The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort,...

41
The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African Middle Stone Age Author(s): Christopher Stuart Henshilwood and Benoît Dubreuil Reviewed work(s): Source: Current Anthropology, Vol. 52, No. 3 (June 2011), pp. 361-400 Published by: The University of Chicago Press on behalf of Wenner-Gren Foundation for Anthropological Research Stable URL: http://www.jstor.org/stable/10.1086/660022 . Accessed: 14/12/2012 19:12 Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at . http://www.jstor.org/page/info/about/policies/terms.jsp . JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range of content in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new forms of scholarship. For more information about JSTOR, please contact [email protected]. . The University of Chicago Press and Wenner-Gren Foundation for Anthropological Research are collaborating with JSTOR to digitize, preserve and extend access to Current Anthropology. http://www.jstor.org This content downloaded on Fri, 14 Dec 2012 19:12:06 PM All use subject to JSTOR Terms and Conditions

Transcript of The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort,...

Page 1: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution ofthe Mind during the African Middle Stone AgeAuthor(s): Christopher Stuart Henshilwood and Benoît DubreuilReviewed work(s):Source: Current Anthropology, Vol. 52, No. 3 (June 2011), pp. 361-400Published by: The University of Chicago Press on behalf of Wenner-Gren Foundation for AnthropologicalResearchStable URL: http://www.jstor.org/stable/10.1086/660022 .

Accessed: 14/12/2012 19:12

Your use of the JSTOR archive indicates your acceptance of the Terms & Conditions of Use, available at .http://www.jstor.org/page/info/about/policies/terms.jsp

.JSTOR is a not-for-profit service that helps scholars, researchers, and students discover, use, and build upon a wide range ofcontent in a trusted digital archive. We use information technology and tools to increase productivity and facilitate new formsof scholarship. For more information about JSTOR, please contact [email protected].

.

The University of Chicago Press and Wenner-Gren Foundation for Anthropological Research are collaboratingwith JSTOR to digitize, preserve and extend access to Current Anthropology.

http://www.jstor.org

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 2: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Current Anthropology Volume 52, Number 3, June 2011 361

� 2011 by The Wenner-Gren Foundation for Anthropological Research. All rights reserved. 0011-3204/2011/5203-0003$10.00. DOI: 10.1086/660022

The Still Bay and Howiesons Poort, 77–59 kaSymbolic Material Culture and the Evolution of the Mind

during the African Middle Stone Age

by Christopher Stuart Henshilwood and Benoıt Dubreuil

Variations in the material culture in Africa in the Late Pleistocene indicate that it was a period ofrapid cultural change not previously observed in the Middle Stone Age. In southern Africa, twotechno-traditions, the Still Bay and the Howiesons Poort, have raised interest because of their relativelyearly cultural complexity. What might have driven the development of the innovative practices andideas between ca. 77,000 and 59,000 years ago? Explanations for the ascent and demise of these twoentities must focus on analyses of recovered materials and in situ features such as hearths and spatialpatterning. They must also consider whether these innovations are likely to have ensued from cognitiveevolution in Homo sapiens and trace the changes in brain organization and cognitive functions thatbest explain them. This article presents an updated review of the Still Bay and Howiesons Poortindustries and argues that innovations during the Late Pleistocene must be related to a previousexpansion of the higher association areas of the temporal and parietal cortices underlying highertheory of mind, perspective taking, and attentional flexibility.

An overview of the material culture of Homo sapiens in south-ern Africa indicates that the Late Pleistocene was a periodduring which innovations were relatively rapidly introduced,survived for a period, and then were lost or replaced by othertechnologies (Henshilwood 2007, 2008b; Henshilwood andMarean 2003, 2006; Henshilwood et al. 2002, 2004; Lombard2005a, 2005b, 2007a; McBrearty and Brooks 2000; Wadley2007; Wadley, Hodgkiss, and Grant 2009; Wurz 2000). Twostratigraphic culture entities, or “techno-traditions” (Wurz etal. 2003), that date to this time period, the Still Bay (SB) andthe Howiesons Poort (HP), have raised particular interest(Henshilwood 2009b), first, because of their early culturalcomplexity and second, because their key lithic markers, thebifacial foliate point and backed segments, respectively, arerestricted only to these techno-traditions during the MiddleStone Age (MSA) in southern Africa. New and accurate datingtechniques have provided reliable ages of 77–72 ka for the SBand 65–59 ka for the HP (Henshilwood et al. 2002; Jacobs,

Christopher Stuart Henshilwood is Professor of African Prehistoryat the Institute for Archaeology, History, Culture and Religion of theUniversity of Bergen (Øysteinsgate 3, 5007 Bergen, Norway) andResearch Professor at the Institute for Human Evolution, Universityof the Witwatersrand, Johannesburg, South Africa ([email protected]). Benoıt Dubreuil is a postdoctoral researchfellow in the Departement de philosophie of the Universite duQuebec a Montreal (Case postale 8888, succursale Centre-ville,Montreal, Quebec H3C 3P8, Canada). This paper was submitted 22IX 09 and accepted 29 IV 10.

Duller, and Wintle 2003; Jacobs and Roberts 2008; Jacobs,Wintle, and Duller 2003; Jacobs et al. 2006, 2008; Tribolo2003; Tribolo, Mercier, and Valladas 2005; Tribolo et al. 2006).Significant finds in SB levels include foliate lithic points, en-graved ochres and bone, marine shell beads, formal bonetools, and a structured use of social space; in HP levels, backedgeometric lithics, advanced projectile systems, engraved os-trich eggshell, bone tools, and evidence of individual domestichearths have been recovered. These innovative technologiesand social practices associated with the SB and the HP (e.g.,d’Errico et al. 2005; Henshilwood et al. 2002, 2004; Lombard2005a, 2005b, 2006a, 2006b; Minichillo 2005; Texier et al.2010; Wadley 2007; Wurz 2000) are likely one part of a trans-formational behavioral mosaic that was also evolving in otherparts of Africa at the same time (Barham 2001; Bouzouggaret al. 2007; d’Errico and Vanhaeren 2009; Henshilwood andMarean 2003; McBrearty and Brooks 2000).

Most explanations for the SB and the HP focus on inter-pretations of the technology and the social meanings of re-covered artifacts (e.g., Henshilwood 2005; Henshilwood et al.2002, 2004; Lombard 2005a, 2005b, 2006a, 2006b, 2007a;Wadley 2007), on in situ features such as hearths and spatialpatterning (e.g., Deacon 1989; Minichillo 2005; Wadley andJacobs 2004, 2006; Wurz 2000), and on the effects of climaticvariability (Henshilwood 2008b; Jacobs and Roberts 2008; Ja-cobs et al. 2008). The SB and HP finds clearly demonstratethat innovative material culture appears, disappears, andreappears in the archaeological record in different forms. This

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 3: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

362 Current Anthropology Volume 52, Number 3, June 2011

pattern suggests that there were major discontinuities in thecultural transmission of innovations and stylistic material cul-ture during the 77–59-ka period in southern Africa and pos-sibly in earlier periods of the MSA as well (Henshilwood,d’Errico, and Watts 2009).

Over the past decade, the SB and HP finds have played acentral role in the debate on the evolution of the humanmind. On the one hand, several authors have claimed thatmaterial culture in the southern African MSA, especially thatof the SB and HP periods, was indicative of cognitive andlinguistic abilities equivalent to those of living H. sapiens(d’Errico et al. 2005; Henshilwood, d’Errico, and Watts 2009;Henshilwood and Dubreuil 2009; Henshilwood et al. 2004;Mellars 2005, 2006; Watts 2009). On the other hand, impor-tant criticisms have recently been raised against the use ofgeneral concepts such as “fully modern cognition” or “mod-ern syntactical language” to account for southern African in-novations during these periods and, in particular, for thepresence of shell beads and abstract engravings to supportthese contentions (Botha 2008, 2009; Coolidge and Wynn2009; Wynn and Coolidge 2007). These criticisms have beenaccompanied by legitimate demands to clarify the alleged linkbetween human cognitive evolution in the MSA and behav-ioral innovations in southern Africa.

Connecting the evolution of behavior with the evolutionof the mind, however, presents its own challenges. Cognitivearchaeologists must learn, and are increasingly learning, tointegrate evidence from various disciplines (primatology, neu-roanatomy, cognitive neurosciences, etc.) and see how it bearson various hypotheses. In recent years, three hypotheses aboutthe evolution of the human mind have become central to thedebate. The first focuses on recursive syntax, the second ontheory of mind, and the third on executive functions. Someauthors have defended one or another version of these threehypotheses in close connection with the archaeological record,while others have rather focused on evidence from compar-ative psychology and neurosciences.

This increasing focus on a restricted set of hypotheses isgood news for all those studying the evolution of the humanmind. It is an indication that interdisciplinary work over thepast two decades has led researchers to identify with increasingaccuracy the space of plausible hypotheses. At the same time,we are still far from having a consensual description of therelationship between the evolution of the mind and the ar-chaeological record.

Our view is that the three hypotheses just mentioned—recursive syntax, theory of mind, and working memory—alloffer imperfect accounts of recent behavioral evolution. Inthe next sections, we explain why and argue that the time hascome for students of human cognitive evolution to go beyondvery general cognitive abilities and to propose increasinglyprecise cognitive mechanisms realized in particular neuralstructures, processes, or organizational patterns. This discus-sion sets the stage for our own proposal, in which behavioralinnovations in the SB and HP periods, and most centrally

the evolution of a stylistic and symbolic component in ma-terial culture, are explained by a previous expansion and re-organization of the higher association areas of the temporaland parietal cortices underlying higher theory of mind, per-spective taking, and working memory.

The Evolution of the Mind:Three Hypotheses

The Evolution of Syntax

Linguistic communication is such a central aspect of humanlife that it seems impossible to explain the evolution of themind without referring to it. The evolution of language is theissue that has received by far the most attention in the broaderdebate on the evolution of human cognition. Within languageevolution, the question of recursion—understood as the ca-pacity to embed a constituent in a constituent of the sametype—is the one that has attracted the most interest. Recur-sion (or its cousin, the function “merge”) has been claimedto be unique to human language (Corballis 2007; Hauser,Chomsky, and Fitch 2002; Suddendorf and Corballis 2007),and its evolution has been linked to the evolution of properlymodern culture in H. sapiens (Bickerton 1990:188, 2000, 2003;Read 2008b:703). The hypothesis, however, faces numerouschallenges. In our view, these challenges are sufficiently seriousto call into question the centrality of recursion in the expla-nation of recent cognitive evolution in the human lineage.

First, the fact that recursion is proper to human languageis not definitely established. On the one hand, it cannot beexcluded that other species are capable of computing recursiveconstructions. The evidence from European starlings providedby Gentner et al. (2006) has been convincingly contested byCorballis (2007), but some doubts remain about the abilityof apes to parse recursive constructions. Research with thebonobo Kanzi has shown that he could grasp important fea-tures of syntax (Savage-Rumbaugh, Shanker, and Taylor1998), although his ability to parse properly recursive con-structions remains open to debate (Corballis 2007:702). Onthe other hand, humans’ ability to compute complex sen-tences is also debatable. It is not clear that the complex syn-tactic trees described in Chomskian formal analysis representproperly what is going on in our brain when we parse complexsentences. Marcus (2009) has argued on the basis of psycho-linguistic data that we instead use partial trees (treelets) todo so.

Second, even if computing linguistic recursion is beyondthe reach of other species and within the reach of living hu-mans, it does not necessarily account for the creativity ofhuman culture. The idea that recursion is a universal featureof human language has been contested (Everett 2005; but seecriticism from Nevins, Pesetsky, and Rodrigues 2009 and theresponse from Everett 2009). Evans and Levinson (2009:442–

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 4: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 363

444) have argued that, in some languages, complex relation-ships between propositions are worked out pragmatically,which suggests that theory of mind may be more importantthan recursive syntax. Moreover, it is not clear that recursivesyntax accounts straightforwardly for the behavioral inno-vations that are most closely associated with the rise of H.sapiens or, in the first place, for the emergence of symbolicallymediated culture. In child development, for instance, the un-derstanding of symbols does not closely parallel the devel-opment of syntax. Children gain some understanding of sym-bols before they can use recursive constructions, but theygrasp other aspects of symbolic culture only much later (Du-breuil 2008 and below).

Third, there are still doubts about how the ability to parsecomplex sentences is realized in the brain and how it relatesto other cognitive abilities. Neuroimaging studies suggest thatparsing complex sentences activates areas of the inferior pre-frontal cortex, including Broca’s area, that are involved in thehierarchical organization and sequencing of both linguisticand nonlinguistic behavior (Koechlin and Jubault 2006; Stoutand Chaminade 2009). In the literature on the evolution ofthe mind, however, recursion has also been linked to othercognitive abilities, often complicating the construction of clearalternative scenarios for the evolution of the modern mind.Lieberman (2005, 2007), for instance, suggests explaining theevolution of recursion by a change in subcortical structurescontrolling motor control. He proposes to relate this changeto the evolution of the FOXP2 gene in modern humans. Lie-berman’s argument is unsatisfying for two reasons. The firstis that the modern version of the FOXP2 gene apparentlyevolved much earlier than modern humans and is also foundin Neanderthals (Krause et al. 2007). The second is thatFOXP2’s impact on behavior is multifaceted and its link withrecursion is, at best, indirect (Fitch, Hauser, and Chomsky2005). It is difficult to determine what would have been thebehavioral and cognitive repertoire of a human populationthat lacked the modern version of the gene.

Other authors, including Hauser, Chomsky, and Fitch(2002) and Suddendorf and Corballis (2007:308), havepointed toward a possible link between syntax and social cog-nition or theory of mind. The parallel is interesting, becausetheory of mind in humans allows for recursive constructions,in which mental states are embedded into mental states, forexample, “I think that he wasn’t sure that she knew that hewould arrive so late.” The parallel is also interesting becausethe development of embedded clauses in children is predictiveof their later success in complex theory-of-mind tasks (deVilliers and Pyers 2002). But despite this developmental link,theory of mind and syntactic recursion are ultimately distinctcognitive abilities, realized in different neural networks.

Recursion has finally been connected with working mem-ory, which imposes limits on the length of utterances that wecan understand (Hauser, Chomsky, and Fitch 2002:1571).Coolidge and Wynn (2007, 2009:216–218), following Aboitizet al. (2006), have suggested a possible link between the evo-

lution of recursion and the expansion of human phonologicalstorage capacity (PSC), which is the capacity to hold utter-ances in attention for a short span of time. They contend thatwhile language is infinitely generative in theory, in practiceour capacity to learn and use recursive syntax is limited byour PSC. An expansion of the PSC would have paved the waynot only for the evolution of recursive speech but also forsubvocal articulatory processing (i.e., inner speech) and re-cursive thinking in general (including theory of mind; Coo-lidge and Wynn 2007:709–710).

This argument is interesting and echoes the proposal ofSavage-Rumbaugh, Shanker, and Taylor (1998:72–73) thatshort-term memory limitations could explain why the lan-guage-trained bonobo Kanzi stumbles with sentences such as“Show me the doggie and the milk.” But Coolidge and Wynndo not want to focus exclusively on the enhancement of thePSC. Indeed, if archaic humans had insufficient phonologicalworking memory to develop spoken language, they mighthave had sufficient visuospatial working memory to developlanguage in the gestural modality. So Coolidge and Wynn(2007:709–710, emphasis added) “believe that a genetic neuralmutation, sometime within the last 100,000 years, enhancedworking memory capacity and/or PSC.” Coolidge and Wynn(2009:216–222) propose a similar picture: the evolution ofmodern cognition could be related to the evolution of re-cursion (through the enhancement of the PSC), and/or theevolution of the pragmatics of speech (through the prior evo-lution of recursion), and/or the evolution of episodic memory.At this point, however, the syntactic hypothesis gets conflatedwith two others: the evolution of theory of mind and theevolution of the executive brain.

The Evolution of Theory of Mind

The second influential explanation of the evolution of themodern mind focuses on social cognition and theory of mind,specifically, on the capacity to attribute mental states to others.Robin Dunbar (1992, 1993, 2003, 2007) has most famouslyargued that the evolution of the mind was prompted not bythe need to solve technological problems but by the need tomaintain social bonds and alliances. We are sympathetic tothe idea that the selective pressure behind human cognitiveevolution was primarily social, but we are unsure about thedetails of his argument.

According to Dunbar (1992, 1993), the relative size of theneocortex in primates is correlated with group size. On thisbasis, Aiello and Dunbar (1993) argue that encephalizationin the human lineage was related to the pressure to formlarger groups and alliances. Unfortunately, there is no evi-dence that the most important phases of encephalization (inPlio-Pleistocene and mid-Pleistocene hominins) were linkedwith an expansion in the size of groups or social networks.Moreover, there are no prima facie reasons to think that groupsize in Pleistocene foragers was not first and foremost con-strained by noncognitive factors, as is generally the case in

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 5: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

364 Current Anthropology Volume 52, Number 3, June 2011

nonhuman primates (Lehmann, Korstjens, and Dunbar2007).

Our most important reservations about the social-brainhypothesis concern the cognitive mechanisms that are putforward as central to the evolution of the modern mind.Dunbar’s (2007) theory builds on philosopher Daniel Den-nett’s (1987) views on intentionality, according to whichthinking about mental states typically involves embeddingmental states into other mental states, as in “I believe thatyou know that he thinks, etc.” There is evidence that humansget rapidly puzzled when they need to decipher situationsinvolving five or more orders of intentionality (Kinderman,Dunbar, and Bentall 1998; Stiller and Dunbar 2007). Althoughthis is an interesting cognitive limitation, it cannot straight-forwardly be used to account for human cognitive evolution.First, there is no evidence that orders of intentionality arecausally connected with the relative size of the neocortex or,as suggested by Dunbar (2007:94), with the volume of thefrontal-lobe gray matter. Second, it is far from obvious thatwe need three, four, or five orders of intentionality to un-derstand symbolic, artistic, or religious behavior (Dunbar2007:95). Third and more important, it is reasonable to pre-sume that our incapacity to go beyond five orders of inten-tionality has nothing to do with social cognition, sensu stricto,but results from the limitations of our domain-general in-telligence.

Indeed, social cognition and theory of mind refer to com-plex cognitive tasks that depend on both domain-specific anddomain-general cognitive mechanisms (Apperly and Butterfill2009; Stone and Gerrans 2006). Domain-general abilities areparticularly important for the most complex social cognitiveabilities, like those that develop later in infancy and are mostlikely to have evolved recently in our lineage. Performance inthe false-belief task, most notably, is strongly correlated withthe ability to inhibit self-perspective and hold in mind op-posing viewpoints on a situation (Carlson, Moses, and Clax-ton 2004). This holds true not only for children but also foradults (Apperly, Samson, and Humphreys 2009; Bailey andHenry 2008). The importance of domain-general cognitionfor theory of mind raises the possibly that the evolution ofthe social brain in fact resulted from the evolution of theexecutive brain.

The Evolution of the Executive Brain

A third influential group of explanations suggests that the keyto the evolution of the modern mind is to be found in theexecutive functions of the brain, that is, its ability to manageand control lower-level psychological processes. Mithen(1996), for instance, links the evolution of the modern mindto that of cognitive fluidity and contends that modern humansbecame increasingly able to establish connections betweenpreviously separated cognitive modules. Deacon (1997), forhis part, explains the emergence of symbolic behavior throughthe evolution of the executive functions of the brain, which

would have made possible the construction of properly sym-bolic relations between indexical tokens. In several works,Coolidge and Wynn (2001, 2004, 2005, 2009) propose to ex-plain the evolution of the modern mind by the evolution ofworking memory, that is, the capacity to temporarily hold inattention recently processed information and to keep it avail-able for further processing (see also Read 2008b; Rossano2007).

The biggest challenge of these approaches is to move be-yond very general accounts of executive functions and linkprecise cognitive mechanisms with likely neural and behav-ioral evolutionary processes. Coolidge and Wynn (2001, 2004,2005) have made an important step in this direction by in-troducing Baddeley and Hitch’s (1974) model of workingmemory into the debate on human cognitive evolution. Intheir initial model, Baddeley and Hitch distinguished threefunctional components of working memory: (1) the centralexecutive, responsible for attention, selection, and inhibition;(2) the phonological store, responsible for vocal and subvocalrehearsal of verbal and acoustic stimuli; and (3) the visuo-spatial sketchpad, in charge of integration and temporary stor-age of visual (what) and spatial (where) information.

As we saw above, Coolidge and Wynn (2007) suggest thatthe expansion of the phonological store might be responsiblefor the evolution of recursion and, incidentally, of modernculture and cognition. In other papers (Coolidge and Wynn2001, 2004, 2005; Wynn, Coolidge, and Bright 2009), theyfocus rather on the importance of the central executive in theevolution of the modern mind.

We have raised some doubts above about the explanatorypower of recursion, whose nature, neural basis, and importancefor behavior are still widely debated. The working-memoryhypothesis raises problems of a different nature. On the onehand, a strong case can be made that working memory (or atleast its central executive component) allows for much greatercognitive and behavioral fluidity in humans than in apes (Read2008b). On the other hand, focusing on the central executivedoes not get us very far, because the concept itself is very broadand can be further subdivided into multiple functional sub-components (e.g., inhibitory control, attention flexibility, goalselection, goal maintenance, and planning; for a review, seeJurado and Rosselli 2007). Moreover, even if it is granted thatthe central executive is expanded in modern humans comparedto that in apes, we need specific arguments to link its expansionto behavioral evolution in modern H. sapiens. Our view is thatwe might be able to design more detailed scenarios for theevolution of the central executive if we turn to what we knowof the evolution of the brain itself.

The Evolution of the Prefrontal Cortex

The executive functions of the brain are in large part realizedin the prefrontal cortex (PFC). It is thus not surprising thatan important part of the discussion on the evolution of themodern mind has been focused on this region (Amati and

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 6: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 365

Shallice 2007; Ardila 2008; Cela-Conde et al. 2004; Coolidgeand Wynn 2001; Deacon 1997; Dunbar 2007; Mithen 1996;Noack 2006, 2007; Noble and Davidson 1996; Read 2008b;Rossano 2007). The comparative study of the frontal lobe inhumans and apes reveals interesting similarities and dissim-ilarities. Although the human frontal lobe underwent a dra-matic expansion in absolute size during human evolution, itdoes not represent a larger proportion of the neocortex inmodern humans than in apes (Semendeferi et al. 2002). Ithas been argued that the expansion of the frontal lobe inabsolute terms was of more benefit to the PFC—and especiallyto the frontopolar cortex (BA [Brodmann area] 10)—than tothe premotor and primary motor areas, but the point is stilldebated (Rilling 2006; Semendeferi et al. 2001; Sherwood,Subiaul, and Zawidzki 2008).

It is not easy to determine the significance of the absoluteexpansion of the frontal lobe for the evolution of humancognition and behavior. Although the human PFC might notbe exceptionally large for a primate of our size, its expansionin the human lineage was accompanied by changes that mighthave favored higher connectivity within the PFC and withother regions of the neocortex. One such change is the in-creased level of gyrification observed in the modern humanPFC (Rilling and Insel 1999). Another is the higher proportionof white matter, which is thought to have facilitated long-distance connections with other regions of the brain (Schoe-nemann, Sheehan, and Glotzer 2005). A final example is thevariation in the number and density of cortical minicolumns,which might have determined new patterns of connectivity(Casanova and Tillquist 2008). All these changes, it shouldbe noted, are likely to have coincided with encephalization,although they might also have occurred in the absence ofencephalization.

It is generally assumed that the expansion of the PFC inabsolute terms must have coincided with significant changesin cognition and behavior. On the one hand, the expansionof the orbitofrontal and medial frontal areas might have ledto an increased capacity to integrate emotionally and moti-vationally relevant information. On the other hand, the ex-pansion of the dorsolateral areas could have come with en-hanced capacity for goal selection, goal maintenance, andinhibitory control (Ardila 2008).

Although the functional reorganization of the PFC likelyexplains important behavioral changes in the human lineage,we contend that it cannot be so easily invoked to explainbehavioral evolution in southern Africa during the SB andHP periods. Indeed, the bulk of encephalization in the humanlineage occurred not only significantly earlier than the SB andHP periods but also well before the evolution of anatomicallymodern humans. Despite the limited number of specimens,it is reasonable to think that the encephalization quotient ofmid-Pleistocene hominins in Africa and Europe was broadlywithin the range of modern H. sapiens (Rightmire 2004,2008). If there was no relevant change in the size of the PFC,however, could there have been relevant changes in shape?

On the one hand, the study of the inner median-sagittal pro-file of the PFC reveals no significant change during the past500,000 years (Bookstein et al. 1999). On the other hand, acomparison with more archaic specimens reveals wider frontallobes in modern H. sapiens, but this trend is even more pro-nounced in Neanderthals (raising intriguing questions aboutits potential cognitive significance; Bruner and Holloway2010). Certainly, the study of endocasts is not likely to revealany possible reorganization of the PFC. However, becausesome of the most specific features of the modern PFC (gy-rification, proportion of white matter, number of microcol-umns) interact with size and shape, it is reasonable to thinkthat most features of the modern human PFC were alreadyin place in Homo heidelbergensis (albeit more recent changescannot be excluded).

Another line of evidence in favor of this conclusion comesfrom the archaeological record. In children, the latest-maturing parts of the PFC are the dorsolateral areas, whichare involved in inhibitory control, goal maintenance, and thecapacity to resist temptation (Gazzaniga et al. 2002; Gogtayet al. 2004; Knoch et al. 2006; Koechlin and Summerfield2007; Sanfey et al. 2003). If the latest-maturing areas are alsothe latest-evolving areas, then a modernlike capacity for in-hibition and goal maintenance would suggest a modernlikeorganization of the PFC in mid-Pleistocene hominins. Manybehaviors that appear during the mid-Pleistocene provide evi-dence of H. heidelbergensis’s unprecedented capacity to stickto long-term norms of cooperation, despite the temptationto free-ride (Dubreuil 2010b). Homo heidelbergensis providesthe first evidence of a subsistence strategy based on large-game hunting (and thus on extensive meat sharing), cooking,and the controlled use of fire. We also find in H. heidelbergensisthe first convincing evidence of a modernlike organization oflife-history phases (Robson and Wood 2008). Because pro-longed infancy in modern humans has been shown to dependon large intergenerational transfers, changes in life history arelikely to be informative about changes in the ability to co-operate (Kaplan and Gurven 2005). Indeed, costly intergen-erational transfers can be secured only if individuals are ableto stick to social norms and long-term goals in the face ofcompeting motivations, that is, if they have a strong capacityfor inhibitory control and goal maintenance. Behavioral andmorphological data for mid-Pleistocene specimens thereforesuggest that the PFC had already undergone its most signif-icant reorganization long before the evolution of modern H.sapiens and thus that its evolution might not the best can-didate to explain the behavioral innovation in southern Africaduring the SB and HP periods.

