Modern hunting behavior in the early Middle Paleolithic: Faunal remains from Misliya Cave, Mount...

22
Modern hunting behavior in the early Middle Paleolithic: Faunal remains from Misliya Cave, Mount Carmel, Israel Reuven Yeshurun * , Guy Bar-Oz, Mina Weinstein-Evron Zinman Institute of Archaeology, University of Haifa, Mount Carmel, 31905 Haifa, Israel Received 12 December 2006; accepted 10 May 2007 Abstract Understanding the behavioral adaptations and subsistence strategies of Middle Paleolithic humans is critical in the debate over the evolution and manifestations of modern human behavior. The study of faunal remains plays a central role in this context. Until now, the majority of Le- vantine archaeofaunal evidence was derived from late Middle Paleolithic sites. The discovery of faunal remains from Misliya Cave, Mount Car- mel, Israel (>200 ka), allowed for detailed taphonomic and zooarchaeological analyses of these early Middle Paleolithic remains. The Misliya Cave faunal assemblage is overwhelmingly dominated by ungulate taxa. The most common prey species is the Mesopotamian fallow deer (Dama mesopotamica), followed closely by the mountain gazelle (Gazella gazella). Some aurochs (Bos primigenius) remains are also present. Small-game species are rare. The fallow deer mortality pattern is dominated by prime-aged individuals. A multivariate taphonomic analysis demonstrates (1) that the assemblage was created solely by humans occupying the cave and was primarily modified by their food-processing activities; and (2) that gazelle carcasses were transported complete to the site, while fallow deer carcasses underwent some field butchery. The new zooarchaeological data from Misliya Cave, particularly the abundance of meat-bearing limb bones displaying filleting cut marks and the acquisition of prime-age prey, demonstrate that early Middle Paleolithic people possessed developed hunting capabilities. Thus, modern large-game hunting, carcass transport, and meat-processing behaviors were already established in the Levant in the early Middle Paleolithic, more than 200 ka ago. Ó 2007 Elsevier Ltd. All rights reserved. Keywords: Middle Paleolithic; Middle Pleistocene; Taphonomy; Zooarchaeology; Levant; Hunting; Multivariate taphonomic approach; Skeletal-element transport Introduction The Middle Paleolithic (MP) period in Eurasia and North Africa and its sub-Saharan African contemporary, the Middle Stone Age (MSA), are widely considered to offer the time frame in which behaviorally modern humans evolved. Whether closer to the end of the MP/MSA, at ca. 50 ka, or ear- lier during this ~200,000-year-long period, the emergence of modern human behavior is a major issue of debate for this pe- riod (e.g., Bar-Yosef, 1995; Klein, 1999; Stringer, 2002; Shea, 2003; Mellars, 2006; and see papers in Mellars and Stringer, 1989; Akazawa et al., 1998; Hovers and Kuhn, 2006). The term ‘‘modern human behavior’’ lacks a uniform definition in the literature (see discussions in Henshilwood and Marean, 2003; Bower, 2005). We define it quite broadly as representing behavior, decision-making, and lifeways similar to those ob- served among present-day hunter-gatherers, as well as unam- biguously anatomically and behaviorally modern humans in later prehistory (cf. Binford, 1985; Klein, 1989; Gaudzinsky and Roebroeks, 2000; Chase, 2003; Speth, 2004a). The ‘‘mod- ern human behavior’’ debate concerns the timing in which we find modern behavior unequivocally implied by the archaeo- logical evidence. Some researchers have suggested that modern human behavior prevailed during the MP/MSA and even before that (e.g., Deacon, 1989; Chase and Dibble, * Corresponding author. Fax: þ972 4 8249 876. E-mail addresses: [email protected] (R. Yeshurun), guybar@ research.haifa.ac.il (G. Bar-Oz), [email protected] (M. Weinstein- Evron). 0047-2484/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved. doi:10.1016/j.jhevol.2007.05.008 Available online at www.sciencedirect.com Journal of Human Evolution 53 (2007) 656e677

Transcript of Modern hunting behavior in the early Middle Paleolithic: Faunal remains from Misliya Cave, Mount...

Available online at www.sciencedirect.com

Journal of Human Evolution 53 (2007) 656e677

Modern hunting behavior in the early Middle Paleolithic: Faunalremains from Misliya Cave, Mount Carmel, Israel

Reuven Yeshurun*, Guy Bar-Oz, Mina Weinstein-Evron

Zinman Institute of Archaeology, University of Haifa, Mount Carmel, 31905 Haifa, Israel

Received 12 December 2006; accepted 10 May 2007

Abstract

Understanding the behavioral adaptations and subsistence strategies of Middle Paleolithic humans is critical in the debate over the evolutionand manifestations of modern human behavior. The study of faunal remains plays a central role in this context. Until now, the majority of Le-vantine archaeofaunal evidence was derived from late Middle Paleolithic sites. The discovery of faunal remains from Misliya Cave, Mount Car-mel, Israel (>200 ka), allowed for detailed taphonomic and zooarchaeological analyses of these early Middle Paleolithic remains. The MisliyaCave faunal assemblage is overwhelmingly dominated by ungulate taxa. The most common prey species is the Mesopotamian fallow deer(Dama mesopotamica), followed closely by the mountain gazelle (Gazella gazella). Some aurochs (Bos primigenius) remains are also present.Small-game species are rare. The fallow deer mortality pattern is dominated by prime-aged individuals. A multivariate taphonomic analysisdemonstrates (1) that the assemblage was created solely by humans occupying the cave and was primarily modified by their food-processingactivities; and (2) that gazelle carcasses were transported complete to the site, while fallow deer carcasses underwent some field butchery.The new zooarchaeological data from Misliya Cave, particularly the abundance of meat-bearing limb bones displaying filleting cut marksand the acquisition of prime-age prey, demonstrate that early Middle Paleolithic people possessed developed hunting capabilities. Thus, modernlarge-game hunting, carcass transport, and meat-processing behaviors were already established in the Levant in the early Middle Paleolithic,more than 200 ka ago.� 2007 Elsevier Ltd. All rights reserved.

Keywords: Middle Paleolithic; Middle Pleistocene; Taphonomy; Zooarchaeology; Levant; Hunting; Multivariate taphonomic approach; Skeletal-element transport

Introduction

The Middle Paleolithic (MP) period in Eurasia and NorthAfrica and its sub-Saharan African contemporary, the MiddleStone Age (MSA), are widely considered to offer the timeframe in which behaviorally modern humans evolved.Whether closer to the end of the MP/MSA, at ca. 50 ka, or ear-lier during this ~200,000-year-long period, the emergence ofmodern human behavior is a major issue of debate for this pe-riod (e.g., Bar-Yosef, 1995; Klein, 1999; Stringer, 2002; Shea,

* Corresponding author. Fax: þ972 4 8249 876.

E-mail addresses: [email protected] (R. Yeshurun), guybar@

research.haifa.ac.il (G. Bar-Oz), [email protected] (M. Weinstein-

Evron).

0047-2484/$ - see front matter � 2007 Elsevier Ltd. All rights reserved.

doi:10.1016/j.jhevol.2007.05.008

2003; Mellars, 2006; and see papers in Mellars and Stringer,1989; Akazawa et al., 1998; Hovers and Kuhn, 2006). Theterm ‘‘modern human behavior’’ lacks a uniform definitionin the literature (see discussions in Henshilwood and Marean,2003; Bower, 2005). We define it quite broadly as representingbehavior, decision-making, and lifeways similar to those ob-served among present-day hunter-gatherers, as well as unam-biguously anatomically and behaviorally modern humans inlater prehistory (cf. Binford, 1985; Klein, 1989; Gaudzinskyand Roebroeks, 2000; Chase, 2003; Speth, 2004a). The ‘‘mod-ern human behavior’’ debate concerns the timing in which wefind modern behavior unequivocally implied by the archaeo-logical evidence. Some researchers have suggested thatmodern human behavior prevailed during the MP/MSA andeven before that (e.g., Deacon, 1989; Chase and Dibble,

657R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

1990; Kaufman, 1999; McBrearty and Brooks, 2000; Roe-broeks, 2001; d’Errico, 2003; Hovers and Belfer-Cohen,2006), while others have placed its appearance only aroundthe transition between the Middle and Upper Paleolithic(e.g., Binford, 1985; Gargett, 1999; Hoffecker, 1999; Klein,2000; Mellars, 2005). Recently, studies of food-acquisitionmethods, raw-material-procurement strategies, intrasite spatialdistribution, and unique, nonutilitarian objects have all con-tributed to the debate over the appearance of modern behaviorin the MP/MSA record. These studies have illuminated the dif-ferences or the similarities in behavior between MP/MSA hu-mans and their Upper Paleolithic/Late Stone Age successors(e.g., Stiner, 1993, 1994; Farizy et al., 1994; Marean, 1998;Milo, 1998; Gamble, 1999; Henshilwood et al., 2002; Kauf-man, 2002; Hovers et al., 2003; Henry et al., 2004; Kleinet al., 2004; Adler et al., 2006; Assefa, 2006; Kuhn and Stiner,2006; Speth, 2006; Wadley, 2006).

Zooarchaeological studies of the MP/MSA assemblagesconstitute an important source of knowledge about human sub-sistence and behavior, which makes them fundamental to themodern behavior debate. Three types of data are frequentlyused to infer human subsistence behavior: prey-species com-position, skeletal-element frequency, and ungulate mortalitypatterns. Claims for low technological and social skills inthe MP/MSA period were based on the paucity of dangerousprey species, such as suids and the African buffalo (Synceruscaffer), in the archaeological record of this period; the rarityof evidence for specialized, monospecific hunting strategies;and the sporadic rather than systematic occurrence of marineresources and small game. All of these traits occur in the ar-chaeofaunas of later periods (Klein, 1989, 1998; Klein andCruz-Uribe, 2000; Klein et al., 2004; Mellars, 2004). Theabundance of head and foot parts in some MP and MSA assem-blages, and of carnivore ravaging marks on the bones, rein-forced claims for scavenging of these nutritionally poorcarcass parts from carnivore kills as an important subsistencestrategy in the MP/MSA (Binford, 1984, 1985, 1988; but seeGrayson and Delpech, 1994). Stiner (1994: 236e270) arguedfor a scavenging-oriented strategy in the Italian MP based ona bias toward head parts in assemblages predating 55 ka. Fi-nally, ungulate mortality patterns dominated by juvenile and se-nile animals have been used to demonstrate scavenging fromcarnivore kills, as carnivores usually hunt the weakest animalsin a herd, whereas later humans and recent hunter-gatherers typ-ically focus on prime-age adults (Stiner, 1990, 1994: 288e315).

These views have recently been challenged by new studiesof additional MP/MSA faunal assemblages and by reassess-ments of several old assemblages. It has been demonstratedthat many MP/MSA assemblages display a level of taxonomicdiversity that is similar to that of assemblages found in laterstratigraphic horizons of the same sites or in the nearby geo-graphical area (e.g., Chase, 1989; Kaufman, 2002; Graysonand Delpech, 2003). Some cases of MP specialized huntingwere documented, at least as a seasonal strategy (e.g., Adleret al., 2006; Gaudzinsky, 2006; Costamagno et al., 2006),while nonspecialized hunting has been shown to exist inmuch later periods and can be attributed to changes in

ecological parameters or seasonality rather than human hunt-ing capability (Grayson and Delpech, 2002; Bar-Yosef,2004). Similarly, the preference for ‘‘docile’’ over dangerousanimals and the acquisition of ungulates rather than smallmarine or terrestrial game is viewed as representing sensibleadaptations in periods of presumed low human populationdensities. It is widely accepted that the lack of populationpressure makes it unnecessary to broaden the dietary spectrum(Stiner et al., 1999, 2000; Speth, 2004b). Multiple publishedMP ungulate mortality patterns show a marked dominanceof prime adults, indicating systematic hunting, not scavengingfrom carnivore kills (e.g., Speth and Tchernov, 1998; Gaudzin-sky and Roebroeks, 2000; Steele, 2004; Stiner, 2002a, 2005;Bar-Oz and Adler, 2005). Also, and importantly, the previousreports of head- and toe-dominated skeletal-element profiles inMP/MSA sites, which contributed to the scavenging hypothe-sis, were criticized on both analytical and taphonomicgrounds. Marean and colleagues (Marean and Kim, 1998;Marean and Assefa, 1999; Pickering et al., 2003; Mareanet al., 2004) argued that these studies either ignored fragmentsof limb-bone shafts or could not study them because of incom-plete retention during and after the excavation. Since shaftfragments are among the densest parts in the skeleton (Lamet al., 1999, 2003) and do not particularly attract ravaging car-nivores (Marean and Spencer, 1991; see also Figure 9 inMarean and Cleghorn, 2003), they survive in higher ratesthan limb-bone ends. Their exclusion from the analysis willtherefore bias the inferred pattern against meat- and marrow-rich limb bones and create an artificial ‘‘head and foot’’ pat-tern, which may be falsely interpreted as evidence of humanscavenging. Actually, most MP/MSA assemblages that weresystematically collected and analyzed have producedskeletal-element profiles dominated by meaty limbs, indicat-ing that humans had first access to the carcass, presumablyas a consequence of hunting (Marean and Kim, 1998; Mareanand Assefa, 1999; Pickering et al., 2003; Marean et al., 2004).

The vast majority of these zooarchaeological (and other)studies dealt with the later MP/MSA (sites younger than 100ka). If we accept the hypothesis that modern human behaviorprevailed during the MP, we should search the earlier MP re-cord for its origins and early evolution. However, the criticalperiod of the early Middle Paleolithic (EMP) is relatively un-known in many areas, including the Levant.

Taphonomic and zooarchaeological studies of the later MPsites in the Levant are an important contribution to the debateover MP subsistence strategies and hunting capabilities, usu-ally providing strong support for the notion that MP humanswere similar in these respects to later humans (e.g., Rabino-vich, 1990; Rabinovich and Tchernov, 1995; Speth and Tcher-nov, 1998, 2001, in press; Rabinovich and Hovers, 2004;Speth, 2004b; Speth and Clark, 2006). In contrast, zooarchaeo-logical evidence concerning the Levantine early MP is ex-tremely rare. Although some rich archaeological layers fromthis period have been excavateddmost notably layer D inTabun Cave, Mount Carmel, and Rosh Ein Mor in the Negevhighlanddthe preservation of faunal remains was usually poor(e.g., Garrod and Bate, 1937; Jelinek et al., 1973; Tchernov,

658 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

1976). A notable exception is Stiner’s (2005) comprehensivestudy of Hayonim Cave, dated to 140e220 ka (Rink et al.,2004; Mercier et al., 2007). The ungulate mortality patternsand skeletal-part profiles of Hayonim Cave indicate systematichunting of gazelle, fallow deer, and aurochs, as well as trans-port of complete or nearly complete animals back to the cave.The long stratigraphic sequence of Hayonim Cave (EMP to theEpipaleolithic period) shows significant expansion of humandietary breadth during the periods postdating the EMP, basedon the types of small game captured. This trend was inter-preted as reflecting human population growth and subsequentpressure on the environment, meaning that the EMP popula-tion lived under the carrying capacity, visiting the site at largeintervals, and thus was small and dispersed (Stiner et al., 1999;Stiner, 2005; see also Meignen et al., 2006).

