The Missing Mousterian

comparing frequencies of particular classes of objects (in- cluding named types ... classes that constitute our basic units of analysis were origi- nally defined in ...... 1995]), but the evidence as presented does not seem to war- rant the claim that ...
638KB taille 1 téléchargements 359 vues
Current Anthropology Volume 47, Number 5, October 2006

777

The Missing Mousterian by Harold L. Dibble and Shannon P. McPherron Data from the Middle Paleolithic site of Pech de l’Aze´ IV (France) indicate the deliberate production of small flakes that appear to have been used in an unretouched state. Such materials have not typically been recognized as being intended end products, though various techniques for their manufacture are present in many Middle Paleolithic industries. This raises not only the question of their distribution and possible use during that time but also the larger issue of how Paleolithic archaeologists recognize lithic classes that are both analytically and behaviorally meaningful.

A significant component of Middle Paleolithic industries seems to have been overlooked in the current systematics. What we attempt to show here is that during this time there was a deliberate production of very small flakes that were intended for use in their unretouched state. If our findings are correct, then it is possible that these flakes represent a significant aspect of Middle Paleolithic industrial variability and behavior even though such artifacts are typically unrecognized in Paleolithic assemblages and certainly do not form part of standard descriptive analyses. In large part, this study is about more than just small flakes, since the demonstration of significance raises a larger question as to how archaeologists recognize and define their basic units of analysis. This is a serious issue in Middle Paleolithic archaeology, since many if not most analyses proceed either by comparing frequencies of particular classes of objects (including named types and other more informal classes) or by examining the technological processes underlying the manufacture of certain classes. In either case a beginning assumption is that the classes or types under investigation are somehow “real” in the sense of being behaviorally meaningful or otherwise interpretable because the processes underlying their production are understood. In Middle Paleolithic archaeology, most of the typological classes that constitute our basic units of analysis were originally defined in the first half of the twentieth century or even earlier, though it was Bordes (1961a) who standardized many of their definitions. Unfortunately, however, there have been only sporadic attempts to examine critically each of the types to determine whether they have any real behavioral signifiHarold L. Dibble is Professor of Anthropology at the University of Pennsylvania (Philadelphia, PA 19104, U.S.A. [[email protected] .edu]). Shannon P. McPherron is Research Scientist in the Department of Human Evolution of the Max Planck Institute for Evolutionary Anthropology in Leipzig, Germany. The present paper was submitted 31 III 05 and accepted 16 II 06.

cance (Debe´nath and Dibble 1994; cf. Bisson 2000). Thus, while most Paleolithic archaeologists are willing to accept that scrapers, for example, were deliberately manufactured, it is much less certain that all of the some 18 scraper subtypes are truly distinct (Dibble 1995a and citations therein) or that other types, such as pseudo-microburins, notched triangles, alternate retouched beaks, rabots, or inverse choppers, were deliberately manufactured to represent specific kinds of objects. It is important, therefore, to begin to develop methods by which evaluations of existing types or artifact classes or the proposal of new ones can be made.

Methodological Background: Recognizing Deliberate Products in Lithic Assemblages Before beginning the presentation of data that suggest the deliberate production of small flakes, it is useful to outline some of the methodological problems that underlie such a demonstration. There are ways to identify deliberately produced end products in lithic assemblages, but all of them require particular characteristics or contexts that can be used to set them apart from the general background of knapping debris. For example, the presence of retouch strongly suggests that pieces exhibiting it were deliberately selected and their edge(s) modified in order to perform a specific task. By analyzing what characteristics particular kinds of retouched pieces have in common and in what ways they differ from nonretouched pieces, it is sometimes possible to arrive at some idea of the criteria used by prehistoric flintknappers for their selection. Likewise, the clear and unambiguous presence of use wear on pieces would also argue for deliberate selection for use. Finally, by noting that particular forms are found in specific contexts (caches or burials, for instance), one can make the argument that they served specific purposes. All of these approaches, however, especially in the context

䉷 2006 by The Wenner-Gren Foundation for Anthropological Research. All rights reserved 0011-3204/2006/4705-0003$10.00

778

of Middle Paleolithic assemblages, are fraught with problems. First, knowing that a blank was selected for retouching or for use does not necessarily mean that the original blank was purposefully manufactured to have those specific characteristics—useful flakes can just as easily be selected from a pile of knapping debris long after their original manufacture (McDonald 1991). It is also recognized now that criteria for discard are much more of a factor governing what enters the archaeological record than are criteria for use. This is one of the main lessons from the work of Frison (1968) and subsequent researchers who investigated changes in artifact form due to resharpening and rejuvenation. This ultimately led to the recognition of the “finished artifact fallacy” (Davidson and Noble 1993), which basically refers to the fact that what we find in the ground is not so much what the prehistoric knappers wanted as what they wanted to throw away. Even objects found in special contexts have to be treated with some skepticism. For example, at the site of Fonte´chevade (France), the original excavator found many quartzite cobbles in the sediments of the cave. Because these cobbles could not have originated in the surrounding limestone, she concluded that they must have been deliberately imported to serve as hammerstones (Henri-Martin 1957). Subsequent analyses have shown that they washed into the cave naturally through a chimney in the back (Dibble et al. 2006). The goal here is to demonstrate that the production of a particular class of objects—small flakes—was an intentional and deliberate result and not simply an unintended by-product of the knapping of other products. This demonstration is admittedly much harder than the examples given above. In the first place, all chipped stone technologies result in flakes of all sizes, and a major proportion are relatively small because of processes that include the preparation of platforms on cores and the retouching of tools. Therefore, the simple presence of small flakes, in any frequency, cannot be used to demonstrate that they represent, in themselves, desired end products. Another problem, as described below, is that the small flakes discussed here are not retouched and do not show any obvious signs of having been used. Moreover, they occur spatially throughout the assemblage, and a large number of them are identical to typical knapping by-products. Thus, the argument to be presented here that they were intentionally produced will be based on analyses not of these small flakes themselves but of the pieces from which they were struck.

Small Flakes as Desired End Products: Data from Pech de l’Aze´ IV The present study is based on analysis of material from the site of Pech de l’Aze´ IV, a Middle Paleolithic site located in southwestern France in the department of the Dordogne. Pech IV, one of a series of Lower and Middle Paleolithic sites located close to one another that have been excavated since the midnineteenth century (see McPherron and Dibble 2000; McPherron, Soressi, and Dibble 2001), is a collapsed cave orig-

Current Anthropology Volume 47, Number 5, October 2006

inally excavated from 1970 to 1977 by Bordes (1975). Overall he recovered about 90,000 lithic artifacts and another 30,000 faunal remains (McPherron and Dibble 2000). From 2000 to 2003 we performed new excavations at the site in an effort to collect fresh dating samples and to study its geological and taphonomic history. A final report on this work, including dating of the complete sequence, is under way. The data presented here are based on the combined collections, and the correspondence between the old and new stratigraphic sequences is shown in table 1 (see Dibble, Raczek, and McPherron 2005 for a complete discussion of how these two collections were combined). One of the more interesting assemblages to come from this site is one that Bordes named the “Asinipodian” (fig. 1), defined originally (Bordes 1975) by the presence of very small Levallois cores and Kombewa cores (flakes whose bulb of percussion has been removed [see Bre´zillon 1968; Debe´nath and Dibble 1994]) and the general absence of retouched tools. From our own analysis of the two collections it is apparent that a third class of object, truncated-faceted pieces (Schroeder 1969), can also be considered a significant element in this assemblage. Although, to the best of our knowledge, the term “Asinipodian” has never been applied to any other assemblage, all of the elements that make up this particular industry are known together or singly in many other Middle Paleolithic contexts. There are several lines of evidence that suggest that all three of these classes—Levallois cores, Kombewa cores, and truncatedfaceted pieces—were used to produce very small flakes, although the specific techniques differed considerably. Two technologies characteristic of the Asinipodian—Levallois and Kombewa—are critical for this demonstration. By its very definition (Debe´nath and Dibble 1994), Levallois is seen, at least in part, as a way of predetermining, through preparation of the core surface, both the size and shape of the resulting flake. Thus, when we find unmistakable Levallois flakes or cores that are exceptionally small, it is difficult to Table 1. Stratigraphic Correlations between Bordes’s and McPherron/Dibble’s Excavations

Bordes’s Level F1-F3 F4 G, H1-H2 I1 I2 J1 J2-J3 J3A-J3B J3C X N/A Y-Z

Dibble/McPherron’s Level

Global Level

3A 3B 4A 4B 4C 5A 5B 6A 6B N/A 7 8

I-A I-B II-A II-B II-C III-A III-B IV-A IV-B – – V

Note: Global Level IV-A corresponds to the Asinipodian as described by Bordes.

Dibble and McPherron The Missing Mousterian

779

Figure 1. Examples from Pech de l’Aze´ IV of the lithic classes described in the text. 1–4, Levallois cores (2 and 3 made from a small knob of raw material); 5–7, truncated-faceted pieces; 8–10, Levallois flakes; 11, Kombewa core, 12, Kombewa flake. All objects from Level IV-A.

argue that, despite their size, their production was anything other than deliberate. Likewise, Kombewa technology is described (Bre´zillon 1968; Tixier, Inizan, and Roche 1980) as a deliberate method of flake production in which the naturally convex surface of the flake’s interior surface at the bulbar end was exploited as a surface for the removal of a small flake. The resulting flakes are clearly identifiable because they appear

to have two “interior” surfaces, one being the interior surface of the flake used as a core. Truncated-faceted pieces have achieved recognition only relatively recently and have previously gone under a number of different names (Dibble 1984; Nishiaki 1985). They are flakes that exhibit a truncation on one edge which in turn served as a platform for the removal of even smaller flakes

780

Figure 2. Frequency distributions of maximum scar lengths of Levallois cores, Kombewa cores, “other” cores, and truncatedfaceted pieces from Level IV-A.

