Social learning and innovation in contemporary hunter gatherers

313 294 0
Social learning and innovation in contemporary hunter gatherers

Đang tải... (xem toàn văn)

Tài liệu hạn chế xem trước, để xem đầy đủ mời bạn chọn Tải xuống

Thông tin tài liệu

Replacement of Neanderthals by Modern Humans Series Hideaki Terashima Barry S.  Hewlett Editors Social Learning and Innovation in Contemporary Hunter-Gatherers Evolutionary and Ethnographic Perspectives Replacement of Neanderthals by Modern Humans Series Edited by Takeru Akazawa Research Institute, Kochi University of Technology Kochi 782-8502, Japan akazawa.takeru@kochi-tech.ac.jp Ofer Bar-Yosef Department of Anthropology, Harvard University Cambridge, Massachusetts 02138, USA obaryos@fas.harvard.edu The planned series of volumes will report the results of a major research project entitled “Replacement of Neanderthals by Modern Humans: Testing Evolutionary Models of Learning”, offering new perspectives on the process of replacement and on interactions between Neanderthals and modern humans and hence on the origins of prehistoric modern cultures The projected volumes will present the diverse achievements of research activities, originally designed to implement the project’s strategy, in the fi elds of archaeology, paleoanthropology, cultural anthropology, population biology, earth sciences, developmental psychology, biomechanics, and neuroscience Comprehensive research models will be used to integrate the discipline-specifi c research outcomes from those various perspectives The series, aimed mainly at providing a set of multidisciplinary perspectives united under the overarching concept of learning strategies, will include monographs and edited collections of papers focusing on specifi c problems related to the goals of the project, employing a variety of approaches to the analysis of the newly acquired data sets Editorial Board Stanley H Ambrose (University of Illinois at Urbana-Champaign), Kenichi Aoki (Meiji University), Emiliano Bruner (Centro National de Investigacion Sobre la Evolution Humana), Marcus W Feldman (Stanford University), Barry S Hewlett (Washinton State University), Tasuku Kimura (University of Tokyo), Steven L Kuhn (University of Arizona), Yoshihiro Nishiaki (University of Tokyo), Naomichi Ogihara (Keio University), Dietrich Stout (Emory University), Hiroki C Tanabe (Nagoya University), Hideaki Terashima (Kobe Gakuin University), Minoru Yoneda (University of Tokyo) More information about this series at http://www.springer.com/series/11816 Hideaki Terashima • Barry S Hewlett Editors Social Learning and Innovation in Contemporary Hunter-Gatherers Evolutionary and Ethnographic Perspectives Editors Hideaki Terashima Faculty of Humanities and Sciences Kobe Gakuin University Kobe Japan Barry S Hewlett Department of Anthropology Washington State University Vancouver Washington USA Replacement of Neanderthals by Modern Humans Series ISBN 978-4-431-55995-5 ISBN 978-4-431-55997-9 DOI 10.1007/978-4-431-55997-9 (eBook) Library of Congress Control Number: 2016953113 # Springer Japan 2016 This work is subject to copyright All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed The use of general descriptive names, registered names, trademarks, service marks, etc in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication Neither the publisher nor the authors or the editors give a warranty, express or implied, with respect to the material contained herein or for any errors or omissions that may have been made Printed on acid-free paper This Springer imprint is published by Springer Nature The registered company is Springer Japan KK Photo Gallery Aka in Central African Republic (photos by Barry Hewlett & Bonnie Hewlett) Photo Gallery Baka in the Republic of Cameroon (photos by H Terashima and N Kamei) Photo Gallery Inuit, Yolngu, and San (photos by K Omura, S Kubota, and K Imamura) Preface The RNMH Project and the Study of Social Learning in Modern Hunter–Gatherers An interdisciplinary 5-year project entitled “Replacement of Neanderthals by Modern Humans: Testing Evolutionary Models of Learning” (RNMH) was carried out from 2010 to 2015 and funded by Japan’s Ministry of Education, Culture, Sports, Science, and Technology (Grant-in-Aid for Scientific Research on Innovative Areas, Grant No 22101001) With a team of social-cultural and biological anthropologists, developmental and clinical psychologists, a behavioral geneticist, and primatologists to contribute to the project, we investigated the learning abilities and behavior of contemporary hunter–gatherers living in various environments Marvelous developments in genetics in recent years have revealed that modern humans (Homo sapiens, hereafter referred to simply as “Sapiens”) originated in Africa around 200 ka (thousand years ago), then around 100 ka they began to spread out of Africa and into Eurasia They arrived in central and western Europe by 45–47 ka and it was there they came into contact with Neanderthals The Neanderthals were a highly advanced human species supposed to have evolved from Homo heidelbergensis (also supposed to be the ancestor of Sapiens) They thrived in Europe for about 300,000 years and adapted to the cold weather during the glacial epoch However, they appear to have disappeared by ca 40 ka, 5,000–7,000 years after the appearance of Sapiens on the continent There remains an intriguing mystery: why and how did the Neanderthals go extinct and Sapiens survive? What determined the fates of the two advanced hominins? Many researchers have been studying this problem for decades and exchanging heated debates on the possible causes of the demise of Neanderthals, but no decisive conclusion has yet been reached When considering the characteristics of modern humans, we usually think of our advanced cognitive capacity—highly flexible and capable of symbolic thought and language Working memory and the executive function of the human brain have been garnering particular attention recently Thus, one of the simplest scenarios of the replacement might be that the Sapiens out-competed Neanderthals due to the advantage of cognitive superiority, perhaps allowing greater breadth and efficiency in hunting in gathering or advantages in interspecies in combat, although there is no substantive evidence of violent confrontation or battle between the two populations In any case, the development of higher cognitive abilities has doubtlessly contributed to the success of modern humans, but there seems to be little evidence to justify the assumption of a sudden increase in our cognitive abilities and advances in brain function, including language use, at the time of the replacement Because the replacement in Europe seems to have happened so rapidly, it is doubtful that these cognitive advances occurred at that time From the standpoint of neurobiology and population genetics, it would be very difficult or impossible for such significant differences in cognition to evolve in the span of just 5000–7000 years and permeate the entire Sapiens population Rather, the rapidity of the replacement suggests ix x that the differences in Sapiens’ cognition evolved earlier, probably before they left Africa Higher cognitive capacity would had to have evolved prior to its expression in the development of tangible innovations such as new lithic industries, efficient subsistence strategies, and flexible and effective social organization There have been many factors proposed so far by researchers regarding the differences between the two populations, such as their physical, social, and other adaptive capacities in addition to the cognitive abilities mentioned above Those factors include differences in average body size and musculature, energy expenditure, birthrate and mortality, demographic patterns, subsistence systems, child development patterns, material culture such as clothing and stone tool technologies, behavioral adaptations to variable environmental conditions, movement of game animals, and social structures All of those factors influenced the competition for survival to various degrees, but it is difficult to point out any one or combination of these as the primary catalyst(s) for the replacement The RNMH project focuses instead on differences in the two species’ capacities for learning, particularly social learning and innovative learning, to address the replacement problem This approach is more parsimonious because learning abilities account for many of the possible differences listed above Knowledge about how to construct and use effective clothing and tools in various environments, for example, results from the accumulation of technical and ecological know-how gathered over multiple generations Learning and the social behavior that supports learning are the most important factors in the foundation of the human capacity to develop cultural adaptations for survival in various types of environments and ecologies RNMH proposes a hypothesis called the “learning hypothesis” that suggests there were innate differences in learning ability between Neanderthals and Sapiens that might have divided the fates of the two populations About 2.5 million years ago, a hominin group known as Homo habilis began to make stone tools in Africa It was the beginning of lithic technology and the distinctive cultural development of our human ancestors, and since then culture has become the keystone of human adaptation not only in the area of technology but also in social and subsistence domains Once cultural behavior was established as a basic human quality, the creation and transmission of culture became humans’ preeminent trait In our learning hypothesis, learning is sorted into two types: (1) individual learning, i.e., learning on one’s own through trial and error, drawing solely on one’s own ideas, and (2) social learning, i.e learning from others through imitation, being taught, or another process The Neanderthals had advanced lithic culture, but it was very conservative They continued to reproduce the same types of stone tools for almost 200,000 years, which suggests they were very good at social learning but did not have much ability to innovate On the other hand, the Sapiens invented various lithic industries after arriving in Europe, which could be a product of their aptitude for innovative individual learning The Neanderthals’ learning behavior, characterized by concentration on social learning but not on innovation, seems to have been adaptive to places where environmental conditions were rather stable from generation to generation The key difference may have been the flexibility of learning strategies in Sapiens, allowing them to switch between and effectively combine individual and social learning in quickly changing environments The final phase of the glacial epoch when the replacement occurred was characterized by a climate that fluctuated widely and rapidly between cold and warm, an environment that may have favored Sapiens’ learning strategies over that of Neanderthals This flexibility would have enabled them to quickly solve adaptive problems and thus to move swiftly and successfully into novel environments as they spread across the globe A wide range of research is needed to test the learning hypothesis In the RNMH project, six research teams (A01, A02, B01, B02, C01 and C02) were organized under a steering committee that gathered archaeologists, paleoanthropologists, social-cultural anthropologists, developmental psychologists, geneticists, climatologists, paleoecologists, neuroscientists, and Preface Preface xi others for collaborative interdisciplinary research Each team’s specific objects were as follows: A01: Archaeological research of the learning behaviors of the Neanderthals and early modern humans A02: Research on human learning behavior based on fieldwork among hunter–gatherers B01: Research on evolutionary models of human learning abilities B02: Reconstructing the distribution of Neanderthals and modern humans in time and space in relation to past climatic changes C01: Reconstruction of fossil crania based on three-dimensional surface modeling techniques C02: Functional mapping of learning activities in archaic and modern human brains It is indispensable to clarify the learning patterns in ancient and modern hunting and gathering societies for the demonstration of the learning hypothesis A01 investigated archaeological evidence, artifacts and traces of living sites indicative of past learning behaviors of the Middle and Upper Paleolithic humans Studies in experimental archaeology and ethnoarchaeology were also conducted to interpret ancient traces of learning A02 investigated contemporary hunter–gatherers’ learning behavior, their social and individual learning, mainly through children’s everyday activities, to discern the characteristic learning behavior of modern humans The study of hunter–gatherers has been one of the main themes in anthropology since its birth in the nineteenth century, and this way of life is believed to be the closest approximation in the contemporary world of ancient living conditions While it is not acceptable or accurate to assert a one-to-one relationship between the lives of contemporary hunter–gatherers