The work reported here adds EMP zooarchaeological andtaphonomic data and interpretations by presenting in detailthe fully recovered and thoroughly studied faunal assemblagefrom Misliya Cave, Mount Carmel, Israel (Weinstein-Evronet al., 2003a). We aim to reconstruct the taphonomic historyof the assemblage and human subsistence behavior at thesitednamely mode of prey acquisition, prey-transport patterns,

and food processingdand use these new data to deepen ourunderstanding of the early Middle Paleolithic of MountCarmel.

Misliya Cave and its settings

Misliya is a collapsed cave located in the western cliff ofMount Carmel, northern Israel, approximately 12 km southof Haifa and 7 km north of the major Paleolithic site of NahalMe’arot (Wadi el-Mughara: the caves of Tabun, Jamal, el-Wad, and Skhul; Fig. 1) at an elevation of ca. 90 m abovemodern sea level. The present setting of the site is withinthe Mediterranean climatic zone, with hot, dry summers andcool, humid winters. The average annual temperature is19 �C and the mean annual rainfall is approximately600 mm. The vegetation surrounding the cave is a Mediterra-nean maquis. Misliya Cave is placed at an ecotonal setting, po-sitioned where the cliff of Mount Carmel meets the openexpanses of the coastal plain [see Weinstein-Evron (1998)for description of the similar setting of the Nahal Me’arotcaves]. This setting enabled the exploitation of locally avail-able resources from different biotopes, including forested hills,

Fig. 1. Location map of Misliya Cave and other MP sites mentioned in the text, with plan of the upper terrace excavation and section view of the site. Excavated

squares (1 m2) are highlighted. The EMP faunal assemblage discussed here originates from squares KeN/9e10 and I10 on the upper terrace of the site, highlighted

by dark gray.

659R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

mountain slopes and inner valleys, the coastal plain with itsmarshes and kurkar (aeolianite) ridges, and the nearby WadiSefunim, located some 500 m north of the site. These Mediter-ranean microenvironments contain diverse types of vegetation(Naveh, 1984). A perennial spring (’Ein Sefunim) exists todayin the upper Wadi Sefunim, at a distance of ca. 1.5 km fromthe site. Other springs and increased flow in the wadis mayhave been available in the vicinity of the site before the twen-tieth-century capture of most natural water resources (Inbarand Ayal, 1980; cf. Survey of Palestine, 1942).

The site was discovered by Brotzen in 1925 (Olami, 1984)but was not excavated until the recent campaign. Severalsurveys conducted at the site came up with Middle and LowerPaleolithic findings (Olami, 1984; Weinstein-Evron and Kauf-man, 1998), and a geophysical survey indicated the presenceof thick archaeological deposits (Weinstein-Evron et al.,2003b). Consequently, a multiyear excavation project at Mis-liya Cave was launched in 2001 (Weinstein-Evron et al.,2003a). The site consists of three ‘‘terraces’’ that are over-looked by a 10e15-m-high cliff wall, the remnants of a col-lapsed cave ceiling (Fig. 1). The archaeological sediments inthe three terraces are strongly cemented (brecciated), exceptfor the eastern part of the upper terrace, where brecciatedlayers change laterally into soft sediments, forming an areaof ca. 15 m2, that are more amenable to excavation.

The ongoing excavations at the site have yielded two cul-tural complexes: MP (Mousterian) and late Lower Paleolithic(Acheulo-Yabrudian) (Weinstein-Evron et al., 2003a; Zaidneret al., 2006). The focus of the excavation has been the ‘‘softarea’’ in the upper terrace, where a rich MP layer was un-earthed over ca. 10 m2. This soft, compacted brown layer(mainly squares KeN/9e10 and I10; Fig. 1) lies on a massiverock-fall and is partially sealed by brecciated sediments(Squares IeN10). Squares KeN9 underlie recent terra rosasediments. Excavation of this soft layer produced numerouslithic artifacts and bone fragments, as well as a well-definedhearth located in squares LeM/8e9. The hearth, which hasbeen only partially excavated, is composed of spatially con-centrated black and gray ash lenses, distinct from the sur-rounding brown sediments.

The ‘‘soft’’ area of Misliya Cave was excavated using 1-m2

squares and 5-cm spits. The hearth unit was excavated usinga number of spatial units in order to separate sediments dis-playing different colors and consistencies, which may repre-sent different burning intensities. The faunal materialreported in this work originates from the ‘‘soft’’ area, whichfacilitated the complete retrieval of faunal fragments. This isin contrast to the cemented sediments of the site, from whichit was difficult to retrieve and clean faunal remains (Bar-Ozet al., 2004). No sedimentological or typological differenceswere found within this 50-cm-thick layer, so we regard allof it as one faunal sample representing the subsistence econ-omy of the MP inhabitants of Misliya.

The lithic analysis of the MP assemblages of Misliya Cave(Weinstein-Evron et al., 2003a) showed that they all belong tothe early Levantine MP (‘‘Tabun D-type,’’ e.g., Garrod andBate, 1937; Jelinek, 1982; Bar-Yosef, 1998). The same

cultural phase in the nearby Tabun Cave was dated by thermo-luminescence (TL) to ca. 190e260 ka (Mercier and Valladas,2003, and references therein). A single optically stimulated lu-minescence sample from the brecciated sediments of squareO17 yielded an age of 130� 33 ka, considered to be a mini-mum age due to signal saturation (Weinstein-Evron et al.,2003a). Preliminary TL dates on burned Middle Paleolithicflint artifacts from Misliya Cave suggest that they are olderthan 200 ka (N. Mercier and H. Valladas, personal communi-cation, 2006), thus corroborating the typological date andbroadly assigning the site to marine isotope stage (MIS) 7. Pal-ynological and microfaunal evidence from nearby Tabun Caveand from Hayonim Cave in the western Galilee indicate a gen-erally humid climate in the EMP of the Mediterranean zone ofIsrael (Jelinek et al., 1973; Tchernov, 1998). Marine oxygen-isotopic data from cave speleothems in central and northernIsrael indicate increased precipitation and relatively warmconditions in MIS 7 (Bar-Matthews and Ayalon, 2001; Bar-Matthews et al., 2003). It should be noted, however, thatmany climatic fluctuations occurred during this long period,so for now we are unable to tie the Misliya Cave habitationlayers to a particular climatic setting with any precision.

Significantly, Misliya Cave is one of the very few ‘‘TabunD-type’’ Mousterian sites in the Levant to be systematicallyexcavated using modern retrieval methods and produce a size-able faunal assemblage, and only the second EMP site in theregion to undergo detailed taphonomic and zooarchaeologicalanalysis (the first was Hayonim Cave; Stiner, 2005). The anal-ysis, instrumental to the understanding of human subsistenceand behavior in this critical period of modern human evolu-tion, is presented in the following sections.

Faunal-analysis procedures

The research protocol applied here included the identifica-tion of all skeletal elements, determination of ungulate mortal-ity patterns, and systematic documentation of bone-surfacemodifications and mode of bone fragmentation. Thesemethods were used to minimize the potential recovery, identi-fication, and interpretation biases in the analysis (e.g., Kleinand Cruz-Uribe, 1984; Lyman, 1994; Marean et al., 2004).We used a multivariate taphonomic approach (Behrensmeyer,1991; Bar-Oz and Dayan, 2003; Bar-Oz, 2004; Bar-Oz andMunro, 2004) to discern the taphonomic history of the assem-blage, with special reference to human subsistence behavior.

During fieldwork, every bone fragment longer than 2.5 cmwas plotted using three coordinates and its orientation and in-clination were recorded. The plotted bone fragments werebagged separately on site. All of the remaining excavated sed-iment was collected and wet-sieved through 5-mm and 1-mmmeshes. All faunal remains that were not plotted in the fieldwere hand-picked from the 5-mm sieves and kept. The follow-ing procedures concern both piece-plotted and sieve-recoveredfaunal remains.

All faunal remains were immersed in acetic acid (10%) forca. 1e3 hours to remove carbonate concretions, which usuallyobscured the bone surfaces. This cleaning procedure did not

660 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

cause damage to bone surfaces, as the concretions usually dis-solved in the acid, exposing the undamaged outer cortex ofbones. Mechanical removal of the concretions, which couldhave damaged bone surfaces, was unnecessary. Followingthis procedure, the bones were soaked in KOH solution tobuffer the acid, and then washed and dried. All bones weresaved, and the identifiable specimens were taken for analysis.Identifiable elements included long-bone articular ends; long-bone shaft fragments with diagnostic zones (Stiner, 2002b,2004) and/or indicative characteristics such as thickness andmorphology of the cross section and medullary cavity (Barbaand Dominguez-Rodrigo, 2005); and teeth, cranial fragments,ribs, vertebrae, and all other recognizable bone fragments.

Anatomical identifications and taxonomic affiliation of thefaunal remains were based on the taphonomic and zooarchaeo-logical comparative collections of the Laboratory of Zoo-archaeology, University of Haifa, and the Department ofEvolution, Systematics and Ecology, Hebrew University ofJerusalem. Classification of cervid bones was achieved usingLister’s (1996) criteria. Due to the high fragmentation of thebone assemblage, the majority of the identified specimenswere assigned to one of three ungulate body-size classes:small, medium, and large. The small-sized ungulate class al-most entirely comprises gazelles; the medium-sized ungulateclass is dominated by fallow deer but includes the remainsof red deer, wild boar, and goat; and the large-sized ungulateclass is entirely composed of aurochs remains (Table 1).

All identified specimens (henceforth NISP; fragmentswhose precise location in the skeletal element, or portionthereof, can be determined and quantified, and can be assignedto species or size class) were recorded according to skeletal el-ement (e.g., proximal shaft of a humerus) and coded accordingto Lam et al.’s (1999: Fig. 1) scan-site codes (e.g., a proximalshaft of a humerus was coded as HU2). In addition, the loca-tion of each element (e.g., dorsolateral) and its completeness(i.e., percentage of the portion of the element that is repre-sented) was documented (following Klein and Cruz-Uribe,1984). In recording limb-shaft fragments, we used both ‘‘diag-nostic zones’’ (following Stiner, 2002b, 2004) and other mor-phological characteristics of the shaft fragments (e.g., Barbaand Dominguez-Rodrigo, 2005). The former refer to, for ex-ample, nutrient foramina, tuberosities and grooves, and por-tions thereof, and the latter refer to indicative characteristicssuch as cortical thickness and morphology of the cross sectionand medullary cavity. The completeness of these morphologi-cal traits was quantified by assigning percentage of complete-ness. This enabled us to compute the minimum number of

elements (MNE; Lyman, 1994) and minimum animal units(MAU) of every fraction of an element and each skeletal ele-ment. This procedure was designed to achieve a maximumaccuracy of the MNE count, in light of recent critiques of iden-tification procedures that are biased against shaft fragments(Marean and Kim, 1998; Pickering et al., 2003; Mareanet al., 2004). Since fragments whose precise location in theskeletal element could not be determined and quantifiedwere not included in the NISP, all identified specimens con-tributed to the MNE counts.

All identified specimens were systematically examined forbone-surface modifications using a stereoscopic microscopewith a high-intensity oblique light source, at 10� magnifica-tion, following the procedure described by Blumenschineet al. (1996). In this study, the NISP serves as a representativesample for bone-surface-modification data, as it includesnumerous long-bone shaft fragments and all other skeletal el-ements. We searched for cut marks (Binford, 1981); hammer-stone percussion marks, including conchoidal notches (Bunn,1981; White, 1992; Capaldo and Blumenschine, 1994; Picker-ing and Egeland, 2006) and percussion pits and striations (Blu-menschine and Selvaggio, 1988; Blumenschine et al., 1996;Pickering and Egeland, 2006); carnivore punctures, scoring,and digestion marks (Binford, 1981; Stiner, 1994); rodentgnaw marks (Brain, 1981; Rabinovich and Horwitz, 1994);biochemical (root) marks (Behrensmeyer, 1978; Binford,1981; Dominguez-Rodrigo and Barba, 2006); trampling marks(Behrensmeyer et al., 1986; Fiorillo, 1989; Oliver, 1989;Fisher, 1995); abrasion (Shipman, 1981; Shipman and Rose,1988); and weathering (Behrensmeyer, 1978). Burning was re-corded by bone color, following Stiner et al. (1995), andburned shaft fragments were recorded according to their exter-nal and internal surfaces, in order to discern the stage of burn-ing [i.e., fleshed bone, defleshed bone, or cracked bone,following Cain (2005)].

Mode of bone fragmentation was recorded for each shaftfragment that retained a portion of epiphysis or the shaftnear an epiphysis to determine the stage at which the boneswere broken (i.e., fresh-green vs. old-dry). The morphologyof the fracture angle and fracture outline was recorded follow-ing Villa and Mahieu (1991). We also recorded the percentageof shaft circumference (Bunn, 1983) to demonstrate the com-plete retrieval of the faunal remains (Marean et al., 2004).

The age structure of the main prey species was analyzed onthe basis of tooth eruption and wear, following Stiner (1994,2005). Due to small sample size, each isolated tooth or toothrow was assigned to one of three age classes (juvenile, prime,

Table 1

Ungulate body-size classes in the Misliya Cave assemblage

Size class Weight range

(kg)

Species in the Misliya assemblage Code

Gazelle size 15e30 Gazella gazella, Capreolus capreolus Gg size

Fallow deer size 60e250 Dama mesopotamica, Cervus elaphus, Sus scrofa, Capra sp. Dm size

Aurochs size 800e1000 Bos primigenius Bp size

Notes: The dominant species in each class appears in bold. Weight ranges are from Mendelssohn and Yom-Tov (1999) and Nowak (1999).

661R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

and old; see also Steele, 2005, for a detailed discussion of age-presentation methods). Deciduous or unworn teeth (dp4 or un-worn M3) were counted as juvenile; M3s or P4s in which thecrown height retained more than 50% of its original heightwere recorded as prime-adult; and M3s or P4s that retainedless than 50% of their original crown height were recordedas old-adult. When crown height could not be accurately mea-sured, we used the occlusal-surface scoring scheme (Grant,1982; Brown and Chapman, 1990).

Results

State of the faunal assemblage

The recovered bone assemblage of early Middle PaleolithicMisliya Cave is heavily fragmented and is dominated by ungu-late bone and tooth fragments. Complete bone elements arerare. No articulated bones were found and some fragment-refitting attempts failed. Overall, 13% of the plotted bonesand a much smaller (but uncounted) fraction of the faunalfragments collected following wet-sieving of the sedimentswere identified. This relatively low percentage is due to the in-tense fragmentation of the assemblage. A total of 1685 identi-fied specimens were recovered from the excavated area, and ofthese, 1584 are mammalian remains. The average concentra-tion of the faunal material in the excavation area is 565 iden-tified bone fragments per cubic meter. Notably, all excavationsquares contained numerous bone fragments and no ‘‘bone-poor zones’’ were encountered (e.g., Stiner et al., 2001).

Approximately one-third of the identified specimens(n¼ 494) were plotted and recorded in situ. The plotted bones(both identified and unidentified) were dispersed in randomorientations. Furthermore, the majority of bones were alignedhorizontally or nearly so, and those that were inclined show nouniform direction of inclination (Table 2). These observations,coupled with the well-defined hearth, the compacted sedi-ments, and the sharp and ‘‘fresh’’ state of the chipped flint ar-tifacts, imply that the Misliya Cave faunal assemblageremained in primary context sensu lato. The bone samplethus represents in situ human habitation, unmodified by suchagents as fluvial action or karstic activity.

