 1980s Yes Yes 57 Ain Maarouf Morocco 33.600 −5.300 OA 1950–1951, 1955, 1982 Yes 58 Bizmoune Morocco 31.670 −8.580 C 59 Bouknadel Morocco 34.107 60 Chaperon Rouge Morocco 61 Contrebandiers Cave 62 Dar es Soltan I Bone Taph Butch Tools Comments LPH 5–60 Chron No M Unknown Yes Yes No Garcea (2001)* No MP 400 BChron No V Unknown No NA No Cheddadi (1986) and Geraads (2012)* No No MP 700 BChron Yes M, B Unknown No NA No Geraads (2012) and Geraads et al. (1992)* 2007, No 2014–present Yes Yes LPH Undated Chron, No BChron, Type V Unknown Yes Yes No Bouzouggar et al. (2017) and Fernandez et al. (2015)* −6.751 C 1950s No Yes Yes LP Undated BChron No M, B 10,000s No Yes No Dellali (1968), Ennouchi (1953), Michel (1988), and Michel (1992)* 33.917 −6.833 OA 1970s No Yes Yes LP 28.2 Chron No U Unknown No NA No Nespoulet et al. (2008)* Morocco 33.922 −6.962 C 1955–1957; 1967–1975; 1997; 2003; 2007–2011 No Yes Yes LPH 10–120 Chron Yes V 10,000s Yes Yes Yes Dibble et al. (2012*, 2013), Hallett (2018), Jacobs et al. (2011), Steele and ÁlvarezFernández (2011), and Texier et al. (1988) Morocco 33.979 −6.898 C 1937–1938; 2005, 2008, 2012 No Yes Yes LPH 43 >120 Chron Yes U Unknown No NA Yes Barton et al. (2009) and Bouzouggar et al. (2018)* References (continued) Table 1 (continued) Year(s) Type excavated Source site Country Lat Long 63 Dar es Soltan II Morocco 33.979 −6.898 C 64 Doukkala I Morocco 33.890 65 Doukkala II Morocco 66 El Harhoura 1 67 El Harhoura 2 Hominin remains Vert Fauna Size Undated Type Yes U Unknown No NA No Debénath (1976)*, Ferembach (1976), and Nespoulet et al. (2008) LPH Undated BChron No V 10,000s No Yes No Michel (1992) * Yes LPH > 45 Chron No V 10,000s No Yes No Michel (1990, 1992)* Yes Yes LPH 32–66 Chron Yes M 10,000s Yes Yes No Aouraghe (2000), Aouraghe and Abbassi (2002), Aouraghe et al. (2012), Bailon and Aouraghe (2002), Debénath (1982), Debénath (1992), Jacobs et al. (2012)*, and Raynal and Occhietti (2012) Yes Yes LPH 10–120 Chron No V 10,000s Yes Yes Yes Ben Arous et al. (2019), Campmas (2012), Jacobs et al. (2012)*, Michel et al. (2009), Nouet et al. (2015), Stoetzel (2009), and Stoetzel et al. (2011, 2012) ESA MSA LSA Age 1969–1970s No Yes Yes LPH −7.000 C 1980s No Yes Yes 33.891 −7.001 C 1980s No Yes Morocco 33.954 −6.924 C 1977 No Morocco 33.954 −6.924 C 1978; 1996; No 2001–present Dates (ka) Dating Bone Taph Butch Tools Comments References Year(s) Type excavated Source site Country Lat Long 68 El Mnasra Morocco 33.928 −6.954 C 69 Fouarat Morocco 34.333 70 Grotte des Gazelles Morocco 71 Grotte des Pigeons (Taforalt) 72 73 Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments LPH 65–120 Chron Yes V 10,000s Yes Yes Yes Campmas (2012), El Hajraoui (1993), and Jacobs et al. (2012)* No PLEP Undated BChron No V Unknown No NA No Choubert et al. (1948) and Clark (1967)* Yes Yes LP 13.5– 23.5 Chron No V 1000s Yes No No Bougariane et al. (2010), Campmas and Daujeard (2020),* and Daujeard et al. (2011) No Yes Yes LPH 18–83 Chron Yes V 10,000s Yes Yes Yes Barton et al. (2016), Bouzouggar et al. (2007), Stephens et al. (2019), Taylor et al. (2011), and Turner et al. (2020)* No Yes Yes MP- 83–170 LP Chron No M Unknown No Yes No Hutterer (2010), Nami and Moser (2010), Richter et al. (2010), Tomasso and Rots (2018)* 1960s, 2000s No Yes No MP Chron Yes V 1000s Yes No Amani and Geraads (1993), Geraads et al. (2013)*, Hublin et al. (2017), and Richter et al. (2017) ESA MSA LSA Age 1990–2001; No 2004–present Yes Yes −6.417 OA 1940s Yes No 33.530 −7.830 C 2005 No Morocco 34.811 −2.408 C 1944–1947; 1950–1955; 1969–1977; 2003–2020 Ifri n’Ammar Morocco 34.785 −3.092 C 1997–2005 Jebel Irhoud Morocco 31.855 −8.873 C 315 Yes References (continued) Table 1 (continued) Year(s) Type excavated MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 1930s, 1940s No Yes Yes LP 35–60 Chron Yes M, B 10,000s Yes Yes No Hawkins (2001) and Wrinn and Rink (2003)* −7.733 C 1980s–1990s No Yes No LP 50 BChron No V 1000s Yes No No Daujeard et al. (2011), Geraads (2012), Raynal et al. (2008), and Raynal and Occhietti (2012)* 33.600 −7.600 C 1996; 2005–2009 Yes No No MP 520–720 Chron Yes V 10,000s Yes Yes No Daujeard et al. (2020)* Morocco 34.558 −1.874 C 1979–1986; 1995–1998; 2007–2010 No Yes Yes LPH 7.8–135 Chron No M, B 100s Yes Yes No Doerschner et al. (2016)* and Michel (1992) Salé Morocco 34.000 −6.750 OA 1970s No Yes No MP 300 BChron Yes V Unknown No NA No Geraads (2012) and Clark (1967)* 79 Sidi Abderrahmane Grotte des Ours Morocco 33.581 −7.683 C 1992–present Yes No No EPMP Undated BChron, No Type V Unknown No NA No Raynal et al. (2001) and Raynal et al. (2010)* 80 Sidi Abderrahmane Les Littorines Morocco 33.571 −7.702 C 1950s No No EPMP Undated BChron V Unknown No NA No Arambourg and Biberson (1956), Debénath et al. (1982), Sbihi Alaoui et al. (2007), and Raynal et al. (2016)* Source site Country Lat Long 74 Mugharet el ’Aliya Morocco 35.760 −5.939 C 75 Oulad Hamida 2 La Grotte des Felins Morocco 33.533 76 Ouled Hamida 1 Grotte de Rhinoceros Morocco 77 Rhafas 78 ESA Yes Yes References Year(s) Type excavated MSA LSA Age Dates (ka) 1978–present Yes No No EP 1000 −7.680 C 1978–present Yes No No 35.900 −5.400 RS 2000s– present No Yes Tunisia 36.250 8.750 OA 1942. 1946–1948 Yes Adrar Bous Niger 20.300 8.950 OA 86 Gobero Niger 17.083 9.517 87 Ihò Eleru (Iwo Eleru) Nigeria 7.441 5.125 Source site Country Lat Long 81 Thomas Quarry 1 Niveau L Morocco 33.580 −7.680 C 82 Thomas Quarry I Morocco Hominid Cave 33.580 83 Benzú Spain 84 Sidi Zin 85 ESA Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments References Chron, Yes BChron, PMag, Type M Unknown Yes Yes No Ben Arous et al. (2019), Daujeard et al. (2016), Daujeard et al. (2020)*, Raynal et al. (2002), and Rhodes et al. (2006) MP- 700 LP Chron, Yes BChron, PMag, Type M Unknown Yes Yes No Ben Arous et al. (2019), Daujeard et al. (2016), Daujeard et al. (2020)*, Raynal et al. (2010), and Raynal et al. (2002) No MP- 70–250 LP Chron No V Unknown Yes Yes No Ramos-Muñoz et al. (2016) and Ramos et al. (2008)* Yes No MP Undated Type No V Unknown No NA No Belhouchet (2002), Clark (1967)*, and Vaufrey (1950) 1960s, 1970s No Yes No LPH Undated Type No V 10s Yes Yes No Clark et al. (2008)*, and GiffordGonzales and Parham (2008) OA 2000, 2003, 2005–2006 No No Yes LPH 8.2–16.5 Chron Yes V Unknown No NA Yes Human remains are Holocene Sereno et al. (2008)* RS 1964–1965; 2019 No No Yes LPH 3–16 Yes V 10s NA No poor preservation and few in number Allsworth-Jones et al. (2010)*, Harvati et al. (2011), and Shaw and Daniels (1984) Chron No (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Hominin remains Vert Fauna Size 88 Abu Hugar Sudan 12.704 34.132 OA 1924 No Yes No MP- 133 LP Chron, Type No M Unknown No No No 89 Affad 105 Sudan 18.042 31.202 OA 2012–2014 No Yes No LP 16 Chron No M 10s Yes No No Osypińska and Osypiński (2016)* 90 Affad 110 Sudan 18.036 31.169 OA 2012–2014 No Yes No LP 15–21 Chron No M 1000s Yes No No Osypińska and Osypiński (2016)* 91 Affad 111 Sudan 18.045 31.203 OA 2012–2014 No Yes No LP 16 Chron No M 100s Yes No No Osypińska and Osypiński (2016)* 92 Affad 23 Sudan 18.030 31.185 OA 2012–2014 No Yes No LP 15–21 Chron No M 1000s Yes Yes No Osypińska and Osypiński (2016), and Osypinski et al. (2016)* 93 Atbara river at site 047 Sudan 15.085 35.955 OA 2005–2006 Yes No No EPMP > 128 Chron, PMag No M, R 10s No No No MSA seems to be Abbate et al. in upper layers, (2010)* but no excavations there 94 Kashm el-Girba KG 15 Sudan 15.003 35.954 OA 1981–1982 No No Yes LP 11–12 Chron No M Unknown No No No Taxonomic presence/absence only noted Marks (1987)*, and Peters (1989) 95 Kashm el-Girba KG 16 Sudan 15.003 35.954 OA 1981–1982 No No Yes LP 11–12 Chron No M Unknown No No No Taxonomic presence/absence only noted Marks (1987)*, and Peters (1989) 96 Kashm el-Girba KG 73 Sudan 15.003 35.955 OA 1981–1982 No No Yes LP 11–12 Chron No M Unknown No No No Taxonomic presence/absence only noted Marks (1987)* and Peters (1989) Dating Bone Taph Butch Tools Comments Clark information based on unpublished work by H. Zeigert References Bate (1951), Clark (1988), McDermott et al. (1996)* Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 97 Kashm el-Girba KG 74 Sudan 15.002 35.954 OA 1981–1982 No No Yes LP 11–12 Chron No M Unknown No No No Taxonomic presence/absence only noted 98 Opposite Wadi Halfa 6B27 Sudan 21.840 31.253 OA 1962–1965 No Yes No LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 99 Opposite Wadi Halfa 6B28 Sudan 21.840 31.253 OA 1962–1965 No No Yes LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 100 Opposite Wadi Halfa 6B29 Sudan 21.840 31.253 OA 1962–1965 No No Yes LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 101 Opposite Wadi Halfa 6B31 Sudan 21.840 31.253 OA 1962–1965 No Yes? Yes? LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 102 Opposite Wadi Halfa 6B35+6G26 Sudan 21.840 31.253 OA 1962–1965 No No Yes LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun (2018)* marks not specified 103 Opposite Wadi Halfa 6B36 Sudan 21.840 31.253 OA 1962–1965 No No Yes LP Undated Type Yes M, B, F 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 104 Opposite Wadi Halfa 6G18 Sudan 21.840 31.253 OA 1962–1965 No Yes No LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 105 Opposite Wadi Halfa 6G27 Sudan 21.840 31.253 OA 1962–1965 No Yes No LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 106 Opposite Wadi Halfa 6G30 Sudan 21.840 31.253 OA 1962–1965 No Yes No LP Undated Type No M 10s Yes Yes No Site with butchery Yeshurun marks not (2018)* specified 107 Wadi Umm Rahau, HP766 Sudan 18.830 32.030 OA 2006–2007 No Yes Yes LP 18–45 Chron No M 100s No No No Gautier et al. (2012)* 108 Ishango 11 DRC −0.014 29.603 OA 1935–1936; 1950; 1986 No No Yes LPH 2–25 Chron Yes M, B, R, F 1000s No No Yes Brooks et al. (1995)* and Peters (1990) References Marks (1987)* and Peters (1989) (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age 109 Ishango 14 DRC −0.138 29.588 OA 1986 No No Yes 110 Kakontwe DRC −10.983 26.700 C 1957 No Yes 111 Katanda 16 DRC −0.092 29.599 OA 1982–1990 No 112 Katanda 2 DRC −0.096 29.597 OA 1986–1990 113 Katanda 9 DRC −0.093 29.599 OA 114 Matupi DRC 1.231 29.823 115 Senga 5A DRC −0.049 116 Barogali Djibouti 11.171 Source site Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments LPH 16.5–22 Chron No M, B, R, F Unknown No No Yes Yes? LPH Undated BChron Yes M Unknown No No No Yes Yes LP 82–173 Chron, No BChron, Type M, F 1000s No No Yes Brooks et al. (1995)* and Yellen et al. (1995) Yes Yes Yes MP- 82–590 LP Chron, No BChron, Type M, F 100s No No Yes Brooks et al. (1995)* and Yellen et al. (1995) 1982–1990 No Yes Yes LP 82–173 Chron, No BChron, Type M, F 1000s Yes No Yes C 1973–1974 No Yes? Yes LPH 3 - > 40 Chron Yes M, B, R, F 1000s Yes Yes No 29.617 OA 1985–1986 Yes No No EP 2000– 2300 BChron No M, R, F 1000s Yes Yes No Cut marks controversial Harris et al. (1987)*, Stewart (2004), and Tappen and Harris (1995) 41.931 OA 1985–1987 Yes No No EP 1300– 1600 Chron No M 100s Yes Yes? No Single elephant associated with many stone tools Berthelet and Chavaillon (2001)* References Brooks et al. (1995)* and Peters (1990) Collection year not provided, but likely close to 1957; not clear if excavated Taphonomy does not include bone surface modification Oakley et al. (1977)* and Van Neer (1990) Brooks et al. (1995)*, Yellen (1996), and Yellen et al. (1995) Van Neer (1981), Van Neer (1984)* Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating 117 Aduma A1 Ethiopia 10.389 40.509 OA 1976 (survey); 1992 No Yes No MP- > 100 LP 118 Aduma A2 VP 1/1 Ethiopia 10.436 40.529 OA 1976 (survey); 1992 No Yes No 119 Aduma A4 Ethiopia 10.438 40.532 OA 1976 (survey); 1992 No Yes 120 Aduma A8 Ethiopia 10.396 40.538 OA 1976 (survey); 1992 No 121 Aduma A8A Ethiopia 10.396 40.538 OA 1976 (survey); 1992 122 Bodo Ethiopia 10.711 40.530 OA 123 Bouri Daka Member Ethiopia 10.290 40.526 124 Bouri Formation, Ethiopia Hata Member 10.260 125 Bouri Herto BOU-A19B 10.260 Ethiopia Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Chron, Type No M, R, F 10s No No No Erosion of bone surfaces and adhering matrix Yellen et al. (2005)* MP- 80–100 LP Chron, Type Yes M, B, R, F 10s No No No Erosion of bone surfaces and adhering matrix Yellen et al. (2005)* No MP- 80–100 LP Chron, Type No M, R, F 10s No No No Erosion of bone surfaces and adhering matrix Yellen et al. (2005)* Yes No MP- 80–100 LP Chron, Type No M, R, F 10s No No No Erosion of bone surfaces and adhering matrix Yellen et al. (2005)* No Yes No MP- 80–100 LP Chron, Type No M, R, F 10s Yes Yes No “Unquestionable” Yellen et al. cutmarks on (2005)* hippopotamus and crocodile 1976–1978 Yes No No MP 600 Chron, BChron Yes M, R, F 10s No No No Only taxonomically informative specimens reported Conroy et al. (1978)*, Kalb et al. (1980), and Millard (2008) OA 1997 Yes No No MP 800– 1000 Chron, Yes BChron, PMag, Type M 100s No Yes No Only taxonomically informative specimens reported Asfaw et al. (2002)* and Vrba (1997) 40.568 OA 1996–1998 Yes No No EP 2500 Chron, PMag Yes M, B, R, F 100s No Yes No 40.557 OA 1997 No Yes No MP 154–160 Chron No M 100s No Yes No References de Heinzelin et al. (1999)* and Sahle et al. (2017) Hippo with cutmarks on cranium Clark et al. (2003)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 126 Bulbula 1 s1 (Aneno Melka Gademotta) Ethiopia 7.819 38.698 OA 2008–2010 No No Yes LPH 11.5– 13.5 127 Bulbula 1 s3 Ethiopia 7.819 38.698 OA 2009–2010 No Yes No LP 128 Bulbula 1 s4 Ethiopia 7.819 38.698 OA 2009–2010 No No Yes 129 Bulbula 4 Ethiopia 7.816 38.694 OA 2008–2009 No Yes 130 Deka Wede 1 Ethiopia 7.793 38.691 OA 1979; 2010 No 131 Fejej, FJ-1 Ethiopia 4.721 36.466 OA 1989, 1992–1993, 1997–1999 132 Gadeb 2A Ethiopia 7.135 39.330 OA 133 Gadeb 2B Ethiopia 7.135 39.330 134 Gadeb 2C Ethiopia 7.135 135 Gadeb 2E Ethiopia 136 Gadeb 8D Ethiopia Source site Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Chron No M, B 10,000s No Yes No Bon et al. (2013)* and Lesur et al. (2016) Undated Type No M 1s No No No Bon et al. (2013)* and Lesur et al. (2016) LPH 13–15 No V 100s No No No Bon et al. (2013)* and Lesur et al. (2016) No LP Undated Type No V 10s No No No Yes No LP 30 Chron No V Unknown No No No Yes No No EP 1896– 1950 Chron, PMag No M, R, F 1000s No No No Human-induced fracturing, “bone splinters” Asfaw et al. (1991)*, Barsky et al. (2011), and de Lumley and Yonas (2004) 1975, 1977 Yes No No EP 700– 1450 Chron, BChron No M, B 10s No No No Fragmentary de la Torre (2011)* OA 1975, 1977 Yes No No EP 700– 1450 Chron, BChron No M, B 10s No No No Fragmentary de la Torre (2011)* 39.330 OA 1975, 1977 Yes No No EP 700– 1450 Chron, BChron No M, B 10s No No No Fragmentary de la Torre (2011)* 7.135 39.330 OA 1975, 1977 Yes No No EP 700– 1450 Chron, BChron No M, B 10s No No No Fragmentary de la Torre (2011)* 7.135 39.330 OA 1975, 1977 Yes No No EP 700– 1450 Chron, BChron No M, B 10s No No No Fragmentary de la Torre (2011)* Chron Bone fragments rare References Bon et al. (2013)* Bon et al. (2013)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age 137 Gadeb 8E Ethiopia 7.135 39.330 OA 1977 Yes No No 138 Gadeb 8F Ethiopia 7.135 39.330 OA 1975, 1977 Yes No 139 Goda Buticha (Serkema) Ethiopia 9.542 41.629 C 2008, 2011 No 140 Gona, Busidima North (BSN 12) Ethiopia 11.074 40.466 OA 1999–2001 141 Gona, Dana Aoule North (DAN5) Ethiopia 11.087 40.526 OA 142 Gona, Dana Aoule North 1 (DAN1) Ethiopia 11.085 40.536 143 Gona, Dana Ethiopia Aoule North 2 (DAN2 + DAN 2 South) 11.085 144 Gona, East Gona Ethiopia 13 (EG13) 11.119 Source site Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments EP 700– 1450 Chron, BChron No M, B 10s No No No Small quantity de la Torre (2011)* and Clark and Kurashina (1979) No EP 700– 1450 Chron, BChron No M, B 100s No No No Alleged hippo butchery de la Torre (2011)* and Clark and Kurashina (1979) Yes Yes LPH 5 - > 30 Chron Yes M 100s Yes Yes No Taphonomy of micromammals only Assefa et al. (2014)* and Stoetzel et al. (2018) Yes No No EP 1260 Chron, PMag Yes M, B, R, F 100s No No No Semaw et al. (2020)* 2000–2017 Yes No No EP 1500– 1600 Chron, PMag Yes M, B, R, F 1000s Yes Yes No Cáceres et al. (2015b), and Semaw et al. (2020)* OA 1999–2003 Yes No No EP 2500– 2600 Chron No U 10s No No No DomínguezRodrigo et al. (2005)* 40.535 OA 1999–2003 Yes No No EP 2000– 2400 Chron No U 10s No Yes No Cáceres et al. (2015b), and DomínguezRodrigo et al. (2005)* 40.533 OA 1999–2003 Yes No No EP 2500– 2600 Chron No M 10s No Yes No DomínguezRodrigo et al. (2005)* and Quade et al. (2008) References (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age 145 Gona, Ounda Gona South 12 (OGS12) Ethiopia 11.102 40.522 OA 2003, 2007–2008 Yes No No EP 146 Gona, Ounda Gona South 6 (OGS6) Ethiopia 11.102 40.530 OA 2000–2003 Yes No No 147 Gona, Ounda Gona South 7 (OGS7) Ethiopia 11.100 40.530 OA 2000, 2010 Yes No 148 Gona, West Gona 9 (WG9) Ethiopia 11.120 40.529 OA 1999–2003 Yes 149 Gotera Ethiopia 5.057 36.837 OA 1974 150 Hadar, A.L. 666 Ethiopia 11.157 40.610 OA 1994, 1999 Source site Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 1500– 1600 Chron No M, R, F 100s No Yes No EP 2500– 2600 Chron, PMag No M 10s No Yes No No EP 2500– 2600 Chron, PMag No M 10s No No No No No EP 2000– 2400 Chron No M 10s No Yes No DomínguezRodrigo et al. (2005)* No Yes No LP Undated Type No M, B, R 10s No No No Chavaillon et al. (1985)* and Geraads and Guillemot (1985) Yes No No EP 2350– 2360 Yes M 10s No No No Chron Site mentioned in 2004 paper, exact decade not specified References Cáceres et al. (2015b), Quade et al. (2008), and Semaw et al. (2018)* Cáceres et al. (2015b), DomínguezRodrigo et al. (2005)*, Quade et al. (2008), Semaw et al. (2003) Fragmented and poorly preserved surfaces Highly fragmented but some good surface preservation Cáceres et al. (2015b), DomínguezRodrigo et al. (2005)*, Quade et al. (2008), Semaw et al. (2003), and Stewart and Murray (2020) Campisano (2012)* and Kimbel et al. (1996) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age 151 Hadar, A.L. 894 Ethiopia 11.157 40.610 OA 2000–2002 Yes No No EP 152 Halibee Ethiopia 10.780 40.390 OA 1975–1978 No Yes No 153 Konso, KGA4-A3 Ethiopia 5.428 37.418 OA 1997 Yes No 154 Konso, KGA6-A1 Ethiopia 5.406 37.429 OA 1996–1997, 2002–2003, 2013 Yes 155 Ledi-Geraru, Bokol Dora 1 (BD1) Ethiopia 11.399 40.868 OA 2013, 2015 156 Melka Kunture Garba III Ethiopia 8.699 38.591 OA 1970–1978; 2011 Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 2350– 2360 Chron No M 100s Yes No No Campisano (2012)*, DomínguezRodrigo and MartinezNavarro (2012), Reed and Geraads (2012) LP 54–106 Chron Yes U Unknown No No No Negash et al. (2011)* No EP 1450 Chron Yes M 10,000s Yes Yes No Beyene et al. (2015), Beyene et al. (2013), Echassoux (2012), Katoh et al. (2000)*, Suwa et al. (2020), and Suwa et al. (2003) No No EP 1750 Chron No M 100s Yes Yes No Beyene et al. (2013, 2015), Echassoux (2012), Katoh et al. (2000)*, Suwa et al. (2020), and Suwa et al. (2003) Yes No No EP 2580– 2610 Chron, PMag No M, B, R 100s Yes No No Braun et al. (2019)* No Yes No MP < 700 Chron, Type Yes M 1000s No No No Di Vincenzo et al. (2015)* and Mussi et al. (2014) References (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 157 Melka Kunture Garba XIII Ethiopia 8.704 38.596 OA 2007 Yes No No EPMP 800– 1000 158 Melka Kunture Gombore I Ethiopia 8.705 38.601 OA 1965–1967 Yes No No EP 159 Melka Kunture Gombore I-gamma Ethiopia 8.705 38.600 OA 1970–1995; 2011 Yes No No 160 Melka Kunture Gombore II Ethiopia 8.704 38.600 OA 1970–1995; 2011 Yes No 161 Melka Kunture Karre I Ethiopia 8.707 38.604 OA 1980–1981 Yes 162 Melka Kunture Simbiro III Ethiopia 8.710 38.567 OA 1973–1976 163 Melka Kunture, Garba IV Ethiopia 8.706 38.599 OA 1972–1982, 2005 Source site Vert Fauna Size Bone Taph Butch Tools Comments Chron, No BChron, PMag M 10s No No No Most fauna are hippos Gallotti et al. (2014)* 1600– 1900 Chron, Yes BChron, PMag M, R 1000s No No No Sources vary on if the site was discovered in 1965 or 1967; “Fragmentary” Di Vincenzo et al. (2015)* and Piperno et al. (2009) EP 1300 Chron, No BChron, PMag M 100s No No No Bones are “numerous and varied” Chavaillon and Berthelet (2004) and Gallotti et al. (2014)* No MP 700–850 Chron, Yes BChron, PMag M 100s Yes Yes No Hippo butchery site with footprints on top Altamura et al. (2018), Gallotti et al. (2014)*, and Mussi et al. (2016) No No EPMP Undated Type No M 100s No No No Berthelet and Chavaillon (2004) and Di Vincenzo et al. (2015)* Yes No No EPMP > 878 Chron No M 100s No No No Beaudet et al. (2015)* and Chavaillon and Berthelet (2004) Yes No No EP 1719 Chron, PMag Yes M, R 1000s Yes Yes No Dating Hominin remains Described as intentionally fractured References Di Vincenzo et al. (2015)*, Fiore and Tagliacozzo (2004), Gallotti (2013), Morgan et al. (2012), and Piperno et al. (2009) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 164 Melka Wakena MW2 Ethiopia 7.081 39.263 OA 2015–2017 Yes No No EP 1340– 1620 Chron No V 10s Yes Yes? No Poor preservation, Hovers et al. but hippo tibia (2021)* may be percussed 165 Melka Wakena MW5 Ethiopia 7.088 39.270 OA 2015–2017 Yes No No EP 1340– 1620 Chron No V 10s Yes No No Poor preservation Hovers et al. (2021)* 166 Mochena Borago Ethiopia 6.904 37.759 C 1998; 2000–2001; 2006–2007 No Yes Yes LP 50–70 Chron No M 1000s No No No Poor bone preservation Brandt et al. (2012)* 167 Omo Kibish BNS Ethiopia 5.408 35.900 OA 2002–2003 No Yes No MP > 104 Chron No U Unknown No No No Faunal remains are “sparse” Sisk and Shea (2008)* 168 Omo Kibish KHS Ethiopia 5.403 35.930 OA 2002–2003 No Yes No MP 195 Chron No M, B 1s No No No Sisk and Shea (2008)* 169 Porc Epic Ethiopia 9.573 41.886 C 1929; 1933, 1974; 1975–1976; 1998 No Yes Yes LPH > 43 Chron Yes M 10,000s Yes Yes No Assefa (2006)* 170 Shinfa-Metema 1 Ethiopia 12.625 36.005 OA 2002–2018 No Yes No LP 31–46 Chron, Type No M, B, R, F, A 1000s Yes Yes No Fauna described in detail in a dissertation only Kappelman et al. (2014)* and Davis (2019) 171 Shungura Formation Member F, Omo A42/OMO 79 Ethiopia 5.061 36.020 OA 2008 Yes No No EP 2270– 2320 Chron No U 1000s Yes No No Possible hominin or crocodile bone surface modifications Maurin et al. (2017)* 172 Yavello Shelter 1 Ethiopia 4.905 38.087 RS 1942 No Yes Yes LP Undated Type No M 10s No No No Clark (1945)* 173 Guli Waabayo Somalia 2.993 44.304 RS 1944 No No Yes LP 6–14 Chron No M 10,000s No No No Reid et al. (2019)* 174 Rifle Range (Moga Marinta) Somalia 2.805 44.078 OA 1945, 1989 No Yes Yes LPH 4.5 > 40 Chron, Type No M, B, R, F, A 1000s No No No Jones et al. (2018)* 175 Gud-Gud Somaliland 10.680 48.045 C 1980, 1982 No Yes Yes LPH > 40 Chron, Type No U Unknown No No No 176 Laas Geel Shelter 7 Somaliland 9.700 44.420 RS 2003–2004 No Yes Yes LPH 1–42 Chron No M, B 100s No No Source site Dates (ka) No “Faunal remains” noted as present References Brandt and Brook (1984)* Gutherz et al. (2014)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age 177 Midhishi 2 Somaliland 10.611 47.900 C 1980, 1982 No Yes Yes LPH > 40 178 Dilit 01 (GcJh6) Kenya 3.470 35.776 OA 2009–2010, 2013 No No Yes LP 179 Enkapune Ya Muto (Twilight Cave GtJi12) Kenya −0.833 36.150 RS 1982, 1987 No Yes Yes LPH 180 FwJi 2 East Turkana Kenya 4.291 36.254 OA 1991 No Yes No 181 FxJj 61 East Turkana Kenya 4.043 36.538 OA 1992 No Yes 182 GaJj 17 East Turkana Kenya 3.985 36.427 OA 1991 No 183 Kanjera South (Exc 1; KS-1, KS-2, KS-3) Kenya −0.340 34.538 OA 1995–1997, 2000–2001 184 Kapthurin GnJh21 HMS Kenya 0.570 35.979 OA 185 Kapthurin GnJh23 LS Kenya 0.564 35.980 186 Kapthurin Hominid Site EKG (GnJh19) Kenya 0.572 187 Kilombe Mountain GqJh13A Kenya −0.075 Source site Dates (ka) Hominin remains Vert Fauna Size Chron, Type No V Unknown No No No 6–36 Chron No M, B, F 1000s Yes Yes Yes 0 - > 40 Chron Yes M 1000s Yes Yes No MP- Undated Type LP No M, F Unknown No No No Kelly and Harris (1992)* No MP- Undated Type LP No M, R Unknown No No No Kelly and Harris (1992)* Yes No MP- Undated Type LP No M, R, F Unknown No No No Kelly and Harris (1992)* Yes No No EP 2000 M 1000s Yes Yes No Bishop et al. (2006), Ferraro et al. (2013)*, Plummer et al. (1999), and Plummer et al. (2009) 1983; 1984; 1986 Yes No No MP 509–538 Chron, PMag No M 10s No No No Cornelissen et al. (1990)* OA 1984–1985 Yes No No MP 509–538 Chron, PMag Yes M, R, F 10s No No No Cornelissen et al. (1990)* 35.975 OA 1966 Yes No No MP 509–538 Chron, PMag Yes M 10s No No No Few bones Cornelissen et al. (1990)*, Leakey et al. (1970) 35.845 OA 2011–2017 Yes No No EP 1814 No V Unknown No No No “Multiple” bones Hoare et al. (2021)* Dating BChron, No PMag Chron, BChron Bone Taph Butch Tools Comments References “Faunal remains|” Brandt and noted as present Brook (1984)* Prendergast and Beyin (2018)* Macro and micromammals reported Marean (1992a)* and Marean et al. (1994) Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 188 Kilombe Volcano Kenya GqJh12A −0.073 35.845 OA 2011–2017 Yes No No EP 1814 189 Kilombe Volcano Kenya GqJh12C −0.074 35.845 OA 2011–2017 Yes No No EP 190 Kokito 01 (West Turkana, GcJh11) Kenya 3.431 35.757 OA mid-2010s No No Yes 191 Kokito 02 (West Turkana, GcJh12) Kenya 3.436 35.758 OA mid-2010s No No 192 Koobi Fora, FwJj1 Kenya 4.338 36.347 OA 1972–1973 Yes 193 Koobi Fora, FwJj14N Kenya 4.307 36.353 OA 1997–2004 194 Koobi Fora, FwJj14S Kenya 4.307 36.353 OA 195 Koobi Fora, FwJj20 Kenya 4.364 36.426 196 Koobi Fora, FxJj1 Kenya 4.090 36.461 Source site Country Hominin remains Vert Fauna Size Chron, BChron No V Unknown No No No Hoare et al. (2021)* 1814 Chron, BChron No V Unknown No No No Hoare et al. (2021)* LPH 10–18 Chron No M, F 1000s Yes Yes Yes Prendergast and Beyin (2018)* Yes LPH 14 Chron No M, B, F 1000s Yes Yes Yes Prendergast and Beyin 2018* No No EP 1520 Chron No M, R, F Unknown No No No Yes No No EP 1520 Chron No M, R, F 1000s Yes Yes No Braun et al. (2010)*, Brown et al. (2006), and Pobiner et al. (2008) 1997–2004 Yes No No EP 1520 Chron No M, R 1000s Yes Yes No Braun et al. (2010)*, Brown et al. (2006), and Pobiner et al. (2008) OA 2005 Yes No No EP 1950 Chron, PMag No M, B, R, F 1000s Yes Yes No Bamford (2011) and Braun et al. (2010)* OA 1969–1972 Yes No No EP 1869 Chron No M, R, F 100s No No No Dating Bone Taph Butch Tools Comments Highly fragmented bone but identifiable teeth Fauna very fragmentary References Harris and Isaac (1997)*, Isaac and Behrensmeyer (1997), and Mana et al. (2019) Isaac and Harris (1997)* and McDougall and Brown (2006) (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Kenya 4.172 36.384 OA 1972–1973 Yes No No EP 1520 Chron No M, R 100s No No No Bunn (1997), Harris and Isaac (1997)*, and Mana et al. (2019) 198 Koobi Fora, Kenya FxJj18IH (Ingrid Herbich Site) 4.161 36.387 OA 1973 Yes No No EP 1520 Chron No M 100s No No No Bunn (1997), Harris and Isaac (1997)*, and Mana et al. (2019) 199 Koobi Fora, FxJj18NS Kenya 4.161 36.387 OA 1972 Yes No No EP 1520 Chron No M 100s No No No Only small fragments Harris and Isaac (1997)* and Mana et al. (2019) 200 Koobi Fora, FxJj20AB Kenya 4.096 36.416 OA 1973–1974, 2010–2015 Yes No No EP 1500 Chron No M, B, R 1000s No No No Very fragmentary Bunn (1997), Harris and Isaac (1997)*, and Hlubik et al. (2017) 201 Koobi Fora, FxJj20E Kenya 4.096 36.416 OA 1972–1974, 1978–1979 Yes No No EP 1480 Chron Yes M, R, F 1000s No No No Highly fragmented fauna with poor surface preservation Bunn (1997), Harris and Isaac (1997)*, and McDougall and Brown (2006) 202 Koobi Fora, FxJj20M Kenya 4.096 36.416 OA 1972–1974, 1978–1979 Yes No No EP 1480 Chron No M, R, F 1000s Yes Yes No Crumbly bones, not wellpreserved surfaces, “chalky” Bunn (1997), Harris and Isaac (1997)*, and McDougall and Brown (2006) 203 Koobi Fora, FxJj20S Kenya 4.096 36.416 OA 1978 Yes No No EP 1480 Chron No M, F 100s No No No Fragmented, and with poor surface preservation Bunn (1997), Harris and Isaac (1997)*, McDougall and Brown (2006) Source site 197 Koobi Fora, FxJj17 References Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 204 Koobi Fora, FxJj3 (HAS/ Hippo and Artifact Site) Kenya 4.094 36.388 OA 1970–1972, 1979 Yes No No EP 1869 Chron No M, R, F 100s No No No 205 Koobi Fora, FxJj38E Kenya 4.143 36.363 OA 1973–1974 Yes No No EP 1760 Chron Yes M 10s No No No 206 Koobi Fora, FxJj38SE Kenya 4.143 36.363 OA 1973–1974 Yes No No EP 1760 Chron Yes M 100s No No No Encrusted in carbonates and nodules Bunn (1997), Harris and Isaac (1997)*, and Mana et al. (2019) 207 Koobi Fora, FxJj50 Kenya 4.141 36.406 OA 1977–1979 Yes No No EP 1480 Chron No M, B, R, F 1000s Yes Yes No Some bone surfaces deteriorating Bunn et al. (1980)*, Bunn (1997), Harris and Isaac (1997), and McDougall and Brown (2006) 208 Koobi Fora, FxJj64 Kenya 4.086 36.345 OA 1979 Yes No No EP 1480 Chron No M, F 100s Yes Yes No Some bone surfaces “leached and pitted” Bunn (1997), Harris and Isaac (1997)*, and McDougall and Brown (2006) 209 Koobi Fora, GaJi14 Kenya 3.943 36.291 OA 1998–2004 Yes No No EP 1500 Chron No M, R, F 1000s Yes Yes No 210 Lainyamok Kenya −1.782 36.187 OA 1976; 1984 Yes Yes No MP 330–392 BChron Yes M 1000s Yes No No 211 Lenderut (HaJjl) Kenya −2.120 36.240 OA 1987 Yes No No MP 500–800 BChron, No Type M Unknown No No No Source site Poorly preserved bone References Isaac and Harris (1997)* and McDougall and Brown (2006) Bunn (1997), Harris and Isaac (1997)*, and Mana et al. (2019) Brown et al. (2006), Bunn et al. (1980)*, and Pobiner et al. (2008) Inferred to not be homininaccumulated Faith et al. (2012) and Potts et al. (1988b)* Barthelme (1991)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size 212 Lukenya Hill GvJm16 Kenya −1.469 37.076 RS 1971 No Yes Yes LP 2 - >18 Chron Yes V Unknown No No No 213 Lukenya Hill GvJm19 Kenya −1.472 37.077 RS 1970s No No Yes LP 4–13 Chron No M 100s Yes Yes No 214 Lukenya Hill GvJm22 Kenya −1.474 37.076 RS 1970–1973 No Yes Yes LPH 1 - >46 Chron No M 100s Yes Yes No Marean (1992b) and Kusimba (2001)* 215 Lukenya Hill GvJm46 Kenya −1.477 37.073 OA 1977–1978 No No Yes LP 19–21 Chron No M 100s Yes Yes No Marean (1992b) and Kusimba (2001)* 216 Lukenya Hill GvJm62 Kenya −1.476 37.074 RS 1979–1981; 1994 No No Yes LPH 4–21 Chron No M 100s Yes Yes No Kusimba provides Marean (1992b) two dates, 1978 and Kusimba and 1979, as the (2001)* start of excavations 217 Marmonet Drift (GtJi15) Kenya −0.755 36.175 OA 2001–2010; 2013 No Yes Yes MP- 35–225 H Chron, Type No V Unknown No No No Only mention of bone is in a dissertation Ambrose (2002) and Slater (2016)* 218 Ntuka River 3 Ntumot (GvJh11) Kenya −1.344 35.913 OA Unknown No Yes Yes LPH Undated Chron, Type No V Unknown No No No Only mention of bone is in a dissertation Ambrose (2002), Moutsiou (2011)*, and Slater (2016) 219 Ntuka River 4 Norikiushin (GvJh12) Kenya −1.325 35.942 OA Unknown No Yes No LP Undated Type No U Unknown No No No Mentioned in DomínguezRodrigo et al. 2007d as Middle Pleistocene but with GuJh12 designation DomínguezRodrigo et al. (2007c) and Slater (2016)* 220 Olorgesailie B11-500 Kenya −1.600 36.446 OA 2010 Yes No No MP 499–615 Chron, No BChron, Type U 10s No No Bone surfaces poorly preserved Brooks et al. (2018) and Potts et al. (2018)* Source site Bone Taph Butch Tools Comments No References Kusimba (2001)* Reported fauna is taxonomically identifiable Marean (1992b) and Kusimba (2001)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age 221 Olorgesailie BOK-1E Kenya −1.588 36.439 OA 2007–2009 No Yes No MP 222 Olorgesailie BOK-2 Kenya −1.588 36.440 OA 2008–2011 No Yes No 223 Olorgesailie BOK-3 Kenya −1.588 36.438 OA 2005–2007 No Yes 224 Olorgesailie BOK-4 Kenya −1.588 36.439 OA 2006–2010 No 225 Olorgesailie D14-10 Kenya −1.570 36.426 OA 2007 226 Olorgesailie GOK-1 Kenya −1.597 36.414 OA 227 Panga ya Saidi Kenya −3.208 40.002 228 Prolonged Drift (GrJi11) Kenya −0.485 229 Rusinga Island Nyamita Main Kenya −0.424 Source site Dates (ka) Vert Fauna Size Bone Taph Butch Tools Comments 295–320 Chron, No BChron, Type M, B, A 100s No Yes No Brooks et al. (2018)* and Potts et al. (2018) MP 295–320 Chron, No BChron, Type M, B, R, F, A 10,000s No Yes No Brooks et al. (2018)* and Potts et al. (2018) No MP 295–320 Chron, No BChron, Type M 10s No No No Brooks et al. (2018)* and Potts et al. (2018) Yes No MP 295–320 Chron, No BChron, Type M 10s No No No Brooks et al. (2018)* and Potts et al. (2018) Yes No No MP 499–615 Chron, No BChron, Type U 1s No No No 2002, 2004–2007 No Yes No MP 295–320 Chron, No BChron, Type M 10s No No No Brooks et al. (2018)* and Potts et al. (2018) C 2010–2013 No Yes Yes LPH 0–78 Chron Yes M 1000s No No Yes d’Errico et al. (2020)* and Shipton et al. (2018) 37.067 OA 1975 No Yes Yes MP- > 30 LP Chron, Type No V Unknown No No No Only mention of bone is in a dissertation Merrick (1975) and Moutsiou (2011)* 34.163 OA 2008–2009; 2013 No Yes No LP Chron No M 10s No No No “obvious” butchery marks identified Blegen et al. (2017)* < 49 Dating Hominin remains No Poor preservation References Brooks et al. (2018)* and Potts et al. (2018) (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 230 Rusinga Island Wakondo Bovid Hill Kenya −0.419 34.175 OA 2007 (survey); 2009–2010 No Yes No LP 68 Chron No M 100s Yes Yes No Blegen et al. (2017)* and Jenkins et al. (2017) 231 Simbi Kenya −0.248 35.086 OA 1984–1990 No Yes No MP 36–45 Chron, Type No M 100s No No No McBrearty (1992)* and Blegen et al. (2021) 232 Songhor Kenya 0.028 35.231 OA 1981 No Yes No MP- > 94 LP Chron, Type No M 100s No No No McBrearty (1981)* and Blegen et al. (2021) 233 West Turkana, Lokalalei 1 (LA1) Kenya 3.954 35.788 OA 1987, 1991 Yes No No EP 2340 Chron No M 1000s No No No Possible stone Kibunjia (1994) tool cut marks and Tiercelin and possible et al. (2010)* percussion groove 234 West Turkana, Lokalalei 2C (LA2C) Kenya 3.949 35.779 OA 1997 Yes No No EP 2340 Chron No M, B, R, F 100s No No No Poor preservation and encrustation of bone surfaces 235 Chole Rockshelter Tanzania −3.045 32.533 RS 1968 No Yes Yes LPH Undated Type Yes M, R, F 1000s No No Yes Soper and Golden (1969)* 236 Isimila G18 Tanzania −7.896 35.605 OA 1957–1958 Yes No No MP 220–330 Chron, Type No M 100s No No No Cole and Kleindienst (1974), Coryndon et al. (1972), and Howell et al. (1962)* Source site References Roche et al. (1999) and Tiercelin et al. (2010)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age 237 Isimila G23 Tanzania −7.898 35.604 OA 1957–1958 Yes No No MP 238 Isimila H20 Tr 3 Tanzania −7.898 35.605 OA 1957–1958 Yes No No 239 Isimila H20 Tr 6 Tanzania −7.898 35.605 OA 1957–1958 Yes No 240 Isimila H21 Tanzania −7.898 35.605 OA 1957–1958 Yes No Source site Dates (ka) Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 222–330 Chron, Type No M 100s No No No Cole and Kleindienst (1974), Coryndon et al. (1972), and Howell et al. (1962)* MP 221–330 Chron, Type No M 100s No No No Cole and Kleindienst (1974), Coryndon et al. (1972), and Howell et al. (1962)* No MP 222–330 Chron, Type No M 100s No No No Cole and Kleindienst (1974), Coryndon et al. (1972), and Howell et al. (1962)* No MP 222–330 Chron, Type No M 100s No No No Cole and Kleindienst (1974), Coryndon et al. (1972), and Howell et al. (1962)* Dating References (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age 241 Isimila K14 Tanzania −7.896 35.607 OA 1957–1958 Yes No No MP 242 Kisese II Tanzania −4.492 35.812 RS 1935; 1956; 2013–2020 No Yes Yes 243 Kuumbi Cave Tanzania −6.347 39.519 C 2005; 2013? No No 244 Lake Eyasi WB9 Tanzania −3.544 35.273 OA 2004–2005 No 245 Lake Manyara, Makuyuni 101 Tanzania −3.571 36.119 OA 2007–2009 246 Lake Manyara, Makuyuni 4 Tanzania −3.555 36.092 OA 247 Loiyangalani (HcJd-1) Tanzania −2.739 34.876 OA Source site Dates (ka) Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 222–330 Chron, Type No M 100s No No No LPH 4 - > 46 Chron Yes M 1000s No No No Burials likely Holocene Marean and GiffordGonzalez (1991) and Tryon et al. (2018)* Yes LPH 1–20 Chron Yes M, F 1000s Yes Yes Yes Human remains are Holocene Prendergast et al. (2016) and Shipton et al. (2016)* Yes No MP 88–132 Chron, BChron Yes M, R, F 100s Yes Yes No DomínguezRodrigo et al. (2007c)* and DomínguezRodrigo et al. (2008) Yes No No MP 270–630 BChron, No Type M 10s No No No Giemsch et al. (2018)* 2007–2009 Yes No No MP 270–630 BChron, Yes Type M 10s No No No Giemsch et al. (2018)* 1979; 2003–2005 No Yes Yes LPH 65 M, R, F 1000s Yes Yes No Masele (2020) and Thompson (2005)* Dating Chron No References Cole and Kleindienst (1974), Coryndon et al. (1972), and Howell et al. (1962)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating 248 Magubike Tanzania −7.763 35.473 RS 2005–2006, 2008 No Yes Yes MP- 0–280 H 249 Mlambalasi Tanzania −7.591 35.500 RS 2006 No Yes Yes LPH 250 Mumba Tanzania −3.541 35.296 RS 1934–1936, 1979, 1981, 2005, 2014–2018 No Yes 251 Nasera Tanzania −2.742 35.358 RS 1932; 1975–1976; 2018 No 252 Njarasa Cave Tanzania −3.541 35.296 RS 1935–1936 No Source site Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Chron, Type Yes M, R 1000s Yes Yes No Collins and Willoughby (2010), Willoughby (2012)*, and Willoughby et al. (2018) 0.5–13 Chron Yes M 1000s Yes Yes No Collins and Willoughby (2010), Sawchuk and Willoughby (2015), and Willoughby (2012)* Yes MP- 0–132 H Chron Yes M, B, R, F, A 100,000s No No No Bushozi et al. (2020), Lehmann (1957), Mehlman (1979), and Prendergast et al. (2007)* Yes Yes LPH 2–61 Chron No M 100s No No No Yes Yes LPH > 40 Chron Yes M, B, R, F 1000s No Yes No Only taxonomically informative specimens reported References MartÍn-Perea et al. (2020)*, Mehlman (1977), Tryon and Faith (2016) Bader et al. (2020)* (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age 253 Olduvai Gorge Bed III JK ferruginous sand Tanzania −2.986 35.376 OA 1968–1971 Yes No No EP 254 Olduvai Gorge Bed III JK grey sand Tanzania −2.986 35.376 OA 1968–1971 Yes No No 255 Olduvai Gorge Bed III JK pink siltstone Tanzania −2.986 35.376 OA 1968–1971 Yes No 256 Olduvai Gorge Bed IV HEB Tanzania −2.994 35.352 OA 1968–1971 Yes 257 Olduvai Gorge Bed IV HEB East Tanzania −2.995 35.352 OA 1968–1971 258 Olduvai Gorge Bed IV HEB West Tanzania −2.994 35.352 OA 259 Olduvai Gorge Bed IV Intermediate Channel Tanzania −2.995 35.378 OA Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 800– 1100 Chron Yes M, B, R, F, A 10,000s No No Yes Specific layer of bone tool unspecified Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) EP 800– 1100 Chron Yes M, B, R, F, A 1000s No No Yes