with that of our human ancestors, it is also inappropriate to think that the research of hunting and gathering societies can shed no light on the reconstruction of ancient human conditions Appropriate and deliberate collaboration between socio-cultural anthropology and archaeology, paleoanthropology, and other related fields could help reconstruct the behaviors of ancient humans Team B01 conducted a theoretical study of the learning hypothesis by describing and analyzing mathematical evolutionary models They simulated and compared various learning strategies to find out what conditions might have led to the expansion of social learners or individual learners in specific societies Team B02 reconstructed the distribution of the Neanderthals and the Sapiens in time and space during 20–200 ka and also reconstructed the environments of those populations, including climatic conditions and ecological settings, in order to make comparisons of the differences in adaptation of each population to each environment The learning hypothesis does not necessarily postulate a large and sudden cognitive jump; however, there are apparent morphological differences between the crania of Neanderthals and Sapiens Therefore, it is crucially important to understand the relationship between brain morphology and its functions Team C01 tried to reconstruct the fossil crania and brains of Neanderthals and ancient modern humans, and C02 utilized fMRI in an attempt to identify the brain sites supposed to relate to various learning activities Learning behavior has essential importance for human culture and evolution There is, however, a huge difference between the learning done in formal school settings in modernized societies and that in hunting and gathering societies in the past as well as present Our study of social learning has been conducted mainly among contemporary hunter–gatherers in various natural and social environments and has revealed characteristics crucial to maintaining their culture, livelihood, and joie de vivre Social-cultural anthropology has methodologically avoided the unilineal cultural evolutionary approach for decades because of the misuse of Darwinian theory, but recent theoretical and methodological developments provide insights into social learning in humans as well as research problems of the RNMH project 25 Evolutionary Locus of the Neanderthal Between Chimpanzees and Modern Humans: A support These motivations look quite odd evolutionarily, because the learning style most nonhuman animals use to adapt to their environment is almost always “individual learning.” Social learning among social animals, e.g., observational and imitative learning, is still individual learning because the models whom individuals imitate not make any direct, active attempts to facilitate learning (not even the effort makes their actions visible to learners) Learners are just learning by themselves using another individual’s behavior as a cue On the other hand, in teaching, individuals conduct special altruistic behavior via social interaction to prompt learners to learn This type of learning behavior is costly and risky for both learners and teachers because it requires the expenditure of extra energy not necessary in individual learning, because learning episodes may subtract from more direct opportunities to obtain personal benefit and because both individuals may be exposed to dangers they would not otherwise be exposed to Therefore, there must be both proximal and distal benefits to teaching and being taught for educational behavior to be advantageous Why modern humans seek being taught despite their capacity of individual and observational learning? What is the motivation for modern humans to teach and learn by teaching? The psychological structure of the motivation to teach and to be taught might be a cue to investigate both proximal and distal benefits to this behavior In the next sections, I discuss the results of investigations into two research questions: (1) How are the psychological states of teaching motivation related to those of individual and imitative (observational) learning situations? And (2) how is the motivation to teach related to altruism and general factor of personality (GFP), which are thought to be evolutionary relevant psychological traits? 303 25.4.5 Motivation to Teach and to Be Taught Compared with Individual and Observational Learning Situations In 2014, I conducted an experimental study (Ando 2014) investigating motivations and related psychological states of individual, imitative, and educational or instructed learning Participants in this experiment were 20 Japanese university students Each engaged in three styles of learning when solving an entangled puzzle: individual learning, in which they solved a puzzle alone without any assistance from others; imitative learning, in which they solved a puzzle alone without any assistance from others but were permitted to imitate a model who had learned the target puzzle in advance; and instructed learning, in which participants solved a puzzle with a teaching agent who was permitted to teach and a learner was permitted to request being taught how to finish a puzzle In the imitative and instructed learning situations, the participants experienced both the model/instructor role and the learner role and were asked about their attitude toward each role, model, and learner and about their motivation to learn Figure 25.7 illustrates causal relationships between extracted psychological factors as indicated by path analysis Motivation to teach is significantly correlated with direct and indirect attitudes about establishing a good partnership with learners when participants played the role of instructor and paid attention to learners This result indicates that in psychological motivations to teach and to learn from teachers are complementary, that is, two sides of the same coin It is also suggested that the easier and the less challenging a task become for an instructor (expert), the more an instructor pays attention to learners (novices) who start engaging in the task Fig 25.7 Causal relationship between teaching motivation and related psychological factors 304 Educational behavior is not one-sided; a teacher’s motivation to teach is complemented by the learner’s motivation to be taught This finding suggests that the motivation to teach and motivation to be taught are supported, at least partially, by a common cognitive mechanism 25.4.6 Two-Factor Structure of Motivation to Teach In another study (Ando 2013), I investigated distal reasons for the evolution of psychological structures underlying educational motivation in order to clarify these structures’ evolutionary relationship to altruistic behavioral tendencies and general factors of personality (GFPs) The work of understanding altruism, and reciprocal altruism and indirect reciprocity specifically, in sophisticated evolutionary terms was being advanced by Hamilton (1964), Williams (1966), Trivers (1971), and Nowak and Sigmund (1998) In this study, the Self-Report Altruism Scale Distinguished by the Recipient (SRAS-DR) (Oda et al 2013) was used to measure three aspects of altruism: (1) altruism toward family members who share genes (kin selection), (2) altruism toward friends and acquaintances who can possibly reciprocate (direct reciprocity), and (3) altruism toward strangers (indirect reciprocity) General factor of personality (GFP) is a single principal component (Musek 2007) consisting of multiple desirable social traits (i.e., emotional stability, extraversion, openness to cultural experiences, cooperative tendency, and conscientiousness—the “Big Five” factors) This component, or dimension, is hypothesized to have evolved in humans under heavy selection pressures for social adaptation (Rushton et al 2008, 2009) In this study, principal components of the Big Five factor scores were used to assess general factor of personality in 967 adult participants from the Tokyo area The psychological structure of motivation to teach was investigated through a 36-item questionnaire that addressed various aspects of attitudes toward teaching and feelings, shown in Table 25.1 The questionnaire produced ten primary factors, as well as two secondary factors— “controlling/enlightenment” motivation and “supportive/ altruistic” motivation to teach It was to be expected that “supportive/altruistic” motivation was significantly correlated with all three aspects of altruism An interesting finding was that “controlling/enlightenment” motivation was significantly correlated with indirect reciprocity This two-factor structure was still maintained even after removing, by regression analysis, GFP and general factor of altruism (the latter was calculated as a principal component of the three dimensions of the SRAS-DR) These two residual factors, without general evolved psychological tendencies (general factor of personality and altruism), can be interpreted as components specific for teaching (knowledge J Ando Table 25.1 Second factor structure of educational motivation and its relationship with altruism Factor Controlling/ enlightenment 0.745 0.714 Factor Altruistic/ supportive 0.003 0.473 First factors Self-seeking Educational content Elation 0.685 0.463 Obedience À0.585 À0.305 Attention 0.580 0.155 Self-teaching 0.533 À0.127 Teaching 0.245 0.855 Aspiring teacher 0.274 0.760 Disliking teacher 0.127 À0.727 Empathy 0.297 0.422 Correlations with altruism scores (*p < 0.05, **p < 0.01) Kinship 0.008 0.322** Direct À0.010 0.357** Indirect 0.231* 0.266** transmission) Therefore, it is reasonable to postulate that teaching motivation factors consist of evolutionarily adaptive general factors and teaching-specific factors From an evolutionary standpoint, the finding that the motivation to teach contains elements of a desire to control and to enlighten suggests this educational behavior may have evolved not just as altruistic and supportive behavior but also for its usefulness in maintaining control over social order through inculcation This aspect of educational motivation is parallel to a general trend in nonhuman social animals, which control members of their community through aggressiveness, physical power, and authority to keep a high status and distribute resources such as food to maintain the social order/hierarchy It seems reasonable to consider that educational behavior may have emerged as a strategy to distribute not food, but knowledge Recent findings that suggest teaching behavior emerges very early in human ontogeny (at around one and a half years old) (Akagi 2004; Liszkowski et al 2006; Mong and Hashiya 2014) also support its evolutionary importance in H sapiens 25.5 Concrete and Abstract Thinking In this, the last section, the location of Neanderthals on the cognitive continuum between chimpanzees and modern humans will be considered in terms of Piagetian developmental stage model A cognitive test to measure formal analytic reasoning ability in Western countries, called Raven’s Progressive Matrices, was administered to adult Baka Pygmy hunter-gatherers in Africa as a tool in this consideration As discussed in Sect 25.1, the Baka are some of the most spatially and culturally distant peoples in relation to Western 25 Evolutionary Locus of the Neanderthal Between Chimpanzees and Modern Humans: A industrial culture, and their hunter-gathering lifestyle is closer than any other extant to that of the Neanderthals Pascual-Leone, a neo-Piagetian psychologist, postulates that the quantitative development of working memory capacity can explain the development of cognitive stages, proposed by Piaget as sensorimotor (from birth to years old approximately), preoperational (2–6), concrete operational (6–12), and formal operational (after 12) (Pascual-Leone 1980) I suggest it is possible to apply this ontogenic theory of cognitive development to a phylogenic model of cognitive abilities and learning strategies As Jane Goodall discovered, chimpanzees can make and use tools to fish for ants and to crack nuts (Goodall 1969) These abilities can be acquired by coordinating psychomotor response system with mental representations (i.e., not by signs or symbols as sapiens can use, but through mental representations of one’s own body movements) using a WM of just one chunk In contrast, as described by Piaget, cognitive ability in humans becomes increasingly complex and powerful over the course of development A nonlogical preoperational stage, in which children lack the ability to manipulate “conservation task,” is succeeded by the middle concrete operational stage in which logical operation is only possible for “here and now” matters Finally, the formal operational stage sees the development of a cognitive capacity for abstract and formal logico-mathematical operations (Piaget 1970) The question then arises, if humans reach cognitive maturity at the formal operational stage, what constituted cognitive maturity for Neanderthals? Did their cognitive ability reach its pinnacle at a preoperational stage? The concrete operational stage? To solve this question, we will consider the cognitive developmental characteristics of hunter-gatherers, as they possess the same