Species representation and body-size classes

The identified faunal assemblage in Misliya Cave is over-whelmingly dominated by ungulate bone and tooth fragments(Table 3). Most of the specimens could not be confidently as-signed to genus or species, and consequently were identified toone of the ungulate size groups defined previously (Table 1).The species-specific elements are mainly teeth for the fallowdeer and aurochs, and teeth and long-bone articular ends forgazelle. Thus, data on the relative frequency of species in Mis-liya is presented using molar and premolar teeth only, whichare durable and usually readily identifiable.

The most common species in Misliya Cave, based on dentalNISP (Table 3) are Mesopotamian fallow deer (Dama mesopo-tamica, 45%) and mountain gazelle (Gazella gazella, 32%)

and to a lesser extent, aurochs (Bos primigenius, 12%). Otherungulate species, represented only by a few specimens, in-clude wild boar (Sus scrofa), red deer (Cervus elaphus), roedeer (Capreolus capreolus), and goat (Capra sp.). Smallgame includes several bones of hare-sized mammalsdwhichcould not be identified with much certainty except for one

Table 2

The depositional state of the faunal remains in the center of the ‘‘soft’’ exca-

vation area and near the cave wall (square I10)

Center Cave wall

n % n %

Orientation

EeW 617 27% 207 38%

NeS 600 26% 148 27%

NEeSW 514 22% 92 17%

NWeSE 557 24% 96 18%

Inclination angle

0e25 1836 87% 411 85%

26e65 211 10% 64 13%

66e90 71 3% 8 2%

Direction of inclination

N 172 17% 35 15%

NE 125 12% 15 7%

E 138 14% 37 16%

SE 101 10% 27 12%

S 96 9% 44 19%

SW 102 10% 19 8%

W 132 13% 37 16%

NW 151 15% 15 7%

Notes: Data include all plotted bones (i.e., fragments �2.5 cm in maximum

dimension).

Table 3

Taxonomic composition of the Misliya Cave assemblage

Species NISP (MNI) % NISP teeth %

Mammals (NISP¼ 1584)

Dama mesopotamica 91 (7) 38% 79 45%

Gazella gazella 98 (6) 40% 56 32%

Bos primigenius 26 (3) 11% 21 12%

Sus scrofa 11 (1) 5% 4 2%

Capra sp. 8 (2) 3% 6 3%

Cervus elaphus 6 (2) 2% 6 3%

Capreolus capreolus 1 (1) 0% 1 1%

Procavia sp. 1 (1) 0% 1 1%

Speciesþ size class

Dm-size 897 (12) 57%

Gg-size 575 (6) 36%

Bp-size 57 (3) 4%

Fetus/neonatal ungulate 15 (3) 1%

Small mammals 13 (2) 1%

Reptiles (NISP¼ 50)

Testudo graeca Boneþ shell 15þ 35 (4)

Birds (NISP¼ 51)

Alectoris chuckar 11 (4)

Other small/medium bird 12

Struthio camelus (eggshell only) 28

Notes: Mammals are listed by total NISP and by molar and premolar teeth

NISP. MNI values are given in parentheses. Ungulate size classes include

the NISP of the dominant animal in each class.

662 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

hyrax (Procavia sp.) mandibledas well as tortoises and birds.This category is dominated by 50 tortoise (Testudo graeca)bone and shell fragments, and by some partridge (Alectorischuckar) remains. Also present are 28 fragments of ostrich(Struthio camelus) eggshell. No fish or marine mollusk re-mains were found. Notably, carnivore remains are entirelyabsent.

Ungulate mortality profiles

The ungulate mortality pattern was assessed by tooth erup-tion and wear. The fallow deer dental sample was the largestand the only one with a satisfactory, albeit small, samplesize (n¼ 10). The fallow deer mortality pattern is heavilydominated by prime-age individuals according to two possibledentition sequences (Table 4). It seems that prime-age gazellesand mostly juvenile aurochs were also acquired, but the sam-ples for these taxa are very small (Table 4).

Bone-surface modifications

We systematically checked all identified bone fragments forsurface modifications, following Blumenschine et al.’s (1996)protocol. Following the cleaning procedure outlined above, thebone surfaces in the assemblage generally show good preser-vation, enabling us to record various types of inconspicuoussurface marks. Table 5 summarizes the bone-surface-modifica-tion data for gazelle and fallow deer size classes. Few speci-mens (2%) display a weathering stage higher than 2 [ofBehrensmeyer’s (1978) six weathering stages], and abrasiondamage is very rare (NISP¼ 3, 0.2%). This indicates rapidburial of the faunal material in favorable conditions and lackof fluvial or other geological postdepositional bone transport.

The most abundant bone-surface modifications are rootmarks (73% for both fallow deer and gazelle classes) andtrampling striations (21% for fallow deer, 12% for gazelle).Furthermore, we have found that bone fragments with rootmarks and trampling marks are significantly longer than frag-ments that do not bear these modifications in both size classes(Student’s t-test for root marks: fallow deer, t¼ 4.4, p< 0.001;gazelle, t¼ 6.7, p< 0.001; for trampling: t¼ 6.7, p< 0.001and t¼ 7.1, p< 0.001, respectively). This result suggeststhat the larger the surface area of a bone fragment, the strongerwill be the impact of root activity and trampling on it. Themarked decrease in trampling modifications on the smaller ga-zelle bone fragments may be explained by this phenomenon.

Table 4

Mortality profiles of fallow deer, gazelle, and aurochs from Misliya Cave,

using tooth eruption and wear

Dentition sequence Juvenile Prime Old

Dama mesopotamica unworn M3eworn M3 2 7 1

unworn M3eworn P4 2 5 2

Gazella gazella dp4eworn M3 0 3 0

dp4eworn P4 0 1 1

Bos primigenius dp4eworn M3 2 1 0

Evidence for carnivore-induced modifications is extremelyrare in the Misliya assemblage, despite the systematic micro-scopic analysis. Only three specimens bearing carnivore toothpits were found, and these remains are fallow-deer-sized. Inaddition, no digested specimens or specimens with ruggededges were found in the assemblage. Similarly, only a singlerodent-gnawed specimen was found (Table 5). These veryrare occurrences of animal-induced bone modifications sug-gest that carnivores played a minimal, if any, role in creatingand modifying the Misliya Cave assemblage. We also notethat no carnivore remains were retrieved during excavation.

Several dozen identified specimens bear cut (butchery)marks (3% of fallow-deer-sized and 4% of gazelle-sized re-mains; based on NISP, excluding teeth). Their location andfunction (following Binford, 1981) are detailed in Table 6.All butchery marks associated with the fallow deer size classare attributed to dismemberment, filleting, periosteal cleaning,or evisceration, whereas the gazelle size class also exhibitsthree cut-marked phalanges, probably attributable to skinning.The proportional distribution of cut marks by anatomical unitshows that gazelle and fallow deer were probably butchered ina similar fashion, which is not surprising for these similarlybuilt ungulates. However, unlike gazelle, fallow deer skeletal

Table 5

Values of key taphonomic variables for fallow deer and gazelle size classes

Dm-size Gg-size

Total NISP 897 575

NISP excluding teeth

(limb-shaft NISP)

818 (511) 502 (285)

Density-mediated attrition

and fragmentation

Correlation BMD*MAU y¼ 0.34x� 0.09 y¼ 0.31xþ 0.02

rs¼ 0.48, p < 0.001 rs¼ 0.39, p¼ 0.001

MNE of cranial

bones to teeth

1/3 4/6

P/D humerus MNE 1/2 1/5

P/D tibia MNE 0/4 2/4

Complete astragali

(total astragali)

2 (7) 5 (6)

Total NISP/MNE 4.02 3.44

Bone-surface modification

% Trampling 21 12

% Root marks 73 73

% Weathering (�3) 3 2

% Abrasion 0.2 0.2

% Carnivore gnaw 0.3 0

% Rodent gnaw 0 0.2

Human subsistence behavior

Abundant skeletal elements mandibles, limbs heads, limbs, pelves

% Cut marks by

NISP (MNE)

3 (8) 4 (9)

% Hammerstone marks

by NISP (MNE)

6 (18) 4 (17)

% Green fractures

by NISP (MNE)

60 (78) 69 (80)

% Burned NISP 21 23

Correlation FUI*MAU y¼�1E-05xþ 0.34 y¼�4E-05xþ 0.59

rs¼�0.32, p¼ 0.25 rs¼�0.37, p¼ 0.18

Correlation

marrow*NISP/MNE

y¼ 0.04xþ 3.3 y¼ 0.06xþ 3.64

rs¼ 0.667, p¼ 0.05 rs¼ 0.43, p¼ 0.24

Notes: P¼ proximal, D¼ distal.

663R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

extremities (toe and head parts) bear no marks (Fig. 2). Thedistribution and functions of cut marks suggest that all butch-ery activities of gazelle were performed on site, while fallowdeer underwent only dismemberment and filleting on site.We should, however, treat this interpretation with caution be-cause skinning operations are less likely to produce butcherymarks than dismemberment or filleting, and because the inter-pretation of butchery marks on the metapodial shafts is notclear-cut, considering the differences between caribou (uponwhich Binford’s butchery model is based) and the Mesopota-mian fallow deer [see Abe (2005) for documentation of differ-ential skinning treatment of caribou feet, which carry valuablefur].

Hammerstone percussion marks (including pits, striations,and percussion notches) were identified on 33 fallow-deer-class specimens (6% of total limb NISP) and on 12 gazelle-class specimens (4%). The frequency of hammerstone-marked

Table 6

Cut-mark location and function (following Binford, 1981) in the three major

size classes

Element Dm-size Gg-size Bp-size

n Function n Function n Function

Mandibular condyle 1 D

Thoracic vertebrae 1 D

Lumbar vertebrae 1 D

Ribs 1 E 2 D, E

Humeral shaft 2 F, D 1 D 1 F

Distal humerus 1 D

Proximal radius 1 D

Radial shaft 1 F 2 D, F

Proximal ulna 1 D

Ilium 1 D

Femoral shaft 1 F 1 F

Tibial shaft 1 F 3 F

Proximal calcaneus 1 F

Proximal metapodial 7 D 2 D

Metapodial shaft 4 C, D (3) 2 D

Phalanges 3 S

Notes: n indicates the number of fragments with cut marks. Function codes are

as follows: S¼ skinnning, D¼ dismemberment, F¼ filleting, C¼ periosteum

cleaning, E¼ evisceration.

0%

1%

2%

3%

4%

5%

6%

7%

8%

head axial upperlimbs

inter.limbs

lowerlimbs

feet

Cu

tm

arked

N

IS

P

Dm-size Gg size

Fig. 2. Cut-mark frequency by anatomical region: fallow deer vs. gazelle size.

‘‘Head’’ includes the skull and mandible; ‘‘axial’’ includes the vertebrae, ribs,

and pelvis; ‘‘upper limbs’’ refers to the scapula, humerus, and femur; ‘‘inter-

mediate limbs’’ refers to the radius and tibia; ‘‘lower limbs’’ refers to the meta-

podials; and ‘‘toes’’ refers to phalanges and sesamoids.

specimens within each limb bone varies greatly (Table 7). Themarks are abundant on the humeri and femora of the fallowdeer class and on the radii of the gazelle class, but rarer onthe tibiae and metapodials of the two groups. This phen-omenon is not easily explained. Calculating the MNE ofhammerstone-marked specimens factors out the differentialfragmentation of limb bones. Still, MNE-based frequenciesshow the same general trend as the NISP-based frequencies,namely remarkable differences between the limb bones (Table7). However, the location of the marks is repeated in a consis-tent manner within each limb bone (Table 7), indicatingsystematic marrow processing by the inhabitants of the site(cf. Munro and Bar-Oz, 2005; Pickering and Egeland, 2006).

Limb-bone fracture patterns

Analysis of breakage patterns (fracture angle and fractureoutline; following Villa and Mahieu, 1991) of 124 fallow-deer-sized and 65 gazelle-sized long-bone shaft fragments in-dicates that the majority of bones were broken while fresh, asis evident from the high proportion of green-bone fractures(oblique angle and curved, or V-shaped, outline; Table 8). Inboth size classes, the forelimb bones display more green frac-tures than the hindlimb bones. This could result from the gen-erally greater level of fragmentation of the long-bone elementsfrom the hindlimb (based on the NISP/MNE index; Table 7),which might have inflated the number of dry breaks on hin-dlimb elements. It is highly probable that the green breaksin the assemblage reflect intentional opening of the medullarcavity for marrow because 10% of the specimens in thislimb-bone subsample display hammerstone percussion marks,yet none displays signs of carnivore activity (Table 8). The ev-idence for systematic marrow processing at Misliya Cave isfurther strengthened by the shaft-circumference data. Almostall shaft fragments in the assemblage, regardless of bodysize, retained less than half of their original circumference(94%; total limb-shaft NISP¼ 710).

Skeletal-element frequencies

The NISP and MNE values for every element in each sizeclass are presented in Appendix A. Figure 3 shows theskeletal-element representation of the gazelle (Gg), fallowdeer (Dm), and aurochs (Bp) size classes (%MAU). The fallowdeer and gazelle skeletal profiles are biased against the verte-bral column and neck but are relatively rich in fore- and hin-dlimbs. Antler/horn and toes also appear in low frequenciesfor both taxa. The fallow deer size class shows high valuesfor metatarsals, moderate values for other long bones and man-dibles, and low frequencies of skulls, pelves, and scapulae. Onthe other hand, the gazelle skeletal-element profile is rich incrania and mandibles, pelves, and long bones. Head partsand tibiae are common among the aurochs remains.

The differences between the two most abundant size clas-ses, the fallow deer and the gazelle, are worth noting. Gazellemandibles and skulls are found in similar numbers, whereas inthe fallow deer size class, the skull is underrepresented. Also,

664 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Table 7

Characteristics of hammerstone-marked specimens within each limb bone: mark frequency compared with the fragmentation (NISP/MNE) of that bone (all spec-

imens) and anatomical location of hammerstone marks on each limb bone

Element Dm-size Gg-size

Marked

NISP

% NISP

marked

% MNE

marked

NISP/

MNE

Bone portion Aspect of

bone

Marked

NISP

% NISP

marked

% MNE

marked

NISP/

MNE

Bone portion Aspect of

bone

Humerus 7 16% 30% 4.30 2 midshaft,

5 distal shaft

5 caudal,

2 cranial

1 2% 9% 3.91 Distal shaft Caudal

Radius 2 4% 8% 4.38 1 midshaft,

1 proximal shaft

All medial 4 11% 22% 4.00 2 proximal

shaft, 1 midshaft,

1 distal shaft

All cranial

Femur 14 19% 42% 6.17 8 proximal shaft,

6 midshaft

10 cranial,

4 lat/med

1 2% 13% 6.88 Proximal shaft Lateral

Tibia 4 5% 38% 10.13 All midshaft 1 cranial,

3 lat-caud

3 5% 20% 5.90 All distal shaft All caudal

Metapodial 6 2% 6% 7.47 All midshaft All lat/med 3 3% 18% 8.64 All midshaft All lat/med

the pelvis is fairly well represented in the gazelle class, unlikethe fallow deer class (Fig. 3). This suggests a different trans-port pattern for the two animals, in which small ungulateprey (gazelle) were brought to the site more often as completecarcasses than medium-sized prey (fallow deer). In order todiscern whether this pattern resulted from differential humantransport and to isolate possible preservation biases, we as-sessed the role of skeletal-part attrition and postdepositionalprocesses in shaping the observed skeletal-element profiles.

