Поиск по сайту




Пишите нам: info@ethology.ru

Follow etholog on Twitter

Система Orphus

Новости
Библиотека
Видео
Разное
Кросс-культурный метод
Старые форумы
Рекомендуем
Не в тему

список статей


Branching, blending, and the evolution of cultural similarities and differences among human populations

Mark Collardad, Stephen J. Shennanbd, Jamshid J. Tehranicd

1. Introduction

2. Materials and methods

3. Results

4. Discussion

5. Conclusions

Acknowledgment

References

Copyright

1. Introduction

The processes responsible for producing the cultural similarities and differences among human populations have long been the focus of debate in the social sciences, as has the corollary issue of linking cultural data with the patterns reconstructed by historical linguists and by biologists working with human populations (e.g., Bellwood, 1996, Boas, 1940, Boyd & Richerson, 1985, Cavalli-Sforza & Feldman, 1981, Cavalli-Sforza et al., 1988, Durham, 1991, Goodenough, 1999, Hurles et al., 2003, Jones, 2003, Kirch & Green, 1987, Kroeber, 1948, Lumsden & Wilson, 1981, Mesoudi et al., 2004, Moore, 1994, Morgan, 1870, Petrie, 1939, Renfrew, 1987, Renfrew, 1992, Rivers, 1914, Romney, 1957, Schmidt, 1939, Smith, 2001, Smith, 1933, Terrell, 1988, Welsch et al., 1992, Whaley, 2001). Currently, debate is focused on two competing hypotheses, which have been termed the branching hypothesis (also known as the “genetic,” “demic diffusion,” or “phylogenesis” hypothesis) and the blending hypothesis (also known as the “cultural diffusion” or “ethnogenesis” hypothesis; Bellwood, 1996, Collard & Shennan, 2000, Guglielmino et al., 1995, Hewlett et al., 2002, Kirch & Green, 1987, Moore, 1994, Moore, 2001, Romney, 1957, Tehrani & Collard, 2002). Other models have been proposed (e.g., Boyd, Borgerhoff Mulder, Durham, & Richerson, 1997), but to date, these have received little attention in the literature.

According to the branching hypothesis, cultural similarities and differences among human populations are primarily the result of a combination of within-group information transmission and population fissioning. The strong version of the hypothesis suggests that Transmission Isolating Mechanisms, or TRIMS (Durham, 1992), impede the transmission of cultural elements among contemporaneous communities. TRIMS are akin to the barriers to hybridisation that separate species and include language differences, ethnocentrism, and intercommunity violence (Durham, 1992). The branching hypothesis predicts that the similarities and differences among cultures can be best represented by the type of branching tree diagram that is used in biology to depict the relationships among species (Fig. 1). The hypothesis also predicts that there will be a strong association between cultural variation and linguistic and biological patterns (e.g., Ammerman & Cavalli-Sforza, 1984, Bellwood, 1996, Bellwood, 2001, Cavalli-Sforza et al., 1994, Cavalli-Sforza et al., 1988, Diamond & Bellwood, 2003, Kirch & Green, 1987, Kirch & Green, 2001, Renfrew, 1987).


View full-size image.

Fig. 1. Example cladogram.


In contrast, supporters of the blending hypothesis (e.g., Dewar, 1995, Moore, 1994, Moore, 2001, Terrell, 1988, Terrell, 2001, Terrell et al., 1997, Terrell et al., 2001) believe that it is unrealistic “to think that history is patterned like the nodes and branches of a comparative, phylogenetic, or cladistic tree” (Terrell et al., 1997, p. 184). The appropriate way to represent relationships among human populations, according to this view, is not as a branching tree but as a braided stream, with different channels flowing into one another, then splitting again. The basis of this argument is that humans have always interacted, and thus ideas, innovations, goods, and cultural practices, not to mention genes, have constantly flowed from one community to another. To the extent that language is an exception to this, it is because of the mutual accommodation of individuals' idiolects to one another that is required if speakers are to understand each other. The blending hypothesis predicts that the similarities and differences among cultures can best be represented by a maximally connected network or reticulated graph (Terrell, 2001). It also predicts that there will be a close relationship between cultural patterns and the frequency and intensity of contact among populations, the usual proxy of which is geographic proximity.

It has been asserted that blending has been the dominant cultural evolutionary process in the ethnohistorical period and is likely to have always been more significant than branching in cultural evolution (e.g., Dewar, 1995, Moore, 1994, Moore, 2001, Terrell, 1988, Terrell, 2001, Terrell et al., 1997, Terrell et al., 2001). The pervasiveness of human interaction obviously cannot be denied. In the words of Bellwood (1996, p. 882), “humans flourish in interactive groups, and total isolation of any human group has been very rare in prehistory.” However, in our view, whether blending is the dominant cultural process is open to question. First, the archaeological record frequently demonstrates the existence of long-lasting cultural traditions with recognisable coherence, despite evidence for the extensive movement of materials and artifacts across boundaries (e.g., Pétrequin, 1993). Second, ethnographic work indicates that in noncommercial settings, cultural transmission is often both vertical and conservative, with children learning skills from their parents with relatively little error (e.g., Childs & Greenfield, 1980, Greenfield, 1984, Greenfield et al., 2000, Hewlett & Cavalli-Sforza, 1986, Shennan & Steele, 1999). Third, recent work in psychology suggests that humans may possess evolved cognitive mechanisms that lead them to interact preferentially with individuals who are similar to themselves (Buston & Emlen, 2001) and to be prejudiced against individuals from unfamiliar ethnic groups (Gil-White, 2001, Schaller et al., 2003). Fourth, empirical and theoretical research suggests that, as counterintuitive as it may seem, interaction between people can actually lead to the emergence of cultural barriers and distinctions where none previously existed (e.g., Barth, 1969, Hodder, 1982, McElreath et al., 2003). Lastly, most contributions to the branching/blending debate published to date have focused on cultural evolution in specific regions of the world often over relatively short spans of time rather than dealing with it as a general phenomenon (e.g., Borgerhoff Mulder, 2001, Collard & Shennan, 2000, Guglielmino et al., 1995, Hewlett et al., 2002, Jordan & Shennan, 2003, Kirch & Green, 1987, Shennan & Collard, 2005, Tehrani & Collard, 2002, Welsch et al., 1992). A few papers have addressed the debate's key issues in global terms, but in these works either the evidence discussed is anecdotal (e.g., Moore, 1994, Moore, 2001, Terrell, 1988, Terrell, 2001) or the analyses reported are informal (e.g., Jones, 2003). As such, it is currently unclear from an empirical perspective whether cultural evolution is dominated by branching or blending processes.