Specific layer of bone tool unspecified Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) No EP 800– 1100 Chron Yes M, B, R, F, A 100s No No Yes Specific layer of bone tool unspecified Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) No No MP 500–800 Chron Yes M, B, R, F, A 1000s No No Yes Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) Yes No No MP 500–800 Chron Yes M, B, R, F, A 100s No No No Leakey and Roe (1994)* and Njau et al. (2020) 1968–1971 Yes No No MP 500–800 Chron Yes M, B, R, F, A 1000s No No No Leakey and Roe (1994)* and Njau et al. (2020) 1968–1971 Yes No No MP 500–800 Chron Yes M, B, R, F 1000s No No No Leakey and Roe (1994)* and Njau et al. (2020) References Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 260 Olduvai Gorge Bed IV PDK Trench IV Tanzania −2.993 35.376 OA 1968–1971 Yes No No MP 261 Olduvai Gorge Bed IV PDK Trenches I-III Tanzania −2.993 35.376 OA 1968–1971 Yes No No 262 Olduvai Gorge Tanzania Bed IV WK East A −2.993 35.371 OA 1968–1971 Yes No 263 Olduvai Gorge Tanzania Bed IV WK East C −2.993 35.371 OA 1968–1971 Yes 264 Olduvai Gorge Bed IV WK Hippo Cliff Tanzania −2.993 35.370 OA 1968–1971 265 Olduvai Gorge Bed IV WK Lower Channel Tanzania −2.992 35.368 OA 266 Olduvai Gorge Bed IV WK Upper Channel Tanzania −2.993 35.371 267 Olduvai Gorge Masek FLK Tanzania −2.988 35.348 Source site Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 500–800 Chron Yes M, F 100s No No Yes Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) MP 500–800 Chron Yes M, F 1000s No No No Leakey and Roe (1994) and Njau et al. (2020) No MP 500–800 Chron Yes M, B, R, F, A 1000s No No Yes Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) No No MP 500–800 Chron Yes M, B, R, F, A 1000s No No No Leakey and Roe 1994*; Njau et al. 2020 Yes No No MP 500–800 Chron Yes M, R, F 100s No Yes Yes Channel or Hippo part of WK unspecified for butchery and tools Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) 1968–1971 Yes No No MP 500–800 Chron Yes M, B, R, F 1000s No Yes Yes Channel or Hippo part of WK unspecified for butchery and tools Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) OA 1968–1971 Yes No No MP 500–800 Chron Yes M, B, R, F 1000s No Yes Yes Channel or Hippo part of WK unspecified for butchery and tools Leakey and Roe (1994)*, Njau et al. (2020), and Pante et al. (2020) OA 1968–1971 Yes No No LP 500–800 Chron Yes M, B, R, F 1000s No No No Dating References Leakey and Roe (1994)* and Njau et al. (2020) (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 268 Olduvai Gorge Naisiusiu Beds Tanzania −2.971 35.330 OA 1931; 1969 No No Yes LP 10–62 Chron No M, B 10s No No No 269 Olduvai Gorge Ndutu Beds, Locality 26 (HdJe2) Tanzania −2.996 35.419 OA 1989–1990 No Yes No MP 260–500 Chron, Type No M, B Unknown No No No Type site Locality Eren et al. 26 from Hay (2014)* and Mabulla (1990) 270 Olduvai Gorge Trench 168 Tanzania −2.952 35.261 OA 2014–2016 Yes No No EP 2000 Chron No M 10s Yes No No Stollhofen et al. (2021)* 271 Olduvai Gorge, FLK Levels 10–12 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1840 Chron No M 10s Yes No No DomínguezRodrigo and Organista (2007), Hay (1976), and Leakey (1971)* 272 Olduvai Gorge, FLKW (FLK West) Tanzania −2.991 35.351 OA 2013 Yes No No EP 1644– 1698 Chron No M, R 1000s Yes Yes Yes Diez-Martín et al. (2015), Leakey (1971)*, and Yravedra et al. (2017) 273 Olduvai Gorge, BK Tanzania −2.995 35.344 OA 1935, 1952–1957, 1963, 2006, 2012 Yes No No EP 1340 Chron, PMag Yes M, B, R, F, A 10,000s Yes Yes Yes Backwell and d’Errico (2004), DomínguezRodrigo et al. (2014), Egeland (2010), Egeland and DomínguezRodrigo (2008), Hay (1976), Leakey (1971)*, Organista et al. (2019), and Renne et al. (2011) Source site References Eren et al. (2014) and Leakey et al. (1972)* Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 274 Olduvai Gorge, DK 1–3 Tanzania −2.983 35.383 OA 1931, 1960, 1962, 1963 Yes No No EP 1850 Chron No M, B, R, F, A 1000s Yes Yes Yes 275 Olduvai Gorge, EF-HR Tanzania −2.982 35.386 OA 1931, 1963, 2009–2013 Yes No No EP 1480– 1720 Chron No M, R 10,000s Yes No No Although fragmented and poorly preserved, possible percussion and bone tools de la Torre et al. (2013) and Leakey (1971)* 276 Olduvai Gorge, FC West Tanzania −2.992 35.352 OA 1931, 1963 Yes No No EP 1530 Chron Yes M, B, R, F, A 100s Yes No Yes Surface preservation poor Backwell and d’Errico (2004), Egeland (2010), Egeland and DomínguezRodrigo (2008), Hay (1976), and Leakey (1971)* 277 Olduvai Gorge, FLK 22 (Zinjanthropus) Tanzania −2.991 35.351 OA 1931–2, 1959–1962 Yes No No EP 1840 Chron Yes M, R, F, A 10,000s Yes Yes No Hay (1976), Leakey (1965, 1971)*; Parkinson (2018), and Potts (1988) 278 Olduvai Gorge, FLK Level 13 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1840 Chron No M 100s Yes Yes No DomínguezRodrigo and Organista (2007), Hay (1976), and Leakey (1971)* References Backwell and d’Errico (2004), Egeland (2007b, 2010), Hay (1976), Leakey (1971)*, and Potts (1988) (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 279 Olduvai Gorge, FLK Level 15 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1840 Chron No M 100s Yes No No DomínguezRodrigo and Organista (2007), Hay (1976), and Leakey (1971)* 280 Olduvai Gorge, FLK N 1–2 Tanzania −2.991 35.351 OA 1960, 1962 Yes No No EP 1750– 1760 Chron No M, R, A 1000s Yes Yes No DomínguezRodrigo and Barba (2007a, b), Hay (1976), and Leakey (1971)* 281 Olduvai Gorge, FLK N 3–4 Tanzania −2.991 35.351 OA 1960, 1962 Yes No No EP 1750– 1760 Chron No M, R, A 1000s Yes Yes No DomínguezRodrigo et al. (2007b), Hay (1976), and Leakey (1971)* 282 Olduvai Gorge, FLK N 5 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1750– 1760 Chron No M, R, A 1000s Yes Yes No Egeland (2007a, 2010), Hay (1976), Leakey (1971)*, and Organista et al. (2019) 283 Olduvai Gorge, FLK N 6 Tanzania −2.991 35.351 OA 1960, 1962 Yes No No EP 1750– 1760 Chron No M, A 1000s Yes No No DomínguezRodrigo et al. (2007a), Hay (1976), Leakey (1971)*, and Potts (1988) 284 Olduvai Gorge, FLK NN 3 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1840 Chron Yes M, R, F, A 1000s Yes Yes No DomínguezRodrigo et al. (2007b), Hay (1976), Leakey (1971)*, and Potts (1988) References Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 285 Olduvai Gorge, FLK NN1 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1840 Chron No M 100s Yes No No Barba and DomínguezRodrigo (2007), Hay (1976), and Leakey (1971)* 286 Olduvai Gorge, FLK NN2 Tanzania −2.991 35.351 OA 1960–1962 Yes No No EP 1840 Chron No M, R, F, A 100s Yes Yes No Egeland (2010), Egeland (2007c), Hay (1976), Leakey (1971)*, and Potts (1988) 287 Olduvai Gorge, HWK E 1 Tanzania −2.993 35.353 OA 1931, 1962–1963 Yes No No EP 1660– 1750 Chron No M, B, R, F 100s Yes Yes Yes Backwell and d’Errico (2004), Hay (1976), Leakey (1971)*, and Monahan (1996) 288 Olduvai Gorge, HWK E 2 Tanzania −2.993 35.353 OA 1931, 1962–1963 Yes No No EP 1660– 1750 Chron No M, R, F, A 100s Yes Yes No Hay (1976), Leakey (1971)*, and Monahan (1996) 289 Olduvai Gorge, HWK E 3–5 Tanzania −2.993 35.353 OA 1931, 1962–1963 Yes No No EP 1530– 1660 Chron No M 100s Yes No No Egeland and DomínguezRodrigo (2008), Hay (1976), and Leakey (1971)* 290 Olduvai Gorge, HWK EE Tanzania −2.993 35.353 OA 1971–1972, 2009–2015 Yes No No EP 1700 Chron No M, R 10,000s Yes Yes Yes de la Torre et al. (2018), McHenry and Stanistreet (2018), Leakey (1971)*, Pante et al. (2018, 2020), and Pante and de la Torre (2018) References (continued) Table 1 (continued) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age 291 Olduvai Gorge, JK2 Tanzania −2.986 35.376 OA 1931–1932, 1961–1962 Yes No No EP 292 Olduvai Gorge, MNK Main Site Tanzania −2.994 35.349 OA 1962–1963 Yes No No 293 Olduvai Gorge, SHK Tanzania −2.994 35.346 OA 1953, 1955, 1957, 1962, 2009–2011, 2012 Yes No 294 Olduvai Gorge, TK Tanzania −2.985 35.364 OA 1931–1932, 1963 Yes No Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 800– 1150 Chron Yes M 10,000s Yes Yes No Hay (1994), Kleindienst (1973), Leakey (1971)*, Pante (2013), and Yravedra et al. (2020) EP 1530 Chron Yes M, R, F, A 1000s Yes Yes Yes Backwell and d’Errico (2004), de la Torre et al. (2021), Egeland and DomínguezRodrigo (2008), Hay (1976), and Leakey (1971)* No EP 1340– 1540 Chron Yes M, B, R, F 1000s Yes Yes Yes Backwell and d’Errico (2004), Diez-Martín et al. (2014), DomínguezRodrigo et al. (2014), Egeland and DomínguezRodrigo (2008), Hay (1976), and Leakey (1971)* No EP 1330– 1480 Chron No M, A 100s Yes Yes No Surface preservation poor References Egeland and DomínguezRodrigo (2008), Hay (1976), Leakey (1971)*, and Yravedra et al. (2016) Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 295 Olduvai Gorge, TKSF Thiongo Korongo Tanzania −2.984 35.365 OA 1963, 2010 Yes No No EP 1353 Chron No M 100s Yes Yes No Panera et al. (2019)* 296 Olduvai Gorge, Victoria Cabrera Tanzania −2.992 35.353 OA 2017 No Yes No LP 70–90 Chron Yes M, R 100s Yes No No MaílloFernández et al. (2019)* 297 Peninj, ST Site Complex Tanzania −2.406 36.118 OA 1995-? Yes No No EP 1500 Chron, PMag No M, B, R, F 1000s Yes Yes No de la Torre et al. (2008)* and DomínguezRodrigo et al. (2002) 298 Ruhanga Rockshelter Tanzania −1.846 31.693 RS 2002 No Yes Yes LPH Undated Type No V 10s No No No Kwekason and Chami (2003)* 299 Munyama Uganda 0.174 33.233 C 1968 No No Yes LPH 10–15 No U Unknown No No No Van Noten (1971)* 300 Nsongezi Uganda Nyabusora facies Mile 8.5 trench −0.988 30.753 OA 1934–1935 Yes No No MP Undated Type No U Unknown No No No Cole (1967)* and O’Brien (1939) 301 Dungo V Angola −12.666 13.162 OA 1998 Yes No No MP 585–786 Chron No M 100s Yes No No Whale skeletons x Gutierrez et al. 2 “faily complete” (2001) and Lebatard et al. (2019)* 302 Leba – Horizons III-V Angola −15.083 13.259 C 1950, 1974, No 2018–present Yes No LP Undated Type No M, B 1000s Yes No Yes? No photo of tools, de Matos and Gautier notes Pereira (2020)* and Gautier (1995) 303 Hora 1 Malawi −11.659 33.644 RS 1950, 2016–2019 No Yes LPH > 9.5 Chron Yes M 10,000s No Yes Yes “Large quantities” Miller et al. of bone (2021)* 304 Mazinga 1 Malawi −11.723 33.702 RS 2017–present No Yes? Yes LPH > 9.5 Chron Yes V 10,000s No Yes Yes “Large quantities” Miller et al. of bone (2021)* 305 Caimane Cave Mozambique −26.228 32.151 C 1982 Yes Yes LPH Undated Type Yes V Unknown No No No No No Chron References Bicho et al. (2018) and Morais (1984)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size 306 Chicaza Rockshelter Mozambique −12.845 35.154 RS 2014 No No Yes LP 0–24 Chron No M Unknown No No No Bicho et al. (2016)* 307 Ngalue Cave Mozambique −12.859 35.198 C 2003–2008 No Yes Yes LP 42–105 Chron No M Unknown No No No Mercader et al. (2009)* 308 Nhamababwa Mozambique −18.550 34.879 C 2010 No Yes No LPH Undated Type No M 1000s No No No Mercader and Sillé (2013)* 309 Nhamissimbiti Mozambique −18.532 34.893 C 2010 No Yes Yes LPH 106–114 Chron, Type No M 100s No No No Good preservation Mercader and Sillé 2013* 310 Kabwe Zambia −14.432 28.452 C 1908, 1921 (selected) No Yes No MP 300 Chron, BChron Yes M Unknown No No Yes Barham (2002), Cooke (1963), Grün et al. (2020), and Trinkaus (2009)* 311 Kalemba – Horizons O-G Zambia −14.121 32.499 RS 1971 No Yes Yes LP 15.5 >37 Chron Yes M 100s No No Yes. Phillipson (1976)* 312 Leopard’s Hill – Spits 17–40 Zambia −15.250 28.450 C 1958 No No Yes LP 17–22 Chron Yes M 100s No No Yes Klein (1984)* 313 Mumbwa – Units Zambia IV-XIV −14.967 27.033 C 1925, 1930, 1939, 1993–1996 No Yes Yes LP 15 >170 Chron, BChron Yes M, R 100s Yes No Yes Barham (2000)* 314 Twin Rivers Kopje Zambia −15.562 28.131 OA 1954, 1956, 1997, 1999 No Yes No MP- 0–270 LP Chron Yes M 1000s Yes No No Avery (2003), Barham et al. (2000), Bishop and Reynolds (2000) and Clark and Brown (2001)* 315 Cave of Bees – Levels IV-VI Zimbabwe −20.504 28.510 C 1975 No No Yes LP Chron Yes M, B, R, F, A 10,000s No No Yes Source site 10.5–13 Bone Taph Butch Tools Comments Human remains are not Pleistocene References Walker (1995)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 316 Chelmer Spruit Zimbabwe −19.585 28.306 OA 1940s No Yes No MP 317 Hwange Main Camp Zimbabwe −18.723 26.952 OA 1990s No Yes No 318 Nswatugi – Members C-D Zimbabwe −20.536 28.475 C 1920s (test), 1932, 1975 No Yes 319 Pomongwe – Late Pleistocene deposits Zimbabwe −20.547 28.514 C 1960–1963, Late 1970s No 320 Redcliff Zimbabwe −19.010 29.460 C 1964–1965 321 Drotsky’s Cave (Gcwihaba) Botswana −20.200 21.250 RS 322 ≠Gi Botswana −19.632 21.010 323 White Paintings Shelter Botswana −18.771 324 Sehonghong Lesotho 325 Amis 10 Namibia Source site Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Undated BChron No M 100s No No No LP 100 Chron, Type No M 10s No No No Yes LP 10–41 Chron Yes M, B, R 10,000s No No Yes Walker (1995)* Yes Yes LP 11–16 Chron No M, R 100s No No Yes Walker (1995)* No Yes Yes? LP 25.5 >40 Chron Yes M, B, R 1000s No No No 1969, 1991 No No Yes LP 5.5 >12 Chron No M, B 1000s No No No Robbins et al. (1996b) and Yellen et al. (1987)* OA 1968–1969 No Yes Yes LPH Undated Type No M 100s No No Yes Brooks (1984) and Brooks and Yellen (1977a)* 21.748 RS 1988–1993 No Yes Yes LP 2–66 Chron No M, R 10,000s No No Yes Robbins et al. (2000)* −29.460 28.470 RS 1971, 1992, No 2014–present Yes Yes LP 1–57 Chron No M, B, R, F 10,000s No No Yes Mitchell (1994)*, Pargeter et al. (2017), and Plug and Mitchell (2008a, b) −21.224 14.465 RS 1984–1987 Yes Yes LPH Undated Type No M, R 100s Yes Yes No Breunig (1989)* and Van Neer and Breunig (1999) No Dating References Bond and Summers (1951)* and Cooke and Wells (1951) “Small” Assemblage size is only the published portion Haynes and Klimowicz (2009)* Cruz-Uribe (1983) and Klein (1978b)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age 326 Amis 11 Namibia −21.224 14.465 RS 1984–1987 No Yes Yes 327 Apollo 11 Namibia −27.701 17.117 RS 1969, 1972, 2007 No Yes 328 Bremen 1C Namibia −25.767 17.150 RS 1970 No 329 Erb Tanks Namibia −22.370 15.052 RS 2011–2012 330 Haalenberg Namibia −26.583 15.467 RS 331 Nos Namibia −25.417 15.583 332 Omungunda 99/1 Namibia −17.972 333 Oruwanje 95/1 −18.290 Source site Namibia Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments LPH 2 - >44 Type Yes M 100s Yes Yes No Breunig (1989)* and Van Neer and Breunig (1999) Yes LP 28–71 Chron No M, B, R, F 10,000s No No Yes Thackeray (1979), Vogelsang et al. (2010), and Wendt (1972)* Yes Yes LPH 3 - > 48 Chron No M, R, A 100s No No Yes Unclear if bone tools are Pleistocene Cruz-Uribe and Klein (1983) and Wendt (1972)* No Yes Yes LPH 5–130 Chron No M 100s No No No Poor preservation in Pleistocene Marks (2018) and McCall et al. (2011)* 1976 No Yes Yes LPH 0 - > 40 Chron No M, R 10s No No No Thackeray (1979) and Vogelsang (1998)* RS 1970 No No Yes LPH 0–22 Chron No M, R 10s No No No Thackeray (1979) and Vogelsang (1998)* 13.784 RS 1999 No No Yes LPH 0–15.5 Chron No M Unknown No No Yes Unclear if bone tools are Pleistocene, ostrich eggshell flask aperture is Pleistocene Vogelsang and Eichhorn (2011)* 13.572 RS 1995–1996 No No Yes LPH 0.5–11.7 Chron No M, R 100s Yes Yes Unclear if bone tools are Pleistocene Albrecht et al. (2001)* Yes References Source site Country Lat Long Year(s) Type excavated ESA MSA LSA Age 334 Oshilongo (Numas Gorge) Namibia −21.074 14.442 OA 1984–1987 No Yes Yes LPH 335 Ovizorombuku 98/6 Namibia −17.020 13.239 RS 2000 No No Yes 336 Pockenbank Namibia −27.130 16.310 RS 1969, 2015 No Yes 337 Zais Namibia −24.067 16.217 C 1930, 1968 No 338 Zebrarivier Namibia −24.310 17.160 C 1972 339 Anyskop_ Blowout South Africa −32.970 18.114 OA 1970s, 2001–2002 340 Blombos Cave South Africa −34.413 21.218 C Dates (ka) Hominin remains Vert Fauna Size Undated Type No V Unknown No No No LPH 2–12.7 No V Unknown No No No Vogelsang and Eichhorn (2011)* Yes LP Undated Type No M, B, R 100s No No Yes Thackeray (1979) and Wendt (1972)* Yes Yes LPH Undated Type No M, R 100s No No Yes No Yes Yes LP 12 - >48 Chron No M, B, R, A 100s No No No Yes Yes Yes MP- 182 LP No M, R 100s No Yes No 1991–1992, No 1999–present Yes Yes LP Yes M, R 10,000s Yes Yes Yes Dating Chron Chron, BChron 2 - >130 Chron Bone Taph Butch Tools Comments Poor preservation in MSA layers Unclear if bone tools are Pleistocene References Van Neer and Breunig (1999) and Vogelsang and Eichhorn (2011)* Cruz-Uribe and Klein (1983) and Wendt (1972)* Avery (1982b, 1983) and Cruz-Uribe and Klein (1983) NISP = 345 Kandel and Conard (2012)* Badenhorst et al. (2016), Henshilwood et al. (2001a)*, Jacobs et al. (2020), Reynard and Henshilwood (2019), and Thompson and Henshilwood 2011, 2014a, b (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 341 Boomplaas South Africa −33.383 22.183 C 1974–1979 No Yes Yes LP 2–66 Chron No M 10,000s Yes Yes Yes 342 Border Cave South Africa −27.025 31.989 C 1934, 1941–1942, 1970–1971, 1987, 2015–2017 No Yes Yes MP- 42–227 LP Chron Yes M Unknown No No Yes 343 Buffelskloof South Africa −33.304 21.520 RS 1974–1975 No No Yes LP 4–22 Chron No M, B, R 100s No No Yes Klein (1978a) and Opperman (1978)* 344 Bundu Farm South Africa −29.751 22.207 OA 1998–2003 No Yes No MP >145 Chron No M, B 100s Yes Yes No Hutson (2018) and Kiberd (2006)* 345 Bushman Rock Shelter South Africa −30.633 24.583 RS 1965, 1967–1975 No Yes Yes LP 9.5 >53 Chron No M, B, R, F, A 10,000s Yes Yes Yes Badenhorst and Plug (2012) and Louw (1969)* 346 Byneskranskop South Africa −34.579 19.452 C 1974, 1976 No No Yes LP 2–17 Chron Yes M, R Unknown No No Yes 347 Cave of Hearths South Africa −24.141 29.199 C 1947–1954 Yes Yes No MP- Undated Type LP Yes M, B, R 1000s No No Source site No References Avery (1977), Deacon (1979)*, Faith (2013), Klein (1978c), and Pargeter et al. 2018 “Moderate quantities” of bone “Substantial quantities” of bone Avery (1982a), Backwell et al. (2018), Butzer et al. (1979)*, and Klein (1977) Avery (1982b), Klein (1981), Loftus et al. (2016), Marean et al. (2000)*, and Schweitzer and Wilson (1978) Cooke (1962) and Cooke (1988)* Country Lat Long Year(s) Type excavated 348 Cooper’s Cave South Africa −26.029 27.718 C 349 Cornelia Beds South Africa −27.233 28.850 OA 350 De Kelders Cave 1 South Africa −34.550 19.371 351 Diepkloof South Africa −32.372 352 Dikbosch I South Africa 353 Drimolen, Drimolen Main Quarry (DMQ) South Africa Source site MSA LSA Age Dates (ka) 1938, 1954, Yes 2001–present No No EP 1375 1930s, 1953, 1998 Yes No No EP C 1969, 1973, 1992–1995 No Yes Yes 18.456 RS 1973, 1986, 1999–2013 No Yes −28.398 23.549 RS 1973 No −25.985 27.710 C 1992–present Yes ESA Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Chron, BChron Yes M 10,000s Yes No No 990– 1070 PMag Yes M 100s No No No Brink et al. (2012), Butzer et al. (1974), and Cooke (1974)* LP 70 Chron Yes M, R 100,000s Yes Yes No Armstrong (2016), Avery (1982b), Klein and Cruz-Uribe (2000a, b), Marean et al. (2000)*, and Schwarcz and Rink (2000) Yes LP 52–100 Chron Yes M, B, R 1000s Yes Yes No Miller et al. (2013)*, Steele and Klein (2013), Tribolo et al. (2013), Val (2019), and Val et al. (2020) No Yes LP 3–13.5 Chron No M, B, R 100s No No Yes Humphreys (1974)* No No EP 1789– 2673 Chron, BChron Yes M 1000s No No Yes Adams et al. (2016), Pickering et al. (2019), and Stammers et al. (2018)* Dating Only a subset of the fauna was analyzed taphonomically; taphonomy only superficially reported References de Ruiter et al. (2009), Pickering et al. (2019), Sutton et al. (2017), and Stammers et al. (2018)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age South Africa −33.665 18.428 OA 1973, 1975, 1997–2001 Yes No No MP 355 Elands Bay Cave South Africa −32.318 18.318 C 1970–1978, 2011 No No Yes 356 Elandsfontein South Africa −33.083 18.250 OA 1950s, 1960s, 1980s, 1990s, 2008–2012 Yes No 357 Equus cave South Africa −27.614 24.630 C 1978, 1982 No 358 Erfkroon (Orangia Terrace) South Africa −28.520 25.350 OA 2006–2014 359 Erfkroon EFK-1 South Africa −28.869 25.594 OA 360 Erfkroon EFK-2 South Africa −28.529 25.354 OA Source site 354 Duinefontein (DFT) 2 Dates (ka) Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 250–500 Chron, BChron No M, B, R 1000s Yes Yes No LP 13.5– 236 Chron No M, B, R, F 1000s Yes No No No MP 600– 1000 BChron, Yes PMag M, R 10,000s Yes Yes No Braun et al. (2013), Forrest et al. (2018), Klein et al. (2007)*, and Stammers et al. (2018) Yes Yes LP 2–27 Chron Yes M 10,000s No No Yes Avery and Avery (2011), Grine and Klein (1985)*, and Klein et al. (1991) No Yes Yes LP 163 Chron No M Unknown No No No 1996–1999 No Yes Yes LP 25 Chron No M 100s No No No Brink et al. (2015) and Churchill et al. (2000)* 1996–1999 No Yes No LP 113 Chron No M 100s No No No Brink et al. (2015) and Churchill et al. (2000)* Dating References Cruz-Uribe et al. (2003) and Klein et al. (1999)* Taphonomy for micromammals only Avery (1987), Klein and Cruz-Uribe (1987), Klein and Cruz-Uribe (2016), and Porraz et al. (2016)* “Large sample” of Brink et al. bone (2015)* Country Lat Long Year(s) Type excavated 361 Florisbad South Africa −28.768 26.070 OA 362 Geelbek_Alice South Africa −33.190 18.153 363 Geelbek_Bay35 South Africa −33.191 364 Geelbek_Equus South Africa 365 Geelbek_Homo Source site Dating Hominin remains Vert Fauna Size Chron Yes M, R, A 100s Yes No No Brink (1988), Churchill et al. (2000), Kuman et al. (1999)*, and Toffolo et al. (2015) LP 4.5–16.5 Chron No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) Yes LP Undated Type No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) Yes Yes LP Undated Type No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) No Yes Yes LP 19–27 Yes M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) 1998–2005 No Yes Yes LP Undated Type No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) OA 1998–2005 No Yes Yes LP Undated Type No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) OA 1998–2005 No Yes Yes LP 11 No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) ESA Dates (ka) MSA LSA Age 1932, 1952, No 1980s, 1990s Yes Yes MP- 15–279 LP OA 1998–2005 No Yes Yes 18.153 OA 1998–2005 No Yes −33.178 18.162 OA 1998–2005 No South Africa −33.191 18.153 OA 1998–2005 366 Geelbek_Loop South Africa −33.179 18.164 OA 367 Geelbek_Mrose South Africa −33.181 18.163 368 Geelbek_Rhino South Africa −33.180 18.152 Chron Chron Bone Taph Butch Tools Comments References (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 369 Geelbek_Snoek South Africa −33.180 18.158 OA 1998–2005 No Yes Yes LP 370 Geelbek_Stella South Africa −33.185 18.156 OA 1998–2005 No Yes Yes 371 Geelbek_Toaster South Africa −33.183 18.154 OA 1998–2005 No Yes 372 Gladysvale South Africa −25.540 27.450 C 1992–2002 Yes 373 Grassridge Rockshelter South Africa −31.342 26.511 RS 1979, 2014–2016 374 Groot Kloof GK-B South Africa −24.183 28.350 C 375 Heuningneskrans South Africa −24.360 30.390 376 Hoedjiespunt 1 −33.001 17.953 Source site South Africa Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Undated Type No M, B, R 100s Yes Yes Yes LP 5–11 No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) Yes LP Undated Type No M, B, R 100s Yes Yes Yes Fuchs et al. (2008)* and Kandel and Conard (2012) Yes No MP- 570 LP >1000 Chron, Yes BChron, PMag M 10,000s No No No Avery (1995), Berger et al. (1993)*, Lacruz et al. (2002), and Pickering et al. (2007) No Yes Yes LP 7–43 Chron M, B Unknown No No Yes 2005 No Yes Yes MP 154–780 Chron, No BChron, Type M Unknown No No No Curnoe et al. (2006)* RS 1968, 2018 No No Yes LP 8 - >30 Chron Yes M 10,000s No No Yes Beaumont (1981)*, Klein (1984), and Porraz and Val (2019) OA 1994–1996, 2011 No Yes Yes LP 100–130 Chron Yes M Unknown No No No Dating Chron No NISP = 879 Bone is “abundant” Most bone comes from older paleontological deposits References Fuchs et al. (2008)* and Kandel and Conard (2012) Ames et al. (2020) and Opperman (1988)* Kyriacou et al. (2015) and Matthews et al. (2006)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 377 Jubilee shelter South Africa −25.420 27.550 RS 1980s No No Yes LP 1 - >29 378 Kalkbank South Africa −23.493 29.211 OA 1954, 1966 No Yes Yes? 379 Kathu Pan South Africa −27.395 23.030 C 1978–1982, 2014–2016 Yes Yes 380 Klasies River South Africa Mouth Main Site −34.108 24.390 C 1966–68, No 1984–1995, 2015–present 381 Klipdrift South Africa −34.452 20.724 C 2010–15 382 Klipfonteinrand South Africa −32.072 19.130 RS 1969, 2011–12 383 Knysna Heads South Africa −34.083 23.066 384 Kromdraai A South Africa −26.011 27.750 Source site Hominin remains Vert Fauna Size Chron, Type No M, B, R, A Unknown No Yes Yes MP- >17 LP Chron, Type No M, B, R 1000s Yes Yes Yes Hutson and Cain (2008)* and Mason et al. (1958) Yes MP- 10–464 LP Chron No M 100s No No No Klein (1988)* and Lukich et al. (2020) Yes Yes LP 57–110 Chron Yes M Unknown Yes Yes No No Yes Yes LP 11–71 Chron Yes M, R 10,000s Yes Yes Yes Discamps et al. (2020)*, Henshilwood et al. (2014), and Reynard et al. (2016) No Yes Yes LP 14–22 Chron No M, B, R 100s No No No Mackay et al. (2020)* C 2013–present No Yes Yes LP 18–39 Chron No M 10,000s No No No Matthews et al. (2019)* C 1947, 1993–2002 No No EP <1800 BChron, No PMag M, B, R 1000s Yes No No Braga et al. (2017)*, Brain (1981), Kuman et al. (1997), McKee (1996), and Stammers et al. (2018) Yes Dating Bone Taph Butch Tools Comments 79.25 kg of bone 182.7 kg of bone References Turner (1986) and Wadley (1986)* Avery (1987), Binford (1984), Klein (1976)*, Turner (1989), and van Pletzen-Vos et al. (2019) (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated 385 Kromdraai B South Africa −26.011 27.750 C 386 Lincoln Cave South Africa −25.960 27.750 387 Melkhoutboom South Africa −33.190 388 Nelson Bay Cave South Africa 389 Pinnacle Point 13B Source site South Africa ESA Dates (ka) Dating 1600– 1800 BChron, Yes PMag MSA LSA Age 1938–1944, Yes 1955–1956, 1977–1980, 1993–2002, 2002–present No No EP C 1997, 1998 Yes Yes No MP- 115–252 Chron LP 25.470 C 1930 No No Yes LP 2–15 −34.100 23.317 C 1964–1965, 1970–1971 No Yes Yes LP −34.207 22.090 C 2000–2012 No Yes No Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments References M, B, R 10,000s Yes No Yes Braga et al. (2017)*, Fourvel et al. (2018), Herries and Adams (2013), and Stammers et al. (2018) Yes M 100s No No No Reynolds et al. (2007)*, Reynolds and Kibii (2011), and Reynolds et al. (2003) Chron Yes M 10s No No Yes Assemblage is “small” Deacon (1969)*, Hewitt (1931), and Klein (1974) 5.5–23 Chron Yes M Unknown No No No Assemblage is “rich” Avery (1982b), Klein (1972)*, Klein (1983b), and Loftus et al. (2016) MP- 91–162 LP Chron Yes M, R 10,000s Yes No Yes Jacobs (2010), Marean et al. (2010)*, Matthews et al. (2009), Rector and Reed (2010), and Thompson (2010) Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) 390 Pinnacle Point 5–6 South Africa −34.206 22.090 C 2012–2017 No Yes No LP 54–99 391 Plover’s lake South Africa −25.584 27.464 C 1989, 2002–2004 No Yes No LP 392 Pniel 6 South Africa −28.382 24.303 OA 1984, 2000 Yes Yes 393 Putslaagte 8 South Africa −32.536 18.849 RS 2010 No 394 Rose Cottage Cave South Africa −29.130 27.280 C 1943–46, 1962, 1987–1997 395 Sea Harvest Site South Africa −32.979 17.989 C 396 Shongweni South Africa −29.515 30.432 397 Sibhudu (Sibudu) South Africa −29.523 398 Spitzkloof South Africa −28.518 Source site Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments Chron, Type No M, B, R, A 10,000s Yes Yes No Armstrong (2016), Marean et al. (2010)*, Matthews et al. (2020), and Smith et al. (2018) 63–88 Chron Yes M 10,000s Yes Yes No de Ruiter et al. (2008)* No MP- >120– LP 500 Chron, Type No M, B, R 1000s Yes Yes No Hutson (2018)* Yes Yes LP 17–75 Chron No M, B, R, F Unknown No No No No Yes Yes LP 27–95 Chron No M, B 100s No Yes Yes 1980 No Yes No LP > 40 Chron Yes M Unknown No No No Most bone comes from older paleontological deposits Grine and Klein (1993)* C 1971, 1981, No No Yes LP 2–23 Chron No M, B, R, F 100s No No Yes Human remains “rumored” to be present Davies (1975)* 31.086 RS 1983, 1998–2011 No Yes Yes LP 39–77 Chron Yes M, B, R, F, A 10,000s Yes Yes Yes Cain (2006), Clark (2019), Clark and Plug (2008), Collins (2016), Jacobs et al. (2008), and Wadley and Jacobs (2004)* 17.047 RS 2010–11 No Yes Yes LP Undated Type No M, R 1000s Yes Yes No Dewar and Stewart (2012)* Dating 11.7 kg of bone References Mackay et al. (2015)* Loftus et al. (2019); Plug and Engela (1992), and Wadley (1997)* (continued) Table 1 (continued) Country Lat Long Year(s) Type excavated 399 Sterkfontein 5 South Africa −26.050 27.700 C 400 Strathalan Cave B South Africa −30.592 28.232 401 Swartkrans Member 1 South Africa −26.059 402 Swartkrans Member 2 South Africa 403 Swartkrans Member 3 404 Umhlatuzana Shelter Source site MSA LSA Age Dates (ka) 1956–1958, Yes 1966–present No No EP 1784 C 1988–1991 No Yes No LP 27.665 C 1948, Yes 1967–present No No −26.059 27.665 C 1948, Yes 1967–present No South Africa −26.059 27.665 C 1948, Yes 1967–present South Africa −29.482 30.452 RS 1985 ESA No Vert Fauna Size Bone Taph Butch Tools Comments Chron, Yes BChron, PMag M, B, R 10,000s Yes Yes Yes 21–24 Chron No M Unknown No No No EP 1600– 2300 Chron Yes M, B 1000s Yes Yes Yes Brain (1981), Pickering et al. (2008), and Stammers et al. (2018)* No EP 1100– 1700 Chron Yes M, B, R 1000s Yes Yes Yes Brain (1981), Herries and Adams (2013), Pickering et al. (2008), and Stammers et al. (2018)* No No EP 960 Chron Yes M, B, R 100s Yes Yes Yes Brain (1981), Gibbon et al. (2014), Pickering et al. (2008), and Stammers et al. (2018)* Yes Yes LP 3–45 Chron No M, B 100s No No Yes Kaplan (1989*, 1990) Dating Hominin remains References Brain (1981), Kuman (1994), Pickering et al. (2019), Pickering (1999), Stammers et al. (2018), and Toussaint et al. (2003)* 3870.9 g of bone Opperman and Heydenrych (1990)* Country Lat Long Year(s) Type excavated ESA MSA LSA Age Dates (ka) Dating Hominin remains Vert Fauna Size Bone Taph Butch Tools Comments 405 Varsche Rivier 003 South Africa −31.524 18.604 RS 2009, 2011 No Yes Yes LP 3–61 Chron No M, B, R 1000s No Yes No Steele et al. (2012,2016)* 406 Vlakkraal South Africa −28.500 26.040 OA 1940? Yes Yes No MP Undated Type No M, B, F 100s No No No Cooke (1962) and Wells et al. (1942)* 407 Waterfall Bluff South Africa −31.434 29.823 RS 2015–present No Yes? Yes LPH 10.7–32 Chron No M, F 10,000s No No Yes Fisher et al. (2020)* 408 Wonderwerk Cave South Africa −27.846 23.555 C 1978–1996, Yes 2007–present Yes Yes EPLP 0>1070 Chron, PMag No M, R, F 10,000s No No No Avery (2007), Beaumont and Vogel (2006)*, Brink et al. (2016), Chazan et al. (2020), and Rüther et al. (2009) 409 Ysterfontein (YFT1) South Africa −33.351 18.144 RS 2002–2008 Yes No LP >46–132 Chron No M, B 1000s No No No Avery et al. (2008), Halkett et al. (2003)*, and Klein et al. (2004) Source site No References (RS Rock Shelter, C Cave, OA Open Air), years of excavation, chronological information (PL Pliocene, EP Early Pleistocene, MP Middle Pleistocene, LP Late Pleistocene, H Holocene), dating methods (Chron Chronometric, BChron Biochronology, PMag Paleomag, Type Typology), if hominin remains are likely associated (Yes/No), reported vertebrate faunal remains (U Unspecified, V Vertebrates not specified, M Mammal, B Bird, R Reptile, F Fish, A Amphibian), assemblage sizes (orders of magnitude), if a taphonomic study was done (Yes/No), if butchery is noted (Yes/No), if bone tools are present (Yes/No), any comments, and references for the faunal assemblages specifically. References that may or may not be about faunal assemblages but refer to the source for the map location are indicated with an asterisk 2020 J. C. Thompson et al. Fig. 2 Distributions of time spans of Pleistocene archaeofaunal assemblages in different regions of Africa. Black bars represent the time spans within which the assemblages could date, not the total span that they actually cover. Gray bars represent assemblages where only minimum or maximum ages are available, indicating that spans are probably longer. Dotted bars represent assemblages marked as “undated” in Table 1. Longer spans typically represent poor chronological resolution rather than long spans of coverage (for example, an assemblage coarsely listed as being Early–Middle Pleistocene would span from 0.126 to 2.58 Ma) The Zooarchaeology of Pleistocene Africa 2021 remains within a site. For open-air sites, geological formations are divided into members that provide the opportunity to assign stone artifacts and fossils to a general time range, and, sometimes, these are reported together as a single assemblage. Within these members, outcrops may be designated as geological beds based on their stratigraphic provenience or may be assigned locality identifiers based on their horizontal provenience. Multiple findspots or excavations may exist within a single locality, and if the excavation is deep enough, it may cross-cut different beds or even different members. Similarly, the concept of a “site” can vary. It may refer to a naturally constrained area such as the space under the overhang of a rock shelter, a particularly dense accumulation on an open landscape where many artifacts and fossils occur, or even the boundaries of the excavated area itself. The term “locality” is sometimes used synonymously with “site”, or it may reference a larger general area within which a number of “sites” are located. These differences in reporting make it impossible to produce truly quantitative comparisons between regions, for example, differences in the densities of faunal accumulation per unit time. However, it is possible to discern general patterns in preservation, research emphasis, and reporting that summarize current knowledge about Pleistocene archaeofaunal research. For our purposes, we have defined an assemblage as the smallest, spatially constrained, aggregate of faunal remains from a Pleistocene-age context. For example, if a multilayered site contains archaeofaunal remains from different parts of the Pleistocene, these are indicated as a single assemblage and we have adjusted the associated timeframes to encompass the span of the entire assemblage. If multiple excavated sites from a single locality are published, each with its own name and faunal list, then each one constitutes an “assemblage,” even if they overlap in time. A summary by country and region of the assemblages in our compilation is provided in Fig. 3. For each region, we offer a brief history of Pleistocene zooarchaeological research and discernible patterns in site distributions and research emphases Site Types by Region and Country 5 1 1 1 20 1 51 55 52 2 1 1 Northern Northern-Cent. Cent. Cave Horn Rock shelter 4 11 15 2 33 2 Eastern Malawi Angola Uganda Kenya 3 1 8 Tanzania Somalia 5 Somaliand Ethiopia DRC 1 3 Djibouti Sudan Niger Nigeria Spain Tunisia Libya 1 Morocoo 2 Egypt 2 4 2 3 1 2 Southern-Central Southern South Africa 1 22 Lesotho 2 23 Namibia 1 1 Botswana 1 1 2 1 Zambia 37 1 1 Zimbabwe 9 1 Mozambique 1 Algeria 100% 90% 80% 70% 60% 50% 40% 30% 20% 10% 0% Open Air Fig. 3 Counts of different site types by region and country. Note that Benzú is geographically in northern Africa, but in the political territory of Spain 2022 J. C. Thompson et al. for the general regions shown in Fig. 1, which have the most available information: northern Africa, the Horn, eastern Africa, southern-central Africa, and southern Africa. There is also a combined section for the data-poor regions of western Africa, northern-central Africa, and central Africa. Northern Africa Here, northern Africa is defined as the modern countries of Algeria, Egypt, Libya, Morocco, and Tunisia (Fig. 4). These include portions of the geophysical and geopolitical regions of the Sahara, Sahel, Maghreb, and Nile valley. With 84 assemblages and spanning 7,402,246 km2, this region has approximately 1.1 Pleistocene faunal assemblages for every 100,000 km2. Geopolitically, the Maghreb refers to Morocco, the disputed territory of Western Sahara, Mauritania, Algeria, Tunisia, and Libya. The original usage of the term “Maghreb” refers to the Atlas Mountains and the coastal areas of Morocco, Algeria, and Tunisia, where maghrib is literally translated from Arabic to mean the west. We follow this usage here but acknowledge that these areas share the vegetation and rainfall regimes known as the Mediterranean Coastal Biotic Zone (Happold & Lock, 2013). There is a long history of archaeological fieldwork in the Maghreb dating from the 1880s. Following establishment of the French colonial territory in Algeria in 1830, French scientific missions were sent for investigation. As the French North African territory expanded in 1881 with the formation of the French protectorate of Tunisia, and again in 1912 with the formation of the French protectorate in Morocco, French-led scientific missions expanded. Gabriel Carrière (1886) published the first descriptions of Mousterian archaeological findings from the Maghreb, including pedunculate (or tanged) stone tools from Eckmuhl Cave in northwestern Algeria, Fig. 4 Map of sites with Pleistocene archaeofaunas in northern Africa. Grayscale is topography, with lighter colors being higher elevation. Numbers reference sites in Table 1. White dots with black centers indicate other sites in the neighboring regions, numbered in each of their respective sections. Major rivers and lakes are indicated in blue The Zooarchaeology of Pleistocene Africa 2023 located on the Mediterranean coast. After work from 1919 to 1920, Maurice Reygasse published on pedunculates from Bir el Atir, an inland open-air site in Algeria in the Oued (wadi) Djebbana, and recognized them as belonging to the Mousterian. In 1922, following further excavations at a site neighboring Bir el Atir in the Oued Djebbana, where more pedunculates were found, Maurice Reygasse (1921–1922) coined the term “Aterian” to describe the stone tools, making the Bir el Atir pedunculates the type specimens of the Aterian, a marker of the northern African MSA (Middle Paleolithic/Mousterian). Bir el Atir was then destroyed (Morel, 1974a) by a meandering river in the Oued Djebbana. Unfortunately, the collections included faunal materials, charcoal, ochre, and a rich lithic assemblage that were never thoroughly published, and little else is known about this site. In the 1930s, archaeologists working in the Maghreb focused more on stratified contexts than did their predecessors, who had primarily collected artifacts from surface contexts. Breuil and Frobenius (1931) described the locations