The ratio of skull bones to maxillary teeth (based on MNE;after Stiner, 1994) indicates underrepresentation of skull bonesin both gazelle and fallow deer classes (Fig. 4). These ratiossuggest that decomposition and fragmentation reduced someof the less resistant skull elements to unidentifiable fragments.It also indicates a greater frequency of destruction of fallowdeer skulls compared with gazelle skulls (bone/tooth MNEis 1:3 for fallow deer and 2:3 for gazelle). Gazelle skull

MNE values were determined by the petrous bone, usuallyfound complete and easily identified to size class (Bar-Ozand Dayan, 2007), while no fallow deer petrous bones werefound. The paucity of the well-preserved and easily identifiedpetrous bones in the fallow deer class probably indicates thattheir skulls were seldom imported to the site. The ratios ofproximal (low density) to distal (high density) gazelle- and fal-low-deer-sized humeri and tibiae show a similar trend (basedon MNE; after Binford, 1981), demonstrating that denser dis-tal parts are more represented in both taxa (Table 5; Fig. 4).

The complete recovery and identification of limb-boneshaft fragments in Misliya enabled us to compare the ratiosof MNE values for limb-bone ends and limb-bone shafts. Gen-erally, limb-bone shafts are much denser than limb-bone ends(Lam et al., 1999, 2003). In Misliya Cave, bone-shaft MNEvalues are greater than bone-end MNE values for nearly alllimb bones of both size classes, with the only exception being

Table 8

Characteristics of a limb-bone subsample composed of shafts with attached epiphysis or near-epiphyseal shafts: fracture patterns and selected bone-surface-

modification data

Dm size Gg size

Humerus Radius Metacarpal Femur Tibia Metatarsal Humerus Radius Metacarpal Femur Tibia Metatarsal

n 20 15 15 27 20 26 12 9 4 16 12 12

Fracture angle

Oblique 16 13 11 20 11 19 10 8 4 13 7 7

Right 4 2 4 7 9 7 2 1 0 3 5 5

Fracture outline

Curved 12 12 11 14 11 18 10 7 4 11 8 6

Transverse 8 3 4 13 9 8 2 2 0 5 4 6

Shaft circumference

<50% 20 15 14 27 19 25 12 9 4 16 11 12

�50% 0 0 1 0 0 1 0 0 0 0 0 0

100% 0 0 0 0 1 0 0 0 0 0 1 0

Hammerstone marks

n 5 0 0 7 2 0 1 2 0 1 1 0

% 25% 0% 0% 26% 10% 0% 8% 22% 0% 6% 8% 0%

Carnivore gnaw

n 0 0 0 0 0 0 0 0 0 0 0 0

% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0% 0%

Notes: Fracture angle and outline after Villa and Mahieu (1991) and shaft circumference after Bunn (1983).

665R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

0%

20%

40%

60%

80%

100%

horn

/ant

ler

skul

l

man

dibl

e

cerv

ical

thor

acic

lum

bar

ribs

scap

ula

hum

erus

radi

us+u

lna

carp

als

met

acar

pus

pelv

is+s

acru

m

fem

ur

tibia

tars

als

met

atar

sus

toes

MA

U

Gg-size Dm-size Bp-size

Fig. 3. Skeletal-element frequency of gazelle, fallow deer, and aurochs size classes. Data are from Appendix A.

Fig. 4. MNE values of skull bones and dentitions and of long-bone portions

for fallow deer and gazelle size classes. Metapodia are presented separately

because of their higher MNE values.

the gazelle metapodials (Fig. 4). The most abundant limb bonein the small sample of aurochs remains, the tibia (MNE¼ 4),shows the same pattern (Appendix A). This result indicatesdifferential survivorship of long-bone parts, with the densershaft fragments determining the MNE of the complete bonein almost all cases. Assuming that limb bones were introducedcomplete to the site, the underrepresentation of their endscompared to their shafts indicates a significant density-medi-ated destruction of long bones. In addition, we found a signif-icantly positive relationship between bone survivorship(%MAU) and bone mineral density (BMD1,2 values for rein-deer, Rangifer tarandus; Lam et al., 1999) for elements inthe fallow deer (Spearman’s r¼ 0.48, p< 0.001) and gazelle(Spearman’s r¼ 0.39, p¼ 0.001) size classes (Fig. 5; regres-sion equations are provided in Table 5). These results furtherindicate a pronounced density-mediated bias in skeletal-partrepresentation.

The degree of bone fragmentation is quite high in the as-semblage. Interestingly, positive correlations were found toexist between the bone fragmentation (NISP:MNE ratios, fol-lowing Lyman, 1994) of marrow-bearing elements from ga-zelle and fallow deer size classes and their bone-marrowindex values [based on Maddrigal’s (2004) marrow valuesfor white-tailed deer (Odocoileus virginianus) and Bar-Ozand Munro’s (2007) for Gazella gazella; see Fig. 6]. The fal-low deer class showed a significant correlation (Spearman’sr¼ 0.667, p¼ 0.05) and the gazelle class displayed a nonsig-nificant positive trend (Spearman’s r¼ 0.43, p¼ 0.24). Thisresult suggests that some of the fragmentation in the assem-blage is due to deliberate breakage of bones to extract marrow.The selective destruction of bones for marrow extraction isalso demonstrated by the high rate of fresh-bone breakage(discussed above).

The intensity of bone fragmentation is not uniform acrossskeletal elements and size classes. The NISP:MNE ratiosshow that elements in the fallow deer class are generally morefragmented than gazelle elements (Fig. 7), implying that thesmall ungulate is better preserved than the medium ungulate.The completeness of the astragalus is considered to be a goodmeasure of postdepositional destruction (following Marean,

666 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

1991). This measure, too, points to greater fragmentation of themedium ungulate, even though caution is required because ofthe small sample size. In the fallow deer class, there are onlytwo complete astragali out of seven, whereas five gazelle astra-gali out of six are complete or nearly so (Table 5).

The Misliya Cave assemblage does not show a positive re-lationship between bone survivorship and food-utility index(FUI, based on the weight of useable tissue of Rangifer taran-dus; Metcalfe and Jones, 1988) for both fallow deer (Spear-man’s r¼�0.316, p¼ 0.25) and gazelle (Spearman’sr¼�0.368, p¼ 0.177) classes (Fig. 8; regression equationsare provided in Table 5). This result is hardly surprising, giventhe amount of density-mediated attrition that masks human-transport patterns and because some of the most nutritious el-ements in the skeleton are also the least likely to be preserved.The densest high-surviving elements (head and limb bones)are generally represented more than the low-surviving

(a)

0%

20%

40%

60%

80%

100%

0.2 0.7 1.2BMD

0.2 0.7 1.2BMD

MA

U

(b)

0%

20%

40%

60%

80%

100%

MA

U

Fig. 5. Relationship between bone mineral density (BMD1,2 values for Rangi-

fer tarandus; Lam et al., 1999) and bone survivorship (%MAU) for all ele-

ments in the fallow deer (a) and gazelle (b) size classes.

elements, regardless of their nutritional value (Fig. 8). Amongthe high-surviving limb-bone elements, there is no apparentrank of representation between more nutritious and less nutri-tious bones (Fig. 8; note separate regression lines for thehigh-survival and the low-survival sets). This result is to be ex-pected, given the similar characteristics of different limbbones when ease of dismemberment and transport are con-cerned. Therefore, the possibility of selective transport to Mis-liya Cave should be discerned by comparing elements withsimilar densities and different structures, and consequently,different nutritional values and return rates (see Marean andCleghorn, 2003; Cleghorn and Marean, 2004).

Burning processes

The Misliya Cave assemblage contains a high proportionof bones burned to various degrees. Burning incidence is

(a)

PH1PH2

HUMC R/U

MT FE

TI

MAN

1

2

3

4

5

6

7

8

9

10

11

0 50 100 150 200Marrow (KCAL)

NIS

P/M

NE

(b)

MT

HUR/U

TI

FEMC

MAN

PH1

PH2

1

2

3

4

5

6

7

0 20 40 60Marrow (KCAL)

NIS

P/M

NE

Fig. 6. Relationship between marrow content (for Odocoileus virginianus: Mad-

drigal, 2004; for Gazella gazella: Bar-Oz and Munro, 2007) and the fragmenta-

tion (NISP/MNE) of marrow-bearing bones of fallow deer (a) and gazelle (b) size

classes. Abbreviations are as follows: FE¼ femur; HU¼ humerus; MAN¼mandible; MC¼metacarpal; MT¼metatarsal; PH1¼ proximal phalanx;

PH2¼ intermediate phalanx; R/U¼ radiusþ ulna; TI¼ tibia.

667R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

similar for both fallow deer and gazelle classes (21% and23% of NISP, respectively). Burning occurs at higher fre-quencies on the unidentified specimens (ca. 40%, based ona sample of 2334 unidentified fragments). The spatial distri-bution of burned NISP is neither uniform nor random. Acomparison of three excavation unitsdinside the hearth(squares L8e9), around the hearth, and near the cave wall(square I10; see plan in Fig. 1)dshows that the burningfrequency differs significantly between the excavation areas.As expected, the frequency is higher inside the hearth(NISP¼ 26, 46% burned) than around it (NISP¼ 124, 18%)and near the cave wall (NISP¼ 221, 23%) (c2¼ 8.76,p¼ 0.01).

Further investigation of burning patterns reveals that limb-bone ends are somewhat more burned than limb-bone shafts inboth size classes (Table 9). It could be that this pattern resultedfrom the roasting of meaty limbs while the epiphyses wererelatively more exposed due to dismemberment (e.g., Speth,2000). However, the proportions of shaft fragments displayingequal burning intensity on their exterior and interior surfaces ishigh (Table 9), indicating that most of the burning resultedfrom nonnutritive incidents that occurred following defleshingand breakage of long bones, presumably for marrow consump-tion. Burning frequency and food utility are positively and sig-nificantly correlated for the gazelle class (Spearman’sr¼ 0.58, p¼ 0.05), but not so for the fallow deer class (Spear-man’s r¼ 0.25, p¼ 0.44; Table 9). This difference may par-tially stem from the higher burning of fallow deer head andtoe parts (Fig. 9). In addition, burned bone fragments in thefallow deer class are significantly shorter than unburned frag-ments (Student’s t-test: t¼ 4.00, p< 0.001; Table 9), indicat-ing that exposure to fire played some role in increasingfragmentation damage to fallow deer elements. This patternwas not observed for gazelle, where burned bone fragmentsare similar in length to unburned fragments (t¼ 0.02;p¼ 0.981).

MP

FE

SKTIMAN

PE

CESPPH2

LUPH1H/A

TH

HURI R/U

CARAS PH3 CA

0

2

4

6

8

10

0 2 4 6 8 10Dm size

Gg

size

Fig. 7. Bivariate scatterplot of NISP:MNE ratios of skeletal parts from gazelle

and fallow deer size classes. The diagonal line represents equal fragmentation.

Abbreviations are as follows: AS¼ astragalus; CA¼ calcaneus; CAR¼ carpals;

CE¼ cervical vertebra; FE¼ femur; H/A¼ hornþ antler; HU¼ humerus;

LU¼ lumbar vertebra; MAN¼mandible; MP¼metapodials; PE¼ pelvis;

PH1¼ proximal phalanx; PH2¼ intermediate phalanx; PH3¼ distal phalanx;

RI¼ rib; R/U¼ radiusþ ulna; SK¼ skull; SP¼ scapula; TH¼ thoracic verte-

bra; TI¼ tibia.

Discussion

Taphonomic history of the Misliya Cave assemblage

The early Middle Paleolithic faunal assemblage of MisliyaCave was created solely by humans. This is evident by the un-gulate-dominated taxonomic composition and the abundanceof butchery marks, burning signs, and hammerstone

(a)

THRI

SP

CE

LU

SK

PH PE

TI

HUMC

MANR/U

FE

MT

0%

20%

40%

60%

80%

100%

0 2000 4000 6000FUI

MA

U(b)

FE

TI

HU

R/U

PEMT

MANSK

PH

MC

SP

LURI

THCE0%

20%

40%

60%

80%

100%

0 2000 4000 6000FUI

MA

U

Fig. 8. Relationship between food-utility index (FUI: Metcalfe and Jones,

1988) and bone survivorship (%MAU) for bones in (a) fallow deer and (b) ga-

zelle size classes. High-survival bones are indicated by squares and low-sur-

vival bones are indicated by small diamonds. Separate regression lines are

shown for the high-survival set and the low-survival set in each graph. Abbre-

viations are as follows: CE¼ cervical vertebra; FE¼ femur; HU¼ humerus;

LU¼ lumbar vertebra; MAN¼mandible; MC¼metacarpals; MT¼metatar-

sals; PE¼ pelvis; PH¼ phalanges; RI¼ rib; R/U¼ radiusþ ulna; SK¼ skull;

SP¼ scapula; TH¼ thoracic vertebra; TI¼ tibia.

668 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

percussion marks on bone surfaces. Since the assemblagelacks carnivore remains, and carnivore gnaw marks wererarely detected, it appears that carnivores played virtually norole in creating and modifying the assemblage. This phenom-enon is known from other Levantine MP cave sites (e.g.,Amud Cave: Rabinovich and Hovers, 2004; Dederiyeh Cave:Griggo, 2004; and, to some extent, Hayonim Cave: Stiner,2005). It can be explained by intensive human occupationand carcass processing, which denied carnivores access tothe faunal refuse (e.g., Bunn et al., 1988; Yellen, 1991;Bunn, 1993) or made it unattractive to carnivores due to roast-ing (e.g., Lupo, 1995). In contrast, many carnivore remainsand ravaging modifications have been found in other Levan-tine MP caves, such as Kebara and Qafzeh, which probablyserved as carnivore dens at certain times (Bar-Yosef et al.,1992; Dayan, 1994; Rabinovich et al., 2004).

Fluvial transport and other geological processes character-istic of cave sites can also be ruled out as a primary or second-ary agent of assemblage formation, because the bone surfaces

Table 9

Characteristics of burned specimens from fallow deer and gazelle classes from

Misliya Cave

Dm size Gg size

NISP % NISP %

Limb burning

Equally burned 90 93% 71 89%

More burned outside 3 3% 6 8%

More burned inside 4 4% 3 4%

Burned bone end 29 26% 29 32%

Burned bone shaft 79 20% 58 30%

Correlation FUI*element-

burning frequency

y¼ 8E-06xþ 0.18 y¼ 4E-05xþ 0.14

rs¼ 0.25, p¼ 0.443 rs¼ 0.58, p¼ 0.048

Burned-fragment length

(mm)

n¼ 177 n¼ 125

Mean¼ 31.86 Mean¼ 25.86

SD¼ 16.27 SD¼ 11.28

Unburned-fragment length

(mm)

n¼ 613 n¼ 393

Mean¼ 37.90 Mean¼ 25.82

SD¼ 21.62 SD¼ 13.12

Burned- vs. unburned-

fragment lengths (t-test)

t¼ 4.00, p < 0.001 t¼ 0.02, p¼ 0.981

0%5%

10%15%20%25%30%35%40%

Bu

rn

ed

N

IS

P

Dm-size Gg-size

head axial upperlimbs

inter.limbs

lowerlimbs

feet

Fig. 9. Relative frequencies of burned NISP by anatomical unit for fallow deer

and gazelle size classes. See the caption of Fig. 2 for the elements included in

each anatomical region.

are generally well preserved and do not exhibit signs ofabrasion, rolling, or subaerial weathering. Moreover, bone ori-entations are random, most of the bone fragments were hori-zontally aligned or nearly so, and burned fragments aremore abundant inside the hearth. All these lines of evidenceimply that the Misliya Cave faunal assemblage representsthe discard of human-acquired animal carcasses, followingtheir processing for meat and marrow.