The Evolution of the Temporoparietal Areas

Although it has been the focus of most discussions on theevolution of the modern mind, the PFC is not the only brainarea dedicated to high-level cognition. Several areas of the tem-poral and parietal cortices are involved in complex cognitive

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 7: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

366 Current Anthropology Volume 52, Number 3, June 2011

tasks and might have undergone changes in recent humanevolution. Wynn, Coolidge, and Bright (2009), for instance,suggest that the advent of modern brain morphology is asso-ciated with functional changes in the parietal cortex, in additionto changes in the frontal cortex. The parietal cortex is activatedin several complex cognitive tasks, including visuomotor con-trol, task switching, attention, spatial working memory, mentalrotation, and mental imagery (Wynn, Coolidge, and Bright2009:75). Their argument is based on Emiliano Bruner’s geo-metric, morphometric analyses, according to which the mod-ernization of human cranium was accompanied by a nonal-lometric increase in parietal volume (Bruner 2003, 2004;Bruner, Manzi, and Arsuaga 2003).

This new hypothesis has the advantage of extending thediscussion to areas outside the PFC. One limitation is thatthe parietal cortex in modern humans is not unexpectedlylarge for a primate of our size (Lieberman 2008; Pearson2008). The posterior parietal areas, activated in tasks requiringselective attention, are expanded in relative terms in humans,but this expansion apparently occurred early in the humanlineage, probably in australopithecines (Holloway, Clark, andTobias 2004). For these reasons, we prefer to remain cautiousconcerning a possible expansion and functional reorganiza-tion of the parietal cortex along with the evolution of modernH. sapiens, although this possibility must be considered.

Lieberman, McBratney, and Krovitz (2002) have advancedanother viewpoint on the modernization of cranial mor-phology. They argue that the most important feature of themodern H. sapiens braincase is its overall globular shape,which they link with an expansion of the volume of the tem-poral lobes, which are both longer and wider in modern hu-mans than in archaic humans. Just as is the case for theparietal lobe, the temporal cortex is activated in a wide rangeof complex cognitive tasks, including visual and phonologicalworking memory and theory of mind, as well as semanticsand concept formation. However, there is a general consensusthat the temporal lobes, in contrast to the parietal lobe, aredisproportionately enlarged in modern humans compared tothose in apes (Rilling and Seligman 2002; Semendeferi andDamasio 2000). This enlargement is primarily connected toa relative increase in white matter, suggesting enhanced con-nectivity within the temporal lobes and with other regions ofthe brain, including the frontal and parietal lobes (Schenker,Desgouttes, and Semendeferi 2005).

The relative and absolute enlargement of the temporal cor-tex in recent human evolution might have multiple causesand effects. Although it is currently impossible to provide adetailed story of its significance, archaeological and neuro-logical data can give us some hints. On the basis of humanand nonhuman-primate comparative data, Ungerleider,Courtney, and Haxby (1998) argue that encephalization inthe human lineage came with a displacement of temporal andparietal cortical areas dedicated to visuospatial working mem-ory. Compared to that in nonhuman primates, the dorsalstream of visual working memory—the “where” stream,

which processes information about objects’ location inspace—has a more superior location in human parietal cortex.The ventral stream—the “what” stream, which processes in-formation about object recognition—has a more inferior lo-cation in the temporal cortex. The displacement of the twostreams of visuospatial working memory coincides with arelative expansion of the higher association areas located alongthe Sylvian fissure.

Comparative studies also suggest an important expansionin humans of the arcuate fasciculus, the white-matter fibertract that arches above the Sylvian fissure and connects thehigher associational cortex of the temporal and parietal lobes(including Wernicke’s area) to Broca’s area in the frontal cor-tex (Rilling et al. 2008). The arcuate fasciculus and otherwhite-matter pathways that run in the temporal lobes aregenerally presented as connecting lexical and semantic infor-mation (stored in the temporal cortex) with syntactic pro-cessing abilities (mostly realized in the inferior frontal cortex,including Broca’s area; Friederici 2009).

In sum, we suggest that it is in the higher association areasof the temporal cortex, or in the underlying white matter thatconnects it with the parietal and frontal cortex, that we shouldlook for the functional reorganization that paved the way tosymbolic behavior in the southern African archaeological rec-ord.

The Temporoparietal Junction, Theory ofMind, and the Executive Brain

The presence of a stylistic and symbolic component in ma-terial culture is probably the most striking feature of the SBand HP periods in southern Africa. Before we examine thearchaeological record in detail, we explain first how the higherassociation areas of the temporoparietal cortex are related tocomplex cognitive abilities such as theory of mind and work-ing memory and second how an expansion of these regionscould account for the emergence of styles and symbols inhuman evolution.

The functional organization of the temporoparietal areasand, in particular, of the temporoparietal junction (TPJ) hasrecently been debated. Some propose that the TPJ is the seatof general metacognitive computational processes that are in-volved in reorienting attention (Decety and Lamm 2007;Mitchell 2008). Others argue that the TPJ is functionally dif-ferentiated and that a specific region within it is involved inprocessing mental states in false-belief tasks and intricatemoral judgments (Saxe and Kanwisher 2003; Saxe and Powell2006; Scholz et al. 2009; Young et al. 2007). It is clear in bothcases, however, that the TPJ in general is the seat of high-level cognitive mechanisms that allow us to contrast differentperspectives on objects or events (Aichhorn et al. 2006; Mitch-ell 2009).

This function is different from the more primitive repre-sentation of biological motions and goal-directed actions,which are realized not in the TPJ but in a network that in-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 8: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 367

Figure 1. Main regions of the social brain: amygdala, anterior cingulatedcortex (ACC), anterior insula (AI), inferior frontal gyrus (IFG), intra-parietal sulcus (IPS), medial prefrontal cortex (mPFC), posterior superiortemporal sulcus (pSTS), and temporoparietal junction (TPJ). Reprintedwith permission from Blakemore (2008:268).

cludes the neighboring posterior superior temporal sulcus(pSTS in fig. 1), the intraparietal sulcus (IPS), and the inferiorfrontal gyrus (IFG; Blakemore 2008; Gobbini et al. 2007;Mitchell 2009; Tunik et al. 2007). It is also different from therepresentation of the affective features of the social situation,which is realized in a network that connects limbic structuressuch as the amygdala to the anterior insula (AI), the anteriorcingulated cortex (ACC), and the medial prefrontal cortex(mPFC).

There is a growing consensus in comparative psychologythat both young children (less than 4 years of age) and apesfail to reach a stable representation of false beliefs, althoughthey both understand intentions and goal-directed actions(for reviews, see Apperly and Butterfill 2009; Call and To-masello 2008). It suggests that the TPJ matures late in infancyand that it underwent a functional reorganization in the hu-man lineage. The maturation of the TPJ in infancy is alsolikely to be responsible for children’s success in other cognitiveabilities that bear some resemblance to the understanding offalse beliefs. One of them is called the “level-2 perspective-taking” ability and refers to the capacity of children of 4 and5 years of age to describe an object from another point ofview (Flavell 1992); while level-1 perspective taking involvesthe capacity to understand what others see, level-2 perspectivetaking is about imagining how things look from another per-spective. Another is the ability to distinguish appearance and

reality, when children understand that an object can actuallybe different from how it appears (Flavell, Flavell, and Green1983).

These tasks have a common feature in that they all involveone’s ability to represent conflicting perspectives on the world.The development of this ability in children arguably has animpact on other cognitive functions. This is the case of cat-egorization. Very young children can learn to use differentconcepts to refer to the same object. Before the age of 4 or5, however, they are incapable of switching from one nameto another or of moving an object across basic-subordinatecategories. They struggle, for instance, when they have to lookat the bunny simultaneously as a rabbit or to the rabbit asan animal, although they use the concepts with the rightextension (Perner et al. 2002).

The representation of perspectives is also likely to be caus-ally linked to the evolution of episodic memory. Before 4 or5 years of age, children rely on semantic memory to encodeaspects of situations but cannot remember situations as theyexperienced them (Tulving 1972, 2002). Interestingly, it is alsoaround that age that “childhood amnesia,” our inability toremember our first years of life (Freud 1966), ends. Childhoodamnesia can be explained by the fact that younger preschool-ers are unable to encode (or recall) memories from the per-spective of the self (Perner and Ruffman 1995).

Finally and most relevant to our point, the cognitive de-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 9: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

368 Current Anthropology Volume 52, Number 3, June 2011

Figure 2. Finer-grained evolutionary scenarios connecting the evolutionof the temporoparietal areas to the emergence of styles and symbols.

velopment that happens around 4 and 5 years of age comeswith important changes in the way children interact withsymbolic culture (although children do have a certain un-derstanding of symbols before that age). Very young chil-dren—and apes—are very good at learning iconic relation-ships and can recognize fairly abstract drawings orphotographs. Human children between the ages of 2 and 3years can use a scale model or a video to retrieve a hiddenobject in the real world (DeLoache 2002, 2004). Between 3and 4 years of age, they can also understand that the meaningof a picture is defined partially by the intention of the personwho draws it and not only by its appearance (Bloom andMarkson 1998). Things get more complicated when it comesto more complex tasks, especially the production of symbols.Young preschoolers do not typically produce recognizable rep-resentations in their drawings. Similarly, only around the ageof 4 or 5 do children learn to represent a number of objectswith a written number (Zhou and Wang 2004). If youngerchildren can understand what written words refer to, theyhave no clue as to the source of the meaning. Typically, theythink that moving a word referring to a picture in front ofanother picture can alter the meaning of the word (Bialystok2000).

The higher cognitive processes that are essential to theoryof mind and perspective taking are not realized only in thetemporoparietal areas. Activation in the mPFC, for instance,is strongly associated with theory-of-mind tasks (Blakemore2008). Given the morphological data on the anatomicallymodern cranium presented above, however, as well as the factthat the mPFC matures earlier than the lateral regions of thePFC, we doubt that the evolution of the mPFC best explainsmore recent behavioral innovations in Homo sapiens.

Although our focus on the temporoparietal cortex is orig-inal, it presents an interesting link with two of the hypotheses

discussed above. The temporoparietal areas are involved inboth complex theory-of-mind tasks (e.g., verbal false-belieftasks) and working-memory tasks (e.g., attentional flexibility).Our hypothesis is compatible with different finer-grained ac-counts of the impact of a change in the temporoparietal cortexon these higher cognitive abilities (fig. 2).

One possibility is that the reorganization of the TPJ affectedtwo different mechanisms, one specifically dedicated to theoryof mind and the other to more general attentional flexibility(a component of working memory; fig. 2a). A second pos-sibility is that the change in the TPJ affected a general atten-tional mechanism implicated, among others, in complextheory-of-mind tasks (fig. 2b). A third scenario is that theexpansion of the temporoparietal cortex did not primarilycause a functional reorganization of the TPJ but rather camewith improved connectivity with the regions of the prefrontalcortex implicated in theory of mind and working memory(fig. 2c). These differing scenarios are all compatible with thebasic idea that the expansion of the temporoparietal cortexis responsible for the stylistic and symbolic innovations in thesouthern African MSA material culture.

Explaining Southern African Innovations

The core of our argument is that several innovations in thesouthern African archaeological record are characterized, ata minimum, by an interest in the appearance of objects, thatis, in the way objects look from different perspectives. Inaddition, we argue that several artifacts likely provide evidenceof symbolically mediated culture that depends on higher the-ory of mind. Our definition is that a symbolically mediatedculture is one in which individuals understand that artifactsare imbued with meaning and that these meanings are con-strued and depend on collectively shared beliefs. This criterion

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 10: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 369

is crucial. It explains how human norms and conventions

differ from the ritualized behaviors found in nonhuman pri-

mates. Apes, for instance, can easily learn how the first step

of a movement sequence predicts the following steps. For this

reason, this first step can become ritualized and used by one

individual to signal an intention to another. Infant chimpan-

zees, for instance, can touch the back of their mother to signal

their intention to be carried (Tomasello 2008:22–23). The

ritualization of intention-movement gestures requires not a

theory of mind but simply an understanding that others can

read our intentions in a partial realization of a movement.

Why is symbolically mediated culture different from this?

Why is higher theory of mind needed for a personal orna-

ment, for instance, to come to be realized as a symbol of

social status (e.g., signaling that “I am a spouse” or “I am a

member of the so-and-so clan”)? Much confusion stems from

the fact that we can learn a lot about symbolic artifacts with-

out complex theory of mind. Young human children and

nonhuman primates can recognize social statuses and develop

rich sets of expectations about the behavior of different cat-

egories of people. Two- and 3-year-old children, for instance,

easily recognize the uniform of firefighters and know what

kind of behaviors are associated with them.

But children’s understanding, however impressive, does not

reflect full integration into the world of symbolic culture.

What is the difference in adults? Adults not only develop

expectations about the behavior of people with uniforms but

also understand that the rights and duties that come with

wearing a uniform are collectively agreed upon (Searle 1995).

In other words, they understand that the ascription of a status

is causally dependent on what people think of the uniform.

It is important to stress that this additional variable (the

point of view of others) not only makes symbolically mediated

culture richer but also is precisely what makes it relevant for

us to imbue objects with symbolic meaning in the first place.

Children develop expectations about the behavior of other

agents (or types of agents), but these expectations are highly

idiosyncratic and are mostly based on previous interactions

with these agents or types of agents.

What children do not perceive is that adults create symbolic

artifacts as shortcuts for collectively shared meanings. It is not

up to each of us, for instance, to decide what expectations we

should have about firefighters’ uniforms—it is a collective mat-

ter, and it becomes so only because we are able to explicitly

coordinate our viewpoints with those of others. This implies

higher theory of mind, which is realized in the higher associ-

ation areas of the temporoparietal cortex, among other areas.

In the next section, we present an updated review of the

SB and HP archaeological record and explain how the be-

havioral innovations that have been documented for these

periods reflect a concern for style and, in many cases, plausible

evidence of symbolically mediated culture.

The Southern African ArchaeologicalRecord

The Still Bay

Sites containing the typical marker of the SB, the bifacialfoliate point (laurel-leaf shaped), are relatively rare in south-ern Africa; these include nine known cave/shelter sites and atleast five open sites (fig. 3). The principal cave or shelter sitesare Blombos Cave (Henshilwood 2008a; Henshilwood et al.2001b), Dale Rose Parlour (Goodwin and Peers 1953; Schir-mer 1975), Diepkloof Rock Shelter (Rigaud et al. 2006), Hol-low Rock Shelter (Evans 1994), Paardeberg (Wurz 2000),Peers Cave (also known as Skildergatkop, Peers’s Shelter, andB/102; Anthony 1963, 1967; Jolly 1948; Malan 1955; Peers1929; Peers and Goodwin 1953), and the sites immediatelyadjacent to Peers Cave known as Tunnel Cave (Malan 1955)and Trappieskop (note that the Trappieskop sites include B/102, C/103 [known as Nero’s Cave], and G/107 but not DaleRose Parlour; see Goodwin and Peers 1953:59). Sibudu Caveis the single known SB site in KwaZulu-Natal (Wadley 2007).Open or dune sites where SB bifacials have been recoveredbut are now difficult to accurately locate include the “CapeTown dunefields” in Maitland (Dale 1870), Cape Hangklip(Gatehouse 1954), Blombosch Sands (Heese 1933, n.d.),Kleinjongensfontein (Heese 1933, n.d.), and various sites inWellington, Cape Province (Heese, n.d.).

The SB sites are beyond the dating range for the 14Cmethod, and until 2002 the age of this techno-tradition wasunknown (Henshilwood et al. 2001b, 2002). The SB levels atBlombos Cave have been dated with a number of methods,including thermoluminescence (TL), optically stimulated lu-minescence (OSL), and electron spin resonance; (Henshil-wood et al. 2002; Jacobs, Duller, and Wintle 2003; Jacobs,Wintle, and Duller 2003; Jacobs et al. 2006; Jones 2001; Tri-bolo 2003; Tribolo et al. 2006). The MSA levels at the BlombosCave site are divided into three phases: M1, M2, and M3 (fig.4). A hiatus level composed of undisturbed aeolian sand abovethe M1 phase is dated by OSL to between and69 � 5

ka (Henshilwood et al. 2002; Jacobs, Duller, and Win-70 � 5tle 2003; Jacobs, Wintle, and Duller 2003; Jacobs et al. 2006;fig. 4). An OSL age of ka was obtained for the72.7 � 3.1upper part of the M1 phase (Jacobs, Duller, and Wintle 2003;Jacobs, Wintle, and Duller 2003). The TL ages for the M1phase are and ka (Tribolo et al. 2006). The74 � 5 78 � 6upper M2 phase also contains SB bifacials, although in lesserquantities than in M1, and the OSL age for the layers thatcontain bifacial points is ka (Jacobs et al. 2006).76.8 � 3.1This latter date should be regarded as the terminus post quemfor the SB levels at the Blombos Cave site. The lower layersin the M2 phase (layers CG, CGAA, CGAB, and CGAC) donot contain bifacial points or bone tools and have an age of

ka.84.6 � 5.8Apollo 11 is the single SB site recorded in Namibia, and

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 11: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

370 Current Anthropology Volume 52, Number 3, June 2011

Figure 3. Location of key Still Bay and Howiesons Poort sites in southernAfrica. AA, Aar 1; AP, Apollo 11; BBC, Blombos Cave; BC, Border Cave;BP, Boomplaas; B1C, Bremen 1C; CH, Cave of Hearths; DRP, Dale RoseParlour; DRS, Diepkloof; HAA, Haalenberg; HAS, Ha Soloja; HR, High-lands; HRS, Hollow Rock Shelter; HP, Howiesons Poort; KKH, KleinKliphuis; KLP, Klipfonteinrand; KP, Kathu Pan; KRM, Klasies River;MLK, Melikane; MON, Montagu Cave; MOS, Moshebi’s Shelter; NBC,Nelsons Bay Cave; NT, Ntloana Tsoana; OAK, Oakleigh; PAR, Paardeberg;PC, Peers Cave; POC, Pockenbanck; PP, Pinnacle Point; RCC, Rose Cot-tage Cave; SEH, Sehonghong; SIB, Sibudu; UMH, Umhlatuzana; WK,Wonderwerk Cave (after Jacobs et al. 2008:733). Tunnel Cave and Trap-pieskop are directly adjacent to Peers Cave.

the age of the SB here, obtained by the OSL method, is givenas ka (Jacobs and Roberts 2008; Jacobs et al. 2008).70.7 � 2.1At Sibudu Shelter, KwaZulu-Natal, the SB levels are dated at

ka by the OSL method (Jacobs and Roberts 2008,70.5 � 2.0Jacobs et al. 2008). The OSL age of the SB facies at Diepkloofrock shelter in the Western Cape is given as between

and ka (Jacobs and Roberts 2008; Ja-70.9 � 1.8 73.6 � 2.1cobs et al. 2008). Ages obtained for the same SB layers atDiepkloof with the TL method are inexplicably older andrange from to ka (layers Kerry–Kate) and99 � 10 118 � 11from to ka (layers Kim–Larry; Tribolo et108 � 9 129 � 11al. 2009).

In summary, the dating evidence for the SB levels in south-

ern Africa from the OSL method indicates a range from ca.70 ka (Apollo 11, Diepkloof, Sibudu) to ka (Blom-76.8 � 3.1bos Cave; Jacobs et al. 2006). The TL results from Diepkloofplace the SB at ca. 25,000–50,000 years older.

The “lance-head” in elongated laurel-leaf form typifies theSB type-artifact, according to Heese (n.d.), and these bifacialpoints are now accepted as the fossile directeurs of this techno-tradition in southern Africa (Henshilwood et al. 2001b; Villaet al. 2009). At the Umhlatazuna site in KwaZulu-Natal, bi-facials similar to those of the SB type are reported (Kaplan1990), and it now seems that their stratigraphic provenienceis clear and that they are associated with layers with an ageof ca. 71 ka (Lombard et al. 2010). Hollow-based points from

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 12: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 371

Figure 4. Layout of interior of Blombos Cave, showing excavated squaresand ages for the Middle Stone Age phases M1, M2, and M3.

the same site are found in the post-HP facies and have anage of ca. 55–50 ka. The tear-shaped points of the PietersburgComplex (Sampson 1974) in the Free State bear similaritiesto the SB points. Age estimates for the Pietersburg Complexsuggest that artifacts were being made there at about the sametime as those in the SB (e.g., Grun and Beaumont 2001).Elsewhere in Africa, the bifacial-type points recovered aremostly distinctive, for example, the Aterian notched bifacials(Bouzouggar, Kozłowski, and Otte 2002). The relationship ofthese complexes in central, east, and north Africa to the SBtechno-tradition in southern Africa is not clear.

At most SB sites, there is a preference for making thesefrom fine-grained raw materials, especially silcrete (e.g., Evans1994; Henshilwood et al. 2001b; Minichillo 2005; Villa et al.2009). The early evidence suggested that the SB bifacial pointswere restricted to the coastal belt of the Western Cape Prov-ince (Minichillo 2005:132), but the recent recovery of SBbifacial points at Sibudu (Wadley 2007) indicate that thistechno-tradition was more widespread in southern Africa. TheSB bifacials both identify that a SB phase is present at a siteand indicate, on the basis of current evidence, that an age ofca. 77–70 ka is likely for the levels that contain this techno-tradition (Jacobs and Roberts 2008; Jacobs et al. 2006, 2008).Bifacial roughouts correspond to various stages of the re-duction process, and flakes produced by thinning and shapingthe bifacials are a constant presence in SB levels (Henshilwoodet al. 2001b; Minichillo 2005; Villa et al. 2009).

Bifacial points likely served as multifunctional tools (Mini-

chillo 2005), and a macrofracture analysis of the BlombosCave bifacials by Lombard (2007a) indicates they were usedas both spear points and knives. The SB bifacials are oneinstance of innovation where the makers clearly showed con-cern for the appearance of the finished object, and it seemspossible that these tools also served a symbolic role (Hen-shilwood et al. 2001b). Minichillo (2005:133) makes the pointthat the interpretations of the SB techno-tradition by someof the earliest archaeologists at the Cape should now be re-garded in a different light. For example, Burkitt (1928) in-terpreted the SB as marking the arrival of “modern people”on the subcontinent, and SB artifacts were frequently com-pared with the bifacials of the European Solutrean (e.g., Good-win 1929; Heese 1933). Certainly, these scholars got theirtiming wrong and also showed their bias toward a Europeanorigin, but their basic conclusion remains valid. The SB bi-facials represent the uppermost level of technical skill andmethod of manufacture as well as their makers’ often carefulselection for the best-quality fine-grained raw materials (Hen-shilwood et al. 2001b; Villa et al. 2009). These selected rawmaterials may have been chosen partially for their outstandingflaking properties, but their selection may also have beenrelated to adding exchange value to these tools (see Wiessner1983) and the promotion of good social relations (Mellars1996). The latter point might suggest that human groupsduring the SB were integrated in structured exchange net-works similar to those found in living foragers, but at presentthere is insufficient evidence to confirm this proposition.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 13: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

372 Current Anthropology Volume 52, Number 3, June 2011

Ochre, an iron-rich mineral, is found in almost all StoneAge shelter occupations in southern Africa after 100 ka, andthe evidence suggests that it was used as a pigment (Barham1996, 2001; Clark 1989; Henshilwood et al. 2001b; Knight,Powers, and Watts 1995; Singer and Wymer 1982; Wadley2005b; Watts 2009). Ochre was also an important componentin compound adhesives made from red ochre mixed withplant gum that were used to haft or attach stone implementsto handles in the MSA (Wadley 2005a; Wadley, Hodgkiss, andGrant 2009:9590; Wadley, Williamson, and Lombard 2004).Wadley, Hodgkiss, and Grant (2009) stress that the ochre wasnot added for its color but because it played a key role inchanging the adhesive compounds from an acidic to a lessacidic pH, in the dehydration of the adhesive near wood fires,and in changes to the mechanical workability and electrostaticforces of the compound. The ability to produce these complexadhesives relied on multilevel operations and advanced work-ing memory. Wadley, Hodgkiss, and Grant (2009:9590) sug-gest that this archaeological evidence provides a new way toclearly recognize complex cognition in the MSA without nec-essarily invoking the concept of symbolism. In their view, itdemonstrates that there was an overlap between the cognitiveabilities of modern people and those of people in the MSA.

Neanderthals in Europe are also reported to have made abirch pitch that can be produced only at temperatures of 300�–400�C. Koller, Baumer, and Mania (2001) suggest that thesepitches were made with deliberate intent and show consid-erable technical capabilities and that the Neanderthal tech-nical skills, in this instance, were comparable to those ofmodern Homo sapiens.

Early examples of ochre use before 100 ka are reportedfrom Kapthurin (Kenya; McBrearty and Brooks 2000), TwinRiver (Zambia; Barham 2001), and Pinnacle Point Cave 13B(South Africa; Marean et al. 2007). At SB sites, ochre is re-ported from Blombos Cave (Henshilwood et al. 2001b), Hol-low Rock Shelter (Evans 1994), Dale Rose Parlour (Schirmer1975), Diepkloof (Rigaud et al. 2006), and Sibudu (Wadley2007). Evidence of scraping or grinding is observed on manypieces, probably to produce powder, and some pieces havecrayon or pencil shapes as a result of direct application ofthe ochre to an abrasive surface (e.g., Clark et al. 1984; Hen-shilwood et al. 2001b; Rigaud et al. 2006; Singer and Wymer1982:117; Soriano, Villa, and Wadley 2007; Watts 2002, 2009).The presence of ochre at these archaeological sites is inter-preted by some researchers as early evidence of the symbolicuse of colorants (Henshilwood 2009a; Henshilwood, d’Errico,and Watts 2009; Henshilwood et al. 2001b, 2002; Watts 1999,2002, 2009). The available evidence supports a multifunc-tional interpretation for the use of ochre (Lombard 2007b).

At Blombos Cave, hundreds of pieces of ochre more than1 cm in length, many bearing signs of utilization (fig. 5a),have been recovered (M1 and upper M2 phases; Henshilwoodet al. 2001b; Watts 2009). Six deliberately engraved piecescome from the M1 phase and two from the upper M2 phase(Henshilwood, d’Errico, and Watts 2009; Henshilwood et al.

2002). On these pieces, it is clear that the designs were madewith deliberate intent, and they arguably are among the mostcomplex and best formed of claimed early abstract represen-tations (d’Errico et al. 2003; Henshilwood, d’Errico, and Watts2009; Lewis-Williams and Pearce 2004; Mellars 2006; Mithen2006).