In this paper, we report a study that goes some way towards rectifying the latter situation. In the study, we assessed how tree-like patterns in cultural data sets are compared with patterns in biological data sets. Essentially, we fitted the bifurcating tree model that biologists use to represent the relationships of species to a group of data sets pertaining to cultural phenomena such as artifacts and rituals, and to a group of biological data sets that have been used to reconstruct the relationships of species and higher level taxa. We then compared the average fit between the cultural data sets and the model with the average fit between the biological data sets and the model. Given that the biological data sets can be confidently assumed to have been structured by speciation, which is a branching process, our assumption was that, if the blending hypothesis is correct and cultural evolution is dominated by blending processes, the fit between the bifurcating tree model and the cultural data sets should be significantly worse than the fit between the bifurcating tree model and the biological data sets. Conversely, if the blending hypothesis is incorrect and cultural evolution is dominated by branching processes, the fit between the bifurcating tree model and the cultural data sets should be no worse than the fit between the bifurcating tree model and the biological data sets.

2. Materials and methods

Our first step was to obtain biological and cultural data sets suitable for phylogenetic analysis. Acquiring the biological data sets was straightforward, as they are readily available in the literature and many of them can also be downloaded from on-line databases, such as TreeBASE (Sanderson, Donoghue, Piel, & Eriksson, 1994). Accordingly, we were able to assemble a group of 21 biological data sets. An effort was made to include a broad range of taxa and characters. Thus, the biological data sets included DNA data for lizards, lagomorphs, and carnivores, morphological data for fossil hominids, seals, and ungulates, and behavioural data for bees, seabirds, and primates. Currently, cultural data sets suitable for phylogenetic analysis are much less easy to come by than their biological counterparts. We had six data sets in our possession from previous work conducted on this topic by researchers associated with the AHRB Centre for the Evolutionary Analysis of Cultural Behaviour (Collard & Shennan, 2000, Croes et al., 2005, Jordan & Shennan, 2003, Tehrani, 2004, Tehrani & Collard, 2002, Venti, 2004). To these, we were able to add 14 data sets from the literature (Barnett, 1937, Barnett, 1939, Driver, 1937, Drucker, 1937, Drucker, 1941, Drucker, 1950, Gifford, 1940, Gifford & Kroeber, 1937, Jorgensen, 1969, Moylan et al., in press, Steward, 1941, Stewart, 1941, Welsch et al., 1992, O'Brien et al., 2001). This gave us a total of 20 cultural data sets to work with. Details of the biological and cultural data sets are provided in Table 1, Table 2, respectively. Copies of the NEXUS files will be made available on request.

Table 1.

Biological data sets analysed in this study

Data set Source Notes
Austalasian teal mtDNA Kennedy and Spencer (2000) Downloaded from TreeBASE. Data for ATPase 6, ATPase 8 and 12S genes.
Corbiculate bee behaviour Noll (2002)
Pelecaniforme bird behaviour Kennedy, Spencer, and Gray (1996)
Anoles lizards morphology Guyer and Savage (1986) Downloaded from TreeBASE.
Primate behaviour DiFiore and Rendall (1994)
Strepsirhine primate morphology Yoder (1994)
Fossil hominid morphology Lieberman, Wood, and Pilbeam (1996)
New World monkey morphology Horowitz, Zardoya, and Meyer (1998) Craniodental data.
Ungulate morphology O'Leary and Geisler (1999) Downloaded from TreeBASE. Data from Runs 5 and 6.
Phalacrocoracid bird mtDNA Kennedy, Gray, and Spencer (2000) Provided by Martyn Kennedy, University of Otago. Data for 12S, ATPase 6 and 8 genes.
Phocid seal morphology Bininda-Edwards and Russell (1996) Downloaded from TreeBASE.
Hawaiian fruit fly mtDNA Baker and DeSalle (1997) Downloaded from TreeBASE. Data from “all genes” analysis.
Hominoid primate cranial morphology Collard and Wood (2000) Qualitative data set.
Carnivore mtDNA Wayne et al. (1997) Downloaded from TreeBASE.
Mammal mtDNA with emphasis on Malagasy primates Yang and Yoder (2003) Downloaded from the website of Anne Yoder, Yale University. Data for COII and cytochrome b genes.
Carnivore mtDNA with emphasis on Malagasy taxa Yoder et al. (2003) Downloaded from the website of Anne Yoder, Yale University. Data for cytochrome b gene.
Mammal mtDNA Yoder and Yang (2000) Downloaded from the website of Anne Yoder, Yale University.
Insectivore mtDNA Stanhope et al. (1998) Downloaded from TreeBASE. Data for 12S-16S genes.
Lagomorph mtDNA Halanych and Robinson (1999) Downloaded from TreeBASE. Data for 12S gene.
Hominoid primate soft-tissue morphology Gibbs, Collard, and Wood (2002)
Anolis lizard mtDNA Jackman, Larson, de Queiroz, and Losos (1999) Downloaded from TreeBASE. Data for ND2 and tRNA genes.
Table 2.

Cultural data sets analysed in this study

Data set Source Notes
Gulf of Georgia Salish food taboos and prescriptions Barnett (1939)
Neolithic pottery Collard and Shennan (2000)
Californian Indian basketry Jordan and Shennan (2003)
Eastern North American projectile points O'Brien et al. (2001)
Coast and inland Salish cultural practices Jorgensen (1969)
New Guinea material culture Welsch et al. (1992)
Turkmen weaving designs Tehrani and Collard (2002)
Northwest Coast tribal religion and ritual Drucker (1950)
Early Christian doctrinal beliefs Venti (2004)
Iranian tribal weavings Tehrani (2004)
Northwest Coast archaeology Croes et al. (2005) Stone, bone-antler, and shell artifacts
Pomo structures Gifford and Kroeber (1937)
Oregon Coast tribal puberty rites Barnett (1937)
Southern Sierra Nevada tribal death and mourning practices Driver (1937)
Nevada Shoshoni tribal mutilations Steward (1941)
Southern California tribal body- and dress-related practices Drucker (1937)
Yuman-Piman warfare-related practices Drucker (1941)
Apache-Pueblo houses Gifford (1940)
African cultural practices Moylan et al. (in press) Downloaded from the website of Monique Borgerhoff Mulder, University of California-Davis.
Northern Paiute birth rituals Stewart (1941)