Human processing behavior in Misliya Cave was generallysimilar for both fallow deer and gazelle. Following dismem-berment, some body parts were roasted, as indicated by thehigher frequency of burned epiphyses. Long bones were thenfilleted and broken for marrow using a hammerstone, creatingabundant percussion marks and green-bone fractures. The highfrequency of equally burned exterior and interior surfaces ofshaft fragments may point to some bone discard into the hearthfollowing consumption of meat and marrow. The significantconcentration of burned bones inside the well-defined hearthand the correlation between burning intensity and food utilityfor gazelles indicate that the observed burning patterns in Mis-liya Cave relate mostly to human activities and are not a prod-uct of postdepositional processes (e.g., Stiner et al., 1995;Shahack-Gross et al., 1997).

The significant density-mediated attrition of the assemblageprobably stems from destruction of porous bone parts as a re-sult of human bone processing, as well as plant activity andtrampling. It is difficult to assess the impact of root activityon the faunal remains because this subject has not beenthoroughly investigated (see Lyman, 1994: 375e377; Domi-nguez-Rodrigo and Barba, 2006, and references therein). Itis reasonable to assume that root activity will enhance thecracking of bone, especially those that have been cracked byhammerstone and burned. Similarly, human trampling andtrampling caused by sediment compaction may crack bonesof small and medium ungulates (Fiorillo, 1989; Fisher,1995). The high occurrence of trampling marks in the assem-blage provides a partial explanation for its high level offragmentation.

Density-mediated attrition processes affecting the MisliyaCave assemblage facilitated the survival mainly of limb-bone shaft fragments and teeth, and eliminated most long-bone ends, vertebrae, ribs, and skull bones. Denser parts ofthese bones do exist in the assemblage (long-bone shafts, pel-vic shafts, and the maxillary teeth and petrosum), indicatingthat the ‘‘deleted’’ elements were indeed brought to the sitebut were differentially preserved.

It appears that fallow-deer-sized (medium ungulate) ele-ments in the Misliya assemblage were less frequently pre-served and more fragmented than gazelle-sized (smallungulate) elements. Does this phenomenon relate to differen-tial human treatment of prey size, or does it result from differ-ential postdepositional processes? Some hunter-gatherers tendto break bones of large animals more intensively than bones ofsmall animals because it takes more effort to extract their mar-row or to fit them into the cooking/roasting facilities (e.g., Bar-tram and Marean, 1999; Klein et al., 1999; Speth, 2000).However, we found great similarity in human processing of

669R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

the two animals, as is evident from the similar frequencies ofbutchery marks, hammerstone percussion marks, and green-bone fractures by %NISP and by %MNE, which factors outdifferential fragmentation (Table 5). Because larger bones inthe assemblage display significantly more trampling and rootmarks than smaller bones, it seems that the better preservationof the small ungulate is not linked to any differential humantreatment of fallow deer and gazelle carcasses. Rather, itmay be linked to stronger destruction processes affecting theremains of the larger ungulate, perhaps because of their greatersurface area (but see Lyman, 1994: 397e398, and referencestherein).

In spite of the significant effect of density-mediated attri-tion on skeletal-element frequency in Misliya, some humantransport patterns may be discerned. The gazelle skeletal-ele-ment frequency is quite balanced and points to transport ofbulky heads and pelves and thus of at least some of the axialskeleton. Conversely, fallow deer skulls and axial parts wereseldom transported, as evidenced by the underrepresentationof dense maxillary teeth, the petrosal part of the cranium,and the pelvic shafts. The emerging scenario is that of trans-port of complete gazelle carcasses and their butchery on site,as opposed to some field butchery of the fallow deer and trans-port of high-utility, low-bulk elements (limbs and marrow-richmandibles) back to the site. This difference in transport patternmay be attributed to the difference in size; the mountain ga-zelle is a small ungulate, weighing about 20 kg, and the Mes-opotamian fallow deer weighs 70e100 kg (Mendelssohn andYom-Tov, 1999). The inhabitants of Misliya Cave appear tohave carried the small gazelles complete for processing onsite, but preferred to transport the higher-return-per-weightportions (i.e., the least bulky body parts) of the heavier fallowdeer.

Hunting and subsistence patterns in the early MiddlePaleolithic

The early Middle Paleolithic inhabitants of Misliya Cavepossessed well developed large-game-hunting capabilities.This is inferred from three major lines of evidence. First,meat-bearing bones are abundant in the assemblage. Fallowdeer are represented by meat- and marrow-rich elements thatare easy to dismember and transport (i.e., limbs and mandi-bles) and gazelles are represented by virtually all high-surviv-ing body parts. This pattern of representation suggests thathumans had primary access to the carcass and could transportit whole to the site or choose the most desirable elements totransport, as is the case for recent foragers (e.g., the Hadza:O’Connell et al., 1988, 1990; Monahan, 1998). Second, fillet-ing cut marks were found on meaty limb-bone shafts, indicat-ing that meat was still attached to the bones at the time it wasacquired by humans. This also indicates primary access to thecarcass, by either aggressive scavenging or by hunting (e.g.,Blumenschine, 1986, 1988, 1995; O’Connell et al., 2002;Dominguez-Rodrigo, 2002; Dominguez-Rodrigo and Picker-ing, 2003). Third, clear evidence for hunting is provided bythe fallow deer mortality pattern, which is dominated by

prime-age adults. Taking prime-adult prey is widely consid-ered to be a uniquely human hunting strategy. This strategyis not employed by carnivores, which either hunt the weakestmembers of the herd (i.e., the juvenile or senile individuals) ortarget prey opportunistically (Stiner, 1990, 1991, 1993, 1994;Steele, 2004). Thus, the Misliya Cave mortality pattern indi-cates human acquisition of prey by hunting, with virtuallyno carnivore involvement. This pattern is unlikely to representa taphonomic bias against the remains of younger individuals(e.g., Munson and Marean, 2003) because it was recorded bydurable molar and premolar teeth and because the assemblagewas not ravaged by carnivores, did not undergo fluvial trans-port, and was fully screened and collected.

The conclusion we have reached for the early Middle Pa-leolithic assemblage of Misliya Cave, namely that the inhabi-tants of the site systematically hunted prime-age ungulates andtransported chosen carcass parts back to the site, accords withrecent taphonomic studies of other Levantine Middle Paleo-lithic assemblages. Analyses of late Middle Paleolithic assem-blages from Kebara Cave (Speth and Tchernov, 1998, 2001, inpress; Speth and Clark, 2006) and Amud Cave (Rabinovichand Hovers, 2004) show that humans had primary access tothe most nutritive carcass parts, and also point to the impor-tance of bulk in their transport decisions, meaning that skullsand axial elements are frequently underrepresented comparedto mandibles and limbs. Butchery marks are abundant onmeaty parts, and age profiles are usually prime-dominated. Im-portantly, these patterns were also found in the early MiddlePaleolithic (ca. 140e220 ka) of Hayonim Cave (Stiner,2005). The combination of these recent data sets and interpre-tations with those from the present study strongly supports thenotion that, at least as early as the early Middle Paleolithic, ca.250 ka, Levantine foragers acquired large game by systematichunting and transported the carcasses, or the desired partsthereof, to their ‘‘home bases’’ (e.g., Isaac, 1978). Similar con-clusions have been reached by several studies of MP/MSAsubsistence worldwide, using skeletal-element profiles andbutchery-mark data (e.g., Marean, 1998; Marean and Kim,1998; Milo, 1998; Marean et al., 2000; Bar-Oz and Adler,2005; Assefa, 2006; Cleghorn, 2006) or focusing on ungulatemortality profiles (e.g., Gaudzinsky and Roebroeks, 2000;Hoffecker and Cleghorn, 2000; Adler et al., 2006). Most ofthese studies describe late Middle Paleolithic assemblages,but the few detailed early Middle Paleolithic case studiesfrom the Levant and the northern Mediterranean coast donot differ in these respects. Rather, they display hunting strat-egies and transport patterns similar to later Middle Paleolithicassemblages (e.g., Lazaret Cave: Valensi and Psathi, 2004;Hayonim Cave: Stiner, 2005).

Some insights concerning Middle Paleolithic subsistencemay be gained from discussing the role of very large andvery small game in the economy of the period. Klein (1998,1999; Klein and Cruz-Uribe, 2000) pointed out that MiddleStone Age people generally refrained from hunting large,dangerous animals that display fierce defense tactics, presum-ably because of their inferior weapon technology. In the Le-vant, hunting of dangerous aurochs and wild boar is well

670 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

demonstrated in the late Middle Paleolithic of Kebara Cave(Speth and Tchernov, 1998, in press), and hunting of aurochsis evident in the early Middle Paleolithic of Hayonim Cave(Stiner, 2005). In Misliya Cave, however, wild boar is repre-sented by a single, juvenile individual and aurochs is repre-sented by only three individuals, two of which are juveniles.

Were aurochs hunted by the early Middle Paleolithic hu-mans in Misliya? Its skeletal-element profile, rich in head partsbut also tibiae and some pelvic parts, is ambiguous. However,the presence of a filleting mark on the humeral shaft of an au-rochs indicates primary access to this meat-rich body part. Theenormous body size of the aurochs (800e1000 kg; Nowak,1999) may have caused the Misliya humans to target juvenilesand/or to adopt a different transport approach than with thesmaller ungulates, namely filleting the animal in the fieldand discarding most bones there, taking only selected mar-row-rich bones, such as the tibia (cf. Monahan, 1998). Thistransport strategy would mean that the aurochs remains are un-derrepresented in the assemblage relative to their real eco-nomic importance (cf. LMP Amud Cave for a similarsuggestion; Rabinovich and Hovers, 2004). Nevertheless, sys-tematic hunting of aurochs is not clearly evident in MisliyaCave, in contrast to the great economic importance of aurochsin Hayonim Cave (Stiner, 2005). Rather than representing theearly Middle Paleolithic humans’ hunting deficiencies (e.g.,Klein, 1998, 1999), we echo Speth’s (2004a) opinion thatthe acquisition of ‘‘docile’’ animals more frequently than dan-gerous ones is a very sensible adaptation for Middle Paleo-lithic hunters. They may have preferred to successfully huntsmaller, prime-age ungulates that presumably supplied all ofthe protein and fat needs for a small forager group.

It appears that the meat diet of early Middle PaleolithicMisliya humans consisted almost exclusively of high-rankedungulate prey, which overwhelmingly dominates the faunal as-semblage. The fraction of lower-ranked prey (small mammalsand tortoises) in the assemblage is very small. Stiner and col-leagues (Stiner et al., 1999, 2000; see also Stiner, 2001; Stinerand Munro, 2002) suggested that reliance on slow-moving andslow-reproducing small game (e.g., tortoise) implies low pop-ulation densities. Conversely, reliance on fast-moving and fast-reproducing small game (e.g., hare and partridge) suggestsgreater human population densities, which caused an overex-ploitation of the easy-to-catch and slowly reproducing smallgame and forced humans to capture faster animals at highercosts (Stiner et al., 1999, 2000). According to this idea, theearly Middle Paleolithic of Hayonim Cave, rich in tortoise re-mains, was interpreted as a period of small, dispersed humanpopulations (Stiner, 2005). This assertion was supported byother evidence concerning the density of animal remains, in-trasite structure, site dispersion in the landscape, and lithictechnology (Hovers, 2001; Meignen et al., 2006). Conversely,the end of the Middle Paleolithic, as represented in KebaraCave, was viewed as a period of population growth and in-creasing predation pressure by Paleolithic humans, impliedby the phasing out of large-bodied prey, increased frequenciesof juvenile ungulates, and a faster rate of debris accumulation(Speth, 2004b; Speth and Clark, 2006; Meignen et al., 2006).

Not only does the Misliya Cave assemblage lack fast smallgame, it displays a very low abundance of small game in gen-eral. Consequently, it supports the notion that the early MiddlePaleolithic population in the Levant did not overexploit its en-vironment and did not need to frequently broaden its diet toinclude small game, much less to capture fast-running smallanimals. This may suggest, according to Stiner et al.’s(1999, 2000) approach, that the humans of Misliya Cave con-stituted a small and/or relatively mobile group of foragers ina landscape sparsely populated by other human groups.

However, this interpretation should be taken with cautionbecause other aspects of the early Middle Paleolithic habita-tion of Misliya Cave do not necessarily agree with thesmall-game-based interpretation. Namely, the density of lithicartifacts and bone fragments in Misliya is high; a well-definedhearth with a depth of more than 20 cm indicates repeatedburning in the same spot; and no sterile layers were detectedwithin the Early Middle Paleolithic sediments (unpublisheddata). Overall, the early Middle Paleolithic habitation in Mis-liya Cave does not seem ephemeral, in contrast to the adapta-tion suggested for Hayonim Cave (Stiner, 2005; Meignenet al., 2006). However, we should note that our interpretationis preliminarydat present the excavated area of Misliya issmall and the accumulation per TL years is not yet known. Ei-ther different environmental parameters that are not easily de-termined (i.e., topographical location or climatic setting) orhuman foraging decisions unknown to us may account forthe currently observed variance among these broadly contem-poraneous sites.

Conclusions

The characteristics of the Misliya Cave assemblage shedlight on the hunting strategies, transport patterns, and car-cass-processing behavior of early Middle Paleolithic humansin the Levant. Thus far, almost no zooarchaeological and taph-onomic data were available from the Mount Carmel caves(e.g., Tabun) in this critical period, and little is known in gen-eral. We have established that Misliya Cave humans systemat-ically hunted prime-aged ungulates, transported complete ordisarticulated carcasses back to their camp depending on car-cass size, roasted the meat, and cracked long bones for mar-row. The data from Misliya, discussed in relation to otherMP/MSA assemblages, should be considered in their widerevolutionary context. Namely, do they attest to modern humanbehavior and ways of life? As with lithic procurement strate-gies or intrasite patterns, subsistence strategies are not directproxies of human cognitive ability and ‘‘modernity.’’ This isbecause of the difficulty in isolating their causedthe appear-ance of modern behavior or simply ecological adaptations re-lated to, for instance, more intensified foraging in response toclimate changes (e.g., Henshilwood and Marean, 2003; Speth,2004a). Direct proxies for modern-human-like behavior areunique nonutilitarian items that can be seen as art objects ordecorations. These artifacts are abundant in the Upper Paleo-lithic but also appear sporadically in the MP/MSA, supportingthe notion that humans were already behaviorally ‘‘modern’’

671R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

by that period (e.g., McBrearty and Brooks, 2000, and refer-ences therein; Henshilwood et al., 2002, 2004).

Even though archaeofaunal studies may not contribute asdirectly as ‘‘symbolic objects’’ to solving the modern humandebate, their importance for illuminating the long-term pro-cesses of human subsistence and lifeways, which are wellreflected by carcass-acquisition patterns, transport, and pro-cessing, cannot be negated. These aspects have been inten-sively studied experimentally and ethnographically, andconsequently, they are understood better than some enigmaticmanifestations of symbolic behavior. The preservation biasesusually found in Lower Paleolithic sites make the period pre-dating the Misliya Cave habitation poorly known zooarchaeo-logically in Eurasia, although some cases of sizeable faunalassemblages both preserved and primarily accumulated by hu-mans do occur (e.g., Goren-Inbar et al., 1994; Carlos-Diezet al., 1999; Rivals et al., 2004, 2006; Gopher et al., 2005;Chazan and Horwitz, 2006).