Arguments that engraved ochres are restricted only to theBlombos site and that this means that they had no symbolicmeaning beyond the site are now being challenged by newevidence. Engraved patterns on stone and ochre have recentlybeen reported from a number of other MSA sites, for example,at Klein Kliphuis (MacKay and Welz 2008) in the WesternCape and Apollo 11 and Pockenbank (Vogelsang 1998; Vo-gelsang et al. 2010) in Namibia. At the Diepkloof Rock Shelter,Western Cape, 270 pieces of ostrich eggshell (OES) deliber-ately engraved with repetitive patterns, including hatched mo-tifs, were recovered from HP layers with an age of ca. 60 ka(Texier et al. 2010; see “The Howiesons Poort”).

The practice of engraving OES in the HP seems restricted,as far as we know, to the Diepkloof site. That the practice ofengraving in general was apparently confined to only a fewsites might suggest that these sites held a special “power ofplace” on the landscape for some MSA people. Deacon (1988)reports this was the case in the Later Stone Age (LSA) for anumber of rock art sites in the Northern Cape Province. Itis now reasonable to contend that the engraved ochres fromthe Blombos SB levels—and arguably those from other SBsites—and the decorated eggshell from Diepkloof providefirm evidence of early stylistic elaboration in southern Africabefore ca. 60 ka (Henshilwood and Dubreuil 2009; Henshil-wood, d’Errico, and Watts 2009; Mackay and Welz 2008; Watts2009).

Information from SB assemblages was in the past regardedas mostly one-dimensional because only stone artifacts wererecovered during excavations. Four SB sites, three of themrecently excavated, have fauna and organic material preserved.These include Blombos Cave, where fauna and some organicsare well preserved (Henshilwood et al. 2001b); Diepkloof RockShelter (C. Poggenpoel, personal communication), and Si-budu, although the faunal results from the latter two sites areunpublished. Peers Cave, excavated first in the 1920s andfinally in the 1960s, did yield some fauna in the SB/HP units,but only a minute portion was retained (Peers and Goodwin1953; Volman 1981). At Blombos Cave, shellfish, fish, and arange of terrestrial and marine animals were hunted, trapped,and collected during the SB occupation. The overall subsis-tence pattern at Blombos Cave signifies that no clear dis-tinction can be made between LSA and SB subsistence pat-terns at this site (Henshilwood 2008a; Henshilwood et al.2001b); the implication is that during the SB, the subsistencemode was essentially modern.

Formal bone tools, regarded as a distinctive material culturemarker of behavioral modernity in the Upper Palaeolithic inEurasia (e.g., Clark 1989; Klein and Edgar 2002; A. I. Thack-eray 1992), are rare at MSA sites (Backwell and d’Errico 2005;

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 14: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 373

Figure 5. Material culture from the Still Bay levels at Blombos Cave:engraved ochres (a), bone tools (b), and Nassarius kraussianus shell beads(c). Photographs by Francesco d’Errico. Scale bar p 1 cm.

Backwell, d’Errico, and Wadley 2008; d’Errico and Henshil-wood 2007; Henshilwood et al. 2001a). More than 30 bonetools have been recovered from the SB levels at Blombos Cave,including awls and formal “points” (d’Errico and Henshil-wood 2007; Henshilwood and Sealy 1997; Henshilwood et al.2001a, 2001b; fig. 5b), and a further 20 have yet to be de-scribed. The majority are awls made on long-bone shaft frag-ments, further manufactured by scraping and then used topierce soft material, such as leather, and to pierce shells tomake beads (d’Errico and Henshilwood 2007; d’Errico et al.2005). At least five points, completely finished by careful pol-ishing after being shaped by scraping, are probably spearheadsmade for hafting. It is noteworthy that points are treateddifferently from awls. The high polish on these points mayhave served to give them a distinctive appearance, an “addedvalue,” as is reported for projectile points made by the Kal-ahari San (Wiessner 1983). It seems apparent that, here again,the makers showed concern for style and finish. Also note-worthy is the presence of a bone fragment from the SB levelsthat is marked with eight parallel lines. Microscopic analysisof these lines indicates that they are the result of deliberateengraving and possibly were made for symbolic purposes. Anumber of other bone tools carry similar markings (d’Errico,Henshilwood, and Nilsson 2001; d’Errico and Henshilwood2007; Henshilwood et al. 2001a).

The formal ordering or division of living space is regardedby some as a further marker of behavioral modernity (Deacon2001; Henshilwood and Marean 2003, 2006; Kolen 1999;

Wadley 2001b, 2006). This may involve the way that areaswithin a cave or open site were used, for example, the locationof particular hearths, perhaps “owned” by the women withina family group (Deacon 2001), or specific areas used mainlyfor sleeping (bedding hollows) or for tool making. Wadley(2001b:215) describes the use of space in a campsite as beingsimilar to the use of artifactual style. Both require the com-munication of “cultural theory” taught through language. Theappropriate use of space within a cultural group must betaught, and the way that space was used and divided at MSAcampsites was likely passed on across generations. However,archaeological evidence for the physical alteration of camp-sites or caves in the MSA is limited. Wadley (2006) presentsevidence of spatial patterning in the distribution of MSAhearths at Rose Cottage Cave, and Deacon (2001) indicatesthat living space, in particular hearths in the MSA levels atKlasies River, were formally organized.

On the eastern side within Blombos Cave (fig. 6), a chamberfilled with more than 75 cm of unexcavated deposit and ca.15 cm of airspace extends for more than 4 m in a northeasterlydirection. During the SB occupation of the site, the inhabi-tants constructed a 90-cm-long barrier, using quartzite andcalcrete cobbles, to apparently block off this chamber fromtheir living area. The base of the barrier lies on layer CB/CCand extends into layer CA. Layer CC is dated at 72.7 � 3.1ka (Jacobs et al. 2006).

The construction of this barrier may simply have been tokeep out drafts emanating from a possible second entrance

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 15: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

374 Current Anthropology Volume 52, Number 3, June 2011

Figure 6. A 50-cm portion of the barrier of quartzite and calcrete cobblesconstructed along the rear cave wall in the Still Bay levels at BlombosCave. A ca. 4-m-long chamber (arrow) filled with unexcavated depositand an airspace of 10 cm extend to the northeast beyond the barrier andunder the wall.

to the cave that is not currently visible. We think that thisbarrier might also have served to self-consciously and delib-erately alter the space used by the SB people. It may haveacted to separate their social/working area in the open partof the cave from “the other,” a drafty, cramped space. In otherwords, the barrier may represent a manipulation of socialspace by people seeking to delineate the boundary betweensocial or public and nonsocial space and also to improve theconditions in their living space.

Personal ornaments occur in the SB in southern Africa andat sites with similar ages in North Africa and the Near East(Bar-Yosef Mayer, Vandermeersch, and Bar-Yosef 2009; Bou-zouggar et al. 2007; d’Errico, Vanhaeren, and Wadley 2008;d’Errico et al. 2005; Henshilwood et al. 2004; Vanhaeren etal. 2006). The discovery of marine shell beads of the SB atBlombos Cave (d’Errico et al. 2005; Henshilwood et al. 2004;fig. 5c) and at Sibudu (d’Errico, Vanhaeren, and Wadley 2008)has added a new dimension to modern-human-behavior de-bates. An analysis of 41 of the recovered Nassarius kraussianus“tick” shells from Blombos Cave shows that they were care-fully pierced with a bone tool to create a keyhole perforation(d’Errico et al. 2005; Henshilwood et al. 2004). These per-forations are anthropogenic and deliberate. The beads werethen strung, perhaps on cord or sinew, and worn as a personalornament. In addition, the beads provide insights into SBtechnology, including the ability to drill, the use of cord or

gut for threading, and the probable tying of knots to securethe beads. Repeated rubbing of the beads against one anotherand against the cord resulted in discrete use-wear facets oneach bead that are not observed on these shells in their naturalenvironment. These use-wear patterns are the principal factorthat defines the shells as beads. Microscopic residues of ochreoccur inside some of the beads and may result from deliberatecoloring or from transfer when worn (d’Errico et al. 2005;Henshilwood et al. 2004). At Sibudu, six Afrolittorina africanashells were recovered from the SB levels. Three of these hadbeen deliberately pierced, and the most likely interpretationis that they were purposely collected to be transformed intopersonal ornaments (d’Errico, Vanhaeren, and Wadley 2008).

The wearing and display of personal ornaments in the MSA,behavior recently unknown for this time, was not idiosyn-cratic, as is demonstrated by the presence of beads at a numberof MSA sites in Africa (Bouzouggar et al. 2007; d’Errico etal. 2005, 2009; Henshilwood et al. 2004; Vanhaeren et al.2006). At Blombos Cave, discrete groups of beads with wearpatterns and coloring specific to that group were recoveredfrom various levels and squares within the site. This patterningsuggests that at Blombos Cave a number of individuals worebeads, perhaps on their person or attached to clothing orother artifacts. In the literature on the evolution of the mind,the wearing of beads has also been taken as a proxy for mod-ern syntactical language, which would have been essential for

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 16: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 375

the sharing and transmission of the symbolic meaning ofpersonal ornaments within and between groups and also overgenerations (d’Errico et al. 2005; Henshilwood, d’Errico, andWatts 2009; Henshilwood et al. 2004; Mellars 2005, 2006;Watts 2009).

This last interpretation has been contested recently, and adebate has ensued regarding the exact implications of thewearing of beads for the evolution of human cognition (Botha2008, 2009; Coolidge and Wynn 2009; Henshilwood and Du-breuil 2009; Malafouris 2008; Wynn and Coolidge 2007) andgroup identity (Wadley 2001b). Wynn and Coolidge (2007:88) suggest that, at a minimum, personal ornaments indicatethe presence of theory of mind and a capacity to pay attentionto one’s appearance and personal identity. Malafouris (2008:406), in a similar vein, argues that Blombos beads providestronger evidence of “self-awareness” than of symbolism.These arguments are interesting because they both accept thewearing of beads as evidence of a capacity to understand theperspective of others relative to oneself, but at the same theyboth regard the inference that beads act as symbols as un-warranted.

We agree in part with this interpretation. At a minimum,the wearing of beads is one of the many innovations of theSB and HP periods that provide evidence of a concern forstyle and of the appearance of innovative material culture.The emergence of these behaviors, we suggest, is best ex-plained by the impact of an expansion of the later-maturingtemporoparietal areas on higher theory of mind and onperspective-taking abilities. Two things can be said regardingthe inference that beads are symbols. The first is that the useof personal ornaments was not idiosyncratic, since beadscome from different levels and squares. This distribution isdifficult to explain unless we accept that beads functioned assymbols, that is, that their function depended on collectivelyagreed-upon meaning (and not on the idiosyncratic interestof individuals in their appearance). If one contests the sym-bolic interpretation of beads, one has to find a more parsi-monious way to account for the fact that different people atdifferent times used similar personal ornaments.

The second point is that there are good reasons to believethat the same cognitive abilities are required to imbue objectswith symbolic meaning and to develop a concern for one’spersonal appearance. Indeed, both depend on the capacity toinhibit one’s own perspective and pay attention to potentiallyconflicting views on objects. Both depend on the late-maturing areas in the prefrontal and temporoparietal areas aswell as on the connections between them. This parallel ex-plains why young children not only have little concern fortheir appearance but also have great difficulty understandinghow the symbolic meaning of objects depends on collectivelyagreed-upon meaning.

Thus, in our view, the inference that beads can act as sym-bols is parsimonious. Even if it is rejected, however, the cog-nitive implications that we draw from the SB period stand

because of the intimate connection between theory of mind,symbolic behavior, and concern for one’s appearance.

The Howiesons Poort

Accurate modeling of the transition from the SB to the HPis uncertain, but OSL ages indicate that the HP phase maybe restricted, at least at some sites, to the period from ca. 65to 59 ka (Jacobs and Roberts 2008; Jacobs et al. 2008; butsee Tribolo et al. 2009). HP sites are more numerous andgeographically more widely distributed than SB sites in south-ern Africa. Most occur south of the Zambezi River, and 32sites are reported from this region (Jacobs et al. 2008). Thisgeographical limitation, Deacon (1992:181) suggests, definesthe area in which there was a shared concept of artifact styleand the related exchange of information. The principal HPsites are shown in figure 3 (for site details and references, seeHenshilwood 2008b; Jacobs and Roberts 2008; Jacobs et al.2008; Lombard 2005b).

HP-like assemblages are reported in other parts of southernAfrica and north into central and eastern Africa. Examplesare the Bambatan/Tshangulan in Zimbabwe (Barham 2001;Deacon 1995), a Bambatan site in Botswana dated at ca. 70–80 ka (Brooks et al. 1990), and the Mumba site in Tanzania(McBrearty and Brooks 2000). A similar industry is also re-ported by Ambrose at the Norikiushan site in Kenya (Mellars2006). In the absence of dates for many of these northerlysites, it is difficult to relate their age to those of sites insouthern Africa, but Mellars (2006:9384) speculates that theHP, and possibly the SB, may have originated in the easternor central parts of Africa, although there is limited evidenceto support this theory.

An age of about 70 ka for the HP at Klasies River is in-dicated by Deacon (1989, 1992), but a series of dates basedon the thorium-uranium method give an average of about65 ka (Vogel 2001). A suite of luminescence dates for Klasiesindicate an even younger age of about 50–60 ka (Tribolo 2003;Valladas et al. 2005). Francis Thackeray’s (1992, 2007) esti-mates of 58–48 ka, based on microfauna, oxygen and deu-terium isotope ratios, and marine mollusks, are in line withthe luminescence dates. Thorium-uranium dates of 70–60 kaare reported for the HP levels at Boomplaas (Vogel 2001).Luminescence dates of 74–60 ka (Parkington 1999) were pro-vided for the HP at Diepkloof, but it has been argued thatthese dates may be too old (Rigaud et al. 2006). Tribolo etal. (2009), in a new study, however, suggest that at the Diep-kloof site the ages of the HP levels extend from ca. 60 to 96ka.

OSL dates for the HP levels at Sibudu Cave fall within theca. 64–49-ka range (Wadley 2001a; Wadley and Jacobs 2004),and those at Rose Cottage Cave are in the 60–55-ka range(Tribolo 2003; Valladas et al. 2005). Gibson, Wadley, and Wil-liamson (2004) suggest a date of 60–56 ka for the HP at thelatter site. A considerably older date for the HP of 80 ka,based on electron spin resonance, is reported for Border Cave

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 17: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

376 Current Anthropology Volume 52, Number 3, June 2011

(Grun and Beaumont 2001). Tribolo (2003) suggests that thisdate may be too old, and Feathers (2002) indicates his supportfor a younger range of dates for the HP of around 60–55 ka.

The HP deposits at eight sites were recently resampled byJacobs et al. (2008; also Jacobs and Roberts 2008), and theirages were calculated by OSL methods. These sites are Apollo11, Diepkloof Rock Shelter, Klein Kliphuis, Melikane, NtloanaTsoana, Rose Cottage Cave, Sibudu, and Klasies Cave 1a. Onthe basis of these ages, Jacobs et al. suggest that the HP startedat ca. 64.8 ka and ended at 59.5 ka. When these results arecompared with the range of 71.9–71 ka suggested by theseauthors for the SB sites they dated—Apollo 11, DiepkloofRock Shelter, and Sibudu—the inference is that there is a ca.6,700-year gap between these two techno-traditions (Jacobset al. 2008:734). A subsequent set of ages for Diepkloof, cal-culated by the TL method, suggests that the HP levels areconsiderably older at this site (Tribolo et al. 2009), and theseages are not concordant with the Jacobs et al. OSL ages. Tri-bolo et al. report ages ranging from to ka for60 � 6 96 � 10layers associated with HP artifacts. The reasons for the dif-ferences in the ages reported by Jacobs et al. (2008) andTribolo et al. (2009) have not been resolved. In summary, themost consistent ages for HP sites fall within the range 65–59ka.

Segments, blades, and a near-microlithic technology thatwas likely linked to the manufacture of hafted composite toolsare considered by some to be key aspects linking HP lithicswith behavior considered “modern” (Ambrose 2001; Deacon1992; Henshilwood and Marean 2003; Lombard 2005b; Wad-ley 2001b; Wurz 1999, 2000). The short, thin blade blankswith high-angled, off-center platforms that were retouched tomake these backed pieces are seen as characteristic of theindustry (fig. 7; Lombard 2005b; Wurz 2000:138). It seemslikely that these backed pieces were hafted with compoundresins (see Wadley, Hodgkiss, and Grant 2009) in much thesame way as smaller segments were hafted to form arrow tipsin the LSA (fig. 7). Their larger size in the HP suggests thatthey were used in the construction of spear heads and mayhave been replaced when broken (Deacon 1992:180; Villa etal. 2010), although their use as tips for arrows launched froma bow cannot be ruled out (Lombard and Phillipson 2010;Mellars 2006; Wadley and Mohapi 2008). The manufactureof backed segments or crescents, hafting, standardization ofshape and size, the use of resins, and the ability to repaircomposite tools through the replacement of parts are directlycomparable to characteristics of the Wilton period after about12 ka (Deacon 1992).

Blade-based industries are relatively common in the AfricanMSA and also in the European Middle Palaeolithic, but noneof these can be considered microlithic, and backed pieces areabsent. The presence of these features by at least 65 ka in theHP points to an early sophisticated technology that wouldnot be out of place in an LSA or Upper Palaeolithic context(Ambrose 2001; Deacon 1992; Lombard 2007a; Mellars 2006;Minichillo 2005; Wurz 2000). This does not suggest, however,

that this “precocious” industry originated elsewhere (see Mel-lars 2006) or that it resulted from the arrival in the region ofpeople with a different lifeway, as suggested by Singer andWymer (1982). To the contrary, Wurz’s (2000) analysis of thelithics from the HP levels at Klasies River shows them to bean integral part of the MSA in the region. There is also clearlya change in the HP technology over time (Soriano, Villa, andWadley 2007). It is nevertheless of interest that although theterm HP is used to describe a lithic industry that is geograph-ically restricted to southern Africa, Ambrose (2001) points tosimilar lithic industries in eastern and central Africa. Reliabledates are lacking from many of the latter sites, so at this stagethe relationship, temporal and cultural, between these eastern-African industries and the HP is not clearly understood.

Flint knappers in the HP selected for fine-grained raw ma-terials like silcrete and quartz to make the bladelet blanks,while quartzite remained their dominant choice on the coastbut not necessarily inland (Soriano, Villa, and Wadley 2007).This can be seen, for example, at Klasies River and NelsonBay Cave (Minichillo 2005). The reasons for the selection andtransport of nonlocal raw materials, often over long distances,continue to generate debate. Adding value through rarity isone argument, particularly if a system of reciprocal exchangewas in place (Tanaka 1982; Wiessner 1983). Janette Deacon(1995) argues that this was the case during the HP and thatthe exchange of backed artifacts made from exotic raw ma-terials provides clear evidence of symbolism. Ambrose andLorenz (1990) suggest that because people were simply moremobile, perhaps driven by climatic conditions, they also hadgreater access to nonlocal materials.

As is the case for the SB, ochre is ubiquitous at HP sites.Compared to earlier substages of the MSA, there is an increasein both color variability and quantity (Watts 2002; Wurz2000). Much of the ochre was worked to produce powder,which was probably extensively used as a symbolic coloringagent (Deacon 1995; Wurz 2000). At Diepkloof and HollowRock Shelter, pieces of ochre with patterns of incisions similarto those found at Blombos Cave are reported in the MSAlevels (Rigaud et al. 2006:841). An escalating presence of ochrein the SB and the HP suggests the increasing presence ofbehavior mediated by symbolism (Henshilwood 2009a; Hen-shilwood and Dubreuil 2009).

Residue studies on backed pieces from the HP at RoseCottage Cave and Sibudu and on other MSA lithics showconcentrations of ochre and plant material near proximal andmedial portions (Gibson, Wadley, and Williamson 2004; Lom-bard 2007b; Wadley 2005a; Wadley, Hodgkiss, and Grant2009). If ochre assists with the hafting process, as Wadley(2005a; also Wadley, Hodgkiss, and Grant 2009; Wadley, Wil-liamson, and Lombard 2004) suggests for these two sites, thenthe mineral may have been used in the same way to haftbacked pieces or in the manufacture of armatures at otherHP sites. Another use suggested for ochre is to assist in thehealing process (Velo 1984). The iron compounds in ochre,in particular the iron salts, have a powerful astringent and

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 18: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 377

Figure 7. Key items of material culture from the Still Bay (SB) andHowiesons Poort (HP) sites in southern Africa. OES p ostrich eggshell.Photographs courtesy of Francesco d’Errico, Marlize Lombard, and Mar-ian Vanhaeren; image of engraved ostrich eggshells from Diepkloof afterTexier et al. (2010; http://www.pnas.org/content/107/14/6180).

antiseptic effect and are used by Aboriginal peoples in Aus-tralia to arrest hemorrhaging and promote healing.

Bone tools are rarely reported at HP sites, but there areexceptions. A bone point and other bone tools were foundat Sibudu (Backwell, d’Errico, and Wadley 2008), and bonetools and decorated bone were also found at Apollo 11 (Vo-gelsang 1998). At Klasies River, Singer and Wymer (1982)found a bone point (SAM-AA 42160) in Cave 1a in Layer 19,but it is not thought to belong to the MSA layer in which itwas found.

A reexamination of this point by d’Errico and Henshilwood(2007) shows that at the proximal end the piece has beenreduced, suggesting that it was used as a double-ended point.

Also, the manufacturing technique and morphology of thepiece indicate that it is similar to LSA and San bone arrowpoints. In Layer 20 in Cave 1a, Singer and Wymer (1982)report a bone fragment with marks on one surface. A reex-amination of the piece (d’Errico and Henshilwood 2007)shows that it is a midshaft fragment of a large limb bone witha weathered surface and polish in places. Four parallel lineswere incised on this surface with a sharp lithic point, andalthough they were made after the piece had been weathered,their antiquity is supported by their condition. These lineswere deliberately engraved, suggesting that bone was beingused at this time to embody symbolic markings.

At Peers Cave, Jolly recovered a bone point from the talus

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 19: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

378 Current Anthropology Volume 52, Number 3, June 2011

in ca. 1947 that was probably left during the Peers excavationin the 1920s. It is mostly covered in a thick layer of manganese,but traces of manufacture are visible in the form of longi-tudinal grooves produced during scraping (d’Errico and Hen-shilwood 2007). The exposed areas are polished, and the di-mensions of the tool are similar to those of one of the bonepoints from the SB levels at Blombos. Carbon and nitrogenanalysis of the Peers point and several bones from the MSAand LSA levels at the same site demonstrate that the Peerspoint derives from the MSA levels. It likely derives from eitherthe SB or the HP (d’Errico and Henshilwood 2007). Theevidence suggests that bone tools were uncommon during theHP.

A notable find in the ca. 60-ka HP levels at Diepkloof RockShelter are 270 pieces of OES (fig. 7) deliberately marked withrepetitive patterns, including hatched motifs (Parkington etal. 2005; Rigaud et al. 2006:842; Texier et al. 2010). Similarmarked OES is not reported from other HP sites in southernAfrica. Texier et al. (2010) claim that the pieces from Diep-kloof attest to an engraving tradition that is the earliest reliableevidence of what is a widespread modern practice. The re-petitive production of patterns on an everyday object, prob-ably used to carry water, signifies that individual and collectiveidentities and individual expressions are being clearly com-municated (Texier et al. 2010). This symbolic behavior is nodifferent, they contend, from that of LSA and extant com-munities in southern Africa that still make abstract lineardepictions on functional items (eggshell containers) with theclear intention of symbolic communication. These engravedpieces are the most obvious examples of stylistic elaborationand symbolic markings in the HP period, and they comple-ment the engraved ochres recovered from the SB and earlierlevels at Blombos Cave and other MSA sites.

Discussion

Southern Africa and the Evolution of “Modern” Cognition

The archaeological record of the SB and HP periods offers,for the moment, the richest evidence in favor of early yetmultiple behavioral innovations in Homo sapiens. We haveargued that these innovations offer convincing evidence ofan interest in the appearance of artifacts as well as plausibleevidence of symbolically mediated culture. We have also ar-gued that their emergence in the archaeological record is bestexplained by reorganization in modern H. sapiens of the tem-poroparietal areas implicated in higher theory of mind andattentional flexibility or of their connection with other partsof the brain dedicated to higher cognition.

Here we wish to emphasize what this hypothesis does notimply for the larger debate on human cognitive evolution.First, we do not claim that the SB and HP periods are theearliest phases of behavioral innovations in the MSA. Severalputative instances of symbolic behavior, for instance, have

been documented earlier in the MSA, and others almost cer-tainly will be (Barham 2007; Vanhaeren et al. 2006; White etal. 2003). To date, however, the SB and the HP industries stillprovide the best-documented evidence.

A second limitation concerns the evolution of the “mod-ern” mind. Although we believe that the authors of the SBand HP industries were broadly similar to living humans interms of cognitive abilities, we cannot exclude more recentcognitive and neural evolution in modern H. sapiens. Coolidgeand Wynn (2009:240–242), for instance, emphasize that thepresence of symbolically mediated behavior in the SB perioddoes not strictly imply “modern” cognition. They recognizethat the wearing of beads implies a form of theory of mind,but they contend that only the Upper Palaeolithic offers clearevidence of modern cognition, which they associate with en-hanced working memory and executive functions.

We concur with Coolidge and Wynn that it is impossible,on the basis of the SB and HP archaeological record alone,to exclude more recent changes in working memory and ex-ecutive functions. Nevertheless, we doubt that such changescould have been very significant. The behavior of Middle andLate Pleistocene hominins reveals more working-memoryability than suggested by Coolidge and Wynn. As mentionedabove, we already find in Homo heidelbergensis advancedforms of cooperative feeding and breeding. This implies thepresence of a capacity for self-control, which in modern hu-mans depends on the executive functions realized in the late-developing areas of the dorsolateral PFC (Dubreuil 2010b).The emergence of a stylistic and symbolic component in ma-terial culture—because it depends on higher theory of mindand perspective taking—also implies advanced forms of work-ing memory and executive functions. Coolidge and Wynnseem to neglect this point in their discussion of SB personalornaments:

The kinds of social categories implied by the Blombos beads

(I belong to group A, as opposed to group B, with whom

we exchange spouses; group C with whom we fight; group

D who are occasional but unreliable allies, etc.) are not

abstract, and do not combine elements from distinct cate-

gories. Neither are they “superordinate” categories—cate-

gories of categories—which too might implicate executive

functions . . . . So, although the Blombos beads imply cat-

egory formation process, they did not, we believe, require

executive functions and enhanced working memory. Coo-

lidge and Wynn (2009:242)

We think that the categorization process described heredoes not capture the cognitive operation underlying symboluse in humans. As mentioned above, very young children areable to develop expectations of the kind “members of groupA wear those kinds of beads and behave in this way.” Evenbaboons are able to understand group membership and todevelop specific expectations toward the behavior of groupmembers (Cheney and Seyfarth 2007). But baboons do notimbue artifacts with symbolic meaning, and there is little

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 20: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 379

reason to believe that young children would create symbolicculture if they were left on their own.