cognitive ability as people living in Western, educated, industrialized, rich, and democratic (WEIRD) societies, but live a lifestyle more similar to that of the Neanderthals One objective is to understand whether there might be any differences in cognitive development between those living in a society with formal education, where logico-mathematical thought and scientific knowledge are prized, and those living in a society where formal education is absent The issue then becomes whether it is possible to extrapolate a continuum of within-species differences to interspecies differences between chimpanzees, Neanderthals, and modern humans The author administered Raven’s Progressive Matrices to Baka Pygmy adults living in Lomie´, southeastern Cameroon, in August 2013 This cognitive test battery is frequently used to measure fluid intelligence or heuristic problem-solving ability and consists of an array of two-by-two or two-bythree geometric configurations, each of which follows certain logical rules Each configuration contains a blank space in the pattern, and participants are asked to choose from among alternative “patches” to fill the whole and complete the logical pattern (see Fig 25.8) 305 Fig 25.8 Raven’s Progressive Matrices and response distribution (percentages and numbers in parentheses) 306 These figures illustrate some excerpts from Raven’s items and the distribution of answers chosen by Baka participants (in parentheses) Strikingly, “correct” answer rates were extremely low in all items In even the simplest item, A1 (Fig 25.8a), where the congruent pattern around the blank should be completely selected, the rate of correct answers was only 32 % In A8 (Fig 25.8b), it seems participants did not try to count the number of lines, and in Ag4 and Ag5 (Fig 25.8c, d), instead, they tended to select complete items which constitute gestalt patterns This way of thinking is called the gestalt algorithm, in contrast to the analytic one (Hunt 1974) Even when they were asked why they chose them, they just said “joko (good)” or “this looks like a soccer ball (in the case of A1).” They might prefer a “good shape” or an image associated with personal experiences (e.g., watching or playing soccer) Here, they don’t appear to be at the formal operational or even concrete operational stage From a Western perspective, the Baka mode of thought does not appear analytic and logical but rather associative and gestalt These results are in line with the findings of vast number of cross-cultural cognitive studies conducted in non-Western cultures (e.g., Dasen 1977, 1984; Cole and Brunner 1971; Scribner and Cole 1981; Rogoff 2003) Hunter-gatherers’ way of thinking looks drastically different from that of WEIRD people because of cultural context, value judgment, education, and language differences It is not correct or accurate to characterize that their cognitive ability as inferior to that of WEIRD people in an absolute sense However, when we apply the Piagetian developmental theory to this situation in order to interpolate the cognitive ability of the Neanderthal in between chimpanzees (at the sensorimotor stage) and H sapiens (at the formal operational stage), it may be reasonable to speculate that Neanderthals did not surpass the preoperational or even the last substage of sensorimotor stage, because hunter-gatherers’ performance (not competence) is mainly located between preoperational (Oesterdiekhoff 2012) and concrete operational stages (Dasen 1977) when considered under the Western paradigm It should be noted that this interpretation is not an ethnocentric claim, but one made for the sake of discerning on a theoretical level, the level learning ability possessed by Neanderthals Hunter-gatherers are fully capable of abstract and symbolic reasoning appropriate to cultural context (Cole and Scribner 1974), but this does not mean the Neanderthals were at concrete operational level When compared to the cognitive level of chimpanzees, which are at the sensorimotor stage, the preoperational way of thinking and behavior of the Neanderthals must have been a great J Ando advantage in adapting to their natural and social circumstances It is plausible to assume that in the process of evolution, the symbol-manipulating abilities of hominids gradually transitioned from the sensorimotor stage, in which symbolic representations are not discriminated from physical kinetic movements, to the preoperational stage, in which symbolic operations are manipulated in reference to specific situations It is clear that the H sapiens hunter-gatherers can think logically in a concrete or even formal operational level when we look at their cultural products and listen to their dialogues, but this mode of thought is not necessarily implemented in unfamiliar and artificial tests such as the Raven’s Matrices tasks The concrete operational way of thinking in even contexts independent from natural situations probably became dominant in humans with the advent of agricultural and pastoralism Formal operational thinking likely emerged within the context of the industrial revolution, at which point more stress was placed on analytic and logical capacities as people sought to understand underlying mechanisms of the natural world through scientific inquiry When we assume a genetic, evolutionary, and developmental continuum of learning strategies as depicted in Fig 25.2, the Neanderthals must have possessed a way of thinking in the range of preoperational cognitive ability, characterized by self-centeredness and a lack of reversibility (i.e., the ability to recognize that numbers or objects can be changed and returned to their original condition) in thinking and perhaps storytelling in an anecdotal and narrative way This tendency would be consistent with two other hypothesized characteristics of Neanderthal cognition mentioned above: (1) a lack of WM capacity and (2) a poor theory of mind coupled with limited ability to read facial expressions 25.6 Conclusion The model of learning ability as a genetic, evolutionary, and developmental continuum postulated in this paper, illustrated in Fig 25.9, is an attempt to describe the relationships between three species—chimpanzees, Neanderthals, and H sapiens—on a single sociocultural dimension The H sapiens range is divided into three cultural stages, the hunter-gatherer stage, the agricultural stage before the industrial revolution, and the WEIRD stage (Lancy 2010), where technology, science, and information are dominant 25 Evolutionary Locus of the Neanderthal Between Chimpanzees and Modern Humans: A 307 Fig 25.9 Summary picture of the hypothesis In this model, the Neanderthal is hypothesized as a species able to use tools but incapable of symbolic, segmental, or manipulative use of language (i.e., Mithen’s (1996) concept of Hmmmm) and possesses only rudimentary theory of mind Cognitively, their working memory capacity had expanded from one to two “chunks” quantitatively and maybe qualitatively, but their cognitive manipulation was preoperational and gestalt, concrete, anecdotal, and personal rather than analytic, abstract, universal, and logical The limited learning ability of the Neanderthals might have been to a modest theory of mind, but, thanks to this limitation, they could concentrate on individual, physical behavior to master tool shaping and manipulation skills via individual learning or sometimes, if necessary, observational learning and emulation of another’s behavior It is unlikely that they taught each other Because teaching (overt instruction) rarely occurs even in modern human hunter-gatherers, it is difficult to assume that Neanderthals did this as a primary way of learning It may not have been until the agricultural and pastoral era that H sapiens acquired knowledge through educational activities and systems as a custom These systems, particularly agriculture, necessitated more systematic educational practices to enable the transmission of more complex methods of controlling aspects of nature It was only after the industrial revolution in the West that the institutionalized educational systems became standard in communities Social learning among Neanderthals was, therefore, not by education or even full-blown imitation using theory of mind, but rather emulation of others’ repetitive behavior Perhaps individual learning focused on one’s own behavior was dominant due to a poverty of ToM Because Neanderthal’s WM capacity likely consisted of only two “chunks” (rather than three to five in H sapiens), they could conduct behavior in a complex, hierarchical, and accumulative way, but did not possess the cognitive abilities that enable one to exchange, integrate, or even create and accumulate knowledge via social interaction 308 The hypotheses proposed in the current paper need to be tested empirically More anthropological research on cognition among hunter-gatherers is necessary to verify the universality of both educational behavior and the Piagetian cognitive level of H sapiens Molecular genetic evidence of autism and a macrocephalic brain from molecular archeology may support (or deny) the “limited theory of mind” hypothesis Neuroimaging studies of that compare how people’s brains fire, while being taught as opposed to individual and observational learning will provide additional circumstantial evidence Finally, other hominids, such as australopithecines, Homo erectus, and Homo heidelbergensis, are other candidate species that may possibly have been able to teach However, we currently have even less evidence as basis to speculate than we for Neanderthals It is possible to infer that some precursors which lead to an ability to teach had emerged in those ancient hominids Satisfactory answers to the questions presented in this discussion will depend on the emergence of more archeological evidence References Akagi K (2004) Can one year-olds teach?: active teaching when observing others who cannot solve problems Jpn J Dev Psychol 15(3):366–375 Ando J (2009) Evolutionary and genetic bases of education: an adaptive perspective Annu Rep Educ Psychol Jpn 48:235–246 Ando J (2012a) Is educational behavior evolutionary product or cultural product?: an example of “educational behavior” of hunter-gatherer, Baka pigmy Paper presented at the 54th annual conference of Japanese society of educational psychology Ryukyu University, Okinawa, 23 Nov 2012 Ando J (2012b) On “Homo educans” hypothesis In: Watanabe S (ed) CARLS series of advanced study of logic and sensibility Keio University Press, Tokyo, pp 147–156 Ando J (2013) A bridge between cognitive ability and personality: a longitudinal (14 year) twin study of working memory and GFP Paper presented at the 54th annual meeting of Japan society of social psychology, Okinawa International University, Okinawa, June 1998 Ando J (2014) Comparisons between individual, imitative and instructed learning In: Akazawa T, Ogihara N, Tanabe H, Terashima H (eds) Dynamics of learning in Neanderthals and modern humans, vol 2, Cognitive and physical perspectives Springer, Tokyo, pp 17–24 Ando J, Ono Y, Wright M (2001) Genetic structure of spatial and verbal working memory Behav Genet 31:615–624 Baddeley A (1986) Working memory Oxford University Press, Oxford Bruner E (2008) Comparing endocranial form and shape differences in modern humans and Neanderthals: a geometric approach Paleoanthropol 93À106 Bruner E, Manzi G, Arsuaga JJ (2003) Encephalization and allometric trajectories in the genus Homo: evidence from the Neanderthal and modern lineages Proc Natl Acad Sci U S A 100(26):15335–15340 Byrne RW, Rapaport LG (2011) What are we learning from teaching? Anim Behav 82:1207–1211 Caro TM, Hauser MD (1992) Is there teaching in nonhuman animals? Q Rev Biol 67:151–174 J Ando Cole M (2010) What’s culture got to with it?: educational research as a necessarily interdisciplinary enterprise Educ Res 39:461–470 Cole M, Brunner JS (1971) Cultural differences and inferences about psychological processes Am Psychol 26:87–876 Cole M, Scribner S (1974) Culture and thought Wiley, New York Coolidge FL, Wynn T (2005) Working memory, its executive functions, and the emergence of modern thinking Camb Archaeol J 15(1):5–26 Courchesne E (2004) Brain development in autism: early overgrowth followed by premature arrest of growth Ment Retard Dev Disabil Res Rev 10(2):106–111 Courchesne E, Carper R, Akshoomoff N (2003) Evidence of brain overgrowth in the first year of life in autism J Am Med Assoc 290:337–344 Cowan N (2000) The magical number in short-term memory: a reconsideration of mental storage capacity Behav Brain Sci 24:87–185 D’Esposit M, Detre JA, Alsop DC, Shin RK, Atlas S, Grossman M (1995) The neural basis of the central executive system of working memory Nature 378(6554):279–281 D’Esposito M, Postle BR, Rypma B (2000) Prefrontal cortical contributions to working memory: evidence from event-related fMRI studies Exp Brain Res 133(1):3–11 Dasen PR (1977) Piagetian psychology: cross-cultural contribution Gardner, New York Dasen PR (1984) The cross-cultural study of intelligence: Piaget and the Boule Int J Psychol 19:407–434 Diamond A (2013) Executive functions Annu Rev Psychol 64:131–168 Ericsson KA, Kintsch W (1995) Long-term working