(“facets”) from either the interior or the exterior surface (Schroeder 1969; Debe´nath and Dibble 1994). They have been variously interpreted as a means of basal thinning for hafting (Coon 1951), as tools in themselves (Leakey 1931, 99–100; Dibble 1984), or as cores (Schroeder 1969; Solecki and Solecki 1970; see also Debe´nath 1988; Goren-Inbar 1988; Hovers 2006; Dibble and McPherron 2006). While both Levallois and Kombewa are considered techniques for producing flakes, it is more difficult to assume that truncated-faceted pieces are also used to produce flakes rather than being tools in themselves or pieces exhibiting hafting modification. However, at Pech IV there do appear to be at least two aspects of underlying similarity for all three artifact classes. The first is the distribution of sizes of the flakes removed from these pieces, and the second is their frequency in the sequence of levels of Pech IV. All three classes of objects—Levallois and Kombewa cores and truncated-faceted pieces—exhibit negative flake scars representing flakes that were removed from them, and the length of these scars (the longest if there were several) was recorded (on Levallois cores the measured scar represents the length of the principal, or preferential, removal). The data on maximum flake scar lengths are presented in figure 2, and two things are important to emphasize here. While the average maximum flake scars on Levallois and other cores are statistically significantly larger than those on truncated-faceted pieces and Kombewa cores (t p 11.159, d.f. p 1935, P ! .00001), there is a high degree of overlap, and all of these techniques produced very small flakes. In fact, the median scar length on the Levallois cores is only 24.9 mm, which means that fully half of the maximum scar lengths—and therefore the length of the final flakes produced from them— were less than this. The median flake scar on Kombewa cores is even smaller, 19.63 mm. Thus, if both Levallois and Kombewa cores were designed for the deliberate production of flakes, then the flakes they produced were quite small. By inference, the similarity in the distribution of scar lengths between them and the truncated-faceted pieces and “other”

Current Anthropology Volume 47, Number 5, October 2006

cores suggests that the same goal could be ascribed to the latter classes as well. Further evidence for a functional relationship among these classes is the fact that in the stratigraphic sequence of Pech IV, the percentage of Kombewa, truncated-faceted pieces, and small (less than 30 mm in maximum dimension) cores, including Levallois cores, show the same spike in occurrence in the Asinipodian (fig. 3). It could be that these classes possessed complementary functions during the time of that occupation (small flakes from Levallois cores being used for one thing while Kombewa cores or flakes and truncated-faceted pieces were used for others). However, what these elements essentially share is the evidence of similar kinds of removals, which reached a peak during this particular time. There is no evidence that the flakes produced by these various techniques were intended to be fashioned into retouched tools. While it is virtually impossible to recognize flakes removed from a truncated-faceted piece, examination of the Levallois and Kombewa flakes shows that they were not generally retouched (only 68 out of 884 Levallois flakes and 3 out of 21 Kombewa flakes). This is not surprising, however. Many studies (Geneste 1985; Meignen 1988; Dibble and Holdaway 1990; Dibble 1995a, 1995b) have shown that in the Middle Paleolithic flakes with retouch are generally among the larger flakes in the assemblage, and data from the Asinipodian and other layers from Pech de l’Aze´ IV (fig. 4 and table 2) continue to support this. Given that the average lengths of the flake removals presented earlier represent, on average, only half the length of the retouched tools, the small sizes of these flakes, including those from the truncated-faceted pieces, probably made them unlikely to be selected for subsequent retouch. The fact remains, however, that they were manufactured, which is a strong argument that the intent must have been to use them in an unretouched state.

Potential Alternative Hypotheses While it is clear that small flakes were being produced, it is

Figure 3. Relative proportions of Kombewa flakes, truncatedfaceted pieces, and small cores (less than 30 mm in length) through the Pech de l’Aze´ IV sequence. The spikes in all three occur in the Asinipodian Level IV-A.

Dibble and McPherron The Missing Mousterian

Figure 4. Average lengths of tools (columns) and flakes (solid line) from the Pech de l’Aze´ IV sequence.

nonetheless relevant to ask whether their small size was something that was desired in and of itself or a consequence of something else. For example, the size of stone tools can be affected by raw-material size and the degree to which cores are reduced. Four lines of evidence suggest that neither of these is playing a significant role in the Asinipodian, however. First, the overall sizes of flakes and tools in the Asinipodian level are, on average, not smaller than in levels that do not show similarly high percentages of these small core techniques (see fig. 4). Four of the assemblages in the Pech IV sequence have greater average flake lengths (II-B, IIC, III-A, III-B), while six have small averages. In other words, in Level IV-A, where a significant number of small flakes were being produced by these three techniques, larger flakes were also being produced in sufficient numbers and sizes to maintain a relatively high average length of all flakes in the level. Unfortunately, there is no independent means of assessing rawmaterial availability and size during the various occupations. However, given that the overall average flake size in Level IVA is similar to that seen in the other levels, it would not appear that raw-material size or variability was any different during the occupation of that level. Second, there is no evidence that the overall intensity of utilization of the lithic resources was significantly higher during the Asinipodian occupation. There are two measures of overall intensity of utilization—the ratio of blanks to cores (how many flakes are removed from a core) and the ratio of tools to flakes (how many flakes are converted into retouched tools). The Asinipodian expresses values of both of these ratios the same as or even lower than those for the other assemblages at the site (fig. 5 and table 3), which suggests that cores were not being more reduced at that time and that more of the larger flakes were not being retouched into tools, leaving behind only small flakes. While both of these ratios can be affected by other processes (see Dibble 1995c), especially import and export of material into and from the site, there is

781

no evidence that these processes were any different during the occupation of Level IV-A. Third, many of the small Levallois cores from the Asinipodian levels were made on very small knobs of flint struck from larger nodules (see fig. 1, 2 and 3) and are not, therefore, the end products of continued reduction of originally large nodules. Finally, while it is possible that some of the Levallois cores may have produced larger flakes in earlier stages of reduction, this is not the case for Kombewa cores and truncated-faceted pieces, which almost always retain the overall form of the original flake. While there is no evidence that the emphasis on small flake production was a consequence of raw materials or a high degree of intensity of utilization of lithic resources, an alternative possibility is that the removals were taken from Kombewa cores and truncated-faceted pieces in order either to prepare those pieces for a specific use or as a means of thinning them for hafting. Arguments against hafting of the truncated-faceted pieces have already been made (see, e.g., Dibble 1984, 29). In part these are based on the fact that truncatedfaceted removals occur randomly with regard to specific classes of tools (scrapers, points, notches, etc.) and that many truncated-faceted pieces exhibit multiple removals—for example, from the proximal and distal ends or from both lateral margins. These same results hold true for the material from Pech de l’Aze´ IV, where 50 out of 196 truncated-faceted pieces exhibit such multiple removals. Having such removals from opposite margins of the piece is inconsistent with normal hafting modification. Likewise, there are no visible traces of hafting wear or residues on these pieces or on the Kombewa cores, though more thorough microscopic examination of the pieces would be critical in addressing this problem. Nonetheless, the presence of the small removals from Levallois and other cores cannot be explained by hafting or other use of the cores. Table 2. Summary Statistics for the Lengths of Complete Retouched Tools and Unretouched Flakes Retouched Tools

I-A I-B II-A II-B II-C III-A III-B IV-A IV-B V

Unretouched Flakes

Mean

N

S.D.

Mean

N

S.D.

T

43.22 42.05 50.79 53.76 58.52 56.08 51.66 43.88 40.04 48.42

266 520 117 84 615 108 286 336 157 287

13.05 13.46 14.48 14.13 16.08 13.67 16.01 13.49 11.07 14.08

33.49 34.36 35.56 38.52 39.82 37.72 38.85 36.79 34.23 36.56

3,066 4,246 406 192 3,344 759 1,293 4,056 1,395 1,837

9.57 10.22 11.92 12.11 12.56 11.91 12.69 11.34 9.43 11.42

⫺10.08 ⫺11.66 ⫺13.96 ⫺14.26 ⫺47.93 ⫺18.29 ⫺22.94 ⫺12.36 ⫺4.25 ⫺18.21

P ! ! ! ! ! ! ! ! ! !

.0001 .0001 .0001 .0001 .0001 .0001 .0001 .0001 .0001 .0001

Note: Length is taken from the point of percussion to the most distal end of the piece. In all of the assemblages, retouched pieces are significantly larger than unretouched ones.

782

Figure 5. Ratios of blank to core (the number of complete and proximal flakes and tools divided by the number of cores) and tool to flake (the number of complete and proximal tools divided by the number of complete and proximal flakes). Although both of these measures can reflect other processes, they also monitor the intensity of utilization of the lithic resources through either increasing core reduction or increasing tool production, and both can result in smaller average sizes of the unretouched flakes. The fact that the Asinipodian level does not show relatively high values of either suggests that these processes were not responsible for the emphasis on small flake production at that time.

There are many reasons, then, to conclude that these Asinipodian levels from Pech de l’Aze´ IV reflect an emphasis on the production of very small flakes and that these flakes were produced in a variety of ways. There is virtually no question that Levallois is a means of flake production, and the same is probably true—though not as certain—for the Kombewa technique. These two techniques share important characteristics with each other and with truncated-faceted pieces, and together all three classes resulted in very small removals. Moreover, the small size of these removals does not appear to be related to other factors, such as raw-material size or degree of core reduction, and there is no evidence that they represent hafting or other deliberate modification of the original piece. Our conclusion is, therefore, that their small size was, in itself, a desired feature and that they were produced for use in their natural, unretouched state.

Discussion One possible objection to the arguments presented here is that the small flakes appear to be too small to have been used. There is, of course, the possibility that they were hafted, perhaps into composite tools, and there is evidence for hafting in the Middle Paleolithic (see Churchill 2001 for a review). These studies are based on larger flakes (for example, Levallois points) or other retouched tools, however, and there are no systematic studies of Middle Paleolithic artifacts of the sizes being emphasized here. However, examination of these pieces revealed no signs of hafting modification or residues. Hafting of the small flakes is not, however, the only way they could have been used. Alternatively, as Moncel (2003, 46) notes in the context of a different kind of industry, small

Current Anthropology Volume 47, Number 5, October 2006

artifacts may simply have been gripped in the hand. Recent studies of Neandertal hand morphology have shown that their hands were as capable as our own in terms of the range of movement (see Niewoehner, Weaver, and Trinkaus 1997; Niewoehner 2001 and citations therein) but also show traits beyond modern variation, including “unusually hypertrophied hand musculature, significantly increased mechanical advantages across many joints, unusually broad fingertips, and unusual shapes and orientations of some of their CMC [carpometacarpal] joints” (Niewoehner 2001, 2979). There is continued debate, however, about what kinds of behaviors would have generated these patterns. Generally, these traits are interpreted as adapted for a greater power grip (gripping a hammerstone, for instance). However, when the direction of force is considered along with its magnitude, it seems that Neandertals were not as well adapted for the kind of oblique joint reaction forces that are involved, for instance, in holding the handle of a hammer (Niewoehner 2001). Thus, Niewoehner concludes (p. 2980) that, despite some evidence of hafting that would imply oblique power grips, Neandertal hands were actually better suited to use “either with handheld stone flakes or with flakes hafted into the distal rather than the lateral aspects of handles” (see also Villemeur 1994). While these interpretations should not be taken as conclusive evidence that small flakes were hand-held, they are presented here to show that there is no a priori reason to conclude that the small size of these flakes would have prevented their use. There is, however, a very important question as to the function of these very thin and sharp flakes. There are suggestions (Niewoehner 2001) that small flakes may have been used for woodworking; however, preliminary microwear analyses of 15 pieces from Pech IV yielded only two that showed signs of having been used for short-term cutting of soft materials (W. Banks, personal communication), and this may reflect a more expedient production and use of such flakes. This is clearly an area of research that should be pursued further. However, a lack of understanding on our part of the

Table 3. Counts of Complete Cores, Retouched Tools, Unretouched Flakes, and Total Blanks (Retouched or Not) Level

Cores

Tools

Flakes

Total Blanks

I-A I-B II-A II-B II-C III-A III-B IV-A IV-B V

182 216 18 16 166 46 163 693 287 126

351 655 141 103 769 132 358 522 224 360

3,999 6,025 590 301 5,044 1,091 2,222 7,736 2,708 3,214

4,350 6,680 731 404 5,813 1,223 2,580 8,258 2,932 3,574

Note: Counts of both tools and flakes are based on complete or proximal pieces (which accounts for the number of platforms present) and do not include medial or distal fragments.