In what sense are adults different? Adults understand thatthe symbolic meaning of an artifact is a function of whatothers think of these artifacts. Not only do they associatesymbolic markers with particular behavioral expectations (asyoung children and nonhuman primates do); they also un-derstand that their expectations must vary, depending on whatother people think of these markers. The cognitive abilityimplied here is higher theory of mind, which, as develop-mental psychology shows, is very demanding in terms ofworking memory and executive functions.

To be sure, working memory goes beyond theory of mind,and we are happy to agree that the presence of personal or-naments provides no indisputable evidence that the authorsof the SB and HP industries had exactly the same working-memory abilities as living humans. Given the history of mod-ern human populations, however, we doubt that it is parsi-monious to imagine significant differences.

The Ascent and Demise of the SB and the HP

In southern Africa, the SB and the HP represent two re-markable phases that provide vital information about the evo-lution of modern behavior in H. sapiens. Whether these be-havioral developments contributed directly or indirectly tothe expansion about 80–60 ka of H. sapiens within and outsideAfrica and their subsequent dispersal across the Old Worldremains a debatable issue. However, it does raise the questionof the origins of the SB and the HP as well as that of thecause of the behavioral innovations associated with thesetechno-traditions.

Insights about these origins might come from comparisonswith other African regions. In particular, the similarities ofthe SB and HP lithics to industries elsewhere and the use ofNassarius shells as beads in northwestern and southern Africabetween ca. 85 and 75 ka (d’Errico et al. 2009) seem unlikelyto be coincidental. If cultural complexity is linked or en-couraged by ecological variability (e.g., Alvard 2003; Potts1996; Richerson and Boyd 1998), then the reported droughtconditions that persisted from ca. 135 to 70 ka (Scholz et al.2007) in the tropical and subtropical zones in eastern andcentral Africa may have been one stimulant for populationsto migrate into southern or northern Africa (Henshilwood2008b; Marean and Assefa 2005; Mellars 2006; Wadley 2007).It is conceivable that the SB techno-tradition of bifacial pointsemanated from farther north because these points bear atleast some resemblance to the lanceolates of the final Lupem-ban-Tshitolian tradition from Twin Rivers in Zambia, with aprovisional age estimate of ∼95 ka or later (Clark and Brown2001). The Aterian levels at Dar es-Soltan I, on the Atlanticcoast of Morocco, contain thin, bifacially flaked lithic pointsand evidence of personal ornaments (d’Errico et al. 2009) anduse of red ochre. OSL dates for these levels (Barton et al.2009) indicate an MIS 5 age that is coincident with the age

of the SB in southern Africa or even older. The distinct phys-ical similarity between the Zambian lanceolates and those ofthe Aterian and the SB may be more than coincidental (seeClark and Brown 2001:325).

Whether a pan-African cultural tradition was then in placeis speculative, but it is possible that there was a drift of ideasfrom south to north, and vice versa, that may have coincidedwith the movement of people (Henshilwood 2008b). Thissame drift of ideas may have occurred during the HP period,and local variants of the SB and the HP in the western Capemay have developed in situ. If humans migrated to the south-ern and western Cape coast during the SB and HP periods,then an increase in population levels may have stimulatedinnovation and cultural complexity. A problem with thismodel is the low number of SB sites that have been found,suggesting that the human population was not large at thistime.

Conclusion

During just the past decade, our understanding of the MiddleStone Age (MSA) and, in particular, of the SB and the HPhas grown exponentially. Examining all the processes thatinteracted and generated the cultural variants that define theSB and the HP is beyond our relatively limited knowledge ofthese periods, but one avenue that has been pursued in thepast and is worth pursuing in the future is the relationshipbetween material culture and behavioral evolution.

The innovative technologies and social practices recordedat archaeological sites during the period 77–59 ka in the west-ern Cape are only one part of a behavioral montage that alsospread across other regions in Africa at this time. Rapid ad-vances in human cognition after ca. 200 ka were manifestedin material-culture practices not previously observed in theMSA. After ca. 100 ka, symbolically mediated behavior seemsstrongly allied to material culture (Henshilwood, d’Errico, andWatts 2009). The innovations that were manifested duringthe SB and HP periods may have been driven by group orindividual endeavor to adapt or adopt technology to copewith variable environments, or so-called variability selection(Alvard 2003:142).

Once the capacity for complex human culture was in place,the process that followed is unlikely to have been random.Cultural novelty is never random but is generated and assim-ilated both strategically and contextually (Gabora 2001:219).Innovation is a reflection of the accumulated knowledge ofindividuals, the circumstances they find themselves in, andthe social structure in which they are embedded. Innovationswill thus have a much better than chance probability of beingfitter than their predecessors (see Boyd and Richerson 1988;Shennan 2002).

Theoretically, this may explain why the key features of theHP and SB material cultures appear so different. What workedfor one group at 75 ka may not have worked for another at60 ka. The incremental, marginal modifications built up over

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 21: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

380 Current Anthropology Volume 52, Number 3, June 2011

many generations contributed to the complexities of subsis-tence systems, material culture, and languages during the SBand the HP. We are able to work now only with a fractionof the knowledge that drove these processes then, but amidthis complexity it seems that climatic determinants, demo-graphics, and resource distribution, as is the case today, weremajor drivers in promoting innovation and variability in ma-terial culture in the past.

A second avenue that is worth pursuing is the identificationof the cognitive and neural changes that underpinned thebehavioral innovations in southern Africa and probably else-where in Africa during the Late Pleistocene. Tracing likelychanges in cognitive and neural organization in extinct hom-inin populations is a difficult task. Our view is that progressis possible only through the integration of all the relevantmethods, including comparative neuropsychology, develop-mental neuropsychology, and paleoneurology. Many hypoth-eses about the evolution of the mind are proposed withoutdue integration of these methods. As a consequence, someauthors tend to focus on archaeological and psychologicaldata but neglect entirely neural evolution. Others, by contrast,consider comparative or developmental neurosciences but paylittle attention to the details of material culture. Given thewell-known limitations of the different approaches to the evo-lution of the mind and behavior, only the thorough integra-tion of approaches is likely to constrain hypotheses in aninteresting way.

This article builds on this strategy of integration to criticizethree influential hypotheses about the evolution of the mind.Recursion, theory of mind, and executive functions, as ha-bitually formulated, cannot properly account for the evolutionin material culture in the SB and HP periods. The most centralinnovative feature of material culture during these periodswas the presence of a symbolic component or, at a very min-imum, a concern for the appearance of objects, persons, andliving spaces manifested in engraved ochres and ostrich egg-shells, shell beads, finely made bifacials, bone tools, and stan-dardized backed segments as well as in the formal orderingof living spaces.

The hypothesis that we propose implies that both the ca-pacity to engage in symbolically mediated culture and concernfor the appearance of things evolved as a consequence of oneand the same cognitive and neural change. In our view, theintegration of paleoneurology, comparative neuropsychology,and developmental neuropsychology suggests that this changewas linked to a reorganization of the temporoparietal areasimplicated in theory of mind, perspective taking, and atten-tional flexibility—or to improved connectivity of these regionswith the prefrontal cortex.

Acknowledgments

Financial support was provided to C.S.H. by a European Re-search Council Advanced Grant, TRACSYMBOLS no. 249587,

awarded under the FP7 program at the University of Bergen,Norway, and by a National Research Foundation/Departmentof Science and Technology funded Chair at the University ofthe Witwatersrand, South Africa. The funding for B.D. camefrom postdoctoral grants from the Fonds Quebecois de Re-cherche sur la Societe et la Culture and the Social Sciences andHumanities Council of Canada. The authors thank Matt Ros-sano, John Shea, Lyn Wadley, and one anonymous reviewer forcomments on a previous version of the article.

Comments

Frederick L. Coolidge and Thomas WynnPsychology Department, University of Colorado at Colo-rado Springs, 1420 Austin Bluffs Parkway, ColoradoSprings, Colorado 80918, U.S.A. ([email protected])/An-thropology Department, University of Colorado at Colo-rado Springs, 1420 Austin Bluffs Parkway, ColoradoSprings, Colorado 80918, U.S.A. 25 I 11

Henshilwood and Dubreuil have raised the level of discourseon the evolution of the modern mind by providing both anexplicit neurocognitive model for executive functions and athorough review of the archaeological evidence in its support.We agree with their central conclusion that executive functionsare the key neurocognitive ability enabling modern thinking.Here we would like to highlight some of the differences betweentheir approach to executive functions and ours.

We also differ on key points of archaeological inference, inparticular the bridging arguments about beads, human activ-ities, and cognition (Botha 2008, 2010; Wynn 2009). First,although the inference that Still Bay shell beads were orna-ments is reasonable, it is not, in fact, the only possibility. Thebeads may have been tokens on a tallying device, a string ofbeads used to keep track of quantities. Such a use would haveequally interesting cognitive implications that are quite dif-ferent from those of ornament use (Overmann, Wynn, andCoolidge, forthcoming; Wynn, Coolidge, and Overmann,forthcoming). Although one could argue that this tally in-terpretation is less parsimonious than the ornament inter-pretation, both interpretations effectively account for the ar-chaeological evidence. Our second disagreement with the beadchain of inference falls on the link between ornament andconscious symbol use. Often, anthropological literature failsto define symbolism or defines it vaguely, acting as if “I knowit when I see it but cannot define it.” Of course, there isdanger in such failures or vagueness. Henshilwood and Du-breuil employ the words “symbol(s),” “symbolic culture,”“stylistic . . . innovation,” “symbolic component,” “symbol-ically mediated culture,” and “symbolic behavior” withoutclear definitions until well into their discussion. They finallyoffer the following definition of symbolically mediated cul-ture: “one in which individuals understand that artifacts are

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 22: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 381

imbued with meaning and that these meanings are construedand depend on collectively shared beliefs.” They later reinforcethis definition with the contention that the function of sym-bols is dependent on their collectively agreed-upon meaning.One of the problems with such definitions is that they excludeany private meanings for a particular artist or that, as CarlJung claimed, symbols represent the unknowable. Thus, forexample, doodling or cave doodles, which might have heldvery personal meanings for their makers, would not be sym-bolic according to Henshilwood and Dubreuil because theydo not have collectively shared meanings. Such personalmeanings were undoubtedly a precursor to collectively sharedmeanings. We hold to our contention that private meaningsfor personal ornaments and stylistic engravings, requiringonly basic theory of mind and basic self-awareness (a la Ma-lafouris), might better explain their appearance, disappear-ance, and reappearance in the SB and HP archaeological pe-riods. As Henshilwood and Dubreuil have implied, somethingmay have served as a barrier to their transmission, thus re-sulting in their spotty appearance. Why not the absence ofcollectively shared meanings? Instead, it may have been theirprivate, idiosyncratic, or group-delimited meanings that hin-dered their reliable transmission across place and time.

Henshilwood and Dubreuil challenge us to find “a moreparsimonious way to account for the fact that different peopleat different times used similar personal ornaments.” Doesshared agreement about appearance require that people imbueobjects with symbolic meaning? We think not. Many of ourstudents sport body piercings, tattoos, and jewelry. None ad-mits to understanding any meaning to these items. They wearthem because their friends and acquaintances wear them. Yes,they are indexes of membership in some vague group, butthey have no explicit or even implicit meaning beyond asimple association of appearance and social role. It does re-quire self-awareness and awareness of others’ perspective, butnot “inhibit[ing] one’s own perspective” or paying “attentionto potentially conflicting views on objects.” The parsimoniousexplanation is that the beads were indexes, not symbols. Yes,perhaps the Blombos people believed that wearing beadsmeant high status or motherhood, but they need not havedone so. This distinction between index and imbued meaningis not, we think, trivial. Symbolic use of objects in the modernworld relies on both. A Christian cross stands for the res-urrection, but this symbolic meaning builds on the indexicallink between cross and crucifixion. It seems unlikely to usthat arbitrary, imbued meaning arose before indexical use.Indeed, indexical use of objects such as beads may have pro-vided the scaffold on which arbitrary meaning developed.

Henshilwood and Dubreuil confuse our argument aboutthe referential role of beads with their contention that beadsare evidence of modern syntactical language. We have notdenied some referential role for shell beads. We do deny thatshell beads are prima facie evidence of modern syntacticallanguage. Botha (2010) has made clear the dangers of thisinferential leap. Beads were made intentionally. Yes. Beads

were worn. Yes. Beads may have had some personal or limitedinterpersonal (collectively shared) meanings. Yes. Beads areevidence of modern syntactical language. No! Not necessarily.Taking shell beads as evidence of syntactical language is anunjustified inferential leap, and the authors’ implication thatfull behavioral modernity is attached to shell bead makingand incised ochre is unjustified as well.

Henshilwood and Dubreuil also claim that “our focus onthe temporoparietal cortex is original.” We would argue thatit is not. Wilkins (2009) argued the importance of the parietal-occipital-temporal junction (POT) plus Wernicke’s area,which she noted were areas of the brain that are “necessary,if not sufficient, for language” (275). She contended that theneural mosaic comprising Broca’s area and the POT initiallyevolved to improve the neurological control of the hand andthumb but subsequently became available for language andother higher cognitive exaptations after the divergence of thehominid lineage.

Perhaps the weakest of the authors’ suppositions is relatedto their quest to provide “increasingly precise cognitive mech-anisms realized in particular neural structures, processes, ororganizational patterns.” While we think that they appropri-ately evaluated the roles of recursion (inadequate to explainsymbolic thinking alone), theory of mind (the social brainresulted from the evolution of the executive brain), and ex-ecutive functions (complex and imprecisely delineated) in theemergence of symbolically mediated culture, the crux of theirproposal of “more precise” neural substrate and mechanismsis to proffer the temporoparietal juncture and a “higher theoryof mind” (emphasis added). We would contend that offeringthe temporoparietal juncture is not more precise neuronally,and perhaps a greater lacuna is that they never defined whatthey meant by “higher” theory of mind. We do not disagreethat the POT may be critical to modern thinking. Bruner(2010) noted that the hypertrophy of the superior parietallobes in modern humans (compared to Neandertals) was alsoaccompanied by the displacement of the inferior parietal lobesinto Wernicke’s area. Henshilwood and Dubreuil ignore boththe hypertrophy and the displacement in Bruner’s work.However, as we noted, the greater problem with their con-tentions is the vagueness of their hypothesis that symbolicallymediated culture required “higher” theory of mind. We wouldrefer Henshilwood and Dubreuil to Baron-Cohen’s (1995)work on levels of theory of mind. Briefly, he hypothesized amulticomponent theory of mind consisting of four indepen-dent skills: detection of the intentions of others, detection ofeye direction, shared attention, and a final component calledthe “theory of mind module.” The final component, whoseonset in humans is thought to develop by the age of 4, con-tains a complex set of social-cognitive rules and, combinedwith the other three components, creates the full-fledged,adultlike theory of mind. Although Baron-Cohen’s final com-ponent may be criticized for a lack of specificity, we have noidea whatsoever what Henshilwood and Dubreuil mean by“higher” theory of mind.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 23: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

382 Current Anthropology Volume 52, Number 3, June 2011

In summary, this is a noble attempt to propose “precisecognitive mechanisms realized in particular neural struc-tures,” and it is heuristic to review the roles that recursion,executive functions, and theory of mind might play in theemergence of symbolic culture, but Henshilwood and Du-breuil do not succeed overall in proposing a more precisecognitive mechanism or a more precise neural structure.

Michael C. CorballisDepartment of Psychology, University of Auckland, PrivateBag 92019, Auckland 1142, New Zealand ([email protected]). 5 I 11

Henshilwood and Dubreuil seek a neurological explanationfor the apparently abrupt increase in cultural and technolog-ical sophistication from around 77 ka. The difficulty is thatthere is no obvious discontinuity in measurable aspects ofbrain anatomy, and brain size evidently reached a peak inNeanderthals, not in Homo sapiens (Wood and Collard 1999).Some authors, including Chomsky (2010), have argued for amutation, or some other process causing a “rewiring” of thebrain, that gave rise to syntactic language within the past100,000 years, but such arguments are post hoc and in de-fiance of Darwinian principles underlying the evolution ofcomplex structures (Pinker and Bloom 1990).

From an evolutionary perspective, it seems more likely thatlanguage and other recursive cognitive functions, such as the-ory of mind and mental time travel, evolved during the Pleis-tocene (Corballis 2009) and did not appear as a sudden evo-lutionary leap within the past 100,000 years. For most of thisperiod, however, language may have been based primarily onmanual gestures, perhaps with increasing vocal accompani-ment. The gestural theory has a long history, ranging fromCondillac (1971 [1746]) to Hewes (1973), and is increasinglyendorsed (e.g., Arbib 2005; Armstrong and Wilcox 2007; Cor-ballis 2002; Donald 1991; Rizzolatti and Sinigaglia 2008; To-masello 2008). The critical change that occurred within thepast 100,000 years, before the dispersal from Africa, may havebeen the emergence not of language but of speech. Of course,manual gestures persist in signed languages and as an accom-paniment to speech, but effective communication need notdepend on hand movement. Speech may perhaps be regardedas an early example of miniaturization, with the burden ofcommunication restricted to the mouth, freeing the rest ofthe body for manufacture of objects and manipulation of theenvironment.

The emergence of speech would have involved anatomicalchanges, including changes to the vocal tract, and corticalcontrol over vocalization. Whether such changes were presentin Neanderthals remains a matter of contention; Lieberman(2007) has maintained that the Neanderthals would have beenincapable of speech, but this has been challenged (Boe et al.2007; but see also de Boer and Fitch 2010). One possibility

is that a mutation of the FOXP2 gene permitted autonomousvocalization, but there is also controversy as to whether thismutation occurred within the past 100,000 years (e.g., Coopet al. 2008; Enard et al. 2002) or whether it was also presentin Neanderthals. It is also possible that the final emergenceof autonomous speech was a cultural invention (Corballis2002), as was the later invention of writing.

From the perspective of evolutionary psychology (e.g.,Tooby and DeVore 1987), it is likely that the changes thatgave rise to language, and eventually to speech, were incre-mental and took place mainly during the Pleistocene. Facialgestures may have supplemented manual gestures, with voic-ing added to render the partly invisible gestures of the tongueaccessible to the receiver. This might have culminated in au-tonomous vocal gestures, with a nonobligatory manual ac-companiment, in our own species within the past 200,000years. The advantages of vocal over manual communicationwere probably practical rather than linguistic, since signedlanguages appear to have all of the linguistic sophisticationof spoken ones. Speech not only frees the rest of the bodyfor nonlinguistic functions but also allows communication atnight, and it is less demanding of energy resources than ismanual communication. Although signed languages includenoniconic gestures, speech may well have given added impetusto abstract representations, leading more readily to such de-velopments as counting and mathematics—beginning, per-haps, with the wearing of beads!

As the example of writing illustrates, changes in the mediumof communication can have profound influences on humanculture. This is further illustrated by the invention of the print-ing press and most recently of the cellphone and the internet.The earlier discovery of the power of speech over the powerof gesture may have been the most profound of all. The ar-gument presented here is not necessarily in contradiction tothe various possibilities discussed by Henshilwood and Du-breuil, but it may suggest some alternative avenues for explo-ration.

Iain DavidsonDepartment of Archaeology, School of Humanities, Univer-sity of New England, Armidale, New South Wales 2350,Australia ([email protected]). 25 I 11

Inference about the evolution of cognition can take severalforms (for some discussion of this, see Barnard 2010; David-son 2010a). The most secure starts with a model of the cog-nition of the last common ancestor of humans and other apesand then looks at the differences between modern humansand modern nonhuman apes. The third step is to produce amodel of how cognitive processes must work if natural se-lection took our ancestors from that ancestor to ourselves(e.g., Barnard et al. 2007). This model should then be testedagainst the evidence of the archaeological and paleoanthro-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 24: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 383

pological record. Bridging arguments are often needed toshow how archaeological materials demonstrate the cognitionneeded to bring them about (e.g., see how the recognitioncriteria for working memory may be assessed individually forone archaeological problem in Davidson 2010b). Similarly,the paleoanthropological record is most generally thought tocontribute to discussions of cognition through argumentsabout aspects of brain function and mind (e.g., Bruner 2010).The evidence of changes in the size and shape of the externalform of the brain needs arguments linking it to cognition,and these must recognize that the brains of our ancestorswere different from ours. If we are to avoid circular reasoning,however, the only evidence of the way in which those differentbrains functioned is through inference from the behavioralevidence in the archaeological record. Finally, there is a wholebody of theory about how the brain comes to be how it isin both developmental and evolutionary terms. One impor-tant element of this is the question of the impact of learningand changes in the learning environment through time onbrain and cognitive function (for the impact of changes inlife history, see, e.g., Nowell and White 2010).

These are complex issues. Noble and I (Davidson and Noble1989; Noble and Davidson 1991, 1996) struggled hard todevelop a fruitful interaction between archaeology and psy-chology, requiring reassessment of our positions in both dis-ciplines and the successive development of different aspectsof the required thinking. In this we emphasized that languagewas fundamental to human cognition and that the importantaspect of language was the emergence of symbols, broadlyconceived as nothing more than things that stand for some-thing other than themselves (with the exception of the words“symbol” and “word”). Crucially, one important feature ofhuman cognition is the symbolic representation of conceptsin thought before their production. These authors do notseem to embrace this general meaning of the role of symbolsin the construction of key artifacts of the period, as Hen-shilwood did not in an earlier publication (Davidson 2003;Henshilwood and Marean 2003), a criticism that went un-answered then.

Unfortunately, use of the word “language” in this argumentallowed some to think that what we were emphasizing wasthe importance of syntax. Henshilwood was memorably takento task, at a Stellenbosch conference, for not providing thetheoretical underpinnings for such an argument (Botha 2009,2010), although Holloway (1969) had seen the necessity forsuch linking arguments long ago. These authors (Henshil-wood and Dubreuil 2009) did not answer Botha’s criticisms,and here they have shifted away from syntax but still fail tomake the linkage between empirical studies and theoreticalissues of relating the brain evidence securely to cognition andof connecting either brain or cognition to the archaeologicalevidence. I agree that there must be a connection somewhere,but it is not found in this paper.

Several claims appear exaggerated, detracting from the mer-its of the argument. This can be demonstrated in two Stel-

lenbosch conference papers published in a companion volumeto the one in which these authors’ previous collaborative pa-per was published. It is not true that no one has relatednumbers of orders of intentionality to the relative size of theneocortex (Dunbar 2009). Similarly, the authors claim thattheir “focus on the temporoparietal cortex is original,” butWilkins (2009) made this point, reiterating an earlier argu-ment (Wilkins and Wakefield 1995). We are all guilty of miss-ing key publications relevant to our work, but some havemore excuse than others. There seems to be a pattern hereof selective citation and the avoidance of difficult issues.

The finds from Blombos have been meticulously excavatedand documented, and for this Henshilwood deserves lastingcredit. Those finds are of the greatest importance (althoughI have heard some suggest that southern Africa is peripheral),but this paper does not help anyone understand why.

Peter GardenforsCognitive Science, Filosofiska institutionen, Lunds Universi-tet, Box 188, 221 00 Lund, Sweden ([email protected]). 24 I 11

Before presenting my central argument, I want to commenton the notion of a symbol. Henshilwood and Dubreuil definea symbolically mediated culture as “one in which individualsunderstand that artifacts are imbued with meaning and thatthese meanings are construed and depend on collectivelyshared beliefs.” This definition is a variation on the traditionalone within anthropology and archaeology.

Henshilwood and Dubreuil’s definition should be con-trasted with the Saussurian notion of symbol as commonlyused within linguistics, philosophy, and computer science: asymbol is an arbitrary sign used to refer to some entity. Indiscussions of the evolution of language, this definition seemsmore appropriate.

The two meanings of “symbol” are often not kept separatein the literature. Henshilwood and Dubreuil indicate a con-nection between the two when they write that “the wearingof beads has also been taken as a proxy for modern syntacticallanguage.” However, ornaments are not evidence of symboluse in the Saussurian sense, at least not in any direct way. AsHurford (2006) has pointed out, the two meanings are indeedradically different: a Saussurian symbol (typically a word) isarbitrary, is (and must be) learned, is common to the com-munity, is cheap to produce, and is ephemeral (at least whenspoken). In contrast, the anthropological/archaeological sym-bol (typically an artifact) is iconic, invented, individualistic,nonreferential, enduring, and often costly to produce.

My central argument concerns the connection, albeit in-direct, between ornaments and the evolution of language. Inearlier writings, I have focused on the role of cooperation inthe emergence of Saussurian symbols (Gardenfors 2004, 2008,forthcoming; Gardenfors and Osvath 2010). Two forms of

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 25: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

384 Current Anthropology Volume 52, Number 3, June 2011

cooperation seem to be uniquely human and are central tothe evolution of symbolic language: cooperation toward futuregoals and indirect reciprocity. Since future goals are not im-mediately present, symbolic reference is required in order tocommunicate about them. Any evidence of cooperation to-ward future goals in the archaeological record is also evidenceof the presence of Saussurian symbols. From the argumentsof Gardenfors (forthcoming) and Dubreuil (2010b), it seemsreasonable to assume that this kind of cooperation was presentin Homo heidelbergensis, for example, in relation to large-gamehunting (Thieme 1999), as, indeed, Henshilwood and Du-breuil suggest.

More critical for my argument is cooperation in the formof indirect reciprocity. Reciprocal altruism (“You scratch myback, I’ll scratch yours”) is found in several animal species.Indirect reciprocity is a more extreme form of altruism: “Ihelp you, because somebody else will help me.” The condi-tions for this form of cooperation to evolve as an evolution-arily stable strategy have recently been modeled by Nowakand Sigmund (2005). Their model is based on the notion ofindividual reputation, a key factor in determining with whomto cooperate.

Verbal communication is not the only method for signalingand maintaining reputation; ornamental artifacts and bodymarkings can also contribute. In particular, the costly natureof ornaments may be an indicator of the status of the wearer.The cost might be the time invested in producing the artifact,or it might be the danger or effort involved in obtaining theraw materials. (This mirrors to some extent Zahavi’s [1975]handicap principle in biology.) The evidence presented byHenshilwood and Dubreuil indicates that individual reputa-tion was indeed important in the techno-traditions of StillBay and Howiesons Poort.