and findings from surface and cave contexts that had been discovered prior to 1930. Early Pleistocene archaeofaunal assemblages are rare, but vertebrate faunal remains found in association with Oldowan (ESA) stone tools from the Ain Hanech Formation in eastern Algeria were first described in the 1930s and 1940s (Arambourg, 1979). Within the Ain Hanech Formation, three fossiliferous localities have since been discovered and include Ain Hanech, Ain Boucherit, and El Kherba (Sahnouni et al., 2011). These three localities range in age from ~1.77 to 2.4 Ma and contain vertebrate assemblages that have been described in detail both taxonomically and taphonomically (Sahnouni et al., 2018). Other archaeofaunal assemblages in Algeria that are Early Pleistocene in age include Erg Tihodaine, Lac Karar, and Mansourah (Boule, 1900; Chaid-Saoudi et al., 2006; Thomas, 1977). These are open-air sites that have not been subjected to detailed taphonomic study, and the total vertebrate assemblage sizes are currently unknown. In Morocco, the open-air site of Fouarat contains a Late Pliocene/Early Pleistocene vertebrate archaeofaunal assemblage that was excavated in the 1940s (Choubert et al., 1948) and has not been described in taphonomic detail. Sidi Zin is an open-air site in Tunisia that was discovered and partially excavated in the 1940s; it contains Middle Pleistocene deposits (Belhouchet, 2002) and a vertebrate archaeofaunal assemblage that has yet to be studied taphonomically (Vaufrey, 1950). The collapsed caves of the Thomas Quarry near Casablanca, Morocco, were discovered in 1978 (Raynal et al., 2002), and excavations have continued to the present. Their archaeofaunas have been described in detail both taphonomically and taxonomically and date to the Early–Middle Pleistocene, between ~1 Ma and 500 ka (Daujeard et al., 2020). Another notable cave site dating to the Middle Pleistocene in Morocco is Jebel Irhoud. Ennouchi (1962) described a hominin cranium from a pocket of a barite mine in Jebel Irhoud, Morocco. The red clay sediments were described as containing pebbles, no trace of a lithic industry, and disarticulated faunal skeletons typical of Lower Paleolithic contexts. Subsequent work has shown that the site is in fact Middle Pleistocene and contains a lithic assemblage dated to ~315 ka (Richter et al., 2017). The vertebrate faunal assemblage from Jebel Irhoud 2024 J. C. Thompson et al. has also since been described in detail both taxonomically and taphonomically (Amani & Geraads, 1993; Geraads et al., 2013; Hublin et al., 2017). In the Nile valley, extensive surveys and excavations in the 1970s and 1980s revealed Middle Pleistocene and Late Pleistocene archaeofaunas from several openair sites that were in proximity to Pleistocene fluvial and lacustrine systems. Kharga Oasis KO10C contains Middle Pleistocene vertebrate archaeofaunal remains, although the butchery patterns and taphonomy of the assemblage are unknown (Caton-Thompson & Gardner, 1932). Bir Tarfawi and Bir Sahara East localities are Middle-to-Late Pleistocene open-air sites that were excavated in the 1970s and 1980s by Angela Close, Romuald Schild, and Fred Wendorf (Wendorf et al., 1993). Bir Tarfawi BT-14 stands out as having a rich Late Pleistocene archaeofaunal assemblage with well-preserved large mammals, birds, fish, and amphibians that have been described in detail (Gautier, 1993; Kowalski et al., 1993; Van Neer, 1993). In the Wadi Kubbaniya localities that were excavated in the 1970s and 1980s, large samples of vertebrate and invertebrate archaeofaunas were identified and described by Achilles Gautier and Wim Van Neer (1989). They note that the Late Pleistocene large mammal bones in the Wadi Kubbaniya assemblage are highly fractured and weathered and that it is difficult to determine whether humans consumed a high number of avian and molluscan remains recovered from the Wadi Kubbaniya localities. There are few Late Pleistocene archaeofaunal assemblages in Libya, though the cave site of Haua Fteah stands out as having a large, well-preserved, and taxonomically rich faunal assemblage that has informed our understanding of Pleistocene human foraging patterns and butchery activity in northern Africa (Klein & Scott, 1986). The majority of important Pleistocene archaeofaunas in the Maghreb date to the Late Pleistocene. Taforalt Cave (or Grotte des Pigeons, Beni-Snassen massif, Morocco) was first excavated by Armand Ruhlmann from 1944 to 1947 (Barton et al., 2016; Ruhlmann, 1945), then by Jean Roche from 1951 to 1955 (Roche, 1953), and again by Roche from 1969 to 1972 (Roche, 1973). From 2003 to present, renewed excavations at Taforalt were carried out by Abdeljalil Bouzouggar and Nick Barton, and the findings include Pleistocene to Holocene vertebrate (Turner et al., 2020) and invertebrate (Bouzouggar et al., 2007; Taylor et al., 2011) archaeofaunas as well as a Late Pleistocene human burial area (Humphrey et al., 2019). Several more cave sites in Morocco were discovered, excavated, and described in the 1950s, even if they were excavated earlier. For example, Armand Ruhlmann (1951) describes excavations at the cave site of Dar es-Soltan I on the Atlantic coast of Morocco, undertaken from 1937 to 1938. Archaeologically sterile layers were observed between Aterian layers, and the cave sequence was capped by a 1.5-m shell midden (Barton et al., 2009; Ruhlmann, 1951). Roche discovered both El Mnasra and Contrebandiers Cave in the 1950s and described the stratigraphy of Contrebandiers Cave based on a first sounding in the cave from 1955 to 1957 (Vallois & Roche, 1958). From 2007 to 2011, a joint Moroccan–American excavation led by Mohamed Abdeljalil El Hajraoui and Harold Dibble at Contrebandiers revealed 20,000+ piece-plotted vertebrate and invertebrate faunal remains, suggesting a much larger total faunal collection. The vertebrate and invertebrate faunal The Zooarchaeology of Pleistocene Africa 2025 remains have been described in detail both taphonomically and taxonomically (Dibble et al., 2012; Hallett, 2018; Steele & Álvarez-Fernández, 2011). Following test excavations in the 1960s by Roche, El Mnasra was excavated in the 1990s and from 2004 to present by Roland Nespoulet and El Hajraoui (Nespoulet et al., 2008). The taphonomic and taxonomic analyses of the vertebrate and invertebrate remains have been published in detail by Emilie Campmas (Campmas, 2017; Campmas et al., 2015). Other Late Pleistocene sites in the Atlantic coastal region of Morocco include El Harhoura 1 and 2, which were first excavated by André Debénath in 1977 and were noted to contain shellfish (Mytilus, Patella, and Purpura) in their stratified deposits (Stoetzel et al., 2014). El Harhoura 2 was excavated again from 2001 to present by Nespoulet and El Hajraoui (Stoetzel et al., 2014). Invertebrates (Nouet et al., 2015), small-bodied vertebrates (Stoetzel et al., 2011, 2012), and large-bodied vertebrates (Campmas et al., 2017; Michel et al., 2009) from El Harhoura 2 have been described in detail, and taphonomic analyses of the vertebrate remains are extensive (Campmas et al., 2017). Tamar Hat, a rock shelter in Algeria, was originally excavated in the 1930s by Camille Arambourg (Arambourg et al., 1934) and then excavated again in the 1970s by a British team (Saxon et al., 1974) focused on the Iberomaurusian (LSA) deposits. Merzoug and Sari (2008) described the vertebrate remains from Tamar Hat in detail and found that humans predominantly accumulated Barbary sheep (Ammotragus lervia) that make up roughly half of the archaeofaunal assemblage. Late Pleistocene archaeofaunas in the Maghreb tend to include large-, medium-, and small-bodied mammals as well as birds, reptiles, fish, and mollusks that were accumulated in cave sites and rock shelters. In the Nile valley, nearly all Late Pleistocene archaeofaunas are known from open-air sites that contain medium- and smallbodied mammals, birds, fish, and freshwater mollusks. In contrast to the large and taxonomically rich Late Pleistocene archaeofaunas from sheltered sites in the rest of northern Africa, Early and Middle Pleistocene archaeofaunal assemblages have mainly been recovered from open-air sites and collapsed caves and are dominated by large-bodied mammals. Northern-Central and Central Africa Unlike the long research traditions of northern Africa, research on Pleistocene fossil localities in northern-central and central Africa has been sporadic (Fig. 5). Most known sites are from Sudan, which has a number of promising archaeological and fossil localities (Harcourt-Smith et al., 2012), including sites that are reported but not excavated (Chaix et al., 2000). The central African equatorial rainforest region has a particularly poor published faunal record, especially from Pleistocene deposits, and, thus, there is no clear history of Pleistocene faunal research here (Cornelissen, 2013). Northern-central Africa covers a land area of 5,612,479 km2, and central 2026 J. C. Thompson et al. Fig. 5 Map of sites with Pleistocene archaeofaunas in northern-central and central Africa. Grayscale is topography, with lighter colors being higher elevation. Numbers reference sites in Table 1. White dots with black centers indicate other sites in the neighboring regions, numbered in each of their respective sections Africa covers 4,731,263 km2. With 23 and 8 assemblages, respectively, they each have fewer than 0.5 Pleistocene archaeofaunal assemblages per 100,000 km2. Recent work in these regions includes excavations at localities such as Site 047 in Sudan that preserves Early Pleistocene artifacts in the same deposits as mammalian fossils, but it is not clear whether the two are in association (Abbate et al., 2010). Prominent Middle Pleistocene hominin-bearing sites such as Singa suffer from the same contextual problems (McDermott et al., 1996). The majority of reported sites that do have excavated Pleistocene faunas and artifacts found together, including those across from Wadi Halfa (Yeshurun, 2018) and HP766 at Wadi Umm Rahau (Gautier et al., 2012), have subsequently been destroyed by flooding associated with works along the Nile. The Zooarchaeology of Pleistocene Africa 2027 In the western part of the region in Niger, only two open-air sites with Late Pleistocene archaeofaunas are known: Adrar Bous and Gobero. Adrar Bous, a locality around a large granitic ring, was surveyed and excavated in the 1960s and 1970s by J. Desmond Clark and Diane Gifford-Gonzalez, and several Aterian (MSA) localities were discovered. A small amount of vertebrate archaeofaunal remains were discovered at the Diatomite 2 locality that contains Aterian artifacts. These remains are poorly preserved, although the younger Holocene vertebrate remains from Adrar Bous are well-preserved and have been described in detail both taxonomically and taphonomically (Gifford-Gonzales & Parham, 2008). Gobero was excavated in the 2000s and contains a large cemetery of Holocene human remains near the paleolake Gobero in the Sahara and a small amount of terminal Pleistocene vertebrate remains (Sereno et al., 2008). The terminal Pleistocene faunas from Gobero accumulated during an arid interval prior to the Holocene expansion of the paleolake Gobero and are associated with transient hunter-gatherers (Sereno et al., 2008). In Nigeria, the site of Ihò Eleru yielded a human calvarium dated to the very end of the Pleistocene (Stojanowski, 2014). Deposits at the site also contained fish otoliths, mollusks, and animal bones (Allsworth-Jones et al., 2010). Situations such as these strongly indicate that even forested regions of Africa should not be overlooked based on the allegedly low chance of bone preservation. Similarly, excavations in the Democratic Republic of Congo (DRC) and Cameroon have mainly been in deposits that extend only back to the Early Holocene or do not preserve faunal remains in their Pleistocene-aged deposits (de Maret et al., 1987; Mercader & Brooks, 2001; Robert et al., 2003; Van Neer, 1990). A few important exceptions described below occur in the DRC, which show that Pleistocene faunas associated with artifacts do occur in this region in both open-air and shelter contexts. Most of these are clustered near the Semliki River, in the northwestern part of the country. Ishango 11 and 14 in the DRC contain large Pleistocene assemblages, including human remains at Ishango 11 (Crevecoeur et al., 2016; Peters, 1990). The nearby locality of Katanda preserves not only a rare Pleistocene faunal assemblage but also osseous harpoons that date to ~95 ka (Yellen et al., 1995). Nearby, the locality of Senga 5A has an Early Pleistocene fauna associated with stone tools and a single controversial marked bone. The mark has been interpreted as a butchery mark, but the assemblage has a complex taphonomic history and generally poorly preserved bone surfaces (Tappen & Harris, 1995). Matupi Cave is an important cave site in the DRC that has a large and well-described faunal assemblage, but little is known about its archaeological associations (Van Neer, 1984; Van Noten, 1977). The lone reported site in the southern part of the DRC is the breccia deposit at Kakontwe, which appears to contain a Late Pleistocene assemblage (dated through biochronology) in association with MSA tools (Cooke, 1957). There is a single hominin tooth in the breccia, but the whole fossil assemblage has only been cursorily described. 2028 J. C. Thompson et al. The Horn The majority of sites in Africa with Pleistocene archaeofaunal assemblages occur in the depositional basins of the East African Rift System (EARS), which has a long history of archaeological, paleontological, and geological research. The Horn of Africa, positioned at the northern end of the Rift, contains the modern countries of Djibouti, Eritrea, Ethiopia, Somalia, and the self-declared country of Somaliland (Fig. 6). With 62 assemblages and spanning 1,954,156 km2, the Horn has approximately 3.2 Pleistocene faunal assemblages for every 100,000 km2. It is dominated Fig. 6 Map of sites with Pleistocene archaeofaunas in the Horn of Africa. Grayscale is topography, with lighter colors being higher elevation. Numbers reference sites in Table 1. White dots with black centers indicate other sites in the neighboring regions, numbered in each of their respective sections The Zooarchaeology of Pleistocene Africa 2029 by the Somali–Masai Bushland Biotic Zone but also features high-elevation Afromontane–Afroalpine Biotic Zone flora and fauna along the highland shoulder of the EARS (Happold & Lock, 2013). Although Pleistocene deposits have been surveyed in Djibouti, there are few published archaeofaunal records (Gutherz et al., 2015). A key exception is the locality of Barogali, an Early Pleistocene site with the remains of a single elephant associated with stone tools (Berthelet & Chavaillon, 2001). Eritrea contains the important Middle Pleistocene locality of Buia, but faunas and artifacts are mainly surfacecollected rather than excavated (Delfino et al., 2018). A behavioral association between lithics and faunal remains derives from taphonomic analyses that report butchery marks from at least one site (Fiore et al., 2004), but the materials were not excavated in situ, so it does not appear in Table 1 or on the associated maps. Archaeological research has been sporadic in Somalia and Somaliland, which contain a number of important Late Pleistocene rock shelters. Excavations conducted in the 1980s resulted in large faunal assemblages from southern Somalia, but the region has since become too politically unstable for continued work. The export of the Rifle Range and Gogoshiis Qabe collections before this happened, however, has resulted in recent analyses of their faunal assemblages (Jones et al., 2018; Reid et al., 2019). This work shows that at the end of the Late Pleistocene and into the Holocene, human hunting of small antelope near inselbergs may suggest increasingly localized or intensive use of animal resources. In Somaliland, to the north, there have been recent new excavations, and, although faunal remains are preserved at sites such as Laas Geel 7, they are not reported in detail (Gutherz et al., 2014). In Ethiopia, research began in the Lower Omo River Valley in the 1930s by the French vertebrate paleontologist Camille Arambourg. This was followed by largerscale work in the 1960s and 1970s by the International Omo Research Expedition (IORE) led by F. Clark Howell and Yves Coppens; research in this area was renewed starting in 2006 by the Omo Research Group Expedition (ORGE) led by JeanRenaud Boisserie et al. (2007). Although fauna is abundant in the Omo Shungura Formation, most faunal and lithic assemblages within it are not spatially associated; the only locality with associated fauna and lithics (Omo A42/Omo 79) does not have definitive evidence of hominin modification of the fauna (Maurin et al., 2017). Other localities in the Lower Omo Valley with Pleistocene fauna include Konso (Beyene et al., 2015) and FT-1 at Fejej (Asfaw et al., 1991). Clear behavioral association between the fauna and lithics is evidenced at Konso (KHA6-A1 and KGA4-A3), including an excavated butchered bone (Echassoux, 2012) and a surface-collected bone handaxe from KGA13-A1 (Sano et al., 2020), but a behavioral link between the fauna and lithics at Fejej has not been demonstrated (Barsky et al., 2011). The Omo Kibish Formation also contains numerous Middle Pleistocene archaeological sites in association with fauna as well as some of the earliest fossils of Homo sapiens, but faunal remains are poorly preserved in excavated contexts (Sisk & Shea, 2008). The majority of Plio-Pleistocene sites in Ethiopia come from farther north, in the Awash Valley. The Lower Awash region of the Afar is best known for its Pliocene fossil hominins, such as Australopithecus spp. (Johanson, 2004). However, it also 2030 J. C. Thompson et al. contains several early sites with stone artifacts and the majority of archaeofaunas older than ~2 Ma that meet our criteria. Bokol Dora 1 from the Ledi-Geraru research area has associated artifacts and fauna dating to at least 2.6 Ma but with faunal surfaces too poorly preserved to identify potential butchery marks (Braun et al., 2019). Eight excavated sites at Gona, dating to ~2.6 Ma, have associated artifacts and fauna, and seven of these have reported butchery marks on fragmented bones, mainly from ungulates; a ninth site has only butchery-marked bones (DomínguezRodrigo et al., 2005). Some Pleistocene sites, such as OGN3 and OGS3 at Gona, have reported butchery marks on fauna but not from excavated contexts (Cáceres et al., 2015b). The DIK-55 site at Dikika, Ethiopia, also contains fossils with reported butchery marks from surface collections, but, at 3.4 Ma, these are Pliocene in age (McPherron et al., 2010). Although the adjacent Hadar research area is best known for its Pliocene hominins, two Pleistocene archaeological sites at A.L. 666 and A.L. 894 have small faunal assemblages associated with lithics at ~2.4 Ma but with no reported butchery marks (Kimbel et al., 1996). The Afar, in general, also has strong potential for recovery of Middle Pleistocene archaeofaunas, for example, at Asbole (Geraads et al., 2004) and Busidima–Telalak (Alemseged & Geraads, 2000). However, although these localities have artifacts and fossils, they have not yet produced excavated assemblages with the two kinds of materials in direct association. The Middle Awash, with a stratigraphic succession that spans the Pleistocene, has been a key region for the recovery of archaeofaunas of many ages. Initial work in the 1970s led by Maurice Taieb resulted in the discovery of the region’s paleontological and archaeological potential, including the recovery of a hominin skull at Bodo, now commonly assigned to Homo heidelbergensis (Conroy et al., 1978). J. Desmond Clark initiated a major paleoanthropological project there in 1981 (Clark et al., 1984), which later continued under the direction of Tim White (White et al., 2003). Although this region has numerous archaeological sites, many with associated faunal remains, the latter are usually reported in a single faunal list that also includes surface-collected remains that are likely paleontological. For example, in the Middle Pleistocene Daka Member in the Bouri “peninsula,” “…archaeological sites are abundant. Bone modifications characteristic of the butchery of large mammals by hominids scar several equid, bovid and hippo postcrania” (Asfaw et al., 2002: 317). In the earlier deposits at Bouri, the majority of excavations have focused on localities with hominin remains (Clark et al., 2003). These include sites that date to ~2.4 Ma with potential butchery-marked bones but no directly associated lithics (Sahle et al., 2017). Other rich localities in the Middle Awash have reported faunal lists and Pleistocene archaeological materials, for example, at the Middle Pleistocene locality of Andalee. However, there are no formal excavation reports that allow individual assemblages of artifacts to be tied to individual assemblages of faunal remains (Kalb et al., 1982). A number of Middle Pleistocene sites at nearby Aduma have been excavated with a primarily archaeological focus, but the majority of the fauna is heavily weathered or encrusted with matrix; only the site A8A has “unquestionable” cut marks on hippopotamus and crocodile bones (Yellen et al., 2005: 43). The Zooarchaeology of Pleistocene Africa 2031 The Upper Awash Valley of Ethiopia, Melka Kunture, first surveyed in 1963 by Jean Chavaillon, has produced a number of important faunal and lithic assemblages in direct association, which range in age from ~1.7 to 0.2 Ma (Piperno et al., 2009). Fauna is preserved in association with stone artifacts but is not individually reported for each site (Geraads, 2004). In fact, many of the key Middle and Late Pleistocene archaeological sites in Ethiopia have sporadic faunal preservation without large assemblages in direct association with excavated archaeological sites, for example, at most of the earlier Gademotta MSA sites in the Ziway–Shala Basin (Smith et al., 2019) and at Gadeb to the east (de la Torre, 2011). Some well-known localities, such as Lake Besaka and Aladi Springs, date to the terminal Pleistocene and have faunal remains in archaeological contexts (Clark & Williams, 1978). However, details of the fauna, exact provenience, and archaeological associations are not reported, so it is not clear which excavations that penetrated into the Pleistocene deposits contain bone. Shinfa-Matema 1, a Late Pleistocene open-air site near the border of Sudan and a rare occurrence outside of the EARS proper, has produced excavated assemblages of mammals, birds, reptiles, fish, and amphibians, as of 2020 only described in a dissertation (Davis, 2019). This summary shows that open-air sites in the Horn often suffer from poor association between specific assemblages of fossil bones and stone artifacts in an excavated context. The spatial association is usually clearer in cave and rock shelter (closed) sites, but this does not equal a behavioral association between the two because many non-human agents can also accumulate or modify bones at archaeological sites (Brain, 1981). Even with these caveats, only 7% (N=4) of assemblages in our compilation from Ethiopia are from closed sites. The French prehistorians Henri de Montfried and Père Teilhard de Chardin produced the first Pleistocene faunal assemblages from closed sites in Ethiopia. The work that they began in 1929 at Porc-Epic Cave, near Dire Dawa in northern Ethiopia, resulted in a large and well-preserved assemblage from MSA and LSA deposits (Assefa, 2006). The site was subsequently excavated several more times, including by Clark, who had first excavated a rock shelter site in the southern part of the country at Yavello in 1942 (Clark, 1945). Another early work, such as the one at Gorgora Rock Shelter near Lake Tana, produced stratified LSA and MSA deposits, but there is no mention of bone preservation (Leakey, 1943). More recently, excavations at Mochena Borago, southwestern Ethiopia, have produced large artifactual assemblages and at least 1500 faunal specimens, with MSA deposits dating between ~50 ka and 70 ka (Brandt et al., 2012). However, bone from Pleistocene layers is described as being poorly preserved. Goda Buticha, near Porc-Epic Cave, is the most recently reported cave site with Pleistocene archaeofauna. Excavations in 2008 and 2011 yielded large mammal (including human) and micromammal remains (Stoetzel et al., 2018) as well as engraved ostrich eggshells (Assefa et al., 2018). Goda Buticha was discovered during a formal survey and test-pitting program, which shows the potential for future work at closed Ethiopian sites (Assefa et al., 2014). 2032 J. C. Thompson et al. Eastern Africa Here, we define eastern Africa as comprising the modern-day countries of Burundi, Kenya, Rwanda, Tanzania, and Uganda (Fig. 7). Zooarchaeological work has an uneven history across these countries, and the majority of well-studied assemblages Fig. 7 Map of sites with Pleistocene archaeofaunas in eastern Africa. Grayscale is topography, with lighter colors being higher elevation. Numbers reference sites in Table 1. White dots with black centers indicate other sites in the neighboring regions, numbered in each of their respective sections The Zooarchaeology of Pleistocene Africa 2033 derive from Kenya and Tanzania. With 123 assemblages and spanning 1,832,311 km2, eastern Africa has approximately 6.7 Pleistocene faunal assemblages for every 100,000 km2 – more than twice as many as the next highest concentrations in the Horn and southern Africa. The reported archaeofaunal assemblages from Rwanda and Burundi are limited to the Holocene (Van Neer, 1990). Uganda has two Pleistocene assemblages, one of which is the Late Pleistocene site of Munyama Cave on Buvuma Island (Van Noten, 1971). This site appears to have substantial significance for the region, as its deposits date back to at least ~20 ka. However, the presence of a faunal assemblage is only implied by the organic preservation that facilitated the 14C dates. There is also an important artifactual and paleontological Middle Pleistocene locality in southern Uganda at Nsongezi, first investigated in the 1930s by Terence Patrick O’Brien. The co-occurrence of both lithics and fauna in an excavated context is not clear, although they are potentially associated at the Nyabusora facies Mile 8.5 trench (Cole, 1967). The earliest archaeological work undertaken in eastern Africa began with Louis and Mary Leakey’s first expedition to Olduvai Gorge (known also by its local name Oldupai), Tanzania. Between 1931 and 1947, they conducted a “preliminary survey” to establish “a sequence of evolutionary stages of culture within the geological horizons exposed in the Gorge” (Leakey, 1965: ix), document the geological history of Olduvai, and locate as many sites as possible. They then undertook a second phase of work at Olduvai beginning in 1949 and progressing into more intensive excavations by 1951. Mary Leakeys and her team’s discoveries of diagnostic hominin fossils at Olduvai beginning in 1959 established eastern Africa as an important area for human evolutionary research and led to the first application of radiometric dating to the hominin fossil record (Leakey et al., 1961). The bulk of the Leakeys’ excavations in Beds I and II, dating between about 1.85 and 1.34 Ma (Pante et al., 2020), were carried out between 1960 and 1963, but their work there continued through the mid-1980s. Since then, various teams have undertaken research at Olduvai Gorge. Because of the large numbers of fossils and often excellent bone surface preservation, particularly in the assemblage from level 22 (the “Zinj” level) at the site of FLK with over 40,000 fossils, Olduvai Gorge dominates the zooarchaeological record of the Early Pleistocene of eastern Africa (e.g., Domínguez-Rodrigo et al., 2007b; Parkinson 2018; Potts 1988). Olduvai Gorge also holds an important place in the history of African zooarchaeological research because of the initial identification of butchery marks on Early Pleistocene fossils from FLK “Zinj” in 1981 (Bunn, 1981; Potts & Shipman, 1981). Although Olduvai Gorge is best known for its Early Pleistocene deposits and Oldowan stone tools, good bone preservation is also attested in association with Acheulean artifacts in its Middle Pleistocene Beds III and IV, including bone tools (Pante et al., 2020). Bone is also well-preserved in association with MSA artifacts in the overlying Ndutu Beds and LSA artifacts in the Late Pleistocene Naisiusiu Beds (Leakey et al., 1972). Although the Leakeys excavated at least one site with fauna from the Naisiusiu Beds (Leakey et al., 1972), the majority of sites postdating c. 500 ka have been only surveyed for surface collections, with an emphasis on stone tools (Eren et al., 2014; Mabulla, 1990). 2034 J. C. Thompson et al. Beginning in 1969 but in earnest in 1970, at the invitation of Richard Leakey, Glynn Isaac directed archaeological research at Koobi Fora, on the east side of Lake Turkana in northern Kenya (Clark, 1997). Faunal studies there were undertaken by Henry Bunn (1997). Several sites with fauna and ESA lithics, or butchery-marked fossils without lithics, were excavated and studied in the late 1990s and 2000s, under the auspices of the Koobi Fora Field School (Braun et al., 2010; Pobiner et al., 2008). Two other sites on the west side of Lake Turkana preserve early fauna and lithics together: Lokalalei 1 – with fauna that have some possible but not definitive bone surface modifications – and Lokalalei 2C – with fauna with poorly preserved surfaces (Kibunjia, 1994). Kanjera South in western Kenya has the earliest large, well-preserved faunal assemblage with numerous butchered carcasses clearly associated with ESA lithics, mainly from small-sized ungulates (Ferraro et al., 2013). At sites like GaJi5 and FwJj70 at Koobi Fora, rare hominin-butchered fauna been found, but only on the surface. These sites are not included in Table 1 because they have no in situ fauna from an excavated context (Bunn, 1997; Merritt et al., 2018). However, there is much potential for new localities in regions that have previously received only cursory attention. For example, a new sequence spanning the Early to Late Pleistocene at Kilombe Mountain includes three excavated sites with faunal remains in association with Oldowan tools (Hoare et al., 2021). The majority of Middle and Late Pleistocene sites with fauna in Kenya are located in the southwestern part of the country, which also features Lakes Baringo, Victoria, and Naivasha. In Tanzania, most sites cluster in the north-central area in or near Olduvai Gorge, Lake Eyasi, Lake Manyara, and the Serengeti (Fig. 7). As was the case for Early Pleistocene sites near Lake Turkana, sites in Tanzania with fauna and artifacts are in many localities, but it is rare to find them in excavated contexts. For example, in spite of much investigation by a team led by Mary Leakey in 1966, only a few sites revealed faunal remains in secured contexts at the important Middle Pleistocene localities in the Kapthurin Formation near Baringo (Cornelissen, 1995). Near Lake Victoria, the excavated site of Aringo 3 yielded MSA artifacts but faunal remains were recovered only from the surrounding outcrops and therefore provide mainly paleoenvironmental information (Faith et al., 2015). At the nearby Bovid Hill, the dense accumulation of the extinct wildebeest-like antelope Rusingoryx atopocranion in association with a few lithics and bones with butchery marks has suggested the possibility of tactical hunting during the MSA (Jenkins et al., 2017). At the Late Pleistocene locality of Kibogo in Kenya, surface artifacts and fossils erode from volcanic tuffs that constrain the age between 36 ka and 12 ka, but no sites have been excavated (Faith et al., 2020). At the famous Acheulean locality of Isimila in Tanzania, only one geological layer is fossiliferous, and most excavations did not produce large assemblages (Howell et al., 1962). At other localities, large assemblages have likely been produced and are associated with well-studied artifacts, but there are almost no reported faunal data. For example, at Ntumot and Ntuka River in Kenya, Pleistocene 14C dates imply that there is unreported fauna preserved there (Ambrose, 2002). At Marmonet Drift and Prolonged Drift, also in Kenya, the presence of faunal remains in MSA layers is only mentioned briefly in dissertations (Merrick, 1975; Slater, 2016). The Zooarchaeology of Pleistocene Africa 2035 Lake Turkana also has an important Late Pleistocene record. Many localities with MSA artifacts also have associated faunal remains on the surface, but, again, they rarely occur together in excavated contexts (Shea & Hildebrand, 2010). There is a wealth of Early Holocene sites near the lake containing aquatic and terrestrial faunal records, especially on the eastern side, and some have date probability distributions that extend into the terminal segment of the Late Pleistocene (Beyin et al., 2017; Prendergast & Beyin, 2018). Lothagam and Lowasera are the best known of the Holocene localities, each comprising multiple sites with excellent faunal preservation and rich assemblages of terrestrial and aquatic taxa in association with barbed bone points (Stewart, 1989). On the western side of the lake, recent investigations have begun to push this chronology more firmly into the Late Pleistocene at sites such as Kokito 01 and 02 and Dilit 01 (Prendergast & Beyin, 2018). Overall, the highest concentration of open-air archaeofaunal assemblages in Kenya is at Lake Turkana, which has produced 26 of 51 (51%) open-air assemblages in our compilation. The Tanzanian record is dominated by Olduvai Gorge, which accounts for 42 of 55 (76%). Closed sites in eastern Africa (five from Kenya and eight from Tanzania) are dominated by rock shelters found near inselbergs or under overhangs of metamorphic rock that produce acidic sediments not always conducive to bone preservation. Lukenya Hill is a locality in Kenya with four of the five rock shelter sites and has undergone substantial investigation of both lithics and fauna (Marean, 1992b; Tryon et al., 2015). The other rock shelter with wellpublished Pleistocene fauna is Enkapune Ya Muto (Marean et al., 1994), while the only excavated Pleistocene cave site in Kenya is Panga ya Saidi, on the coast (Shipton et al., 2018). This site has produced faunal evidence of shifting exploitation of closed-habitat fauna and a bone tool industry dating to at least ~20 ka (d’Errico et al., 2020). Excavated rock shelter sequences in Tanzania are less clustered, although there is a concentration of sites investigated in the 1930s by Ludwig and Margit KohlLarsen in the Eyasi Basin – for example, Ostrich Cave and Njarasa Cave – that have been largely published in German monographs, with new investigations of fauna mentioned only from Njarasa Cave (Bader et al., 2020). The best known of these sites, Mumba, was further investigated in the 1970s (Mehlman, 1979), and its fauna has subsequently been reinvestigated, re-dated, and new analyses conducted on the existing materials (Prendergast et al., 2007). Similarly, there are new investigations at Kisese II (Tryon et al., 2018). The only published rock shelter sites in the southern part of the country are Magubike and Mlambalasi, and these have produced both human and faunal remains (Collins & Willoughby, 2010). As in Kenya, there is only a single cave site in Tanzania with Late Pleistocene fauna and it is on the coast: Kuumbi Cave on the island of Zanzibar (Prendergast et al., 2016). This assemblage documents the impacts of rising sea levels and island formation on local faunal communities, resulting in the eventual extirpation of many of the large mammals present in the Pleistocene. 