Dibble and McPherron The Missing Mousterian

way Middle Paleolithic implements were used is not in itself a compelling argument against those pieces’ having been a significant part of the general technological component. There is, for example, still considerable debate on the functions of most Paleolithic implements, from bifaces, which could have been tools or cores or something else entirely (e.g., Davidson and Noble 1993; Kohn and Mithen 1999; Roberts and Parfitt 1999), to scrapers and so-called points (Gordon 1993; Dibble 1995a) and notches and denticulates (which can also reflect taphonomic disturbances [Bordes and Bourgon 1951; McBrearty et al 1998; Nielson 1991]). A better understanding of the spatial and particularly the environmental distribution of small flake technologies could help address this, and experimental studies would be useful to address the relative merits of small, thin flakes versus the edges exhibited on other Middle Paleolithic elements, be they large unretouched flakes, scrapers, or notches.

Further Implications If further research supports the conclusion that Middle Paleolithic hominins deliberately manufactured and used small flakes, it would have significant implications for the interpretation of Mousterian assemblage variability. In the Asinipodian, the production of small flakes was based on three particular techniques, and, to the best of our knowledge, in this regard this particular assemblage appears to be unique. However, it is known that many other Middle Paleolithic industries exhibit one or more of these techniques. For example, Moncel (2003 and citations therein) reports on other Middle Paleolithic industries that exhibit small cores, and truncated-faceted pieces clearly exist (under a variety of names) in assemblages from Europe (Turq and Marcillaud 1976; Roth, Lenoir, and Dibble 1995), the Levant (Schroeder 1969; Solecki and Solecki 1970; Crew 1976), the Zagros (Dibble 1984; Dibble and Holdaway 1990), and Africa (Leakey 1931). Kombewa technique is also widespread through Europe and Africa (Owen 1938; Dauvois 1981) though it is virtually absent in the Zagros or Levantine Mousterian. Like many other aspects of Middle Paleolithic industries, small flake production may be very widespread, though it may be associated with several different techniques. Furthermore, there may be other, still unrecognized techniques for small flake production. One distinct possibility is Clactonian notches. Rather than the notch itself, the flake that was removed from the notch may have been the intended end product, since it would possess many of the same characteristics as the small flakes made with the other techniques described above. Another possibility is handaxes. While most have focused on these artifacts as tools with heavy-duty cutting edges, it is also widely acknowledged that they are a ready source of sharp flakes whose morphology would appear to be quite similar—that is, small, thin, and extremely sharp. While we are not concluding that handaxes existed for the sole purpose of flake production, it could very well represent

783

an additional purpose for their manufacture and transport. Our point, though, is only that there are many possible techniques that can be used to produce small flakes, and this opens the door for new interpretations of particular tool types. Moreover, it is possible that small flake production may represent an important aspect of variability in the Middle Paleolithic. As has been noted for some time (Dibble 1988a; Dibble and Rolland 1992), most Middle Paleolithic assemblages contain more or less the same range of major classes of tool types and technologies but differ in the relative frequencies of those classes. Likewise, all of the techniques described here occur in most if not all Middle Paleolithic assemblages but vary in their relative proportions. In this regard, the Asinipodian appears to represent an extreme emphasis but probably only one end of a continuum of variability among Middle Paleolithic assemblages. Thus, it is likely that such production, along with the production of other retouched tools, bifaces, and particular technologies, represents a significant axis of assemblage variability and (as is still true of the others) one that definitely needs explanation. Having said that, we do not concur with Bordes (1975) in elevating the status of the Asinipodian to the level of another named facies of the Mousterian. Rather, for whatever reason, activities at this particular time and place were suited for this kind of production to a degree not previously observed. Aside from what it means for Middle Paleolithic behavior, the production of small flakes also raises important methodological issues in terms of techniques of excavation as well as the analysis and description of Middle Paleolithic assemblages. It is common practice to provenience, number, and analyze lithic objects only above a minimum size—usually 2 to 2.5 cm. The flakes that were being deliberately produced by the techniques described here were often much smaller and thus easily overlooked. Even if such cutoffs are lowered, the problem of identifying deliberately produced small flakes among the debris produced through core reduction and tool production still remains. While flakes removed from Levallois and Kombewa cores are generally recognizable, it is less clear for those coming from truncated-faceted pieces and perhaps other techniques of small flake production as well.

Conclusion One of the most important contributions to Paleolithic archaeology was the standardization of the typology for lithic artifacts put forward by Bordes during the middle of the last century (Bordes 1961a). His typology of flake tools consisted of 63 named types, most of them defined by the type and the placement of retouch on flakes and some by the technology used to create them (Debe´nath and Dibble 1994). His second major contribution (see Sackett 1981) was to apply this typology to Mousterian assemblages of northern and southwestern France, and on that basis he developed the assemblage systematics—the various Mousterian facies—that are still recognized and used to this day. The significance of these con-

784

tributions is undeniable given the tremendous amount of research that they stimulated concerning the factors that underlie variability in the types themselves (e.g., Dibble 1995a; McPherron 1994, 1995) and the assemblages that are composed of them (e.g., Bordes 1961b; Binford and Binford 1966; Mellars 1969, 1996; Rolland 1981; Rolland and Dibble 1990). Much of this work has also focused on fundamental anthropological concerns, using these same lithic types as evidence for the evolution of language (e.g., Holloway 1981; Gowlett 1984; Dibble 1989), culture (Holloway 1969; see Gibson and Ingold 1993), and intelligence (Wynn 1979, 1985). As is true in any empirical discipline, research tends to be focused on units of analysis defined on the basis of particular criteria (see papers in Ramenofsky and Steffen 1998). Thus, in the case of Paleolithic archaeology, for a long time these units were largely retouched pieces, and with the exception of Levallois flakes and cores most unretouched pieces were related to the status of “waste” or by-products. After all, it was assumed for many years that retouched tools represented the desired end products, and so naturally they were the focus of most research. This situation has changed over the past two decades, however, as interest in the underlying technology of blank production has increased (Marks and Volkman 1983; Boe¨da 1986; see papers in Dibble and Bar-Yosef 1995). Nonetheless, even in a technological approach, where the emphasis is on reconstructing the processes that resulted in particular flake forms, the need still exists to understand what the intended results were (Bar-Yosef and Dibble 2005). To put it simply, it is irrelevant to know how certain things were made unless we know what their producers were trying to make. In a similar vein, while we desperately need to know more about the functions of these objects, such studies also require that potentially informative classes of material be identified. Thus, this paper is not at all about whether a new “type” should be added to Bordes’s type list but focuses on the problem of being able to recognize meaningful units in the archaeological record. What we have tried to do here is to present arguments based on several lines of evidence that Middle Paleolithic hominins deliberately produced small flakes for specific functions. If this conclusion is true, then this fact has been overlooked for virtually the entire history of the discipline. It represents, then, a component of Middle Paleolithic technology and behavior that is, in effect, missing from our concepts of what constitute behaviorally significant aspects of Middle Paleolithic industries. While future work may clarify the specific role that theses flakes played in Middle Paleolithic adaptation, that they have been overlooked for nearly a century and a half reminds us once again of how little we truly understand about those past lifeways.

Acknowledgments We thank several colleagues, including the seven anonymous reviewers of this manuscript, for their helpful comments and

Current Anthropology Volume 47, Number 5, October 2006

suggestions. Funding for the excavation and analysis of Pech de l’Aze´ IV was provided by the Leakey Foundation and the National Science Foundation (Grant No. BNS0073856).

Comments Nick Ashton Department of Prehistory and Europe, British Museum, 56 Orsman Road, London N1 5QJ, UK (nashton@thebritish museum.ac.uk). 28 IV 06 Dibble and McPherron have identified some interesting patterns in the lithic data from Pech de l’Aze´ IV. Their arguments centre on the distinctiveness of Level IV-A, in particular the production of Levallois and Kombewa cores and truncatedfaceted pieces, and the idea that these pieces represent the intentional production of small unretouched flakes. This opens up questions relating to Neanderthal power grip or hafting and the effects of a previously unrecognized category of tool on interpretations of Mousterian assemblage variability. The data that they present indicate that all cores are smaller in Level IV-A, flake scars from intended products generally vary between 10 and 35 mm in maximum length, average flake lengths are similar for all levels at Pech de l’Aze´ IV (means 33–40 mm), there is less conversion of flakes into tools in Level IV-A, and there are fewer flakes per core in Level IV-A. Dibble and McPherron argue that because of the similarity in flake sizes between the different levels, raw-material size was also similar. However, the less intense working of the cores in Level IV-A despite their smaller size suggests that nodule size was also smaller. The size of flakes removed from a core will depend on the size of the original nodule, the intensity of working, and the type of technology deployed. Therefore the question is to what extent the observed pattern is based on the use of different technology, which itself may have been determined by differences in the size and, importantly, the shape of the nodules. A further problem lies in how certain it is that the removals that have been measured were intended products. With Levallois, this is commonly assumed, but when some of these (3–4%) were as small as 10–15 mm (fig. 2) doubts must arise about the intentions of the makers. Could some of these cores have been in the process of repreparation when a decision was made to discard them? From a Lower Palaeolithic perspective, knapping frequently seems to go beyond the point of utility. One example is at High Lodge (Suffolk, UK [Ashton et al. 1992]), where a refitting knapping scatter shows the reduction of a partly worked core through the removal of 30 further flakes. Several flakes are missing from the middle of the knapping sequence, but most of the final flakes have been discarded with the scatter, presumably as waste products. It would seem that the core was rapidly flaked and discarded