Henshilwood and Dubreuil argue that imbuing objects withsymbolic meaning and concerning oneself with personal ap-pearance both “depend on the capacity to inhibit one’s ownperspective and to pay attention to potentially conflictingviews on objects.” Both these capacities are indicators of ex-ecutive cognitive functions and advanced theories of mind.Henshilwood and Dubreuil’s mention of “collectively sharedbelief” in defining symbolically mediated culture implies thatthe members of such a culture must have a well-developedtheory of mind. As I argue (Gardenfors 2008), the evolutionof Saussurian symbols also depends critically on an advancedtheory of mind. Both the Saussurian and the anthropological/archaeological approaches to symbols therefore require a the-ory of mind that goes beyond the capacities of nonhomininspecies (as currently understood). The upshot is that, wheninterpreting the significance of the archaeological record, itmight be more rewarding to focus on the complexity of thetheory of mind required (see, e.g., Gardenfors 2007) and theimplications for what forms of cooperation are possible thanon the presence of symbols.

In conclusion, I suggest that the ornaments of Still Bay andHowiesons Poort are best understood as evidence that indirect

reciprocity based on reputation—along with the requisite the-ory of mind—were present in these techno-traditions.

Kathleen R. GibsonDepartment of Neurobiology and Anatomy, University ofTexas Medical School–Houston, 6516 M. D. AndersonBoulevard, Suite 371, Houston, Texas 77035, U.S.A.([email protected]). 24 I 11

Henshilwood and Dubreuil provide a superb review of SouthAfrican archaeological finds and of certain aspects of neuralfunction and intelligence. They ignore, however, the impli-cations of the profound plasticity of the human brain, a plas-ticity that assures that developing human brains are func-tionally responsive to existing cultural conditions (Gibson2002, 2005). Their entire review, moreover, is based on twoquestionable assumptions. (1) Modern intelligence derivesfrom a few key cognitive advances, each of which appearedat a distinct point in human evolution in response to theemergence of a new neocortical area, tract, or organization.(2) These key human neural and cognitive events emerged insynchrony with and were causally responsible for specificevents in the archaeological record, such as the first appear-ance of material symbols and/or blade tools.

Assumptions that the only neural evolutionary events per-tinent to the evolution of intelligence occurred in the neo-cortex or in tracts interconnecting cortical regions are clearlyerroneous (Gibson 1999; Gibson and Jessee 1999). The cere-bellum, basal ganglia, and hippocampus experienced two- tothree-fold increases in absolute size during human evolution,and humans, in comparison to apes, have greater numbersof neuronal fibers that descend from the neocortical motorareas to spinal cord and brain stem motor nuclei. The cer-ebellum contributes to higher cognitive and linguistic func-tions in a major way (De Smet et al. 2007). Along with thebasal ganglia and the premotor cortex, the cerebellum alsomediates procedural learning, a type of learning needed forthe mastery of efficient and novel motor skills such as dance,sport, speech, and tool use. The hippocampus mediates se-mantic memory and contributes to our understandings ofspatial relationships. Direct projections from the neocortexto subcortical motor neurons partially underlie our abilitiesto make precise movements of the hands and speech organs.Hence, human, as opposed to ape, behavioral capacities reflectthe expansion of multiple cortical and subcortical structuresand include advances not only in cognitive and social domainsbut in motor and motor learning domains as well (Gibsonand Jessee 1999).

Brains and guts are both metabolically expensive organs(Aiello and Wheeler 1995), so much so that any major changein brain size requires reduced gut size and hence increasedreliance on high-energy, easy-to-digest foods. At variouspoints in prehistory, this meant increased reliance on meat,

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 26: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 385

tubers, roots, shellfish, fowl, and fish. These changes, in turn,demanded both technological and social advances. From earlyhominin times, the acquisition and preparation of high-energy foods probably demanded digging sticks and cuttingtools. Ultimately, containers, the control of fire, and tools forcapturing and killing animals or fish also became mandatoryin virtually all cultures, as did male/female divisions of labor,food sharing, and adult provisioning of young. The archae-ological record documents many, but not all, of these changes.Both Neandertals and the earliest anatomically modern hu-mans, however, had slightly larger brains, on average, thanmodern humans, and hence must have already evolved met-abolically critical technological and social intelligences. Thekey question is whether either or both of these taxa alreadypossessed essentially modern brain functions or whethersomething new was needed to propel the development of fullymodern intelligence. Arguments that more was needed restprimarily on the absence of evidence of material symbolismin Neandertals and the earliest modern humans. However,many human populations with fully modern intelligence andlanguage create material symbols only from perishable ma-terials such as feathers, bark, wood, and plant fibers. Thus,an absence of material symbols capable of surviving the rav-ages of time for 40,000 years or more cannot be used asevidence that any culture, modern or ancient, truly lackedsymbolism, theory of mind, working memory, or any otheressential component of human intelligence.

The manufacture of beads and the creation of most artwork, perishable or not, requires tools, artifacts, and tech-niques such as awls, cordage, extracted pigments, and weav-ing. Art and material symbolism are thus not merely cognitiveor social advances; they are also technological endeavors thatbuild on preceding technological advances and require themotor and procedural learning skills necessary for skilled ob-ject manipulation. Moreover, the initial appearances of ma-terial symbolism represented only the first of a seeminglynever-ending series of cultural revolutions (e.g., Neolithic,Industrial, computer). This “sapient paradox” (Renfrew 1996)is not explicable in terms of repeated emergences of newneural structures. It is explicable in terms of a plastic braincapable of both absorbing previous technical advances (Ma-lafouris and Renfrew 2010) and building on them throughthe interactions of multiple cognitive, motor, and social ca-pacities (Gibson and Jessee 1999). Parsimony suggests thatall critical capacities were in place by the time brains reachedtheir modern size. Shell beads are simply the earliest incon-trovertible material manifestation of this intelligence.

Lambros MalafourisArchaeology, Anthropology and Human Evolution, KebleCollege, Oxford OX1 3PG, United Kingdom ([email protected]). 1 II 11

The main argument of Henshilwood and Dubreuil can bebroken down into three major parts. (1) A wide range ofbehavioral innovations have been identified in the archaeo-logical remains from southern Africa during the Late Pleis-tocene. (2) These innovations testify to the evolution of asymbolic component in material culture. (3) This stylistic orsymbolic component can be explained in cognitive terms byway of a previous brain reorganization of the temporoparietalareas implicated in theory of mind and perspective taking.

To start, Henshilwood and Dubreuil offer, no doubt, animportant and timely synthesis of the available archaeologicalevidence from the Still Bay and Howiesons Poort periods insouthern Africa. In trying to make sense of the evidence andreveal its cognitive implications, Henshilwood and Dubreuilrightly suggest, it is necessary to move beyond general abilitiesand look for “precise cognitive mechanisms realized in par-ticular neural processes.” The interdisciplinary integration ofall the relevant methods is thus needed. I welcome the di-rection and inspiration that Henshilwood and Dubreuil’s in-tegrative approach provides in current archaeological think-ing. Where I disagree with their analysis, however, is with theunderlying “internalist” and “brainbound” logic that seemsto characterize the way they perceive the role of this meth-odological integration in the archaeology of mind. In partic-ular, the big challenge that Henshilwood and Dubreuil pro-pose is to find the possible links between mind and behavior,more specifically for their purposes here, between symbolicthinking and material culture. Trying to meet this challenge,Henshilwood and Dubreuil go halfway, by drawing attentionto the importance of identifying the cognitive and neuralchanges that might have underpinned the innovations of theLate Pleistocene, but they fall short of recognizing the problemof our limited archaeological understanding of the nature ofmaterial signification (Malafouris 2007, 2010b). As a result,their examination of the archaeological evidence simply re-iterates the usual conventional assumptions about the putative“symbolic” character and stylistic function of certain cate-gories of artifacts. But why, for example, does a series ofdeliberately incised lines suggest or embody symbolic mean-ing? When or how are the markings symbolic? No satisfactoryanswer is given to these basic questions.

To my mind, the presence of a symbolic component as “themost central innovative feature of material culture duringthese periods” should have been the end rather than the start-ing point of Henshilwood and Dubreuil’s analysis. Rather thanbeing taken for granted, how and why a “deliberately en-graved” ochre or a “shell bead” comes to stand for somethingelse in an explicit, conscious way—that is, as an arbitrarysign—is precisely what must be established and accountedfor. Take, for instance, the example of personal decoration.The authors claim that “at a minimum,” the wearing of beadsprovides evidence of a concern for style. They suggest thatthe appearance of beads indicates the presence of developedtheory of mind and a capacity to understand the perspectiveof others relative to oneself. For reasons that I discuss exten-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 27: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

386 Current Anthropology Volume 52, Number 3, June 2011

sively elsewhere, in my approach to early personal decorationI prefer to emphasize the issue of “self-awareness” (Malafouris2008). Nonetheless, I think that Henshilwood and Dubreuil’sinterpretation offers a very interesting and complementaryway of approaching the issue of early personal decoration.What I find problematic, however, is how they use their pro-posed association between changes in prefrontal and tem-poroparietal areas and changes in higher theory of mind andperspective-taking abilities to warrant their claim for the sym-bolic use of beads. I think that their argument at this pointis unsatisfying. I cannot see anything “parsimonious” aboutthe inference that beads can act as symbols, although I dorecognize them as potentially powerful, enactive material signs(Malafouris 2007, 2008) capable of bringing forth a new con-ception of selfhood and perspective taking rather than re-flecting an already formed symbolically equipped humanmind. More is needed before a material sign can be identifiedas a symbol. Any attempt toward a cognitive archaeology ofsymbolism must be grounded on a clear framework of ma-terial semiotics. It is something of a missed opportunity, there-fore, that the authors do not go on to explain how their“strategy of integration” is likely to constrain hypotheses inthis direction.

I would like to end with a final critical comment. Ourattempt at building bridges between brain and culture caneasily mislead us to adopt a sterile “neurocentric” or “inter-nalist” stance on human cognitive becoming. While Hen-shilwood and Dubreuil emphasize the archaeological evi-dence, they do not escape this trap, since they obviously thinkthat it is neural biological change (by way of reorganizationin the human brain) that drives cultural innovation. I believethat this one-directional flow of cause and effect violates thesynergetic embodied nature of the relationship between cog-nition and material culture. In fact, I argue that to prioritize(causally or ontologically) brain reorganization over behav-ioral or material changes is grossly misleading. I contend thatin building a neuroarchaeology of mind, our overarching re-search assumption should be that human intelligence “spreadsout” beyond the skin into culture and the material world(Clark 1997; Malafouris 2010a, 2010c). Our primary focusshould be on the study of the inextricable enfolding of brainand culture, avoiding the reduction of the one to the other.

Dwight ReadDepartment of Anthropology and Department of Statistics,UCLA, Los Angeles, California 90095, U.S.A. ([email protected]). 24 I 11

Henshilwood and Dubreuil provide a useful and timely over-view of the implications of the Still Bay and Howiesons Poorttechno-traditions for early cultural complexity. Their obser-vation that single-factor, mental-capacity explanations do notaccount for the richness of the material products produced

during the Middle Stone Age is well taken. The evolution ofhuman mental processes is almost surely due to multiple,interconnected neurological evolutionary trajectories, includ-ing brain functioning in one domain exapted as part of brainfunctioning in other domains. The question, then, is notwhether recursion, executive function, or theory of mind isthe primary causal driver giving rise to the cognitive abilitiesof modern Homo sapiens. Rather, we need to understand theinterplay among these factors, as shown by the evidence ofevolutionary changes in the trajectory leading to modern H.sapiens. In so doing, we also need to take into account socialconditions that affected further development of one or an-other of our cognitive abilities.

With regard to recursion, the important question is notwhether all languages are syntactically recursive (although thatis an important linguistic question) but the role of recursionin expressing complex relations among ideas and events. Ev-erett (2005, 2009) may be right that the Piraha language ofAmazonia lacks a recursive syntax, but as he points out, thisdoes not mean that recursion is not part of their discourse,as there are “recursive groupings in Piraha, but of ideas ratherthan sentences” (Everett 2009:436). As others have noted, “weexpect those languages [without syntactic recursion] to showrecursion at other levels” (Hollebrandse and Roeper 2007, asquoted in Everett 2009:438). One of these other levels is theincorporation of recursion in lithic technology production,in accordance with expansion in the size of working memoryduring the ancestry of H. sapiens (Read and van der Leeuw2008). In contrast, the limited size of working memory inPan precludes recursion (Read 2008b), and we do not seeevidence of the incorporation of a recursive procedure, ratherthan repetitive flake removals, before the development of theLevallois technique for flake production (Read and van derLeeuw 2008). Recursion became integral to artifact produc-tion with the flourishing prismatic-blade technology in theUpper Paleolithic, and we find in that period a plethora oftools produced using recursive production techniques (Hof-fecker 2007).

The authors make the important point that a critical junc-ture occurred in our cognitive evolution once our ancestorsconceptualized that individual awareness of the meaning ofartifacts—perhaps through the tasks that may be performedwith them—are also meanings common to others as well. Incontrast, even if each chimpanzee cognizes the meaning of atermite stick through its functionality for obtaining termitesfrom a termite mound, the result is shared meaning in onlya statistical sense. The authors note that the critical step goesfrom statistically shared, individual meanings about objectsto awareness that others also hold the same meaning. Sharedmeaning in this sense is a prerequisite to the development ofsymbolically mediated culture. However, the “cognitive dis-tance” from awareness that others hold the same meaning forobjects to symbolically mediated culture is large, and heretheir argument becomes more speculative. The perforatedshells at Blombos suggest a conscious notion of self, as op-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 28: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 387

posed to nonself, as argued by Malafouris (2008), but thelogical negation of self is undifferentiated and does not forma conceptual opposition that is part of symbolically mediatedculture. Contrary to the assertion of Coolidge and Wynn(2009) that categories such as spouse exchange “are not ab-stract” (242, as quoted by Henshilwood and Dubreuil), theformation of categories in opposition, such as wife-giver ver-sus wife-taker, is not simply through the logical negation ofa category but requires the abstract, cultural construction ofcategories in opposition (see El Guindi’s theory of mediatingstructures; El Guindi and Read 1979). In brief, equating (1)shared meaning with (2) categories in opposition with (3)symbolically mediated culture drastically oversimplifies theconceptual distance involved when going empirically from (a)wearing shells on a cord as an indicator of self as opposedto nonself to, for example, (b) a concept of marriage versusreproduction as a basis for kin relations and then to (c) socialorganization based on a culturally constructed kinship system(Read 2001, 2007) that includes the formation of superor-dinate categories of categories and relations of relations. Nor,as the authors point out, can development of conceptual com-plexity be reduced to demographic growth, as has been sug-gested (e.g., Powell, Shennan, and Thomas 2009), since, asthey state, the population size in southern Africa was smallduring the relevant time period. The demographic model isalso contradicted by data from hunter-gatherer groups (Col-lard, Kemery, and Banks 2005; Read 2006, 2008a) and usesinvalid assumptions (Read 2011). Rather than reduction, weneed, as Henshilwood and Dubreuil note, “an integration ofapproaches.”

Matt RossanoDepartment of Psychology, Box 10831, Southeastern Loui-siana University, Hammond, Louisiana 70402, U.S.A.([email protected]). 4 I 11

I commend the authors on an interesting paper that movesdiscussion of the evolution of modern cognition forward. Ihave a question and a suggestion that I hope will continuethat forward motion. First, the question: What is it preciselythat is new and significant about the perspective-taking abilitymanifested in the Still Bay (SB) and Howiesons Poort (HP)industries? It cannot simply be the ability to appreciate thatanother’s perspective on a situation is different and leads toa different knowledge state. Studies with chimpanzees showthat in competitive situations, they appear to understand howanother’s visual viewpoint affects what they know (where foodis hidden and where it is not). Thus, the human/chimpanzeelast common ancestor had some minimal perspective-takingabilities.

These may or may not have included the ability to appre-ciate that others make judgments about the self based on self-appearance or the self’s possessions and products. Evidence

of ochre use (some of it very likely used to enhance self-appearance) and carefully crafted symmetrical hand axes sub-stantially precedes the SB and the HP. When people paintedtheir bodies with ochre or created aesthetically pleasing handaxes, were they not as concerned with how they or theirproducts appeared to others as the makers of the SB/HP in-dustries? Would the authors argue that the hand axe makernever “mentalized” his creation in the same way that the SB/HP artisans did? In other words, the hand axe maker did notunderstand the hand axe as something that others cognitivelyor emotionally evaluated; rather, he simply recognized thatwhen he created a certain visual/physical form in stone, hereceived a positive reaction from others. Maybe so, but whatcan we point to in the archaeological record to support thisinterpretation?

Second, I would like to suggest a way of theoretically sit-uating the beads and other self-reflective artifacts of the SB/HP that can circumvent the unproductive “are they symbolicor not” debate. I suggest that we start thinking about themas an incremental and necessary step toward the wholly ar-bitrary, cognitively demanding, all-encompassing symbolismuniquely associated with our species. Symbolism is a conceptthat encourages dichotomous thinking: you either have it oryou do not, and if you have it, you have modern cognition.But this is too simplistic. There are different levels of refer-ential thinking that make varying cognitive demands. Beadsand other self-reflective artifacts may be understood as sym-bolic in the sense that they stand for something else (e.g., theclan to which one belongs), but they may be best understoodas what Peirce called “indexes.” An index gains its power byvirtue of a consistent temporal or spatial connection to itsreferent (smoke indicates fire, tears indicate sadness, etc.). Ifa certain clan always wears a certain bead, then those beads“stand for” the clan, indexically. True symbols (in the Peirciansense) are more arbitrary and rarely if ever occur in the samecontext as their referent. We regularly use words such as“drudgery” and “angel” with no concrete referent in sight (ormaybe even possible).

Beads are, however, an artificially constructed index. Attheir inception, an arbitrary connection was made: somebodydecided that this bead would serve as an indicator of thatclan. From that point, spatial and temporal reinforcementcould keep the connection going and minimize the cognitiveload. This combination of arbitrariness supported by constantspatial/temporal reinforcement may have served as an im-portant bridging step in moving from simple, naturallyoccurring indexes (smoke-fire) to entirely culturally con-structed, naturally unsupported, cognitively demanding, all-encompassing symbol systems within which human social lifeis fully and uniquely immersed.

An enhancement of perspective-taking ability may be nec-essary before the arbitrary bead-to-clan connection can bemade. The motivation to use an object as a personal or social-group representation may first require the capacity to appre-ciate that another holds the object in high positive regard. If

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 29: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

388 Current Anthropology Volume 52, Number 3, June 2011

I can see that you “like” and “admire” the object, then I mightbe able to redirect that positive regard toward me and mysocial group by adopting it as a personal/social representation.Thus, what we are seeing in the SB/HP are not yet the complexsymbol systems of modern human cognition but an importantmove in that direction: arbitrarily constructed indexes thathelp span the cognitive gap from minimally symbolically ca-pable cognition to fully symbolically immersed cognition.

Modern cognition is not an either/or proposition but some-thing incrementally achieved. Therefore, it should not be sur-prising to find evidence of bits and pieces of it in the ar-chaeological record and in nonhuman species. An enhancedcapacity for perspective taking may indeed be a critical foun-dational piece of modern cognition, but what exactly doesthis enhancement entail, and how do we situate that abilityon the evolutionary road to uniquely human thought?

Lyn WadleySchool of Geography, Archaeology and EnvironmentalStudies, and the Institute for Human Evolution, Universityof the Witwatersrand, P.O. WITS 2050, Johannesburg,South Africa ([email protected]). 20 I 11

This paper is a welcome and fascinating addition to the lit-erature engaging cognitive archaeology. I limit my commentsto a few generic issues that arise from the central premisesof this paper. These premises are that (1) theory of mind wasenabled by functional reorganization of the brain that tookplace during temporoparietal evolution; (2) theory of mindis needed for symbolic manifestations of material culture; and(3) beads and engraved ochre and ostrich eggshell are sym-bolic manifestations.

Complex cognitive capacities, such as those demonstratedby people today, are the result of developments in the brainthrough time; there can be no controversy over this realiza-tion. However, when we start to ask questions (When, why,and how did these enhancements take place, and what ismodern thinking?), either the answers are not forthcomingor they are convoluted.

If we accept that cognition is an evolutionary modification,then it must be an inherited genotype, but we do not yetknow which genetic changes are applicable (Welshon 2010).Genotypic changes certainly influenced cognitive develop-ments, but a lot more research is necessary to understandhow this might have happened to Homo sapiens about 100,000years ago. Neandertals are strictly not relevant to this dis-cussion because they did not live in Africa, yet studies of theirgenes are prescient for African hominids. The fact thatlanguage-related gene FOXP2 is shared by Neandertals andmodern humans implies that it was acquired by a commonancestor before the era that interests us here. A recent, andremarkable, genetic discovery is that Neandertals may havehad kin relationships based on patrilocal mating behavior (still

the most common form today; Lalueza-Fox et al. 2011). Thisbodes well for the possibility that genetic studies will even-tually provide information about forms of social behavior thatlink directly and exclusively to H. sapiens that we recognizein Africa from 100,000 years ago. Regrettably, this has nothappened yet.

Brain encephalization took place much earlier than theperiod under discussion in this paper, and at present thereseems little hope of finding morphological evidence ofchanges leading to cognitive enhancements within the past100,000 years. It seems far more likely, as Henshilwood andDubreuil concede and as Mithen (1996) concluded some yearsago, that neural connectivity between multifarious compo-nents of the brain facilitated advanced cognition. The chal-lenge is to recognize and localize neural changes chronolog-ically (Bruner 2010). The absence of a reliable chronology forthe cognitive developments proposed in the Henshilwood andDubreuil paper is, through no fault of theirs, a major im-pediment. They compensate in the only way that they can:by using circumstantial evidence for the recognition of com-plex cognition and by indirectly dating relevant material cul-ture through dosimetric methods applied to sediments or therocks found in them.

This brings me to the second of their central premises,which is that theory of mind is the cognitive capacity nec-essary for creating symbolic material culture and sharing themeanings of these symbolic items. Theory of mind and socialcognition are said to enable people to attribute mental statesto others, to inhibit self-perspective, and to hold in mindopposing viewpoints. Orders of intentionality have been usedelsewhere to define theory of mind, but Henshilwood andDubreuil exclude this attribute from their definition becausesymbolic behavior can be understood without it. As will beclear from the discussion above, the first appearance of theoryof mind cannot be calibrated chronologically, and eventhough this faculty may be linked to temporoparietal evo-lution, it is also not possible to tie it securely to any recentevolutionary changes in the brain.

Symbolic artifacts are said to have been created as shortcutsfor collectively shared meanings. The consensus view is thatbeads, engraved ochre, and engraved ostrich eggshell are ex-amples of symbolic manifestations in the Middle Stone Ageof Africa (for different opinions, see, e.g., Botha 2008; Wynnand Coolidge 2009), although complex cognition can also beinferred from behavior that has little to do with symbolism(Wadley 2010). Henshilwood and Dubreuil point out thatcomplex theory of mind is not needed to comprehend allsymbolism. Small children (and even nonhuman primates)differentiate social status and categories of people; for in-stance, 2- or 3-year-old children easily recognize firefighters’uniforms, and they know that the behavior associated withwearing this uniform is collectively agreed upon. However,these children cannot create symbolism, even though they canread it. I should be pleased if Henshilwood and Dubreuilcould produce bridging theory and methodology to link the

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 30: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 389

concept of theory of mind with the creation, rather than thereading, of symbolic material culture.

Henshilwood and Dubreuil have come some way to sen-sitizing their readers to research that needs to be undertaken.Every such contribution to cognitive archaeology helps toclear the mist obscuring our view of the mountain before us.One day, when the peaks are sharply in focus, we shall, likeearly travellers, be empowered to plan an ascent or perhapsa strategy for chipping a path to our destination. In brief,there is a lot of work ahead.

Reply

We first want to thanks all commentators for their valuableinputs. We have the impression that the study of cognitiveevolution has made rapid progress in recent years and thatthis is in large part because of the work done by many of ourcommentators. Forums like this one provide an invaluableopportunity to raise the level of the debate, both by clarifyingold issues and by presenting new arguments and evidence.

In this paper, we are indeed tackling an old issue: how onecan infer symbolic material culture from the archaeologicalrecord and how this emergence can be related to the evolutionof the human brain and cognition. Our argument is that wecan infer symbolic culture from the diffusion—in SB and HPmaterial culture—of an interest in the appearance of thingsand that this evolution must be related to the evolution ofan attentional mechanism located in the TPJ and essential tohigher theory of mind. The aim of our paper is to addressprevious criticisms about inferring symbolism from the ar-chaeological record, but we understand from the commentsthat we still have some way to go. We begin by answeringcomments on the nature of symbols and their foundation insocial cognition. We then move to the possibility of relatingcognitive evolution to the SB and HP material culture andconclude by clarifying the implications of our arguments forthe evolution of brain and language.

What is a symbol? At the conceptual level, much of thecontention relates to the notions of symbol and theory ofmind. Matt Rossano notes that understanding symbolism interms of either/or may be unproductive and suggests usingcategories that do not encourage dichotomous thinking. Wewholeheartedly agree with this view and, in fact, have writtenour paper as an attempt to move in this direction. The chal-lenge, however, is to find ways of characterizing types of sym-bol use that make sense at the cognitive level. We found someuseful suggestions in the replies to our paper. Rossano andCoolidge and Wynn, for instance, make use of C. S. Peirce’sdistinction between indices and symbols. Can this distinctionhelp clarify the issues we address?

In Peirce’s view, indices get their meaning by way of phys-ical association with their objects. In his classical text (Peirce

1978 [1931]), he presents guideposts and pronouns as typicalinstances of indices. Symbols, by contrast, get their meaningthrough convention or by some form of tacit contract withina group of speakers. Peirce mentions individual words suchas “give,” “bird,” or “marriage” as typical instances of sym-bols. Peirce’s distinction is useful, but interpreting it at thecognitive level is tricky. The most influential attempt to dothis has been proposed by Deacon (1997). Deacon argues thatunderstanding indices is made possible by the mechanism ofassociative learning. Understanding symbols, by contrast, re-quires something more; it is what Deacon (1997) describesas the capacity to go “beyond the mess of associations” (89).

Deacon’s book advanced our understanding of cognitiveevolution in several ways, but it also introduced some con-fusion into the field. The problem is that both indices andsymbols, as understood by Peirce, implied more at the cog-nitive level than simple associative learning. Most centrally,understanding that a guidepost or a pronoun refers to some-thing—in the same way that we understand the meanings ofwords like “give,” “bird,” or “marriage”—implies the capacityto read the referential intentions of others. Reading inten-tions—the most basic form of theory of mind—goes beyondassociative learning. It implies causal knowledge, in which“intentions” function as underlying causes of observable be-havior (Gopnik and Schulz 2007). This is why children graspthe meaning of indices and single words pretty much at thesame time, around their first birthday or as soon as they beginto grasp referential intentions of others.