memory Psychol Rev 102:211–245 Evans PD, Gilbert SL, Mekel-Bobrov N, Vallender EJ, Anderson JR, Vaez-Azizi LM, Tishkoff SA, Hudson RR, Lahn BT (2005) Microcephalin, a gene regulating brain size, continues to evolve adaptively in humans Science 309(5741):1717–1720 Flynn JR (1987) Massive IQ gains in 14 nations: what IQ tests really measure Psychol Bull 101:171–191 Franks NR, Richardson T (2006) Teaching in tandem-running ants Nature 439:153 Friedman NP, Miyake A, Robinson JL, Hewitt JK (2011) Developmental trajectories in toddlers’ self-restraint predict individual differences in executive functions 14 years later: a behavioral genetic analysis Dev Psychol 47(5):1410–1430 Fujisawa KK, Ozaki K, Ozaki K, Yamagata S, Kawahashi I, Ando J (2012) The genetic and environmental relationships between head circumference growth in the first year of life and sociocognitive development in the second year: a longitudinal twin study Dev Sci 15:99–112 Goodall JVL (1969) My friends the wild chimpanzee National Geographic Society, Washington, DC Hamilton WD (1964) The genetical evolution of social behavior I, II J Theor Biol 7:1–52 Hewlett BS, Fouts HN, Boyette AH, Hewlett BL (2011) Social learning among Congo Basin hunter–gatherers Philos Trans R Soc B Biol Sci 366(1567):1168–1178 Hoffmann W, Schmeichel BJ, Baddeley AD (2012) Executive functions and self-regulation Trends Cogn Sci 16(3):174–180 Hoppitt WJE, Brown GR, Kendal R, Rendell L, Thornton A, Webster MM, Laland KN (2008) Lessons from animal teaching Trends Ecol Evol 23(9):486–493 Hunt E (1974) Quote raven? nevermore In: Gregg LW (ed) Knowledge and cognition Lawrence Erlbaum Press, Oxford Ishida H (2005) Kaseki kataru (Fossils speak) In: Akazwa T (ed) Neanderutahru no shoutai-Karera no ‘nayami’ ni semaru (Real identity of the Neanderthals: an approach toward their ‘annoyance’ Asahi Shinnbunn Shuppan, Tokyo, pp 165–184 25 Evolutionary Locus of the Neanderthal Between Chimpanzees and Modern Humans: A Kamei N (2010) Mori no chiisana hantah tachi-shuryou saishuumin no kodomo no minnzokushi (Little hunters in forests: folklore of hunter-gatherer children) Kyoto University Press, Kyoto Kawachi M (2005) Thinking about how they grew In: Akazwa T (ed) Neanderutahru no shouti-Karera no ‘nayami’ ni semaru (Real identity of the Neanderthals: an approach toward their ‘annoyance’ Asahi Shinnbunn Shuppan, Tokyo, pp 205–236 Klein RG, Edger B (2002) The dawn of human culture Nevaumont Publishing Company, New York Kline MA (2015) How to learn about teaching: an evolutionary framework for the study of teaching behavior in humans and other animals Behav Brain Sci 38:1–71 Lancy DL (2010) Learning “from nobody”: the limited role of teaching in folk models of children’s development Child Past 3:79–106 Liszkowski U, Carpenter M, Striano T, Tomasello M (2006) 12 and 18 month olds point to provide information for others J Cogn Dev (2):173–187 Matsuzawa T (2011) Souzou suru Chikara- Chimpanzee ga oshietekureta ningen no kokoro (An ability to imagine: what chimpanzees teach us about the human mind) Iwanami Shoten, Tokyo Mekel-Bobrov N, Gilbert SL, Evans P, Vallender EJ, Anderson JR, Hudson RR, Tishkoff SA, Lahn BT (2005) Ongoing adaptive evolution of ASPM, a brain size determinant in Homo sapiens Science 309(5741):1720–1722 Mithen SJ (1996) The prehistory of the mind: a search for the origins of art, religion, and science Thames and Hudson, London Mithen SJ (2005) The singing Neanderthals: the origins of music, language, mind and body Harvard University Press, Cambridge Mong X, Hashiya K (2014) Pointing behavior in infants reflects the communication partner’s attentional and knowledge states: a possible case of spontaneous informing PLoS One 9(9):e107579 Musek J (2007) A general factor of personality: evidence for the Big One in the five-factor model J Res Pers 41(6):1213–1233 Nowak AM, Sigmund K (1998) Evolution of indirect reciprocity by image scoring Nature 393:573–577 Oda R, Dai M, Niwa Y, Ihobe H, Kiyonari T, Takeda M, Hiraishi H (2013) Self-Report Altruism Scale Distinguished by the Recipient (SRAS-DR): validity and reliability Jpn J Psychol 84(1):28–36 Oesterdiekhoff GW (2012) Was pre-modern man a child? The quintessence of the psychometric and developmental approaches Intelligence 40:470–478 309 Pascual-Leone J (1980) Constructive problems for constructive theories: the current relevance of Piaget’s work and a critique of informationprocessing simulation psychology In: Kuwe R, Spada H (eds) Developmental models of thinking Academic, New York, pp 263–296 Piaget J (1970) Piaget’s theory In: Mussen PH (ed) Carmichael’s manual of child psychology, vol Wiley, New York, pp 703–732 Premack DG, Woodruff G (1978) Does the chimpanzee have a theory of mind? Behav Brain Sci 1(4):515–526 Redcay E, Courchesne E (2005) When is the brain enlarged in autism? A meta-analysis of all brain size reports Biol Psychiatry 58:1–9 Robins DL, Fein D, Barton ML (1999) The modified checklist for autism in toddlers (M-CHAT) Self-published, Storrs Rogoff B (2003) The cultural nature of human development Oxford University Press, New York Rushton JP, Bons TA, Hur Y-M (2008) The genetics and evolution of the general factor of personality J Res Pers 42(5):1173–1185 Rushton JP, Bons TA, Ando J, Hur Y-M, Irwing P, Vernon PA, Petrides KV, Barbaranelli C (2009) A general factor of personality from multitrait-multimethod data and cross-national twins Twin Res Hum Genet 12:356–365 Scribner S, Cole M (1981) The psychology of literacy Harvard University Press, Cambridge Strauss S, Ziv M (2011) What nonhuman animal teaching teaches us about teaching In: Watanabe S (ed) CARLS series of advanced studies of logic and sensibility, vol Keio University Press, Tokyo, pp 363–371 Strauss S, Ziv M (2012) Teaching is a natural cognitive ability for humans Mind Brain Educ 6(4):186–196 Thornton A, McAuliffe K (2006) Teaching in wild meerkats Science 313:227–229 Thornton A, Raihani NJ (2008) The evolution of teaching Anim Behav 75:1823–1836 Tomasello M (1999) The cultural origins of human cognition Harvard University Press, Cambridge Trivers RL (1971) The evolution of reciprocal altruism Q Rev Biol 46:35–57 Williams GC (1966) Adaptation and natural selection Princeton University Press, Princeton Yamamoto S, Humle T, Tanaka M (2013) Basis for cumulative cultural evolution in chimpanzees: social learning of a more efficient tooluse technique Plos One doi:10.1371/journal.pone.0055768 Zimmer C (2006) Evolution: the triumph of an idea Harper Perennial, New York Reflections on Hunter-Gatherer Social Learning and Innovation 26 Hideaki Terashima 26.1 The RNMH Project and Rethinking the Replacement Our mission in the RNMH project was to elucidate and compare the learning behaviors of modern humans and Neanderthals in order to test the “learning hypothesis” – that modern humans’ superior learning ability was the crucial factor in their success and the concurrent demise of Neanderthals in Europe We utilized three sources to test the learning hypothesis: (1) contemporary modern humans, particularly children in foraging cultures, (2) archaeological evidence of Neanderthal learning and culture, and (3) great apes like chimpanzees and bonobos who are our closest relatives by virtue of shared common ancestors As mentioned in Chap 1, the study of the learning behaviors of hunter-gatherer children has been far less popular than that of nonhuman primates We have conducted an investigation into the learning behavior of modern hunter-gatherer children to discover in their everyday activities the social, ecological, behavioral, and psychological factors that may ensure the fidelity of cultural transmission and expand opportunities for innovative behavior, considering the foundations of human culture Participant observation, interviews, and psychological experiments were carried out among various foraging peoples Data on Neanderthal technologies, cognitive abilities, social life, physical conditions, and subsistence activities were gathered so that comparisons could be made between their learning behaviors and those of sapiens Making this comparison with so little data available on Neanderthals has been a difficult and unique challenge in sociocultural anthropological studies H Terashima (*) Faculty of Humanities and Sciences, Kobe Gakuin University, Kobe, Japan e-mail: terasima@kh.rim.or.jp Scholarly interest in Neanderthals has continued to grow since the emergence of a new theory on the dispersal of the anatomically modern humans (Homo sapiens, hereafter referred to as “sapiens”), based on advanced genetic studies done in the 1980s, which suggested there may have been interaction between the species According to the theory, sapiens emerged in Africa ca 150–200 ka and began to disperse from there to Eurasia from around 100 ka and met with the closely related but slightly different hominins, the Neanderthals, in Europe Interaction between Neanderthals and sapiens has since garnered special interest and become the subject of intense academic debate Recently, genetic studies of the Neanderthal genome have become possible, producing many intriguing discoveries, including evidence of interbreeding with sapiens (Green et al 2010; Paăaăbo 2014a, b; Sankararaman et al 2012) In addition, an innovative radiocarbon dating technique called ultrafiltration has enabled more accurate and reliable dating of archaeological objects (Higham et al 2010, 2014; Higham 2011; Sano 2012) This method has revealed that both Neanderthals and sapiens lived side by side in Europe for several millennia However, controversy over the subject persists as many mysteries regarding differences in cognitive ability, life history, social life, and behavior between the two hominin groups remain unsolved Here I will first show several important points revealed in the RNMH project An archaeological information database called “NeanderDB” was constructed with the collaboration of teams A01 and B02 (see Preface for the mission of each team) They accumulated data on the archaeological record between 200 and 20 ka, covering the areas of Europe, Asia, and Oceania For Europe, records on 737 sites, 2,151 archaeological horizons, and 3,453 radiometric dates obtained using the methodology mentioned above were input through January 30, 2015 Analyses of cultural transitions in both space and time were then conducted for both species using the NeanderDB and other sources of evidence We found the largely accepted narrative in which sapiens evolved in # Springer Japan 2016 H Terashima, B.S Hewlett (eds.), Social Learning and Innovation in Contemporary Hunter-Gatherers, Replacement of Neanderthals by Modern Humans Series, DOI 10.1007/978-4-431-55997-9_26 311 312 Africa, migrated to Europe, and outcompeted Neanderthals with advanced technologies and a modern behavioral repertoire (Mellars 2006) needs reconsideration Sapiens migrated into Europe in at least three waves: the first was around 47 ka, the second around 45 ka, and the third around 43 ka Neanderthal residential sites had already been diminishing by the time of the first arrival of the sapiens, probably due to the severe cold of Marine Isotope Stage (MIS4, ca 71–57 ka) (Sano 2012) There seem to have been no significant differences between the two species lithic cultures when sapiens first arrived (Kadowaki 2013, 2014; Kadowaki et al 2015), and neither were there differences in their environments The climate in Europe at that time was fluctuating more abruptly than in Africa and Siberia due to climatic events such as the Heinrich event (HE5) and the Dansgaard–Oeschger (DO) event (Chan et al 2014) After the third wave of migration, the sapiens moved into Mediterranean coastal areas where Neanderthals liked to dwell and developed the proto-Aurignacian culture there, characterized by advances in lithic culture such as the use of backed bladelets attached to the spear points (Sano and Omori 2015) The density of proto-Aurignacian sites increased rapidly thereafter, while the Neanderthals, losing their favorite niches, moved into mountainous areas with steep slopes and a colder climate Their population decreased rapidly and finally disappeared by 40 ka (Bar-Yosef 2013; Higham et al 2014; Sano 2013; Sano and Omori 2015) The relationship between the newcomers and the indigenous Neanderthals does not seem to have been broadly antagonistic In some places, the latter adopted some technology from the former and interbreeding took place Recent theoretical studies of human demography predict that only a small demographic advantage would have been sufficient for sapiens numbers to grow rapidly and lead to the demise of the Neanderthals (Zubrow 1989; Snodgrass and Leonard 2009; Sørensen and Leonard 2001) A simulation study revealed that a mere % difference in mortality could result in the extinction of one hominin group in as few as 30 generations (Zubrow 1989) One possible explanation for sapiens’ success and the Neanderthals’ demise, which is the core of the learning hypothesis, would be that sapiens were capable of developing survival strategies in response to the drastic climatic changes during the last glacial epoch, while the Neanderthals were not Questions then arise of when, where, and