Dibble and McPherron The Missing Mousterian

and several flakes were then selected for use elsewhere. Despite a different level of planning and production in Levallois technology, there is no reason that the same process in terms of automated, procedural knapping and tool selection (either blanks or unretouched flakes) was not taking place. In other words, just as we have to beware of the “finished artefact fallacy,” we should also be aware of the “finished core fallacy.” With the other techniques there must also be doubts about intention. The apparent removal of Kombewa flakes can also be a by-product of the knapping process, where a single blow can remove both the flake and a flake resembling a Kombewa flake from the same point of impact. This is potentially the case in figure 1 (11 and 12). In terms of the original and broader definition of Kombewa, in which flakes were removed primarily from the lateral edges and ventral face of another flake, there is a reasonable certainty of their having been intentionally knapped (Owen 1938). With truncated-faceted pieces, as Dibble and McPherron acknowledge, it is not known whether they were tools or cores. It would have been interesting to see more figures for each of the levels, such as average lengths for each major artefact category. For example, how do the average lengths of Levallois and Kombewa flakes tie in with the figures obtained from the cores? One intriguing statistic that can be derived from the figures presented is the apparent production of up to 11 Levallois flakes from each core, including roughly one per core that is retouched (see my table 1). This in itself is perhaps not surprising, but it underlines the problems of identifying intentional Levallois products as opposed to by-products. In contrast to this there is a deficit of 40% (14) in the number of Kombewa flakes in relation to Kombewa cores. Have these flakes been taken away, or does this reflect a problem of identification? The suggestion that unretouched flakes (perhaps sometimes small) were used is not in doubt, but whether there was planned production of small flakes remains unanswered. In the absence of use-wear, any attempt to quantify the extent to which unretouched flakes were used is clearly problematic. As I am sure Dibble and McPherron would agree, this also draws attention to the problem of using simply quantitative differences in tool types between assemblages when attempting to identify cultural groupings in the Mousterian.

Abdeljalil Bouzouggar National Institute of Archaeological Sciences and Heritage, Av. John Kennedy, 10 000 Rabat, Morocco (bouzouggar@ menara.ma). 29 IV 06 Despite the wealth of evidence for the Middle Palaeolithic in France, surprisingly little is known about the Asinipodian (Bordes 1975), and Dibble and McPherron’s paper is helpful in redressing this imbalance. It represents a fresh opportunity to discuss and share information about small cores and flakes,

785

Table 1. Levallois and Kombewa Flakes, Retouched Flakes, and Cores from Level IV-A at Pech de l’Aze´ IV

Levallois Kombewa

Flakes

Retouched Flakes

Cores

816 18

68 3

80 35

Note: Source: Dibble and McPherron’s figures 2 and 3 and text.

a topic I myself have been studying for many years, though in a totally different geographical, chronological, and cultural context (Bouzouggar 1997a, 1997b; Bouzouggar, Kozlowski, and Otte 2002). I especially welcome their work because the data concerning this technology come from Pech de l’Aze´ IV, a cave with a long stratified record of human activity and well-preserved organic remains. The site was recently excavated by Dibble and McPherron, and I had the opportunity in 2002 to observe the high level of precision of field recording using the total-station system—a system I have subsequently adopted in my excavations in Morocco. However, there are a number of points arising from their paper that require further explanation and clarification. To begin with, the Asinipodian is still not fully explained in terms of its technological, chronological, or palaeoenvironmental contexts. As far as the technology is concerned, for Dibble and McPherron the Asinipodian is a technology exhibiting basically Levallois and Kombewa techniques and the truncated-faceted method. Each of these techniques produces small flakes. It is equally important, however, to know the function of these small flakes. Secondly, it would be inappropriate to extrapolate the definition of the Levallois to the small Levallois cores without controlled experiments. I am not suggesting creating a new term but simply pointing to the need to verify whether the same definition fits the small cores or not. Another area for discussion concerns the distinction based on technological criteria between the small Levallois flakes and those originating from truncated-faceted cores (McPherron, Soressi, and Dibble 2001, 21). Here again replication studies are needed, as well as analysis of the use wear on both flakes and cores (especially Levallois and Kombewa cores) of this technology. Larger nodules are not abundant around Pech de l’Aze´ (Dibble 1985, 392) but a persistent problem concerns the independent means of estimating past raw-material availability, an issue raised by Dibble (1991) many years ago. However, very little evidence exists for the occupational intensity of the site. The increasing number of small cores and flakes during this occupation could be connected with the fact that the site was a potential source of residual raw material, as has been suggested for the Middle Paleolithic of layer 10 at Roca Dels Bous northeastern Spain (Mora, de la Torre, and Moreno 2004), and probably exploited for making expedient tools. Without seeking connections between the different worlds of European Neandertals and North African Homo sapiens, small cores also exist in the Aterian (a subdivision of the

786

Middle Paleolithic/Middle Stone Age in North Africa) [Antoine 1938 and Tixier 1958–59]), and I have suggested a pattern of exploitation based on experimental knapping in which it appears that small pebbles were broken into two pieces with the anvil technique before the preparation of the Levallois core surface (Bouzouggar 1997a, 1997b). In the Aterian, some of the small flakes were transformed into pedunculate pieces that were probably hafted but left no wear traces (de Igreja Araujo, personal communication). Another idea that needs to be verified is whether small cores (especially Levallois examples) were themselves made as tools because the scar of the preferential removal would allow easy manipulation with the thumb (Antoine 1938, 19). I am in basic agreement with Dibble and McPherron in that they do not attempt to consider the Asinipodian as a new cultural facies, since it is likely that the Mousterian of the Old World is probably a highly flexible technology. Much of the interest in the Old World Mousterian lithic technology has tended to concentrate on the very narrow thematic issue of retouched tools rather than considering a broader set of variables such as the presence of small flakes and cores. As mentioned above, Pech de l’Aze´ IV continues to offer rich sources of multiproxy data for palaeoenvironmental and chronological studies (McPherron, Soressi, and Dibble 2001). An examination of the detailed palaeoenvironmental evidence should allow fuller consideration of questions regarding the rates and scales of climatic and landscape change and how these might have impacted human behavior. This is probably an important gap in our knowledge concerning the Asinipodian, and I hope it will be filled with new studies now in progress.

Steven E. Churchill Department of Biological Anthropology and Anatomy, Box 90383, Duke University, Durham, NC 27708, U.S.A. ([email protected]). 30 IV 06 Dibble and McPherron are to be commended for an interesting article on the possible deliberate production of very small flakes in the Mousterian at Pech de l’Aze´ IV and a worthwhile discussion of the difficulties inherent in identifying the correct units of analysis when interpreting the lithic record. They focus on the importance of recognizing small flake production in the Mousterian as a prerequisite for understanding assemblage variability, but I would say that its importance goes beyond the way we interpret Mousterian assemblages. When it comes to the European record, the production and use of microliths—Mode 5 technology as identified by Desmond Clark (1968)—has been confidently identified only in Upper Paleolithic assemblages. Relatively large microliths (2.0–2.5 cm in length, similar in size to the possibly purposeful small flakes at Pech IV), appear late in the Middle Stone Age of southern Africa (see McBrearty and Brooks

Current Anthropology Volume 47, Number 5, October 2006

2000), and the production of composite tools with hafted microliths is generally thought to be part of the modern human behavioral “package” that was brought into Europe in the hands of Homo sapiens. Documenting microlithic technology among Neandertals (the presumed makers of the Mousterian at Pech IV) would contribute to our understanding of their cognitive abilities and adaptive strategies and remove from consideration one of the technological behaviors that seemingly separates modern humans from Neandertals. Thus the discovery of mircolithic technology in the French Mousterian would be of great significance. But here the old saw about extraordinary claims’ requiring extraordinary evidence applies, and the evidence in this case does not seem up to the claim. Dibble and McPherron recognize that small flakes are a ubiquitous by-product of lithic reduction and therefore their presence in an assemblage is not evidence in and of itself that small flakes were the desired end product of knapping. Bona fide microlithic tools of the Upper Paleolithic and Mesolithic of Europe, even of the earliest Upper Paleolithic, are recognizable end products of reduction because they show evidence of use in the form of either retouch or edge wear (e.g., retouched Dufour bladelets in the basal Aurignacian). Dibble and McPherron are not able to present such direct evidence in the Pech de l’Aze´ case. The evidence they do cite—the presence of very small Levallois cores, Kombewa cores, and truncated-faceted pieces—is problematic. As they note, the truncation and faceting of flakes may represent basal thinning to facilitate hafting, and the Kombewa “cores”—flakes that are considered “cores” on the basis of a single flake removal— are also parsimoniously interpreted as flakes subjected to basal trimming. The suggestion that truncated-faceted flakes and Kombewa cores served for the production of microliths would be more compelling if all of these pieces were systematically examined and found to be free of traces of hafting resins or wear polish. In both cases one has to wonder why knappers would have gone to the trouble to produce flakes from which they then struck one or two microliths rather than simply driving small flakes off of small nodules. The interpretation of Kombewa flakes is equally problematic, and more evidence is needed to demonstrate that they are not simply debitage from reduction of the bulb of percussion (3 of 21 Kombewa flakes show signs of retouch, which is consistent with the intensity of retouch on Levallois flakes from the assemblage and may support the claim that they are intentionally produced microliths, but what percentage of the Kombewa flakes evinces use wear?). Finally, the existence of small Levallois cores is not in and of itself compelling evidence for microlith production. Dibble and McPherron argue that the generally small size of the flake scars on these cores reveals a desire on the part of the knapper to produce small flakes. But what matters is not the size of core preparation scars (which simply reflects the average size of debitage) but rather the length of the privileged flake. Judging from the cores illustrated in figure 1, the smaller of the two cores illustrated would have produced

Dibble and McPherron The Missing Mousterian

privileged flakes on the order of about 3 to 4 cm in length— small but not microlithic. It does raise the question why Neandertals at Pech de l’Aze´ were interested in relatively small flakes (Dibble and McPherron make clear that it is not a case of limitations in the size of the raw materials as it is, for example, in the pebble-derived Italian Pontinian [Kuhn 1995]), but the evidence as presented does not seem to warrant the claim that there is an important microlithic component to the Mousterian that we have been missing all these years. The suggestion of a microlithic component to Mousterian assemblages is an interesting one, and further investigation of this possibility may ultimately lead to the kind of direct evidence required to substantiate the claim.