2036 J. C. Thompson et al. Southern-Central Africa Southern-central Africa is a region that today includes Angola, Malawi, Mozambique, Zambia, and Zimbabwe (Fig. 8). Much of this part of Africa falls within the Zambezian Woodland Biotic Zone, which is today a wide band of open woodland separating the savanna ecosystems in eastern and southern Africa (Happold & Lock, 2013). This open woodland ecozone would have provided Pleistocene hominins with diverse resource options during shifting climatic conditions, but there is much less understood about its evolution than about the evolution of grasslands in adjacent regions. With 20 assemblages and spanning 3,534,149 km2, southern-central Africa has approximately 0.6 Pleistocene faunal assemblages for every 100,000 km2. There remain today important temporal and geographical gaps in the distribution of published Pleistocene archaeofaunas from southern-central Africa. Much of the early research is essentially paleontological. Zooarchaeological investigations explicitly addressing human behavior began in the late 1970s at Redcliff Cave, after its excavations in 1964–1965 by Charles Kimberlin “Bob” Brain and Cranmer Fig. 8 Map of sites with Pleistocene archaeofaunas in southern-central Africa. Grayscale is topography, with lighter colors being higher elevation. Numbers reference sites in Table 1. White dots with black centers indicate other sites in the neighboring regions, numbered in each of their respective sections The Zooarchaeology of Pleistocene Africa 2037 Kenrick (Klein, 1978b). The only fossils associated with ESA stone tools are reported from Dungo V in Angola and consist of cetacean remains that might represent the oldest case of scavenging of a marine mammal (Gutierrez et al., 2001). However, no taphonomic evidence is published in support of this interpretation. Remarkably, faunal remains are not preserved in the ESA waterlogged deposits at sites A and B of Kalambo Falls in Zambia, despite the occurrence of fossil wood (Clark, 2001). Dungo V, Kabwe (Cooke, 1963), Twin Rivers Kopje (Bishop & Reynolds, 2000), and Chelmer Spruit (Cooke & Wells, 1951) are the only sites with excavated archaeofauna of the Middle Pleistocene age, while most of the available material from southern-central Africa dates to the Late Pleistocene. Five closed sites with Late Pleistocene archaeofaunas are reported in Mozambique. Ngalue Cave (Mercader et al., 2009) and Chicaza Rockshelter (Bicho et al., 2016), both in the northern Niassa region, are the best reported of these. In Sofala, Nhamababwa and Nhamissimbiti both have associated MSA lithics, and deposits at Nhamissimbiti are preliminarily dated to the earlier part of the Late Pleistocene (Mercader & Sillé, 2013). However, no fauna from these sites has yet been published in detail. Caimane Cave, near Maputo, has a faunal assemblage in association with MSA, LSA, and potentially also ESA artifacts (Bicho et al., 2018). Although open-air sites in Malawi have been reported with both artifacts and fauna, such as the Late Pleistocene Mwanganda’s Village “elephant butchery” site, and various localities in the Early Pleistocene Chiwondo Beds, these associations were the result of geological processes rather than human behavior (Juwayeyi & Betzler, 1995; Wright et al., 2014). Renewed excavations in rock shelter sites in northern Malawi have yielded a new site with the Pleistocene archaeofauna (Mazinga 1) and pushed the chronology of Hora 1 (originally excavated by J. Desmond Clark in 1950) into the Late Pleistocene (Miller et al., 2021). In the southern part of the country, the site of Malowa has been reported as having MSA artifacts (Denbow, 1973), but none of the layers date to the Pleistocene (Juwayeyi, 2011). Available studies in the region focus almost entirely on taxonomy, which highlight that the faunal communities of the eastern Zambezian Woodland Biotic Zone include both eastern and southern elements that probably entered the area during interstadial and stadial phases, respectively (cf. Bromage et al., 1995). Taphonomic studies of Pleistocene assemblages, in contrast, are extremely rare. In some cases, only taxonomically identifiable elements (mainly teeth and horn cores) have been reported and inferences about human behavior are uncommon. Body part representation data have been published for Mumbwa (Barham, 2000), Leopard’s Hill (Klein, 1984), Cave of Bees, Nswatugi, Pomongwe (Walker, 1995), and Redcliff (Cruz-Uribe, 1983). Selective transport of large bovids has been proposed at Leopard’s Hill (Klein, 1984), but different skeletal elements are tallied in an aggregate form, and this does not allow tests of correlation between element representation, nutritional utility, and body size (cf. Marean & Cleghorn, 2003). Bone surface modifications, when observed, are usually reported anecdotally. Quantitative assessments of BSMs are available only for Mumbwa and a portion of the Twin Rivers fauna. At Mumbwa, a single observed cut mark occurs in the Holocene deposits, while the only indication of human activities recorded on the 2038 J. C. Thompson et al. Pleistocene bones are a few instances of burning (Klein & Cruz-Uribe, 2000a). At Twin Rivers, no anthropogenic BSMs were recognized (Bishop & Reynolds, 2000). This leaves open the possibility that some of these fossil assemblages were entirely or partially accumulated through non-human agency. For instance, both Redcliff and Leopard’s Hill include abundant suids, including neonates, which may have been accumulated by carnivores (Cruz-Uribe, 1983; Klein, 1984). At Kabwe, the famous hominin skull has been directly dated to ~300,000 BP, but the large mammals are probably much older (Grün et al., 2020). The absence of taphonomic data on the fossil fauna prevents any conclusion about the relationship between hominin and other mammal remains at Kabwe. Therefore, generating high-resolution taphonomic analyses of new or existing fossil faunal collections from southern-central Africa should be a research priority. Southern Africa The southern African region comprises the modern-day countries of Botswana, Eswatini, Lesotho, Namibia, and South Africa (Fig. 9). With 89 assemblages and spanning 3,298,021 km2, southern Africa has approximately 2.7 Pleistocene faunal assemblages for every 100,000 km2. Almost all of these are in the modern country of South Africa. There are no reported assemblages from Eswatini and only one – Sehonghong Rockshelter – from Lesotho that preserves fauna from Pleistocene deposits (Stewart et al., 2012). In Botswana, fossiliferous cave deposits at Gcwihaba and Nqumtsa in Western Ngamiland have only one reported Pleistocene archaeofauna, from Drotsky’s Cave (Gcwihaba) (Pickford, 1990). The open-air site of ≠Gi, also in Ngamiland, is a rare case of preserved fauna at an open MSA site in southern Africa (Brooks, 1984; Brooks & Yellen, 1977b). Although fossils and MSA artifacts were found together on the surface at Kudiakam Pan, test excavations did not recover any materials in situ (Robbins, 1989). Similarly, Ngxaishini Pan is described as “the only documented site with Acheulian tools and faunal remains in Botswana,” but no archaeological excavations are reported with these materials (Robbins & Murphy, 1998: 55). Excavated sites with MSA deposits, such as Rhino Cave (Robbins et al., 1996a) and Depression Shelter (Robbins, 1990) in the Tsodilo Hills, only preserve bone in the Holocene. White Paintings Shelter, also in the Tsodilo Hills, contains Pleistocene fauna; however, like all other assemblages from Botswana of this age, no detailed taphonomic work on this fauna has been reported to date. The most well-dated site with Pleistocene fauna in Namibia is Apollo 11, which extends in age back to at least ~71 ka (Vogelsang et al., 2010). The faunal assemblage is large, and it remains the only Pleistocene archaeofauna in Namibia for which there has been any reported isotopic work (Vogel, 1983). Although it has been described taxonomically by Thackeray (1979), taphonomic work is limited to only a general description. There is also a large MSA-associated assemblage at Zebrarivier (Avery, 1983; Cruz-Uribe & Klein, 1983), and bones are reported from The Zooarchaeology of Pleistocene Africa 2039 Fig. 9 Map of sites with Pleistocene archaeofaunas in southern Africa. Grayscale is topography, with lighter colors being higher elevation. Numbers reference sites in Table 1. White dots with black centers indicate other sites in the neighboring regions, numbered in each of their respective sections undated MSA deposits at the open site of Oshilongo near the rock art-rich region of Brandberg Mountain (Van Neer & Breunig, 1999). Several sites in Namibia preserve bones that may be from undated Pleistocene LSA layers but do not preserve deeper MSA layers and so cannot be given a Pleistocene age based on typology. For example, there are faunal remains from MSA “contact” deposits from Bremen 1C (Vogelsang, 1998), described by CruzUribe and Klein (1983). At Etemba 14, Schmidt (2011: 88) describes the bone sample as “small and heavily fragmented” and not preserved in the primary MSA layer. Faunal and human remains are preserved in the layer immediately above this (CruzUribe & Klein, 1983), implying a Pleistocene age, but the LSA materials therein are not dated. Two other shelters near the Erongo Mountains, Cymot and Davib-Ost, have reported MSA artifacts and fauna, but it is unclear whether they are in association (Schmidt, 2011). Sandelowsky and Viereck (1969) refer to poorly preserved organics in lower layers at both sites. At sites such as Erb Tanks, faunal preservation in the lowest layers is poor, but there are a few bone fragments recovered from LSA deposits with chronologies that include the latest Pleistocene (McCall et al., 2011). Some rock shelters, such as Oruwanje 95/1 (Albrecht et al., 2001), Ovizorombuku 2040 J. C. Thompson et al. 98/6, and Omungunda 99/1 (Vogelsang & Eichhorn, 2011), contain deposits with fauna that just barely extend into the Late Pleistocene. At other sites, the presence of Pleistocene archaeofaunas is likely but ambiguously reported. Thackeray (1979) indicates that the fauna from several Namibian sites (Fackelträger, Haalenberg, Tiras 5, Aar I, Aar II, and Nos), originally excavated by Wendt (1972), do not preserve fauna in MSA layers. However, Plug and Clark (2008) refer to avian remains from MSA deposits at both Fackelträger and Haalenberg. Confirmation of Pleistocene fauna at Haalenberg was later provided by Vogel and Visser (1981), who report that MSA deposits dated to ~40 ka from a spit that produced ostrich eggshell, tortoise, and bovid remains (Thackeray, 1979). Vogel and Visser (1981) also report an age of ~22 ka from Nos based on ostrich eggshells. Vogelsang (1998) describes a very small amount of MSA material from Tiras 5, but because Thackeray (1979) does not separate the faunal assemblage by layer, it is not possible to know whether any of it derives from those deposits. Vogelsang (1998) emphasizes that there is no fauna preserved in the MSA deposits at Aar I (Schlangengrotte) and Aar II, although the possibility remains that there is terminal Pleistocene fauna associated with LSA deposits at one or both sites. South Africa has a long history of archaeological research including the formal designation in 1929 by John Goodwin and Clarence van Riet Lowe of the “Earlier,” “Middle,” and “Later” Stone Ages, which is still used today across sub-Saharan Africa (Goodwin & Van Riet Lowe, 1929). However, most of this work has not been on faunal remains until more recently, and there is uneven temporal and spatial detail. In South Africa, research is heavily focused on the Western Cape and areas of the southern and eastern coastline, with patchy research activity across the interior. Lime mining led to the earliest discovery of Australopithecus at the site of Taung, announced by Raymond Dart in 1924, which launched a century of exploration of Pliocene and Early Pleistocene sites in South Africa. This included sites in today’s Cradle of Humankind World Heritage Area, such as Sterkfontein in 1936, followed by discoveries of hominins by Dart, Robert Broom, and John Robinson at Makapansgat Limeworks, Swartkrans, and Kromdraai in the 1940s (Thackeray, 2016). The pioneering faunal analyses at several of these sites were done by Dart, who assumed that hominins were responsible for the patterns of skeletal part representation and breakage of the fauna; he posited an “osteodontokeratic” (bone– tooth–horn) culture to explain the lack of stone tools (Dart, 1949). This early work was followed by C.K. “Bob” Brain’s, who used observations of modern bone modifiers to demonstrate that carnivore activity was a better explanation for these fossil accumulations (Brain, 1981). Other sites yielding hominins and/or rich Early and/or Middle Pleistocene faunal assemblages, but not associated with lithics or evidence of butchery, cluster in or around the Cradle of Humankind karst systems, including Bolt’s Farm Cave System, Gondolin, Haasgat, Malapa, Minnaar’s Cave, Rising Star (Thackeray, 2016), and Motsetse (Berger & Lacruz, 2003). Excavations at Gladysvale have yielded fossil hairs, potentially including hominin hairs, inside Middle Pleistocene hyena The Zooarchaeology of Pleistocene Africa 2041 coprolites (Taru & Backwell, 2013) but only a single handaxe with a tenuous association with the fauna (Hall et al., 2006). However, several sites in the Cradle of Humankind have produced both stone and bone tools in association with Pleistocene fauna: Sterkfontein Member 5, Swartkrans Members 1–3, Drimolen Main Quarry, and Kromdraai B (Stammers et al., 2018). The only Early Pleistocene archaeofaunas reported from southern Africa outside of the immediate area of the Cradle of Humankind are the Cornelia Beds open locality in the Free State (Brink et al., 2012) and Wonderwerk Cave in the Northern Cape (Brink et al., 2016). Outside the Cradle of Humankind, the majority of foundational archaeological investigations did not translate into large archaeofaunal assemblages until Ronald Singer and John Wymer’s excavations at the Klasies River Mouth complex, initially studied by Richard Klein (1976). Work in the 1950s through the 1970s also resulted in the recognition of important Early, Middle, and Late Pleistocene deposits in the dune fields of the Western Cape that were later investigated: Elandsfontein (Klein et al., 2007), Geelbek (Fuchs et al., 2008), and Duinefontein (Klein, Avery, et al., 1999). As with the Early Pleistocene, faunal material from the Middle Pleistocene of southern Africa has been exclusively described from the country of South Africa. A majority of these sites are open-air, two are cave sites, and one is a sinkhole. Fewer than half of these have been subjected to taphonomic study, although most researchers have published general observations of the assemblages. Most of the discussions have revolved around the taxonomy of extinct and extant species and reconstructions of the environment, although some research has addressed hominin behaviors. Florisbad and Vlakkraal (Orange Free State), within 10 km of each other, have overall similar taxonomic diversity, although there is the presence of different extinct taxa at each site (Wells et al., 1942). Vlakkraal has not been chronometrically dated, and the presence of MSA stone tool forms (the “Maselspoort” culture) has been argued to link the two sites (Wells et al., 1942). Taxonomic diversity at Bundu Farm (Northern Cape) suggests that grassland environments were widespread (Hutson, 2018; Kiberd, 2006). Hominin BSMs are relatively rare in these sites, suggesting a palimpsest of carnivore behaviors with occasional hominin contributions (Hutson, 2018; Hutson & Cain, 2008; Kandel & Conard, 2012; Klein et al., 1999a, 2007). Paleoenvironmental work has suggested that Florisbad, Duinefontein 2, and Bundu Farm were in close proximity to water, drawing prey and predators alike (Hutson, 2018; Kiberd, 2006; Klein et al., 1999a; Kuman et al., 1999). Hominins may have regularly scavenged from the remains of herbivores (Klein et al., 1999a), although it is not well-understood whether this was in complement to hunting of larger mammals or a dominant subsistence behavior (Hutson, 2018). For example, reanalysis of the contentious Kalkbank (Limpopo) assemblage demonstrates that multiple agents of accumulation contributed to the palimpsest (Hutson & Cain, 2008). A comprehensive taphonomic analysis of the Middle Pleistocene deposits at Pinnacle Point 13B revealed only small proportions of small fauna, marine mammals, and tortoises when compared to later MSA sites on different parts of the coastline (Thompson, 2010). Changing abundances of tortoise remains over time at the nearby site of Blombos Cave support the inference 2042 J. C. Thompson et al. that over the course of the MSA, people exploited diverse resources opportunistically as their abundances changed in connection with local environments (Thompson & Henshilwood, 2014a). A considerably higher number of sites across southern Africa come from Late Pleistocene assemblages, and these have made significant contributions to both our knowledge of past hominin behavior and the development of zooarchaeological methodology. As with other regions, most Late Pleistocene assemblages in southern Africa are from caves or rock shelters (73%, N = 55), but, unlike in other regions, about a third of these are located within 15 km of the modern coastline. These coastal sites have received the most detailed descriptions over the last 20 years, for example (from west to east): Diepkloof (Steele & Klein, 2013), Die Kelders Cave 1 (Klein & Cruz-Uribe, 2000b; Marean et al., 2000), Blombos Cave (Badenhorst et al., 2016; Discamps & Henshilwood, 2015; Reynard et al., 2014; Thompson & Henshilwood, 2011, 2014b), Pinnacle Point (Rector & Reed, 2010; Thompson, 2010), Nelson Bay Cave (Klein, 1972, 1983a), and Klasies River Mouth (Binford, 1984; Klein, 1976; van Pletzen-Vos et al., 2019). Almost all coastal or near-coastal sites are along the western and southern coasts; only Sibhudu (formerly Sibudu) and Waterfall Bluff are located to the east (Clark, 2011, 2017; Collins, 2015). Thus, work on Late Pleistocene archaeofaunas in southern Africa is weighted toward the southern African coast, in particular the South-West Cape (Fynbos) Biotic Zone (Happold & Lock, 2013). Work at some of these Late Pleistocene sites in southern Africa has substantially contributed to long-standing discussions about human behavior and population dynamics. In a seminal study on Byneskranskop and Die Kelders 1 (Western Cape), Klein and Cruz-Uribe (1983) argued that an increase in human population density at the end of the Pleistocene depressed tortoise size, reflected in smaller-sized tortoises in the LSA compared to those in the MSA. Similar results from Ysterfontein 1, Elands Bay Cave, and Blombos Cave also show decreases in shellfish size over time (Avery et al., 2008; Henshilwood et al., 2001b; Klein et al., 1999b, 2004; Klein & Cruz-Uribe, 1987, 2016). This technique has been applied to zooarchaeological assemblages both elsewhere in Africa and at Paleolithic sites in other regions (Steele & Álvarez-Fernández, 2011). More controversially, Late Pleistocene faunal assemblages in southern Africa have also been used to suggest that MSA people were less effective at exploiting resources than LSA people (Klein & Cruz-Uribe, 1996). Klein (1992) has argued that humans in southern Africa were not proficient hunters until the LSA, citing a lack of “dangerous” fauna, such as wild pigs and buffalo. A multisite comparison of species richness failed to demonstrate any difference between MSA and LSA assemblages (Faith, 2008), sparking a debate about the application of statistical methods to faunal data (Faith, 2011a; Weaver et al., 2011). A subsequent study also showed that when controlling for environment and available prey, there is no difference in the capture rates of “dangerous” prey (Dusseldorp, 2010). As with other regions of Africa, the description and publication of faunal material has been inconsistent and unstandardized, which makes comparisons challenging. For example, the fauna from the Robberg deposits at Melkhoutboom (Eastern The Zooarchaeology of Pleistocene Africa 2043 Cape) is only described in the most general terms and without NISP, due to a purportedly small sample size (Klein, 1974), while fauna from the ESA/MSA layers at the Cave of Hearths (Limpopo) are reported as present/absent (Cooke, 1962). Many descriptions include an NISP for taxonomic (and/or element) identification but do not quantify the size of the overall assemblage (e.g., Thackeray, 1979). Issues with dating early excavations have also hampered the comparative analyses and understanding of human subsistence behavior and environmental pulses. This has spurred researchers to revisit key sites to better constrain their chronology, such as at Sehonghong (Lesotho) (Pargeter et al., 2017), Boomplaas (Western Cape) (Pargeter et al., 2018), and Apollo 11 (Namibia) (Vogelsang et al., 2010). Reporting of anthropogenically accumulated smaller fauna has complemented the discussion of larger vertebrates in South Africa more than in other regions. Foragers consumed birds as early as 77 ka at Sibhudu (Val et al., 2016) and occasionally at Diepkloof (Val, 2019). Tortoise remains are frequently recovered from sites across southern Africa and have been studied in detail at two cave sites and an open-air locality to better understand their role in hominin subsistence (Avery et al., 2004; Thompson, 2010; Thompson & Henshilwood, 2014b). Taphonomic work along the southwest African coast has also revealed small mammal exploitation for food (Armstrong, 2016) and skinning of carnivores potentially for use as pelts (Val et al., 2020). As researchers continue to return to early excavations and reanalyze faunal assemblages, as well as recover new ones, the southern African region promises to remain a leader in our understanding of Pleistocene zooarchaeology in Africa. Discussion Geographic Patterns The only geographic region to have no reported Pleistocene archaeofaunas is western Africa. Available data in other regions are concentrated on the depositional basins of the EARS and the Nile valley, in fossil dunes, tufas, and breccias in southern and central Africa, and in caves and rock shelters concentrated on the northern and southwestern coastlines. Depositional context is the biggest factor influencing fossil preservation, which is best in the presence of environments that have rapid burial, shelter from the elements, and neutral or slightly basic sediments (Grupe & Harbeck, 2015). Caves and shelters formed even on acidic rocks can preserve bone if there are high concentrations of bone, ash, and shell that provide a localized chemical buffer, but they tend to collapse over time. This means that Early Pleistocene sites are almost always open-air, near paleolakes and ancient river systems that would have provided the rapid burial conditions necessary for preservation of bone; the few closed-site examples derive from sinkholes and collapsed cave systems with extremely complex depositional histories. The result is that 2044 J. C. Thompson et al. environmental and behavioral information from faunal remains at different points in the Pleistocene is strongly biased toward hominin activities that took place in specific parts of the landscape (Fig. 10). The history of archaeological research in each region and the subsequent research efforts have further shaped the distribution pattern of data. At localities such as Lake Turkana or Olduvai Gorge, decades of research efforts have revealed rich assemblages. We treated each archaeofaunal assemblage as a separate entry in our table, even if it is from a different site or excavation area within the same locality as other assemblages. This places heavier weighting on these places relative to other localities that have rich archaeological and faunal records but where the focus has been on survey (rather than excavation), recovery of hominin fossils (rather than in situ faunal assemblages), or there has only been sporadic work or an emphasis on reporting lithics rather than fauna. However, some patterns are still clear. In the early twentieth century, fossiliferous sites often came to the attention of archaeologists through serendipitous discovery (e.g., mining operations, guano extraction). This was followed by a collection of the most obvious fossils and artifacts, often from surfaces. Where excavations ensued, they were typically large in scale and executed quickly, resulting in coarse chronological and spatial resolutions. Associated documentation of the deposits from these early excavations is variable and unstandardized, and, in some cases, it is unclear what sieve size was used, if at all. For some sites, for example, Jebel Irhoud in Morocco (Hublin et al., 2017), Porc-Epic Cave in Ethiopia (Assefa, 2006), Hora 1 in Malawi (Miller et al., 2021), and Klasies River Main site in South Africa (van Pletzen-Vos et al., 2019), archaeologists have revisited old collections with new analytical approaches or Site Types by Region and Chronology 100% 2 90% 20 80% 7 14 1 40% 3 21 1 1 8 4 5 50% 3 15 33 54 29 37 1 9 5 7 NorthernCentral Cave Horn Rock shelter 2 3 Eastern SouthernCentral Early Late Early Middle Late Central Early Middle Late Middle Late Early Middle Early Early 1 0% Middle 1 5 Late 2 Early 22 Northern 9 Middle 6 Late 3 13 3 Middle 20% 26 Late 1 30% 10% 14 35 70% 60% 1 Southern Open Air Fig. 10 Distribution of sites with Early, Middle, or Late Pleistocene faunal assemblages by region and type. Shows only sites where the sub-epoch is probable or confirmed (not just “possible”). If an assemblage has a chronology that spans two sub-epochs, then it may appear twice The Zooarchaeology of Pleistocene Africa 2045 carried out new excavations with finer recovery. However, lack of standardization in recovery and analytical protocols continues to make direct comparisons difficult. There is also evidence that important Pleistocene archaeofaunal assemblages have been recovered with more detailed approaches but not yet studied. Some reports only mention the preservation of bone or describe only selected specimens that are part of a larger assemblage, such as a hippopotamus bearing cut marks at Bouri Herto BOU-A19B (Clark et al., 2003). Others provide short taxonomic lists with presence/absence only, for example, at the newly excavated Melka Wakena complex in Ethiopia (Hovers et al., 2021). Where these derive from excavations, they likely represent much larger assemblages because elements identified to species typically comprise a small proportion of what is preserved. However, where taxonomic lists are drawn from mixed excavation and surface collections, it is likely that curated assemblages are smaller. In these cases, the most taxonomically identifiable specimens are collected from the surface and can constitute substantial faunal lists, even if the excavated sample found in situ and actually in association with stone artifacts is small. In general, cave and rock shelter sites tend to produce larger numbers of recovered specimens. Thus, a smaller excavation area in these types of sites will produce a relatively larger assemblage in comparison to most open-air sites (Fig. 11). Assemblage Size by Site Type 100% 90% 15 80% 70% 1 46 60 10 60% 86 60 50% 3 40% 9 30% 20% 1 27 13 17 10% 17 25 13 5 100s 10s 0% 100,000s 10,000s 1000s Cave Rock shelter 1s Unknown Open Air Fig. 11 Distribution of sites with different orders of magnitude of reported fragments. Some were approximated, for example, if the statement was “few,” it would be approximated to 10s. Specific cases are noted in the “Comments” column of Table 1. Note that 34 of the closed sites and 46 of the open sites in our sample had completely unknown/unreported assemblage sizes 2046 J. C. Thompson et al. Historical trends in excavation and analysis of archaeofaunal assemblages show that early collecting prior to the 1920s typically ignored the eastern and southern regions. These became prominent localities for Pleistocene fauna after the Leakeys began work at Olduvai Gorge in the 1930s and archaeological investigations became more commonplace in general in southern Africa. Work that began in eastern Africa in the 1960s and in the Horn in the 1970s came to dominate the number of sites to produce faunal remains in association with Pleistocene archaeological evidence. Notably, this was contemporaneous to a slowing down of the work in the southerncentral African region. The production of new archaeofaunal assemblages from northern and southern Africa, in contrast, has continued steadily since the 1930s. Early excavations of sites across Africa favored taxonomically identifiable faunal remains. Taphonomic analyses were sporadic in early reports, typically limited to observations on fragmentation of an assemblage (normally high, especially within closed-site contexts), and, occasionally, the presence/absence of anthropogenic or other modifications. More recently, southern Africa has become a region known for the number of large faunal assemblages that have received careful taphonomic analyses. Of the 409 assemblages in our sample, 156 (~38%) have undergone taphonomic analyses, sufficient to ascertain their basic depositional histories. As a proportion of the total number of assemblages, the northern-central region has benefited the most, but this is because of the efforts of a small number of people working on assemblages with small quantities of bone from Sudan (Osypińska & Osypiński, 2016; Yeshurun, 2018). Table 2 shows the number of assemblages from each region for which such work remains to be done, by assemblage size. Although a substantial number have unknown assemblage sizes, there are many assemblages with only a few 10s or 100s of fossils for which this work has not yet been done – but where the small assemblage size would make it highly efficient to return to them to do so. Table 2 Numbers and proportions of assemblages per region for which detailed taphonomic analyses have not been reported. Total assemblages refer to all assemblages in each region. Headers refer to the sizes of assemblages (orders of magnitude) Region Northern Northerncentral Central 100,000s 10,000s 1000s 100s 10s 0 (0%) 7 (8%) 4 (5%) 12 5 (6%) (14%) 0 (0%) 0 (0%) 0 (0%) 1 (4%) 2 (9%) 0 (0%) 0 (0%) Horn 0 (0%) 2 (3%) Eastern 1 (1%) 2 (2%) Southerncentral Southern 0 (0%) 4 (20%) 9 (10%) 0 (0%) 2 1 (25%) (13%) 5 (8%) 10 (16%) 17 22 (14%) (18%) 2 5 (10%) (25%) 6 (7%) 19 (21%) 1s 0 (0%) 0 (0%) 0 (0%) 0 (0%) 21 2 (34%) (3%) 13 1 (11%) (1%) 1 (5%) 0 (0%) 3 (3%) 0 (0%) Total Unknown assemblages 26 (31%) 84 6 (26%) 23 2 (25%) 8 7 (11%) 62 16 (13%) 123 4 (20%) 20 12 (13%) 89 The Zooarchaeology of Pleistocene Africa 2047 The effects of revisiting sites in our analysis with taphonomic methods could potentially be transformative. Of the 156 assemblages in our sample where detailed taphonomic work has been done, 120 (76.9%) show definitive evidence for hominin butchery. In some cases, the absence of butchery marks likely means that there is no behavioral association between the lithics and fauna, for example, at Lainyamok in Kenya (Potts et al., 1988a). In other cases, poor preservation of surfaces may make it difficult to ever know whether hominins butchered the fauna, for example, at many of the Aduma sites (Yellen et al., 2005) or the earliest Oldowan site at Bokol Dora 1 (Braun et al., 2019), both in Ethiopia. Because these kinds of detailed studies are not available for most sites, assemblages with both fauna and artifacts in the same excavation – including Lainyamok – meet the criteria for inclusion in Table 1. Assemblages can have multiple accumulators, some of which may be hominin, but it is probable that many of the “archaeofaunas” presented here have only some components that are relevant to the question of hominin subsistence behavior. Key Debates Since the early 1980s, both new discoveries and re-analyses of previously excavated Pleistocene African archaeofaunas have sparked persistent debates about theoretical linkages between fossil traces and ancient behavior, taphonomic methods, and interpretation of bone surface marks. These have led to significant advances in zooarchaeological methodology more broadly. Three of the most impactful have been the “hunting–scavenging” debate, debate over whether the zooarchaeological record can be extended into the Pliocene, and the “shaft debate.” The “hunting–scavenging debate” framework traditionally characterized Early Pleistocene hominin foragers as either “marginal scavengers” of only marrow and flesh scraps left from carnivore kills (playing only a minor role in the accumulation of these archaeofaunas) or competent hunters of at least smaller-sized prey and aggressive scavengers of larger-sized prey (and the major accumulators of these archaeofaunas) (Pobiner, 2020 and references therein). Evidence has mostly derived from fragmented and butchery-marked bones from Koobi Fora (Kenya) and Olduvai Gorge (Tanzania), although there has also been a related argument regarding the Late Pleistocene fauna from Klasies River Main Site in South Africa. As one of the largest and best preserved Early Pleistocene archaeofaunas, FLK “Zinj” at Olduvai has played a central role in this debate. Bunn and Kroll (1986) considered the assemblage to have been accumulated by hominin hunters, while Binford (1988) argued that hominins scavenged the remains. This debate, which primarily used skeletal part abundances and the presence/absence of stone tool cut marks, led to subsequent work by Blumenschine (1988, 1995), who developed a way to use hammerstone percussion marks and carnivore tooth marks to understand whether hominins or carnivores first accessed a carcass. This approach is based on the idea that relative proportions of these marks in archaeofaunal assemblages can be compared to proportions of marks in modern experimental assemblages to untangle the 2048 J. C. Thompson et al. timing of carcass access (James & Thompson, 2015). Initial work appeared to demonstrate late access by hominins and was later expanded upon to include other agents and in different orders of access. This set of arguments and subsequent work was summarized by DomínguezRodrigo (2002), who later instigated a new debate about the FLK 22 “Zinj” assemblage. Domínguez-Rodrigo and Barba (2007a) argued that some marks identified as tooth marks were actually those made by fungi and bacteria, which changed the relative proportions of tooth marks and thus suggested that hominins had earlier access than previously inferred. Most recently, Parkinson (2018) has reanalyzed this assemblage and concluded that hominins at FLK 22 had early access to carcasses of smaller prey, acquired through hunting, and possibly larger prey as well, acquired through hunting or aggressive scavenging. Evidence from Kanjera South, various sites at Koobi Fora, and additional sites at Olduvai Gorge all indicate that, in addition to early access to smaller mammals, Early Pleistocene hominins in eastern Africa (and likely throughout the continent), “practiced a sophisticated and flexible strategy of carcass acquisition that would have most likely included active scavenging of medium to large mammal carcasses when they could intimidate and outnumber carnivores, as well as less frequent passive scavenging” (Pobiner, 2020: 12). These hominins also likely incorporated diverse smaller animals into their diets, including aquatic resources (Braun et al., 2010). Debates about how to best identify BSMs have also had implications for the interpretation (and identification) of the earliest archaeological record. Marks on bones from the Pliocene DIK-55 site were described by McPherron et al. (2010) as butchery marks, implying that hominins used stone tools to butcher large mammals much earlier than the oldest Oldowan sites. Identification of the marks as hominininflicted has been contested on the basis of their morphology and context (Domínguez-Rodrigo et al., 2010; Sahle et al., 2017), which has led to major efforts to revise how BSMs are identified. These efforts employ computer automation, morphometric analyses, and new statistical modeling to take a probabilistic approach to identifying the agency behind BSMs (Domínguez-Rodrigo, 2019; Harris et al., 2017; Pante et al., 2017). This methodological debate, begun with African archaeofaunas, has far-reaching implications across zooarchaeology – especially where unexpected claims rest on small samples of modified bones. The “shaft debate” revolves around problems of how to best quantify skeletal parts in archaeofaunal assemblages (Yravedra & Domínguez-Rodrigo, 2009) and highlights how critical it is to include less identifiable long bone shaft fragments in analyses. Realization of the importance of shafts to quantifying skeletal part frequencies represents a major methodological and theoretical development in zooarchaeology, which was rooted in the analysis of African assemblages. Here, the Klasies River Mouth cave complex (Eastern Cape) has played a particularly prominent role. The Klasies Main Site assemblage was initially studied by Klein (1976), but as at Olduvai’s Zinj site (Binford, 1981), Binford’s (1984) reanalysis at Klasies suggested that most of the large mammal remains had been scavenged, rather than hunted – in this case, by MSA people. Turner (1989) then drew attention to the fact that bone fragments had been sorted in the field and the less identifiable specimens were discarded. These less identifiable fragments would have included the shafts of The Zooarchaeology of Pleistocene Africa 2049 long bones. Klein (1989) argued that this had no effect on the analysis and that differences in small and large bovid representations were attributable to transport decisions by ancient foragers. Based on ethnographic and experimental observations, Bartram et al. (1999) argued that excavator bias resulting in discarded long bone shafts at Klasies had indeed influenced the outcome of analyses, by retaining only the most identifiable fragments. Differential fragmentation between larger and smaller animals, potentially from scavenging carnivores, would mean that bones from smaller animals were likely to preserve in a more complete – and therefore more identifiable – state. The same long bone elements from larger animals would be mainly represented by shafts, and these had been discarded at Klasies during its earliest excavations. They argued that this led to the “Klasies pattern,” in which smaller bovids were represented by more complete skeletons and larger bovids were represented mainly by skeletal parts from the heads and feet. Because head and foot elements have the least nutritive value in a skeleton, this could lead to the interpretation that hominins were marginal scavengers of what was left after carnivore kills. Klein et al. (1999b) have argued that carnivore ravaging on this scale must be demonstrated for each site before investing substantial time in identifying shaft fragments. Outram (2001) points to the representation of other skeletal elements from the same anatomical regions as important pieces of information about mode of acquisition. Newly recovered faunal remains from Klasies, which do include shaft fragments, confirm that the “Klasies pattern” was indeed likely caused by discarding these specimens (van Pletzen-Vos et al., 2019). Zooarchaeological work at Klasies and nearby sites along the southern African coast, particularly from assemblages with a more complete recovery, has continued to provide details about the diversity of the ways in which MSA people accumulated and modified faunal remains within complex taphonomic systems (Marean et al., 2000; Milo, 1998; Reynard et al., 2016; Thompson, 2010; Thompson & Henshilwood, 2011). These studies have revealed that MSA hunters did take dangerous animals and that they did transport meaty parts from both large and small faunas. However, as with modern hunter-gatherers, MSA subsistence behavior cannot be characterized by a single strategy across either time or space. These revelations show how the “shaft debate” has had major implications for interpretation of archaeofaunal assemblages that carry across time periods and different depositional contexts. Concluding Remarks The longest zooarchaeological records in the world derive from eastern Africa and the Horn, which have seen significant archaeological research since the 1950s. However, the majority of archaeofaunas are from open-air localities