how sapiens developed successful survival strategies for these conditions and why the Neanderthals did not or could not Differences in cognitive ability may immediately come to mind, but the nature of these differences, if they did in fact exist, remains poorly understood There are two schools of thought The first posits that there were large biologically based differences in cognitive ability and brain function between the two groups and that H Terashima Neanderthals lacked the characteristics critical for innovation and adaptive behavior (Coolidge and Wynn 2005; Wynn and Coolidge 2004, 2007, 2011; Klein and Edger 2002; Mithen 1996; Ambrose 2010; Pearce et al 2013; Boquet-Appel and Tuffreau 2009) The second suggests cognitive differences were unimportant and other factors such as demographic and social complexity (Zilha˜o 2007, 2013), division of labor (Kuhn and Stiner 2006), assimilation and interbreeding (Villa and Roebroeks 2014), artifact preservation biases (Langley et al 2008), demographic and social complexity (d’Errico and Stringer 2011), or gradual enhancements of working memory within the genus Homo (Martı´n-Loeches 2010) precipitated the replacement The learning hypothesis of the RNMH project presupposes that the cultural differences between the two species were caused by their difference in brain functions, i.e., biologically defined capacities Teams C01 and C02 carried out a comparative investigation into the brain morphology and functions of the two species (Tanabe et al 2014a) The fossil crania of ancient hominins and the brains situated in those crania were reconstructed with high accuracy, drawing on a mathematical reconstruction method (Kondo et al 2014; Ogihara et al 2014) The results show that in ancient sapiens the parietal region and both hemispheres of the cerebellum were larger than in the Neanderthals’ brains, whereas widespread occipital regions were larger in the Neanderthals The volume of the frontal region of both species turned out to be almost similar contrary to popular assumptions (Kubo et al 2014a, b; Tanabe et al 2014b) Recently, researchers have found the cerebellum plays an important role in higher cognition such as language, working memory, and social abilities (Imamizu and Kawato 2012; Penhune and Steele 2012; Bower and Parsons 2003) and suggested that the sapiens, who have a larger cerebellum, may have a more efficient brain structure A recent fMRI investigation of the brain functions of modern sapiens by C02 has revealed that there is a positive correlation between the size of the cerebellum and cognitive functions, especially in regard to language capacity and attention (Tanabe et al 2014b) However, the relationship between brain functions and morphological characteristics is still far from clear Many regions of the brain can carry out multiple functions, and the connections between functions and particular brain regions should not be glossed over (Healy and Rowe 2006) As far as archaeological evidence goes, it was only after the replacement of Neanderthals by sapiens that the rapid development of symbolic and artistic behavior appeared and technological innovations such as spear-throwers spread widely In Africa, modern behavior seems to have appeared sporadically in various sites from around 70–100 ka, but was not sustained for long periods (McBrearty and Brooks 2000; Kadowaki 2013, 2014; d’Errico and Stringer 2011) Judging from 26 Reflections on Hunter-Gatherer Social Learning and Innovation 313 archaeological materials, there seems to have been no critical behavioral difference between Neanderthals and sapiens before the beginning of the replacement Neither the groups of researchers who postulate the cognitive superiority of sapiens nor those who doubt that premise have been able to show evidence of biological difference; therefore, assertions about the cognitive difference between the two species based on archaeological artifacts remain speculative (Nishiaki 2014) Finally, there are also arguments based on other factors, such as population demography, which could have more influence on producing and spreading innovations than any cognitive difference (Powell et al 2009; Shennan 2001) What has become clear through the RNMH project is that it might be difficult to judge whether the biological differences in learning ability between the two species were the primary cause of the replacement and that the experience of rapid environmental changes and new social environments, shaped in part by Neanderthal populations, spurred sapiens to develop new technologies and social relationships This does not deny the importance of learning capacity per se but emphasizes environmental influences that might have activated or augmented existing capacities for learning and innovation In humans, culture has in many respects supplanted biology as the primary means of interacting with the environment In addition to brain function, the co-development of various social, cultural, and biological factors that influence learning behavior is indispensable for the innovative potential of our species (Coward and Grove 2011; Terashima 2013) in some cases forager children have been shown to produce substantial amounts of food (Hagino and Yamauchi, Chap 11, Hawkes et al 1995) Forager children spend their time playing and exploring their natural environment and often learn complex foraging techniques by imitating adult activities (Hagino and Yamauchi 2014; Kamei 2005, Imamura, Chap 15) Vertical transmission from parents to children is most commonly seen as the mode of cultural transmission in infancy and early childhood Once children are weaned and join play groups, horizontal (peer-to-peer) transmission increases In adolescence, the range of daily activities expands, and oblique transmission (transmission from adults other than parents) increases, for example, when special knowledge or skills are learned from experts (Hewlett 2013, Hewlett, Chap 17) These switches in prevalent mode of transmission, called the multistage learning pattern (Reyes-Garcı´a et al., Chap 4), partly match a theoretical model called the “learning schedule hypothesis” proposed by team B01 (Aoki et al 2012; Mesoudi and Aoki 2015; Lehmann et al 2013; Wakano and Miura 2014) This model, constructed of a mixture of social and individual learning, suggests social learning pertaining to the accumulation of specific subsistence-related skills and knowledge should be prevalent early in life and that innovative individual learning that contributes to improvement of one’s own culture should follow in later life stages It appears forager children generally learn all the subsistence skills necessary to be successful adults by age 10 Unfortunately, almost nothing is clear about the development pattern and learning behavior of Neanderthal children For adult Neanderthals, we have the following traditional picture of their life: they had a robust, heavily muscular body due to an adaptation to the cold climate, the sustainment of which made it necessary to consume large amounts of food (Sørensen and Leonard 2001; Snodgrass and Leonard 2009) Hunting of large game, including bison, rhinos, and horses, was the solution, and so women and children likely engaged in hunting rather than collecting plants and small animals as modern hunter-gatherers (Kuhn and Stiner 2006) Due to the dangers of large game hunting, they may have been injured often and had shorter life spans (Smith et al 2010; Caspari and Lee 2004) These dangers may have favored shorter childhood and adolescent stages and faster maturation than in sapiens If so, the ontogeny of Neanderthal learning might also have been different from that of sapiens’ children, putting them at a disadvantage in terms of learning capacity (Yamauchi, Chap 24) In particular, individual learning in late childhood might have been affected more than the social learning of subsistence knowledge and skills earlier in life There is, however, another view, which asserts Neanderthal children were well protected and cared for Even if their 26.2 The Quest for Social Learning in Sapiens and Neanderthals Our research on learning among contemporary huntergatherers has revealed a crucial relationship between children’s physical development and learning strategies As is often noted, modern humans have a unique development pattern, characterized by a long childhood lasting for about 15 years before becoming adults, including a special period called “adolescence,” when the body and the mentality develop at an accelerated rate in preparation for adulthood (Bogin 1997) In spite of the importance of the relationship between this development pattern and learning behavior, there is yet a paucity of research on the matter Extended childhood means increased costs of child-rearing for parents, which must be compensated for by a significant benefit From an evolutionary perspective, the long childhood of sapiens appears important as a time for learning the skills and knowledge necessary to be an efficient adult forager The data obtained from our fieldwork shows that, contrary to the usual pattern in agricultural societies, hunter-gatherer parents place few labor demands on their children, although 314 lives were difficult, short, and dangerous, this would not be an exception compared to other early hominins, and there is no evidence that the life of a Neanderthal child was unusually “brutish” (Spikins et al 2014) Neanderthal children appear to have been buried with flint artifacts and animal bones more often than their adult counterparts, indicating children may have been given much attention, love, and care by adults The scarcity of data on the lives and behavior of children in the archaeological record may be due to researchers’ biased interest in the lives of adults (Shea 2006; Baxter 2005) It is necessary to expand our attention to children’s lives – especially to the social behavior of Neanderthal children – in order to make relevant comparisons between sapiens and Neanderthal childhoods Our research has also clarified the universality of teaching in humans Of the various modes of social learning, teaching seems to have special value since it takes place rarely in nonhuman animals and is conspicuously absent in higher nonhuman primates like chimpanzees, orangutans, and bonobos These animals show little or no teaching behavior in natural settings, while those raised by humans can demonstrate teaching behaviors Many philosophers and educators, from Jean-Jacques Rousseau and Immanuel Kant in the eighteenth century to contemporary cognitive psychologists such as Bateson (1972) and Tomasello (1999) and evolutionary philosopher Kim Sterelny (2012), assert that education, or teaching, is what “humanizes” children On the other hand, in ethnographic accounts of foragers, many authors have written that children were supposed to learn on one’s own and the parents did not feel any obligation to teach their offspring (Lancy 2010) Certainly, in hunting and gathering societies, we have recognized that children most often learn through observation and imitation, with little or no instruction from teachers in contrast to the formal education systems familiar in modern industrial societies Adults in hunting and gathering societies are usually reluctant to give instruction, much less orders, to their children in accordance with the high-value place on egalitarianism and autonomy (Hewlett, Chap 1) However, careful observation reveals that adults occasionally offer demonstration or instruction to children, sometimes through gestures such as pointing to draw attention to something or through evaluations of a child’s performance, scaffolding, and so on Such instruction may be given explicitly or implicitly in the course of daily interaction Teaching may also take place as a natural part of various communal activities in which children participate (Paradise and Rogoff 2009, Sonoda, Chap 9) Although teaching exacts some cost in terms of time and energy from those who teach, its universality in modern humans suggests its importance from an evolutionary perspective Teaching enables the transmission of opaque, complex, and difficult knowledge and skills with high fidelity and is one of the most important H Terashima factors in establishing cumulative culture (Csibra and Gergely 2006; Lewis and Laland 2012) Cognitive psychologists Csibra and Gergely (2006) claim that teaching or pedagogy is a special communication system for social learning and the evolutionary design of a basic human adaptation It is thus logical that teaching is embedded in every instance of social interaction and communication The universality of teaching alone, however, does not inform us much about our learning capacity Even in the twenty-first century, we cannot say we have a clear understanding of teaching or education Competition exists between a seemingly infinite range of teaching methods and education policies in modern industrial societies Within small-scale societies, ethnographers have documented any number of teaching strategies tailored to the specific cultural contexts in which they are practiced For example, Omura (Chap 23) observes that in Inuit society, adults begin to tease their 