The early Middle Paleolithic data from Misliya Cave joinmany later MP/MSA faunal studies indicating that middleand late Pleistocene foragers practiced ‘‘modern’’ and ecolog-ically well-adapted hunting strategies, transport decisions, andbutchery virtually indistinguishable from those of later groupsin human prehistory. Thus, the Misliya Cave archaeofaunalstudy deepens our understanding of Paleolithic lifeways andindicates that early Middle Paleolithic humans in Mount

Carmel displayed modern subsistence behavior. Future re-search should fine-tune the zooarchaeological and taphonomicdata in order to decipher the archaeofaunal signature of the de-velopment of more specific hunting and subsistence strategies(e.g., long-range hunting tactics or carcass sharing) againsttheir technological, social, and environmental background,thereby improving our understanding of Paleolithic humansocieties.

Acknowledgements

We thank Naomi Cleghorn, Matt Hill, Erella Hovers, Dan-iel Kaufman, and Nimrod Marom for their important com-ments on previous versions of this manuscript. DickBruggeman provided editorial assistance. We also thankRivka Rabinovich for her help using the comparative collec-tion of the Department of Evolution, Systematics and Ecol-ogy at the Hebrew University of Jerusalem. The researchwas generously supported by the Dan David Foundationthrough the ‘‘Dan David Expedition: Searching for the Ori-gins of Modern Homo sapiens,’’ the L.S.B. Leaky Founda-tion, The Irene Levi Sala CARE ArchaeologicalFoundation, and the Faculty of Humanities at the Universityof Haifa. The research was funded in part by the Israel Sci-ence Foundation (grant 147/04).

Appendix A. NISP and MNE values for each bone portion in the three major size classes in the Misliya Cave assemblage

DmþDm size GgþGg size BpþBp size

NISP MNE NISP MNE NISP MNE

Horn/antler 1 1 2 1 0 0

Skull 31 3 30 6 11 2

Frontal 3 2

Zygomatic 1 1

Temporal bulla 2 2

Petrosum 8 8

Occipital 1 1

Maxilla-bone 2 1 1 1

Maxilla-teeth (total

maxilla)

23 18 (6) 18 16 (10) 11 9 (4)

Mandible 81 13 67 12 17 6

Mandible-teeth [total

mandible]

57 [13] 57 [12] 17 13 [6]

DN1 4 3

DN2 6 3 3 3

DN3 2 1 2 1

DN4 5 3 3 2

DN5 3 1 3 1

DN6 1 1

DN7 2 2 1 1

DN8 8 5 2 2

Cervical vertebrae 8 1 7 3 1 1

AX1 3 1

CE1 5 1 3 2

CE2 3 1 1 1 1 1

Thoracic vertebrae 7 2 10 3 0 0

TH1 6 2 4 3

TH2 1 1 3 3

TH OTHER 3 2

(continued on next page)

672 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Appendix A (continued )

DmþDm size GgþGg size BpþBp size

NISP MNE NISP MNE NISP MNE

Lumbar vertebrae 22 7 14 4 1 1

LU1 16 7 7 4 1

LU2 2 1 5 3

LU OTHER 4 1 2 1

Ribs 45 11 38 10 0 0

RI1 4 4 6 6

RI2 4 4 5 5

RI3 22 11 20 10

RI4 10 4 8 5

RI5 7 6 2 1

Scapula 6 1 4 2 1 1

SP1 2 2

SP2 1 1 1 1

SP4 2 1 1 1

SP shoulder blade 3 1

SP axial blade 1 1

Humerus 43 10 43 11 1 1

HU1 1 1 1 1

HU2 2 1 6 4

HU3 16 7 12 7 1 1

HU4 19 10 13 11

HU5 6 2 13 5

Radius and ulna 57 13 36 9 1 1

RA1 9 6 6 3

RA2 13 10 5 4 1 1

RA3 18 13 13 6 1 1

RA4 13 10 10 9

RA5 1 1

UL1 1 1

UL2 4 2 3 3

UL-shaft 2 1

Carpals 6 6 4 4 0 0

Cuneiform 1 1 4 4

Scaphoid 4 4

Lunate 1 1

Metacarpals 49 11 19 3 0 0

MC1 12 6 4 3

MC2 2 1 3 3

MC3e4 36 11 12 3

MC5 1 1

Pelvis and sacrum 17 2 15 4 1 1

IL2 6 3 4 3

IS1 4 2 4 3

IS2 3 3

IS3 1 1

PU1 2 2 6 6

SC1 1 1 1 1

SC2 1 1

Femur 74 12 55 8 1 1

FE1 1 1 4 3

FE2 2 1

FE3 22 10 9 4

FE4 36 12 21 8 1 1

FE5 7 5 5 2

FE6 11 4 14 4

Tibia 81 8 59 10 8 4

TI1 2 2 2 2

TI2 25 7 16 10 1 1

TI3 42 8 26 6 4 4

TI4 11 6 10 6 2 2

TI5 4 4 5 4

673R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Appendix A (continued )

DmþDm size GgþGg size BpþBp size

NISP MNE NISP MNE NISP MNE

Tarsals 17 14 11 11 1 1

AS1-2 7 4 6 6 1 1

CA1 1 1

CA3 1 1 1 1

CA4 1 1

Ex-mid cuneiform 6 5 2 2

Distal fibula 2 2 2 2

Metatarsals 134 23 35 8 3 2

MR1 25 8 17 8 2 2

MR2 9 4 3 1

MR3e4 102 23 15 4 1 1

MR5 1 1

Indeterminate

metapodials

73 14 39 7 3 1

MP2 1 1

MP3e4 32 14 16 7

MP5 2 1

MP6 34 10 22 7 3 1

Phalanges 1/2/3/S 48/27/13/52 17/10/6/52 41/19/13/19 15/12/11/19 5/1/0/1 3/1/0/1

Sesamoid 52 52 19 19 1 1

P11 22 14 21 15 3 3

P12 7 4 5 3

P13 25 17 15 11 2 2

P21 4 4 8 8

P22 22 10 12 12 1 1

P23 2 2 3 3

P31 13 6 13 11

Total 897 223 575 167 57 27

%NISP 57% 36% 4%

MNI 12 6 3

Notes: Portion codes follow Lam et al.’s (1999: Figure 1) scan sites. Lines in bold specify the total NISP and MNE values for a specific bone (e.g., tibia) or group of

bones (e.g., cervical vertebrae). The values of the different portions do not necessarily add up to the figures in bold because many specimens have more than one

scan site.

References

Abe, Y., 2005. Hunting and butchery patterns of the Evenki in northern

Transbaikalia, Russia. Ph.D. Dissertation, Stony Brook University, Stony

Brook.

Adler, D.S., Bar-Oz, G., Belfer-Cohen, A., Bar-Yosef, O., 2006. Ahead of the

game: Middle and Upper Palaeolithic hunting behaviors in the southern

Caucasus. Curr. Anthropol. 47, 89e118.

Akazawa, T., Aoki, K., Bar-Yosef, O. (Eds.), 1998. Neandertals and Modern

Humans in Western Asia. Plenum Press, New York.

Assefa, Z., 2006. Faunal remains from Porc-Epic: paleoecological and

zooarchaeological investigation from a Middle Stone Age site in southeast-

ern Ethiopia. J. Hum. Evol. 51, 50e75.

Barba, R., Dominguez-Rodrigo, M., 2005. The taphonomic relevance of the

analysis of bovid long limb bone shaft features and their application to el-

ement identification: study of bone thickness and morphology of the med-

ullary cavity. J. Taphonomy 3, 29e42.

Bar-Matthews, M., Ayalon, A., 2001. Eastern Mediterranean paleoclimate dur-

ing the last 250,000 years as derived from the petrography, mineralogy,

trace element and isotopic composition of cave deposits (speleothems), Is-

rael. Report GSI / 41 /01, Jerusalem.

Bar-Matthews, M., Ayalon, A., Gilmour, M., Matthews, A., Hakesworth, C.J.,

2003. Sea-land oxygen isotopic relationships from planktonic foraminifera

and speleothems in the eastern Mediterranean region and their implications

for palaeorainfall during interglacial intervals. Geochim. Cosmochim. Acta

67, 3181e3199.

Bar-Oz, G., 2004. Epipaleolithic Subsistence Strategies in the Levant. A

Zooarchaeological Perspective. ASPR Monograph Series. Brill, Boston.

Bar-Oz, G., Adler, D.S., 2005. Taphonomic history of the Middle and Upper

Palaeolithic faunal assemblages from Ortvale Klde, Georgian Republic.

J. Taphonomy 3, 185e211.

Bar-Oz, G., Dayan, T., 2003. Testing the use of multivariate inter-site tapho-

nomic comparisons: the faunal analysis of Heftziba in its Epipalaeolithic

cultural context. J. Archaeol. Sci. 30, 885e900.

Bar-Oz, G., Dayan, T., 2007. FOCUS: on the use of the petrous bone for estimating

cranial abundance in fossil assemblages. J. Archaeol. Sci. 34, 1356e1360.

Bar-Oz, G., Munro, N.D., 2004. Beyond cautionary tales: a multivariate taph-

onomic approach for resolving equifinality in zooarchaeological studies.

J. Taphonomy 2, 201e222.

Bar-Oz, G., Munro, N.D., 2007. Gazelle bone marrow yields and Epipalaeo-

lithic carcass exploitation strategies in the southern Levant. J. Archaeol.

Sci. 34, 946e956.

Bar-Oz, G., Weinstein-Evron, M., Livne, P., Zaidner, Y., 2004. Fragments of

information: preliminary taphonomic results from the Middle Palaeolithic

breccia layers of Misliya Cave, Mount Carmel, Israel. In: O’Connor, T.

(Ed.), Biosphere to Lithosphere: New Studies in Vertebrate Taphonomy.

Oxbow Press, Oxford, pp. 128e136.

Bartram, L.E., Marean, C.W., 1999. Explaining the ‘‘Klasies pattern’’: Kua eth-

noarchaeology, the Die Kelders Middle Stone Age archaeofaunas, long

bone fragmentation and carnivore ravaging. J. Archaeol. Sci. 26, 9e20.

Bar-Yosef, O., 1995. The origins of modern humans. In: Levy, T.E. (Ed.), The

Archaeology of Society in the Holy Land. Leicester University Press, Lon-

don, pp. 110e123.

Bar-Yosef, O., 1998. The chronology of the Middle Palaeolithic of the Levant.

In: Akazawa, T., Aoki, K., Bar-Yosef, O. (Eds.), Neandertals and Modern

Humans in Western Asia. Plenum Press, New York, pp. 39e56.

674 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Bar-Yosef, O., 2004. Eat what is there: hunting and gathering in the world of

Neanderthals and their neighbors. Int. J. Osteoarchaeol. 14, 333e342.

Bar-Yosef, O., Vandermeersch, B., Arensburg, B., Belfer-Cohen, A.,

Goldberg, P., Laville, H., Meignen, L., Rak, Y., Speth, J.D., Tchernov, E.,

Tillier, A.-M., Weiner, S., 1992. The excavations in Kebara Cave, Mount

Carmel. Curr. Anthropol. 33, 497e550.

Behrensmeyer, A.K., 1978. Taphonomic and ecological information from bone

weathering. Paleobiology 4, 150e162.

Behrensmeyer, A.K., 1991. Terrestrial vertebrate accumulations. In:

Allison, P.A., Briggs, D.E.G. (Eds.), Taphonomy: Releasing the Data

Locked in the Fossil Record. Plenum Press, New York, pp. 291e335.

Behrensmeyer, A.K., Gordon, K.D., Yanagi, G.T., 1986. Trampling as a cause

of bone surface damage and pseudo-cutmarks. Nature 319, 768e771.

Binford, L.R., 1981. Bones: Ancient Men and Modern Myths. Academic

Press, New York.

Binford, L.R., 1984. Faunal Remains from Klasies River Mouth. Academic

Press, New York.

Binford, L.R., 1985. Human ancestors: changing views of their behavior.

J. Anthropol. Archaeol. 4, 292e327.

Binford, L.R., 1988. Etude taphonomique des restes fauniques de la grotte

Vaufrey. Memoires de la Societe Prehistorique Francaise 19, 535e564.

Blumenschine, R.G., 1986. Early Hominid Scavenging Opportunities: Impli-

cation of Carcass Availability in the Serengeti and Ngorongoro Ecosys-

tems. British Archaeological Reports, Oxford.

Blumenschine, R.G., 1988. An experimental model of the timing of hominid

and carnivore influence on archaeological bone assemblages. J. Archaeol.

Sci. 15, 483e502.

Blumenschine, R.G., 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. J. Hum. Evol. 29,

21e51.

Blumenschine, R.G., Marean, C.W., Capaldo, S.D., 1996. Blind tests of inter-

analyst correspondence and accuracy in the identification of cut marks,

percussion marks, and carnivore tooth marks on bone surfaces. J. Archaeol.

Sci. 23, 493e507.

Blumenschine, R.G., Selvaggio, M.M., 1988. Percussion marks on bone sur-

faces as a new diagnostic of hominid behavior. Nature 333, 763e765.

Bower, J.R., 2005. On ‘‘modern behavior’’ and the evolution of human intel-

ligence. Curr. Anthropol. 46, 121e122.

Brain, C.K., 1981. The Hunters or the Hunted? An Introduction to African

Cave Taphonomy. University of Chicago Press, Chicago.

Brown, W.A.B., Chapman, N.G., 1990. The dentition of fallow deer (Damadama): a scoring scheme to assess age from wear of the permanent molari-

form teeth. J. Zool. Lond. 221, 659e682.

Bunn, H.T., 1981. Archaeological evidence for meat-eating by Plio-Pleisto-

cene hominids from Koobi Fora and Olduvai Gorge. Nature 291,

574e577.

Bunn, H.T., 1983. Evidence on the diet and subsistence patterns of

Plio-Pleistocene hominids at Koobi Fora, Kenya, and Olduvai Gorge,

Tanzania. In: Clutton-Brock, J., Grigson, C. (Eds.), Animals and Archae-

ology 1: Hunters and Their Prey. British Archaeological Reports, Oxford,

pp. 21e30.

Bunn, H.T., 1993. Bone assemblage at base camps: a further consideration of

carcass transport and bone destruction by the Hadza. In: Hudson, J. (Ed.),

From Bones to Behavior: Ethnoarchaeological and Experimental Contribu-

tion to the Interpretation of Faunal Remains. Centre for Archaeological In-

vestigation Occasional Papers no. 21, Illinois, pp. 156e168.

Bunn, H.T., Bartram, L.E., Croll, E.M., 1988. Variability in bone assemblage

formation from Hadza hunting, scavenging and carcass processing.

J. Anthropol. Archaeol. 7, 412e457.

Cain, C.R., 2005. Using burned animal bone to look at Middle Stone Age oc-

cupation and behavior. J. Archaeol. Sci. 32, 873e884.

Capaldo, S.D., Blumenschine, R.J., 1994. A quantitative diagnosis of notches

made by hammerstone percussion and carnivore gnawing on bovid long

bones. Am. Antiq. 59, 724e748.