Next, we measured the fit between the bifurcating tree model and each data set. To do so, we employed an analytical approach from evolutionary biology known as “phylogenetic systematics” or, more commonly, “cladistics.” First presented coherently in the 1950s (Hennig, 1950), cladistics is now the dominant method of phylogenetic reconstruction used in biology (Kitching et al., 1998, Schuh, 2000). Based on a null model in which new taxa arise from the bifurcation of existing ones, cladistics define phylogenetic relationship in terms of relative recency of common ancestry. Two taxa are deemed to be more closely related to one another than either is to a third taxon if they share a common ancestor that is not also shared by the third taxon. The evidence for exclusive common ancestry is evolutionarily novel or “derived” character states. Two taxa are inferred to share a common ancestor to the exclusion of a third taxon if they exhibit derived character states that are not also exhibited by the third taxon. In its simplest form, cladistic analysis proceeds via four steps. First, a character state data matrix is generated. This shows the states of the characters exhibited by each taxon. Next, the direction of evolutionary change among the states of each character is established. Several methods have been developed to facilitate this, including communality analysis (Eldredge & Cracaft, 1980), ontogenetic analysis (Nelson, 1978), and stratigraphic sequence analysis (Nelson & Platnick, 1981). Currently the favoured method is outgroup analysis (Arnold, 1981). Outgroup analysis entails examining a close relative of the study group. When a character occurs in two states among the study group, but only one of the states is found in the outgroup, the principle of parsimony is invoked and the state found only in the study group is deemed to be evolutionarily novel with respect to the outgroup state. Having determined the probable direction of change for the character states, the next step in a cladistic analysis is to construct a branching diagram of relationships for each character. As shown in Fig. 1, this is done by joining the two most derived taxa by two intersecting lines, and then successively connecting each of the other taxa according to how derived they are. Each group of taxa defined by a set of intersecting lines corresponds to a clade, and the diagram is referred to as a cladogram or tree. The final step in a cladistic analysis is to compile an ensemble cladogram from the character cladograms. Ideally, the distribution of the character states among the taxa will be such that all the character cladograms imply relationships among the taxa that are congruent with one another. Normally, however, a number of the character cladograms will suggest relationships that are incompatible. This problem is overcome by generating an ensemble cladogram that is consistent with the largest number of characters and therefore requires the smallest number of ad hoc hypotheses of character appearance or “homoplasies” to account for the distribution of character states among the taxa.

We identified the most parsimonious cladogram for each data set with the aid of the popular phylogenetics computer program PAUP* 4 (Swofford, 1998). In all the analyses, the characters were treated as unordered, and the most parsimonious cladogram was detected via the heuristic search routine. We then used PAUP* 4 to evaluate how well the most parsimonious cladogram explains the distribution of similarities and differences within each data set. The goodness-of-fit measure we used was the retention index, or RI, of Farris, 1989a, Farris, 1989b. Equivalent to the Archie's (1989) homoplasy excess ratio maximum index (Farris, 1989b, Farris, 1991, Archie, 1989), the RI is a measure of the number of homoplastic changes that a cladogram requires that are independent of its length (Farris, 1989a, Farris, 1989b). The RI of a single character is calculated by subtracting the number of character state changes required by the focal cladogram (s) from the maximum possible amount of change required by a cladogram in which all the taxa are equally closely related (g). This figure is then divided by the result of subtracting the minimum amount of change required by any conceivable cladogram (m) from g. The RI of two or more characters is computed as (GS)/(GM), where G, S, and M are the sums of the g, s, and m values for the individual characters. A maximum RI of 1 indicates that the cladogram requires no homoplastic change, and the level of homoplasy increases as the index approaches 0. The RI is a useful goodness-of-fit measure when comparing data sets because, unlike some other measures (e.g., the Consistency Index), it is not affected by number of taxa or number of characters. The RIs for the 21 biological data sets and the 20 cultural data sets are presented in Table 3. Also shown in Table 3 is an RI associated with the most parsimonious cladogram obtained by Darwent and O'Brien (in press) from a Northeastern Missouri projectile point data set.

Table 3.

Goodness-of-fit values associated with most parsimonious cladograms derived from 21 biological and 21 cultural data sets

Data set NT NC PI RI Dataset NT NC PI RI
Austalasian teal mtDNA 7 1172 73 0.94 Gulf of Georgia Salish food taboos and prescriptions 11 77 51 0.57
Corbiculate bee behaviour 23 42 41 0.94 Neolithic pottery 59 35 33 0.71
Pelecaniforme bird behaviour 20 37 36 0.84 Californian Indian basketry 40 219 184 0.71
Anoles lizards morphology 24 18 16 0.79 Eastern North American projectile points 17 8 6 0.70
Primate behaviour 38 34 34 0.73 Coast and inland Salish cultural practices 29 78 75 0.63
Strepsirhine primate morphology 29 43 43 0.72 New Guinea material culture 31 47 47 0.51
Fossil hominid morphology 9 48 48 0.71 Turkmen weaving designs 6 90 56 0.44
New World monkey morphology 20 76 65 0.70 Northwest Coast tribal religion and ritual 18 220 137 0.65
Ungulate morphology 40 123 122 0.70 Early Christian doctrinal beliefs 12 18 15 0.61
Phalacrocoracid bird mtDNA 24 1141 234 0.65 Iranian tribal weavings 10 110 92 0.60
Phocid seal morphology 27 196 184 0.60 Northwest Coast archaeology 48 69 69 0.50
Hawaiian fruit fly mtDNA 17 2550 501 0.50 Pomo structures 20 43 31 0.52
Hominoid primate cranial morphology 6 96 57 0.49 Oregon Coast tribal puberty rites 10 109 39 0.55
Carnivore mtDNA 25 2001 615 0.47 Southern Sierra Nevada tribal death and mourning practices 23 181 138 0.48
Mammal mtDNA with emphasis on Malagasy primates 36 1812 932 0.47 Nevada Shoshoni tribal mutilations 19 48 22 0.78
Carnivore mtDNA with emphasis on Malagasy taxa 35 1140 498 0.47 Southern California tribal body- and dress-related practices 18 98 78 0.52
Mammal mtDNA 31 10806 6049 0.44 Yuman-Piman warfare-related practices 8 185 110 0.69
Insectivore mtDNA 43 2086 866 0.44 Apache-Pueblo houses 20 140 120 0.63
Lagomorph mtDNA 12 739 97 0.39 African cultural practices 35 54 54 0.42
Hominoid primate soft-tissue morphology 5 171 154 0.38 Northern Paiute birth rituals 14 128 86 0.43
Anolis lizard mtDNA 55 1456 866 0.35 Northeastern Missouri projectile points 22 13 ? 0.66

NT, number of taxa; NC, number of characters; PI, number of parsimony informative characters; RI, Retention Index. A maximum RI of 1 indicates that the cladogram requires no homoplastic change, and the level of homoplasy increases as the index approaches 0.