Vladimir Doronichev and Lubov Golovanova Institute for the History of Material Culture, Dvortsovaya nab. 18, 191186 St. Petersburg, Russia ([email protected]). 28 IV 06 Dibble and McPherron’s main goal is to demonstrate that the production of small flakes was a deliberate result and not simply a by-product of the knapping or retouching of other artifacts. Given the impossibility of drawing conclusions from an analysis of the small flakes themselves, they examine the pieces from which small flakes were struck—small Levallois cores, small Kombewa cores, and truncated-faceted pieces. They accept a priori that these three artifact classes were used to produce small flakes while differing from one another in the mode of production. The main problem here is that they represent either desired or undesired/reduced core types, desired or undesired/unfinished tool types. Originally Levallois and Kombewa were described as special types of large flakes and then their definitions were applied to particular flaking techniques for producing them. Similar morphology does not mean that small Levallois and Kombewa pieces are desired cores or represent the same deliberate method of flake production as large Levallois or Kombewa cores. This is best understood when we turn to the third group, truncated-faceted pieces. Contrary to the case with the previous groups, the term “truncated-faceted” was originally applied to small artifacts variously interpreted as results of thinning for hafting, as tools in themselves, or as cores. From the analysis of the Pech de l’Aze´ IV material Dibble and McPherron make two observations that contradict their hypothesis that small Levallois, Kombewa, and truncated-faceted pieces were core types deliberately made to produce small flakes. (1) As figure 2 shows, there is a similarity in the distribution of scar lengths between these three groups and “other” cores, suggesting that small flakes were struck from many if not all cores in this assemblage and therefore that there is no direct correlation between scar length and any core type. (2) Although the Asinipodian Level IV-A has the highest percentage of these small cores, including 50% of the

787

combined distribution of Kombewa and truncated-faceted pieces and small cores throughout the Pech IV sequence (fig. 3), higher percentages of small flakes occur in other assemblages in the sequence (fig. 4). Again, there is no direct correlation between distributions of small cores and small flakes. Finally, the idea of identifying any universal “function” for all morphologically similar lithic pieces seems faulty. On the contrary, one can suggest functional variability within artifact classes in temporally and geographically different archaeological contexts. For example, in the Zagros Mousterian assemblages of the southern Caucasus the truncated-faceted technique was used to accommodate hafting of the bases of points and other tools, while in the eastern Gravettian of the Russian Plain this method was exploited to rejuvenate the cutting edges of Kostenki knives. In Russian Paleolithic archaeology, the truncated-faceted technique was originally identified as a “corelike method of thinning.” Also, the Kombewa technique, though rarely discussed in Russian archaeology, is generally considered a method of reducing the bulbar part of a flake to accommodate hafting. Regarding small Levallois cores one can suggest that some of them were reduced centripetal cores, especially in industries with no large Levallois cores or flakes such as the Asinipodian level of Pech IV. In true Levallois assemblages with a variety of large, medium, and small Levallois flakes and cores, the latter were apparently preferentially used for removing larger flakes of predictable morphology. In any case, the study of small tools/ flakes is related to two general problems: (1) the existence of special (hence culturally specific) lithic technologies for producing small flakes or the deliberate exploitation of small flakes made from scarce high-quality raw material and (2) the absence of most small lithics in most old excavations and the poor preservation of many Paleolithic collections. In general, Dibble and McPherron’s approach is a very important attempt to rejuvenate Paleolithic research paradigms and analytical procedures.

Metin I. Eren Department of Anthropology, Southern Methodist University, Box 750336, Dallas, TX 75275-0336, U.S.A. ([email protected]). 17 IV 06 Dibble and McPherron’s lithic analysis at Pech de l’Aze´ IV shows how detailed reduction models can delineate prehistoric knapping behaviors. Without use-wear or residue analysis, proving that small, unretouched blanks are intentional products is no simple task, and they should be applauded for their efforts. Rather than using a static typology of assumed end products, they employ a dynamic typology of process that identifies an overlooked Middle Paleolithic knapping strategy. Their study has implications for how lithic analysts recognize meaningful type categories, and their methods should be emulated.

788

Still, I have a few concerns. Although they mention that “there is no independent means of assessing raw-material availability and size during the various occupations” at the site, raw-material availability and transport may present larger problems than they concede. Small flake production may indeed have been intentional (after all, small flakes were produced), but intentionally producing a small flake because one’s hand is adapted to holding it and intentionally producing a small flake because one is running out of raw material are very different. Although there is a large number of small cores in Level IV-A (fig. 3), the average tool and blank lengths are not the smallest when compared with other levels. The authors explain that “larger flakes were also being produced in sufficient numbers and sizes to maintain a relatively high average length of all flakes in the level [IV-A].” It would be good to know where these large flakes came from. Are there large unexhausted cores in Level IV-A, and, if so, how many? The presence or absence of large cores may strengthen or weaken the argument. Three scenarios illustrate the point: Scenario 1: Large cores are present. Limited raw-material availability is not responsible for small flake production, since the option for creating larger flakes exists. Scenario 2: Large cores were subsequently transported offsite. Though these produced enough large flakes to increase the overall blank size, the cores were still large enough to be taken away for future use. If this is the case, the blank-tocore ratio will be even lower than figure 5 indicates, again implying that limited raw-material availability is not responsible for small flake production. Scenario 3: If large cores are not present in Level IV-A and were not transported off-site, the only possible explanation for the presence of large flakes is that large cores were reduced into small cores (not including Kombewa or truncated-faceted pieces). This implies raw-material shortage, indicating that many of the small flakes were probably by-products of large flake production and only some of them were intentionally produced toward the end of the reduction sequence. While there is no way to detect scenario 2, the presence or absence of large cores coupled with an analysis of large blanks may help distinguish between scenarios 1 and 3. Of course, there is probably some degree of overlap among them. The authors state, “Most Middle Paleolithic assemblages contain more or less the same range of major classes of tool types and technologies but differ in terms of the relative frequencies of those classes.” If this is indeed the case, it would be interesting to know more about small flake production in other levels at Pech de l’Aze´ IV. While the actual amount of small flake production in other levels is obviously much less than in Level IV-A, how does the intensity of production compare? For example, Levels II-A and II-B exhibit only 34 cores between them, but if, say, 30 of these 34 cores are small cores used for producing small flakes and the intensity of small flake production in Levels II-A and II-B is comparable to that in Level IV-A, new questions arise. Hypothetically, if

Current Anthropology Volume 47, Number 5, October 2006

small flake production intensity is similar between Levels IIA, II-B, and IV-A, and import and export processes are indeed similar among all the levels as stated, why do Levels II-A and II-B have high core-to-blank and blank-to-tool ratios (fig. 5) while IV-A does not? Discussion of other levels may provide insight and context for the patterning seen in Level IV-A. It remains uncertain whether truncated-faceted pieces are cores. Dibble points out elsewhere that there is no consensus on whether they were used for producing small flakes or for producing a particular kind of edge (De´benath and Dibble 1994, 123). One wonders if use-wear or residue analysis on such pieces might distinguish between these two possibilities. Finally, Dibble and McPherron state, “Many of the small Levallois cores from the Asinipodian levels were made on very small knobs of flint struck from larger nodules.” How many exactly? This has implications for how many Levallois cores were reduced from larger sizes and thus how many small flakes are possibly by-products.

Bruce L. Hardy Department of Anthropology, Kenyon College, Gambier, OH 43022, U.S.A. ([email protected]). 30 IV 06 Dibble and McPherron are to be commended for thinking beyond the normal constraints of typology. Although typology has certainly been useful in many ways, it has also served as blinders for Paleolithic research. While it is a necessary evil, it is not an end in itself. While they provide technological evidence for production of small thin flakes from truncated-faceted pieces, Levallois flakes, and Kombewa flakes, the only information they present relative to functional analyses is that “preliminary mirowear analyses of 15 pieces from Pech IV yielded only two that showed signs of having been used for short-term cutting of soft materials.” To be fair, they recognize that this is insufficient, stating that “this is clearly an area of research that should be pursued further.” However, their case for the “missing Mousterian” would have been greatly strengthened had some form of functional analysis (either use-wear or residue analysis) been conducted. Citing Niewohner (2001), Dibble and McPherron suggest that Neandertal hand anatomy would not have been efficient at oblique power grips and that Neandertals therefore would have been more efficient with hand-held than with hafted tools. How efficient does our anatomy have to be in order to perform a task? Our own modern human hands may be less efficient than Neandertal hands in performing tasks involving power grips, but this does not preclude our performing tasks with a power grip. Despite the presumed inefficiency of Neandertals’ oblique power grips, numerous researchers have found evidence for hafting in the Mousterian (e.g., AndersonGerfaud 1990; Beyries 1988; Boe¨da et al. 1996; Hardy et al. 2001; Shea 1988). Furthermore, since Dibble and McPherron

Dibble and McPherron The Missing Mousterian

are willing to recognize nontraditional typological units, it may be worthwhile to consider nontraditional hafting methods as well. For example, small flakes may be placed in a stationary substrate (such as a piece of wood) and the material to be processed passed over them in a manner similar to the use of a rasp or grater (see Kardulias and Yerkes 1996 for a discussion of a similar mounting method for microliths in threshing sledges). While I am certainly not suggesting that Neandertals were using threshing sledges, given the lack of functional evidence for these small flakes this proposed type of “hafting” or mounting arrangement has no more or less evidence to support it than hand-holding. Without other sources of data, both hypotheses are simply possibilities. Experimental replication and use both hafted and unhafted could provide valuable information on the feasibility and use of the “missing Mousterian” flakes. We await functional evidence that will confirm that these small, thin flakes were used as tools.