where assemblages tend to be smaller and less commonly show stratified change over time. Both southern and northern Africa also have long records of research, but, here, Pleistocene archaeofaunas are heavily biased toward closed sites. These are more 2050 J. C. Thompson et al. likely to be stratified but can have intimidatingly large assemblages and likely sample the surrounding faunal assemblage in a different way from open-air sites. These regional differences, along with differences in how assemblages have been collected and reported, make it challenging to make direct comparisons over both space and time. Lack of taphonomic reporting and lack of standardization in how data are reported are also serious problems. Many processes can combine homininaccumulated fauna with fauna accumulated or modified by other agents, such as water flow (abiotic) or carnivores (biotic). However, fewer than half of African Pleistocene zooarchaeological assemblages have undergone detailed taphonomic analyses, and when this work has been done, it is often not executed or reported in the same manner. Thus, taphonomic work is of utmost importance in establishing the association between stone tools (or other evidence of anthropogenic activity) with faunal remains, and, without such analyses, it is only possible to use archaeofaunas to make general statements about the environmental context of past hominin behaviors. Across the continent, Pleistocene faunal assemblages have benefitted substantially over the last two decades from changes in excavation techniques and more careful attention to site formation processes. Although there is still much work to do, researchers have now begun to undertake full-scale taphonomic analyses as a more standard part of their work. However, significant differences remain in how archaeofaunas are recovered between regions, types of excavations (open-air or closed site), and even time period(s) of interest. Open-air localities are still primarily investigated by surveys, using largely paleontological approaches to collect surface fauna instead of attempting to recover it in situ as part of an archaeological excavation. This is especially the case in open deposits where important hominin fossils have been found and where understanding hominin–animal interactions at a site scale remains a lower research priority than finding more complete fossils. In spite of these challenges, research on the African Pleistocene record has made significant contributions to the development of zooarchaeological methodology. Its role has been particularly salient in understanding site formation processes and untangling the impacts of multiple accumulators or modifiers of bone. Debates about how to best identify and quantify bone surface modifications and skeletal parts are ongoing and continue to stimulate new approaches to understanding the record. Future work can fruitfully continue to fill important geographic and temporal gaps in the Pleistocene record by reinvestigating previously recovered assemblages, finding new localities, and revisiting known localities with new recovery methods. Acknowledgments Thanks are owed to Amanuel Beyin and David Wright for organizing the original symposium that formed the basis of this volume. We thank Kenechi Ijemere for his contributions to filling out some data in Table 1. Andrew Kandel and Zara Kanaeva of the project “The Role of Culture in Early Expansions of Humans” (ROCEEH) provided an additional list of Pleistocene sites in Africa containing both lithics and fauna, and their coordinates, from the ROCEEH Out of Africa Database (ROAD). A.B. thanks Gary Haynes (University of Nevada), Romala Govender (Iziko Museums of South Africa), Charity Ndlovu (Natural History Museum of The Zooarchaeology of Pleistocene Africa 2051 Zimbabwe), Manuel Gutierrez (Université de Paris I), Daniela de Matos (Universität Tübingen), Mirriam Tawane (Ditsong Museums of South Africa), and Constance Mulenga (Livingstone Museum) for providing precious information and advice on Pleistocene fossil assemblages from the southern-central region. E.Y.H. thanks Eslem Ben Arous (Max Planck Institute for the Science of Human History) for references on the geochronology of Casablanca area sites, Giulio Lucarini (Instituto di Scienze del Patrimonio Culturale, Consiglio Nazionale delle Ricerche, and Università degli Studi di Napoli “L’Orientale”) for references on locality coordinates in Egypt and Libya, and Eleanor Scerri (Max Planck Institute for the Science of Human History) for information on Adrar Bous. B.P. thanks Justin Adams (Monash University), Erella Hovers (The Hebrew University of Jerusalem), Tiphaine Maurin (Université de Bordeaux), Jackson Njau (Indiana University), Sally Reynolds (Bournemouth University), and Michael Rogers (Southern Connecticut State University) for access to references and/or relevant information on Early Pleistocene assemblages from southern and eastern Africa. Jürgen Richter (University of Cologne) and Wim Van Neer (Royal Belgian Institute of Natural Sciences) were most helpful in assisting us with sourcing references. J. Tyler Faith and Teresa Steele provided constructive reviews of the initial draft. References Abbate, E., Albianelli, A., Awad, A., Billi, P., Bruni, P., Delfino, M., et al. (2010). Pleistocene environments and human presence in the middle Atbara valley (Khashm El Girba, Eastern Sudan). Palaeogeography, Palaeoclimatology, Palaeoecology, 292(1), 12–34. Adams, J. W., Rovinsky, D. S., Herries, A. I., & Menter, C. G. (2016). Macromammalian faunas, biochronology and palaeoecology of the early Pleistocene Main Quarry hominin-bearing deposits of the Drimolen Palaeocave System, South Africa. PeerJ, 4, e1941. Albrecht, M., Berke, H., Eichhorn, B., Frank, T., Kuper, R., Prill, S., et al. (2001). Oruwanje 95/1: A late Holocene stratigraphy in northwestern Namibia. Cimbebasia, 17, 1–22. Alemseged, Z., & Geraads, D. (2000). A new Middle Pleistocene fauna from the Busidima-Telalak region of the Afar, Ethiopia. Comptes Rendus de l’Académie des Sciences, 331, 549–556. Allsworth-Jones, P., Harvati, K., & Stringer, C. (2010). The archaeological context of the Iwo Eleru cranium from Nigeria and preliminary results of new morphometric studies. In P. Allsworth-Jones (Ed.), West African archaeology new developments, new perspectives (British Archaeological Reports) (Vol. S2164, pp. 29–42). Archaeopress. Altamura, F., Bennett, M. R., D’Août, K., Gaudzinski-Windheuser, S., Melis, R. T., Reynolds, S. C., et al. (2018). Archaeology and ichnology at Gombore II-2, Melka Kunture, Ethiopia: Everyday life of a mixed-age hominin group 700,000 years ago. Scientific Reports, 8(1), 2815. Amani, F., & Geraads, D. (1993). Le gisement moustérien du Djebel Irhoud, Maroc: précisions sur la faune et la biochronologie, et description d’un nouveau reste humain. Comptes Rendus de l’Académie des Sciences, 316, 847–852. Amara, A. (1977). Le gisement Es-Sayar. Bou Saâda (Algérie). Libyca Alger, 25, 59–71. Ambrose, S. H. (2002). Small things remembered: Origins of early microlithic Industries in SubSaharan Africa. In R. G. Elston & S. L. Kuhn (Eds.), Thinking small: Global perspective on Micolithization. American Anthropological Association. Ames, C. J., Gliganic, L., Cordova, C. E., Boyd, K., Jones, B. G., Maher, L., et al. (2020). Chronostratigraphy, site formation, and Palaeoenvironmental context of late Pleistocene and Holocene occupations at Grassridge Rock Shelter (Eastern Cape, South Africa). Open Quaternary, 6(1), 5. Aouraghe, H. (2000). Les carnivores fossiles d’El Harhoura 1, Temara, Maroc. L’Anthropologie, 104(1), 147–171. Aouraghe, H., & Abbassi, M. (2002). Les rongeurs du site atérien d’El Harhoura 1 (Témara, Maroc). Quaternaire, 13(2), 125–136. 2052 J. C. Thompson et al. Aouraghe, H., Bougariane, B., & Abbassi, M. (2012). Les lagomorphes du Pléistocène supérieur de la grotte d’El Harhoura 1 (Témara, Maroc). Quaternaire. Revue de l’Association française pour l’étude du Quaternaire, 23(2), 3. Arambourg, C. (1979). Vertébrés villafranchiens d’Afrique du Nord (p. 141). Fondation Singer-Polignac. Arambourg, C., & Biberson, P. (1956). The fossil human remains from the Paleolithic site of Sidi Abderrahman (Morocco). American Journal of Physical Anthropology, 14(3), 467–489. Arambourg, C., Boule, M., Vallois, H., & Verneau, R. (1934). Les grottes paléolithiques des Beni Ségoual (Algérie). In Archives de l’Institut de Paléontologie Humaine (Vol. 13). Édition Masson. Armstrong, A. (2016). Small mammal utilization by Middle Stone Age humans at die Kelders cave 1 and pinnacle point site 5-6, Western Cape Province, South Africa. Journal of Human Evolution, 101, 17–44. Asfaw, B., Beyene, Y., Semaw, S., Suwa, G., White, T. D., & WoldeGabriel, G. (1991). Fejej: A new paleoanthropological research area in Ethiopia. Journal of Human Evolution, 21, 137–143. Asfaw, B., Gilbert, W. H., Beyene, Y., Hart, W. K., Renne, P. R., WoldeGabriel, G., et al. (2002). Remains of Homo erectus from Bouri, Middle Awash, Ethiopia. Nature, 416(6878), 317–320. Assefa, Z. (2006). Faunal remains from Porc-Epic: Paleoecological and zooarchaeological investigations from a Middle Stone Age site in southeastern Ethiopia. Journal of Human Evolution, 51, 50–75. Assefa, Z., Pleurdeau, D., Duquesnoy, F., Hovers, E., Pearson, O., Asrat, A., et al. (2014). Survey and explorations of caves in southeastern Ethiopia: Middle Stone Age and Later Stone Age archaeology and Holocene rock art. Quaternary International, 343, 136–147. Assefa, Z., Asrat, A., Hovers, E., Lam, Y., Pearson, O., & Pleurdeau, D. (2018). Engraved ostrich eggshell from the Middle Stone Age contexts of Goda Buticha, Ethiopia. Journal of Archaeological Science: Reports, 17, 723–729. Avery, D. M. (1977). Past and present distribution of some rodent and insectivore species in the Southern Cape province, South Africa. New information. Annals of the South African Museum, 74(7), 201–209. Avery, D. M. (1982a). The micromammalian fauna from Border Cave, Kwazulu, South Africa. Journal of Archaeological Science, 9, 187–204. Avery, D. M. (1982b). Micromammals as palaeoenvironmental indicators and an interpretation of the Late Quaternary in the southern Cape Province, South Africa. Annals of the South African Museum, 85, 183–374. Avery, D. M. (1983). Micromammals and environmental change at Zebrarivier Cave, central Namibia. Journal of the South West Africa Scientific Society, 38, 79–86. Avery, D. M. (1987). Late Pleistocene coastal environment of the Southern Cape province of South Africa: Micromammals from Klasies river Mouth. Journal of Archaeological Science, 14(4), 405–421. Avery, D. M. (1995). A preliminary assessment of the micromammalian remains from Gladysvale Cave, South Africa. Palaeontologia Africana, 32, 1–10. Avery, D. M. (2003). Early and Middle Pleistocene environments and hominid biogeography; micromammalian evidence from Kabwe, Twin Rivers and Mumbwa Caves in central Zambia. Palaeogeography, Palaeoclimatology, Palaeoecology, 189(1–2), 55–69. Avery, D. M. (2007). Pleistocene micromammals from Wonderwerk cave, South Africa: Practical issues. Journal of Archaeological Science, 34(4), 613–625. Avery, D. M., & Avery, G. (2011). Micromammals in the Northern Cape Province of South Africa, past and present. African Natural History, 7, 9–39. Avery, G., Kandel, A., Klein, R. G., Conard, N., & Cruz-Uribe, K. (2004). Tortoises as food and taphonomic elements in palaeo landscapes. In J.-P. Brugal & J. Desse (Eds.), Petits Animaux et Societes Humaines. Du Complement Alimentaire aux Ressources Utilitaires, XXIVe rencontres internationales d’archéologie et d’histoire d’Antibes (pp. 147–161). Éditions APDCA. The Zooarchaeology of Pleistocene Africa 2053 Avery, G., Halkett, D., Orton, J., Steele, T., & Klein, R. G. (2008). The Ysterfontein 1 Middle Stone Age rock shelter and the evolution of coastal foraging. South African Archaeological Society Goodwin Series, 10, 66–89. Backwell, L. R., & d’Errico, F. (2004). The first use of bone tools: A reappraisal of the evidence from Olduvai Gorge, Tanzania. Palaeontologia Africana, 40(9), 95–158. Backwell, L. R., d’Errico, F., Banks, W. E., de la Peña, P., Sievers, C., Stratford, D., et al. (2018). New excavations at Border Cave, KwaZulu-Natal, South Africa. Journal of Field Archaeology, 43(6), 417–436. Badenhorst, S., & Plug, I. (2012). The faunal remains from the Middle Stone Age levels of Bushman Rock Shelter in South Africa. South African Archaeological Bulletin, 67(195), 16. Badenhorst, S., Van Niekerk, K. L., & Henshilwood, C. S. (2016). Large mammal remains from the 100 ka Middle Stone Age layers of Blombos Cave, South Africa. The South African Archaeological Bulletin, 71, 46–52. Bader, G., Bushozi, P., Will, M., Schmid, V., Val, A., Blessing, M., et al. (2020). Investigating the 1930s Kohl-Larsen Collection from the Lake Eyasi Basin, Tanzania. Mitteilungen der Gesellschaft für Urgeschichte, 29, 93–103. Bahra, N., Djerrab, A., Ruault-Djerrab, M., Semiane, K., & Zedam, R. (2020). Middle Stone Age technology in Algeria: A techno-economic approach case study of the Oued Bousmane site (Djebel Dyr). Quaternary International, 555, 33–46. Bailon, S., & Aouraghe, H. (2002). Amphibiens, chéloniens et squamates du Pléistocène supérieur d’El Harhoura 1 (Témara, Maroc). Geodiversitas, 24(4), 821–830. Bamford, M. K. (2011). Late Pliocene woody vegetation of Area 41, Koobi Fora, East Turkana Basin, Kenya. Review of Palaeobotany and Palynology, 164(3), 191–210. Banks, K. M., Snortland, S., Cummings, L. S., Gatto, M. C., & Usai, D. (2015). The terminal late Palaeolithic in Wadi Kubbaniya, Egypt. Antiquity, 89, 346. Barba, R., & Domínguez-Rodrigo, M. (2007). FLK North North I: “Living floor” or natural accumulation? In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic study of the bed I sites (pp. 217–228). Springer. Barham, L. S. (2000). The Middle Stone Age of Zambia, South-Central Africa. Western Academic and Specialist Press. Barham, L. S. (2002). Systematic pigment use in the Middle Pleistocene of South-Central Africa. Current Anthropology, 43, 181–190. Barham, L. S., Simms, M. J., Gilmour, M., & Debenham, N. C. (2000). Twin Rivers, excavation and behavioural record. In L. S. Barham (Ed.), The Middle Stone Age of Zambia, south-central Africa (pp. 165–216). Western Academic and Specialist Press. Barker, G., Basell, L., Brooks, I., Burn, L., Cartwright, C., Cole, F., et al. (2008). The Cyrenaican Prehistory Project 2008: The second season of investigations of the Haua Fteah cave and its landscape, and further results from the initial (2007) fieldwork. Libyan Studies, 39, 175–221. Barsky, D., Chapon-Sao, C., Bahain, J., Beyene, Y., Cauche, D., Celiberti, V., et al. (2011). The early Oldowan stone-tool assemblage from Fejej FJ-1A, Ethiopia. Journal of African Archaeology, 9(2), 207–224. Barthelme, J. (1991). Kenya: Lenderut: A new Acheulian Site in the Southern Kenya Rift. Nyame akuma, 35, 21–24. Barton, R., Bouzouggar, A., Collcutt, S. N., Schwenninger, J.-L., & Clark-Balzan, L. (2009). OSL dating of the Aterian levels at Dar es-Soltan I (Rabat, Morocco) and implications for the dispersal of modern Homo sapiens. Quaternary Science Reviews, 28(19), 1914–1931. Barton, R. N. E., Bouzouggar, A., Collcutt, S. N., Carrión Marco, Y., Clark-Balzan, L., Debenham, N. C., et al. (2016). Reconsidering the MSA to LSA transition at Taforalt Cave (Morocco) in the light of new multi-proxy dating evidence. Quaternary International, 413, 36–49. Bartram, J., Laurence, E., & Marean, C. W. (1999). Explaining the “Klasies pattern”: Kua ethnoarchaeology, the Die Kelders Middle Stone Age archaeofauna, long bone fragmentation and carnivore ravaging. Journal of Archaeological Science, 26, 9–29. Bate, D. M. (1951). The mammals from Singa and Abu Hugar. Fossil Mammals of Africa, 2(1), 50. 2054 J. C. Thompson et al. Bate, D. M. A. (1955). Vertebrate faunas of Quaternary deposits in Cyrenaica. In C. B. M. McBurney & R. W. Hey (Eds.), Prehistory and Pleistocene Geology in Cyrenaican Libya (pp. 274–291). Cambridge University Press. Bayle, E. (1854). Note sur des ossements fossiles trouvés prés de Constantine. Bulletin de la Société géologique de France 2ème sér, 11(1), 8. Beaudet, A., Zanolli, C., Redae, B. E., Endalamaw, M., Braga, J., & Macchiarelli, R. (2015). A new cercopithecoid dentognathic specimen attributed to Theropithecus from the late Early Pleistocene (c. 1 Ma) deposits of Simbiro, at Melka Kunture, Ethiopian highlands. Comptes Rendus Palevol, 14(8), 657–669. Beaumont, P. B. (1981). The Heuningneskrans Shelter. In E. A. Voigt (Ed.), Guide to archaeological sites in the Northern and Eastern Transvaal (pp. 132–145). Transvaal Museum. Beaumont, P. B., & Vogel, J. C. (2006). On a timescale for the past million years of human history in central South Africa. South African Journal of Science, 102(May/June), 217–228. Belhouchet, L. (2002). Étude technologique des bifaces des niveaux inférieur et moyen de Sidi Zin (Le Kef, Tunisie). Préhistoires Méditerranéennes, 10–11, 97–106. Ben Arous, E., Falguères, C., Tombret, O., El Hajraoui, M. A., & Nespoulet, R. (2019). Combined US-ESR dating of fossil teeth from El Harhoura 2 cave (Morocco): New data about the end of the MSA in Temara region. Quaternary International. Berger, L., & Lacruz, R. (2003). Preliminary report on the first excavations at the new fossil site of Motsetse, Gauteng, South Africa. South African Journal of Science, 99, 279. Berger, L. R., Keyser, A. W., & Tobias, P. V. (1993). Gladysvale: First early hominid site discovered in South Africa since 1948. American Journal of Physical Anthropology, 92(1), 107–111. Bernor, R. L., Armour-Chelu, M. J., Gilbert, H., Kaiser, T. M., & Schulz, E. (2010). Equidae. In L. Werdelin & W. J. Sanders (Eds.), Cenozoic mammals of Africa (pp. 691–727). University of California Press. Berthelet, A., & Chavaillon, J. (2001). The Early Palaeolithic butchery site of Barogali (Republic of Djibouti). In G. Cavarretta, P. Gioia, M. Mussi, & M. R. Palombo (Eds.), The world of elephants (Proceedings of the First International Congress) (pp. 176–179). Consiglio Nazionale delle Ricerche. Berthelet, A., & Chavaillon, J. (2004). Karre I. In J. Chavaillon & M. Piperno (Eds.), Studies on the early Paleolithic site of Melka Kunture, Ethiopia (pp. 211–251). Alvernia and Aquitania Regions: Dipartimento di Scienze Storiche, Archeologiche e Antropologiche dell’Antichità, Università di Roma “La Sapienza”, Istituto Italiano di Preistoria e Protostoria. Beyene, Y., Katoh, S., WoldeGabriel, G., Hart, W. K., Uto, K., Sudo, M., et al. (2013). The characteristics and chronology of the earliest Acheulean at Konso, Ethiopia. Proceedings of the National Academy of Sciences, USA, 110(5), 1584–1591. Beyene, Y., Asfaw, B., Sano, K., & Suwa, G. (2015). Konso-Gardula Research Project volume 2, Archaeological collection: Background and the early acheulean assemblages (Vol. Bulletin no. 48). The University Museum, The University of Tokyo. Beyin, A., Prendergast, M. E., Grillo, K. M., & Wang, H. (2017). New radiocarbon dates for terminal Pleistocene and early Holocene settlements in West Turkana, northern Kenya. Quaternary Science Reviews, 168, 208–215. Bicho, N., Haws, J., Raja, M., Madime, O., Gonçalves, C., Cascalheira, J., et al. (2016). Middle and Late Stone Age of the Niassa region, northern Mozambique. Preliminary results. Quaternary International, 404, 87–99. Bicho, N., Cascalheira, J., Haws, J., & Gonçalves, C. (2018). Middle Stone Age Technologies in Mozambique: A preliminary study of the Niassa and Massingir Regions. Journal of African Archaeology, 16(1), 60–82. Binford, L. R. (1978). Nunamiut ethnoarchaeology. Academic. Binford, L. R. (1981). Bones: Ancient men and modern myths. Academic. Binford, L. R. (1984). Faunal remains from Klasies River mouth. Academic. Binford, L. R. (1988). Fact and fiction about the Zinjanthropus floor: Data, arguments, and interpretations. Current Anthropology, 29, 123–135. The Zooarchaeology of Pleistocene Africa 2055 Bishop, L. C., & Reynolds, S. C. (2000). The Fauna from twin Rivers. In L. Barham (Ed.), The Middle Stone Age of Zambia. South Central Africa (pp. 217–222). Western Academic & Specialist Press Ltd. Bishop, L., Plummer, T., Ferraro, J., Braun, D., Ditchfield, P., Hertel, F., et al. (2006). Recent research into Oldowan hominin activities at Kanjera South, western Kenya. African Archaeological Review, 23(1–2), 31. Blegen, N., Faith, J. T., Mant-Melville, A., Peppe, D. J., & Tryon, C. A. (2017). The Middle Stone Age after 50,000 years ago: New evidence from the Late Pleistocene sediments of the eastern Lake Victoria basin, Western Kenya. PaleoAnthropology, 139, 139–169. Blegen, N., Faith, J. T., & Peppe, D. J. (2021). Tephrostratigraphy of the eastern Lake Victoria Basin including the Nyanza Rift, Kenya: Building a stratigraphic and chronological framework for modern human evolution. Quaternary Science Reviews, 256, 106823. Blumenschine, R. J. (1988). An experimental model of the timing of hominid and carnivore influence on archaeological bone assemblages. Journal of Archaeological Science, 15, 483–502. Blumenschine, R. J. (1995). Percussion marks, tooth marks, and experimental determinations of the timing of hominid and carnivore access to long bones at FLK Zinjanthropus, Olduvai Gorge, Tanzania. Journal of Human Evolution, 29(1), 21–51. Bluszcz, A. (1993). Thermoluminescence dating of deposits from the area of Bir Sahara east and Bir Tarfawi. In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 224–226). Plenum. Boisserie, J.-R., Guy, F., Delagnes, A., Hlusko, L. J., Bibi, F., Beyene, Y., et al. (2007). New palaeoanthropological research in the Plio-Pleistocene Omo Group, Lower Omo Valley, SNNPR (Southern Nations, Nationalities and People Regions), Ethiopia. Comptes Rendus Palevol, 7(7), 429–439. Bon, F., Dessie, A., Bruxelles, L., Daussy, A., Douze, K., Fauvelle-Aymar, F.-X., et al. (2013). Archéologie préhistorique de la partie centrale du Main Ethiopian Rift (bassin lacustre de Ziway-Shala): Contribution à l’établissement de la séquence Late Stone Age d’Afrique orientale. Annales d’Éthiopie, 28, 261–297. Bond, G., & Summers, R. (1951). The Quaternary succession and archaeology at Chelmer near Bulawayo, Southern Rhodesia. South African Journal of Science, 47, 200–204. Bougariane, B., Zouhri, S., Ouchaou, B., Oujaa, A., & Boudad, L. (2010). Large mammals from the Upper Pleistocene at Tamaris I ‘Grotte des gazelles’(Casablanca, Morocco): Paleoecological and biochronological implications. Historical Biology, 22(1–3), 295–302. Boule, M. (1900). Etude paléontologique et archéologique sur la station paléolithique du Lac Karar. L’Anthropologie, 11, 1–21. Bouzouggar, A., Barton, N., Vanhaeren, M., d’Errico, F., Collcutt, S., Higham, T., et al. (2007). 82,000-year-old shell beads from North Africa and implications for the origins of modern human behavior. Proceedings of the National Academy of Sciences, USA, 104(24), 9964–9969. Bouzouggar, A., Kuhn, S., Fernandez, P., Collina-Girard, J., Mouhiddine, M., Hoffman, D., et al. (2017). La grotte de Bizmoune (region d’Essaouira): Une nouvelle sequence atérienne au Maroc sud atlantique. Bulletin d’Archéologie Marocaine, 22, 26–38. Bouzouggar, A., Humphrey, L. T., Barton, N., Parfitt, S. A., Balzan, L. C., Schwenninger, J.-L., et al. (2018). 90,000 year-old specialised bone technology in the Aterian Middle Stone Age of North Africa. PLoS One, 13(10), e0202021. Braga, J., Thackeray, J. F., Bruxelles, L., Dumoncel, J., & Fourvel, J.-B. (2017). Stretching the time span of hominin evolution at Kromdraai (Gauteng, South Africa): Recent discoveries. Comptes Rendus Palevol, 16(1), 58–70. Brain, C. K. (1981). The hunters or the hunted? An introduction to African Cave Taphonomy. University of Chicago Press. Brandt, S. A., & Brook, G. A. (1984). Archaeological and Paleoenvironmental Research in Northern Somalia. Current Anthropology, 25(1), 119–121. Brandt, S. A., Fisher, E. C., Hildebrand, E. A., Vogelsang, R., Ambrose, S. H., Lesur, J., et al. (2012). Early MIS 3 occupation of Mochena Borago rockshelter, southwest Ethiopian high- 2056 J. C. Thompson et al. lands: Implications for Late Pleistocene archaeology, paleoenvironments and modern human dispersals. Quaternary International, 274, 38–54. Braun, D. R., Harris, J. W. K., Levin, N. E., McCoy, J. T., Herries, A. I. R., Bamford, M. K., et al. (2010). Early hominin diet included diverse terrestrial and aquatic animals 1.95 Ma in East Turkana, Kenya. Proceedings of the National Academy of Sciences, USA, 107(22), 10002–10007. Braun, D. R., Levin, N. E., Stynder, D., Herries, A. I. R., Archer, W., Forrest, F., et al. (2013). MidPleistocene Hominin occupation at Elandsfontein, Western Cape, South Africa. Quaternary Science Reviews, 82, 145–166. Braun, D. R., Aldeias, V., Archer, W., Arrowsmith, J. R., Baraki, N., Campisano, C. J., et al. (2019). Earliest known Oldowan artifacts at >2.58 Ma from Ledi-Geraru, Ethiopia, highlight early technological diversity. Proceedings of the National Academy of Sciences, USA, 116(24), 11712–11717. Breuil, H., & Frobenius, L. (1931). L’Afrique préhistorique. Cahiers d’Art. Breunig, P. (1989). Untersuchungen zur Besiedlungsgeschichte des Brandbergs. In H. Pager (Ed.), The Rock Paintings of the Upper Brandberg, part I – Amis Gorge (Vol. 1). Africa Praehistorica. Brink, J. S. (1988). The taphonomy and palaeoecology of the Florisbad spring fauna. Palaeoecology of Africa, 19, 169–179. Brink, J., & Herries, A. (2012). First hominine remains from a ∼1.0 million year old bone bed at Cornelia-Uitzoek, Free State Province, South Africa. Journal of Human Evolution, 63, 527–535. Brink, J. S., Herries, A. I., Moggi-Cecchi, J., Gowlett, J. A., Bousman, C. B., Hancox, J. P., et al. (2012). First hominine remains from a 1.0 million year old bone bed at Cornelia-Uitzoek, Free State Province, South Africa. Journal of Human Evolution, 63(3), 527–535. Brink, J. S., Bousman, C. B., & Grün, R. (2015). A reconstruction of the skull of Megalotragus priscus (Broom, 1909), based on a find from Erfkroon, Modder River, South Africa, with notes on the chronology and biogeography of the species. In Changing climates, ecosystems and environments within arid southern Africa and adjoining regions (pp. 83–106). CRC Press. Brink, J., Holt, S., & Horwitz, L. K. (2016). The Oldowan and Early Acheulean Mammalian Fauna of Wonderwerk Cave (Northern Cape Province, South Africa). African Archaeological Review, 33(3), 223–250. Bromage, T. G., Schrenk, F., & Juwayeyi, Y. M. (1995). Paleobiogeography of the Malawi Rift: Age and vertebrate paleontology of the Chiwondo Beds, northern Malawi. Journal of Human Evolution, 28(1), 37–57. Brooks, A. S. (1984). San land-use patterns, past and present: Implications for southern African prehistory. In M. Hall, G. Avery, D. M. Avery, M. L. Wilson, & A. J. B. Humphreys (Eds.), Frontiers: Southern African archaeology today (Vol. 10, pp. 40–52). British Archaeological Reports. Brooks, A. S., & Yellen, J. E. (1977a). Archaeological excavations at ≠Gi: A preliminary report on the first two field seasons. Botswana Notes and Records, 9, 21–30. Brooks, A. S., & Yellen, J. E. (1977b). Archaeological excavations at Gi: A preliminary report on the first two field seasons. Botswana Notes and Records, 9, 21–30. Brooks, A. S., Helgren, D. M., Cramer, J. S., Franklin, A., Hornyak, W., Keating, J. M., et al. (1995). Dating and context of three Middle Stone Age sites with bone points in the upper Semliki Valley, Zaire. Science, 268, 548–553. Brooks, A. S., Yellen, J. E., Potts, R., Behrensmeyer, A. K., Deino, A. L., Leslie, D. E., et al. (2018). Long-distance stone transport and pigment use in the earliest Middle Stone Age. Science, 360(6384), 90–94. Brown, F. H., Haileab, B., & McDougall, I. (2006). Sequence of tuffs between the KBS Tuff and the Chari Tuff in the Turkana Basin, Kenya and Ethiopia. Journal of the Geological Society, 163(1), 185–204. Bunn, H. T. (1981). Archaeological evidence for meat-eating by Plio-Pleistocene hominids from Koobi Fora and Olduvai Gorge. Nature, 291, 574–577. Bunn, H. T. (1997). The bone assemblages from the excavated sites. In G. Isaac (Ed.), Koobi Fora Research Project, Volume 5: Plio-Pleistocene archaeology (pp. 402–458). Clarendon Press. The Zooarchaeology of Pleistocene Africa 2057 Bunn, H. T., & Kroll, E. M. (1986). Systematic butchery by Plio-Pleistocene hominids at Olduvai Gorge, Tanzania. Current Anthropology, 27, 431–452. Bunn, H. T., Harris, J. W. K., Isaac, G., Kaufulu, Z., Kroll, E. M., Schick, K. D., et al. (1980). FxJj50: An early Pleistocene site in northern Kenya. World Archaeology, 12, 109–136. Bushozi, P. M., Skinner, A., & de Luque, L. (2020). The Middle Stone Age (MSA) technological patterns, innovations, and behavioral changes at Bed VIA of Mumba Rockshelter, northern Tanzania. African Archaeological Review, 37, 1–18. Butzer, K. W., Clark, J. D., & Cooke, H. B. S. (1974). The geology, archaeology and fossil mammals of the Cornelia Beds, OFS. In The geology, archaeology and fossil mammals of the Cornelia (Vol. 9). Memoirs of the Bloemfontein Nasionale Museum. Butzer, K. W., Fock, G. J., Scott, L., & Stuckenrath, R. (1979). Dating and context of rock engravings in southern Africa. Science, 203(4386), 1201–1214. Cáceres, I., Kandi, N., Sahnouni, M., Zoheir, H., & van der Made, J. (2015a). Taphonomy of the Early Pleistocene fossil assemblages from Ain Boucherit (Ain Hanech, Algeria). In M. Sahnouni, S. S. and J. R. Garaizar (Eds.), Proceedings of the II meeting of African prehistory (pp. 172–196). Cáceres, I., Rogers, M. J., Semaw, S., & Leiss, A. C. (2015b). Hominin exploitation of animal resources at the Gona Plio-Pleistocene archaeological sites (Afar, Ethiopia). In M. Sahnouni & J. R. Garaizar (Eds.), Proceedings of the II meeting of African prehistory: Burgos 15–16 April, 2015 (pp. 198–217). Centro Nacional de Investigación sobre la Evolución Humana: CENIEH. Cain, C. (2006). Human activity suggested by the taphonomy of 60 ka and 50 ka faunal remains from Sibudu Cave. Southern African Humanities, 18, 241–260. Campisano, C. J. (2012). Geological summary of the Busidima formation (Plio-Pleistocene) at the Hadar paleoanthropological site, Afar Depression, Ethiopia. Journal of Human Evolution, 62(3), 338–352. Campmas, E. (2012). Caractérisation de l’occupation des sites de la région de Témara (Maroc) au Pléistocène supérieur et nouvelles données sur la subsistance des hommes du Paléolithique moyen d’Afrique du Nord: Exemples des approches taphonomiques et archéozoologiques menées sur les faunes d’El Harhoura 2 et d’El Mnasra. PhD dissertation, Université Bordeaux 1. Campmas, E. (2017). Integrating human-animal relationships into new data on Aterian complexity: A paradigm shift for the north African Middle Stone Age. African Archaeological Review, 34(4), 469–491. Campmas, E., & Daujeard, C. (2020). Co-évolution Hommes-Canidés en Afrique du Nord-Ouest: une longue histoire. In M. Boudadi-Maligne & J.-B. Mallye (Eds.), Relations hommes – canidés de la préhistoire aux périodes modernes (pp. 11–32). Ausonius éditions, collection DAN@ 3, 2020. Campmas, E., Michel, P., Costamagno, S., Amani, F., Stoetzel, E., Nespoulet, R., et al. (2015). Were Upper Pleistocene human/non-human predator occupations at the Témara caves (El Harhoura 2 and El Mnasra, Morocco) influenced by climate change? Journal of Human Evolution, 78, 122–143. Campmas, E., Chakroun, A., & Merzoug, S. (2016). Données préliminaires sur l’exploitation de la malacofaune marine par les groupes ibéromaurusiens de l’abri Alain (Oran, Algérie). PALEO. Revue d’archéologie préhistorique, 27, 83–104. Campmas, E., Michel, P., Costamagno, S., El Hajraoui, M. A., & Nespoulet, R. (2017). Which predators are responsible for faunal accumulations at the Late Pleistocene layers of El Harhoura 2 Cave (Témara, Morocco)? Comptes Rendus Palevol, 16(3), 333–350. Carrière, G. (1886). Quelques stations préhistoriques de la province d’Oran. Bulletin de la Société Géographique et Archéologique, Province d’Oran, VI, 136–154. Caton-Thompson, G. (1952). Kharga oasis in prehistory. Athlone Press. Caton-Thompson, G., & Gardner, E. (1932). The prehistoric geography of Kharga Oasis. The Geographical Journal, 80(5), 369–406. 2058 J. C. Thompson et al. Chaid-Saoudi, Y., Geraads, D., & Raynal, J.-P. (2006). The fauna and associated artefacts from the Lower Pleistocene site of Mansourah (Constantine, Algeria). Comptes Rendus Palevol, 5(8), 963–971. Chaix, L., Faure, M., Guerin, C., & Honegger, M. (2000). Kaddanarti, a lower Pleistocene assemblage from Northern Sudan. In L. Krzyżaniak & K. Kroeper (Eds.), Recent research into the Stone Age of Northeastern Africa (pp. 33–46). Poznań Archaeological Museum. Chavaillon, J., & Berthelet, A. (2004). The archaeological sites of Melka Kunture. In J. Chavaillon & M. Piperno (Eds.), Studies on the early Paleolithic site of Melka Kunture, Ethiopia (pp. 25–80). Alvernia and Aquitania Regions: Dipartimento di Scienze Storiche, Archeologiche e Antropologiche dell’Antichità, Università di Roma “La Sapienza”, Istituto Italiano di Preistoria e Protostoria. Chavaillon, N., Chavaillon, J., Geraads, D., & Guillemot, C. (1985). Gotera. Un site Paléolithique récent d’Ethiopie. Mémoire de l’institut mauritanien de la recherche scientifique, 59, 1–58. Chazan, M., Berna, F., Brink, J., Ecker, M., Holt, S., Porat, N., et al. (2020). Archeology, environment, and chronology of the early Middle Stone Age component of Wonderwerk cave. Journal of Paleolithic Archaeology, 3, 302–335. Cheddadi, A. (1986). Les rongeurs d’âge pléistocène moyen du site de l’Aïn Bahya (Skhirat, Maroc atlantique); Implications stratigraphiques et paléoécologiques. Thèse ENS Souissi. Chibane, S. (2016). Étude taphonomique et archéozoologique de la faune ibéromaurusienne de l’Abri Alain (Oran, Algérie). M2 Arts et Cultures de la Préhistoire et de la Protohistoire. Choubert, G., Ennouchi, E., & Marcais, J. (1948). Contribution à l’étude du Pliocène de la région de Port Lyautey Oued Fouarat. Notes et mémoires du Service géologique, 1(71), 15–30. Churcher, C. S., & Smith, P. E. L. (1972). Kom Ombo: Preliminary report on the fauna of late Paleolithic sites in Upper Egypt. Science, 177, 259–261. Churcher, C. S., Kleindienst, M. R., & Schwarcz, H. P. (1999). Faunal remains from a Middle Pleistocene lacustrine marl in Dakhleh Oasis, Egypt: Palaeoenvironmental reconstructions. Palaeogeography, Palaeoclimatology, Palaeoecology, 154(4), 301–312. Churcher, C. S., Kleindienst, M. R., Wiseman, M. F., & McDonald, M. M. (2008). The Quaternary faunas of Dakhleh Oasis, Western Desert of Egypt. In M. F. Wiseman (Ed.), The oasis papers 2, proceedings of the second international conference of the Dakhleh oasis project (pp. 1–24). Oxbow Books. Churchill, S. E., Brink, J. S., Berger, L. R., Hutchison, R. A., Rossouw, L., Stynder, D., et al. (2000). Erfkroon: A new Florisian fossil locality from fluvial contexts in the western Free State, South Africa. South African Journal of Science, 96(4), 161–163. Clark, J. D. (1945). Short notes on Stone age sites at Yavello, southern Abyssinia. Transactions of the Royal Society of South Africa, 31(1), 29–37. Clark, J. D. (1967). Atlas of African prehistory. University of Chicago Press. Clark, J. D. (1988). The Middle Stone Age of East Africa and the beginnings of regional identity. Journal of World Prehistory, 2(3), 235–305. Clark, J. F. (1997). Foreword. In G. Isaac (Ed.), Koobi fora research project, volume 5: PlioPleistocene archaeology (pp. v–vi). Clarendon Press. Clark, J. D. (2001). Kalambo falls prehistoric site: The earlier cultures, Middle and Earlier Stone Age. Cambridge University Press. Clark, J. L. (2011). The evolution of human culture during the later Pleistocene: Using fauna to test models on the emergence and nature of “modern” human behavior. Journal of Anthropological Archaeology, 30(3), 273–291. Clark, J. L. (2017). The Howieson’s Poort fauna from Sibudu Cave: Documenting continuity and change within Middle Stone Age industries. Journal of Human Evolution, 107, 49–70. Clark, J. L. (2019). The Still Bay and pre-Still Bay fauna from Sibudu Cave: Taphonomic and taxonomic analysis of the macromammal remains from the Wadley excavations. Journal of Paleolithic Archaeology, 2(1), 26–73. The Zooarchaeology of Pleistocene Africa 2059 Clark, J. D., & Brown, K. S. (2001). The Twin Rivers Kopje, Zambia: Stratigraphy, fauna, and artefact assemblages from the 1954 and 1956 excavations. Journal of Archaeological Science, 28, 305–330. Clark, J. D., & Kurashina, H. (1979). Hominid occupation of the East-Central Highlands of Ethiopia in the Plio–Pleistocene. Nature, 282(5734), 33–39. Clark, J. L., & Plug, I. (2008). Animal exploitation strategies during the South African Middle Stone Age: Howiesons Poort and post-Howiesons Poort fauna from Sibudu Cave. Journal of Human Evolution, 54(6), 886–898. Clark, J. D., & Williams, M. A. J. (1978). Recent archaeological research in Southeastern Ethiopia (1974–1975). Annales d’Éthiopie, 11, 19–44. Clark, J. D., Asfaw, B., Assefa, G., Harris, J. W. K., Kurashina, H., Walter, R. C., et al. (1984). Palaeoanthropological discoveries in the Middle Awash Valley, Ethiopia. Nature, 307, 423–428. Clark, J. D., Beyene, Y., Giday, W., Hart, W. K., Renne, P. R., Gilbert, H., et al. (2003). Stratigraphic, chronological and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature, 423, 747–752. Clark, J. D., Gifford-Gonzalez, D., & Batkin, J. (2008). Adrar Bous: Archaeology of a central Saharan granitic ring complex in Niger. Royal Museum for Central Africa. Close, A. E. (1980). Current research and recent radiocarbon dates from northern Africa. Journal of African History, 21, 145–167. Close, A., & Wendorf, F. (1993). The archaeology of the sandsheet. In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 473–487). Plenum. Coetzee, E. M., & van Zinderen Bakker, E. M. (1970). Palaeoecological problems of the Quaternary of Africa. South African Journal of Science, 66(3), 78. Cohen, K. M., Finney, S. C., Gibbard, P. L., & Fan, J. (2013). The ICS international chronostratigraphic chart. Episodes, 36(3), 199–204. Cole, G. H. (1967). The later Acheulian and Sangoan of southern Uganda. In W. W. Bishop & J. D. Clark (Eds.), Background to evolution in Africa (pp. 481–528). Chicago University Press. Cole, G. H., & Kleindienst, M. R. (1974). Further reflections on the Isimila Acheulian. Quaternary Research, 4, 346–355. Collins, B. (2015). The Taphonomy of the final Middle Stone Age Fauna from Sibudu Cave, South Africa. International Journal of Osteoarchaeology, 25(6), 805–815. Collins, B. (2016). Foraging strategies during the final Middle Stone Age occupation at Sibudu Cave, South Africa. Journal of Archaeological Science: Reports, 5, 61–70. Collins, B. R., & Willoughby, P. R. (2010). The faunal analysis of Magubike and Mlambalasi, two MSA-LSA archaeological sites from Iringa District, Tanzania. Journal of Taphonomy, 8(1), 33–68. Conroy, G. C., Jolly, C. J., Cramer, D., & Kalb, J. E. (1978). Newly discovered fossil hominid skull from the Afar depression, Ethiopia. Nature, 276, 67–70. Cooke, H. B. S. (1950). Quaternary fossils from northern Rhodesia. In J. D. Clark (Ed.), The Stone Age cultures of Northern Rhodesia (pp. 137–142). South African Archaeological Society. Cooke, H. B. S. (1957). Short report on a collection of fossil mammals from Kakontwe, Belgian Congo. In Proceedings of the Third Pan-African Congress on Prehistory, 1955 (pp. 100–101). Cooke, H. B. S. (1962). Notes on the faunal material from the Cave of Hearths and Kalkbank. In Prehistory of the Transvaal (pp. 447–453). Witwatersrand University Press. Cooke, H. B. S. (1963). Pleistocene mammal faunas of Africa, with particular reference to southern Africa. African Ecology and Human Evolution, 36, 65–116. Cooke, H. B. S. (1974). The fossil mammals of Cornelia, OFS, South Africa. In K. Butzer, J. D. Clark, & H. B. S. Cooke (Eds.), The geology, archaeology and fossil mammals of the Cornelia Beds, O.F.S (Vol. 9, pp. 63–84). Memoirs of the Bloemfontein Nasionale Museum. Cooke, H. B. S. (1988). Chapter 11: The larger mammals from the Cave of hearths. In R. J. Mason & C. K. Brain (Eds.), Cave of hearths, Makapansgat, Transvaal (pp. 507–522). University of the Witwatersrand. 