In the next stage of the study, we compared the RIs of the 21 biological data sets with the 21 cultural RIs with a view to determining whether they are significantly different. This was accomplished with the Mann–Whitney U test function of SPSS 11.

3. Results

The RIs associated with the most parsimonious cladograms derived from the biological and cultural data sets (Table 2) suggest that the fit between the bifurcating tree model and the cultural data sets is little different from the fit between the bifurcating tree model and the biological data sets. Not only are the averages similar, but also the ranges are comparable. The mean, minimum, and maximum biological RIs are 0.61, 0.35, and 0.94, respectively. The corresponding figures for the cultural RIs are 0.59, 0.42, and 0.78. Thus, the descriptive statistics do not support the hypothesis that blending is more important than branching in cultural evolution. On average, the cultural data sets appear to be no more reticulate than the biological data sets.

The result of the Mann–Whitney U test is in line with the descriptive statistics. The biological and cultural RIs are not significantly different according to the test (Mann–Whitney U=215.5, p=.900). Thus, once again, the hypothesis that blending is more important than branching in cultural evolution is not supported.

4. Discussion

To evaluate the assertion that blending has always been a more important cultural evolutionary process than branching is, we fitted the bifurcating tree model that biologists use to represent relationships among species to a set of cultural data sets and to a set of biological data sets that have been used to reconstruct the relationships of species and higher level taxa. We then compared the average fit between the cultural data sets and the model with the average fit between the biological data sets and the model. What we found was that the goodness-of-fit measures derived from the cultural data sets were not significantly different from the goodness-of-fit measures derived from the biological data sets. Given that the latter can be confidently assumed to have been structured by a branching process, namely, speciation, this implies that branching processes were more important than blending processes in structuring the cultural data sets. Thus, our analysis does not support the suggestion that blending processes have always dominated cultural evolution.

The failure of our analysis to support the claim that blending is the dominant cultural evolutionary process is in line with the region-specific quantitative studies that have been published to date (Borgerhoff Mulder, 2001, Collard & Shennan, 2000, Guglielmino et al., 1995, Hewlett et al., 2002, Jordan & Shennan, 2003, Moore & Romney, 1994, Moore & Romney, 1996, Roberts et al., 1995, Shennan & Collard, 2005, Tehrani & Collard, 2002, Welsch, 1996, Welsch et al., 1992). Several of these studies have focused on cultural variation among villages on the North Coast of New Guinea, using geographic distance and linguistic affinity as proxies for blending and branching, respectively. Using regression and correspondence analysis of presence/absence data, Welsch et al. (1992; see also Welsch, 1996) found that the material culture similarities and differences among the villages are strongly associated with geographic propinquity and unrelated to the linguistic relations of the villages. In contrast, correspondence and hierarchical log-linear analyses of frequency data carried out by Moore and colleagues indicated that geography and language have equally strong effects on the variation in material culture among the villages (Moore & Romney, 1994, Roberts et al., 1995). Moore and Romney (1996) obtained the same result in a reanalysis of the presence/absence data of Welsch et al. using correspondence analysis, thereby accounting for one potential explanation for the difference in findings, namely the use of different data sets. Recent work by Shennan and Collard (2005) supports the assessment of Moore and Romney that a combination of both branching and blending was operating in this case.

Three quantitative studies have examined cultural evolution in African societies: Guglielmino et al., 1995, Borgerhoff Mulder, 2001, Hewlett et al., 2002. The first of these explored the roles of branching, blending and local adaptation in the evolution of 47 cultural traits among 277 African societies. Models of the three processes were generated, and then correlation analyses undertaken in which language was used as a proxy for branching, geographic distance was used as a proxy for blending, and vegetation type was used as a proxy for adaptation. These analyses found that most of the traits fit best the branching model. The distributions of only a few traits were explicable in terms of adaptation to local conditions and even fewer traits supported the blending model. The results of Hewlett et al. were less clear-cut than those of Guglielmino et al., but they nevertheless supported the branching hypothesis. Hewlett et al. investigated the processes responsible for the distribution of 109 cultural attributes among 36 African ethnic groups. Using phenetic clustering and regression analysis, they tested three explanatory models: demic diffusion, which is equivalent to branching, cultural diffusion, which is equivalent to blending, and local invention. Hewlett et al. found that 32% of the cultural attributes could not be linked with an explanatory model, and that the distributions of another 27% of the cultural attributes were compatible with two of the models. Of the remaining cultural attributes, 18% were compatible with demic diffusion, 11% were compatible with cultural diffusion, and just 4% were compatible with local invention. The results of Borgerhoff Mulder's (2001; see also Borgerhoff Mulder, George-Cramer, Eshleman, & Ortolani, 2001) analysis of correlations between cultural traits associated with kinship and marriage patterns in 35 East African societies were more equivocal. In this study, analyses of phylogenetically controlled data supported roughly half the number of statistically significant correlations returned by analyses of phylogenetically uncorrected data. These results failed to support the Borgerhoff Mulder's preferred hypothesis, which is that adaptation to local environments plus diffusion between neighbouring populations erases any phylogenetic signature. Were that the case, then the correlations between different traits in the phylogenetically controlled analysis would have returned very similar results to a conventional statistical analysis of the raw data, which was not the case. However, Borgerhoff Mulder's results also do not lend unqualified support to the branching hypothesis either because a high proportion of correlations remained unaffected by phylogenetic correction. In these cases, the trace of descent is obscured either by a relatively fast rate of cultural evolution and adaptation to local conditions, or by the mixing and merging between cultural groups that has been reported in ethnographic and historical sources on East African societies. Thus, two of the three African studies offer strong support for the branching hypothesis, while the third is equivocal regarding the relative importance of branching and blending.