3–4year-old children with unanswerable questions and dilemmas as a way to teach them how to treat difficult situations in life with a calm mind Musharbash (Chap 14) describes how, in Aborigine societies, adults induce and then relieve fear in toddlers to help them understand their world and how they fit into it One important aspect of teaching is that it is more than what is taught Knowledge and skills the teacher intends to transfer are not being copied by the learner as they are, but are instead reconstructed by the learner based on inference and reference to past experience and existing knowledge base The fact that people can learn more than what they are taught characterizes human learning through teaching (Rancie`re 1991; Polanyi 2009[1967]) The effectiveness of teaching therefore depends as much or more on the active learning behavior of the learner than on the teacher’s intentions and effort As shown in many ethnographical studies, overt teaching effort is rather minimal in contemporary hunting and gathering societies Such a minimalist approach presupposes the learner’s active attitude, which might help enhance the learning capacity of the learner and lead to innovative learning in the future Although archaeological data that might provide insight into the learning behavior of Neanderthals are extremely scarce, we can make some inferences about learning in ancient hominins by looking at patterns in stone toolmaking It is possible to reconstruct stone-knapping techniques used in various archaeological sites through meticulous analysis of lithic flake refitting and patterns of ancient tool production sites In late Upper Paleolithic sites used by sapiens at E´tiolles in France and Shirataki in Hokkaido, Japan, it has been relatively easy for researchers to discern skill differences between highly skilled, intermediate, and novice knappers, who worked together in shared spaces (Pigeot 1990; Takakura 2013) In such working sites, it can be supposed that observation and imitation learning were 26 Reflections on Hunter-Gatherer Social Learning and Innovation practiced by knappers with lesser skills and teaching in the form of demonstration by experts (Shea 2006) Similarly, Stapert (2007) speculates that particular stone tools excavated from several Neanderthal sites may have been produced by children imitating adults’ techniques, as they seem to have been produced by individuals with relatively low levels of dexterity Spikins et al (2014) looked into knapping records in Arcy-sur-Cure, France, reported by Bodu (1990), and found evidence suggesting Neanderthal adults provided children with structured opportunities for learning, as experts, skilled learners, and novices seem to have worked together in close proximity Among contemporary hunter-gatherers, parents and older siblings often provide miniature forms of tools used in subsistence activities such as bows and arrows, hand axes, and digging sticks to children as toys (MacDonald 2010) This custom is thought to help children to get accustomed to the subsistence tools and learn how to use and make them A “micro-hand axe,” only 4.4 cm long and 17 g in weight, was excavated from a Neanderthal site in Rhenen in the Netherlands and supposed to be an instructive toy made by a father for his child (Stapert 2007) Spikins (2012) claims that learning how to make and use hand axes may have been part of adult sculpting of emotional self-control in children All in all, we see more similarity than difference between Neanderthals and ancient sapiens with regard to social learning and stone toolmaking techniques Furthermore, the patterns seen in archaeological remains left by both species also seem to be echoed strongly in the learning behavior observed in today’s hunter-gatherer children, who learn through observation and imitation, often guided by demonstrations and toys given by parents and older siblings Another point discussed in debates concerning the replacement of Neanderthals by sapiens is that of sociality and social networks Briefly, it has put forth that Neanderthal social relations were weaker than those of sapiens Because of their robust body structure and greater energy expenditure, they would have had to engage in large game hunting and live in small groups to avoid serious in-group competition (Kuhn 2013) The smallness of Neanderthal social networks is also suggested by aspects of their lithic culture In Neanderthals, raw materials suitable for lithic artifacts were mostly collected from within groups’ local neighborhoods, whereas sapiens often exploited materials brought from remote places via expansive social networks (Adler et al 2006) Neanderthals seem to have had the ability to exploit distant raw materials when necessary, but did so rarely because there was little or no motivation to so (Spikins et al 2014) On the other hand, ancient sapiens seem to have established large networks and utilized them not only for the circulation of rare materials but probably also for information exchange and to search for mates just as contemporary hunter-gatherers 315 26.3 Cognition, Innovation, and Evolution The nature of Neanderthal learning ability will remain uncertain as long as there is no hard evidence for it, and comparisons with sapiens are therefore still speculative It is also quite difficult to precisely define learning ability per se in both biological and cultural terms because, as Lyn Wadley (2013) maintains, there are recursive relationships between biological cognition, social behavior, and technology Wadley also notes that while some technologies may be a good indicator of complex cognition, that does not mean that cognition necessarily drove innovation in the past any more than it does today The relationship between learning ability and innovation is similar to that between genotype and phenotype, i.e., not a simple one-to-one relationship but one of complex and indirect connections (Sterelny 2001) Sapiens’ biologically based cognitive ability may appear generally to have increased steadily over the course of our evolution, but given the relatively sluggish pace of genetic evolution, it is difficult to suppose we have undergone any major biological advances in intelligence since the emergence of our species roughly 200,000 years ago There is of course massive variation in the pace of technological innovation across human societies The pace of scientific and technological innovation in the Global North has increased exponentially, especially since the industrial revolution of the eighteenth and nineteenth centuries in Europe On the other hand, there are throughout the world many peoples who continue life without any major technological innovations or change Contemporary huntergatherers who continue their traditional ways of life have been referred to by some as culturally and even biologically “inert” or “stationary.” This situation does not, however, have any basis in biology and is instead due to complex interactions between social, geographical, and historical contingencies as well as the ethical and emotional propensity common among foragers to live in the same way as their ancestors did Moreover, though hunter-gatherer societies tend to be remarkably conservative in terms of technological innovation and material culture, they demonstrate ample innovation and change in various nonmaterial domains such as aesthetics, symbolism, myths, music, dancing, and oral traditions (Takada, Chap 8, Peng, Chap 9, Lewis, Chap 12, Tsuru 2001) In addition, studies in ethnoscience have shown foragers continually and actively learn about and advance their knowledge of the natural world (Conklin 1954; Le´vi-Strauss 1962; Ichikawa and Terashima 1996) Human culture is characterized by a cumulative process that has led over many millennia to modern civilization, and the evolution of human learning ability through natural selection seems certain considering the changes in brain morphology and functions through time However, the path 316 of evolution is neither straightforward nor continuous, as it has neither a purpose nor a goal (Sterelny 2001) Le´viStrauss (1952) remarked that “advancing humanity can hardly be likened to appear climbing stairs and, with each movement, adding a new step to all those he has already mounted .it is only occasionally that history is ‘cumulative,’ that is to say, that the scores add up to a lucky combination.” The period of the replacement 50,000 years ago, which saw the intermingling of diverse foraging cultures as sapiens became dispersed throughout Europe, may have been one such rare and lucky combination Although evolution is a process of chance and is unpredictable, it is also true that through scientific investigation of the past, we are able understand something of the complex, interwoven chains of causality that have led to important evolutionary transitions The accumulation and rigorous scientific vetting of empirical and substantive knowledge are the only means we have of understanding the complicated pathway(s) of our evolution Social learning is one of the core drivers of human cultural evolution, and an understanding of the processes of learning and the contexts in which it occurs today can help us make sophisticated inferences about our past as well as provide clues about where we might be headed in the future References Adler DS, Bar-Oz G, Belfer-Cohen A, Bar-Yosef O (2006) Ahead of game: middle and upper palaeolithic hunting behaviors in the Southern Caucasus Curr Anthropol 47(1):89–118 Ambrose SH (2010) Coevolution of composite-tool technology, constructive memory, and language: implications for the evolution of modern human behavior Curr Anthropol 51(Supplement 1):S135– S147 Aoki Aoki K, Wakano JY, Lehmann L (2012) Evolutionarily stable learning schedules and cumulative culture in discrete generation models Theor Popul Biol 81:300–309 Bar-Yosef O (2013) Neanderthals and modern humans: across Eurasia In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans, vol 1, Cultural perspectives Springer Japan, Tokyo, pp 7–20 Bateson (1972) Steps to an ecology of mind The University of Chicago Press, Chicago Baxter JE (2005) The archaeology of childhood: children, gender, and material culture Altamira Press Walnut Creek, Walnut Creek Bodu P (1990) L’Application de la me´thod des remontages a` l’e´tude du mate´riel lithique des premiers niveaux chaˆtelperroniens de la Grotte du Renne a` Arcy-sur-Cure (Yonne) In: Farizy C (ed) Pale´olithique moyen et Pale´olithique supe´rieur ancien en Europe (Me´moires du Muse´e de Pre´hisoire d’Ile-de-France 3), (Nemours, APRAIF), pp 309–312 Bogin B (1997) Evolutionary hypotheses for human childhood Yearb Phys Anthropol 40:63–89 Boquet-Appel J-P, Tuffreau A (2009) Technological responses of Neanderthals to macroclimatic variations (240,000–40,000 BP) Hum Biol 81(2–3):287–307 Bower JM, Parsons LM (2003) Rethinking the “lesser brain” Sci Am 289(2):51–57 H Terashima Caspari R, Lee S-H (2004) Older age becomes common late in human evolution Proc Natl Acad Sci U S A 101(3):10895–10900 Chan WL, Abe-Ouchi A, O’ishi R and Takahashi K (2014) Stadial and interstadial climates in the late Pleistocene, as simulated in experiments with the MIROC climate model Akazawa T and Nishiaki Y (eds) RNMH 2014 the Second International Conference (Program and Abstracts), pp 157–158 Conklin, HC (1954) The relation of Hanuno´o culture to the plant world, PhD thesis, Yale University, New Haven Coolidge FL, Wynn T (2005) The working memory account of Neanderthal cognition: how phonological storage capacity may be related to recursion and the pragmatics of modern speech J Hum Evol 52:707–710 Coward F, Grove M (2011) Beyond the tools: social innovation and hominin evolution Paleoanthropology 2011:111–129 doi:10.4207/ PA.2011.ART46 Csibra G, Gergely G (2006) Social learning and social cognition: the case for pedagogy In: Munakata Y, Jonson MJ (eds) Processes of change in brain and cognitive development: attention and performance Oxford University Press, Oxford, pp 249–274 d’Errico F, Stringer CB (2011) Evolution, revolution or saltation scenario for the emergence of modern cultures? Phil Trans R Soc B 366:1060–1069 doi:10.1098/rstb.2010.0340 Green R et al (2010) A draft sequence of Neandertal genome Science 328:710–722 Hagino I, Yamauchi T (2014) Daily physical activity and time-space using of Pygmy hunter-gatherers’ children in southeast Cameroon In: Akazawa T, Ogihara N, Tababe HC, Terashima H (eds) Dynamics of learning in Neanderthals and modern humans, vol 2, Cognitive and physical perspectives Springer Japan, Tokyo, pp 91–97 Hawkes K, O’Connell JF, Blurton Jones N (1995) Hadza children’s foraging: juvenile dependency, social arrangements, and mobility among hunter-gatherers Curr Anthropol 36:688–700 Healy SD, Rowe C (2006) A critique of comparative studies of brain size Proc R Soc B (2007) 274, 453–464, doi: 10.1098/rspb.2006 3748 Published online 12 Dec 2006 Hewlett BL (2013) “Ekeloko” The spirit to create: innovation and social learning among Aka adolescents of the Central African Rainforest In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans: cultural perspectives Springer Japan, Tokyo, pp 1187–1195 Higham T (2011) European middle and upper