Carlos-Diez, J., Fernandez-Jalvo, Y., Rossel, J., Caceres, I., 1999. Zooarch-

aeology and taphonomy of Aurora Stratum (Gran Dolina, Sierra de Ata-

puerca, Spain). J. Hum. Evol. 37, 623e652.

Chase, P.G., 1989. How different was middle palaeolithic subsistence? A

zooarchaeological perspective on the Middle to Upper Palaeolithic transi-

tion. In: Mellars, P., Stringer, C. (Eds.), The Human Revolution: Behaviou-

ral and Biological Perspectives on the Origins of Modern Humans.

Edinburgh University Press, Edinburgh, pp. 321e337.

Chase, P.G., 2003. Comment on Henshilwood and Marean. Curr. Anthropol.

44, 637e638.

Chase, P.G., Dibble, H.L., 1990. On the emergence of modern humans. Curr.

Anthropol. 31, 58e59.

Chazan, M., Horwitz, L.K., 2006. Finding the message in intricacy: the asso-

ciation of lithics and fauna on Lower Paleolithic multiple carcass sites.

J. Anthropol. Archaeol. 25, 436e447.

Cleghorn, N., 2006. A zooarchaeological perspective on the Middle to Upper

Paleolithic transition at Mezmaiskaya Cave, northwestern Caucasus, Rus-

sia. Ph.D. Dissertation, Stony Brook University, Stony Brook.

Cleghorn, N., Marean, C.W., 2004. Distinguishing selective transport and insitu attrition: a critical review of analytical approaches. J. Taphonomy 2,

43e67.

Costamagno, S., Meignen, L., Beauval, C., Vandermeersch, B., Maureille, B.,

2006. Les Pradelles (Marillac-le-Franc, France): a Mousterian reindeer

hunting camp? J. Anthropol. Archaeol. 25, 466e484.

Dayan, T., 1994. Carnivore diversity in the late Quaternary of Israel. Quatern.

Res. 41, 343e349.

Deacon, H.J., 1989. Late Pleistocene palaeoecology and archaeology in the

southern Cape, South Africa. In: Mellars, P., Stringer, C. (Eds.), The Hu-

man Revolution: Behavioural and Biological Perspectives on the Origins

of Modern Humans. Edinburgh University Press, Edinburgh, pp. 547e564.

Dominguez-Rodrigo, M., 2002. Hunting and scavenging by early humans: the

state of the debate. J. World Prehist. 16, 1e54.

Dominguez-Rodrigo, M., Barba, R., 2006. New estimates of tooth mark and

percussion mark frequencies at the FLK Zinj site: the carnivore-homi-

nid-carnivore hypothesis falsified. J. Hum. Evol. 50, 170e194.

Dominguez-Rodrigo, M., Pickering, T.R., 2003. Early hominid hunting and

scavenging: a zooarchaeological review. Evol. Anthropol. 12, 275e282.

d’Errico, F., 2003. The invisible frontier: a multiple species model for the or-

igin of behavioral modernity. Evol. Anthropol. 12, 188e202.

Farizy, C., David, F., Jaubert, J., 1994. Hommes et Bisons du Paleolithique

Moyen a Mauran (Haute-Garrone). Editions CNRS, Paris.

Fiorillo, A.R., 1989. An experimental study of trampling: implications for the

fossil record. In: Bonnichsen, R., Sorg, M.H. (Eds.), Bone Modification.

University of Maine, Orono.

Fisher, J.W., 1995. Bone surface modifications in zooarchaeology. J. Archaeol.

Meth. Theory 2, 7e68.

Gamble, C., 1999. The Palaeolithic Societies of Europe. Cambridge University

Press, Cambridge.

Gargett, R.H., 1999. Middle Palaeolithic burial is not a dead issue: the view

from Qafzeh, Saint-Cesaire, Kebara, Amud, and Dederiyeh. J. Hum.

Evol. 37, 27e90.

Garrod, D.A.E., Bate, D.M.A., 1937. The Stone Age of Mount Carmel.

Excavations at the Wadi Mughara, vol. I. Clarendon Press, Oxford.

Gaudzinsky, S., 2006. Monospecific or species-dominated faunal assemblages

during the Middle Palaeolithic in Europe. In: Hovers, E., Kuhn, S.L.

(Eds.), Transitions before the Transition: Evolution and Stability in the

Middle Palaeolithic and Middle Stone Age. Springer, New York, pp.

137e147.

Gaudzinsky, S., Roebroeks, W., 2000. Adults only: reindeer hunting at the

Middle Palaeolithic site Salzgitter Lebenstedt, northern Germany. J.

Hum. Evol. 38, 497e521.

Gopher, A., Barkai, R., Shimelmitz, R., Khalaily, M., Lemorini, C.,

Hershkovitz, I., Stiner, M.C., 2005. Qesem Cave: An Amudian site in cen-

tral Israel. J. Isr. Prehist. Soc. 35, 69e92.

Goren-Inbar, N., Lister, A., Werker, E., Chech, M., 1994. A butchered elephant

skull and associated artifacts from the Acheulian site of Gesher Benot

Ya’aqov, Israel. Paleorient 20, 99e112.

Grant, A., 1982. The use of tooth wear as a guide to the age of domestic an-

imals. In: Wilson, B., Griggson, C., Payne, S. (Eds.), Aging and Sexing

Animal Bones from Archaeological Sites. BAR International Series,

Oxford, pp. 91e108.

675R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Grayson, D.K., Delpech, F., 1994. The evidence for Middle Paleolithic scav-

enging from Couche VIII, Grotte Vaufrey (Dordogne, France). J. Archaeol.

Sci. 21, 359e375.

Grayson, D.K., Delpech, F., 2002. Specialized early Upper Palaeolithic hunters

in southwestern France? J. Archaeol. Sci. 29, 1439e1449.

Grayson, D.K., Delpech, F., 2003. Ungulates and the Middle to Upper Paleo-

lithic transition at Grotte XVI (Dordogne, France). J. Archaeol. Sci. 30,

1633e1648.

Griggo, C., 2004. Mousterian fauna from Dederiyeh Cave and comparisons

with fauna from Umm el Tlel and Douara Cave. Paleorient 30,

149e162.

Henry, D.O., Hietala, H.J., Rosen, A.M., Demidenko, Y.E., Usik, V.I.,

Armagan, T.L., 2004. Human behavioral organization in the Middle Palae-

olithic: Were Neanderthals different? Am. Anthropol. 106, 17e31.

Henshilwood, C.S., d’Errico, F., Vanhaeren, M., Van Niekerk, K., Jacobs, Z.,

2004. middle stone age shell beads from South Africa. Science 304, 404.

Henshilwood, C.S., d’Errico, F., Yates, R., Jacobs, Z., Tribolo, C.,

Duller, G.A.T., Mercier, N., Sealy, J.C., Valladas, H., Watts, I.,

Wintle, A.G., 2002. Emergence of modern human behavior: Middle Stone

Age engravings from South Africa. Science 295, 1278e1281.

Henshilwood, C.S., Marean, C.W., 2003. The origin of modern human behav-

ior. Curr. Anthropol. 44, 627e651.

Hoffecker, J.F., 1999. Neanderthals and modern humans in eastern Europe.

Evol. Anthropol. 7, 129e141.

Hoffecker, J.F., Cleghorn, N., 2000. Mousterian hunting patterns in the north-

western Caucasus and the ecology of the Neanderthals. Int. J. Osteoar-

chaeol. 10, 368e378.

Hovers, E., 2001. Territorial behavior in the Middle Palaeolithic of the south-

ern Levant. In: Conard, N.J. (Ed.), Settlement Dynamics of the Middle Pa-

laeolithic and the Middle Stone Age. Kerns Verlag, Tubingen, pp. 123e152.

Hovers, E., Belfer-Cohen, A., 2006. Now you see it, now you don’tdmodern

human behavior in the Middle Palaeolithic. In: Hovers, E., Kuhn, S.L.

(Eds.), Transitions before the Transition: Evolution and Stability in the

Middle Palaeolithic and Middle Stone Age. Springer, New York, pp.

295e304.

Hovers, E., Ilani, S., Bar-Yosef, O., Vandermeersch, B., 2003. An early case of

color symbolism: ochre use by early modern humans in Qafzeh Cave, Is-

rael. Curr. Anthropol. 44, 491e522.

Hovers, E., Kuhn, S.L. (Eds.), 2006. Transitions before the Transition: Evolu-

tion and Stability in the Middle Palaeolithic and Middle Stone Age.

Springer, New York.

Inbar, M., Ayal, E., 1980. Water resources. In: Soffer, A., Kipnis, B. (Eds.),

Atlas of Haifa and Mount Carmel. University of Haifa, Haifa, pp. 26e27.

Isaac, G.L., 1978. The food-sharing behavior of proto-human hominids. Sci.

Am. 238, 90e108.

Jelinek, A.J., 1982. The Middle Palaeolithic in the southern Levant, with com-

ments on the appearance of modern Homo sapiens. In: Ronen, A. (Ed.),

The Transition from Lower to Middle Palaeolithic and the Origin of Mod-

ern Man. BAR International Series, Oxford, pp. 57e104.

Jelinek, A.J., Farrand, W.R., Hass, G., Horowitz, A., Goldberg, P., 1973. New

excavations at the Tabun Cave, Mount Carmel, Israel: a preliminary report.

Paleorient 1, 151e183.

Kaufman, D., 1999. Archaeological Perspectives on the Origins of Modern

Humans: A View from the Levant. Bergin and Garvey, Westport.

Kaufman, D., 2002. Re-evaluating subsistence skills of Levantine Middle and

Upper Palaeolithic hunters: A comparison of the faunal assemblages. Oxf.

J. Archaeol. 21, 217e229.

Klein, R.G., 1989. Biological and behavioural perspectives on modern hu-

man origins in southern Africa. In: Mellars, P., Stringer, C. (Eds.), The

Human Revolution: Behavioural and Biological Perspectives on the Or-

igins of Modern Humans. Edinburgh University Press, Edinburgh, pp.

529e546.

Klein, R.G., 1998. Why anatomically modern humans did not disperse from

Africa 100,000 years ago. In: Akazawa, T., Aoki, K., Bar-Yosef, O.

(Eds.), Neandertals and Modern Humans in Western Asia. Plenum Press,

New York, pp. 509e521.

Klein, R.G., 1999. The Human Career. University of Chicago Press, Chicago.

Klein, R.G., 2000. Archeology and the evolution of human behavior. Evol.

Anthropol. 8, 17e36.

Klein, R.G., Avery, G., Cruz-Uribe, K., Halkett, D., Parkington, J.E.,

Steele, T., Volman, T.P., Yates, R., 2004. The Ysterfontein 1 Middle Stone

Age site, South Africa, and early human exploitation of coastal resources.

Proc. Natl. Acad. Sci. U. S. A. 101, 5708e5715.

Klein, R.G., Cruz-Uribe, K., 1984. The Analysis of Animal Bones from Ar-

chaeological Sites. University of Chicago Press, Chicago.

Klein, R.G., Cruz-Uribe, K., 2000. Middle and Later Stone Age large mammal

and tortoise remains from Die Kelders 1, Western Cape Province, South

Africa. J. Hum. Evol. 38, 169e195.

Klein, R.G., Cruz-Uribe, K., Milo, R.G., 1999. Skeletal part representation in

archaeofaunas: comments on ‘‘Explaining the Klasies pattern: Kua ethno-

archaeology, the Die Kelders Middle Stone Age archaeofauna, long bone

fragmentation and carnivore ravaging’’ by Bartram & Marean. J. Archaeol.

Sci. 26, 1225e1234.

Kuhn, S.L., Stiner, M.C., 2006. What’s a mother to do? The division of labor

among Neandertals and modern humans in Eurasia. Curr. Anthropol. 47,

953e963.

Lam, Y.M., Chen, X., Pearson, O.M., 1999. Intertaxonomic variability in pat-

terns of bone density and the differential representation of bovid, cervid

and equid elements in the archaeological record. Am. Antiq. 64, 343e362.

Lam, Y.M., Pearson, O.M., Marean, C.W., Chen, X., 2003. Bone density stud-

ies in zooarchaeology. J. Archaeol. Sci. 30, 1701e1708.

Lister, A.M., 1996. The morphological distinction between bones and teeth of

fallow deer (Dama dama) and red deer (Cervus elaphus). Int. J. Osteoar-

chaeol. 6, 119e143.

Lupo, K.D., 1995. Hadza bone assemblages and hyena attrition: an ethno-

graphic example of the influence of cooking and mode of discard on the

intensity of scavenger ravaging. J. Anthropol. Archaeol. 14, 288e314.

Lyman, R.L., 1994. Vertebrate Taphonomy. Cambridge University Press,

Cambridge.

Maddrigal, T.C., 2004. The derivation and application of white-tailed deer util-

ity indices and return rates. J. Taphonomy 2, 185e200.

Marean, C.W., 1991. Measuring the postdepositional destruction of bone in ar-

chaeological assemblages. J. Archaeol. Sci. 18, 677e694.

Marean, C.W., 1998. A critique of the evidence for scavenging by Neandertals

and early modern humans: new data from Kobeh Cave (Zagros Mountains,

Iran) and Die Kelders Cave 1 layer 10 (South Africa). J. Hum. Evol. 35,

111e136.

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. J. Hum. Evol. 38, 197e233.

Marean, C.W., Assefa, Z., 1999. Zooarchaeological evidence for the faunal ex-

ploitation behavior of Neandertals and early modern humans. Evol.

Anthropol. 8, 22e37.

Marean, C.W., Cleghorn, N., 2003. Large mammal skeletal element transport:

applying foraging theory in a complex taphonomic system. J. Taphonomy

1, 15e42.

Marean, C.W., Dominguez-Rodrigo, M., Pickering, T.R., 2004. Skeletal ele-

ment equifinality in zooarchaeology begins with method: the evolution

and status of the ‘‘shaft critique.’’ J. Taphonomy 2, 69e98.

Marean, C.W., Kim, S.Y., 1998. Mousterian large mammal remains from Ko-

beh Cave: behavioral implications for Neandertals and early modern hu-

mans. Curr. Anthropol. 39, S79eS113.

Marean, C.W., Spencer, L.M., 1991. Impact of carnivore ravaging on zooarch-

aeological measures of element abundance. Am. Antiq. 56, 645e658.

McBrearty, S., Brooks, A.S., 2000. The revolution that wasn’t: a new interpre-

tation of the origin of modern human behavior. J. Hum. Evol. 39, 453e

563.

Meignen, L., Bar-Yosef, O., Speth, J.D., Stiner, M.C., 2006. Middle Paleolithic

settlement patterns in the Levant. In: Hovers, E., Kuhn, S.L. (Eds.), Tran-

sitions before the Transition: Evolution and Stability in the Middle Paleo-

lithic and Middle Stone Age. Springer, New York, pp. 149e169.

Mellars, P., 2004. Reindeer specialization in the early Upper Palaeolithic: the

evidence from south west France. J. Archaeol. Sci. 31, 613e617.

Mellars, P., 2005. The impossible coincidence: a single-species model for the

origins of modern human behavior in Europe. Evol. Anthropol. 14, 12e27.

676 R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Mellars, P., 2006. Why did modern human populations disperse from Africa

ca. 60,000 years ago? A new model. Proc. Natl. Acad. Sci. U. S. A.