Three other quantitative contributions to the branching/blending debate have been published: Collard & Shennan, 2000, Tehrani, 2004, Jordan & Shennan, 2003. The first of these investigated the relative contribution of branching and blending to cultural evolution by applying phylogenetic techniques from biology to assemblages of pottery from Neolithic sites in the Merzbach valley, Germany. The analyses indicated that, while both branching and blending were involved in generating the patterns observed among the Merzbach pottery assemblages, branching was the dominant process. The study of Tehrani and Collard applied biological phylogenetic techniques to a data set comprising decorative characters from textiles produced by Turkmen tribes between the 18th and 20th centuries. The analyses focused on two periods in Turkmen history: the era in which most Turkmen practiced nomadic pastoralism and were organised according to indigenous structures of affiliation and leadership; and the period immediately following their defeat by Tsarist Russia in 1881, which is associated with the sedentarization of nomadic Turkmen and their increasing dependence on the market. The analyses of Tehrani and Collard indicated that branching was the dominant process in the evolution of Turkmen carpet designs both before and after their incorporation into the Russian Empire. The study of Jordan and Shennan (2003) used multivariate and cladistic methods to examine Californian Indian basketry variation in relation to linguistic affinity and geographic proximity. The analyses suggested that the variation is best explained by blending rather than branching, or rather that linguistic affiliation has not provided a strong canalising force on the distribution of basketry attributes, which appears to be mainly determined by geographical proximity and, therefore, presumably, frequency of interaction.

Thus, the suggestion that blending has always been a more important cultural evolutionary process than branching is also not supported by the region-specific quantitative studies that have been published to date. Blending seems to have been the dominant process in the evolution of the Californian data set, but branching was at least as important as blending in generating the New Guinea, Neolithic, and African data sets, and it was clearly the major process in producing the Turkmen data set.

5. Conclusions

The results of the quantitative comparative study described here do not support the claim that blending processes such as trade and exchange have always been more important in cultural evolution than the branching process of population fissioning (e.g., Dewar, 1995, Moore, 1994, Moore, 2001, Terrell, 1988, Terrell, 2001, Terrell et al., 1997, Terrell et al., 2001). Collectively the cultural data sets in our sample do not differ from the biological data sets in terms of how tree-like they are. The claim that blending has always been more important in cultural evolution than branching is also not supported by the region-specific quantitative assessments of cultural evolution that have been published to date. Blending processes clearly structured some data sets, but branching processes are equally clearly responsible for structuring other data sets. It appears, therefore, that branching cannot be discounted as a process in cultural evolution. This, in turn, suggests that, rather than deciding how cultural evolution has proceeded a priori (e.g., Moore, 1994, Terrell, 1988, Terrell, 2001, Terrell et al., 1997, Terrell et al., 2001), researchers need to ascertain which model or combination of models is relevant in a particular case and why (Shennan & Collard, 2005, Tehrani, 2004).

Acknowledgments

We thank Samantha Banks, Peter Jordan, Martyn Kennedy, Mike O'Brien, and Anne Yoder for assisting us with this project. Useful comments were provided by Stephen Lycett, Martin Daly, Margo Wilson, and two anonymous referees. Lastly, we would like to express our gratitude to the UK's Arts and Humanities Research Board for their ongoing support of our work on cultural evolution.

References

Ammerman & Cavalli-Sforza, 1984 1.Ammerman AJ, Cavalli-Sforza LL. The neolithic transition and the genetics of populations in Europe. Princeton, NJ: Princeton University Press; 1984;.

Archie, 1989 2.Archie JW. A randomization test for phylogenetic information in systematic data. Systematic Zoology. 1989;38:219–252.

Arnold, 1981 3.Arnold EN. Estimating phylogenies at low taxonomic levels. Zeitschrift für Zoologische Systematik und Evolutionsforschung. 1981;19:1–35.

Baker & DeSalle, 1997 4.Baker R, DeSalle R. Multiple sources of character information and the phylogeny of Hawaiian Drosophilids. Systematic Biology. 1997;46:654–673. MEDLINE | CrossRef

Barnett, 1937 5.Barnett HG. Culture element distributions: VII. Oregon coast. Anthropological Records. 1937;1:155–204.

Barnett, 1939 6.Barnett HG. Culture element distributions: IX. Gulf of Georgia Salish. Anthropological Records. 1939;5:221–295.

Barth, 1969 7.Barth F. Introduction. In:  Barth F editors. Ethnic groups and boundaries: The social organisation of culture difference. Boston, MA: Little Brown; 1969;p. 9–38.

Bellwood, 1996 8.Bellwood P. Phylogeny vs. reticulation in prehistory. Antiquity. 1996;70:881–890.

Bellwood, 2001 9.Bellwood P. Early agriculturalist diasporas? Farming, languages, and genes. Annual Review of Anthropology. 2001;30:181–207.

Bininda-Emonds & Russell, 1996 10.Bininda-Emonds ORP, Russell AP. A morphological perspective on the phylogenetic relationships of the extant phocid seals (Mammalia: Carnivora: Phocidae). Bonner Zoologische Monographien. 1996;41:1–256.

Boas, 1940 11.Boas F. Race, language, and culture. Chicago: Chicago University Press; 1940;.

Borgerhoff Mulder, 2001 12.Borgerhoff Mulder M. Using phylogenetically based comparative methods in anthropology: More questions than answers. Evolutionary Anthropology. 2001;10:99–111.

Borgerhoff Mulder et al., 2001 13.Borgerhoff Mulder M, George-Cramer M, Eshleman J, Ortolani A. A study of East African kinship and marriage using a phylogenetically based comparative method. American Anthropologist. 2001;103:1059–1082.

Boyd et al., 1997 14.Boyd R, Borgerhoff Mulder M, Durham WH, Richerson PJ. Are cultural phylogenies possible?. In:  Weingart P,  Mitchell SD,  Richerson PJ,  Maasen S editor. Human by nature. Mahwah, NJ: Lawrence Erlbaum; 1997;p. 355–386.

Boyd & Richerson, 1985 15.Boyd R, Richerson PJ. Culture and the evolutionary process. Chicago: University of Chicago Press; 1985;.

Buston & Emlen, 2001 16.Buston PM, Emlen ST. Cognitive processes underlying human mate choice: The relationship between self-perception and mate preference in Western society. Proceedings of the National Academy of Sciences. 2001;100:8805–8810.

Cavalli-Sforza & Feldman, 1981 17.Cavalli-Sforza LL, Feldman MW. Cultural transmission and evolution: A quantitative approach. Princeton, NJ: Princeton University Press; 1981;.

Cavalli-Sforza et al., 1994 18.Cavalli-Sforza LL, Menozzi P, Piazza A. The history and geography of human genes. Princeton, NJ: Princeton University Press; 1994;.

Cavalli-Sforza et al., 1988 19.Cavalli-Sforza LL, Piazza A, Menozzi P, Mountain J. Reconstruction of human evolution: Bringing together genetic, archaeological and linguistic data. Proceedings of the National Academy of Sciences of the United States of America. 1988;85:6002–6006. MEDLINE | CrossRef

Childs & Greenfield, 1980 20.Childs CP, Greenfield PM. Informal modes of learning and teaching: The case of Zinacanteco weaving. In:  Warren N editors. Studies in Cross-Cultural Psychology. vol. 2:New York: Academic Press; 1980;p. 269–316.