palaeolithic radiocarbon dates are often older than they look: problems with previous dates and some remedies Antiquity 85:235–249 Higham et al (2010) The chronology of the Grotte du Renne (France) and implications for the association of ornaments and human remains within the Chaˆtelperronian Proc Natl Acad Sci U S A 107(47):20234–20239 Higham T et al (2014) The timing and spatiotemporal patterning of Neanderthal disappearance Nature 512:306–309 Ichikawa M, Terashima H (1996) Cultural diversity in the use of plants by Mbuti hunter-gatherers in northeastern Zaire: an ethnobotanical approach In: Kent S (ed) Cultural diversity among twentiethcentury foragers: an African perspective Cambridge University Press, Cambridge, pp 276–293 Imamizu H, Kawato M (2012) Cerebellar internal models: implications for the dexterous use of tools Cerebellum 11:325–335 doi:10 1007/s12311-010-0241-2 Kadowaki S (2013) Issues of chronological and geographical distributions of middle and upper palaeolithic cultural variability in the Levant and implications for the learning behavior of Neanderthals and homo sapiens In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans, vol 1, Cultural perspectives Springer Japan, Tokyo, pp 59–91 Kadowaki S (2014) Shoki Homo sapiensu no gakushu koudou: Afurika to nishi Ajia no kouko kiroku ni motodzuku kousatsu (Learning 26 Reflections on Hunter-Gatherer Social Learning and Innovation 317 behavior of early Homo sapiens: an inquiry based on archaeological evidence in Africa and west Asia) In: Nishiaki Y (ed) Homo sapiensu to Neanderutaru (Homo sapiens and Neanderthals), vol 2, Koukogaku kara mita gakushu kodo (Learning behavior viewed from archaeological perspective) Rokuichi Shobo, Tokyo, pp 3–18 (in Japanese) Kadowaki S, Omori T, Nishiaki Y (2015) Variability in early Ahmarian lithic technology and its implications for the model of a Levantine origin of the protoaurignacian J Hum Evol 82:67–87 Kamei N (2005) Play among Baka children in Cameroon In: Hewlett BS, Lamb ME (eds) Hunter-gatherer childhoods: evolutionary, developmental and cultural perspectives Aldine Transaction, New Brunswick, pp 343–359 Klein RG, Edger B (2002) The dawn of human culture Wiley, New York Kondo O, Kubo D, Suzuki H, Ogihara N (2014) Virtual endocast of qafzeh 9: a preliminary assessment In: Akazawa T, Ogihara N, Tababe HC, Terashima H (eds) Dynamics of learning in Neanderthals and modern humans, vol 2, Cognitive and physical perspectives Springer Japan, Tokyo, pp 183–190 Kubo D et al (2014a) Cerebellar size estimation from endocranial measurements: an evaluation based on MRI data In: Akazawa T, Ogihara N, Tababe HC, Terashima H (eds) Dynamics of learning in Neanderthals and modern humans, vol 2, Cognitive and physical perspectives Springer Japan, Tokyo, pp 209–215 Kubo D, et al (2014b) Estimating the cerebral and cerebellar volumes of Neanderthal and Middle and Upper Paleolithic Homo Sapiens, In: Akazawa T, Nishiaki Y (eds) (2014) RNMH 2014 The second international conference, program and abstracts, pp 116–118 Kuhn SL (2013) Cultural transmission, institutional continuity and the persistence of the Mousterian In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans: cultural perspectives Springer Japan, Tokyo, pp 105–113 Kuhn SL, Stiner MC (2006) What’s a mother do? The division of labor among Neanderthals and modern humans in Eurasia Curr Anthropol 47(6):953–980 Lancy DF (2010) Learning ‘From Nobody’: the limited role of teaching in folk models of children’s development Child Past 3:79–106 Langley MC, Clarkson C, Ulmet S (2008) Behavioural complexity in Eurasian Neanderthal populations: a chronological examination of the archaeological evidence Camb Archaeol J 18(3):289–307 Lehmann L, Wakano JY, Kenichi A (2013) On optimal learning schedules and the marginal value of cumulative cultural evolution Evolution 67–5:1435–1445 Le´vi-Strauss C (1952) Race and history UNESCO, Paris Le´vi-Strauss C (1962) La Panse´e Sauvage Librairie Plon, Paris Lewis HM, Laland KN (2012) Transmission fidelity is the key to the build-up of cumulative culture Phil Trans R Soc B 367:2171–2180 doi:10.1098/rstb.2012.0119 MacDonald K (2010) Learning to hunt In: Lancy DF, Bock J, Gaskins S (eds) The anthropology of learning in childhood AltaMira Press, Plymouth, pp 371–396 Martı´n-Loeches M (2010) Use and abuse of the enhanced-workingmemory hypothesis in explaining modern thinking Curr Anthropol 51(S1):S67–S75 Mcbrearty S, Brooks AS (2000) The revolution that wasn’t: a new interpretation of the origins of modern human behavior J Hum Evol 39:453–563 Mellars P (2006) Why did modern human populations disperse from Africa ca 60,000 years ago? A new model PNAS 103 (25):9381–9386 Mesoudi A, Aoki K (eds) (2015) Learning strategies and cultural evolution during the palaeolithic Springer Japan, Tokyo Mithen S (1996) The prehistory of the mind: a search for the origins of art, religion and science Thames and Hudson Ltd, London Nishiaki Y (2013) “Gifting” as a means of cultural transmission: the archaeological implications of bow-and-arrow technology in Papua New Guinea In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans, vol 1, Cultural perspectives Springer Japan, Tokyo, pp 173–186 Nishiaki Y (2014) Kyujin-shinjin no gakushu kodo wo meguru shomondai (Problems about learning behavior in archaic humans and modern humans) In: Nishiaki Y (ed) Homo sapiensu to Neanderutaru (Homo sapiens and Neanderthals), vol 2, Koukogaku kara mita gakushu (Learning viewed from Archaeology) Rokuichi Shobo, Tokyo, pp 175–185 (in Japanese) Ogihara N, Morita Y, Amano H, Kondo O, Suzuki H, Nakatsukasa M (2014) Application of sliding landmark method for morphological analysis of modern Japanese neurocranial shape In: Akazawa T, Ogihara N, Tababe HC, Terashima H (eds) Dynamics of learning in Neanderthals and modern humans, vol 2, Cognitive and physical perspectives Springer Japan, Tokyo, pp 145152 Paăaăbo S (2014a) The human condition: a molecular approach Cell 157:216226 Paăaăbo S (2014b) Neanderthal man: in search of lost genomes Basic Books, New York Paradise R, Rogoff B (2009) Side by side: learning by observing and pitching in Ethos 37(1):102–138 Pearce E, Stringer C, Dunbar RIM (2013) New insights into differences in brain organization between Neanderthals and anatomically modern humans Proc R Soc B 280:20130168, http://dx.doi.org/10 1098/rspb.2013.0168 Penhune VB, Steele CJ (2012) Parallel contributions of cerebellar, striatal and M1 mechanisms to motor sequence learning Behav Brain Res 226:579–591 Pigeot N (1990) Technical and social actors: flintknapping specialists and apprentices at Magdalenian etiolles Archaeol Rev Camb (1):126–141 Polanyi M (2009[1967]) The tacit dimension Chicago University Press, Chicago Powell A, Shennan S, Thomas M (2009) Late Pleistocene demography and appearance of modern human behavior Science 324:1298–1301 Rancie`re J (1991) The ignorant schoolmaster: five lessons in intellectual Emancipation (Translated by Kristin Ross) Stanford University Press, Stanford Sankararaman S, Patterson N, Li H, Paăaăbo S, Reich D (2012) The date of interbreeding between Neandertals and modern humans PLoS Genet 8(10):e1002947 doi:10.1371/journal.pgen.1002947 Sano K (2012) Yoroppa ni okeru kyusekki bunka hennen to kyujinshinjin koutaigeki (Chronology of Palaeolithic cultures in Europe and the replacement of Neanderthals by modern humans) In: Nishiaki Y (ed) Homo sapiensu to Neanderutaru (Homo sapiens and Neanderthals), vol 2, Kyusekki koukogaku kara mita koutaigeki (The replacement viewed from palaeolithic archaeology) Rokuichi Shobo, Tokyo, pp 38–56 (in Japanese) Sano K (2013) Yoroppa kyujin iseki ni miru gakushu no shoko (The evidence of learning behavior at Neanderthal sites in Europe) In: Nishiaki Y (ed) Homo sapiensu to Neanderutaru (Homo sapiens and Neanderthals), vol 2, Koukogaku kara mita gakushu (The learning behavior viewed from archaeology) Rokuichi Shobo, Tokyo, pp 19–27 (in Japanese) Sano K, Omori T (2015) Yoroppa ni okeru kyujin-shinjin no koutaigeki no purosesu (The process of replacement of Neanderthals by modern humans in Europe) In: Nishiaki Y (ed) Homo sapiensu to Neanderutaru (Homo sapiens and Neanderthals), vol 3, Hito to bunka no koutaigeki (The replacement of humans and cultures) Rokuichi Shobo, Tokyo, pp 20–35 (in Japanese) Shea JJ (2006) Child’s play: reflections on the invisibility of children in the palaeolithic record Evol Anthropol 15:212–216 318 Shennan S (2001) Demography and cultural innovation: a model and its implications for the emergence of modern human culture Camb Archaeol J 11(1):5–16 Smith TM et al (2010) Dental evidence for ontogenetic differences between modern humans and Neanderthals Proc Natl Acad Sci U S A 107(49):20923–20928 Snodgrass JJ, Leonard WR (2009) Neandertal energetics revisited: insights into population dynamics and life history evolution Paleoanthropology 2009:220–237 Sørensen MV, Leonard WR (2001) Neandertal energetics and foraging efficiency J Hum Evol 40:483–495 Spikins PA (2012) Goodwill hunting? Debates over the meaning of handaxe from revisited World Archaeol 44(3):378–392 Spikins PA, Hitches G, Needham A, Rutherford H (2014) The cradle of thought: growth, learning, play and attachment in Neanderthal children Oxf J Archaeol 33(2):111–134 doi:10.1111/ojoa.12030 Stapert D (2007) Neanderthal children and their flints PalArch’s J Archaeol Northwest Eur 1(2):16–39 Sterelny K (2001) Dawkins vs Gould: survival or the fittest Icon Books Ltd, London Sterelny K (2012) The evolved apprentice: how evolution made humans unique The MIT Press, Cambridge, MA Takakura J (2013) Using lithic refitting to investigate the skill learning process: lessons from upper Paleolithic assemblages at the Shirataki sites in Hokkaido, northern Japan In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans: cultural perspectives Springer Japan, Tokyo, pp 151–171 Tanabe HC, Kochiyama T, Ogihara N, Sadato N (2014a) Integrated analytical scheme for comparing the Neanderthal brain to modern human brain using neuroimaging techniques In: Akazawa T, Ogihara N, Tababe HC, Terashima H (eds) Dynamics of learning in Neanderthals and modern humans, vol 2, Cognitive and physical perspectives Springer Japan, Tokyo, pp 203–207 Tanabe HC, Kochiyama T, Sadato N, Ogihara N (2014b) Exploring the difference of brain anatomy and function between Neanderthals and Modern humans: Neuroanatomical and functional neuroimaging approach In: Akazawa T, Nishiaki Y (eds) RNMH 2014 The second international conference (Program and Abstracts), pp 121–123 H Terashima Terashima H (2013) The evolutionary development of learning and teaching strategies in human societies In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans: cultural perspectives Springer Japan, Tokyo, pp 141–150 Tomasello M (1999) The cultural origins of human cognition Harvard University Press, Cambridge Tsuru D (2001) Generation and transition processes in the spirit ritual of the Baka Pygmies in southeastern Cameroon Afr Study Monogr Suppl 27:103–123 Villa P, Roebroeks W (2014) Neandertal demise: an archaeological analysis of the modern human superiority complex PLoS One 9(4): e96424 doi:10.1371/journal.pone.0096424 Wadley L (2013) Recognizing complex cognition through innovative technology in stone age and palaeolithic sites Camb Archaeol J 23 (02):163–183 Wakano JY, Miura C (2014) Trade-off between learning and exploitation: the Pareto-optimal versus evolutionarily stable learning schedule in cumulative cultural evolution Theor Popul Biol 91:37–43 Wynn T, Coolidge FL (2004) The expert Neanderthal mind J Hum Evol 46:467–487 Wynn T, Coolidge FL (2007) Neanderthals became extinct while Homo sapiens prospered: a marked contrast in mental capacities may account for these different fates Am Sci 96:44–51 Wynn T, Coolidge FL (2011) The implications of the working memory model for the evolution of modern cognition Int J Evol Biol 2011, 741357 doi:10.4061/2011/741357 Zilha˜o J (2007) The emergence of ornaments and art: an archaeological perspective on the origins of “behavioral modernity.” J Archaeol Res 15:1–54 Zilha˜o J (2013) Neandertal-modern human contact in western Eurasia: issues of dating, taxonomy, and cultural associations In: Akazawa T, Nishiaki Y, Aoki K (eds) Dynamics of learning in Neanderthals and modern humans, vol 1, Cultural perspectives Springer Japan, Tokyo, pp 21–57 Zubrow E (1989) The demographic modelling of Neanderthal extinction In: Mellars, Stringer (eds) The human revolution: behavioral and biological perspectives on the origins of modern humans Princeton University Press, Princeton ... Japan Contributors Social Learning and Innovation in HunterGatherers Barry S Hewlett Abstract This chapter provides an introduction to social learning and innovation in hunter- gatherers, summarizes... societies and that in hunting and gathering societies in the past as well as present Our study of social learning has been conducted mainly among contemporary hunter gatherers in various natural and social. .. archaeology and ethnoarchaeology were also conducted to interpret ancient traces of learning A02 investigated contemporary hunter gatherers learning behavior, their social and individual learning, mainly