The fact that indices and single words develop at the sametime in infancy raises no problem for Deacon, because hethinks that single words are indices. Symbols, he thinks, aresomething else. But what exactly? This is where things getmore elusive. Deacon (1997:83–90) explains that symbolsproper arise when one grasps the “complex functions,”“higher-order regularities,” and “patterns” between indices.We understand symbols when we go beyond the mess ofassociation. But going beyond the mess of association—thatis, accessing causal knowledge—is already a precondition tounderstanding the referential intentions of indices and singlewords, so that we end up with no clear idea as to why indicesshould be seen as precursors of symbols.

We see Peter Gardenfors’s distinction between Saussurianand anthropological meanings of symbols as a more prom-ising way of going beyond an either/or approach to symbols.From a Saussurian viewpoint, a symbol—or rather a sign—is an arbitrary association between a sound and a thought.This view is different from the traditional anthropological orarchaeological view, on which we focus in the paper and whichis closely connected with the idea that material culture isimbued with shared beliefs and meanings. According to Gar-denfors, both types of symbols require more sociocognitiveskills than are found in great apes. He then suggests linkingSaussurian symbols to human ability for cooperation andanthropological/archaeological symbols with human concernswith reputation and indirect reciprocity. From an evolutionary

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 31: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

390 Current Anthropology Volume 52, Number 3, June 2011

viewpoint, human cooperation and Saussurian symbols wouldhave preceded anthropological/archaeological symbols andconcerns for reputation and indirect reciprocity.

Gardenfors’s argument makes much sense at the cognitivelevel and complements our arguments advantageously. Apesdo not spontaneously come to use Saussurian signs or Peir-cean symbols, but the reason is not an inability to go beyondthe mess of association. In fact, they can read intentions, asRossano reminds us, but they lack the socioaffective and so-ciocognitive abilities underlying human cooperation, includ-ing the disposition to spontaneously share attention and in-tentions with conspecifics (Tomasello 2008). Along withGardenfors, we think that evidence in favor of increased co-operation in the archeological record is evidence of theseabilities as well as evidence of Saussurian signs.

From Peirce to piercings. A crucial step in our argumentis the idea that to imbue material culture with symbolic mean-ing requires more than the social-affective and sociocognitiveabilities underlying cooperation and use of Saussurian signs.Gardenfors’s ideas are in agreement with ours on this point,and we suggest that emphasis should be placed on the com-plexity of theory of mind required for particular behaviors.That is precisely what we tried to do by focusing on “highertheory of mind,” “higher perspective taking,” and the “atten-tional mechanism” realized in the TPJ. But what do theseconcepts refer to?

Coolidge and Wynn note that we do not give a properdefinition of them. We are happy to clarify our view if ourdefinition in the paper is not clear. “Higher theory of mind”simply refers to the understanding of false beliefs and otherabstract mental states (e.g., higher-order desires). It is differentfrom the capacity to read intention that is present in apesand young children. “Higher perspective taking” refers to thetask known as “level-2 perspective taking,” in which the sub-ject has to describe how an object appears from someoneelse’s viewpoint, which is different from the capacity to de-termine whether someone else sees an object (an ability thatRossano reminds us is present in both apes and young chil-dren). Children younger than 4 years of age are well knownto fail both higher theory-of-mind and higher perspective-taking tasks, and the TPJ is well known to be involved inthese task in adults.

Coolidge and Wynn suggest that a focus on this mechanismis not as precise as we claim. It is true that “being precise”is a matter of degree. Our hypothesis is more precise thanthose that simply refer to “working memory,” “theory ofmind,” or “recursion,” but it could be further developed indifferent directions. For instance, we do not know whetherthere is a higher-theory-of-mind module realized in the TPJor whether success in the false-beliefs and level-2 perspective-taking tasks is driven by the development of general cognitiveresources in childhood (see fig. 2). What we know for sure,however, is that there is a crucial change in a child’s abilityto deal with abstract mental states at around 4 or 5 years ofage, that these new abilities build on previously existing

intention-reading abilities, and that they further improve dur-ing childhood and adolescence, along with the developmentof general cognitive abilities and of late-maturing brain areas(including the PFC and the TPJ).

This development of social cognition coincides with radicalchanges in children’s behavior, which are of interest for thestudy of cognitive evolution. One of those crucial changes isan interest in fashion and a desire to look fashionable, whichdevelop relatively late in childhood and of which there is littleevidence among young children. Just like apes, children youn-ger than 4 sometimes alter their appearance to produce aninteresting effect on their audience (e.g., wearing witch orfireman costumes for Halloween), but they are not very goodat making sure that they look good. Typically, they do noteven care about looking good or maintaining their “reputa-tion.” The most obvious reason why is that they lack highertheory of mind, which is a precursor of complex social emo-tions such as pride, shame, and contempt. Young childrenand apes see that others see them, but they do not grasp thatthe way others see them has an impact on where they standin the game of indirect reciprocity described by Gardenfors.

This brings us to piercings and tattoos. We are happy togrant to Coolidge and Wynn that those artifacts have nospecific meaning. People wear them because they think thatthey make them look cool. In exchange, they may grant usthat using piercings, tattoos, and jewels is highly demandingin terms of social cognition. Indeed, young children some-times use body adornments, but their behavior is both idi-osyncratic and cheap. By contrast, teenagers wear costlyadornments (jewels are costly, while tattoos and piercings areboth costly and painful), and they do it in ways that are notidiosyncratic: in the United States, no less than 25% of peopleaged between 18 and 50 have tattoos, and 14% have bodypiercings (Laumann and Derick 2006). These two factors (thefact that body adornments are costly and not idiosyncratic)indicate more than basic self-awareness (e.g., “I see that otherssee me”). They indicate the presence of widely shared stan-dards of coolness, as well as of a strong desire in young Amer-icans to improve their position in the game of reputation.

So are piercings and tattoos indices or symbols? It all de-pends on what we mean. On the one hand, we can argue thatpiercings do not symbolize coolness; they indicate it. Theyare thus better understood as indices rather than as symbols.On the other hand, they do not indicate coolness by way ofphysical association—the distinctive feature of Peircean in-dices—because “coolness” is not a physical property that onecan point to ostensively. Coolness refers to the way a behavioror an object is appraised within a certain group. It takes theform of abstract and socially shared standards that we use tofind our way in the game of reputation. Thus, piercings maynot symbolize coolness in the same way that, for instance,the cross symbolizes the crucifixion, but they certainly reston an arbitrary (and conventional) connection between aphysical object and a shared representation of coolness. Thisfeature seems sufficient for piercings to qualify both as sym-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 32: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 391

bolic (at least in Peirce’s sense) and, crucially for our point,as a proxy of higher theory of mind.

Can beads act as symbols? The clarifications above shouldhelp us respond to comments concerning our inferences fromarchaeology to cognition and, more precisely, concerning thepossibility of turning SB and HP innovations into proxies forsymbolically mediated culture. First, Kathleen Gibson remindsus that the absence of earlier evidence of symbolic behavioris not evidence of absence of the capacity for symbolicthought, which might have been in place “by the time brainsreached their modern size.” Similarly, Matt Rossano notesthat early pigment processing and hand axes’ symmetry mightprovide evidence of an interest in the appearance of materialculture that would precede the SB and HP periods. We per-fectly acknowledge these possibilities. In fact, it is quite un-likely that the SB and the HP will provide the earliest evidenceof the cognitive abilities in which we are interested, althoughearlier evidence remains quite flimsy for the moment. Weshould also emphasize, along with Wadley, that the “produc-tion-comprehension gap” is ubiquitous in social cognitionand that it is highly likely that the capacity to understandsymbolic material culture preceded the creation of symbolicartifacts.

More central are the concerns raised as to whether it ispossible to infer symbolically mediated culture from the SBand the HP. Coolidge and Wynn suggest that Blombos marineshells might not have been used as personal ornaments butcould have served as a tallying device. The proposal is worthexamining, as are its cognitive implications. Nevertheless, theinterpretation of shells as personal ornaments seems morereasonable, because the use of personal ornaments is ubiq-uitous among human foragers, while that of tallying devicesis not.

Coolidge and Wynn also argue that beads “are indexes ofmembership in some vague group, but they have no explicitor even implicit meaning beyond a simple association of ap-pearance and social role.” Malafouris contends that we simplyreiterate the assumption that Blombos artifacts are symbolicwithout defending this point explicitly. In the same vein, Readsuggests that there is significant cognitive distance between“awareness that others hold the same meaning for objects”and symbolically mediated culture.

Let us begin by emphasizing one point that, we hope, willhelp clarify the debate. If by “symbolization” we mean therelationship between the cross and crucifixion—the fact thatthe cross stands for crucifixion—then it is perfectly possiblethat SB and HP engravings and personal ornaments were notsymbolic. It is perfectly possible that they acted in the sameway as piercings, simply indicating coolness or good taste.But our point is that turning an object into an indicator ofcoolness makes sense within a social group only if there issome shared understanding of what is cool and what is not.“Being cool” is not like “being red” or “being a bird.” It isan abstract property that indicates where one stands in themarket of reputation, in other words, in the public space

where we care about the fact that people are looking at us ina certain way. We call this conventional link between an ab-stract property (coolness) and an artifact (personal orna-ments) “symbolic.” We think that this use is in line with thePeircean notion of symbol, although maybe not with anyunderstanding of it.

This may also help explain why we think there is no cog-nitive distance between the type of awareness evidenced bySB and HP material culture and symbolically mediated cul-ture. This can be a matter of debate, but we see no primafacie difference, at the cognitive level, between understandingthat there is conventional link between an artifact and anabstract property (e.g., beads and being cool) and understand-ing that there is one between an artifact and an event (e.g.,the cross and crucifixion). In both cases, it is insufficient tosee that people see things: one must see how people see things.This is higher theory of mind and perspective taking.

How does this relate to brain evolution? Is our focus onthe TPJ original? Davidson and Coolidge and Wynn suggestthat it is not, and they point to Wilkins’s (2009) discussionof the importance of this region for the evolution of language.We confess that there is little originality in proposing that theTPJ played some role in cognitive evolution. Our originalityis limited to our use of this brain area to interpret behavioralchanges in early Homo sapiens.

In addition, Davidson suspects our use of “selective” ci-tation and notes that Dunbar (2009) did relate the relativesize of the neocortex to orders of intentionality. We do notdispute that Dunbar (2009) made this connection, but hisargument is based on the (at least) contentious suggestionthat apes have second-order intentionality (compared to first-order intentionality in monkeys) and that there must havebeen a linear increase in orders of intentionality during hu-man evolution, along with encephalization. We reiterate thatDunbar presents no evidence to support this contention.

More fundamentally, Kathleen Gibson criticizes our focuson the neocortex and argues that subcortical structures werealso crucial to cognitive evolution. Her point is hardly dis-putable, and we would be happy to discuss how potentialsubcortical changes can help to shed light on the evolutionof human behavior, but our intention in this paper was toimprove on the most prominent hypotheses that have beenused to account for some southern African Middle Stone Age(and potentially earlier) behavioral innovations.

Gibson and Malafouris both see in our approach a neu-rocentric bias that ignores brain plasticity. We may be re-sponsible for the misunderstanding here, but we want to em-phasize that our primary intention is to establish a correlationbetween the evolution of the TPJ and behavioral evolutionand that we are more than happy to remain agnostic aboutthe exact causal story that has linked brain and behavioralevolution. Given all the uncertainties about early H. sapiens,it would be immensely naive to argue that there has been asingle causal pathway from brain to behavioral evolution—and we do not do so.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 33: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

392 Current Anthropology Volume 52, Number 3, June 2011

And what about language? Corballis suggests that the be-havioral innovations in which we are interested could be re-lated to the evolution of speech rather than of language. Weare not hostile to the idea that speech evolved relatively latein the human lineage, but we do not spontaneously see howit should be related to the behavioral innovations in the SBand the HP. Gestural language is prima facie perfectly adaptedfor communicating the shared meanings in which we areinterested.

Davidson notes that we shift away from syntax and do notexplain how it can be inferred from higher theory of mindand perspective taking. This is true. In Henshilwood and Du-breuil (2009:59), we claim that syntactic recursion is essentialto express metarepresentations. Botha (2010) argues that thisis not necessarily the case, a point also granted in Dubreuil(2010a:124).

If syntactic recursion is not absolutely needed to expressmetarepresentations, does it mean that it is likely to haveevolved after the SB and HP periods? In Henshilwood andDubreuil (2009:59–60), we give reasons (reiterated in thispaper) to think that this is unlikely. One reason is ontogenetic:recursive constructions appear much earlier in language thandoes higher theory of mind (obviously, we know that ontog-eny does not always recapitulate phylogeny). Another reasonis the inability of cognitive linguists and neuroscientists toexplain what the cognitive and neural foundations of syntacticrecursion are, suggesting that recursion may result from moregeneral features of human cognition (including theory ofmind and working memory). Those are not knock-down ar-guments, but Botha (2010) and Davidson have not (yet) pre-sented any reason to think that they should have no weight.

Along with authors such as Tomasello (2003), we are in-creasingly inclined to see syntax as the outcome of a culturalevolutionary process, triggered by the evolution of social cog-nition, rather than as a cognitive development of its own. Wedo not make this point in this paper, and our argument doesnot hinge on the exact relationship between social cognitionand syntax.

—Christopher Stuart Henshilwood and Benoıt Dubreuil

References CitedAboitiz, Francisco, Ricardo R. Garcıa, Conrado Bosman, and Enzo

Brunetti. 2006. Cortical memory mechanisms and language ori-gins. Brain and Language 98(1):40–56.

Aichhorn, Markus, Josef Perner, Martin Kronbichler, Wolfgang Staf-fen, and Gunther Ladurner. 2006. Do visual perspective tasks needtheory of mind? NeuroImage 30(3):1059–1068.

Aiello, Leslie C., and R. I. M. Dunbar. 1993. Neocortex size, groupsize, and the evolution of language. Current Anthropology 34(2):184–193.

Aiello, Leslie C., and Peter Wheeler. 1995. The expensive-tissue hy-pothesis: the brain and the digestive system in human and primateevolution. Current Anthropology 36(2):199–221. [KRG]

Alvard, Michael S. 2003. The adaptive nature of culture. EvolutionaryAnthropology 12(3):136–149.

Amati, Daniele, and Tim Shallice. 2007. On the emergence of modernhumans. Cognition 103(3):358–385.

Ambrose, Stanley H. 2001. Paleolithic technology and human evo-lution. Science 291(5509):1748–1753.

Ambrose, Stanley H., and Karl G. Lorenz. 1990. Social and ecologicalmodels for the MSA in South Africa. In The emergence of modernhumans: an archaeological perspective. Paul Mellars, ed. Pp. 3–33.New York: Cornell University Press.

Anthony, Barbara W. 1963. Report on excavations at Peer’s Cave,Fish Hoek, C. P. March–June 1963. Unpublished manuscript inthe collections of the South African Museum, Cape Town.

———. 1967. Excavations at Peers Cave, Fish Hoek, South Africa.Palaeoecology of Africa 2:58–59.

Apperly, Ian A., and Stephen A. Butterfill. 2009. Do humans havetwo systems to track beliefs and belief-like states? PsychologicalReview 116(4):953–970.

Apperly, Ian A., Dana Samson, and Glyn W. Humphreys. 2009. Stud-ies of adults can inform accounts of theory of mind development.Developmental Psychology 45(1):190–201.

Arbib, Michael A. 2005. From monkey-like action recognition tohuman language: an evolutionary framework for neurolinguistics.Behavioral and Brain Sciences 28(2):105–168. [MCC]

Ardila, Alfredo. 2008. On the evolutionary origins of executive func-tions. Brain and Cognition 68(1):92–99.

Armstrong, David F., and Sherman E. Wilcox. 2007. The gesturalorigin of language. Oxford: Oxford University Press. [MCC]

Backwell, Lucinda, and Francesco d’Errico. 2005. The origin of bonetool technology and the identification of early hominid culturaltraditions. In From tools to symbols: from early hominids to modernhumans. Francesco d’Errico and Lucinda Backwell, eds. Pp. 238–275. Johannesburg: Witwatersrand University Press.

Backwell, Lucinda, Francesco d’Errico, and Lyn Wadley. 2008. MiddleStone Age bone tools from the Howiesons Poort layers, SibuduCave, South Africa. Journal of Archaeological Science 35(6):1566–1580.

Baddeley, Alan D., and Graham Hitch. 1974. Working memory. InThe psychology of learning and motivation: advances in research andtheory, vol. 8. Gordon H. Bower, ed. Pp. 47–89. New York: Aca-demic Press.

Bailey, Phoebe E., and Julie D. Henry. 2008. Growing less empathicwith age: disinhibition of the self-perspective. Journal of Geron-tology: Series B 63(4):P219–P226.

Barham, Lawrence. 1996. Recent research on the Middle Stone Ageat Mumbwa Caves, central Zambia. In Aspects of African archae-ology: papers from the 10th Congress of the Pan African Associationfor Prehistory and Related Studies. Gilbert Pwiti and Robert Soper,eds. Pp. 191–200. Harare: University of Zimbabwe Publications.

———. 2001. Central Africa and the emergence of regional identityin the Middle Pleistocene. In Human roots: Africa and Asia in theMiddle Pleistocene. Lawrence S. Barham and Kate Robson-Brown,eds. Pp. 65–80. Bristol: Western Academic & Specialist.

———. 2007. Modern is as modern does? technological trends andthresholds in the south-central African record. In Rethinking thehuman revolution: new behavioural and biological perspectives onthe origin and dispersal of modern humans. Paul Mellars, KatieBoyle, Ofer Bar-Yosef, and Chris Stringer, eds. Pp. 165–176. Cam-bridge: McDonald Institute for Archaeological Research.

Barnard, Philip J. 2010. Current developments in inferring cognitivecapabilities from the archaeological traces left by stone tools:caught between a rock and a hard inference. In Stone tools andthe evolution of human cognition. April Nowell and Iain Davidson,eds. Boulder: University Press of Colorado. [ID]

Barnard, Philip J., David J. Duke, Richard W. Byrne, and Iain Da-vidson. 2007. Differentiation in cognitive and emotional meanings:an evolutionary analysis. Cognition & Emotion 21(6):1155–1183.[ID]

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 34: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 393

Baron-Cohen, Simon. 1995. Mindblindness: an essay on autism andtheory of mind. Cambridge, MA: MIT Press. [FLC/TW]

Barton, R. N. E., Abdeljalil Bouzouggar, Simon N. Collcutt, Jean-Luc Schwenninger, and Laine Clark-Balzan. 2009. OSL dating ofthe Aterian levels at Dar es-Soltan I (Rabat, Morocco) and im-plications for the dispersal of modern Homo sapiens. QuaternaryScience Reviews 28(19–20):1914–1931.

Bar-Yosef Mayer, Daniella, Bernard Vandermeersch, and Ofer Bar-Yosef. 2009. Shells and ochre in Middle Paleolithic Qafzeh Cave,Israel: indications for modern behavior. Journal of Human Evo-lution 56(3):307–314.

Bialystok, Ellen. 2000. Symbolic representation across domains inpreschool children. Journal of Experimental Child Psychology 76(3):173–189.

Bickerton, Derek. 1990. Language and species. Chicago: University ofChicago Press.

———. 2000. Resolving discontinuity: a minimalist distinction be-tween human and non-human minds. American Zoologist 40(6):862–873.

———. 2003. Symbol and structure: a comprehensive frameworkfor language evolution. In Language evolution. Morten H. Chris-tiansen and Simon Kirby, eds. Pp. 77–93. Oxford: Oxford Uni-versity Press.

Blakemore, Sarah-Jayne. 2008. The social brain in adolescence. Na-ture Reviews Neuroscience 9(4):267–277.

Bloom, Paul, and Lori Markson. 1998. Capacities underlying wordlearning. Trends in Cognitive Science 2(2):67–73.

Boe, Louis-Jean, Jean-Louis Heim, Kiyoshi Honda, Shinji Maeda,Pierre Badin, and Christian Abry. 2007. The vocal tract of newbornhumans and Neanderthals: acoustic capabilities and consequencesfor the debate on the origin of language. A reply to Lieberman(2007a). Journal of Phonetics 35(4):564–581. [MCC]

Bookstein, Fred L., Katrin Schafer, Hermann Prossinger, Horst Seid-ler, Martin Fieder, Chris Stringer, Gerhard W. Weber, et al. 1999.Comparing frontal cranial profiles in archaic and modern Homoby morphometric analysis. The Anatomical Record 257(6):217–224.

Botha, Rudolf. 2008. Prehistoric shell beads as a window on languageevolution. Language and Communication 28:197–212.

———. 2009. Theoretical underpinnings of inferences about lan-guage evolution. In The cradle of language. Rudolf Botha and ChrisKnight, eds. Pp. 93–111. Oxford: Oxford University Press.

———. 2010. On the soundness of inferring modern language fromsymbolic behaviour. Cambridge Archaeological Journal 20(3):345–356.

Bouzouggar, Abdeljalil, Nick Barton, Marian Vanhaeren, Francescod’Errico, Simon Collcutt, Tom Higham, Edward Hodge, et al. 2007.82,000-year-old shell beads from North Africa and implicationsfor the origins of modern human behavior. Proceedings of theNational Academy of Sciences of the USA 104(24):9964–9969.

Bouzouggar, Abdeljalil, Janusz K. Kozłowski, and Marcel Otte. 2002.Study of the Aterian lithic assemblages from El Aliya cave in Tan-gier (Morocco). L’Anthropologie 106(2):207–248.

Boyd, Robert, and Peter J. Richerson. 1988. Culture and the evolu-tionary process. Chicago: University of Chicago Press.

Brooks, Alison, P. Edgar Hare, Julie E. Kokis, Gifford H. Miller, R.D. Ernst, and Fred Wendorf. 1990. Dating Pleistocene archaeo-logical sites by protein diagenesis in ostrich eggshell. Science248(4951):60–64.

Bruner, Emiliano. 2003. Fossil traces of the human thought: paleo-neurology and the evolution of the genus Homo. Rivista di An-tropologia (Journal of Anthropological Sciences) 81:29–56.

———. 2004. Geometric morphometrics and paleoneurology: brainshape evolution in the genus Homo. Journal of Human Evolution47(5):279–303.

———. 2010. Morphological differences in the parietal lobes within

the human genus: a neurofunctional perspective. Current Anthro-pology 51(suppl. 1):S77–S88. [FLC/TW, ID, LW]

Bruner, Emiliano, and Ralph L. Holloway. 2010. A bivariate approachto the widening of the frontal lobes in the genus Homo. Journalof Human Evolution 58(2):138–146.

Bruner, Emiliano, Giorgio Manzi, and Juan Luis Arsuaga. 2003. En-cephalization and allometric trajectories in the genus Homo: evi-dence from the Neandertal and modern lineages. Proceedings ofthe National Academy of Sciences of the USA 100(26):15335–15340.

Burkitt, Miles C. 1928. South Africa’s past in stone and paint. London:Cambridge University Press.

Call, Josep, and Michael Tomasello. 2008. Does the chimpanzee havea theory of mind? 30 years later. Trends in Cognitive Sciences 12(5):187–192.

Carlson, Stephanie M., Louis J. Moses, and Laura J. Claxton. 2004.Individual differences in executive functioning and theory of mind:an investigation of inhibitory control and planning ability. Journalof Experimental Child Psychology 87(4):299–319.

Casanova, Manuel F., and Christopher R. Tillquist. 2008. Encephal-ization, emergent properties, and psychiatry: a minicolumnar per-spective. Neuroscientist 14(1):101–118.

Cela-Conde, Camilo J., Gisele Marty, Fernando Maestu, Tomas Ortiz,Enric Munar, Alberto Fernandez, Miquel Roca, Jaume Rossello,and Felipe Quesney. 2004. Activation of the prefrontal cortex inthe human visual aesthetic perception. Proceedings of the NationalAcademy of Sciences of the USA 101(16):6321–6325.

Cheney, Dorothy L., and Robert M. Seyfarth. 2007. Baboon meta-physics: the evolution of a social mind. Chicago: University of Chi-cago Press.

Chomsky, Noam. 2010. Some simple evo devo theses: how true mightthey be for language? In The evolution of human language. RichardK. Larson, Viviane Deprez, and Hiroko Yamakido, eds. Pp. 45–62. Cambridge: Cambridge University Press. [MCC]

Clark, Andy. 1997. Being there: putting brain, body and world togetheragain. Cambridge, MA: MIT Press. [LM]

Clark, J. Desmond. 1989. The origins and spread of modern humans:a broad perspective on the African evidence. In The human rev-olution: behavioural and biological perspectives on the origins of mod-ern humans. Paul Mellars and Chris Stringer, eds. Pp. 566–588.Edinburgh: Edinburgh University Press.

Clark, J. Desmond, and K. S. Brown. 2001. The Twin Rivers Kopje,Zambia: stratigraphy, fauna, and artefact assemblages from the1954 and 1956 excavations. Journal of Archaeological Science 28(3):305–330.

Clark, J. Desmond, Kenneth D. Williamson, Joseph W. Michels, andCurtis A. Marean. 1984. A Middle Stone Age occupation site atPorc Epic Cave, Dire Dawa (east-central Ethiopia). African Ar-chaeological Review 2:37–71.

Collard, Mark, Michael Kemery, and Samantha Banks. 2005. Causesof toolkit variation among hunter-gatherers: a test of four com-peting hypotheses. Canadian Journal of Archaeology 29(1):1–19.[DR]

Condillac, Etienne B. de. 1971 (1746). An essay on the origin of humanknowledge. Gainesville, FL: Scholars’ Facsimiles and Reprints.[MCC]

Coolidge, Frederick L., and Thomas Wynn. 2001. Executive functionsof the frontal lobes and the evolutionary ascendancy of Homosapiens. Cambridge Archaeological Journal 11(2):255–260.

———. 2004. A cognitive and neuropsychological perspective on theChatelperronian. Journal of Anthropological Research 60(1):55–73.

———. 2005. Working memory, its executive functions, and theemergence of modern thinking. Cambridge Archaeological Journal15(1):5–26.

———. 2007. The working memory account of Neandertal cogni-tion: how phonological storage capacity may be related to recur-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 35: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

394 Current Anthropology Volume 52, Number 3, June 2011

sion and the pragmatics of modern speech. Journal of HumanEvolution 52(6):707–710.

———. 2009. The rise of Homo sapiens: the evolution of modernthinking. Chichester: Wiley-Blackwell.

Coop, Graham, Kevin Bullaughey, Francesca Luca, and Molly Prze-worski. 2008. The timing of selection of the human FOXP2 gene.Molecular Biology and Evolution 25(7):1257–1259. [MCC]

Corballis, Michael C. 2002. From hand to mouth. Princeton, NJ:Princeton University Press. [MCC]

———. 2007. Recursion, language, and starlings. Cognitive Science31(4):697–704.

———. 2009. The evolution of language. Proceedings of the New YorkAcademy of Sciences 1156:19–43. [MCC]

Dale, Langham. 1870. Stone implements in South Africa. CapeMonthly Magazine 1:23–29.