Collard & Shennan, 2000 21.Collard M, Shennan SJ. Ethnogenesis versus phylogenesis in prehistoric culture change: A case-study using European Neolithic pottery and biological phylogenetic techniques. In:  Renfrew C,  Boyle K editor. Archaeogenetics: DNA and the population prehistory of Europe. Cambridge: McDonald Institute for Archaeological Research; 2000;p. 89–97.

Collard & Wood, 2000 22.Collard M, Wood BA. How reliable are human phylogenetic hypotheses?. Proceedings of the National Academy of Sciences of the United States of America. 2000;97:5003–5006. MEDLINE | CrossRef

Croes et al., 2005 23.Croes D, Kelly KM, Collard M. Cultural historical context of Qwu?gwes (Puget Sound, USA): A preliminary investigation. Journal of Wetland Archaeology. 2005;5:141–154.

Darwent & O'Brien, in press 24.Darwent, J., & O'Brien, M. J. (in press). Using cladistics to construct lineages of projectile points from Northeastern Missouri. In C. P. Lipo, M. J. O'Brien, M. Collard, S. J. Shennan (Eds.) Mapping our ancestors: Phylogenetic methods in anthropology and prehistory. Hawthorne, NY: Aldine de Gruyter.

Dewar, 1995 25.Dewar RE. Of nets and trees: Untangling the reticulate and dendritic in Madagascar's prehistory. World Archaeology. 1995;26:301–318.

Diamond & Bellwood, 2003 26.Diamond J, Bellwood P. Farmers and their languages: The first expansions. Science. 2003;300:597–603. CrossRef

DiFiore & Rendall, 1994 27.DiFiore A, Rendall D. Evolution of social organization: A reappraisal for primates by using phylogenetic methods. Proceedings of the National Academy of Sciences of the United States of America. 1994;91:9941–9945. MEDLINE | CrossRef

Driver, 1937 28.Driver HE. Culture element distributions: VI. Southern Sierra Nevada. Anthropological Records. 1937;1:53–154.

Drucker, 1937 29.Drucker P. Culture element distributions: V. Southern California. Anthropological Records. 1937;1:1–52.

Drucker, 1941 30.Drucker P. Culture element distributions: XVII. Yuman-Piman. Anthropological Records. 1941;6:91–230.

Drucker, 1950 31.Drucker P. Culture element distributions: XXVI. Northwest Coast. Anthropological Records. 1950;9:157–294.

Durham, 1991 32.Durham WH. Co-evolution: Genes, culture, and human diversity. Stanford: Stanford University Press; 1991;.

Durham, 1992 33.Durham WH. Applications of evolutionary culture theory. Annual Review of Anthropology. 1992;21:331–355.

Eldredge & Cracraft, 1980 34.Eldredge N, Cracraft J. Phylogenetic patterns and the evolutionary process: Method and theory in comparative biology. New York: Columbia University Press; 1980;.

Farris, 1989a 35.Farris JS. The retention index and homoplasy excess. Systematic Zoology. 1989;38:406–407. CrossRef

Farris, 1989b 36.Farris JS. The retention index and the rescaled consistency index. Cladistics. 1989;5:417–419.

Farris, 1991 37.Farris JS. Excess homoplasy ratios. Cladistics. 1991;7:81–91.

Gibbs et al., 2002 38.Gibbs S, Collard M, Wood BA. Soft tissue anatomy of the extant hominoids: A review and phylogenetic analysis. Journal of Anatomy. 2002;200:3–49.

Gifford, 1940 39.Gifford EW. Culture element distributions: XII. Apache-Pueblo. Anthropological Records. 1940;4:1–207.

Gifford & Kroeber, 1937 40.Gifford EW, Kroeber AL. Culture element distributions: IV. Pomo. University of California Publications in American Archaeology and Ethnology. 1937;37:117–254.

Gil-White, 2001 41.Gil-White FJ. Are ethnic groups biological “species” to the human brain?. Current Anthropology. 2001;42:515–554. CrossRef

Goodenough, 1999 42.Goodenough WH. Outline of a framework for a theory of cultural evolution. Cross-Cultural Research. 1999;33:84–107.

Greenfield, 1984 43.Greenfield PM. A theory of the teacher in the learning activities of everyday life. In:  Rogoff B,  Lave J editor. Everyday cognition. Cambridge, MA: Harvard University Press; 1984;p. 117–138.

Greenfield et al., 2000 44.Greenfield PM, Maynard AE, Childs CP. History, culture, learning, and development. Cross-Cultural Research. 2000;34:351–374.

Guglielmino et al., 1995 45.Guglielmino CR, Viganotti C, Hewlett B, Cavalli-Sforza LL. Cultural variation in Africa: Role of mechanisms of transmission and adaptation. Proceedings of the National Academy of Sciences of the United States of America. 1995;92:7585–7589. MEDLINE | CrossRef

Guyer & Savage, 1986 46.Guyer C, Savage JM. Cladistic relationships among Anoles (Sauria: Iguanidae). Systematic Zoology. 1986;35:509–531.

Halanych & Robinson, 1999 47.Halanych KM, Robinson TJ. The utility of cytochrome b and 12S rDNA data for phylogeny reconstruction of leporid (Lagomorpha) genera. Journal of Molecular Evolution. 1999;48:369–379. MEDLINE | CrossRef

Hennig, 1950 48.Hennig W. Grundzüge einer Theorie der Phylogenetischen Systematik. Berlin: Deutscher Zentralverlag; 1950;.

Hewlett & Cavalli-Sforza, 1986 49.Hewlett BS, Cavalli-Sforza LL. Cultural transmission among Aka pygmies. American Anthropologist. 1986;88:922–934.

Hewlett et al., 2002 50.Hewlett BS, de Silvestri A, Guglielmino CR. Semes and genes in Africa. Current Anthropology. 2002;43:313–321. CrossRef

Hodder, 1982 51.Hodder I. Symbols in action: Ethnoarchaeological studies of material culture. Cambridge: Cambridge University Press; 1982;.

Horowitz et al., 1998 52.Horowitz I, Zardoya R, Meyer A. Platyrrhine systematics: A simultaneous analysis of molecular and morphological data. American Journal of Physical Anthropology. 1998;106:261–281. MEDLINE | CrossRef

Hurles et al., 2003 53.Hurles ME, Matisoo-Smith E, Gray RD, Penny D. Untangling Oceanic settlement: The edge of the unknowable. Trends in Ecology and Evolution. 2003;18:531–540. CrossRef

Jackman et al., 1999 54.Jackman TR, Larson A, de Queiroz K, Losos JB. Phylogenetic relationships and tempo of early diversification in Anolis lizards. Systematic Biology. 1999;48:254–285. CrossRef

Jones, 2003 55.Jones D. Kinship and deep history: Exploring connections between culture areas, genes and language. American Anthropologist. 2003;105:501–514.