Ngày đăng: 14/05/2018, 12:40

Từ khóa liên quan

Mục lục

  • Preface

    • The RNMH Project and the Study of Social Learning in Modern Hunter-Gatherers

    • Contents

    • Contributors

    • 1: Social Learning and Innovation in Hunter-Gatherers

      • 1.1 Introduction

        • 1.1.1 Why Hunter-Gatherers?

        • 1.1.2 What Is Social Learning?

        • 1.1.3 Why Children?

        • 1.2 The Social-Cognitive Learning Environment of Hunter-Gatherers

          • 1.2.1 Foundational Schema

          • 1.2.2 Physical and Social Setting: Demography of Forager Social Learning

          • 1.2.3 Social-Emotional Setting: Cultural Practices that Impact Social Learning

            • 1.2.3.1 Intimacy

            • 1.2.3.2 Self-Directed

            • 1.2.3.3 Trust of Others

            • 1.2.3.4 Play

            • 1.3 From Whom and How Do Children Learn?

              • 1.3.1 From Whom Do Children Learn?

                • 1.3.1.1 Previous Hunter-Gatherer Studies on from Whom Children Learn

                • 1.3.2 How Do Children Learn?

                  • 1.3.2.1 Evolutionary Approaches

                  • 1.3.2.2 Social-Cultural Anthropology and Participatory Approaches

                  • 1.3.2.3 Social Learning and Play

                  • 1.3.2.4 Social Learning and Embodiment

                  • 1.3.2.5 Social Learning and Language

                  • 1.3.2.6 Indigenous Approaches to Social Learning

                  • 1.4 Innovation

Tài liệu cùng người dùng

Tài liệu liên quan