Peter Hiscock Department of Anthropology, School of Archaeology and Anthropology, Australian National University, Canberra, ACT, Australia ([email protected]). 5 V 06 Perceived anomalies often promote reassessment, and discussion of the seemingly obscure Asinipodian industry with data from Pech de l’Aze´ IV has enabled Dibble and McPherron to raise fundamental questions about Middle Palaeolithic research. Debates over behavioural “modernity” often invoke depictions of diversity and planning represented in Middle Palaeolithic industries, making the claim that knappers often made large numbers of small, usable flakes and that they deliberately pursued reduction strategies producing small flakes a challenging one. Dibble and McPherron observe that categories of cores and retouched flakes have similar-sized negative flake scars and covary in abundance through the archaeological sequence. This is important and forms the basis for questioning purported functional and ideational differences between such categories. Their analysis is compelling for Pech IV. Indeed, they may underestimate the similarity, since measuring only the longest scars on the surface of “principal” removals on Levallois cores probably exaggerates scar-size differences with Kombewa/truncated-faceted pieces. Dibble and McPherron argue that the size similarity of flakes struck from those objects implies that (a) the principal role of Levallois, Kombewa, and truncated-faceted objects was to produce small flakes and (b) such flakes were frequently produced and used in the Middle Palaeolithic without further retouching. These propositions are plausible, although further wear/residues studies are desirable. In late Holocene Australia (e.g., Hiscock 1996, 2006), small flake production and use was a dominant technological strategy, and there is no obvious reason such a strategy could

789

not have been advantageous for Mousterian knappers. I therefore accept the inference that at some sites and in some periods the manufacture of small flakes was a feature of Middle Palaeolithic activities. Further discussion about the level and nature of intentionality is warranted. Dibble and McPherron, like conventional typologists, assert that without knowledge of the knapper’s goal descriptions of manufacturing processes are “irrelevant.” This is a radical statement, because many questions about past economy and technology can be answered without knowledge of the knapper’s thoughts. Yet the primacy of mental depictions of knapping is central to past typological frameworks and to this reconsideration. Dibble and McPherron recognize that the detection of purpose is seldom easy, raising the methodological quandary of how to recognize intentionality behind material entities created in alien cultural systems. The criteria used to illustrate how deliberateness can be perceived can be applied only in some situations and are not unambiguous. Middle Palaeolithic assemblages in the Dordogne contain evidence for extensive scavenging and recycling of artefacts over long time periods, and elsewhere there are examples of retouching and flake selection that cannot be taken to reveal the nature of knapper’s intentions in any direct way (see Hiscock 2004). Scar-size similarity on objects belonging to several typological categories is employed by Dibble and McPherron as evidence that they were producers of flakes for use. However, overlaps in the size of flakes removed from each class could reveal that artefacts created for different reasons shared characteristics. If some Kombewa or truncated-faceted pieces were tools, others may have yielded flakes that were used; these intentions need not covary with typological categories, creating problems of equifinality for analysts. While transverse or notched scrapers from Middle Palaeolithic assemblages have significantly smaller average scar sizes, some specimens have scars more than 40 mm in length. Were these scrapers sometimes tools and at other times producers of usable flakes (see Hiscock and Clarkson n.d.)? These are examples of how a multiplicity of intentions might exist within a single typological category and why there is no need to explain all specimens in an implement or core class by reference to a single goal. The two models Dibble and McPherron discuss— removing flakes to use versus removing them to attain another goal—need not be viewed as mutually exclusive; they may both be correct for different specimens or assemblages. Similar systems of core/flake reduction to that of Pech IV could have been employed for different ends as the contexts of foraging life varied. Small-flake production may have created tool edges in one situation, prepared core morphologies in another, flakes for use as tools in a third, or all of these in another. Behaviours involved in making and using stone artefacts may have been diverse, and that diversity may not be adequately characterized by trying to match typological categories with particular knapping intentions. An alternative approach, using technology and wear/residue to infer ancient behaviours

790

for each artefact and assemblage, can document rigid, normative practices where they exist but also recognize flexible strategies for artefact production and use where they occurred. Observations of systematic small flake production are valuable in expanding debates about the diversity and complexity of early lithic technologies. Dibble and McPherron have offered us a fresh and provocative hypothesis that will help stimulate a much-needed reassessment of Middle Palaeolithic technology and life ways.

Simon Holdaway Department of Anthropology, Private Bag 92019, University of Auckland, Auckland, New Zealand (sj.holdaway@ auckland.ac.nz). 2 IV 06 A Mousterian facies characterized by small flakes, manufactured using a variety of core technologies that span Levallois, Kombewa, and truncated-facetted technologies, that lacks large numbers of tools but retains core technologies for the production of large flakes is a curious phenomenon. If this assemblage occurred in a different time and place, such as the Australian late Pleistocene, it would look decidedly nonMousterian. To be sure, the “Asinipodian” apparently lacks the range of formal tool types that characterizes more recent assemblages, those identified as Upper Paleolithic, but so do a range of assemblages from late Pleistocene contexts from places other than western Europe. If, as Dibble and McPherron imply, the “Asinipodian” from Pech de l’Aze´ IV is unusual only in that archaeologists have failed to identify the small flake component in other assemblages they have studied, one is left to wonder just how different the Middle and Upper Paleolithic really are, at least on the basis of stone artifact comparisons. But at the heart of Dibble and McPherron’s article is a more fundamental critique of approaches to Paleolithic studies. The authors comment that most studies of the Middle Paleolithic stone artifacts are based on one of two approaches: comparing the frequencies of different classes of stone artifacts (i.e., types) or examining the technological sequence of production for certain classes of artifacts (i.e., chaıˆne ope´ratoire). Despite claims to the contrary, for some of us who work outside the European Paleolithic it is very difficult to see how these two approaches are substantially different. Typological analysis limits consideration to a select group of retouched tools and “predetermined” flakes whose identification can be traced back to the nineteenth century. Studies of the technology of production likewise limit analysis to certain “predetermined” forms whose identification has a long history. Both forms of analysis rest on the assumption that the artifacts identified for analysis are those that convey the most useful information for interpreting Middle Paleolithic stone artifacts. Dibble and McPherron’s study challenges this assumption by identifying a group of artifacts that were “intended” but have

Current Anthropology Volume 47, Number 5, October 2006

largely been ignored until now. If they are able so easily to identify a group of artifacts ignored for over a century, one may well ask what else is left unobserved by Middle Paleolithic systematics of either variety. Working in places and times in which the “intention” of stone knappers in the past is much less clear, the systematics of the Middle (or, indeed, the Upper) Paleolithic has seemed very neat and tidy. Retouched tools form clear typological sequences. Technological sequences can be reconstructed from raw-material source through to discarded artifact as a prescribed series of events. In comparison, the Australian record with which I deal seems much messier. We have few retouched tools on which to build a typological systematics and lack the ability to “see” intention through the study of single technological reduction sequences. Certainly we can study rawmaterial sources, core reduction strategies, and artifact discard, but which of the numerous flake end products were the ones the Aboriginal knappers intended to produce? Australian studies lack the depth of history of the French Paleolithic, so perhaps we lack the number of studies needed to make clear which artifacts retain the most information and therefore should be the focus of our investigations. But Dibble and McPherron’s work suggests an alternative explanation. The reason that the “Asinipodian,” with its small flakes, multiple core types, and few tools with many possible “intended” end products, looks familiar to us is that it has become, through Dibble and McPherron’s work, the type of messy assemblage that we are familiar with, one that fits not at all well into Paleolithic systematics. Is the orderly progression of Paleolithic assemblages, whether characterized by tool type frequencies or technological reduction sequences, a product of the data or of the methods employed to characterize them? If Australianists were to limit their analysis to a handful of retouched types or “intended” end products, might not our systematics look neat and orderly as well? This is the challenge raised by Dibble and McPherron’s paper, one that will inspire us all to reassess what neat and orderly systematics really means.

Wesley A. Niewoehner Department of Anthropology, California State University, San Bernardino, CA 92407, U.S.A. ([email protected]). 28 IV 06 Dibble and McPherron present an interesting case that small unretouched flakes are erroneously ignored as a source of information on Neandertal behavior. The fact that Neandertals were physically capable of working with small flakes is confirmed by the Pontinian, a facies of the Mousterian in Italy (Taschini 1979). Dibble and McPherron probably chose not to discuss the Pontinian because there is no disagreement that these are “real” tools; nevertheless, there is valuable information here. Most of the tools, produced from flint peb-

Dibble and McPherron The Missing Mousterian

bles, are quite small (ca. 3.5 cm in length) and retouched (Kuhn 1992, 1995). I am unaware of evidence that the Pontinian flakes were regularly hafted, so we can assume that they approach the lower limit of flake-sizes that Neandertals were able to use in an unhafted state. It appears to me that their hand skeletal anatomy would not have prevented them from working with the even smaller pieces indicated by the Asinipodian material. We must keep in mind, though, that we have no direct way of assessing a crucial limiting factor of the Neandertals’ ability to use diminutive pieces of stone— their degree of neuromuscular control of fine finger movements. That said, in their total morphological pattern Neandertal hands are exquisitely suited for habitually and effectively using small unhafted flakes. Anatomical correlates of this conclusion are osteological indications of generally hypertrophied hand musculature that likely included the digital flexor muscles and dorsal interosseous muscles, both of which aid in finger flexion. Add to this their expanded distal tuberosities (producing wide fingertips) and they present a functional complex that would have conferred the ability to grasp firmly objects against any of the fingertips. Additionally, if flakes were preferentially held between the thumb and the side of the index finger, their aforementioned hypertrophied finger musculature, combined with powerful flexion of the thumb, must have imbued Neandertals with a viselike pinch grip. Thus I agree that Neandertals were likely capable of using these small flakes, but only a properly constructed archaeological analysis can hope to settle the fundamental question: If the flakes were intentionally produced as tools, then what were Neandertals doing with minuscule flakes that could not be accomplished with larger ones? One interesting aspect of Neandertal hand functional morphology is the maintenance of significant mechanical advantages at the base of the thumb (i.e., the metacarpophalangeal joint) and on both the ulnar and radial sides of the wrist relative to Upper Paleolithic specimens (Niewoehner n.d.; Trinkaus and Villemeur 1991). This indicates that when it came to performing manipulative behaviors that required thumb circumduction and flexion, cupping of the palm, flexion of the wrist, and adduction of the hand, Neandertals had (or required) significant strength advantages relative to Upper Paleolithic peoples. This discussion and my interpretation of the literature on microwear analyses of Mousterian tools suggest to me that if Neandertals occasionally used hafted flakes they were probably set into the distal aspect of the handle, because using a tool hafted in this fashion demands significant muscular stabilization of the wrist. An additional inference (which really qualifies as speculation) is that many aspects of Neandertal hand anatomy, including those mentioned above, may relate to a high frequency of woodworking (Niewoehner 2001). Dibble and McPherron mention that preliminary microwear analyses fail to support this interpretation, but it appears to me that Neandertals were habitually engaging in behaviors that necessitated repetitive hand motions in the same plane such as scraping, whittling, cutting, and possibly

791

percussion (Niewoehner n.d.). Of course, many upper limb behaviors require these motions, from working wood to scraping skins to flint knapping. It merely strikes me that another “missing” piece of the Mousterian may be a reliance on worked wood. Here it is clear that the limits of skeletal and lithic microwear analyses intersect, preventing us from making detailed inferences regarding specific behaviors. Dibble and McPherron’s assertion that the Asinipodian represents a geographically widespread Neandertal predilection for deliberately making small flakes using a variety of knapping techniques deserves serious investigation if only because doing so requires that we take a fresh look at how they go about conducting lithic analyses. In this context, Dibble and McPherron rightly question the validity of the fundamental assumption that analyzing frequencies of artifact classes or types necessarily gets us closer to understanding Neandertal behavior. The use of behaviorally meaningful units of analysis should bring us to a more comprehensive understanding of Late Pleistocene human behavioral variability through the improved integration of skeletal with archaeological analyses. Both disciplines have a long way to go given that we are often faced with dichotomous behavioral interpretations (e.g., Niewoehner 2001) and it is difficult to determine whether one (or both) of the interpretations should be discounted or whether there is a basis for their reconciliation.