Jordan & Shennan, 2003 56.Jordan P, Shennan SJ. Cultural transmission, language, and basketry traditions amongst the Californian Indians. Journal of Anthropological Archaeology. 2003;22:42–74.

Jorgensen, 1969 57.Jorgensen JG. Salish language and culture. Bloomington, IN: Indiana University; 1969;.

Kennedy et al., 2000 58.Kennedy M, Gray RD, Spencer HG. The phylogenetic relationships of the shags and cormorants: Can sequence data resolve a disagreement between behaviour and morphology?. Molecular Phylogenetics and Evolution. 2000;17:345–359. MEDLINE | CrossRef

Kennedy & Spencer, 2000 59.Kennedy M, Spencer HG. Phylogeny, biogeography, and taxonomy of Australasian Teals. Auk. 2000;117:154–163.

Kennedy et al., 1996 60.Kennedy M, Spencer HG, Gray RD. Hop, step and gape: Do the social displays of the Pelecaniformes reflect phylogeny?. Animal Behaviour. 1996;51:273–291. CrossRef

Kirch & Green, 1987 61.Kirch PV, Green RC. History, phylogeny, and evolution in Polynesia. Current Anthropology. 1987;28:431–456. CrossRef

Kirch & Green, 2001 62.Kirch PV, Green RC. Hawaiki, Ancestral Polynesia: In essay in historical anthropology. Cambridge: Cambridge University Press; 2001;.

Kitching et al., 1998 63.Kitching IJ, Forey PL, Humphries CJ, Williams DM. Cladistics: The theory and practice of parsimony analysis. Oxford: Oxford University Press; 1998;.

Kroeber, 1948 64.Kroeber AL. Anthropology: Race, language, culture, psychology, prehistory. New York: Brace; 1948;.

Lieberman et al., 1996 65.Lieberman DE, Wood BA, Pilbeam DR. Homoplasy and early Homo: An analysis of the evolutionary relationships of H. habilis sensu stricto and H. rudolfensis. Journal of Human Evolution. 1996;30:97–120. CrossRef

Lumsden & Wilson, 1981 66.Lumsden CJ, Wilson EO. Genes, mind, and culture: The co-evolutionary process. Cambridge, MA: Harvard University Press; 1981;.

McElreath et al., 2003 67.McElreath R, Boyd R, Richerson PJ. Shared norms and the evolution of ethnic markers. Current Anthropology. 2003;44:122–129. CrossRef

Mesoudi et al., 2004 68.Mesoudi A, Whiten A, Laland KN. Is human cultural evolution Darwinian? Evidence reviewed from the perspective of The Origin of Species. Evolution. 2004;58:1–11. MEDLINE | CrossRef

Moore & Romney, 1994 69.Moore CC, Romney AK. Material culture, geographic propiniquity, and linguistic affiliation on the North Coast of New Guinea: A reanalysis of Welsch, Terrell, and Nadolski (1992). American Anthropologist. 1994;96:370–396.

Moore & Romney, 1996 70.Moore CC, Romney AK. Will the “real” data please stand up? Reply to Welsch (1996). Journal of Quantitative Anthropology. 1996;6:235–261.

Moore, 1994 71.Moore JH. Putting anthropology back together again: The ethnogenetic critique of cladistic theory. American Anthropologist. 1994;96:370–396.

Moore, 2001 72.Moore JH. Ethnogenetic patterns in native North America. In:  Terrell JE editors. Archaeology, language and history: Essays on culture and ethnicity. Wesport: Bergin and Garvey; 2001;p. 30–56.

Morgan, 1870 73.Morgan LH. Systems of consanguinity and affinity of the human family. Washington, DC: Smithsonian Institution Press; 1870;.

Moylan et al., in press 74.Moylan, J. W., Graham, C. M., Borgerhoff Mulder, M., Nunn, C. L., Håkansson, T. (in press). Cultural traits and linguistic trees: Phylogenetic signal in East Africa. In C. P. Lipo, M. J. O'Brien, M. Collard & S. J. Shennan (Eds.), Mapping our ancestors: Phylogenetic methods in anthropology and prehistory. Hawthorne, NY: Aldine de Gruyter.

Nelson, 1978 75.Nelson G. Species and taxa: Systematics and evolution. In:  Otto D,  Endler JA editor. Speciation and its consequences. Sunderland, MA: Sinauer Associates; 1978;p. 60–81.

Nelson & Platnick, 1981 76.Nelson G, Platnick N. Systematics and biogeography: Cladistics and vicariance. New York: Columbia University Press; 1981;.

Noll, 2002 77.Noll FB. Behavioral phylogeny of corbiculate apidae (Hymenoptera; Apinae), with special reference to social behavior. Cladistics. 2002;18:137–153.

O'Brien et al., 2001 78.O'Brien MJ, Darwent J, Lyman RL. Cladistics is useful for reconstructing archaeological phylogenies: Paleoindian points from the southeastern United States. Journal of Archaeological Science. 2001;28:1115–1136.

O'Leary & Geisler, 1999 79.O'Leary MA, Geisler JH. The position of Cetacea within Mammalia: Phylogenetic analysis of morphological data from extinct and extant taxa. Systematic Biology. 1999;48:455–490. MEDLINE | CrossRef

Pétrequin, 1993 80.Pétrequin P. North wind, south wind: Neolithic technological choices in the Jura Mountains, 3700–2400 BC. In:  Lemonnier P editors. Technological choices. London: Routledge; 1993;p. 36–76.

Petrie, 1939 81.Petrie WMF. The making of Egypt. London: Sheldon; 1939;.

Renfrew, 1987 82.Renfrew C. Archaeology and language: The puzzle of Indo–European origins. London: Cape; 1987;.

Renfrew, 1992 83.Renfrew C. Archaeology, genetics and linguistic diversity. Man (New Series). 1992;27:445–478.

Rivers, 1914 84.Rivers WHR. The history of Melanesian society. Cambridge: Cambridge University Press; 1914;.

Roberts et al., 1995 85.Roberts JM, Moore CC, Romney AK. Predicting similarity in material culture among New Guinea villages from propinquity and language: A log-linear approach. Current Anthropology. 1995;36:769–788. CrossRef

Romney, 1957 86.Romney AK. The genetic model and Uto-Aztecan time perspective. Davidson Journal of Anthropology. 1957;3:35–41.