2060 J. C. Thompson et al. Cooke, H. B. S., & Wells, L. H. (1951). Fossil remains from Chelmer, near Bulawayo, southern Rhodesia. South African Journal of Science, 47(7), 205–209. Cornelissen, E. (1995). Indications of post Acheulean (Sangoan) in the Kapthurin Formation, Baringo, Kenya. L’Anthropologie, 99(1), 55–73. Cornelissen, E. (2013). Hunting and gathering in Africa’s tropical forests at the end of the Pleistocene and in the early Holocene. In P. Mitchell & P. Lane (Eds.), The Oxford handbook of African archaeology (pp. 403–418). Oxford University Press. Cornelissen, E., Boven, A., Dabi, A., Hus, J., Yong, K. J., Keppens, E., et al. (1990). The Kapthurin Formation revisited. African Archaeological Review, 8(1), 23–75. Coryndon, S. C., Gentry, A., Harris, J. M., Hooijer, D. A., Maglio, V. J., & Howell, F. C. (1972). Mammalian remains from the Isimila prehistoric site, Tanzania. Nature, 237(5353), 292–292. Crevecoeur, I. (2012). The Upper Paleolithic human remains of Nazlet Khater 2 (Egypt) and past modern human diversity. In J.-J. Hublin & S. P. McPherron (Eds.), Modern origins: A north African perspective (pp. 205–219). Springer. Crevecoeur, I., Brooks, A., Ribot, I., Cornelissen, E., & Semal, P. (2016). Late Stone Age human remains from Ishango (Democratic Republic of Congo): New insights on Late Pleistocene modern human diversity in Africa. Journal of Human Evolution, 96, 35–57. Cruz-Uribe, K. (1983). The mammalian fauna from Redcliff Cave, Zimbabwe. South African Archaeological Bulletin, 38(137), 7–16. Cruz-Uribe, K., & Klein, R. G. (1983). Faunal remains from some Middle and Later Stone Age sites in South West Africa. South West Africa Scientific Society, 36(37), 91–114. Cruz-Uribe, K., Klein, R. G., Avery, G., Avery, M., Halkett, D., Hart, T., et al. (2003). Excavation of buried Late Acheulean (Mid-Quaternary) land surfaces at Duinefontein 2, Western Cape Province, South Africa. Journal of Archaeological Science, 30(5), 559–575. Curnoe, D., Herries, A., Brink, J., Hopley, P., Van Reyneveld, K., Henderson, Z., et al. (2006). Discovery of Middle Pleistocene fossil and stone tool-bearing deposits at Groot Kloof, Ghaap escarpment, Northern Cape province. South African Journal of Science, 102, 180–184. d’Errico, F., Pitarch Martí, A., Shipton, C., Le Vraux, E., Ndiema, E., Goldstein, S., et al. (2020). Trajectories of cultural innovation from the Middle to Later Stone Age in Eastern Africa: Personal ornaments, bone artifacts, and ocher from Panga ya Saidi, Kenya. Journal of Human Evolution, 141, 102737. Dachy, T., Briois, F., Marchand, S., Minotti, M., Lesur, J., & Wuttmann, M. (2018). Living in an Egyptian oasis: Reconstruction of the Holocene archaeological sequence in Kharga. African Archaeological Review, 35(4), 531–566. Dart, R. A. (1949). The predatory implement technique of the Australopithecines. American Journal of Physical Anthropology, 7, 1–16. Daujeard, C., Geraads, D., Raynal, J.-P., Mohib, A., Gallotti, R., & Sbihi-Alaoui, F.-Z. (2011). Carnivores et/ou hommes dans deux sites moustéro-atériens de Dar Bouazza (Casablanca, Maroc): les données de la taphonomie. In Prédateur dans tous leurs états. Évolution, Biodiversité, Interactions, Mythes, Symboles. XXXIe rencontres internationales d’archéologie et d’histoire d’Antibes (pp. 49–64). APDCA. Daujeard, C., Geraads, D., Gallotti, R., Lefèvre, D., Mohib, A., Raynal, J.-P., et al. (2016). Pleistocene hominins as a resource for carnivores: A c. 500,000-year-old human femur bearing tooth-Marks in North Africa (Thomas Quarry I, Morocco). PLoS One, 11(4), e0152284–e0152284. Daujeard, C., Falguères, C., Shao, Q., Geraads, D., Hublin, J.-J., Lefèvre, D., et al. (2020). Earliest African evidence of carcass processing and consumption in cave at 700 ka, Casablanca, Morocco. Scientific Reports, 10(1), 4761. Davies, O. (1975). Excavations at Shongweni South Cave: The oldest evidence to date for cultigens in southern Africa. Annals of the Natal Museum, 22(2), 627–662. Davis, C. (2019). Foraging Along Blue Highways: Seasonality and Subsistence Strategies in the Middle Stone Age of Ethiopia. PhD dissertation, The University of Texas at Austin. The Zooarchaeology of Pleistocene Africa 2061 de Heinzelin, J., Clark, J. D., White, T. D., Hart, W. S., Renne, P. R., WoldeGabriel, G., et al. (1999). Environment and behavior of 2.5-million-year-old Bouri hominids. Science, 284, 625–629. de la Torre, I. (2011). The Early Stone Age lithic assemblages of Gadeb (Ethiopia) and the developed Oldowan/early Acheulean in East Africa. Journal of Human Evolution, 60(6), 768–812. de la Torre, I., Mora, R., & Martínez-Moreno, J. (2008). The early Acheulean in Peninj (Lake Natron, Tanzania). Journal of Anthropological Archaeology, 27(2), 244–264. de la Torre, I., Benito-Calvo, A., Arroyo, A., Zupancich, A., & Proffitt, T. (2013). Experimental protocols for the study of battered stone anvils from Olduvai Gorge (Tanzania). Journal of Archaeological Science, 40(1), 313–332. de la Torre, I., Albert, R. M., Arroyo, A., Macphail, R., McHenry, L. J., Mora, R., et al. (2018). New excavations at the HWK EE site: Archaeology, paleoenvironment and site formation processes during late Oldowan times at Olduvai Gorge, Tanzania. Journal of Human Evolution, 120, 140–202. de la Torre, I., Benito-Calvo, A., Martín-Ramos, C., McHenry, L. J., Mora, R., Njau, J. K., et al. (2021). New excavations in the MNK Skull site, and the last appearance of the Oldowan and Homo habilis at Olduvai Gorge, Tanzania. Journal of Anthropological Archaeology, 61, 101255. de Lumley, H., & Yonas, B. (2004). Les sites préhistoriques de la région de Fejej, Sud-Omo, Éthiopie, dans leur contexte stratigraphique et paléontologique. Éditions Recherche sur les civilisations. de Maret, P., Clist, B., & Van Neer, W. (1987). Resultats des premieres fouilles dans les abris de shum laka et d’abeke au nord-ouest du Cameroun. L’Anthropologie, 91(2), 559–584. de Matos, D., & Pereira, T. (2020). Middle Stone Age lithic assemblages from Leba Cave (Southwest Angola). Journal of Archaeological Science: Reports, 32, 102413. de Ruiter, D. J., Brophy, J. K., Lewis, P. J., Churchill, S. E., & Berger, L. R. (2008). Faunal assemblage composition and paleoenvironment of Plovers Lake, a Middle Stone Age locality in Gauteng Province, South Africa. Journal of Human Evolution, 55(6), 1102–1117. de Ruiter, D. J., Pickering, R., Steininger, C. M., Kramers, J. D., Hancox, P. J., Churchill, S. E., et al. (2009). New Australopithecus robustus fossils and associated U-Pb dates from Cooper’s Cave (Gauteng, South Africa). Journal of Human Evolution, 56(5), 497–513. Deacon, H. J. (1969). Melkhoutboom Cave, Alexandria district, Cape Province: A report on the 1967 investigation. Annals of the Cape Provincial Museums, 6(13), 141–169. Deacon, H. J. (1979). Excavations at Boomplaas Cave – A sequence through the Upper Pleistocene and Holocene in South Africa. World Archaeology, 10(3), 241–257. Debénath, A. (1976). La Grotte de Dar-es-Soltane II à Rabat (Maroc): Géologie et préhistoire. Bulletins et Mémoires de la Société d’Anthropologie de Paris, 3(2), 181–182. Debénath, A. (1982). Découverte d’une mandibule humaine atérienne à El Harhoura, province de Rabat. Bulletin d’Archéologie Marocaine, 12, 1–2. Debénath, A. (1992). Hommes et cultures matérielles de l’Atérien marocain. L’anthropologie, 96(4), 711–720. Debénath, A., Raynal, J. P., & Texier, P. J. (1982). Position stratigraphique des restes humains paléolithiques marocains sur la base des travaux récents. Comptes Rendus de l’Académie des Sciences, Série II, 294, 1247–1250. Delfino, M., Candilio, F., Carnevale, G., Coppa, A., Medin, T., Pavia, M., et al. (2018). The early Pleistocene vertebrate fauna of Mulhuli-Amo (Buia area, Danakil Depression, Eritrea). Bollettino della Società Paleontologica Italiana, 57(1), 27–44. Dellali, J. (1968). Contribution à l’étude de la faune des Mammifères fossiles de Bouknadel (Maroc). DES, Géol., Fac. Sci. Bordeaux, dactylogn, 25. Denbow, J. (1973). Malowa Rockshelter, archaeological report (Occasional papers (Vol. 14)). Government Press, Zomba: Malwai Department of Antiquities. Derricourt, R. M. (1971). Radiocarbon chronology for Egypt and North Africa. Journal of Near Eastern Studies, 30(4), 271–292. Dewar, G., & Stewart, B. A. (2012). Preliminary results of excavations at Spitzkloof Rockshelter, Richtersveld, South Africa. Quaternary International, 270, 30–39. 2062 J. C. Thompson et al. Di Vincenzo, F., Rodriguez, L., Carretero, J. M., Collina, C., Geraads, D., Piperno, M., et al. (2015). The massive fossil humerus from the Oldowan horizon of Gombore I, Melka Kunture (Ethiopia, >1.39 Ma). Quaternary Science Reviews, 122, 207–221. Dibble, H. L., Aldeias, V., Alvarez-Fernandez, E., Blackwell, B. A., Hallett-Desguez, E., Jacobs, Z., et al. (2012). New excavations at the site of Contrebandiers Cave, Morocco. PaleoAnthropology, 2012, 145–201. Dibble, H. L., Aldeias, V., Jacobs, Z., Olszewski, D. I., Rezek, Z., Lin, S. C., et al. (2013). On the industrial attributions of the Aterian and Mousterian of the Maghreb. Journal of Human Evolution, 64(3), 194–210. Diez-Martín, F., Sánchez-Yustos, P., Uribelarrea, D., Domínguez-Rodrigo, M., Fraile-Márquez, C., Obregón, R.-A., et al. (2014). New archaeological and geological research at SHK main site (Bed II, Olduvai Gorge, Tanzania). Quaternary International, 322, 107–128. Diez-Martín, F., Sánchez Yustos, P., Uribelarrea, D., Baquedano, E., Mark, D. F., Mabulla, A., et al. (2015). The origin of the Acheulean: The 1.7 million-year-old site of FLK West, Olduvai Gorge (Tanzania). Scientific Reports, 5(1), 17839. Discamps, E., & Henshilwood, C. S. (2015). Intra-site variability in the still bay Fauna at Blombos cave: Implications for explanatory models of the Middle Stone Age cultural and technological evolution. PLoS One, 10(12), e0144866. Discamps, E., Henshilwood, C. S., & van Niekerk, K. L. (2020). Large mammal exploitation during the c. 14-11 ka Oakhurst techno-complex at Klipdrift Cave, South Africa. South African Journal of Science, 116(5–6), 1–7. Doerschner, N., Fitzsimmons, K. E., Ditchfield, P., McLaren, S. J., Steele, T. E., Zielhofer, C., et al. (2016). A new chronology for Rhafas, Northeast Morocco, spanning the North African Middle Stone Age through to the Neolithic. PLoS One, 11(9), e0162280. Domínguez-Rodrigo, M. (2002). Hunting and scavenging by early humans: The state of the debate. Journal of World Prehistory, 16, 1–54. Domínguez-Rodrigo, M. (2019). Successful classification of experimental bone surface modifications (BSM) through machine learning algorithms: A solution to the controversial use of BSM in paleoanthropology? Archaeological and Anthropological Sciences, 11(6), 2711–2725. Domínguez-Rodrigo, M., & Barba, R. (2007a). Five more arguments to invalidate the passive scavenging version of the carnivore-hominid-carnivore model: A reply to Blumenschine et al. (2007a). Journal of Human Evolution, 53, 427–433. Domínguez-Rodrigo, M., & Barba, R. (2007b). A palimpsest at FLK North 1-2: Independent carnivore- and hominid-made bone accumulations. In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 127–164). Springer. Domínguez-Rodrigo, M., & Martinez-Navarro, B. (2012). Taphonomic analysis of the early Pleistocene (2.4 Ma) faunal assemblage from AL 894 (Hadar, Ethiopia). Journal of Human Evolution, 62(3), 315–327. Domínguez-Rodrigo, M., & Organista, E. (2007). Natural background bone assemblages and their ravaging stages in Olduvai Bed I. In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 201–216). Springer. Domínguez-Rodrigo, M., de la Torre, I., de Luque, L., Alcal, L., Mora, R., Serrallonga, J., et al. (2002). The ST site complex at Peninj, West Lake Natron, Tanzania: Implications for early hominid behavioural models. Journal of Archaeological Science, 29(6), 639–666. Domínguez-Rodrigo, M., Pickering, T. R., Semaw, S., & Rogers, M. J. (2005). Cutmarked bones from Pliocene archaeological sites at Gona, Afar, Ethiopia: Implications for the function of the world’s oldest stone tools. Journal of Human Evolution, 48(2), 109–121. Domínguez-Rodrigo, M., Barba, R., de la Torre, I., & R., M. (2007a). A cautionary tale about early archaeological sites: A reanalysis of FLK North 6. In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 101–126). Springer. Domínguez-Rodrigo, M., Barba, R., & Egeland, C. P. (2007b). Deconstructing Olduvai: A taphonomic study of the Bed I sites (Vertebrate paleobiology and paleoanthropology). Springer. The Zooarchaeology of Pleistocene Africa 2063 Domínguez-Rodrigo, M., Barba, R., & Organista, E. (2007c). A taphonomic study of FLK North 3 and 4: A felid-hyaenid and hominid palimpsest. In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 165–190). Springer. Domínguez-Rodrigo, M., Díez-Martín, F., Mabulla, A., Luque, L., Alcalá, L., Tarriño, A., et al. (2007d). The archaeology of the Middle Pleistocene of Lake Eyasi, Tanzania. Journal of African Archaeology, 5(1), 47. Domínguez-Rodrigo, M., Mabulla, A., Luque, L., Thompson, J. W., Rink, J., Bushozi, P., et al. (2008). A new archaic Homo sapiens fossil from Lake Eyasi, Tanzania. Journal of Human Evolution, 54(6), 899–903. Domínguez-Rodrigo, M., Pickering, T. R., & Bunn, H. T. (2010). Configurational approach to identifying the earliest hominin butchers. Proceedings of the National Academy of Sciences, USA, 107(49), 20929–20934. Domínguez-Rodrigo, M., Bunn, H. T., Mabulla, A. Z. P., Baquedano, E., Uribelarrea, D., PérezGonzález, A., et al. (2014). On meat eating and human evolution: A taphonomic analysis of BK4b (Upper Bed II, Olduvai Gorge, Tanzania), and its bearing on hominin megafaunal consumption. Quaternary International, 322, 129–152. Douka, K., Jacobs, Z., Lane, C., Grün, R., Farr, L., Hunt, C., et al. (2014). The chronostratigraphy of the Haua Fteah cave (Cyrenaica, northeast Libya). Journal of Human Evolution, 66, 39–63. Dusseldorp, G. L. (2010). Prey choice during the South African Middle Stone Age: Avoiding dangerous prey or maximising returns? African Archaeological Review, 27, 107–133. Echassoux, A. (2012). Comportements de subsistance et modifications osseuses à l’aube de l’Acheuléen à Konso, Éthiopie. L’Anthropologie, 116(3), 291–320. Egeland, C. (2007a). Zooarchaeology and taphonomy of FLK North 5. In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 191–200). Springer. Egeland, C. (2007b). Zooarchaeology and taphonomy of the DK site. In M. Domínguez-Rodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 229–238). Springer. Egeland, C. P. (2007c). Zooarchaeology and taphonomy of FLK North North 2. In M. DomínguezRodrigo, R. Barba, & C. P. Egeland (Eds.), Deconstructing Olduvai: A Taphonomic Study of the Bed I Sites (pp. 229–238). Springer. Egeland, C. (2010). Zooarchaeological and Taphonomic Perspectives on Hominid and Carnivore Interactions at Olduvai Gorge, Tanzania. PhD dissertation, Indiana University, Bloomington. Egeland, C. P., & Domínguez-Rodrigo, M. (2008). Taphonomic perspectives on hominid site use and foraging strategies during Bed II times at Olduvai Gorge, Tanzania. Journal of Human Evolution, 55(6), 1031–1052. El Hajraoui, M. (1993). Nouvelles découvertes néolithiques et atériennes dans la région de Rabat «Grotte d’El Mnasra». In Mediterrâneo (Vol. 2, pp. 105–121). Ennouchi, E. (1953). La faune des limons rouges de Bou-Knadel (Maroc). Actes du IVème Congrès de l’INQUA. Ennouchi, E. (1962). Un Néanderthalien: l’Homme du Jebel Irhoud (Maroc). L’Anthropologie, 66, 279–299. Eren, M. I., Durant, A. J., Prendergast, M., & Mabulla, A. Z. P. (2014). Middle Stone Age archaeology at Olduvai Gorge, Tanzania. Quaternary International, 322–323, 292–313. Faith, J. T. (2008). Eland, buffalo, and wild pigs: Were Middle Stone Age humans ineffective hunters? Journal of Human Evolution, 55, 24–36. Faith, J. T. (2011a). Ungulate biogeography, statistical methods, and the proficiency of Middle Stone Age hunters. Journal of Human Evolution, 60, 315–317. Faith, J. T. (2011b). Ungulate community richness, grazer extinctions, and human subsistence behavior in southern Africa’s Cape Floral Region. Palaeogeography, Palaeoclimatology, Palaeoecology, 306(3–4), 219–227. 2064 J. C. Thompson et al. Faith, J. T. (2013). Taphonomic and paleoecological change in the large mammal sequence from Boomplaas Cave, Western Cape, South Africa. Journal of Human Evolution, 65(6), 715–730. Faith, J. T. (2014). Late Pleistocene and Holocene mammal extinctions on continental Africa. Earth-Science Reviews, 128, 105–121. Faith, J. T., & Gordon, A. D. (2007). Skeletal element abundances in archaeofaunal assemblages: Economic utility, sample size, and assessment of carcass transport strategies. Journal of Archaeological Science, 34, 872–882. Faith, J. T., & Lyman, R. L. (2019). Paleozoology and Paleoenvironments: Fundamentals, assumptions, techniques. Cambridge University Press. Faith, J. T., & Thompson, J. C. (2013). Fossil evidence for seasonal calving and migration of extinct blue antelope (Hippotragus leucophaeus) in southern Africa. Journal of Biogeography, 40(11), 2108–2118. Faith, J. T., Choiniere, J. N., Tryon, C. A., Peppe, D. J., & Fox, D. L. (2011). Taxonomic status and paleoecology of Rusingoryx atopocranion (Mammalia, Artiodactyla), an extinct Pleistocene bovid from Rusinga Island, Kenya. Quaternary Research, 75(3), 697–707. Faith, J. T., Potts, R., Plummer, T. W., Bishop, L. C., Marean, C. W., & Tryon, C. A. (2012). New perspectives on Middle Pleistocene change in the large mammal faunas of East Africa: Damaliscus hypsodon sp. nov. (Mammalia, Artiodactyla) from Lainyamok, Kenya. Palaeogeography, Palaeoclimatology, Palaeoecology, 361–362, 84–93. Faith, J. T., Tryon, C. A., Peppe, D. J., Beverly, E. J., & Blegen, N. (2014). Biogeographic and evolutionary implications of an extinct Late Pleistocene impala from the Lake Victoria Basin, Kenya. Journal of Mammalian Evolution, 21(2), 213–222. Faith, J. T., Tryon, C. A., Peppe, D. J., Beverly, E. J., Blegen, N., Blumenthal, S., et al. (2015). Paleoenvironmental context of the Middle Stone Age record from Karungu, Lake Victoria Basin, Kenya, and its implications for human and faunal dispersals in East Africa. Journal of Human Evolution, 83, 28–45. Faith, J. T., Rowan, J., & Du, A. (2019). Early hominins evolved within non-analog ecosystems. Proceedings of the National Academy of Sciences, USA, 116(43), 21478–21483. Faith, J. T., Rowan, J., O’Brien, K., Blegen, N., & Peppe, D. J. (2020). Late Pleistocene mammals from Kibogo, Kenya: Systematic paleontology, paleoenvironments, and non-analog associations. Journal of Vertebrate Paleontology, 40, e1841781. Faith, J. T., Du, A., Behrensmeyer, A. K., Davies, B., Patterson, D. B., Rowan, J., et al. (2021). Rethinking the ecological drivers of hominin evolution. Trends in Ecology & Evolution, 36(9), 797–807. Ferembach, D. (1976). Les restes humains de la Grotte de Dar-es-Soltane II (Maroc). Campagne 1975. Bulletins et Mémoires de la Société d’Anthropologie de Paris, 3(2), 183–193. Fernandez, P., Bouzouggar, A., Collina-Girard, J., & Coulon, M. (2015). The last occurrence of Megaceroides algericus Lyddekker, 1890 (Mammalia, Cervidae) during the middle Holocene in the cave of Bizmoune (Morocco, Essaouira region). Quaternary International, 374, 154–167. Fernandez-Jalvo, Y., & Andrews, P. (2016). Atlas of taphonomic identifications: 1001+ images of fossil and recent mammal. Springer. Ferraro, J. V., Plummer, T. W., Pobiner, B. L., Oliver, J. S., Bishop, L. C., Braun, D. R., et al. (2013). Earliest archaeological evidence of persistent hominin carnivory. PLoS One, 8(4), e62174. Fiore, I., & Tagliacozzo, A. (2004). Taphonomic analysis of the bone remains from the Oldowan site of Garba IV. In J. Chavaillon & M. Piperno (Eds.), Studies on the Early Paleolithic site of Melka Kunture, Ethiopia (pp. 639–684). Alvernia and Aquitania Regions: Dipartimento di Scienze Storiche, Archeologiche e Antropologiche dell’Antichità, Università di Roma “La Sapienza”, Istituto Italiano di Preistoria e Protostoria. Fiore, I., Bondioli, L., Coppa, A., Macchiarelli, R., Russom, R., Kashay, H., et al. (2004). Taphonomic analysis of the late Early Pleistocene bone remains from Buia (Dandiero Basin, Danakil Depression, Eritrea): Evidence for large mammal and reptile butchering. Rivista Italiana di Paleontologia e Stratigrafia, 110, 89–97. The Zooarchaeology of Pleistocene Africa 2065 Fisher, E. C., Cawthra, H. C., Esteban, I., Jerardino, A., Neumann, F. H., Oertle, A., et al. (2020). Coastal occupation and foraging during the last glacial maximum and early Holocene at Waterfall Bluff, eastern Pondoland, South Africa. Quaternary Research, 97, 1–41. Forrest, F. L., Stynder, D. D., Bishop, L. C., Levin, N. E., Lehmann, S. B., Patterson, D. B., et al. (2018). Zooarchaeological reconstruction of newly excavated Middle Pleistocene deposits from Elandsfontein, South Africa. Journal of Archaeological Science: Reports, 17, 19–29. Fourvel, J.-B., Thackeray, J. F., Brink, J. S., O’Regan, H., & Braga, J. (2018). Taphonomic interpretations of a new Plio-Pleistocene hominin-bearing assemblage at Kromdraai (Gauteng, South Africa). Quaternary Science Reviews, 190, 81–97. Fuchs, M., Kandel, A. W., Conard, N. J., Walker, S. J., & Felix-Henningsen, P. (2008). Geoarchaeological and chronostratigraphical investigations of open-air sites in the Geelbek Dunes, South Africa. Geoarchaeology: An International Journal, 23(4), 425–449. Gallotti, R. (2013). An older origin for the Acheulean at Melka Kunture (Upper Awash, Ethiopia): Techno-economic behaviours at Garba IVD. Journal of Human Evolution, 65(5), 594–620. Gallotti, R., Raynal, J.-P., Geraads, D., & Mussi, M. (2014). Garba XIII (Melka Kunture, Upper Awash, Ethiopia): A new Acheulean site of the late Lower Pleistocene. Quaternary International, 343, 17–27. Garcea, E. A. A. (2001). Uan Tabu in the Settlement History of the Libyan Sahara (Vol. 2). All’insegna del giglio. Gardner, E. W. (1932). Some problems of the Pleistocene hydrography of Kharga Oasis, Egypt. Geological Magazine, 69(9), 386–421. Gardner, E. W. (1935). The Pleistocene fauna and flora of Kharga Oasis, Egypt. Quarterly Journal of the Geological Society, 91(1–4), 479–518. Gautier, A. (1976). Animal remains from localities near Dishna. In F. Wendorf & R. Schild (Eds.), Prehistory of the Nile Valley (pp. 365–367). Academic. Gautier, A. (1980). Contributions to the archaeozoology of Egypt. In F. Wendorf & R. Schild (Eds.), Prehistory of the Eastern Sahara (pp. 317–343). Academic. Gautier, A. (1993). The Middle Paleolithic archaeofaunas from Bir Tarfawi (Western Desert, Egypt). In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 121–143). Plenum. Gautier, A. (1995). Restes animaux Holocènes et du Paléolithique Moyen (MSA) de la Grotte de Leba sur le Plateau de Humpata (Angola). Archaeofauna, 4, 131–141. Gautier, A., & Van Neer, W. (1989). Animal remains from the late Paleolithic sequence at Wadi Kubbaniya. The Prehistory of Wadi Kubbaniya, 2, 119–161. Gautier, A., Makowiecki, D., Paner, H., & Neer, W. V. (2012). Palaeolithic big game hunting at HP766 in Wadi Umm Rahau, northern Sudan. Journal of African Archaeology, 10(2), 165. Geraads, D. (2004). The Large Mammal Fauna of the Oldowayan sites of Melka-Kunturé, Ethiopia. In V. Eisenmann & G. Petter (Eds.), Studies on the early Palaeolithic site of Melka Kunturé, Ethiopia (pp. 169–192). Istituto Italiano di Preistoria e Protoistoria. Geraads, D. (2010). Biogeographic relationships of Pliocene and Pleistocene North-western African mammals. Quaternary International, 212(2), 159–168. Geraads, D. (2012). The faunal context of human evolution in the late Middle/Late Pleistocene of northwestern Africa. In Modern Origins (pp. 49–60). Springer. Geraads, D., & Guillemot, C. (1985). Faune du site archéologique (Pléistocène supérieur) de Gotera (Gemu Gofa, Ethiopie). In J. Chavaillon & N. Chavaillon (Eds.), Gotera: Un Site Paléolithique Récent d’Ethiopie (Vol. 59). Recherches sur les Civilisations. Geraads, D., Hublin, J. J., Jaeger, J. J., Tong, H., Sen, S., & Toubeau, P. (1986). The Pleistocene hominid site of Ternifine, Algeria: New results on the environment, age, and human industries. Quaternary Research, 25, 380–386. Geraads, D., Amani, F., & Hublin, J.-J. (1992). The Middle Pleistocene site of Ain Maarouf near El Hajeb, Morocco: Discovery of a fossil hominid [Le gisement pleistocene moyen de l’Ain Maarouf pres de El Hajeb, Maroc: Presence d’un hominide]. Comptes Rendus – Academie des Sciences, Serie II, 314, 319–323. 2066 J. C. Thompson et al. Geraads, D., Alemseged, Z., Reed, D., Wynn, J., & Roman, D. C. (2004). The Pleistocene fauna (other than Primates) from Asbole, lower Awash Valley, Ethiopia, and its environmental and biochronological implications. Geobios, 37(6), 697–718. Geraads, D., Amani, F., Ben-Ncer, A., McPherron, S. P., Raynal, J.-P., & Hublin, J.-J. (2013). The rodents from the late Middle Pleistocene hominid-bearing site of J’bel Irhoud, Morocco, and their chronological and paleoenvironmental implications. Quaternary Research, 80(3), 552–561. Gibbon, R. J., Pickering, T. R., Sutton, M. B., Heaton, J. L., Kuman, K., Clarke, R. J., et al. (2014). Cosmogenic nuclide burial dating of hominin-bearing Pleistocene cave deposits at Swartkrans, South Africa. Quaternary Geochronology, 24, 10–15. Giemsch, L., Hertler, C., Märker, M., Quénéhervé, G., Saanane, C., & Schrenk, F. (2018). Acheulean Sites at Makuyuni (Lake Manyara, Tanzania): Results of Archaeological Fieldwork and Classification of the Lithic Assemblages. African Archaeological Review, 35(1), 87–106. Gifford-Gonzales, D., & Parham, J. (2008). Fauna from Adrar Bous and Surrounding Areas. In J. D. Clark, E. G. Agrilla, D. C. Crader, D. Gifford-Gonzales, A. B. Smith, & M. A. J. Williamson (Eds.), Adrar Bous: The archaeology of a granitic ring complex in Central Sahara, Niger (pp. 313–353). Annales in Archaeology, Royal Africa Museum. Gifford-Gonzalez, D. (2018). An introduction to Zooarchaeology. Springer Internatinal Publishing. Goodwin, A. J. H., & Van Riet Lowe, C. (1929). The Stone Age cultures of South Africa. Annals of the South African Museum, 27, 1–289. Greenwood, P. H., & Todd, E. J. (1976). Fish remains from Upper Paleolithic sites near Idfu and Isna. In F. Wendorf & R. Schild (Eds.), Prehistory of the Nile Valley (pp. 383–388). Academic Press. Grine, F. E., & Klein, R. G. (1985). Pleistocene and Holocene human remains from Equus Cave, South Africa. Anthropology, 8, 55–98. Grine, F. E., & Klein, R. G. (1993). Late Pleistocene human remains from the Sea Harvest site, Saldanha Bay, South Africa. South African Journal of Science, 89, 145–145. Grün, R., Pike, A., McDermott, F., Eggins, S., Mortimer, G., Aubert, M., et al. (2020). Dating the skull from Broken Hill, Zambia, and its position in human evolution. Nature, 580(7803), 372–375. Grupe, G., & Harbeck, M. (2015). Taphonomic and diagenetic processes. In W. Henke & I. Tattersall (Eds.), Handbook of paleoanthropology (2nd ed., pp. 417–439). Springer. Gutherz, X., Diaz, A., Ménard, C., Bon, F., Douze, K., Léa, V., et al. (2014). The Hargeisan revisited: Lithic industries from shelter 7 of Laas Geel, Somaliland and the transition between the Middle and Late Stone Age in the Horn of Africa. Quaternary International, 343, 69–84. Gutherz, X., Lesur, J., Cauliez, J., Charpentier, V., Diaz, A., Ismaël, M. O., et al. (2015). New insights on the first Neolithic societies in the Horn of Africa: The site of Wakrita, Djibouti. Journal of Field Archaeology, 40(1), 55–68. Gutierrez, M., Guérin, C., Léna, M., Piedade, M., & da Jesus. (2001). Exploitation d’un grand cétacé au Paléolithique ancien: le site de Dungo V à Baia Farta (Benguela, Angola). Comptes Rendus de l’Académie des Sciences-Series IIA-Earth and Planetary Science, 332(5), 357–362. Hadjouis, D. (1985). Les bovidés (Artiodactyla, Mammalia) du gisement Atérien des phacochères (Alger, Algérie). Interprétations paléontologiques et phylogénétiques. Comptes Rendus de l’Académie des Sciences, Paris, 301, 1251–1254. Hadjouis, D. (1994). Taphonomie des faunes du gisement atérien des phacochères (Alger-Algérie): Actions anthropiques. Artefacts (Viroinval), 9, 183–191. Halkett, D., Hart, T., Yates, R., Volman, T. P., Parkington, J. E., Orton, J., et al. (2003). First excavation of intact Middle Stone Age layers at Ysterfontein, Western Cape Province, South Africa: Implications for Middle Stone Age ecology. Journal of Archaeological Science, 30(8), 955–971. Hall, G., Pickering, R., Lacruz, R., Hancox, J., Berger, L. R., & Schmid, P. (2006). An Acheulean handaxe from Gladysvale cave site, Gauteng, South Africa: Research in action. South African Journal of Science, 102(3–4), 103–105. The Zooarchaeology of Pleistocene Africa 2067 Hallett, E. Y. (2018). Zooarchaeological and Taphonomic analyses of Pleistocene vertebrate Fauna from the Middle and Later Stone Age Occupations at Contrebandiers Cave, Atlantic Coast, Morocco. PhD dissertation, Arizona State University. Hanon, R., Patou-Mathis, M., Péan, S., & Prat, S. (2019). Paleobiodiversity and large mammal associations during the late Pliocene and the early Pleistocene in South Africa. Quaternaire, 30(3), 243–256. Happold, D. C. D., & Happold, M. (1989). Biogeography of montane small mammals in Malawi, Central Africa. Journal of Biogeography, 16(4), 353–367. Happold, M., & Lock, J. M. (2013). The biotic zones of Africa. In J. Kingdon, D. C. D. Happold, T. Butynski, M. Hoffmann, M. Happold, & J. Kalina (Eds.), Mammals of Africa (Vol. 1, pp. 57–74). Bloomsbury. Harcourt-Smith, W., Vincent, F., & DeVoogt, A. (2012). Northern Sudan and paleoanthropology: Historical perspectives and preliminary results from the AMNH 2011 survey. Der Antike Sudan. Harmand, S., Lewis, J. E., Feibel, C. S., Lepre, C. J., Prat, S., Lenoble, A., et al. (2015). 3.3-million-year-old stone tools from Lomekwi 3, West Turkana, Kenya. Nature, 521(7552), 310–315. Harris, J. W. K., & Isaac, G. L. (1997). Sites in the upper KBS, Okote, and Chari members: Reports. In G. Isaac & B. Isaac (Eds.), Koobi fora research project, volume 5: Plio-pleistocene archaeology (pp. 115–236). Clarendon Press. Harris, J. W. K., Williamson, P. G., Verniers, J., Tappen, M. J., Stewart, K., Helgren, D., et al. (1987). Late Pliocene hominid occupation in Central Africa: The setting, context, and character of the Senga 5A site, Zaire. Journal of Human Evolution, 16(7), 701–728. Harris, J. A., Marean, C. W., Ogle, K., & Thompson, J. (2017). The trajectory of bone surface modification studies in paleoanthropology and a new Bayesian solution to the identification controversy. Journal of Human Evolution, 110, 69–81. Harvati, K., Stringer, C., Grün, R., Aubert, M., Allsworth-Jones, P., & Folorunso, C. A. (2011). The Later Stone Age calvaria from Iwo Eleru, Nigeria: Morphology and chronology. PLoS One, 6(9), e24024. Hawkins, A. L. (2001). Getting a Handle on Tangs: Defining the Dakhleh Unit of the Aterian Technocomplex: A Study in Surface Archaeology from Dakhleh Oasis. PhD dissertation, University of Toronto. Hay, R. L. (1976). Geology of Olduvai Gorge. University of California Press. Hay, R. L. (1994). Geology and dating of Beds III, IV, and the Masek Beds. In M. D. Leakey & D. A. Roe (Eds.), Olduvai Gorge: Excavations in Beds III, IV, and the Masek Beds, 1968–1971 (Vol. 5, pp. 8–14). Cambridge University Press. Haynes, G., & Klimowicz, J. (2009). Middle Stone Age evidence from Hwange National Park, north western Zimbabwe. Nyame Akuma, 71, 63–73. Henshilwood, C. S., D’errico, F., Marean, C. W., Milo, R. G., & Yates, R. (2001a). An early bone tool industry from the Middle Stone Age at Blombos Cave, South Africa: Implications for the origins of modern human behaviour, symbolism and language. Journal of Human Evolution, 41(6), 631–678. Henshilwood, C. S., Sealy, J. C., Yates, R., Cruz-Uribe, K., Goldberg, P., Grine, F. E., et al. (2001b). Blombos Cave, southern Cape, South Africa: Preliminary report on the 1992–1999 excavations of the Middle Stone Age levels. Journal of Archaeological Science, 28(4), 421–448. Henshilwood, C. S., van Niekerk, K. L., Wurz, S., Delagnes, A., Armitage, S. J., Rifkin, R. F., et al. (2014). Klipdrift shelter, southern Cape, South Africa: Preliminary report on the Howiesons Poort layers. Journal of Archaeological Science, 45, 284–303. Herries, A. I. R., & Adams, J. W. (2013). Clarifying the context, dating and age range of the Gondolin hominins and Paranthropus in South Africa. Journal of Human Evolution, 65(5), 676–681. Hewitt, J. (1931). Artefacts from Melkhoutboom. South African Journal of Science, 28(540), 48. 2068 J. C. Thompson et al. Hlubik, S., Berna, F., Feibel, C., Braun, D., & Harris, J. W. K. (2017). Researching the nature of fire at 1.5 Mya on the site of FxJj20 AB, Koobi Fora, Kenya, using high-resolution spatial analysis and FTIR spectrometry. Current Anthropology, 58(S16), S243–S257. Hoare, S., Brink, J. S., Herries, A. I. R., Mark, D. F., Morgan, L. E., Onjala, I., et al. (2021). Geochronology of a long Pleistocene sequence at Kilombe volcano, Kenya: From the Oldowan to Middle Stone Age. Journal of Archaeological Science, 125, 105273. Hocine, S. (2016). Le site acheuléen d’Erg Tihodaïne: Caractéristiques technologiques de l’industrie lithique du Pléistocène moyen (Sahara central, Algérie). L’Anthropologie, 120(3), 263–284. Hogue, J. T., & Barton, R. N. E. (2016). New radiocarbon dates for the earliest Later Stone Age microlithic technology in Northwest Africa. Quaternary International, 413, 62–75. Horsburgh, K. A., Orton, J., & Klein, R. G. (2016). Beware the springbok in sheep’s clothing: How secure are the faunal identifications upon which we build our models? African Archaeological Review, 33(4), 353–361. Hovers, E., Gossa, T., Asrat, A., Niespolo, E. M., Resom, A., Renne, P. R., et al. (2021). The expansion of the Acheulian to the Southeastern Ethiopian Highlands: Insights from the new early Pleistocene site-complex of Melka Wakena. Quaternary Science Reviews, 253, 106763. Howell, F., Cole, G., & Kleindienst, M. (1962). Isimila: An Acheulian occupation site in the Iringa Highlands, Southern Highlands Province, Tanganyika. Actes du IVe Congrès Panafricain de Préhistoire et de l’étude du Quaternaire, 40, 43–80. Hublin, J.