103, 9381e9386.

Mellars, P., Stringer, C. (Eds.), 1989. The Human Revolution: Behavioural and

Biological Perspectives on the Origins of Modern Humans. Edinburgh

University Press, Edinburgh.

Mendelssohn, H., Yom-Tov, Y., 1999. Fauna Palestina: Mammalia of Israel.

The Israel Academy of Sciences and Humanities, Jerusalem.

Mercier, N., Valladas, H., 2003. Reassessment of TL age estimates of burnt

flints from the Palaeolithic site of Tabun Cave, Israel. J. Hum. Evol. 45,

401e409.

Mercier, N., Valladas, H., Frojet, L., Joron, J.-L., Ryess, J.-L., Weiner, S.,

Goldberg, P., Meignen, L., Bar-Yosef, O., Belfer-Cohen, A., Chech, M.,

Kuhn, S.L., Stiner, M.C., Tillier, A.-M., Arensburg, B., Vandermeersch, B.,

2007. Hayonim Cave: a TL-based chronology for this Levantine Mousterian

sequence. J. Archaeol. Sci. 34, 1064e1077.

Metcalfe, D., Jones, K.T., 1988. A reconsideration of animal body part utility

indices. Am. Antiq. 53, 486e504.

Milo, R.G., 1998. Evidence for hominid predation at Klasies River Mouth,

South Africa, and its implications for the behavior of early modern hu-

mans. J. Archaeol. Sci. 25, 99e133.

Monahan, C.M., 1998. The Hadza carcass transport debate revisited and its ar-

chaeological implications. J. Archaeol. Sci. 25, 405e424.

Munro, N.D., Bar-Oz, G., 2005. Gazelle bone fat processing in the Levantine

Epipalaeolithic. J. Archaeol. Sci. 32, 223e239.

Munson, P.J., Marean, C.W., 2003. Adults only? A reconsideration of Middle

Palaeolithic ‘prime-dominated’ reindeer hunting at Salzgitter Lebenstedt.

J. Hum. Evol. 44, 263e273.

Naveh, Z., 1984. The vegetation of the Carmel and Nahal Sefunim and the

evolution of the cultural landscape. In: Ronen, A. (Ed.), Sefunim Prehis-

toric Sites, Mount Carmel, Israel. BAR International Series, Oxford, pp.

23e63.

Nowak, R.M., 1999. Walker’s Mammals of the World. vol. II, sixth ed. The

Johns Hopkins University Press, Baltimore.

O’Connell, J.F., Hawkes, K., Blurton Jones, N.G., 1988. Hadza hunting, butch-

ering and bone transport and their archaeological implications. J. Anthro-

pol. Res. 44, 113e161.

O’Connell, J.F., Hawkes, K., Blurton Jones, N.G., 1990. Reanalysis of large

mammal body part transport among the Hadza. J. Archaeol. Sci. 17,

301e316.

O’Connell, J.F., Hawkes, K., Lupo, K.D., Blurton Jones, N.G., 2002. Male

strategies and Plio-Pleistocene archaeology. J. Hum. Evol. 43, 831e872.

Olami, Y., 1984. Prehistoric Carmel. Israel Exploration Society, Jerusalem.

Oliver, J.S., 1989. Analogues and site context: bone damages from Shield Trap

Cave (24CB91), Carbon County, Montana, U.S.A. In: Bonnichsen, R.,

Sorg, M.H. (Eds.), Bone Modification. University of Maine Center for

the Study of First Americans, Orono, pp. 73e98.

Pickering, T.R., Egeland, C.P., 2006. Experimental patterns of hammerstone

percussion damage on bones: implications for inferences of carcass pro-

cessing by humans. J. Archaeol. Sci. 33, 459e469.

Pickering, T.R., Marean, C.W., Dominguez-Rodrigo, M., 2003. Importance of

limb bone shaft fragments in zooarchaeology: a response to: ‘‘On in situ

attrition and vertebrate body parts profiles’’ (2002), by M.C. Stiner. J. Ar-

chaeol. Sci. 30, 1469e1482.

Rabinovich, R., 1990. Taphonomic research of the faunal assemblage from the

Quneitra site. In: Goren-Inbar, N. (Ed.), Quneitra: A Mousterian Site on

the Golan Heights. Qedem Monographs of Archaeology, Jerusalem, pp.

189e219.

Rabinovich, R., Bar-Yosef, O., Vandermeersch, B., Horwitz, L.K., 2004. Hom-

inid-carnivore interactions in the Palaeolithic site of Qafzeh Cave, Israel.

Revue de Paleobiologie, Geneve 23, 627e637.

Rabinovich, R., Horwitz, L.K., 1994. An experimental approach to the study

of porcupine damage to bones. In: Patou-Mathis, M. (Ed.), Taphonomie/

Bone Modification. Editions du CEDRAC, Treignes.

Rabinovich, R., Hovers, E., 2004. Faunal analysis from Amud Cave: prelimi-

nary results and interpretations. Int. J. Osteoarchaeol 14, 287e306.

Rabinovich, R., Tchernov, E., 1995. Chronological, paleocological and tapho-

nomical aspects of the middle palaeolithic site of Qafzeh, Israel. In:

Buitenhuis, H., Uerpmann, H.-P. (Eds.), Archaeozoology of the Near

East II. Backhuis Publishers, Leiden, pp. 5e44.

Rink, W.J., Schwartz, H.P., Weiner, S., Goldberg, P., Meignen, L., Bar-

Yosef, O., 2004. Age of the Mousterian industry at Hayonim Cave, north-

ern Israel, using electron spin resonance and 230TH/234U methods. J. Ar-

chaeol. Sci. 31, 953e964.

Rivals, F., Kacimi, S., Moutousammy, J., 2004. Artiodactyls, favorite game

of prehistoric hunters at the Caune de l’Arago cave (Tautavel, France).

Opportunistic or selective hunting strategies? Eur. J. Wildl. Res. 50,

25e32.

Rivals, F., Testu, A., Moigne, A.-M., De Lumley, H., 2006. The middle Pleisto-

cene argali (Ovis ammon antiqua) assemblages at the Caune de l’Arago (Tau-

tavel, Pyrenees-Orientales, France): Were prehistoric hunters or carnivores

responsible for their accumulation? Int. J. Osteoarchaeol. 16, 249e268.

Roebroeks, W., 2001. Hominid behaviour and the earliest occupation of

Europe: an exploration. J. Hum. Evol. 41, 437e461.

Shahack-Gross, R., Bar-Yosef, O., Weiner, S., 1997. Black-coloured bones in

Hayonim Cave, Israel: differentiating between burning and oxide staining.

J. Archaeol. Sci. 24, 439e446.

Shea, J.J., 2003. Neandertals, competition, and the origin of modern human

behavior in the Levant. Evol. Anthropol. 12, 173e187.

Shipman, P., 1981. Life History of a Fossil: An Introduction to Taphonomy

and Paleoecology. Harvard University Press, Cambridge.

Shipman, P., Rose, J., 1988. Bone tools: an experimental approach. In:

Olsen, S.L. (Ed.), Scanning Electron Microscopy in Archaeology. British

Archaeological Reports, Oxford, pp. 303e335.

Speth, J.D., 2000. Boiling vs. baking and roasting: a taphonomic approach to the

recognition of cooking techniques in small mammals. In: Rowley-Conwy, P.

(Ed.), Animal Bones, Human Societies. Oxbow, Oxford, pp. 89e105.

Speth, J.D., 2004a. News flash: negative evidence convicts Neanderthals of

gross mental incompetence. World Archaeol. 36, 519e526.

Speth, J.D., 2004b. Hunting pressure, subsistence intensification and demo-

graphic change in the Levantine late Middle Palaeolithic. In: Goren-

Inbar, N., Speth, J.D. (Eds.), Human Palaeoecology in the Levantine Corridor.

Oxbow Press, Oxford, pp. 149e166.

Speth, J.D., 2006. Housekeeping, Neandertal-style: hearth placement and

midden formation in Kebara Cave (Israel). In: Hovers, E., Kuhn, S.L.

(Eds.), Transitions before the Transition: Evolution and Stability in

the Middle Palaeolithic and Middle Stone Age. Springer, New York,

pp. 171e188.

Speth, J.D., Clark, J., 2006. Hunting and overhunting in the Levantine late

Middle Palaeolithic. Before Farming 3, 1e42.

Speth, J.D., Tchernov, E., 1998. The role of hunting and scavenging in Nean-

dertal procurement strategies: new evidence from Kebara Cave, Israel. In:

Akazawa, T., Aoki, K., Bar-Yosef, O. (Eds.), Neandertals and Modern

Humans in Western Asia. Plenum Press, New York, pp. 223e239.

Speth, J.D., Tchernov, E., 2001. Neandertal hunting and meat processing in the

Near East: evidence from Kebara Cave (Israel). In: Stanford, C.B.,

Bunn, H.T. (Eds.), Meat Eating and Human Evolution. Oxford University

Press, New York, pp. 52e72.

Speth, J.D., Tchernov, E., in press. The Middle Palaeolithic occupations at Ke-

bara Cave: a faunal perspective. In: Bar-Yosef, O., Meignen, L. (Eds.), Ke-

bara Cave, Mt. Carmel, Israel. The Middle and Upper Palaeolithic

Archaeology. Part 1. Peabody Museum of Archaeology and Ethnology,

Cambridge.

Steele, T.E., 2004. Variation in mortality profiles of red deer (Cervus elaphus)in Middle Palaeolithic assemblages from western Europe. Int. J. Osteoar-

chaeol. 14, 307e320.

Steele, T.E., 2005. Comparing methods for analyzing mortality profiles in

zooarchaeological and palaeontological samples. Int. J. Osteoarchaeol.

15, 404e420.

Stiner, M.C., 1990. The use of mortality patterns in archaeological studies of

hominid predatory adaptations. J. Anthropol. Archaeol. 9, 305e351.

Stiner, M.C., 1991. Introduction: actualistic and archaeological studies of prey

mortality. In: Stiner, M.C. (Ed.), Human Predators and Prey Mortality.

Westview Press, Boulder, pp. 1e13.

Stiner, M.C., 1993. Modern human originsdfaunal perspective. Annu. Rev.

Anthropol. 22, 55e82.

677R. Yeshurun et al. / Journal of Human Evolution 53 (2007) 656e677

Stiner, M.C., 1994. Honor among Thieves: A Zooarchaeological Study of Ne-

andertal Ecology. Princeton University Press, Princeton.

Stiner, M.C., 2001. Thirty years on the ‘‘broad spectrum revolution’’ and Pa-

leolithic demography. Proc. Natl. Acad. Sci. U. S. A. 98, 6993e6996.

Stiner, M.C., 2002a. Carnivory, coevolution, and the geographic spread of the

genus Homo. J. Archaeol. Res. 10, 1e63.

Stiner, M.C., 2002b. On in situ attrition and vertebrate body part profiles. J.

Archaeol. Sci. 29, 979e991.

Stiner, M.C., 2004. A comparison of photon densitometry and computed

tomography parameters of bone density in ungulate body part profiles. J.

Taphonomy 2, 117e146.

Stiner, M.C., 2005. The Faunas of Hayonim Cave (Israel): A 200,000-Year

Record of Paleolithic Diet, Demography and Society. Peabody Museum

of Archaeology and Ethnology, Cambridge.

Stiner, M.C., Kuhn, S.L., Surovell, T.A., Goldberg, P., Meignen, L., Weiner, S.,

Bar-Yosef, O., 2001. Bone preservation in Hayonim Cave, Israel: a macro-

scopic and mineralogical study. J. Archaeol. Sci. 28, 643e659.

Stiner, M.C., Kuhn, S.L., Weiner, S., Bar-Yosef, O., 1995. Differential burning,

recrystalization, and fragmentation of archaeological bone. J. Archaeol.

Sci. 22, 223e237.

Stiner, M.C., Munro, N.D., 2002. Approaches to prehistoric diet breadth, de-

mography and prey ranking systems in space and time. J. Archaeol.

Meth. Theory 9, 181e214.

Stiner, M.C., Munro, N., Surovell, T.A., 2000. The tortoise and the hare: small-

game use, the broad spectrum revolution and Palaeolithic demography.

Curr. Anthropol. 41, 39e73.

Stiner, M.C., Munro, N., Surovell, T.A., Tchernov, E., Bar-Yosef, O., 1999. Pa-

laeolithic growth pulses evidenced by small animal exploitation. Science

283, 190e194.

Stringer, C., 2002. Modern human origins: problems and prospects. Phil.

Trans. R. Soc. 357, 563e579.

Survey of Palestine, 1942. ’Atlit, 1:20,000 Series, Sheet 14e23.

Tchernov, E., 1976. Some late Quaternary faunal remains from the Avdat/

Aqev area. In: Marks, A.E. (Ed.), Prehistory and Paleonvironments in

the central Negev, Israel. The Avdat/Aqev Area, Part 1, vol. I. Southern

Methodist University Press, Dallas, pp. 57e68.

Tchernov, E., 1998. The faunal sequences of the southwest Asian Middle

Palaeolithic in relation to hominid dispersal events. In: Akazawa, T.,

Aoki, K., Bar-Yosef, O. (Eds.), Neandertals and Modern Humans in West-

ern Asia. Plenum Press, New York, pp. 77e90.

Valensi, P., Psathi, E., 2004. Faunal exploitation during the Middle Palaeolithic

in south-eastern France and north-western Italy. Int. J. Osteoarchaeol. 14,

256e272.

Villa, P., Mahieu, E., 1991. Breakage patterns of human long bones. J. Hum.

Evol. 21, 27e48.

Wadley, L., 2006. The use of space in the late Middle Stone Age of Rose Cot-

tage Cave, South Africa. Was there a shift to modern behavior? In: Hov-

ers, E., Kuhn, S.L. (Eds.), Transitions before the Transition: Evolution

and Stability in the Middle Palaeolithic and Middle Stone Age. Springer,

New York, pp. 279e294.

Weinstein-Evron, M., 1998. Early Natufian el-Wad Revisited. ERAUL 77,

Liege.

Weinstein-Evron, M., Kaufman, D., 1998. Nahal Mitla Cliff survey. Hadashot

Arkheologiyot, Excavations and Surveys in Israel 108, 176.

Weinstein-Evron, M., Bar-Oz, G., Zaidner, Y., Tsatskin, A., Druck, D., Porat, N.,

Hershkovitz, I., 2003a. Introducing Misliya Cave: a new continuous Lower/

Middle Palaeolithic sequence in the Levant. Eurasian Prehistory 1, 31e55.

Weinstein-Evron, M., Beck, A., Ezersky, M., 2003b. Geophysical investiga-

tions in the service of Mount Carmel (Israel) prehistoric research. J. Ar-

chaeol. Sci. 30, 1331e1341.

White, T.D., 1992. Prehistoric Cannibalism at Mancos 5MTUMR-2346.

Princeton University Press, Princeton.

Yellen, J.E., 1991. Small mammals: !Kung San utilization and the production

of faunal assemblage. J. Anthropol. Archaeol. 10, 1e26.

Zaidner, Y., Druck, D., Weinstein-Evron, M., 2006. Acheulo-Yarudian han-

daxes from Misliya Cave, Mount Carmel, Israel. In: Goren-Inbar, N.,

Sharon, G. (Eds.), Axe Age: Acheulian Tool-Making from Quarry to Dis-

card. Equinox Publishing, London, pp. 243e266.