Sanderson et al., 1994 87.Sanderson MJ, Donoghue MJ, Piel W, Eriksson T. TreeBASE: A prototype database of phylogenetic analyses and an interactive tool for browsing the phylogeny of life. American Journal of Botany. 1994;81:183.

Schaller et al., 2003 88.Schaller M, Park JH, Faulkner J. Prehistoric dangers and contemporary prejudices. European Review of Social Psychology. 2003;14:105–137.

Schmidt, 1939 89.Schmidt W. The culture historical method of ethnology. New York: Fortuny's; 1939;.

Schuh, 2000 90.Schuh RT. Biological systematics: Principles and applications. Ithaca, NY: Cornell University Press; 2000;.

Shennan & Collard, 2005 91.Shennan SJ, Collard M. Investigating processes of cultural evolution on the North Coast of New Guinea with multivariate and cladistic analyses. In:  Mace R,  Shennan SJ,  Holden CJ editor. The evolution of cultural diversity: A phylogenetic approach. London: University College London Press; 2005;p. 133–164.

Shennan & Steele, 1999 92.Shennan SJ, Steele J. Cultural learning in hominids: A behavioural ecological approach. In:  Box HO,  Gibson KR editor. Mammalian social learning: Comparative and ecological perspectives. Cambridge: Cambridge University Press; 1999;p. 367–388.

Smith, 2001 93.Smith EA. On the coevolution of cultural, linguistic and biological diversity. In:  Maffi L editors. On biocultural diversity: Linking language, knowledge, and the environment. Washington, DC: Smithsonian Institution Press; 2001;p. 95–117.

Smith, 1933 94.Smith GE. The diffusion of culture. London: Watts; 1933;.

Stanhope et al., 1998 95.Stanhope MJ, Waddell VG, Madsen O, de Jong W, Hedges SB, Cleven GC, et al. Molecular evidence for multiple origins of Insectivora and for a new order of endemic African insectivore mammals. Proceedings of the National Academy of Sciences of the United States of America. 1998;95:9967–9972. MEDLINE | CrossRef

Steward, 1941 96.Steward JH. Culture element distributions: XIII. Nevada Shoshoni. Anthropological Records. 1941;4:209–259.

Stewart, 1941 97.Stewart OC. Culture element distributions: XIV. Northern Paiute. Anthropological Records. 1941;4:361–446.

Swofford, 1998 98.Swofford DL. PAUP* 4. Phylogenetic analysis using parsimony (*and Other Methods). Version 4. Sunderland: Sinauer; 1998;.

Tehrani, 2004 99.Tehrani, J. J. (2004). Processes of cultural diversification in the evolution of Iranian tribal craft traditions. Unpublished PhD dissertation, University College London.

Tehrani & Collard, 2002 100.Tehrani JJ, Collard M. Investigating cultural evolution through biological phylogenetic analyses of Turkmen textiles. Journal of Anthropological Archaeology. 2002;21:443–463.

Terrell, 1988 101.Terrell JE. History as a family tree, history as a tangled bank. Antiquity. 1988;62:642–657.

Terrell, 2001 102.Terrell JE. Introduction. In:  Terrell JE editors. Archaeology, language, and history: Essays on culture and ethnicity. Wesport: Bergin and Garvey; 2001;p. 1–10.

Terrell et al., 1997 103.Terrell JE, Hunt TL, Gosden C. The dimensions of social life in the Pacific: Human diversity and the myth of the primitive isolate. Current Anthropology. 1997;38:155–195. CrossRef

Terrell et al., 2001 104.Terrell JE, Kelly KM, Rainbird P. Foregone conclusions? In search of “Papuans” and “Austronesians”. Current Anthropology. 2001;42:97–124. CrossRef

Venti, 2004 105.Venti J. F. (2004). Methods for investigating the evolutionary pattern of adaptive radiations in cultural history: A preliminary cultural evolutionary pattern analysis of early Christianity. Unpublished MA thesis, University College London.

Wayne et al., 1997 106.Wayne RK, Geffen E, Girman DJ, Koepfli KP, Lau LM, Marshall CR. Molecular systematics of the Canidae. Systematic Biology. 1997;46:622–653. MEDLINE | CrossRef

Welsch, 1996 107.Welsch RL. Language, culture and data on the north coast of New Guinea. Journal of Quantitative Anthropology. 1996;196:209–234.

Welsch et al., 1992 108.Welsch RL, Terrell JE, Nadolski JA. Language and culture on the North Coast of New Guinea. American Anthropologist. 1992;94:568–600.

Whaley, 2001 109.Whaley LJ. Manchu-Tungusic culture change among Manchu-Tungusic peoples. In:  Terrell JE editors. Archaeology, language, and history: Essays on culture and ethnicity. Wesport, CT: Bergin and Garvey; 2001;p. 103–124.

Yang & Yoder, 2003 110.Yang Z, Yoder AD. Comparison of likelihood and Bayesian methods for estimating divergence times using multiple gene loci and calibration points, with application to a radiation of cute-looking mouse lemur species. Systematic Biology. 2003;52:705–716. MEDLINE | CrossRef

Yoder, 1994 111.Yoder AD. The relative position of the Cheirogaleidae in strepsirhine phylogeny: A comparison of morphological and molecular methods and results. American Journal of Physical Anthropology. 1994;94:25–46. MEDLINE | CrossRef

Yoder et al., 2003 112.Yoder AD, Burns MM, Zehr S, Delefosse T, Veron G, Goodman SM, et al. Single origin of Malagasy Carnivora from an African ancestor. Nature. 2003;421:734–737. MEDLINE | CrossRef

Yoder & Yang, 2000 113.Yoder AD, Yang Z. Estimation of primate speciation dates using local molecular clocks. Molecular Biology and Evolution. 2000;17:1081–1090. MEDLINE

a Department of Anthropology and Sociology, University of British Columbia, Vancouver, Canada

b Institute of Archaeology, University College London, London, UK

c Department of Anthropology, University College London, London, UK

d AHRB Centre for the Evolutionary Analysis of Cultural Behaviour, University College London, Gower Street, London, UK

Corresponding author. Department of Anthropology and Sociology, University of British Columbia, 6303 NW Marine Drive, Vancouver, Canada V6T 1Z1. Tel.: +1 604 822 4845.

PII: S1090-5138(05)00060-7

doi:10.1016/j.evolhumbehav.2005.07.003



2007:12:08