-J., Ben-Ncer, A., Bailey, S. E., Freidline, S. E., Neubauer, S., Skinner, M. M., et al. (2017). New fossils from Jebel Irhoud, Morocco and the pan-African origin of Homo sapiens. Nature, 546, 289–292. Humphrey, L., Freyne, A., van de Loosdrecht, M., Hogue, J. T., Turner, E., Barton, N., et al. (2019). Infant funerary behavior and kinship in Pleistocene hunter-gatherers from Morocco. Journal of Human Evolution, 135, 102637. Humphreys, A. J. B. (1974). A preliminary report on test excavations at Dikbosch Shelter I, Herbert District, Northern Cape. The South African Archaeological Bulletin, 29(115/116), 115–119. Hutson, J. M. (2018). The faunal remains from Bundu Farm and Pniel 6: Examining the problematic Middle Stone Age archaeological record within the southern African interior. Quaternary International, 466, 178–193. Hutson, J. M., & Cain, C. R. (2008). Reanalysis and reinterpretation of the Kalkbank faunal accumulation, Limpopo Province, South Africa. Journal of Taphonomy, 6(3–4), 399–428. Hutterer, R. (2010). The Middle Palaeolithic vertebrate fauna of Ifri n’Ammar. In M. Nami, J. Moser (Eds.), La Grotte d’Ifri N’Ammar - le Paléolithique Moyen (Tome 2), (pp. 307–314). Kommission für Archäologie Aussereuropaïscher Kulturen des Deutschen Archäologischen Instituts, Bonn, Germany. Isaac, G., & Behrensmeyer, A. K. (1997). Geological context and palaeoenvironments. In G. Isaac & B. Isaac (Eds.), Koobi fora research project, Volume 5: Plio-Pleistocene archaeology (pp. 12–70). Clarendon Press. Isaac, G. L., & Harris, J. W. K. (1997). Sites stratified within the KBS tuff: Reports. In G. Isaac & B. Isaac (Eds.), Koobi fora research project, Volume 5: Plio-Pleistocene archaeology (pp. 71–114). Clarendon Press. Issawi, B. (1993). Contribution to the stratigraphy of the area of Bir Tarfawi and Bir Sahara East. In F. A. Wendorf, R. A. Schild, & A. E. Close (Eds.), Egypt during the last interglacial: The middle Paleolithic of Bir Tarfawi and Bir Sahara East (pp. 11–14). Springer Science+Business Media. Jacobs, Z. (2010). An OSL chronology for the sedimentary deposits from Pinnacle Point Cave 13B—a punctuated presence. Journal of Human Evolution, 59(3–4), 289–305. Jacobs, Z., Wintle, A. G., Duller, G. A. T., Roberts, R. G., & Wadley, L. (2008). New ages for the post-Howiesons Poort, late and final Middle Stone Age at Sibudu, South Africa. Journal of Archaeological Science, 35(7), 1790–1807. Jacobs, Z., Meyer, M., Roberts, R., Aldeias, V., Dibble, H., & El Hajraoui, M. (2011). Singlegrain OSL dating at La Grotte des Contrebandiers (‘Smugglers’ Cave’), Morocco: Improved The Zooarchaeology of Pleistocene Africa 2069 age constraints for the Middle Paleolithic levels. Journal of Archaeological Science, 38(12), 3631–3643. Jacobs, Z., et al. (2012). Single-grain OSL chronologies for Middle Palaeolithic deposits at El Mnasra and El Harhoura 2, Morocco: Implications for Late Pleistocene human–environment interactions along the Atlantic coast of northwest Africa. Journal of Human Evolution, 62(3), 377–394. Jacobs, Z., Jones, B. G., Cawthra, H. C., Henshilwood, C. S., & Roberts, R. G. (2020). The chronological, sedimentary and environmental context for the archaeological deposits at Blombos Cave, South Africa. Quaternary Science Reviews, 235, 105850. James, E. C., & Thompson, J. C. (2015). On bad terms: Problems and solutions within zooarchaeological bone surface modification studies. Environmental Archaeology, 20(1), 89–103. Jenkins, K. E., Nightingale, S., Faith, J. T., Peppe, D. J., Michel, L. A., Driese, S. G., et al. (2017). Evaluating the potential for tactical hunting in the Middle Stone Age: Insights from a bonebed of the extinct bovid, Rusingoryx atopocranion. Journal of Human Evolution, 108, 72–91. Johanson, D. C. (2004). Lucy, thirty years later: An expanded view of Australopithecus afarensis. Journal of Anthropological Research, 60, 465–486. Joleaud, L. (1918). Etudes de géographie zoologique sur la Berbérie. III. Les Hippotraginés. Bull Soc Géogr Archéol Prov Oran, 3(8), 89–81. Jones, M. B., Brandt, S. A., & Marshall, F. (2018). Hunter-gatherer reliance on inselbergs, big game, and dwarf antelope at the Rifle Range Site, Buur Hakaba, southern Somalia∼ 20,000−5,000 BP. Quaternary International, 471, 55–65. Jousse, H. (2017). Atlas of Mammal Distribution through Africa from the LGM (~18KA) to Modern Times – The zooarchaeological record. Archaeopress. Juwayeyi, Y. M. (2011). Ecological pressure and the transition from foraging to agricultural lifestyle on the Shire Highlands, Malawi. Human Ecology, 39(3), 361–371. Juwayeyi, Y., & Betzler, C. G. (1995). Archaeology of the Malawi Rift: The search continues for Early Stone Age occurences in the Chiwondo Beds, northern Malawi. Journal of Human Evolution, 28, 115–116. Kalb, J. E., Wood, C. B., Smart, C., Oswald, E. B., Mabrete, A., Tebedge, S., et al. (1980). Preliminary geology and palaeontology of the Bodo D’Ar hominid site, Afar, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology, 30, 107–120. Kalb, J. E., Jaegar, M., Jolly, C. J., & Kana, B. (1982). Preliminary geology, paleontology and paleoecology of a sangoan site at Andalee, Middle Awash Valley, Ethiopia. Journal of Archaeological Science, 9(4), 349–363. Kandel, A. W., & Conard, N. J. (2012). Settlement patterns during the Earlier and Middle Stone Age around Langebaan Lagoon, Western Cape (South Africa). Quaternary International, 270, 15–29. Kaplan, J. M. (1989). 45,000 years of hunter-gatherer history in Natal as seen from Umhlatuzana Rock Shelter. South African Archaeological Society Goodwin Series, 6, 7–16. Kaplan, J. (1990). The Umhlatuzana Rock Shelter sequence: 100,000 years of Stone Age history. Natal Museum Journal of Humanities, 2, 1–94. Kappelman, J., Tewabe, D., Todd, L., Feseha, M., Kay, M., Kocurek, G., et al. (2014). Another unique river: A consideration of some of the characteristics of the trunk tributaries of the Nile River in northwestern Ethiopia in relationship to their aquatic food resources. Journal of Human Evolution, 77, 117–131. Kato, H., Elbeih, S., Iwasaki, E., Sefelnasr, A., Shalaby, A., Zaghloul, E., et al. (2014). The relationship between groundwater, landuse, and demography in Dakhla oasis, Egypt. Journal of Asian Network for GIS-based Historical Studies, 3, 10. Katoh, S., Nagaoka, S., WoldeGabriel, G., Renne, P., Snow, M. G., Beyene, Y., et al. (2000). Chronostratigraphy and correlation of the Plio-Pleistocene tephra layers of the Konso Formation, southern Main Ethiopian Rift, Ethiopia. Quaternary Science Reviews, 19(13), 1305–1317. 2070 J. C. Thompson et al. Kelly, A. J., & Harris, J. W. (1992). Recent findings of Middle Stone Age material from East Turkana. Nyame Akuma, 38, 29–34. Kiberd, P. (2006). Bundu Farm: A report on archaeological and palaeoenvironmental assemblages from a pan site in Bushmanland, Northern Cape, South Africa. The South African Archaeological Bulletin, 61(184), 189–201. Kibunjia, M. (1994). Pliocene archaeological occurrences in the Lake Turkana basin. Journal of Human Evolution, 27(1–3), 159–171. Kimbel, W. H., Walter, R. C., Johanson, D. C., Reed, K. E., Aronson, J. L., Assefa, Z., et al. (1996). Late Pliocene Homo and Oldowan tools from the Hadar Fromation (Kada Hadar Member), Ethiopia. Journal of Human Evolution, 31, 549–561. Klein, R. G. (1972). The late Quaternary mammalian fauna of Nelson Bay Cave (Cape Province, South Africa): Its implications for megafaunal extinctions and environmental and cultural change. Quaternary Research, 2(2), 135–142. Klein, R. G. (1974). Environment and subsistence of prehistoric man in the Southern Cape province, South Africa. World Archaeology, 5, 249–289. Klein, R. G. (1976). The mammalian fauna of the Klasies River Mouth Sites, Southern Cape Province, South Africa. South African Archaeological Bulletin, 31(123/124), 75–98. Klein, R. G. (1977). The mammalian fauna from the Middle and Later Stone Age (Later Pleistocene) levels of Border Cave, Natal Province, South Africa. South African Archaeological Bulletin, 32, 14–27. Klein, R. (1978a). The vertebrate fauna from the Buffelskloof Rock Shelter. South African Archaeological Bulletin, 33, 35–38. Klein, R. G. (1978b). Preliminary analysis of the mammalian fauna from the Redcliff stone age cave site, Rhodesia. Occassional Papers of the National Museums and Monuments of Rhodesia, 4, 74–80. Klein, R. G. (1978c). A preliminary report on the larger mammals from the Boomplaas Stone Age cave site, Cango Valley, Oudtshoorn District, South Africa. The South African Archaeological Bulletin, 33(127), 66–75. Klein, R. (1981). Archaeological implications of the mammalian fauna from the Later Stone Age cave site of Byneskranskop, southern Cape Province, South Africa. In Omnivorous primates: Gathering and hunting in human evolution (pp. 166–190). Columbia University Press. Klein, R. G. (1983a). Palaeoenvironmental implications of Quaternary large mammals in the fynbos region. In H. J. Deacon, Q. B. Hendey, & J. J. N. Lambrechts (Eds.), Fynbos Palaeoecology: A preliminary synthesis (South African National Scientific Programmes Report) (Vol. 75, pp. 116–138). Cooperative Scientific Programmes, Council for Scientific and Industrial Research. Klein, R. G. (1983b). The stone age prehistory of southern Africa. Annual Review of Anthropology, 12, 25–48. Klein, R. G. (1984). Later Stone Age faunal samples from Heuningneskrans Shelter (Transvaal) and Leopard’s Hill Cave (Zambia). The South African Archaeological Bulletin, 39, 109–116. Klein, R. G. (1988). The archaeological significance of animal bones from Acheulean sites in southern africa. African Archaeological Review, 6, 3–25. Klein, R. G. (1989). Why does skeletal part representation differ between smaller and larger bovids at Klasies River Mouth and other archaeological sites? Journal of Archaeological Science, 6, 363–381. Klein, R. G. (1992). The archeology of modern human origins. Evolutionary Anthropology: Issues, News, and Reviews, 1(1), 5–14. Klein, R. G., & Cruz-Uribe, K. (1983). Stone age population numbers and average tortoise size at Byneskranskop Cave 1 and Die Kelders Cave 1, Southern Cape Province, South Africa. South African Archaeological Bulletin, 38, 26–30. Klein, R. G., & Cruz-Uribe, K. (1987). Large mammal and tortoise bones from Eland’s Bay Cave and nearby sites, Western Cape Province, South Africa. In R. G. Klein, K. Cruz-Uribe, The Zooarchaeology of Pleistocene Africa 2071 J. Parkington, & M. Hall (Eds.), Papers in the Prehistory of the Western Cape, South Africa (pp. 132–164). BAR International Series. Klein, R. G., & Cruz-Uribe, K. (1996). Exploitation of large bovids and seals at Middle and Later Stone Age sites in South Africa. Journal of Human Evolution, 31, 315–334. Klein, R. G., & Cruz-Uribe, K. (2000a). Macromammals and reptiles. In L. S. Barham (Ed.), The Middle Stone Age of Zambia, South Central Africa (pp. 51–56). Bristol Western Academic & Specialist Press. Klein, R. G., & Cruz-Uribe, K. (2000b). Middle and Later Stone Age large mammal and tortoise remains from Die Kelders Cave 1, Western Cape Province, South Africa. Journal of Human Evolution, 38, 169–196. Klein, R. G., & Cruz-Uribe, K. (2016). Large mammal and tortoise bones from Elands Bay Cave (South Africa): Implications for Later Stone Age environment and ecology. Southern African Humanities, 29(1), 259–282. Klein, R. G., & Scott, K. (1986). Re-analysis of faunal assemblages from the Haua Fteah and other Late Quaternary archaeological sites in Cyrenaican Libya. Journal of Archaeological Science, 13, 515–542. Klein, R. G., Cruz-Uribe, K., & Beaumont, P. B. (1991). Environmental, ecological, and paleoanthropological implications of the late Pleistocene mammalian fauna from Equus Cave, Northern Cape Province, South Africa. Quaternary Research, 36(1), 94–119. Klein, R. G., Avery, G., Cruz-Uribe, K., Halkett, D., Hart, T., Milo, R. G., et al. (1999a). Duinefontein 2: An Acheulean site in the western cape province of South Africa. Journal of Human Evolution, 37(2), 153–190. Klein, R. G., Cruz-Uribe, K., & Milo, R. G. (1999b). Skeletal part representation in archaeofaunas: Comments on “Explaining the ‘Klasies Pattern’: Kua Ethnoarchaeology, the Die Kelders Middle Stone Age Archaeofauna, Long Bone Fragmentation and Carnivore Ravaging” by Bartram & Marean. Journal of Archaeological Science, 26, 1225–1234. Klein, R. G., Avery, G., Cruz-Uribe, K., Halkett, D., Parkington, J. E., Steele, T., et al. (2004). The Ysterfontein 1 Middle Stone Age site, South Africa, and early human exploitation of coastal resources. Proceedings of the National Academy of Sciences, USA, 101(16), 5708–5715. Klein, R. G., Avery, G., Cruz-Uribe, K., & Steele, T. E. (2007). The mammalian fauna associated with an archaic hominin skullcap and later Acheulean artifacts at Elandsfontein, Western Cape Province, South Africa. Journal of Human Evolution, 52(2), 164–186. Kleindienst, M. R. (1973). Excavations at site JK2, Olduvai Gorge, Tanzania, 1961-1962: The geological setting. Quaternaria, 17, 145–208. Kowalski, K., Szyndlar, Z., Mlynarski, M., Bocheński, Z., & Rzebik-Kowalska, B. (1993). Remains of small vertebrates from Bir Tarfawi and their paleoecological significance. In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 155–204). Plenum. Kuman, K. (1994). The archaeology of Sterkfontein – Past and present. Journal of Human Evolution, 27, 471–495. Kuman, K., Field, A. S., & Thackeray, J. F. (1997). Discovery of new artefacts at Kromdraai. South African Journal of Science, 93, 187–193. Kuman, K., Inbar, M., & Clarke, R. J. (1999). Palaeoenvironments and cultural sequence of the Florisbad Middle Stone Age hominid site, South Africa. Journal of Archaeological Science, 26, 1409–1425. Kusimba, S. B. (2001). The Early Later Stone Age in East Africa: Excavations and lithic assemblages from Lukenya Hill. African Archaeological Review, 18(2), 77–124. Kwekason, A., & Chami, F. A. (2003). The archaeology of Muleba, south west of Lake Nyanza: A preliminary report. In F. Chami, G. Pwiti, & C. Radimilahy (Eds.), Climate change, trade and modes of production in Sub-Saharan Africa (Studies in the African Past) (Vol. 3, pp. 59–85). Dar es Salaam University Press Ltd. Kyriacou, K., Parkington, J. E., Will, M., Kandel, A. W., & Conard, N. J. (2015). Middle and Later Stone Age shellfish exploitation strategies and coastal foraging at Hoedjiespunt and Lynch Point, Saldanha Bay, South Africa. Journal of Archaeological Science, 57, 197–206. 2072 J. C. Thompson et al. Lacruz, R. S., Brink, J. S., Hancox, P. J., Skinner, A. R., Herries, A., Schmid, P., et al. (2002). Palaeontology and geological context of a Middle Pleistocene faunal assemblage from the Gladysvale Cave, South Africa. Palaeontologia Africana, 38, 99–114. Lam, Y. M., Pearson, O. M., Marean, C. W., & Chen, X. (2003). Bone density studies in zooarchaeology. Journal of Archaeological Science, 30(12), 1701–1708. Leakey, L. S. B. (1943). The Industries of the Gorgora Rock Shelter, Lake Tana. Journal of the East Africa and Uganda Natural History, 17, 199–203. Leakey, L. S. B. (1965). Olduvai Gorge 1951–1961: A preliminary report on the geology and Fauna. Cambridge University Press. Leakey, M. D. (1971). Olduvai Gorge: Volume 3, excavations in Beds I and II, 1960–1963. Cambridge University Press. Leakey, M. D., & Roe, D. A. (1994). The Fauna. In M. D. Leakey & D. A. Roe (Eds.), Olduvai Gorge volume 5: Excavations in Beds III, IV, and the Masek Beds, 1968–1971 (Vol. 5, pp. 130–145). Cambridge University Press. Leakey, L. S. B., Evernden, J. F., & Curtis, G. H. (1961). The age of Bed I, Olduvai Gorge, Tanganyika. Nature, 191, 478–479. Leakey, M., Tobias, P. V., Martyn, J. E., & Leakey, R. E. F. (1970). An Acheulean Industry with prepared core technique and the discovery of a contemporary hominid mandible at Lake Baringo, Kenya. Proceedings of the Prehistoric Society, 35, 48–76. Leakey, M. D., Hay, R. L., Thurber, D. L., Protsch, R., & Berger, R. (1972). Stratigraphy, archaeology, and age of the Ndutu and Naisiusiu beds, Olduvai Gorge, Tanzania. World Archaeology, 3(3), 328–341. Lebatard, A.-E., Bourlès, D. L., Braucher, R., & Team, A. (2019). Absolute dating of an Early Paleolithic site in Western Africa based on the radioactive decay of in situ-produced 10Be and 26Al. Nuclear Instruments and Methods in Physics Research Section B: Beam Interactions with Materials and Atoms, 456, 169–179. Lehmann, U. (1957). Eine jungpleistozäne Wirbeltierfauna aus Ostafrika. Geologie des Schilfsandsteins, 26, 100–140. Leigh, E. G., & Butzer, K. W. (1968). Fossil molluscs from the Kom Ombo plain. In K. W. Butzer & C. L. Hansen (Eds.), Desert and River in Nubia (pp. 509–512). University of Wisconsin Press. Lesur, J., Faith, J. T., Bon, F., Dessie, A., Ménard, C., & Bruxelles, L. (2016). Paleoenvironmental and biogeographic implications of terminal Pleistocene large mammals from the Ziway–Shala Basin, Main Ethiopian Rift, Ethiopia. Palaeogeography, Palaeoclimatology, Palaeoecology, 449, 567–579. Loftus, E., Sealy, J., & Lee-Thorp, J. (2016). New radiocarbon dates and Bayesian models for Nelson Bay Cave and Byneskranskop 1: Implications for the South African Later Stone Age sequence. Radiocarbon, 58(2), 365–381. Loftus, E., Pargeter, J., Mackay, A., Stewart, B. A., & Mitchell, P. (2019). Late Pleistocene human occupation in the Maloti-Drakensberg region of southern Africa: New radiocarbon dates from Rose Cottage Cave and inter-site comparisons. Journal of Anthropological Archaeology, 56, 101–117. Louw, A. (1969). Bushman rock shelter, Ohrigstad, eastern Transvaal: A preliminary investigation, 1965. The South African Archaeological Bulletin, 24, 39–51. Lüdecke, T., Mulch, A., Kullmer, O., Sandrock, O., Thiemeyer, H., Fiebig, J., et al. (2016). Stable isotope dietary reconstructions of herbivore enamel reveal heterogeneous savanna ecosystems in the Plio-Pleistocene Malawi Rift. Palaeogeography, Palaeoclimatology, Palaeoecology, 459, 170–181. Lukich, V., Cowling, S., & Chazan, M. (2020). Palaeoenvironmental reconstruction of Kathu Pan, South Africa, based on sedimentological data. Quaternary Science Reviews, 230, 106153. Lupo, K. (2001). Archaeological skeletal part profiles and differential transport: An ethnoarchaeological example from the Hadza bone assemblages. Journal of Anthropological Archaeology, 20, 361–378. The Zooarchaeology of Pleistocene Africa 2073 Lyman, R. L. (2008). Quantitative paleozoology (Cambridge manuals in archaeology). Cambridge University Press. Mabulla, A. (1990). Preliminary report on an archaeological survey of the Ndutu Beds, Olduvai Gorge, Tanzania. Nyame Akuma, 33, 20–24. Mackay, A., Jacobs, Z., & Steele, T. E. (2015). Pleistocene archaeology and chronology of Putslaagte 8 (PL8) Rockshelter, Western Cape, South Africa. Journal of African Archaeology, 13(1), 71–98. Mackay, A., Cartwright, C. R., Heinrich, S., Low, M. A., Stahlschmidt, M. C., & Steele, T. E. (2020). Excavations at Klipfonteinrand reveal local and regional patterns of adaptation and interaction through MIS 2 in southern Africa. Journal of Paleolithic Archaeology, 3(3), 362–397. Maíllo-Fernández, J.-M., Marín, J., Solano-Megías, I., Uribelarrea, D., Martín-Perea, D., Aramendi, J., et al. (2019). Victoria Cabrera Site: A Middle Stone Age site at Olduvai Gorge, Tanzania. Quaternary International, 526, 129–154. Mana, S., Hemming, S., Kent, D. V., & Lepre, C. J. (2019). Temporal and stratigraphic framework for paleoanthropology sites within East-Central Area 130, Koobi Fora, Kenya. Frontiers in Earth Science, 7, 230. Marean, C. W. (1992a). Hunter to herder: Large mammal remains from the hunter-gatherer occupation at Enkapune Ya Muto rockshelter (Central Rift, Kenya). African Archaeological Review, 10, 65–127. Marean, C. W. (1992b). Implications of late Quaternary mammalian fauna from Lukenya Hill (south-central Kenya) for paleoenvironmental change and faunal extinctions. Quaternary Research, 37, 239–255. Marean, C. W., & Cleghorn, N. (2003). Large mammal skeletal element transport: Applying foraging theory in a complex taphonomic system. Journal of Taphonomy, 1, 15–42. Marean, C. W., & Gifford-Gonzalez, D. (1991). Late Quaternary extinct ungulates of East Africa and paleoenvironmental implications. Nature, 350, 418–420. Marean, C. W., Spencer, L. M., Blumenschine, R. J., & Capaldo, S. D. (1992). Captive hyena bone choice and destruction, the Schlepp effect, and Olduvai archaeofaunas. Journal of Archaeological Science, 19, 101–121. Marean, C. W., Mudida, N., & Reed, K. E. (1994). Holocene paleoenvironmental change in the Kenyan Central Rift as indicated by micromammals from Enkapune Ya Muto rockshelter. Quaternary Research, 41, 376–389. Marean, C. W., Abe, Y., Frey, C. J., & Randall, R. C. (2000). Zooarchaeological and taphonomic analysis of the Die Kelders Cave 1 Layers 10 and 11 Middle Stone Age larger mammal fauna. Journal of Human Evolution, 38(1), 197–233. Marean, C. W., Abe, Y., Nilssen, P. J., & Stone, E. C. (2001). Estimating the minimum number of skeletal elements (MNE) in zooarchaeology: A review and a new image-analysis GIS approach. American Antiquity, 66, 333–348. Marean, C. W., Bar-Matthews, M., Fisher, E., Goldberg, P., Herries, A., Karkanas, P., et al. (2010). The stratigraphy of the Middle Stone Age sediments at Pinnacle Point Cave 13B (Mossel Bay, Western Cape Province, South Africa). Journal of Human Evolution, 59(3–4), 234–255. Marean, C. W., Cowling, R. M., & Franklin, J. (2020). The Palaeo-Agulhas Plain: Temporal and spatial variation in an extraordinary extinct ecosystem of the Pleistocene of the Cape Floristic Region. Quaternary Science Reviews, 235, 106161. Marks, A. E. (1987). Terminal Pleistocene and Holocene hunters and gatherers in the eastern Sudan. African Archaeological Review, 5(1), 79–92. Marks, T. P. (2018). Bedtime for the Middle Stone Age: Land use, strategic foraging, and lithic technology at the end of the Pleistocene in the Namib Desert. PhD dissertation, University of Iowa. MartÍn-Perea, D. M., MaÍllo-FernÁndez, J. M., Medialdea, A., MarÍn, J., Solano-MegÍas, I., Gidna, A., et al. (2020). Revisiting an old profile: An updated geoarchaeological study at Nasera Rockshelter (Tanzania). Journal of Quaternary Science, 35(7), 951–960. 2074 J. C. Thompson et al. Masele, F. (2020). Testing meat-eating by Middle Stone Age hominins at Loiyangalani openair site in Serengeti National Park, Tanzania. Archaeological and Anthropological Sciences, 12(2), 55. Mason, R. J., Dart, R. A., & Kitching, J. W. (1958). Bone tools at the Kalkbank Middle Stone Age site and the Makapansgat australopithecine locality, Central Transvaal. The South African Archaeological Bulletin, 13, 85–116. Matthews, T., Parkington, J. E., & Denys, C. (2006). The taphonomy of the micromammals from the late Middle Pleistocene site of Hoedjiespunt 1 (Cape Province, South Africa). Journal of Taphonomy, 4(1), 11–26. Matthews, T., Marean, C. W., & Nilssen, P. J. (2009). Micromammals from the Middle Stone Age (92–167 ka) at Cave PP13B, Pinnacle Point, south coast, South Africa. Palaeontologia African, 44, 112–120. Matthews, T., Marean, C. W., & Cleghorn, N. (2019). Past and present distributions and community evolution of Muridae and Soricidae from MIS 9 to MIS 1 on the edge of the PalaeoAgulhas Plain (south coast, South Africa). Quaternary Science Reviews, 235, 105774. Matthews, T., Marean, C. W., & Cleghorn, N. (2020). Past and present distributions and community evolution of Muridae and Soricidae from MIS 9 to MIS 1 on the edge of the PalaeoAgulhas Plain (south coast, South Africa). Quaternary Science Reviews, 235, 105774. Maurin, T., Bertran, P., Delagnes, A., & Boisserie, J.-R. (2017). Early hominin landscape use in the Lower Omo Valley, Ethiopia: Insights from the taphonomical analysis of Oldowan occurrences in the Shungura Formation (Member F). Journal of Human Evolution, 111, 33–53. McBrearty, S. (1981). Songhor: A Middle Stone Age site in western Kenya. Quaternaria, 23, 171–190. McBrearty, S. (1992). Sangoan technology and habitat at Simbi, Kenya. Nyame Akuma, 38, 34–39. McBurney, C. B. (1967). The Haua Fteah (Cyrenaica) and the Stone Age of the South-East Mediterranean. Cambridge University Press. McBurney, C. B. M., & Hey, R. W. (1955). A record of two seasons’ geological and archaelogical fieldwork in the Gebel Akhdar Hills, with a summary of prehistoric finds from neighbouring territories. In C. B. M. McBurney & R. W. Hey (Eds.), Prehistory and Pleistocene Geology in Cyrenaican Libya (Vol. 4). Cambridge Unviersity Press. McCall, G. S., Marks, T. P., Thomas, J. T., Eller, M., Horn, S. W., Horowitz, R. A., et al. (2011). Erb tanks: A Middle and Later Stone Age rockshelter in the central Namib Desert, western Namibia. PaleoAnthropology, 2011, 398–421. McDermott, F., Stringer, C., Grün, R., Williams, C. T., Din, V. K., & Hawkesworth, C. J. (1996). New late-Pleistocene uranium–thorium and ESR dates for the Singa hominid (Sudan). Journal of Human Evolution, 31(6), 507–516. McDougall, I., & Brown, F. H. (2006). Precise 40Ar/39Ar geochronology for the upper Koobi Fora formation, Turkana Basin, northern Kenya. Journal of the Geological Society, 163(1), 205–220. McHenry, L. J., & Stanistreet, I. G. (2018). Tephrochronology of Bed II, Olduvai Gorge, Tanzania, and placement of the Oldowan–Acheulean transition. Journal of Human Evolution, 120, 7–18. McKee, J. K. (1996). Faunal turnover patterns in the Pliocene and Pleistocene of southern Africa. South African Journal of Science, 92(3), 111–113. McPherron, S. P., Alemseged, Z., Marean, C. W., Wynn, J. G., Reed, D., Geraads, D., et al. (2010). Evidence for stone-tool-assisted consumption of animal tissues before 3.39 million years ago at Dikika, Ethiopia. Nature, 466(7308), 857–860. Mehlman, M. J. (1977). Excavations at Nasera Rock, Tanzania. Azania, 12, 111–118. Mehlman, M. (1979). Mumba-Höhle revisited: The relevance of a forgotten excavation to some current issues in east African prehistory. World Archaeology, 11(1), 80–94. Meier, R., Sahnouni, M., Medig, M., & Derradji, A. (2003). Human skull from the Taza Locality, Jijel, Algeria. Anthropologischer Anzeiger, 61(2), 129–140. The Zooarchaeology of Pleistocene Africa 2075 Mercader, J., & Brooks, A. S. (2001). Across Forests and Savannas: Later Stone Age Assemblages from Ituri and Semliki, Democratic Republic of Congo. Journal of Anthropological Research, 57(2), 197–217. Mercader, J., & Sillé, P. (2013). Middle and Later Stone Age sites from Sofala, Gorongosa (Central Mozambique). Nyame Akuma, 80, 1–13. Mercader, J., Asmeromb, Y., Bennett, T., Raja, M., & Skinner, A. (2009). Initial excavation and dating of Ngalue Cave: A Middle Stone Age site along the Niassa Rift, Mozambique. Journal of Human Evolution, 57, 63–74. Merrick, H. V. (1975). Change in later Pleistocene Lithic Industries in Eastern Africa. PhD dissertation, University of California at Berkeley. Merritt, S. R., & Davis, K. M. (2017). Diagnostic properties of hammerstone-broken long bone fragments, specimen identifiability, and Early Stone Age butchered assemblage interpretation. Journal of Archaeological Science, 85, 114–123. Merritt, S. R., Mavuso, S., Cordiner, E. A., Fetchenhier, K., & Greiner, E. (2018). FwJj70 – A potential Early Stone Age single carcass butchery locality preserved in a fragmentary surface assemblage. Journal of Archaeological Science: Reports, 20, 736–747. Merzoug, S. (2005). Comportements de subsistance des Ibéromaurusiens d’après l’analyse archéozoologique des mammifères des sites de Tamar Hat, Taza 1 et Columnata (Algérie). Muséum national d’histoire naturelle. Merzoug, S. (2012). Essai d’interprétation du statut économique du Megaceroides algericus durant l’Ibéromaurusien dans le massif des Babors (Algérie). Quaternaire. Revue de l’Association française pour l’étude du Quaternaire, 23(2), 3. Merzoug, S., & Sari, L. (2008). Re-examination of the Zone I material from Tamar Hat (Algeria): Zooarchaeological and technofunctional analyses. African Archaeological Review, 25(1/2), 57–73. Michel, P. (1988). Un nouveau Mellivorinae (Carnivora, Mustelidae) du Pleistocène de Bouknadel (Région de Rabat, Maroc): Mellivora carolae n. sp. Comptes Rendus de l’Académie des Sciences, Paris, 306(13), 935–938. Michel, P. (1990). Contribution à l’étude paléontologique des Vertébrés fossiles du Quaternaire marocain à partir de sites du Maroc atlantique, central et oriental. Institut de paléontologie humaine, Museum national d’histoire naturelle. Michel, P. (1992). Pour une meilleure connaissance du Quaternaire Continental Marocain: les vertébrés fossiles du Maroc Atlantique, Central et Oriental. L’Anthropologie, 96(4), 643–656. Michel, P., Campmas, É., Stoetzel, E., Nespoulet, R., El Hajraoui, M. A., & Amani, F. (2009). La macrofaune du Pléistocène supérieur d’El Harhoura 2 (Témara, Maroc): données préliminaires. L’Anthropologie, 113(2), 283–312. Millard, A. R. (2008). A critique of the chronometric evidence for hominid fossils: I. Africa and the Near East 500–50ka. Journal of Human Evolution, 54(6), 848–874. Miller, C. E., Goldberg, P., & Berna, F. (2013). Geoarchaeological investigations at Diepkloof Rock Shelter, Western Cape, South Africa. Journal of Archaeological Science, 40(9), 3432–3452. Miller, J. M., Keller, H. M., Heckel, C., Kaliba, P. M., & Thompson, J. C. (2021). Approaches to land snail shell bead manufacture in the Early Holocene of Malawi. Archaeological and Anthropological Sciences, 13(3), 37. Milo, R. G. (1998). Evidence for hominid predation at Klasies River Mouth, South Africa, and its implications for the behavior of early modern humans. Journal of Archaeological Science, 25, 99–133. Mitchell, J. P. (1994). Understanding the MSA/LSA transition: The pre-20 000 BP assemblages from new excavations at Sehonghong Rock Shelter, Lesotho. Southern African Field Archaeology, 3(1), 15–25. Moeyersons, J., Vermeersch, P., & Van Peer, P. (2002). Dry cave deposits and their palaeoenvironmental significance during the last 115 ka, Sodmein Cave, Red Sea Mountains, Egypt. Quaternary Science Reviews, 21(7), 837–851. 2076 J. C. Thompson et al. Monahan, C. M. (1996). New zooarchaeological data from Bed II, Olduvai Gorge, Tanzania: Implications for hominid behavior in the Early Pleistocene. Journal of Human Evolution, 31, 93–128. Morais, J. (1984). Mozambican archaeology: Past and present. African Archaeological Review, 2, 113–128. Morel, J. (1974a). La station éponyme de l’Oued Djebbana à Bir-el-ater (est algérien). Contribution à la connaissance de son industrie et de sa faune. L’Anthropologie, 78, 53–80. Morel, J. (1974b). Nouvelles datations absolues de formations littorales et de gisements préhistoriques de l’Est algérien. Bulletin de la Société Préhistorique Française, 71(4), 103–105. Morgan, L. E., Renne, P. R., Kieffer, G., Piperno, M., Gallotti, R., & Raynal, J.-P. (2012). A chronological framework for a long and persistent archaeological record: Melka Kunture, Ethiopia. Journal of Human Evolution, 62(1), 104–115. Morin, E. (2007). Fat composition and Nunamiut decision-making: A new look at the marrow and bone grease indices. Journal of Archaeological Science, 34, 69–82. Moutsiou, T. (2011). The Obsidian Evidence for the Scale of Social Life during the Palaeolithic. PhD dissertation, Royal Holloway, University of London. Mussi, M., Altamura, F., Macchiarelli, R., Melis, R. T., & Spinapolice, E. E. (2014). Garba III (Melka Kunture, Ethiopia): A MSA site with archaic Homo sapiens remains revisited. Quaternary International, 343, 28–39. Mussi, M., Altamura, F., Bonnefille, R., De Rita, D., & Melis, R. T. (2016). The environment of the Ethiopian highlands at the Mid Pleistocene Transition: Fauna, flora and hominins in the 850-700 ka sequence of Gombore II (Melka Kunture). Quaternary Science Reviews, 149, 259–268. Nami, M., & Moser, J. (Eds.). (2010) La grotte d’Ifri n’Ammar: le Paléolithique Moyen (Vol. 2, pp. 307–314). Kommission für Archäologie Außereuropäischer Kulturen. Negash, A., Brown, F., & Nash, B. (2011). Varieties and sources of arefactual obsidian in the Middle Stone Age of the Middle Awash, Ethiopia. Archaeometry, 53(4), 661–673. Nespoulet, R., Debénath, A., El Hajraoui, M., Michel, P., Campmas, E., Oujaa, A., et al. (2008). Le contexte archéologique des restes humains atériens de la région de Rabat-Témara (Maroc): apport des fouilles des grottes d’El Mnasra et d’El Harhoura 2. In Actes des quatrièmes rencontres des Quaternaristes Marocains (RQM4), Le Quaternaire marocain dans son contexte méditerranéen (pp. 356–375). de la Faculté des Sciences d’Oujda. Nicoll, K. (2018). A revised chronology for Pleistocene paleolakes and Middle Stone Age–Middle Paleolithic cultural activity at Bîr Tirfawi–Bîr Sahara in the Egyptian Sahara. Quaternary International, 463, 18–28. Njau, J. K., Herrmann, E. W., Ruck, L., Pante, M., Farrugia, P., Toth, N., et al. (2020). Core stratigraphy constrains Bed IV archaeological record at HEB site, Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology, 552, 109773. Nouet, J., Chevallard, C., Farre, B., Nehrke, G., Campmas, E., Stoetzel, E., et al. (2015). Limpet shells from the Aterian level 8 of El Harhoura 2 cave (Témara, Morocco): Preservation state of crossed-foliated layers. PLoS One, 10(9), e0137162. O’Regan, H. J., Bishop, L. C., Lamb, A., Elton, S., & Turner, A. (2005). Large mammal turnover in Africa and the Levant between 1.0 and 0.5 Ma. Geological Society, London, Special Publications, 247(1), 231–249. Oakley, K. P., Campbell, B. G., & Molleson, T. I. (1977). Catalogue of Fossil Hominids: Africa (Vol. 1, 2n ed.). British Museum (Natural History). O’Brien, T. P. (1939). The prehistory of Uganda protectorate. Cambridge University Press. Opperman, H. (1978). Excavations in the Buffelskloof rock shelter near Calitzdorp, southern Cape. The South African Archaeological Bulletin, 33(127), 18–38. Opperman, H. (1988). An excavation of a Middle Stone Age deposit in Grassridge Rockshelter, Sterkstroom District, Cape Province. Fort Hare Papers, 9, 51–61. The Zooarchaeology of Pleistocene Africa 2077 Opperman, H., & Heydenrych, B. (1990). A 22 000 year-old Middle Stone Age camp site with plant food remains from the north-eastern Cape. The South African Archaeological Bulletin, 45, 93–99. Organista, E., Arriaza, M. C., Barba, R., Gidna, A., Ortega, M. C., Uribelarrea, D., et al. (2019). Taphonomic analysis of the level 3b fauna at BK, Olduvai Gorge. Quaternary International, 526, 116–128. Osypińska, M., & Osypiński, P. (2016). Animal exploitation and behaviour of the latest middle Stone age societies in the middle Nile Valley: Archaeozoological and Taphonomic analysis of late Pleistocene Fauna from the Affad Basin, Sudan. African Archaeological Review, 33(2), 107–127. Osypinski, P., Morley, M. W., Osypinska, M., & Kotarba-Morley, A. M. (2016). Affad 23: Settlement structures and palaeoenvironments in the terminal Pleistocene of the Middle Nile Valley, Sudan. Antiquity, 90(352), 894–913. Outram, A. K. (2001). FOCUS: The scapula representation could be the key: A further contribution to the ‘Klasies Pattern’ debate. Journal of Archaeological Science, 28, 1259–1263. Panera, J., Rubio-Jara, S., Domínguez-Rodrigo, M., Yravedra, J., Méndez-Quintas, E., PérezGonzález, A., et al. (2019). Assessing functionality during the early Acheulean in level TKSF at Thiongo Korongo site (Olduvai Gorge, Tanzania). Quaternary International, 526, 77–98. Pante, M. C. (2013). The larger mammal fossil assemblage from JK2, Bed III, Olduvai Gorge, Tanzania: Implications for the feeding behavior of Homo erectus. Journal of Human Evolution, 64(1), 68–82. Pante, M. C., & de la Torre, I. (2018). A hidden treasure of the Lower Pleistocene at Olduvai Gorge, Tanzania: The Leakey HWK EE assemblage. Journal of Human Evolution, 120, 114–139. Pante, M. C., Muttart, M. V., Keevil, T. L., Blumenschine, R. J., Njau, J. K., & Merritt, S. R. (2017). A new high-resolution 3-D quantitative method for identifying bone surface modifications with implications for the Early Stone Age archaeological record. Journal of Human Evolution, 102, 1–11. Pante, M. C., Njau, J. K., Hensley-Marschand, B., Keevil, T. L., Martín-Ramos, C., Peters, R. F., et al. (2018). The carnivorous feeding behavior of early Homo at HWK EE, Bed II, Olduvai Gorge, Tanzania. Journal of Human Evolution, 120, 215–235. Pante, M., de la Torre, I., d’Errico, F., Njau, J., & Blumenschine, R. (2020). Bone tools from Beds II–IV, Olduvai Gorge, Tanzania, and implications for the origins and evolution of bone technology. Journal of Human Evolution, 148, 102885. Pargeter, J., Loftus, E., & Mitchell, P. (2017). New ages from Sehonghong rock shelter: Implications for the late Pleistocene occupation of highland Lesotho. Journal of Archaeological Science: Reports, 12, 307–315. Pargeter, J., Loftus, E., Mackay, A., Mitchell, P., & Stewart, B. (2018). New ages from Boomplaas Cave, South Africa, provide increased resolution on late/terminal Pleistocene human behavioural variability. Azania: Archaeological Research in Africa, 53(2), 156–184. Parkinson, J. A. (2018). Revisiting the hunting-versus-scavenging debate at FLK Zinj: A GIS spatial analysis of bone surface modifications produced by hominins and carnivores in the FLK 22 assemblage, Olduvai Gorge, Tanzania. Palaeogeography, Palaeoclimatology, Palaeoecology, 511, 29–51. Paulissen, E., & Vermeersch, P. M. (1987). Earth, man and climate in the Egyptian Nile Valley during the Pleistocene. In A. Close (Ed.), Prehistory of Arid North Africa (Vol. 29). SMU Press. Pazdur, A., Pazdur, M. F., & Zastawny, A. (1994). Gliwice radiocarbon dates XII. Radiocarbon, 36(2), 281–302. Peters, J. (1989). Faunal remains and environmental change in central and eastern Sudan from terminal Pleistocene to middle Holocene times. AWLSK, 51(4), 123–148. Peters, J. (1990). Late Pleistocene Hunter-gatherers at Ishango (Eastern Zaire): The faunal evidence. Revue de Paleobiologie, 9, 73–112. Phillips, J. L. (1988). The Upper Paleolithic of the Wadi Feiran, Southern Sinai. Paléorient, 14, 183–200. 2078 J. C. Thompson et al. Phillips, J. (1994). The upper Paleolithic chronology of the Levant and the Nile Valley. In Late Quaternary chronology and paleoclimates of the Eastern Mediterranean (pp. 169–176). Radiocarbon Press. Phillipson, D. W. (1976). The Prehistory of Eastern Zambia. British Institute in Eastern Africa. Pickering, T. R. (1999). Taphonomic interpretations of the Sterkfontein Early Hominid Site (Gauteng, South Africa) reconsidered in light of recent evidence. PhD dissertation, University of Wisconsin. Pickering, T. R., Marean, C. W., & Domínguez-Rodrigo, M. (2003). Importance of limb bone shaft fragments in zooarchaeology: A response to “On in situ attrition and vertebrate body part profiles” (2002), by M.C. Stiner. Journal of Archaeological Science, 30(11), 1469–1482. Pickering, R., Hancox, P. J., Lee-Thorp, J. A., Grün, R., Mortimer, G. E., McCulloch, M., et al. (2007). Stratigraphy, U-Th chronology, and paleoenvironments at Gladysvale Cave: Insights into the climatic control of South African hominin-bearing cave deposits. Journal of Human Evolution, 53(5), 602–619. Pickering, T. R., Egeland, C. P., Domínguez-Rodrigo, M., Brain, C., & Schnell, A. G. (2008). Testing the “shift in the balance of power” hypothesis at Swartkrans, South Africa: Hominid cave use and subsistence behavior in the early Pleistocene. Journal of Anthropological Archaeology, 27(1), 30–45. Pickering, R., Herries, A. I. R., Woodhead, J. D., Hellstrom, J. C., Green, H. E., Paul, B., et al. (2019). U–Pb-dated flowstones restrict South African early hominin record to dry climate phases. Nature, 565(7738), 226–229. Pickford, M. (1990). Some fossiliferous Plio-Pleistocene cave systems of Ngamiland, Botswana. Botswana Notes and Records, 22, 1–15. Piperno, M., Collina, C., Gallotti, R., Raynal, J.-P., Kieffer, G., Le Bourdonnec, F.