Davidson, Iain. 2003. The origin of modern human behavior: a re-view and critique of the models and their test implications (Com-ment on Henshilwood and Marean). Current Anthropology 44(5):637–638. [ID]

———. 2010a. The archaeology of cognitive evolution. Wiley Inter-disciplinary Reviews: Cognitive Science 1(2):214–229. [ID]

———. 2010b. The colonization of Australia and its adjacent islandsand the evolution of modern cognition. Current Anthropology51(suppl. 1):S177–S189. [ID]

Davidson, Iain, and William Noble. 1989. The archaeology of per-ception: traces of depiction and language. Current Anthropology30(2):125–155. [ID]

Deacon, Hilary J. 1989. Late Pleistocene palaeoecology and archae-ology in the southern Cape, South Africa. In The human revolution:behavioural and biological perspectives on the origins of modern hu-mans. Paul Mellars and Chris Stringer, eds. Pp. 547–564. Edin-burgh: Edinburgh University Press.

———. 1992. Southern Africa and modern human origins. Philo-sophical Transactions of the Royal Society B: Biological Sciences337(1280):177–183.

———. 2001. Modern human emergence: an African archaeologicalperspective. In Humanity from African naissance to coming millen-nia: colloquia in human biology and palaeaonthropology. Phillip V.Tobias, Michael A. Raath, Jacopo Moggi-Cecchi, and Gerald A.Doyle, eds. Pp. 217–226. Florence: Florence University Press.

Deacon, Janette. 1988. The power of a place in understanding south-ern San rock engravings. World Archaeology 20(1):129–140.

———. 1995. An unsolved mystery at the Howieson’s Poort namesite. South African Archaeological Bulletin 50(162):110–120.

Deacon, Terrence W. 1997. The symbolic species: the co-evolution oflanguage and the brain. New York: Norton.

de Boer, Bart, and W. Tecumseh Fitch. 2010. Computer models ofvocal tract evolution: an overview and critique. Adaptive Behavior18(1):36–47. [MCC]

Decety, Jean, and Claus Lamm. 2007. The role of the right tempo-roparietal junction in social interaction: how low-level computa-tional processes contribute to meta-cognition. Neuroscientist 13(6):580–593.

DeLoache, Judy S. 2002. Early development of the understandingand use of symbolic artefacts. In Blackwell handbook of cognitivedevelopment. Usha C. Goswami, ed. Pp. 207–226. Oxford: Black-well.

———. 2004. Becoming symbol-minded. Trends in Cognitive Sciences8(2):66–70.

Dennett, Daniel C. 1987. The intentional stance. Cambridge, MA:MIT Press.

d’Errico, Francesco, and Christopher S. Henshilwood. 2007. Addi-tional evidence for bone technology in the southern African MiddleStone Age. Journal of Human Evolution 52(2):142–163.

d’Errico, Francesco, Christopher Henshilwood, Graeme Lawson,Marian Vanhaeren, Anne-Marie Tillier, Marie Soressi, Frederique

Bresson, et al. 2003. Archaeological evidence for the emergence oflanguage, symbolism, and music: an alternative multidisciplinaryperspective. Journal of World Prehistory 17(1):1–70.

d’Errico, Francesco, Christopher S. Henshilwood, and Peter Nilssen.2001. An engraved bone fragment from ca. 75,000-year-old MiddleStone Age levels at Blombos Cave, South Africa: implications forthe origin of symbolism and language. Antiquity 75(2):309–318.

d’Errico, Francesco, Christopher Henshilwood, Marian Vanhaeren,and Karen van Niekerk. 2005. Nassarius kraussianus shell beadsfrom Blombos Cave: evidence for symbolic behaviour in the Mid-dle Stone Age. Journal of Human Evolution 48(1):3–24.

d’Errico, Francesco, and Marian Vanhaeren. 2009. Earliest personalornaments and their significance for the origin of language debate.In The cradle of language. Rudolf Botha and Chris Knight, eds.Pp. 24–60. Oxford: Oxford University Press.

d’Errico, Francesco, Marian Vanhaeren, Nick Barton, Abdeljalil Bou-zouggar, Henk Mienis, Daniel Richter, Jean-Jacques Hublin, Shan-non P. McPherron, and Pierre Lozouet. 2009. Additional evidenceon the use of personal ornaments in the Middle Paleolithic ofNorth Africa. Proceedings of the National Academy of Sciences ofthe USA 106(38):16051–16056.

d’Errico, Francesco, Marian Vanhaeren, and Lyn Wadley. 2008. Pos-sible shell beads from the Middle Stone Age layers of Sibudu Cave,South Africa. Journal of Archaeological Science 35(10):2675–2685.

De Smet, Hyo Jung, Hanne Baillieux, Peter P. De Deyn, Peter Marien,and Philippe Paquier. 2007. The cerebellum and language: the storyso far. Folia Phoniatrica et Logopaedica 59(4):165–170. [KRG]

de Villiers, Jill G., and Jennie E. Pyers. 2002. Complements to cog-nition: a longitudinal study of the relationship between complexsyntax and false-belief-understanding. Cognitive Development17(1):1037–1060.

Donald, Merlin. 1991. Origins of the modern mind: three stages in theevolution of culture and cognition. Cambridge, MA: Harvard Uni-versity Press. [MCC]

Dubreuil, Benoıt. 2008. What do modern behaviors in Homo sapiensmean for the evolution of language? In The evolution of language:proceedings of the 7th International Conference (EvoLang 7). AndrewD. M. Smith, Kenny Smith, and Ramon Ferrer i Cancho, eds. Pp.99–106. Singapore: World Scientific.

———. 2010a. Human evolution and the origin of hierarchies. Cam-bridge: Cambridge University Press.

———. 2010b. Paleolithic public goods games: why human cultureand cooperation did not evolve in one step. Biology and Philosophy25(1):53–73.

Dunbar, R. I. M. 1992. Neocortex size as a constraint on group sizein primates. Journal of Human Evolution 22(6):469–493.

———. 1993. Co-evolution of neocortical size, group size, and lan-guage in humans. Behavioral and Brain Sciences 16(4):681–735.

———. 2003. The social brain: mind, language and society in evo-lutionary perspective. Annual Review of Anthropology 32:163–181.

———. 2007. The social brain and the cultural explosion of thehuman revolution. In Rethinking the human revolution: new be-havioural and biological perspectives on the origins and dispersal ofmodern humans. Paul Mellars, Katie Boyle, Ofer Bar-Yosef, andChris Stringer, eds. Pp. 91–98. Cambridge: McDonald Institute forArchaeological Research.

———. 2009. Why only humans have language. In The prehistoryof language. Rudolf Botha and Chris Knight, eds. Pp. 12–35. Ox-ford: Oxford University Press.

El Guindi, Fadwa, and Dwight W. Read. 1979. Mathematics in struc-tural theory. Current Anthropology 20(4):761–790. [DR]

Enard, Wolfgang, Molly Przeworski, Simon E. Fisher, Cecilia S. L.Lai, Victor Wiebe, Takashi Kitano, Anthony P. Monaco, and SvantePaabo. 2002. Molecular evolution of FOXP2, a gene involved inspeech and language. Nature 418(6900):869–871. [MCC]

Evans, Nicholas, and Stephen C. Levinson. 2009. The myth of lan-

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 36: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 395

guage universals: language diversity and its importance for cog-nitive science. Behavioral and Brain Sciences 32(5):429–492.

Evans, Ursula. 1994. Hollow Rock Shelter, a Middle Stone Age sitein the Cederberg. Southern African Field Archaeology 3(2):63–73.

Everett, Daniel L. 2005. Cultural constraints on grammar and cog-nition in Piraha: another look at the design features of humanlanguage. Current Anthropology 46(4):621–646.

———. 2009. Piraha culture and grammar: a response to some crit-icisms. Language 85(2):405–442.

Feathers, James K. 2002. Luminescence dating in less than ideal con-ditions: case studies from Klasies River Main Site and Duinefon-tein, South Africa. Journal of Archaeological Science 29(2):177–194.

Fitch, W. Tecumseh, Marc D. Hauser, and Noam Chomsky. 2005.The evolution of the language faculty: clarifications and implica-tions. Cognition 97(2):179–210.

Flavell, John H. 1992. Perspectives on perspective-taking. In Piaget’stheory: prospects and possibilities. Harry Beilin and Peter B. Pufall,eds. Pp. 107–139. Hillsdale, NJ: Erlbaum.

Flavell, John H., Eleanor R. Flavell, and Frances L. Green. 1983.Development of the appearance-reality distinction. Cognitive Psy-chology 15(1):95–120.

Freud, Sigmund. 1966. Introductory lectures on psychoanalysis. JamesStrachey, ed. New York: Norton.

Friederici, Angela D. 2009. Pathways to language: fiber tracts in thehuman brain. Trends in Cognitive Sciences 13(4):175–181.

Gabora, L. 2001. Cognitive mechanisms underlying the origin andevolution of culture. PhD thesis, Vrije Universiteit, Brussels.

Gardenfors, Peter. 2004. Cooperation and the evolution of symboliccommunication. In Evolution of communication systems: a com-parative approach. D. Kimbrough Oller and Ulrike Griebel, eds.Pp. 237–256. Cambridge, MA: MIT Press. [PG]

———. 2007. Evolutionary and developmental aspect of intersub-jectivity. In Consciousness transitions: phylogenetic, ontogenetic andphysiological aspects. Hans Liljenstrom and Peter Arhem, eds. Pp.281–385. Amsterdam: Elsevier. [PG]

———. 2008. The role of intersubjectivity in animal and humancooperation. Biological Theory 3(1):1–12. [PG]

———. Forthcoming. The role of cooperation in the evolution ofprotolanguage and language. In The evolution of mind and culture:proceedings of the conference. Gary Hatfield et al., eds. Harrisburg:Pennsylvania State Museum. [PG]

Gardenfors, Peter, and Mathias Osvath. 2010. Prospection as a cog-nitive precursor to symbolic communication. In Evolution of lan-guage: biolinguistic perspectives. Richard K. Larson, Viviane Deprez,and Hiroko Yamakido, eds. Pp. 103–114. Cambridge: CambridgeUniversity Press. [PG]

Gatehouse, R. P. 1954. The prehistoric site at Cape Hangklip. Trans-actions of the Royal Society of South Africa 34(2):335–344.

Gazzaniga, Michael S., Richard B. Ivry, and George R. Mangun, eds.2002. Cognitive neuroscience: the biology of the mind. 2nd edition.New York: Norton.

Gentner, Timothy Q., Kimberly M. Fenn, Daniel Margoliash, andHoward C. Nusbaum. 2006. Recursive syntactic pattern learningby songbirds. Nature 440(7088):1204–1207.

Gibson, Kathleen R. 1999. Social transmission of facts and skills inthe human species: neural mechanisms. In Mammalian sociallearning: comparative and ecological perspectives. Hilary O. Box andKathleen R. Gibson, eds. Pp. 351–366. Cambridge: CambridgeUniversity Press. [KRG]

———. 2002. Customs and cultures in animals and humans: neu-robiological and evolutionary perspectives. Anthropological Theory2:323–339. [KRG]

———. 2005. Epigenesis, brain plasticity, and behavioral versatility:alternatives to standard evolutionary theory. In Complexities: be-yond nature and nurture. Susan McKinnon and Sydel Silverman,eds. Pp. 23–42. Chicago: University of Chicago Press. [KRG]

Gibson, Kathleen R., and Stephen Jessee. 1999. Language evolutionand the expansion of multiple neurological processing areas. InThe origins of language: what nonhuman primates can tell us. Bar-bara J. King, ed. Pp. 189–228. Santa Fe, NM: School for AmericanResearch. [KRG]

Gibson, N. E., Lyn Wadley, and Bonny Williamson. 2004. Microscopicresidues as evidence of hafting on backed tools from the 60,000to 68,000 year-old Howiesons Poort layers of Rose Cottage Cave,South Africa. Southern African Humanities 16:1–11.

Gobbini, Maria I., Aaron C. Koralek, Ronald E. Bryan, Kimberly J.Montgomery, and James V. Haxby. 2007. Two takes on the socialbrain: a comparison of theory of mind tasks. Journal of CognitiveNeuroscience 19(11):1803–1814.

Gogtay, Nitin, Jay N. Giedd, Leslie Lusk, Kiralee M. Hayashi, DeannaGreenstein, A. Catherine Vaituzis, Tom F. Nugent III, et al. 2004.Dynamic mapping of human cortical development during child-hood through early adulthood. Proceedings of the National Academyof Sciences of the USA 101(21):8174–8179.

Goodwin, A. J. H. 1929. The Stone Ages in South Africa. Africa:Journal of the International African Institute 2(2):174–182.

Goodwin, A. J. H., and Bertie Peers. 1953. Two caves at Kalk Bay,Cape Peninsula. South African Archaeological Bulletin 8(31):59–77.

Gopnik, Alison, and Laura Schulz, eds. 2007. Causal learning: psy-chology, philosophy, and computation. New York: Oxford UniversityPress.

Grun, Rainer, and Peter Beaumont. 2001. Border Cave revisited: arevised ESR chronology. Journal of Human Evolution 40(6):467–482.

Hauser, Marc D., Noam Chomsky, and W. Tecumseh Fitch. 2002.The faculty of language: what is it, who has it, and how did itevolve? Science 298(5598):1569–1579.

Heese, Christoph H. T. D. 1933. The evolution of Palaeolithic tech-nique. Annals of the University of Stellenbosch XI(2):1–65.

———. n.d. Collected papers. African Heritage Research Institute,Cape Town.

Henshilwood, Christopher S. 2005. Stratigraphic integrity of the Mid-dle Stone Age levels at Blombos Cave. In From tools to symbols:from early hominids to modern humans. Francesco d’Errico andLucinda Backwell, eds. Pp. 441–458. Johannesburg: WitwatersrandUniversity Press.

———. 2007. Fully symbolic sapiens behaviour: innovation in theMiddle Stone Age at Blombos Cave, South Africa. In Rethinkingthe human revolution: new behavioural and biological perspectiveson the origins and dispersal of modern humans. Paul Mellars, KatieBoyle, Ofer Bar-Yosef, and Chris Stringer, eds. Pp. 123–132. Cam-bridge: McDonald Institute for Archaeological Research.

———. 2008a. Holocene prehistory of the southern Cape, South Africa:excavations at Blombos Cave and the Blombosfontein Nature Reserve.Cambridge Monographs in African Archaeology 75. BAR Inter-national Series S1860. Cambridge: British Archaeological Reports.

———. 2008b. Winds of change: palaeoenvironments, material cul-ture and human behaviour in the Late Pleistocene (c. 77–48 ka)in the Western Cape Province, South Africa. South African Ar-chaeological Society Goodwin Series 10:35–51.

———. 2009a. The origins of symbolism, spirituality and shamans:exploring Middle Stone Age material culture in South Africa. InBecoming human: innovation in prehistoric material and spiritualcultures. Colin Renfrew and Iain Morley, eds. Pp. 29–49. Cam-bridge: Cambridge University Press.

———. 2009b. The “Upper Palaeolithic” of southern Africa: the StillBay and Howiesons Poort techno-traditions. In African genesis:perspectives on hominid evolution. Sally Reynolds and Andrew Gal-lagher, eds. Johannesburg: Witwatersrand University Press.

Henshilwood, Christopher S., Francesco d’Errico, Curtis W. Marean,Richard G. Milo, and Royden Yates. 2001a. An early bone toolindustry from the Middle Stone Age at Blombos Cave, South

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 37: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

396 Current Anthropology Volume 52, Number 3, June 2011

Africa: implications for the origins of modern human behaviour,symbolism and language. Journal of Human Evolution 41(6):631–678.

Henshilwood, Christopher S., Francesco d’Errico, Marian Vanhaeren,Karen van Niekerk, and Zenobia Jacobs. 2004. Middle Stone Ageshell beads from South Africa. Science 304(5669):404.

Henshilwood, Christopher S., Francesco d’Errico, and Ian Watts.2009. Engraved ochres from the Middle Stone Age levels at Blom-bos Cave, South Africa. Journal of Human Evolution 57(1):27–47.

Henshilwood, Christopher S., Francesco d’Errico, Royden Yates, Zen-obia Jacobs, Chantal Tribolo, Geoff A. T. Duller, Norbert Mercier,et al. 2002. Emergence of modern human behavior: Middle StoneAge engravings from South Africa. Science 295(5558):1278–1280.

Henshilwood, Christopher S., and Benoıt Dubreuil. 2009. Readingthe artefacts: gleaning language skills from the Middle Stone Agein southern Africa. In The cradle of language. Rudolf Botha andChris Knight, eds. Pp. 41–60. Oxford: Oxford University Press.

Henshilwood, Christopher S., and Curtis W. Marean. 2003. The or-igin of modern human behavior: a review and critique of themodels and their test implications. Current Anthropology 44(5):627–651.

———. 2006. Remodelling the origins of modern human behaviour:tracing the lineage of modern man. In The prehistory of Africa:tracing the lineage of modern man. Himla Soodyall, ed. Pp. 37–59.Cape Town: Jonathan Ball.

Henshilwood, Christopher S., and Judith Sealy. 1997. Bone artefactsfrom the Middle Stone Age at Blombos Cave, southern Cape, SouthAfrica. Current Anthropology 38(5):890–895.

Henshilwood, Christopher S., Judith C. Sealy, Royden Yates, KathrynCruz-Uribe, Paul Goldberg, Frederick E. Grine, Richard G. Klein,Cedric Poggenpoel, Karen van Niekerk, and Ian Watts. 2001b.Blombos cave, southern Cape, South Africa: preliminary reporton the 1992–1999 excavations of the Middle Stone Age levels.Journal of Archaeological Science 28(4):421–448.

Hewes, Gordon W. 1973. Primate communication and the gesturalorigins of language. Current Anthropology 14(1):5–24. [MCC]

Hoffecker, John F. 2007. Representation and recursion in the ar-chaeological record. Journal of Archaeological Method and Theory14(4):359–387. [DR]

Hollebrandse, Bart, and Tom Roeper. 2007. Recursion and propo-sitional exclusivity. Paper presented at the conference Recursionin Human Language, Normal, IL, April 27–29. [DR]

Holloway, Ralph L., Jr. 1969. Culture: a human domain. CurrentAnthropology 10(4):395–413. [ID]

Holloway, Ralph L., Ronald J. Clark, and Philip V. Tobias. 2004.Posterior lunate sulcus in Australopithecus africanus: was Dartright? Comptes Rendus Palevol 3(4):287–293.

Hurford, Jim. 2006. Diversity in languages, language, genes and sym-bols. Talk presented at the conference The Cradle of Language,Stellenbosch, November 6–10. [PG]

Jacobs, Zenobia, Geoffrey A. T. Duller, and Ann G. Wintle. 2003.Optical dating of dune sand from Blombos Cave, South Africa.II. Single grain data. Journal of Human Evolution 44(5):613–625.

Jacobs, Zenobia, Geoffrey A. T. Duller, Ann G. Wintle, and Chris-topher S. Henshilwood. 2006. Extending the chronology of de-posits at Blombos Cave, South Africa, back to 140 ka using opticaldating of single and multiple grains of quartz. Journal of HumanEvolution 51(3):255–273.

Jacobs, Zenobia, and Richard G. Roberts. 2008. Testing times: oldand new chronologies for the Howieson’s Poort and Still Bayindustries in environmental context. South African ArchaeologicalSociety Goodwin Series 10:8–33.

Jacobs, Zenobia, Richard G. Roberts, Rex F. Galbraith, Hilary J. Dea-con, Rainer Grun, Alex Mackay, Peter Mitchell, Ralf Vogelsang,and Lyn Wadley. 2008. Ages for the Middle Stone Age of southern

Africa: implications for human behavior and dispersal. Science322(5902):733–735.

Jacobs, Zenobia, Ann G. Wintle, and Geoffrey A. T. Duller. 2003.Optical dating of dune sand from Blombos Cave, South Africa: I.Multiple grain data. Journal of Human Evolution 44(5):599–612.

Jolly, Keith. 1948. The development of the Cape Middle Stone Agein the Skildergat Cave, Fish Hoek. South African ArchaeologicalBulletin 3(12):106–107.

Jones, Heather L. 2001. Electron spin resonance dating of toothenamel at three Palaeolithic sites. MSc thesis, McMaster University,Hamilton, Ontario.

Jurado, Marıa B., and Monica Rosselli. 2007. The elusive nature ofexecutive functions: a review of our current understanding. Neu-ropsychology Review 17(3):213–233.

Kaplan, Hillard, and Michael Gurven. 2005. The natural history ofhuman food sharing and cooperation. In Moral sentiments andmaterial interests: the foundations of cooperation in economic life.Herbert Gintis, Samuel Bowles, Robert Boyd, and Ernst Fehr, eds.Pp. 75–113. Cambridge, MA: MIT Press.

Kaplan, Jonathan. 1990. The Umhlatuzana rock shelter sequence:100,000 years of Stone Age history. Natal Museum Journal of Hu-manities 2:1–94.

Kinderman, Peter, Robin Dunbar, and Richard P. Bentall. 1998.Theory-of-mind deficits and causal attributions. British Journal ofPsychology 89(2):191–204.

Klein, Richard G., and Blake Edgar. 2002. The dawn of human culture.New York: Wiley.

Knight, Chris, Camilla Powers, and Ian Watts. 1995. The humansymbolic revolution: a Darwinian account. Cambridge Archaeo-logical Journal 5(1):75–114.

Knoch, Daria, Alvaro Pascual-Leone, Kaspar Meyer, Valerie Treyer,and Ernst Fehr. 2006. Diminishing reciprocal fairness by disruptingthe right prefrontal cortex. Science 314(5800):829–832.

Koechlin, Etienne, and Thomas Jubault. 2006. Broca’s area and thehierarchical organization of human behavior. Neuron 50(6):963–974.

Koechlin, Etienne, and Christopher Summerfield. 2007. An infor-mation theoretical approach to prefrontal executive function.Trends in Cognitive Sciences 11(6):229–235.

Kolen, Jan. 1999. Hominids without homes: on the nature of MiddlePalaeolithic settlement in Europe. In The Middle Palaeolithic oc-cupation of Europe. Wil Roebroeks and Clive Gamble, eds. Pp.139–175. Leiden: University of Leiden.

Koller, Johann, Ursula Baumer, and Dietrich Mania. 2001. High-techin the Middle Palaeolithic: Neandertal-manufactured pitch iden-tified. European Journal of Archaeology 4(3):385–397.

Krause, Johannes, Carles Lalueza-Fox, Ludovic Orlando, WolfgangEnard, Richard E. Green, Hernan A. Burbano, Jean-Jacques Hub-lin, et al. 2007. The derived FOXP2 variant of modern humanswas shared with Neandertals. Current Biology 17(21):1908–1912.

Lalueza-Fox, Carles, Antonio Rosas, Almudena Estalrrich, ElenaGigli, Paula F. Campos, Antonio Garcıa-Tabernero, Samuel Garcıa-Vargas, et al. 2011. Genetic evidence for patrilocal mating behavioramong Neandertal groups. Proceedings of the National Academy ofSciences of the USA 108(1):250–253. [LW]

Laumann, Anne E., and Amy J. Derick. 2006. Tattoos and bodypiercings in the United States: a national data set. Journal of theAmerican Academy of Dermatology 55(3):413–421.

Lehmann, Julia, Amanda H. Korstjens, and R. I. M. Dunbar. 2007.Group size, grooming and social cohesion in primates. AnimalBehaviour 74(6):1617–1629.

Lewis-Williams, J. David, and David G. Pearce. 2004. San spirituality:roots, expression, and social consequences. Walnut Creek, CA: Al-tamira.

Lieberman, Daniel E. 2008. Speculations about the selective basis for

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 38: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 397

modern human craniofacial form. Evolutionary Anthropology17(1):55–68.

Lieberman, Daniel E., Brandeis M. McBratney, and Gail Krovitz.2002. The evolution and development of cranial form in Homosapiens. Proceedings of the National Academy of Sciences of the USA99(3):1134–1139.

Lieberman, Philip. 2005. The pied piper of Cambridge. LinguisticReview 22:289–301.

———. 2007. The evolution of human speech: its anatomical andneural bases. Current Anthropology 48(1):39–66.

Lombard, Marlize. 2005a. Evidence of hunting and hafting duringthe Middle Stone Age at Sibidu Cave, KwaZulu-Natal, South Af-rica: a multianalytical approach. Journal of Human Evolution 48(3):279–300.

———. 2005b. The Howiesons Poort of South Africa: what we know,what we think we know, what we need to know. South AfricanHumanities 17(Dec):33–55.

———. 2006a. Direct evidence for the use of ochre in the haftingtechnology of Middle Stone Age tools from Sibudu Cave. SouthernAfrican Humanities 18(1):57–67.

———. 2006b. First impressions of the functions and hafting tech-nology of Still Bay pointed artefacts from Sibudu Cave. SouthernAfrican Humanities 18(1):27–41.

———. 2007a. Evidence for change in Middle Stone Age huntingbehaviour at Blombos Cave: results of a macrofracture analysis.South African Archaeological Bulletin 62(185):62–67.

———. 2007b. The gripping nature of ochre: the association of ochrewith Howiesons Poort adhesives and Later Stone Age mastics fromSouth Africa. Journal of Human Evolution 53(4):406–419.

Lombard, Marlize, and Laurel Phillipson. 2010. Indications of bowand stone-tipped arrow use 64,000 years ago in KwaZulu-Natal,South Africa. Antiquity 84(325):635–648.

Lombard, Marlize, Lyn Wadley, Zenobia Jacobs, Moleboheng Mo-hapi, and Richard G. Roberts. 2010. Still Bay and serrated pointsfrom Umhlatuzana rock shelter, KwaZulu-Natal, South Africa.Journal of Archaeological Science 37(7):1773–1784.

Mackay, Alex, and Aara Welz. 2008. Engraved ochre from a MiddleStone Age context at Klein Kliphuis in the Western Cape of SouthAfrica. Journal of Archaeological Science 35(6):1521–1532.

Malafouris, Lambros. 2007. Before and beyond representation: to-wards an enactive conception of the Palaeolithic image. In Imageand imagination: a global history of figurative representation. ColinRenfrew and Iain Morley, eds. Pp. 289–302. Cambridge: McDonaldInstitute for Archaeological Research. [LM]

———. 2008. Beads for a plastic mind: the “blind man’s stick” (BMS)hypothesis and the active nature of material culture. CambridgeArchaeological Journal 18(3):401–414.