John J. Shea Anthropology Department, Stony Brook University, Stony Brook, NY 11794-4364, U.S.A. ([email protected]). 29 III 06 Dibble and McPherron argue that Middle Paleolithic assemblages from Pech de l’Aze´ contain unretouched stone artifacts smaller than 25 mm in length that were used as cutting tools. This is not a surprising conclusion. The ethnographic record is rich in accounts of small, unretouched stone tools’ having been used for various cutting tasks. Microwear analyses of countless Pleistocene and Holocene assemblages have shown that many stone tools were used without having been retouched (Odell 2001). African prehistorians have long recognized small tool production in Middle Stone Age assemblages (Ambrose 2002). If this conclusion is a surprise, it will be so only to the steadily diminishing contingent of archaeologists for whom Bordes’s typology is the principal analytical framework for understanding European Middle Paleolithic industrial variability. Misplaced reverence for this typology and analogous typologies in other prehistoric contexts remains a significant obstacle to making the Paleolithic stone tool record relevant to important questions in human origins research (Bisson 2000; Dibble 1995a). The recognition that artifacts smaller than 25 mm were made and used systematically from at least Middle Paleolithic times onward has significant implications for how we ar-

792

chaeologists treat the stone artifacts recovered by our excavations. The overwhelming majority of stone artifacts recovered from most excavations are flakes and flake fragments less 25 mm in length. If there are significant proportions of use-worn stone tools among small flakes and flake fragments, then recovering these data is going to require a vast increase in the amount of time archaeologists spend scrutinizing these artifacts. The all-too-common practice of placing all such “debris” from the same excavation unit together into large plastic bags will also need to be changed, because stone-onstone abrasion can destroy wear traces on the edges of stone tools. I am in full agreement with the larger argument in Dibble and McPherron’s paper—that archaeological “mythology” adversely influences our understanding of the past. The common practice of dividing lithic assemblages into “tools” “byproducts/de´bitage,” and “waste/debris” reflects an essentialist and fundamentally antievolutionary perspective on technology. Stone tool production, like any other aspect of hominin behavior, was under selective pressure from its inception in Late Pliocene times onward. Small stone artifacts constitute potential cutting edges obtained at an energetic cost. It is only reasonable to expect there to have been selective pressure for hominins to recoup this cost by using small stone tools. Thus, using small flakes makes sense in evolutionary terms. The least satisfying aspect of this paper is that, although Dibble and McPherron make a strong case for rejecting rawmaterial scarcity and prolonged curation as factors in the Pech de l’Aze´ sample, they do not propose an affirmative hypothesis for the origins and variability of these artifacts. One possibility, evidently not considered, is that these small tools were produced by children. Aptitude in stone tool production was undoubtedly an important skill for Pleistocene hominins. It seems reasonable to suppose that they practiced such skills from an early age, imitating the gestures of their parents and other adults and producing similar, if smaller, lithic artifacts. Small size and expedient manufacture are characteristics associated with toys in other archaeological contexts (Baxter 2005, 46–50). Residential sites, including caves, and other sites with high archaeological visibility would have been among the few places where children could have knapped stone without interfering with important subsistence tasks. Unless the context for the social transmission of technical knowledge in the past was fundamentally different from that of recent human societies, Middle Paleolithic assemblages from cave sites probably contain substantial numbers of artifacts made and used by children. Is children’s knapping a plausible explanation for the small Mousterian tools that Dibble and McPherron describe? It is hard to say. The proportion of such children’s knapping products probably varies with site layout, raw-material availability, and other situational variables, as well as with differing archaeological recovery techniques. Unfortunately, very few actualistic studies have been conducted on how children affect the lithic record. Studies of children in stone-tool-using pop-

Current Anthropology Volume 47, Number 5, October 2006

ulations or of avocational flint knappers may shed light on the production of small tools in Paleolithic assemblages.

Silvia Toma´sˇkova´ Curriculum in Women’s Studies and Department of Anthropology, 301 Alumni Bldg., CB #3115, University of North Carolina, Chapel Hill, NC 27599-3115, U.S.A. ([email protected]). 2 V 06 By restricting themselves to the Middle Paleolithic Dibble and McPherron overlook literature that has raised similar issues in other contexts with more extensive investigations of actual data sets. Belfer-Cohen and Goren-Inbar, working in the Lower Paleolithic of the Levant, have discussed Levallois and Kombewa techniques specifically as ways of manufacturing both handaxes and unretouched flakes (Belfer-Cohen and Goren-Inbar 1994; Goren-Inbar and Saragusti 1996). Clearly distinguishing between finished tools and production techniques, they address the mental capacities of Lower and Middle Pleistocene hominids with reference to choices of raw materials and flake sizes. More important, they stress the need for attention to reduction sequences over time and between different assemblages to gain the sort of behavioral insights that Dibble and McPherron seem to seek (Belfer-Cohen and Goren-Inbar 1994, 151–3l; for a careful study employing a chaıˆne ope´ratoire approach in a Middle Paleolithic context, see Van Peer and Wurz 2006). Other studies have also dealt with debitage, if not specifically addressing Middle Paleolithic deposits, and suggested the benefits of attention to different regions and time periods beyond the local variants (see, e.g., entries in Andrefsky 2001). Sullivan and Rozen (1985), working in the American Southwest, offered an early thorough analysis of large amounts of core reduction and tool manufacture debitage. Whether or not one accepts their “interpretation-free categories that enhance objectivity and replicability” (Amick and Mauldin 1989; Ensor and Roemer 1989), they came to a general conclusion similar to Dibble and McPherron’s: that certain flakes were most likely intentionally produced for use without any further retouch (p. 766). Moss (1986) carried out a study of Paleolithic European collections, examining 126 pieces of debitage from Pont d’Ambon, 91 pieces from Klithi, and 72 pieces from Pincevent section 36 IV 30, and concluded that the extensive use of unretouched flakes could account for the limited range of tool types, for example, in the Azilian, linking the debitage analysis to a formal tool-typology discussion. Moss also suggested the possibility of different production techniques for debitage as opposed to tools depending on the intended end product. In my own microwear study of debitage from the Paleolithic collections at Pavlov (Czech Republic) and Willendorf II (Austria), I obtained similar results, showing some variation between different layers in the extent to which unretouched flakes showed traces of use (Toma´sˇkova´ 2000). Since both

Dibble and McPherron The Missing Mousterian

collections had a rich variety of Upper Paleolithic formal tools, my main goal was to investigate the broader question of unit analysis and to question the typological character of welldefined assemblages. Including debitage in the total count of lithics changed the proportions of formal tools quite dramatically and altered the “typical character” of both collections. Bisson (2000) has boldly suggested that it is time to abandon Bordes’s nineteenth-century classification, which no longer serves our needs and is actually preventing us from arriving at insight into prehistoric behavior. Dibble and McPherron, in their effort to show that the unmodified flakes at Pech de l’Aze´ IV were made intentionally to be used as tools, end up awarding the flakes the same status as tool types. While they caution that “this paper is not at all about whether a new ‘type’ should be added to Bordes’s type list,” by not actually examining the flakes for use and focusing only on the “reality” of these flakes they end up treating them as a tool-type category. This is unfortunate, given the potential implications of their study for units of analysis. Finally, I disagree with their statement that “it is irrelevant how certain things were made unless we know what things they were trying to make.” It may be the “finished artifact fallacy” that is haunting us here, since the recent turn toward techniques and life histories of artifacts has offered a muchneeded revitalization of lithic debates, following the “how” to its implications of decision making, divisions of labor, learning processes, or cognitive abilities and leaving the “what” behind. Indeed, not all tools may have been as certain as we usually take them to be. In the case of burins I have suggested that it could be irrelevant what the knappers were trying to make in that the technique, the how, may have been the point (Toma´sˇkova´ 2005). Judging from use-wear evidence, the formal category of burin in my study was a stage in tool production or modification as much as an end in itself. However, I could not have come to this conclusion without carrying out a detailed use-wear study. Dibble and McPherron present us with a very suggestive case based on small flakes from Mousterian layers. Absent further analysis of the lithics or the incorporation of other lines of evidence, however, the behavioral implications of their study remain unclear. Are these small flakes an example of a learning process of tool making, their actual use being irrelevant, or were they produced for a specific task and material? As Dibble and McPherron, speaking of use-wear, state, “this is clearly an area of research that should be pursued further.”