-X., et al. (2009). Obsidian exploitation and utilization during the Oldowan at Melka Kunture (Ethiopia). In E. Hovers & D. R. Braun (Eds.), Interdisciplinary approaches to the Oldowan (pp. 111–128). Springer. Plug, I., & Badenhorst, S. (2001). The distribution of macromammals in Southern Africa over the past 30 000 years as reflected in animal remains from archaeological sites (Transvaal Museum Monographs, Vol. 1) (Vol. 12). Transvaal Museum. Plug, I., & Clark, J. L. (2008). In the air: A preliminary report on the birds from Sibudu Cave, KwaZulu-Natal, South Africa. Goodwin Series, 10, 133–142. Plug, I., & Engela, R. (1992). The macrofaunal remains from recent excavations at Rose Cottage Cave, Orange Free State. The South African Archaeological Bulletin, 47, 16–25. Plug, I., & Mitchell, P. (2008a). Fishing in the Lesotho highlands: 26,000 years of fish exploitation, with special reference to Sehonghong Shelter. Journal of African Archaeology, 6(1), 33–55. Plug, I., & Mitchell, P. (2008b). Sehonghong: Hunter-gatherer utilization of animal resources in the highlands of Lesotho. Annals of the Transvaal Museum, 45, 31–53. Plummer, T., Bishop, L. C., Ditchfield, P., & Hicks, J. (1999). Research on Late Pliocene Oldowan Sites at Kanjera South, Kenya. Journal of Human Evolution, 36(2), 151–170. Plummer, T. W., Ditchfield, P. W., Bishop, L. C., Kingston, J. D., Ferraro, J. V., Braun, D. R., et al. (2009). Oldest evidence of toolmaking hominins in a grassland-dominated ecosystem. PLoS One, 4(9), e7199. Pobiner, B. L. (2020). The zooarchaeology and paleoecology of early hominin scavenging. Evolutionary Anthropology: Issues, News, and Reviews, 29(2), 68–82. Pobiner, B. L., Rogers, M. J., Monahan, C. M., & Harris, J. W. K. (2008). New evidence for hominin carcass processing strategies at 1.5 Ma, Koobi Fora, Kenya. Journal of Human Evolution, 55(1), 103–130. Porraz, G., & Val, A. (2019). Heuningneskrans and the Stone Age sequence of the Ohrigstad river catchment on the eastern border of the great escarpment, Limpopo Province, South Africa. South African Archaeological Bulletin, 74, 46–55. Porraz, G., Schmid, V. C., Miller, C. E., Tribolo, C., Cartwright, C. C., Charrié-Duhaut, A., et al. (2016). Update on the 2011 excavation at Elands Bay Cave (South Africa) and the Verlorenvlei Stone Age. Southern African Humanities, 29(1), 33–68. The Zooarchaeology of Pleistocene Africa 2079 Potts, R. (1988). Early hominid activities at Olduvai. Aldine de Gruyter. Potts, R., & Shipman, P. (1981). Cutmarks made by stone tools on bones from Olduvai Gorge, Tanzania. Nature, 291, 577–580. Potts, R., Shipman, P., & Ingall, E. (1988a). Taphonomy, paleoecology, and hominids of Lainyamok, Kenya. Journal of Human Evolution, 17(6), 597–614. Potts, R., Shipman, P., & Ingall, E. (1988b). Taphonomy, paleoecology, and hominids of Lainyamok, Kenya. Journal of Human Evolution, 17, 597–614. Potts, R., Behrensmeyer, A. K., Faith, J. T., Tryon, C. A., Brooks, A. S., Yellen, J. E., et al. (2018). Environmental dynamics during the onset of the Middle Stone Age in eastern Africa. Science, 360(6384), 86–90. Prendergast, M. E., & Beyin, A. (2018). Fishing in a fluctuating landscape: Terminal Pleistocene and early Holocene subsistence strategies in the Lake Turkana Basin, Kenya. Quaternary International, 471, 203–218. Prendergast, M. E., Luque, L., Domínguez-Rodrigo, M., Diez-Martín, F., Mabulla, A. Z. P., & Barba, R. (2007). New excavations at Mumba Rockshelter, Tanzania. Journal of African Archaeology, 5(2), 217. Prendergast, M. E., Rouby, H., Punnwong, P., Marchant, R., Crowther, A., Kourampas, N., et al. (2016). Continental Island formation and the archaeology of Defaunation on Zanzibar, eastern Africa. PLoS One, 11(2), e0149565. Prendergast, M. E., Buckley, M., Crowther, A., Frantz, L., Eager, H., Lebrasseur, O., et al. (2017). Reconstructing Asian faunal introductions to eastern Africa from multi-proxy biomolecular and archaeological datasets. PLoS One, 12(8), e0182565. Quade, J., Levin, N. E., Simpson, S. W., Butler, R., McIntosh, W. C., Semaw, S., et al. (2008). The geology of Gona, Afar, Ethiopia. Geological Society of America Bulletin, 446, 1–31. Ramos, J., Bernal, D., Domínguez-Bella, S., Calado, D., Ruiz, B., Gil, M. J., et al. (2008). The Benzú rockshelter: A middle Palaeolithic site on the North African coast. Quaternary Science Reviews, 27(23–24), 2210–2218. Ramos-Muñoz, J., Cantillo-Duarte, J. J., Bernal-Casasola, D., Barrena-Tocino, A., DomínguezBella, S., Vijande-Vila, E., et al. (2016). Early use of marine resources by Middle/Upper Pleistocene human societies: The case of Benzú rockshelter (northern Africa). Quaternary International, 407, 6–15. Raynal, J.-P., & Occhietti, S. (2012). Amino chronology and an earlier age for the Moroccan Aterian. In J.-J. Hublin & S. P. McPherron (Eds.), Modern origins: A north African perspective (pp. 79–90). Springer. Raynal, J. P., Sbihi Alaoui, F. Z., Geraads, D., Magoga, L., & Mohi, A. (2001). The earliest occupation of North-Africa: The Moroccan perspective. Quaternary International, 75(1), 65–75. Raynal, J.-P., Sbihi Alaoui, F.-Z., Magoga, L., Mohib, A., & Zouak, M. (2002). Casablanca and the earliest occupation of North Atlantic Morocco. Paléorivages de Casablanca. Stratigraphie et Préhistoire ancienne au Maroc atlantique, 13(1), 65–77. Raynal, J.-P., Amani, F., Geraads, D., Graoui, M. E., Magoga, L., Texier, J.-P., et al. (2008). La Grotte des Félins, site paléolithique du Pléistocène supérieur à Dar Bouazza (Maroc). L’Anthropologie, 112(2), 182–200. Raynal, J.-P., Sbihi-Alaoui, F.-Z., & Mohib, A. (2010). Bilan des recherches récentes sur le Paléolithique de Casablanca (Maroc). Les nouvelles de l’archéologie, 120–121, 102–109. Raynal, J.-P., Mohib, A., & Lefévre, D. (2016). Casablanca des origins. In J.-P. Raynal & A. Mohib (Eds.), Préhistoire de Casablanca 1 – La Grotte des Rhinocéros (fouilles 1991 et 1996). Villes et sites archéologiques du Maroc (VESAM) (Royaume du Maroc) (Vol. VI, pp. 11–33). Ministère de la Culture, Institut National des Sciences de l’Archéologie et du Patrimoine. Reade, H., O’Connell, T. C., Barker, G., & Stevens, R. E. (2016). Pleistocene and Holocene palaeoclimates in the Gebel Akhdar (Libya) estimated using herbivore tooth enamel oxygen isotope compositions. Quaternary International, 404, 150–162. Rector, A. L., & Reed, K. E. (2010). Middle and Late Pleistocene faunas of Pinnacle Point and their paleoecological implications. Journal of Human Evolution, 59(3–4), 340–357. 2080 J. C. Thompson et al. Reed, D. N., & Geraads, D. (2012). Evidence for a late Pliocene faunal transition based on a new rodent assemblage from Oldowan locality Hadar AL 894, Afar Region, Ethiopia. Journal of Human Evolution, 62(3), 328–337. Reid, R. E. B., Jones, M., Brandt, S., Bunn, H., & Marshall, F. (2019). Oxygen isotope analyses of ungulate tooth enamel confirm low seasonality of rainfall contributed to the African Humid Period in Somalia. Palaeogeography, Palaeoclimatology, Palaeoecology, 534, 109272. Renne, P. R., Balco, G., Ludwig, K. R., Mundil, R., & Min, K. (2011). Response to the comment by WH Schwarz et al. on “Joint determination of 40K decay constants and 40Ar∗/40K for the Fish Canyon sanidine standard, and improved accuracy for 40Ar/39Ar geochronology” by PR Renne et al. (2010). Geochimica et Cosmochimica Acta, 75(17), 5097–5100. Reygasse, M. (1921–1922). Etudes de Palethnologie Maghrébine (deuxième série). Recueil des Notices et Mémoires de la Société Archéologique Historique et Géographique de Constantine, 53, 159–204. Reynard, J. P., & Henshilwood, C. S. (2019). Environment versus behaviour: Zooarchaeological and taphonomic analyses of fauna from the still bay layers at Blombos cave, South Africa. Quaternary International, 500, 159–171. Reynard, J. P., Henshilwood, C. S., & Badenhorst, S. (2014). Inferring animal size from the unidentified long bones from the Middle Stone Age layers at Blombos Cave, South Africa. Annals of the Ditsong National Museum of Natural History, 4(1), 9–25. Reynard, J. P., Discamps, E., Badenhorst, S., Van Niekerk, K., & Henshilwood, C. S. (2016). Subsistence strategies in the southern Cape during the Howiesons Poort: Taphonomic and zooarchaeological analyses of Klipdrift Shelter, South Africa. Quaternary International, 404, 2–19. Reynolds, S. C., & Kibii, J. M. (2011). Sterkfontein at 75: Review of paleoenvironments, fauna, dating and archaeology from the hominin site of Sterkfontein (Gauteng Province, South Africa). Palaeontologia Africana, 46, 59–88. Reynolds, S. C., Vogel, J. C., Clarke, R. J., & Kuman, K. A. (2003). Preliminary results of excavations at Lincoln Cave, Sterkfontein, South Africa. South African Journal of Science, 99, 286. Reynolds, S. C., Clarke, R. J., & Kuman, K. A. (2007). The view from the Lincoln Cave: Mid-to late Pleistocene fossil deposits from Sterkfontein hominid site, South Africa. Journal of Human Evolution, 53(3), 260–271. Rhodes, E. J., Singarayer, J. S., Raynal, J.-P., Westaway, K. E., & Sbihi-Alaoui, F. Z. (2006). New age estimates for the Palaeolithic assemblages and Pleistocene succession of Casablanca, Morocco. Quaternary Science Reviews, 25, 2569–2585. Richter, D., & Wagner, G. A. (2015). Chronometric methods in paleoanthropology. In W. Henke & I. Tattersall (Eds.), Handbook of paleoanthropology (pp. 317–350). Springer-Verlag. Richter, D., Moser, J., Nami, M., Eiwanger, J., & Mikdad, A. (2010). New chronometric data from Ifri n’Ammar (Morocco) and the chronostratigraphy of the Middle Palaeolithic in the Western Maghreb. Journal of Human Evolution, 59(6), 672–679. Richter, D., Grün, R., Joannes-Boyau, R., Steele, T. E., Amani, F., Rué, M., et al. (2017). The age of the hominin fossils from Jebel Irhoud, Morocco, and the origins of the Middle Stone Age. Nature, 546(7657), 293–296. Robbins, L. H. (1989). The Middle Stone Age of Kudiakam pan. Botswana Notes and Records, 20, 41–50. Robbins, L. H. (1990). The depression site – A Stone Age sequence in the Northwest Kalahari Desert, Botswana. National Geographic Research, 6(3), 329–338. Robbins, L. H., & Murphy, M. L. (1998). The Early and Middle Stone Age. In P. Lane, A. Reid, & A. Segobye (Eds.), Ditswa Mmung: The archaeology of Botswana (pp. 50–64). Pula Press. Robbins, L. H., Murphy, M. L., Campbell, A., & Brook, G. A. (1996a). Excavations at the Tsodilo Hills Rhino Cave. Botswana Notes and Records, 28, 23–45. Robbins, L. H., Murphy, M. L., Stevens, N. J., Brook, G. A., Ivester, A. H., Haberyan, K. A., et al. (1996b). Paleoenvironment and archaeology of Drotsky’s cave: Western Kalahari desert, Botswana. Journal of Archaeological Science, 23(1), 7–22. The Zooarchaeology of Pleistocene Africa 2081 Robbins, L. H., Murphy, M. L., Brook, G. A., Ivester, A. H., Campbell, A. C., Klein, R. G., et al. (2000). Archaeology, palaeoenvironment, and chronology of the Tsodilo Hills White Paintings Rock Shelter, northwest Kalahari desert, Botswana. Journal of Archaeological Science, 27, 1085–1113. Robert, A., Soriano, S., Rasse, M., Stokes, S., & Huysecom, E. (2003). First chrono-cultural reference framework for the West African Paleolithic: New data from Ounjougou, Dogon Country, Mali. Journal of African Archaeology, 1(2), 151–169. Roche, J. (1953). La grotte de Taforalt. L’Anthropologie, 57, 375–380. Roche, J. (1973). Notes sur les activités de chantiers (1972). Bulletin d’archéologie marocaine, 9, 147–153. Roche, H., Delagne, A., Brugal, J. P., Feibel, C., Kibunjia, M., et al. (1999). Early homind stone tool production and technical skill 2.34 myr ago in West Turkana, Kenya. Nature, 399, 57–59. Ruhlmann, A. (1945). Le Paléolithique Marocain. Nouvelle esquisse d’une étude d’ensemble (Vol. Fasc 7). Publications du Services des Antiquités du Maroc. Ruhlmann, A. (1951). La grotte pré historique de Dar Es-Soltan, Institut des Hautes Etudes Marocaines 11. Collections Hespé ris. Rüther, H., Chazan, M., Schroeder, R., Neeser, R., Held, C., Walker, S. J., et al. (2009). Laser scanning for conservation and research of African cultural heritage sites: The case study of Wonderwerk cave, South Africa. Journal of Archaeological Science, 36(9), 1847–1856. Sahle, Y., El Zaatari, S., & White, T. D. (2017). Hominid butchers and biting crocodiles in the African Plio–Pleistocene. Proceedings of the National Academy of Sciences, USA, 114(50), 13164–13169. Sahnouni, M., Hadjouis, D., van der Made, J., Derradji, A.-E.-K., Canals, A., Medig, M., et al. (2002). Further research at the Oldowan site of Ain Hanech, North-Eastern Algeria. Journal of Human Evolution, 43(6), 925–937. Sahnouni, M., Made, J. V. D., & Everett, M. A. (2010). Early North Africa: chronology, ecology, and hominin behavior: insights from Ain Hanech and El-Kherba, northeastern Algeria. Quaternary International, 223, 436–438. Sahnouni, M., Van der Made, J., & Everett, M. (2011). Ecological background to Plio-Pleistocene hominin occupation in North Africa: The vertebrate faunas from Ain Boucherit, Ain Hanech and El-Kherba, and paleosol stable-carbon-isotope studies from El-Kherba, Algeria. Quaternary Science Reviews, 30(11), 1303–1317. Sahnouni, M., Rosell, J., van der Made, J., Vergès, J. M., Ollé, A., Kandi, N., et al. (2013). The first evidence of cut marks and usewear traces from the Plio-Pleistocene locality of El-Kherba (Ain Hanech), Algeria: Implications for early hominin subsistence activities circa 1.8 Ma. Journal of Human Evolution, 64(2), 137–150. Sahnouni, M., Parés, J. M., Duval, M., Cáceres, I., Harichane, Z., van der Made, J., et al. (2018). 1.9-million- and 2.4-million-year-old artifacts and stone tool–cutmarked bones from Ain Boucherit, Algeria. Science, 362(6420), 1297–1301. Sandelowsky, B. H., & Viereck, A. (1969). Supplementary report on the archaeological expedition of 1962 to the Erongo Mountains of South West Africa. Cimbebasia, 1(1), 1–43. Sano, K., Beyene, Y., Katoh, S., Koyabu, D., Endo, H., Sasaki, T., et al. (2020). A 1.4-millionyear-old bone handaxe from Konso, Ethiopia, shows advanced tool technology in the early Acheulean. Proceedings of the National Academy of Sciences, USA, 117(31), 18393–18400. Sari, L. (2020). Diachronic variation in microlith production systems during the Late Pleistocene, Algeria. African Archaeological Review, 37, 1–33. Sari, L., & Kim, K.-J. (2017). Lithic economy and specialized activities among the Iberomaurusian populations of Tamar Hat Rockshelter (Northeastern Algeria). African Archaeological Review, 34(4), 543–556. Sawchuk, E. A., & Willoughby, P. R. (2015). Terminal Pleistocene Later Stone Age human remains from the Mlambalasi rock shelter, Iringa region, southern Tanzania. International Journal of Osteoarchaeology, 25(5), 593–607. 2082 J. C. Thompson et al. Saxon, E. C., Close, A., Cluzel, C., Morse, V., & Shackelton, N. J. (1974). Results of recent investigations at Tamar Hat. Libyca, 22, 49–91. Sbihi Alaoui, F.-Z., Mohib, A., & Raynal, J.-P. (2007). 700 000 à 200 000 ans? Chasseurs et charognards: l’Acheuléen moyen et supérieur des sites des carrières Oulad Hamida 1, Thomas 1 et Sidi Abderrahmane. In F.-Z. S. Alaoui, A. Mohib, & J.-P. Raynal (Eds.), Casablanca il y a un million d’années...Un voyage dans le temps et dans l’espace. Ministère de la Culture. Schild, R., & Wendorf, F. (1993). Middle Paleolithic lakes in the southwestern Desert of Egypt. In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 15–65). Plenum. Schmidt, I. (2011). A Middle Stone Age assemblage with discoid lithic technology from Etemba 14, Erongo mountains, northern Namibia. Journal of African Archaeology, 9(1), 85–100. Schwarcz, H. P., & Grün, R. (1993). Electron spin resonance dating of tooth enamel from Bir Tarfawi. In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 234–237). Plenum. Schwarcz, H. P., & Rink, W. J. (2000). ESR dating of the Die Kelders Cave 1 Site, South Africa. Journal of Human Evolution, 38, 121–128. Schweitzer, F. R., & Wilson, M. L. (1978). A preliminary report on excavations at Byneskranskop, Bredasdorp District, Cape. The South African Archaeological Bulletin, 33, 134–140. Semaw, S., Rogers, M. J., Quade, J., Renne, P. R., Butler, R. F., Dominguez-Rodrigo, M., et al. (2003). 2.6-million-year-old stone tools and associated bones from OGS-6 and OGS-7, Gona, Afar, Ethiopia. Journal of Human Evolution, 45, 169–177. Semaw, S., Rogers, M. J., Cáceres, I., Stout, D., & Leiss, A. C. (2018). The early Acheulean ~1.6–1.2 Ma from Gona, Ethiopia: Issues related to the emergence of the Acheulean in Africa. In R. Gallotti & M. Mussi (Eds.), The emergence of the Acheulean in East Africa and beyond: Contributions in Honor of Jean Chavaillon (pp. 115–128). Springer International Publishing. Semaw, S., Rogers, M. J., Simpson, S. W., Levin, N. E., Quade, J., Dunbar, N., et al. (2020). Co-occurrence of Acheulian and Oldowan artifacts with Homo erectus cranial fossils from Gona, Afar, Ethiopia. Science Advances, 6(10), eaaw4694. Sereno, P. C., Garcea, E. A., Jousse, H., Stojanowski, C. M., Saliège, J.-F., Maga, A., et al. (2008). Lakeside cemeteries in the Sahara: 5000 years of Holocene population and environmental change. PLoS One, 3(8), e2995. Shaw, T., & Daniels, S. (1984). Excavations at Iwo Eleru, Ondo State, Nigeria. West African Journal of Archaeology, 14, 1–269. Shea, J. J., & Hildebrand, E. A. (2010). The Middle Stone Age of West Turkana, Kenya. Journal of Field Archaeology, 35(4), 355–364. Shipton, C., Crowther, A., Kourampas, N., Prendergast, M. E., Horton, M., Douka, K., et al. (2016). Reinvestigation of Kuumbi Cave, Zanzibar, reveals Later Stone Age coastal habitation, early Holocene abandonment and Iron Age reoccupation. Azania: Archaeological Research in Africa, 51(2), 197–233. Shipton, C., Roberts, P., Archer, W., Armitage, S. J., Bita, C., Blinkhorn, J., et al. (2018). 78,000-year-old record of Middle and Later Stone Age innovation in an East African tropical forest. Nature Communications, 9(1), 1832. Sisk, M. L., & Shea, J. J. (2008). Intrasite spatial variation of the Omo Kibish Middle Stone Age assemblages: Artifact refitting and distribution patterns. Journal of Human Evolution, 55(3), 486–500. Slater, P. A. (2016). Change in lithic technological organization strategies during the Middle and Later Stone Ages in East Africa. PhD dissertation, University of Illinois at Urbana-Champaign. Smith, E. I., Jacobs, Z., Johnsen, R., Ren, M., Fisher, E. C., Oestmo, S., et al. (2018). Humans thrived in South Africa through the Toba eruption about 74,000 years ago. Nature, 555, 511. Smith, G. M., Ruebens, K., Gaudzinski-Windheuser, S., & Steele, T. E. (2019). Subsistence strategies throughout the African middle Pleistocene: Faunal evidence for behavioral change and continuity across the Earlier to Middle Stone Age transition. Journal of Human Evolution, 127, 1–20. Soper, R. C., & Golden, B. (1969). An archaeological survey of Mwanza Region, Tanzania. Azania: Archaeological Research in Africa, 4(1), 15–79. The Zooarchaeology of Pleistocene Africa 2083 Stammers, R. C., Caruana, M. V., & Herries, A. I. R. (2018). The first bone tools from Kromdraai and stone tools from Drimolen, and the place of bone tools in the South African Earlier Stone Age. Quaternary International, 495, 87–101. Steele, T. E., & Álvarez-Fernández, E. (2011). Initial investigations into the exploitation of coastal resources in North Africa during the Late Pleistocene at Grotte des Contrebandiers, Morocco. In Trekking the shore (pp. 383–403). Springer. Steele, T. E., & Klein, R. G. (2013). The Middle and Later Stone Age faunal remains from Diepkloof Rock Shelter, Western Cape, South Africa. Journal of Archaeological Science, 40(9), 3453–3462. Steele, T. E., Mackay, A., Orton, J., & Schwortz, S. (2012). Varsche Rivier 003, a new Middle Stone Age site in southern Namaqualand, South Africa. The South African Archaeological Bulletin, 67(195), 108–119. Steele, T., Mackay, A., & Fitzsimmons, K. (2016). Varsche Rivier 003: A Middle and Later Stone Age Site with Still Bay and Howiesons Poort Assemblages in Southern Namaqualand, South Africa. PaleoAnthropology, 64, 100. Steele, T., Álvarez-Fernández, E., & Hallett-Desguez, E. (2019). Special issue: Early personal ornaments: A review of shells as personal ornamentation during the African Middle Stone Age. PaleoAnthropology, 2019, 24–51. Stephens, L., Fuller, D., Boivin, N., Rick, T., Gauthier, N., Kay, A., et al. (2019). Archaeological assessment reveals Earth’s early transformation through land use. Science, 365(6456), 897–902. Stewart, K. M. (1989). Fishing sites of North and East Africa in the Late Pleistocene and Holocene: Environmental change and human adaptation (British archaeological reports) (Vol. 521). British Archaeological Reports Ltd. Stewart, K. M. (2004). Revisiting the Senga 5a fish Fauna, upper Semliki River, Democratic Republic of Congo. Archaeofauna, 13, 145–154. Stewart, K. M., & Murray, A. M. (2020). Pliocene and Pleistocene fishes from Gona, Ethiopia: Inferences for reconstructing freshwater paleoecology. Journal of Vertebrate Paleontology, 40(5), e1819302. Stewart, B. A., Dewar, G. I., Morley, M. W., Inglis, R. H., Wheeler, M., Jacobs, Z., et al. (2012). Afromontane foragers of the Late Pleistocene: Site formation, chronology and occupational pulsing at Melikane Rockshelter, Lesotho. Quaternary International, 270, 40–60. Stoetzel, E. (2009). Les microvertébrés du site d’occupation humaine d’El Harhoura 2 (Pléistocène supérieur, Holocène, Maroc): systématique, évolution, taphonomie et paléoécologie. , Muséum national d’histoire naturelle. Stoetzel, E., Marion, L., Nespoulet, R., El Hajraoui, M. A., & Denys, C. (2011). Taphonomy and palaeoecology of the Late Pleistocene to middle Holocene small mammal succession of El Harhoura 2 cave (Rabat-Témara, Morocco). Journal of Human Evolution, 60(1), 1–33. Stoetzel, E., Denys, C., Bailon, S., El Hajraoui, M., & Nespoulet, R. (2012). Taphonomic analysis of amphibian and squamate remains from El Harhoura 2 (Rabat-Témara, Morocco): Contributions to palaeoecological and archaeological interpretations. International Journal of Osteoarchaeology, 22(5), 616–635. Stoetzel, E., Campmas, E., Michel, P., Bougariane, B., Ouchaou, B., Amani, F., et al. (2014). Context of modern human occupations in North Africa: Contribution of the Témara caves data. Quaternary International, 320, 143–161. Stoetzel, E., Sime, W. B., Pleurdeau, D., Asrat, A., Assefa, Z., Desclaux, E., et al. (2018). Preliminary study of the rodent assemblages of Goda Buticha: New insights on Late Quaternary environmental and cultural changes in southeastern Ethiopia. Quaternary International, 471, 21–34. Stojanowski, C. M. (2014). Iwo Eleru’s place among Late Pleistocene and Early Holocene populations of North and East Africa. Journal of Human Evolution, 75, 80–89. Stollhofen, H., Stanistreet, I. G., Toth, N., Schick, K. D., Rodríguez-Cintas, A., Albert, R. M., et al. (2021). Olduvai’s oldest Oldowan. Journal of Human Evolution, 150, 102910. Sutton, M. B., Kuman, K., & Steininger, C. (2017). Early Pleistocene stone artefacts from Cooper’s Cave, South Africa. The South African Archaeological Bulletin, 72(206), 156–161. 2084 J. C. Thompson et al. Suwa, G., Nakaya, H., Asfaw, B., Saegusa, H., Amzaye, A., Kono, R. T., et al. (2003). PlioPleistocene terrestrial mammal assemblage from Konso, southern Ethiopia. Journal of Vertebrate Paleontology, 23(4), 901–916. Suwa, G., Asfaw, B., Sano, K., & Beyene, Y. (2020). Reply to Barkai: Implications of the Konso bone handaxe. Proceedings of the National Academy of Sciences, USA, 117(49), 30894–30895. Tappen, M., & Harris, J. W. K. (1995). Comment on possible cut marks and taphonomic history of Senga 5A in the Western Rift Valley, Zaire. Journal of Human Evolution, 29(5), 483–486. Taru, P., & Backwell, L. (2013). Identification of fossil hairs in Parahyaena brunnea coprolites from Middle Pleistocene deposits at Gladysvale cave, South Africa. Journal of Archaeological Science, 40(10), 3674–3685. Taylor, V. K., Barton, R. N. E., Bell, M., Bouzouggar, A., Collcutt, S., Black, S., et al. (2011). The Epipalaeolithic (Iberomaurusian) at Grotte des Pigeons (Taforalt), Morocco: A preliminary study of the land Mollusca. Quaternary International, 244(1), 5–14. Texier, J. P., Huxtable, J., Rhodes, E., Miallier, D., & Ousmoi, M. (1988). Nouvelles données sur la situation chronologique de l’Atérien du Maroc et leurs implications. Comptes rendus de l’Académie des sciences. Série 2, Mécanique, Physique, Chimie, Sciences de l’univers, Sciences de la Terre, 307(7), 827–832. Thackeray, J. F. (1979). An analysis of faunal remains from archaeological sites in southern South West Africa (Namibia). South African Archaeological Bulletin, 34, 18–33. Thackeray, J. F. (2016). A history of research on human evolution in South Africa from 1924 to 2016. Revue de Primatologie, 99(7), 279–282. Thomas, P.-É. (1884). Recherches stratigraphiques et paléontologiques sur quelques formations d’eau douce de l’Algérie: rue des Grands-Augustins, 7. Thomas, H. (1977). Géologie et Paléontologie du gisement acheuléen de l’erg Tihodaïne (Ahaggar. Sahara central) (Mém., Vol. XXVII). CRAPE. Thompson, J. C. (2005). The impact of post-depositional processes on bone surface modification frequencies: A corrective strategy and its application to the Loiyangalani site, Serengeti Plain, Tanzania. Journal of Taphonomy, 3(2), 57–80. Thompson, J. C. (2010). Taphonomic analysis of the Middle Stone Age faunal assemblage from Pinnacle Point Cave 13B, Western Cape, South Africa. Journal of Human Evolution, 59(3), 321–339. Thompson, J. C. (2020). Faunal analysis in African archaeology. In C. Shirikure (Ed.), Oxford research encyclopedia of anthropology. Oxford University Press. Thompson, J. C., & Henshilwood, C. S. (2011). Taphonomic analysis of the Middle Stone Age larger mammal faunal assemblage from Blombos Cave, southern Cape, South Africa. Journal of Human Evolution, 60(6), 746–767. Thompson, J. C., & Henshilwood, C. S. (2014a). Nutritional values of tortoises relative to ungulates from the Middle Stone Age levels at Blombos Cave, South Africa: Implications for foraging and social behaviour. Journal of Human Evolution, 67, 33–47. Thompson, J. C., & Henshilwood, C. S. (2014b). Tortoise taphonomy and tortoise butchery patterns at Blombos Cave, South Africa. Journal of Archaeological Science, 41, 214–229. Tiercelin, J.-J., Schuster, M., Roche, H., Brugal, J.-P., Thuo, P., Prat, S., et al. (2010). New considerations on the stratigraphy and environmental context of the oldest (2.34Ma) Lokalalei archaeological site complex of the Nachukui Formation, West Turkana, northern Kenya Rift. Journal of African Earth Sciences, 58(2), 157–184. Tobias, P. V. (1967). The hominid skeletal remains of Haua Fteah. The Haua Fteah (Cyrenaica) and the Stone Age of the South-East Mediterranean. In: C.B.M. McBurney (Ed.), The Haua Fteah (Cyrenaica) and the Stone Age of the South-East Mediterranean, Cambridge University Press, Cambridge, pp. 338–352. Toffolo, M. B., Brink, J. S., & Berna, F. (2015). Bone diagenesis at the Florisbad spring site, Free State Province (South Africa): Implications for the taphonomy of the Middle and Late Pleistocene faunal assemblages. Journal of Archaeological Science: Reports, 4, 152–163. The Zooarchaeology of Pleistocene Africa 2085 Tomasso, S., & Rots, V. (2018). What is the use of shaping a tang? Tool use and hafting of tanged tools in the Aterian of Northern Africa. Archaeological and Anthropological Sciences, 10(6), 1389–1417. Tong, H. (1986). The Gerbillinae (Rodentia) from Tighennif (Pleistocene of Algeria) and their significance. Modern Geology, 10, 197–214. Toussaint, M., Tobias, P. V., Hughes, A. R., Macho, G. A., & Partridge, T. C. (2003). The third partial skeleton of a late Pliocene hominin (Stw 431) from Sterkfontein, South Africa. South African Journal of Science, 99(5), 215–223. Tribolo, C., Mercier, N., Douville, E., Joron, J.-L., Reyss, J.-L., Rufer, D., et al. (2013). OSL and TL dating of the Middle Stone Age sequence at Diepkloof Rock Shelter (South Africa): A clarification. Journal of Archaeological Science, 40(9), 3401–3411. Trinkaus, E. (2009). The human tibia from Broken Hill, Kabwe, Zambia. PaleoAnthropology, 2009, 145–165. Tryon, C. A., & Faith, J. T. (2016). A demographic perspective on the Middle to Later Stone Age transition from Nasera rockshelter, Tanzania. Philosophical Transactions of the Royal Society B, 371, 20150238. Tryon, C. A., Crevecoeur, I., Faith, J. T., Ekshtain, R., Nivens, J., Patterson, D., et al. (2015). Late Pleistocene age and archaeological context for the hominin calvaria from GvJm-22 (Lukenya Hill, Kenya). Proceedings of the National Academy of Sciences, USA, 112(9), 2682–2687. Tryon, C. A., Lewis, J. E., Ranhorn, K. L., Kwekason, A., Alex, B., Laird, M. F., et al. (2018). Middle and Later Stone Age chronology of Kisese II rockshelter (UNESCO World Heritage Kondoa Rock-Art Sites), Tanzania. PLoS One, 13(2), e0192029. Turner, G. (1986). Faunal remains from Jubilee Shelter, Transvaal. The South African Archaeological Bulletin, 41, 63–68. Turner, A. (1989). Sample selection, schlepp effects and scavenging: The implications of partial recovery for interpretations of the terrestrial mammal assemblage from Klasies River Mouth. Journal of Archaeological Science, 16, 1–11. Turner, E., Humphrey, L., Bouzouggar, A., & Barton, N. (2020). Bone retouchers and technological continuity in the Middle Stone Age of North Africa. PLoS One, 15(3), e0230642. Val, A. (2019). New data on avifaunal remains associated with the Middle Stone Age layers from Diepkloof Rock Shelter, Western Cape, South Africa. Journal of Archaeological Science: Reports, 26, 101880. Val, A., de la Peña, P., & Wadley, L. (2016). Direct evidence for human exploitation of birds in the Middle Stone Age of South Africa: The example of Sibudu Cave, KwaZulu-Natal. Journal of Human Evolution, 99, 107–123. Val, A., Porraz, G., Texier, P.-J., Fisher, J. W., & Parkington, J. (2020). Human exploitation of nocturnal felines at Diepkloof Rock Shelter provides further evidence for symbolic behaviours during the Middle Stone Age. Scientific Reports, 10(1), 1–12. Vallois, H. V., & Roche, J. (1958). La mandibule acheuléene de Témara, Maroc. Comptes Rendus de l’Acádemie des Sciences, 246, 3113–3116. Van Neer, W. (1981). Archeozoögische studie van Matupi (Ijzertijd en Late Steentijd) en Kiantapo (Ijzertijd) in Zaire. Katholieke Universiteit te Leuven. Van Neer, W. (1984). Faunal remains from Matupi Cave, an Iron Age and Late Stone Age site in northeastern Zaire. AWLSK, 46(2), 59–76. Van Neer, W. (1989). Contribution to the archaeozoology of Central Africa (Annales Sciences Zoologiques-Musee Royal de l’Afrique Centrale). Musee Royal de l’Afrique Centrale. Van Neer, W. (1990). Les faunes de vertébrés quaternaires en Afrique centrale. In R. Lanfranchti & D. Schwartz (Eds.), Paysages Quaternaires de l’Afrique Centrale Atlantique (pp. 195–220). Institut Français de Recherche Scientifique pour le Développement en Coopération. Van Neer, W. (1993). Fish remains from the last interglacial at Bir Tarfawi (eastern Sahara, Egypt). In F. Wendorf, R. Schild, & A. Close (Eds.), Egypt during the last interglacial (pp. 144–154). Plenum. 2086 J. C. Thompson et al. Van Neer, W., & Breunig, P. (1999). Contribution to the archaeozoology of the Brandberg, Namibia. Cimbebasia, 15, 127–140. Van Noten, F. (1971). Excavations at Munyama cave. Antiquity, 45(177), 56–58. Van Noten, F. (1977). Excavations at Matupi Cave. Antiquity, LI, 35–40. Van Peer, P., Vermeersch, P. M., & Paulissen, E. (2010). Chert quarrying, lithic technology and a modern human burial at the palaeolithic site of Taramsa 1, Upper Egypt (Vol. Vol. 5). Leuven University Press. van Pletzen-Vos, L., Brink, J., Reynard, J., & Wurz, S. (2019). Revisiting Klasies River: A report on the large mammal remains from the Deacon excavations of Klasies River main site, South Africa. South African Archaeological Bulletin, 74(211), 127. Vanhaeren, M., D’Errico, F., Stringer, C., James, S. L., Todd, J. A., & Mienis, H. K. (2006). Middle Paleolithic shell beads in Israel and Algeria. Science, 312, 1785–1788. Vaufrey, R. (1950). La faune de Sidi Zin. In E. G. Gobert (Ed.), Le gisement paléolithiquede Sidi Zin (pp. 41–51). Karthago. Vermeersch, P. M. (2008). Egypt from 50 to 25 ka BP: A scarcely inhabited region. In M. Camps & C. Szmidt (Eds.), The Mediterranean from 50,000 to 25,000 BP: Turning points and new directions (pp. 67–88). Oxbow Books. Vermeersch, P. M., & Van Neer, W. (2015). Nile behaviour and Late Palaeolithic humans in Upper Egypt during the late Pleistocene. Quaternary Science Reviews, 130, 155–167. Vermeersch, P. M., Paulissen, E., Gijselings, G., Otte, M., Thoma, A., Van Peer, P., et al. (1984). 33,000-yr old chert mining site and related Homo in the Egyptian Nile Valley. Nature, 309(5966), 342–344. Vermeersch, P. M., Paulissen, E., Van Peer, P., Stokes, S., Charlier, C., Stringer, C., et al. (1998). A Middle Palaeolithic burial of a modern human at Taramsa Hill, Egypt. Antiquity, 72(277), 475–484. Vermeersch, P. M., Linseele, V., Marinova, E., Van Neer, W., Moeyersons, J., & Rethemeyer, J. (2015). Early and middle Holocene human occupation of the Egyptian eastern desert: Sodmein cave. African Archaeological Review, 32(3), 465–503. Vogel, J. (1983). Isotopic evidence for the past climates and vegetation of southern Africa. Bothalia, 14(3/4), 391–394. Vogel, J., & Visser, E. (1981). Pretoria radiocarbon dates II. Radiocarbon, 23(1), 43–80. Vogelsang, R. (1998). Middle Stone Age Fundstellen in Südwest-Namibia (Vol. 11). Heinrich-Barth-Institut. Vogelsang, R., & Eichhorn, B. (2011). Under the mopane tree: Holocene settlement in northern Namibia (Vol. 24). Heinrich Barth Institut. Vogelsang, R., Richter, J., Jacobs, Z., Eichhorn, B., Linseele, V., & Roberts, R. G. (2010). New excavations of Middle Stone Age deposits at Apollo 11 Rockshelter, Namibia: Stratigraphy, archaeology, chronology and past environments. Journal of African Archaeology, 8(2), 185–218. Vrba, E. (1997). New fossils of Alcelaphini and Caprinae (Bovidae: Mammalia) from Awash, Ethiopia, and phylogenetic analysis of Alcelaphini. Palaeontologica Africana, 34, 127–198. Wadley, L. (1986). Segments of time: A mid-Holocene Wilton site in the Transvaal. The South African Archaeological Bulletin, 41, 54–62. Wadley, L. (1997). Rose Cottage Cave: Archaeological work 1987 to 1997. South African Journal of Science, 93(10), 439–444. Wadley, L., & Jacobs, Z. (2004). Sibudu Cave, KwaZulu-Natal: Background to the excavations of Middle Stone Age and Iron Age occupations. South African Journal of Science, 100(3–4), 145–151. Walker, N. J. (1995). Late Pleistocene and Holocene hunter-gatherers of the Matopos: an archaeological study of change and continuity in Zimbabwe. Societas archaeologica Upsaliensis: Department of Archaeology. Weaver, T. D., Steele, T., & Klein, R. (2011). The abundance of eland, buffalo, and wild pigs in Middle and Later Stone Age sites. Journal of Human Evolution, 60, 309–314. The Zooarchaeology of Pleistocene Africa 2087 Wells, L. H., Cooke, H. B. S., Malan, B. D., Wells, L. H., & Cooke, H. B. S. (1942). The associated fauna and culture of the Vlakkraal thermal springs, OFS. Transactions of the Royal Society of South Africa, 29(3), 203–233. Wendorf, F., & Said, R. (1967). Palaeolithic remains in upper Egypt. Nature, 215(5098), 244–247. Wendorf, F., & Schild, R. (1976). Prehistory of the Nile valley. Academic. Wendorf, F., Schild, R., Close, A. E., Hillman, G. C., Gautier, A., Van Neer, W., et al. (1988). New radiocarbon dates and late Palaeolithic diet at Wadi Kubbaniya, Egypt. Antiquity, 62(235), 279–283. Wendorf, F., Schild, R., & Close, A. E. (1993). Egypt during the last interglacial. Plenum. Wendt, W. E. (1972). Preliminary report on an archaeological research programme in South West Africa. Cimbebasia, 2(1), 1–61. Werdelin, L., & Sanders, W. J. (2010). Cenozoic mammals of Africa. University of California Press. White, T. D., Asfaw, B., DeGusta, D., Gilbert, H., Richards, G. D., Suwa, G., et al. (2003). Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature, 423(6941), 742–747. Willoughby, P. R. (2012). The Middle and Later Stone Age in the Iringa Region of southern Tanzania. Quaternary International, 270, 103–118. Willoughby, P. R., Compton, T., Bello, S. M., Bushozi, P. M., Skinner, A. R., & Stringer, C. B. (2018). Middle Stone Age human teeth from Magubike rockshelter, Iringa Region, Tanzania. PLoS One, 13(7), e0200530. Wright, D. K., Thompson, J. C., Mackay, A., Welling, M., Forman, S., Price, G., et al. (2014). Renewed geoarchaeological investigations of Mwanganda’s Village (Elephant Butchery Site), Karonga, Malawi. Geoarchaeology, 29(2), 98–120. Wrinn, P., & Rink, W. J. (2003). ESR dating of tooth enamel from Aterian levels at Mugharet el ‘Aliya (Tangier, Morocco). Journal of Archaeological Science, 30(1), 123–133. Yellen, J. E. (1996). Behavioural and taphonomic patterning at Katanda 9: A Middle Stone Age site, Kivu Province, Zaire. Journal of Archaeological Science, 6, 915–932. Yellen, J. E., Brooks, A. S., Stuckenrath, R., & Welbourne, R. (1987). A terminal Pleistocene assemblage from Drotsky’s Cave, western Ngamiland, Botswana. Botswana Notes and Records, 19, 1–6. Yellen, J. E., Brooks, A. S., Cornelissen, E., Mehlman, M. J., & Stewart, K. (1995). A Middle Stone Age worked bone industry from Katanda, Upper Semliki Valley, Zaire. Science, 268, 553–556. Yellen, J., Brooks, A., Helgren, D., Tappen, M., Ambrose, S., Bonnefille, R., et al. (2005). The archaeology of Aduma Middle Stone Age sites in the Awash Valley, Ethiopia. PaleoAnthropology, 10(2), 26–100. Yeshurun, R. (2018). Taphonomy of old archaeofaunal collections: New site-formation and subsistence data for the Late Paleolithic Nile Valley. Quaternary International, 471, 35–54. Yravedra, J., & Domínguez-Rodrigo, M. (2009). The shaft-based methodological approach to the quantification of long limb bones and its relevance to understanding hominid subsistence in the Pleistocene: Application to four Palaeolithic sites. Journal of Quaternary Science, 24(1), 85–96. Yravedra, J., Domínguez-Rodrigo, M., Santonja, M., Rubio-Jara, S., Panera, J., Pérez-González, A., et al. (2016). The larger mammal palimpsest from TK (Thiongo Korongo), Bed II, Olduvai Gorge, Tanzania. Quaternary International, 417, 3–15. Yravedra, J., Diez-Martín, F., Egeland, C. P., Maté-González, M. Á., Palomeque-González, J. F., Arriaza, M. C., et al. (2017). FLK West (Lower Bed II, Olduvai Gorge, Tanzania): A new early Acheulean site with evidence for human exploitation of fauna. Boreas, 46(4), 816–830. Yravedra, J., Rubio-Jara, S., Courtenay, L. A., & Martos, J. A. (2020). Mammal butchery by Homo erectus at the Lower Pleistocene Acheulean site of Juma’s Korongo 2 (JK2), bed III, Olduvai Gorge, Tanzania. Quaternary Science Reviews, 249, 106612.