———. 2010a. The brain-artefact interface (BAI): a challenge forarchaeology and cultural neuroscience. Social Cognitive and Affec-tive Neuroscience 5(2–3):264–273. [LM]

———. 2010b. Grasping the concept of number: how did the sapientmind move beyond approximation? In The archaeology of mea-surement: comprehending heaven, earth and time in ancient societies.Iain Morley and Colin Renfrew, eds. Pp. 35–42. Cambridge: Cam-bridge University Press. [LM]

———. 2010c. Metaplasticity and the human becoming: principlesof neuroarchaeology. Journal of Anthropological Sciences 88:49–72.[LM]

Malafouris, Lambros, and Colin Renfrew, eds. 2010. The cognitivelife of things: recasting the boundaries of the mind. Cambridge:McDonald Institute for Archaeological Research. [KRG]

Malan, Berry D. 1955. The archaeology of Tunnel Cave and SkildergatKop, Fish Hoek, Cape of Good Hope. South African ArchaeologyBulletin 10(37):3–9.

Marcus, Gary F. 2009. How does the mind work? insights frombiology. Topics in Cognitive Science 1(1):145–172.

Marean, Curtis W., and Zelalem Assefa. 2005. The Middle and UpperPleistocene African record for the biological and behavioral originsof modern humans. In African archaeology: a critical introduction.Ann B. Stahl, ed. Pp. 93–129. Malden, MA: Blackwell

Marean, Curtis W., Miryam Bar-Matthews, Jocelyn Bernatchez, ErichFisher, Paul Goldberg, Andy I. R. Herries, Zenobia Jacobs, et al.2007. Early human use of marine resources and pigment in SouthAfrica during the Middle Pleistocene. Nature 449(7164):905–908.

McBrearty, Sally, and Alison Brooks. 2000. The revolution that wasn’t:a new interpretation of the origin of modern human behavior.Journal of Human Evolution 39(5):453–563.

Mellars, Paul. 1996. The Neanderthal legacy: an archaeological per-spective from western Europe. Princeton, NJ: Princeton UniversityPress.

———. 2005. The impossible coincidence: a single-species modelfor the origins of modern human behaviour in Europe. Evolu-tionary Anthropology 14(1):12–27.

———. 2006. Why did modern human populations disperse fromAfrica ca. 60,000 years ago? a new model. Proceeding of the NationalAcademy of Sciences of the USA 103(25):9381–9386.

Minichillo, Thomas J. 2005. Middle Stone Age lithic study, SouthAfrica: an examination of modern human origins. PhD disserta-tion, University of Washington, Seattle.

Mitchell, Jason P. 2008. Activity in right temporo-parietal junctionis not selective for theory-of-mind. Cerebral Cortex 18(2):262–271.

———. 2009. Inferences about mental states. Philosophical Trans-actions of the Royal Society B: Biological Sciences 364(1521):1309–1316.

Mithen, Steven. 1996. The prehistory of the mind. London: Thames& Hudson.

———. 2006. The singing Neanderthals: the origins of music, language,mind, and body. Cambridge, MA: Harvard University Press.

Nevins, Andrew, David Pesetsky, and Cilene Rodrigues. 2009. Pirahaexceptionality: a reassessment. Language 85(2):355–404.

Noack, Raymond A. 2006. The frontal feedback model of the evo-lution of the human mind. Part 1: the “pre”-human brain andthe perception-action cycle. Journal of Mind and Behavior 27(3–4):215–242.

———. 2007. The frontal feedback model of the evolution of thehuman mind. Part 2: the human brain and the frontal feedbacksystem. Journal of Mind and Behavior 28(3–4):233–264.

Noble, William, and Iain Davidson. 1991. The evolutionary emer-gence of modern human behaviour: language and its archaeology.Man 26(2):223–253. [ID]

———. 1996. Human evolution, language and mind. Cambridge:Cambridge University Press.

Nowak, Martin, and Karl Sigmund. 2005. Evolution of indirect reci-procity. Nature 437(7063):1291–1298. [PG]

Nowell, April, and Mark J. White. 2010. Growing up in the MiddlePleistocene: life history strategies and their relationship to Acheu-lean industries. In Stone tools and the evolution of human cognition.April Nowell and Iain Davidson, eds. Pp. 67–82. Boulder: Uni-versity Press of Colorado. [ID]

Overmann, Karenleigh A., Thomas Wynn, and Frederick L. Coolidge.Forthcoming. The prehistory of number concept. Behavioral andBrain Sciences. [FLC/TW]

Parkington, John. 1999. Western Cape landscapes. In World prehis-tory: studies in memory of Grahame Clark. John Coles, RobertBewley, and Paul Mellars, eds. Pp. 25–35. Proceedings of the BritishAcademy 99. Oxford: Oxford University Press.

Parkington, John, Cedric Poggenpoel, Jean-Philippe Rigaud, andPierre-Jean Texier. 2005. From tool to symbol: the behaviouralcontext of intentionally marked ostrich eggshell from Diepkloof,Western Cape. In From tools to symbols: from early hominids tomodern humans. Francesco d’Errico and Lucinda Backwell, eds.Pp. 475–492. Johannesburg: Witwatersrand University Press.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 39: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

398 Current Anthropology Volume 52, Number 3, June 2011

Pearson, Osbjorn M. 2008. Statistical and biological definitions ofanatomically modern humans: suggestions for a unified approachto modern morphology. Evolutionary Anthropology 17(1):38–48.

Peers, Bertie. 1929. Preliminary report on the archaeology of FishHoek–Noord Hoek valley. Unpublished manuscript in the collec-tions of the South African Museum, Cape Town.

Peers, Bertie, and A. J. H. Goodwin. 1953. Peers Shelter B/102. SouthAfrican Archaeological Bulletin 8(31):67–77.

Peirce, Charles Sanders. 1978 (1931). The collected papers of CharlesSanders Peirce, vol 2. Charles Hartshorne and Paul Weiss, eds.Cambridge, MA: Harvard University Press.

Perner, Josef, and Ted Ruffman. 1995. Episodic memory and auto-noetic consciousness: developmental evidence and a theory ofchildhood amnesia. Journal of Experimental Child Psychology 59(3):516–548.

Perner, Josef, Sandra Stummer, Manuel Sprung, and Martin Doherty.2002. Theory of mind finds its Piagetian perspective: why alter-native naming comes with understanding belief. Cognitive Devel-opment 17(3–4):1451–1472.

Pinker, Steven, and Paul Bloom. 1990. Natural language and naturalselection. Behavioral and Brain Sciences 13(4):707–784. [MCC]

Potts, Richard. 1996. Humanity’s descent: the consequences of ecologicalinstability. New York: Morrow

Powell, Adam, Stephen Shennan, and Mark G. Thomas. 2009. LatePleistocene demography and the appearance of modern humanbehavior. Science 324(5932):1298–1301. [DR]

Read, Dwight W. 2001. What is kinship? In The cultural analysis ofkinship: the legacy of David Schneider and its implications for an-thropological relativism. Richard Feinberg and Martin Ottenheimer,eds. Pp. 78–117. Urbana: University of Illinois Press. [DR]

———. 2006. Tasmanian knowledge and skill: maladaptive imitationor adequate technology? American Antiquity 71(1):164–184. [DR]

———. 2007. Kinship theory: a paradigm shift. Ethnology 46(4):329–364. [DR]

———. 2008a. An interaction model for resource implement com-plexity based on risk and number of annual moves. AmericanAntiquity 73(4):599–625. [DR]

———. 2008b. Working memory: a cognitive limit to non-humanprimate recursive thinking prior to hominid evolution. Evolution-ary Psychology 6(4):676–714.

———. 2011. The misuse of a mathematical model: the Tasmaniancase (reply to Henrich’s response). Human Complex Systems,UCLA. http://escholarship.org/uc/item/88k8g4rj. [DR]

Read, Dwight, and Sander van der Leeuw. 2008. Biology is only partof the story . . . . Philosophical Transactions of the Royal Society B:Biological Sciences 363(1499):1959–1968. [DR]

Renfrew, Colin. 1996. The sapient behavior paradox: how to test forpotential. In Modelling the early human mind. Paul Mellars andKathleen Gibson, eds. Pp. 11–14. Cambridge: McDonald Institutefor Archaeological Research. [KRG]

Richerson, Peter J., and Robert Boyd. 1998. The Pleistocene and theorigins of human culture: built for speed. Paper presented at the5th Biannual Symposium on the Science of Behaviour: Behaviour,Evolution and Culture, Guadalajara, February 22–24.

Rigaud, Jean-Philippe, Pierre-Jean Texier, John Parkington, and Ced-ric Poggenpoel. 2006. Le mobilier Stillbay et Howiesons Poort del’abri Diepkloof: la chronologie du Middle Stone Age sud-africainet ses implications. Comptes Rendus Palevol 5(6):839–849. [In-cludes abridged English version.]

Rightmire, G. Philip. 2004. Brain size and encephalization in Earlyto Mid-Pleistocene Homo. American Journal of Physical Anthro-pology 124(2):109–123.

———. 2008. Homo in the Middle Pleistocene: hypodigms, variation,and species recognition. Evolutionary Anthropology 17(1):8–21.

Rilling, James K. 2006. Human and nonhuman primate brains: are

they allometrically scaled versions of the same design? EvolutionaryAnthropology 15(2):65–77.

Rilling, James K., Matthew F. Glasser, Todd M. Preuss, XiangyangMa, Tiejun Zhao, Xiaoping Hu, and Timothy E. J. Behrens. 2008.The evolution of the arcuate fasciculus revealed with comparativeDTI. Nature Neuroscience 11(4):426–428.

Rilling, James K., and Thomas R. Insel. 1999. The primate neocortexin comparative perspective using magnetic resonance imaging.Journal of Human Evolution 37(2):191–223.

Rilling, James K., and Rebecca Seligman. 2002. A quantitative mor-phometric comparative analysis of the primate temporal lobe.Journal of Human Evolution 42(5):505–534.

Rizzolatti, Giacomo, and Corrado Sinigaglia. 2008. Mirrors in thebrain. Frances Anderson, trans. Oxford: Oxford University Press.[MCC]

Robson, Shannen L., and Bernard Wood. 2008. Hominin life history:reconstruction and evolution. Journal of Anatomy 212(4):394–425.

Rossano, Matt J. 2007. Did meditating make us human? CambridgeArchaeological Journal 17(1):47–58.

Sampson, C. Garth. 1974. The Stone Age archaeology of southernAfrica. New York: Academic Press.

Sanfey, Alan G., James K. Rilling, Jessica A. Aronson, Leigh E. Ny-strom, and Jonathan D. Cohen. 2003. The neural basis of economicdecisionmaking in the Ultimatum Game. Science 300(5626):1755–1758.

Savage-Rumbaugh, Sue, Stuart G. Shanker, and Talbot J. Taylor. 1998.Apes, language, and the human mind. New York: Oxford UniversityPress.

Saxe, Rebecca, and Nancy Kanwisher. 2003. People thinking aboutthinking people: the role of the temporoparietal junction in theoryof mind. NeuroImage 19(4):1835–1842.

Saxe, Rebecca, and Lindsey J. Powell. 2006. It’s the thought thatcounts: specific brain regions for one component of theory ofmind. Psychological Science 17(8):692–699.

Schenker, Natalie M., Anne-Marie Desgouttes, and Katerina Semen-deferi. 2005. Neural connectivity and cortical substrates of cog-nition in hominoids. Journal of Human Evolution 49(5):547–569.

Schirmer, G. R. 1975. An analysis of lithic material from Dale RoseParlour, Trappies Kop, Kalk Bay, Cape Peninsula. Archaeologyadditional report, Cape Town: University of Cape Town.

Schoenemann, P. Thomas, Michael J. Sheehan, and L. Daniel Glotzer.2005. Prefrontal white matter volume is disproportionately largerin humans than in other primates. Nature Neuroscience 8(2):242–252.

Scholz, Christopher A., Thomas C. Johnson, Andrew S. Cohen, JohnW. King, John A. Peck, Jonathan T. Overpeck, Michael R. Talbot,et al. 2007. East African megadroughts between 135 and 75 thou-sand years ago and bearing on early-modern human origins. Pro-ceedings of the National Academy of Sciences of the USA 104(42):16416–16421.

Scholz, Jonathan, Christina Triantafyllou, Susan Whitfield-Gabrieli,Emery N. Brown, and Rebecca Saxe. 2009. Distinct regions of righttemporo-parietal junction are selective for theory of mind andexogenous attention. PLoS ONE 4(3):e4869.

Searle, John R. 1995. The construction of social reality. New York: FreePress.

Semendeferi, Katerina, Este Armstrong, Axel Schleicher, Karl Zilles,and Gary W. Van Hoesen. 2001. Prefrontal cortex in humans andapes: a comparative study of area 10. American Journal of PhysicalAnthropology 114(3):224–241.

Semendeferi, Katerina, and Hanna Damasio. 2000. The brain andits main anatomical subdivisions in living hominoids using mag-netic resonance imaging. Journal of Human Evolution 38(2):317–332.

Semendeferi, Katerina, A. Lu, Natalie Schenker, and Hanna Damasio.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 40: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

Henshilwood and Dubreuil Still Bay and Howiesons Poort in the Middle Stone Age 399

2002. Humans and great apes share a large frontal cortex. NatureNeuroscience 5(3):272–276.

Shennan, Stephen. 2002. Genes, memes and human history: Darwinianarchaeology and cultural evolution. London: Thames & Hudson.

Sherwood, Chet C., Francys Subiaul, and Tadeusz W. Zawidzki. 2008.A natural history of the human mind: tracing evolutionary changesin brain and cognition. Journal of Anatomy 212(4):426–454.

Singer, Ronald, and John Wymer. 1982. The Middle Stone Age atKlasies River mouth in South Africa. London: University of ChicagoPress.

Soriano, Sylvain, Paola Villa, and Lyn Wadley. 2007. Blade technologyand tool forms in the Middle Stone Age of South Africa: theHowiesons Poort and post-Howiesons Poort at Rose Cottage Cave.Journal of Archaeological Science 34(5):681–703.

Stiller, James, and R. I. M. Dunbar. 2007. Perspective-taking andmemory capacity predict social network size. Social Networks 29(1):93–104.

Stone, Valerie E., and Philip Gerrans. 2006. What’s domain-specificabout theory of mind? Social Neuroscience 1(3–4):309–319.

Stout, Dietrich, and Thierry Chaminade. 2009. Making tools andmaking sense: complex, intentional behaviour in human evolution.Cambridge Archaeological Journal 19(1):85–96.

Suddendorf, Thomas, and Michael Corballis. 2007. The evolution offoresight: what is mental time travel and is it unique to humans?Behavioral and Brain Sciences 30(3):299–313.

Tanaka, Jiro. 1982. Adaptation to arid environment: a comparativestudy of hunter-gatherers and pastoralists in Africa. African StudyMonographs Supplementary Issue 1:1–12.

Texier, Pierre-Jean, Guillaume Porraz, John Parkington, Jean-Philippe Rigaud, Cedric Poggenpoel, Christopher Miller, ChantalTribolo, et al. 2010. A Howiesons Poort tradition of engravingostrich eggshell containers dated to 60,000 years ago at DiepkloofRock Shelter, South Africa. Proceedings of the National Academyof Sciences of the USA 107(14):6180–6185.

Thackeray, Anne I. 1992. The Middle Stone Age south of the LimpopoRiver. Journal of World Prehistory 6(4):385–440.

Thackeray, J. Francis. 1992. Chronology of Late Pleistocene depositsassociated with Homo sapiens at Klasies River Mouth, South Africa.Palaeoecology of Africa 23:177–191.

———. 2007. Sea levels and chronology of Late Pleistocene coastalcave deposits at Klasies River in South Africa. Annals of the Trans-vaal Museum 44:219–220.

Thieme, Hartmu. 1999. Altpalaolithische Holzgerate aus Schoningen,lkr. Helmstedt. Germania 77(2):451–487. [PG]

Tomasello, Michael. 2003. Constructing a language. Cambridge, MA:Harvard University Press.

———. 2008. Origins of human communication. Cambridge, MA:MIT Press.

Tooby, John, and Irven DeVore. 1987. The reconstruction of hominidbehavioral evolution through strategic modeling. In The evolutionof human behavior: primate models. Warren G. Kinzey, ed. Pp. 183–237. Stony Brook: SUNY Press. [MCC]

Tribolo, Chantal. 2003. Apports des methodes de la luminescence ala chronologie de techno-facies du Middle Stone Age associes auxpremiers hommes modernes d’Afrique du Sud. PhD thesis, Uni-versite Bordeaux-1.

Tribolo, Chantal, Norbert Mercier, Madeleine Selo, Helene Valladas,Jean-Louis Joron, Jean-Louis Reyss, Christopher Henshilwood, Ju-dith Sealy, and Royden Yates. 2006. TL dating of burnt lithics fromBlombos Cave (South Africa) and the antiquity of modern be-haviour. Archaeometry 48(2):341–357.

Tribolo, Chantal, Norbert Mercier, and Helene Valladas. 2005.Chronologie des technofacies Howiesons Poort et Still Bay (MiddleStone Age, Afrique du Sud): bilan et nouvelles donnees de laluminescence. Bulletin de la Societe Prehistorique Francais 102(4):855–866.

Tribolo, Chantal, Norbert Mercier, Helene Valladas, Jean-Louis Joron,Pierre Guibert, Yannick Lefrais, Madeleine Selo, et al. 2009. Ther-moluminescence dating of a Stillbay–Howiesons Poort sequenceat Diepkloof Rock Shelter (Western Cape, South Africa). Journalof Archaeological Science 36(3):730–739.

Tulving, Endel. 1972. Episodic and semantic memory. In Organi-zation of memory. Endel Tulving and Wayne Donaldson, eds. Pp.381–403. New York: Academic Press.

———. 2002. Episodic memory: from mind to brain. Annual Reviewof Psychology 53:1–25.

Tunik, Eugene, Nichola J. Rice, Antonia Hamilton, and Scott T. Graf-ton. 2007. Beyond grasping: representation of action in humananterior intraparietal sulcus. NeuroImage 36(suppl. 2):T77–T86.

Ungerleider, Leslie G., Susan M. Courtney, and James V. Haxby. 1998.A neural system for human visual working memory. Proceedingsof the National Academy of Sciences of the USA 95(3):883–890.

Valladas, Helene, Lyn Wadley, Norbert Mercier, Laurence Froget,Chantal Tribolo, Jean-Louis Reyss, and Jean-Louis Joron. 2005.Thermoluminescence dating on burnt lithics from Middle StoneAge layers at Rose Cottage Cave. South African Journal of Science101(3–4):169–174.

Vanhaeren, Marian, Francesco d’Errico, Chris Stringer, Sarah L.James, Jonathan A. Todd, and Henk K. Mienis. 2006. Middle Pa-leolithic shell beads in Israel and Algeria. Science 312(5781):1785–1788.

Velo, Joseph. 1984. Ochre as medicine: a suggestion for the inter-pretation of the archaeological record. Current Anthropology 25(5):674.

Villa, Paola, Marie Soressi, Christopher S. Henshilwood, and VincentMourre. 2009. The Still Bay points of Blombos Cave (South Af-rica). Journal of Archaeological Science 36(2):441–460.

Villa, Paola, Sylvain Soriano, Nicolas Teyssandier, and Sarah Wurz.2010. The Howiesons Poort and MSA III at Klasies River MainSite, Cave 1A. Journal of Archaeological Science 37(3):630–655.

Vogel, John C. 2001. Radiometric dates for the Middle Stone Age inSouth Africa. In Humanity from African naissance to coming mil-lennia. Phillip V. Tobias, Michael A. Raath, Jacopo Moggi-Cecchi,and Gerald A. Doyle, eds. Pp. 261–268. Florence: Florence Uni-versity Press.

Vogelsang, Ralf. 1998. Middle-Stone-Age Fundstellen in Sudwest-Namibia. Africa praehistorica 11. Koln: Heinrich-Barth-Intitut.

Vogelsang, Ralf, Jurgen Richter, Zenobia Jacobs, Barbara Eichhorn,Veerle Linseele, and Richard G. Roberts. 2010. New excavationsof Middle Stone Age deposits at Apollo 11 rockshelter, Namibia:stratigraphy, archaeology, chronology and past environments. Jour-nal of African Archaeology 8(2):185–218.

Volman, Thomas P. 1981. The Middle Stone Age of the southernCape. PhD dissertation, University of Chicago.

Wadley, Lyn. 2001a. Preliminary report on excavations at SibuduCave, Kwazulu-Natal. Southern African Humanities 13:1–17.

———. 2001b. What is cultural modernity? a general view and aSouth African perspective from Rose Cottage Cave. CambridgeArchaeological Journal 11(2):201–221.

———. 2005a. Putting ochre to the test: replication studies of ad-hesives that may have been used for hafting tools in the MiddleStone Age. Journal of Human Evolution 49(5):587–601.

———. 2005b. A typological study of the final Middle Stone Agestone tools from Sibudu Cave, KwaZulu-Natal. South African Ar-chaeological Bulletin 60(182):51–63.

———. 2006. The use of space in the Late Middle Stone Age ofRose Cottage Cave, South Africa. In Transitions before the transi-tions: evolution and stability in the Middle Paleolithic and MiddleStone Age. Erella Hovers and Steven L. Kuhn, eds. Pp. 279–294.New York: Springer.

———. 2007. Announcing a Still Bay industry at Sibudu Cave, SouthAfrica. Journal of Human Evolution 52(6):681–689.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions

Page 41: The Still Bay and Howiesons Poort, 7759 ka: Symbolic ... · The Still Bay and Howiesons Poort, 77–59 ka: Symbolic Material Culture and the Evolution of the Mind during the African

400 Current Anthropology Volume 52, Number 3, June 2011

———. 2010. Compound-adhesive manufacture as a behavioralproxy for complex cognition in the Middle Stone Age. CurrentAnthropology 51(suppl. 1):S111–S119. [LW]

Wadley, Lyn, Tamaryn Hodgskiss, and Michael Grant. 2009. Impli-cations for complex cognition from the hafting of tools with com-pound adhesives in the Middle Stone Age, South Africa. Proceed-ings of the National Academy of Sciences of the USA 106(24):9590–9594.

Wadley, Lyn, and Zenobia Jacobs. 2004. Sibudu Cave, KwaZulu-Natal: background to the excavations of middle stone age and ironage occupations. South African Journal of Science 100(3–4):145–151.

———. 2006. Sibudu Cave: background to the excavations, stratig-raphy and dating. Southern African Humanities 18:1–26.

Wadley, Lyn, and Moleboheng Mohapi. 2008. A segment is not amonolith: evidence from the Howiesons Poort of Sibudu, SouthAfrica. Journal of Archaeological Science 35(9):2594–2605.

Wadley, Lyn, Bonny Williamson, and Marlize Lombard. 2004. Ochrein hafting in Middle Stone Age southern Africa: a practical role.Antiquity 78(301):661–675.

Watts, Ian. 1999. The origin of symbolic culture. In The evolution ofculture: an interdisciplinary view. Robin Dunbar, Chris Knight, andCamilla Power, eds. Pp. 113–145. Edinburgh: Edinburgh CityPress.

———. 2002. Ochre in the Middle Stone Age of southern Africa:ritualised display or hide preservative? South African ArchaeologicalBulletin 57(175):1–14.

———. 2009. Red ochre, body painting, and language: interpretingthe Blombos ochre. In The cradle of language. Rudolf Botha andChris Knight, eds. Pp. 62–92. Oxford: Oxford University Press.

Welshon, Rex. 2010. Working memory, neuroanatomy, and archae-ology. Current Anthropology 51(suppl. 1):S191–S199. [LW]

White, Tim D., Berhane Asfaw, David DeGusta, Henry Gilbert, GaryD. Richards, Gen Suwa, and F. Clark Howell. 2003. PleistoceneHomo sapiens from Middle Awash, Ethiopia. Nature 423(6941):742–747.

Wiessner, Polly. 1983. Style and social information in Kalahari Sanprojectile points. American Antiquity 48(2):253–276.

Wilkins, Wendy K. 2009. Mosaic neurobiology and anatomical plau-sibility. In The prehistory of language. Rudolf Botha and ChrisKnight, eds. Pp. 266–285. Oxford: Oxford University Press.

Wilkins, Wendy K., and Jennie Wakefield. 1995. Brain evolution andneurolinguistic preconditions. Behavioral and Brain Sciences 18(1):161–182. [ID]

Wood, Bernard, and Mark Collard. 1999. The human genus. Science284(5411):65–71. [MCC]

Wurz, Sarah. 1999. The Howiesons Poort backed artefacts from Kla-sies River: an argument for symbolic behaviour. South AfricanArchaeological Bulletin 54(169):38–50.

———. 2000. The Middle Stone Age at Klasies River, South Africa.PhD thesis, University of Stellenbosch.

Wurz, Sarah, Niel J. le Roux, Sugnet Gardner, and Hilary J. Deacon.2003. Discriminating between the end products of the earlier Mid-dle Stone Age sub-stages at Klasies River using biplot. Journal ofArchaeological Science 30(9):1107–1126.

Wynn, Thomas. 2009. Hafted spears and the archaeology of mind.Proceedings of the National Academy of Sciences of the USA 106(24):9544–9545. [FLC/TW]

Wynn, Thomas, and Frederick L. Coolidge. 2007. Did a small butsignificant enhancement in working memory capacity power theevolution of modern thinking? In Rethinking the human revolution:new behavioural and biological perspectives on the origins and dis-persal of modern humans. Paul Mellars, Katie Boyle, Ofer Bar-Yosef,and Chris Stringer, eds. Pp. 79–90. Cambridge: McDonald Institutefor Archaeological Research.

———. 2009. Implications of a strict standard for recognizing mod-ern cognition in prehistory. In Cognitive archaeology and humanevolution. Sophie A. de Beaune, Frederick L. Coolidge, and ThomasWynn, eds. Pp. 117–127. Cambridge: Cambridge University Press.[LW]

Wynn, Thomas, Frederick Coolidge, and Martha Bright. 2009.Hohlenstein-Stadel and the evolution of human conceptualthought. Cambridge Archaeological Journal 19(1):73–84.

Wynn, Thomas, Frederick L. Coolidge, and Karenleigh A. Overmann.Forthcoming. The archaeology of number concept and its impli-cations for the evolution of language. In EvoLang 8. Rudolf Botha,ed. Oxford: Oxford University Press. [FLC/TW]

Young, Liane, Fiery Cushman, Marc Hauser, and Rebecca Saxe. 2007.The neural basis of the interaction between theory of mind andmoral judgment. Proceedings of the National Academy of Sciencesof the USA 104(20):8235–8240.

Zahavi, Amotz. 1975. Mate selection: a selection for a handicap.Journal of Theoretical Biology 53(1):205–214. [PG]

Zhou, Xin, and Bin Wang. 2004. Preschool children’s representationand understanding of written number symbols. Early Child De-velopment and Care 174(3):253–266.

This content downloaded on Fri, 14 Dec 2012 19:12:06 PMAll use subject to JSTOR Terms and Conditions