Gilbert B. Tostevin Department of Anthropology, University of Minnesota, 395 H. H. Humphrey Center, 301 19th Ave. South, Minneapolis, MN 55455, U.S.A. ([email protected]). 3 V 06 Dibble and McPherron should be applauded for tackling an epistemologically difficult task: the identification of specific

793

desired end products from the multitude of artifacts of a Middle Paleolithic assemblage. Not only are their arguments more methodologically rigorous than previous approaches but they see the larger implications of the uniqueness of the assemblage. They are thus able to use this example of unretouched small flakes’ being intended products in Level IV-A of Pech de l’Aze´ IV to broaden archaeologists’ views of what constitutes Middle Paleolithic assemblage variability and, potentially, hominin behavior. The paper is noteworthy for the way the authors demonstrate the intentionality of the production of the small flakes in the assemblage. Given the absence of retouch on these products and the nondiagnostic nature of the removals from truncated-faceted blanks, they rely upon the evidence of the cores, specifically the core-facets or scars, from which the products were removed rather than the products themselves. This is particularly progressive in comparison with other studies that assume that certain products were intended merely because of their situational role in the technological process of flake removal (e.g., Boe¨da 1994 see also Sandgathe 2004) or indeed that all products were intended, an epistemologically easy approach I myself have taken (Tostevin 2003). As a result, this is a refreshing paper, particularly given the authors’ clear concern for the difficulty in making such arguments (Bar-Yosef and Dibble 1995; McPherron 2000). The care Dibble and McPherron take in making their case unfortunately does not solve the general methodological problem for archaeologists working to identify similar products in other contexts. As they point out, it is the coincidence of high frequencies of small-sized Levallois products and Kombewa flakes, both diagnostic even when small, which allows for the identification of the intentionality of these products as desired blanks rather than ubiquitous small waste flakes. In contexts in which the dominant reduction strategy produces less diagnostic products, it would be harder to make the case for intentionality, as Dibble and McPherron note for the case of the products of truncated-faceted pieces. Further, if Level IV-A had 4 evidenced more raw-material exhaustion (a situation which Dibble and McPherron take pains to disprove), their case would have been harder to disassociate from general economization of raw material. Thus, while I agree that “a better understanding of the spatial and particularly the environmental distribution of small flake technologies could help address” the functionality of these small products, this is likely to be a much harder task everywhere except for Pech de l’Aze´ IV. Dibble and McPherron do, however, point to another research direction which could help us understand this phenomenon: “Experimental studies would be useful to address the relative merits of small, thin flakes versus the edges exhibited on other Middle Paleolithic elements, be they large unretouched flakes, scrapers, or notches.” This type of research has perhaps a better chance of success, particularly if the ergonomic and prehensile implications of such flakes (e.g., Tomka 2001) are taken into consideration. Relevant curation

794

models could also prove informative, given the fact that these small blanks fit the cutting-edge-per-gram predictions of Kuhn (1994) for curated tools. Further, experimental aggregate analysis following the direction espoused by Baumler and Davis (2004, 2000) may offer an independent approach to identifying such intentional small-sized debitage in assemblages which lack the unique context of Level IV-A. Specifically, Baumler and Davis’s application of Sullivan and Rosen’s (1985) typology to the aggregate analysis of small-sized debitage (between 0.16 and 2.57 cm) could isolate the products of truncated-faceted pieces from the products of the reduction of larger cores, even when the products fall in the same size (i.e., sieve) range. In sum, Dibble and McPherron are to be congratulated on their courage and methodological care in arguing that another, as yet unrecognized, “axis of assemblage variability” may exist in the Middle Paleolithic record. While most lithic analysts have long considered that any small flake could serve as an expedient cutting edge, the Asinipodian phenomenon does not match the current theoretical concepts related to expediency (e.g., Andrefsky 1994) and so requires another explanation.

A. Turq De la Pre´histoire a` l’Actuel: Culture, Environment, Anthropologie, UMR 5199 du CNRS, Avenue des Faculte´s, 33405 Talence cedex, France ([email protected]). 28 IV 06 The research presented by Dibble and McPherron is part of the current debate over the idea of systematic production of small tool blanks during the Middle Paleolithic. In the context of Level IV-a of Pech de l’Aze´ IV, their argument is positive and convincing. It joins other examples such as Coustal Cave (Fonton, Lhomme, and Christensen 1991) and Les Tares (Geneste and Plisson 1996) and supports the recently proposed hypothesis (Bourguignon, Faivre, and Turq 2004) that the ramified structure of Mousterian lithic production permitted the fabrication of small blanks, which were most often used unretouched. The desired products, around 2 to 3 cm long, were very variable—Levallois flakes, Kombewa flakes, pseudoLevallois points (Bourguignon and Turq 2003), and diverse bladelets (Newcomer and Hivernel-Guerre 1974; Delagnes 1992; Slimak 2004)—and were produced either from large unretouched flakes or from retouched tools. Could this production represent a forgotten Mousterian type or a technical behavior until now rarely or never observed in the Middle Paleolithic? These technical choices were always present but are particularly well developed in the context presented here. Why did this technological component remain unknown for so long? We must remember that for more than 30 years the Bordes method concentrated mainly on retouched tools. Among the unretouched objects, only Levallois flakes, pseudo-

Current Anthropology Volume 47, Number 5, October 2006

Levallois points, and naturally backed knives were included in the type list. Flakes smaller than the smallest retouched tools were considered knapping waste products and simply counted. As a result, Bordes identified only the small Levallois and Kombewa flakes at Pech de l’Aze´ IV and emphasized the particular nature of this industry, calling it the “Asinipodian.” The development of technological, lithological, and traceological methods have renewed our approaches to lithic industries. The smaller number of products relative to cores (Bourguignon and Turq 2003; Bourguignon, Faivre, and Turq 2004) and the presence of use traces have demonstrated that these small flakes were intentional products, which were probably exported and used away from the site of production. Certain objects until now considered as tools, such as notched pieces, denticulates, and notched scrapers, all with Clactonian notches (Bourguignon 1997), could be either tools (sensu stricto) or cores for the production of small flakes. Depending on the site context (occupation duration, activities), the same knapping technique could respond to one or several intentions. Whatever the case, along with Dibble and McPherron we believe that the production of small flakes must be taken into account in the definition of Middle Paleolithic industries.

Philip Van Peer University of Leuven, Redingenstraat 16, 3000 Leuven, Belgium ([email protected]). 20 V 06 I agree with Dibble and McPherron about how little we truly know about Mousterian life ways. In Palaeolithic archaeology’s core business, lithic analysis, some basic methodological tenets still need to be settled. Served by the small-flakes case, Dibble and McPherron illuminate once again the arbitrary composition of the classic type list. In their own words, their paper is about epistemology, and they do make an important point. It is in the demonstration of the intentional nature of small flakes itself that the contribution lies. Clearly, we are dealing with the nature of a basic unit of analysis on which generalizations are built and from which interpretations are ultimately derived. Dibble and McPherron begin by questioning the assumption that the types in present classification systems for the Mousterian represent behaviourally meaningful categories. It is clear from their discussion that they understand the latter as emic categories or intentional “finished forms” (Noble and Davidson 1996), and they are right that this assumption is by no means unproblematic. The ontogenetic nature of categories relating to the dynamic process of lithic technology has been much neglected in the construction and use of typological systems. Dibble (1986, 1988b) himself was the first to argue that the array of side-scraper forms is mostly a typological translation of the Frison effect. Schlanger (1991) has made the point that Middle Palaeolithic tools may really only

Dibble and McPherron The Missing Mousterian

exist when they are in use. Dibble and McPherron’s implication, obviously, is that these systems are heterogeneous assemblages of real types and others, a quality that is accordingly extended to the data they generate. Their point is that behaviourally meaningful categories should be established by inference and not just assumed. They do not, however, question the definition of “type” itself, their concern being to provide a procedure for its consistent implementation. Only the real conceptual categories guiding deliberate manufacture and resulting in discrete groups of objects should receive the status of a type. These are the behaviourally meaningful units of analysis. There is, obviously, an alternative definition of archaeological types in terms of formal categories (Dunnell 1971) constructed on the basis of observational properties of objects whose emic content is irrelevant. The property of “reality” might seem to constitute the major difference between the two definitions, but this depends on the yardstick that one wishes to use; insofar as formal categories enlighten us about Mousterian life ways, they are real and behaviourally meaningful. Furthermore, formal categories are at least analytically meaningful: they represent discrete groups of artefacts with common, objective features formulated according to a coherent underlying logic. Under this definition, the problem of finding the “right” types for, for example, the Mousterian type list is less acute, although their relevance can seemingly be established through an inferential process. The degree to which they reveal structure in the archaeological record is a measure of their utility. As hinted above, however, they have proven not to be behaviourally meaningful unless the nature of the relationship between structure and behaviour is known. The Mousterian in particular has amply shown us that the latter is a matter of interpretation. I agree with Dibble and McPherron that there are probably emic types in the Mousterian, but in most of our analyses we are concerned with formal types. In fact, we must make room for a dual type definition. The discussion of which type definition should be the operational principle for archaeological analysis is an example of the epistemological ambiguities of a discipline with different paradigmatic orientations. One school of thought holds that behaviour can be inferred through formal analysis. The underlying rationale is provided by Clarke’s (1968) definition of material culture as the structured outcome of a system and its behavioural principles. The process of lithic technology varies with the rhythm of the decisions and actions of individuals whose life histories were regulated by general behavioural principles within their populations. Therefore, when formal categories begin to reveal nonrandom properties or structure in the archaeological record, behaviour arrives on the scene. The alternative uses a middle-range approach (Binford 1981) to decode the behaviour at work in the formation of archaeological records; it is behaviour itself that must be made the basis for generalization. I believe that both these positions are legitimate. In both the process of inference is crucial, and that is what Dibble and McPherron show here. They demonstrate that “small flakes” is an analytically useful category, and it is from

795

that quality that their value in the study of Mousterian life ways will emerge. Even if we must acknowledge that, when it comes to the final interpretation of analytical patterns, the logic of our inferential reasoning will often run counter to the way intentions were actually played out tens of thousands of years ago, it is all we have to go by.

Manuel Vaquero Area de Prehistoria, Universitat Rovira i Virgili, Pl. Imperial Tarraco, l, 43005 Tarragona, Spain (manuel. [email protected]). 29 IV 06 Dibble and McPherron’s paper deals with a far-reaching question in the interpretation of Middle Paleolithic lithic assemblages. Small flakes constitute the bulk of most assemblages, but researchers have tended to consider them simply as byproducts of the manufacturing of larger tools. Dibble and McPherron argue that these small items possibly had an important role in the technical behavior of Neandertals and were sometimes intended end products used in an unretouched state. The small-remains question is essential for understanding the cultural formation processes of lithic assemblages, and Dibble and McPherron’s paper should therefore be considered a valuable contribution to the debate on Middle Paleolithic assemblage variability. I fully agree that the deliberate production of small flakes was a fundamental component of the technical activities carried out at Middle Paleolithic sites. In my studies of Spanish Mousterian assemblages, especially those from the Abric Romanı´ (Vaquero et al. 2001), I have found that the recurrent detachment of small flake series was a common knapping activity, clearly suggesting that they were desired end products. In some archaeological levels from the Abric Romanı´, the majority of the core reduction sequences were aimed at obtaining small products. Some of these sequences have been reconstructed through refits, showing the recurrent and exclusive nature of this small-sized production. Nevertheless, some aspects of the paper merit discussion, especially those related to the archaeological criteria for proposing the preferential production of small flakes in Level IVA. Dibble and McPherron do not pay appropriate attention to certain contradictions in their data. They point out that the “Asinipodian” Level IV-A is characterized by regular production of small flakes and that this distinguishes it from the rest of the archaeological levels. It seems, however, that the proportion of small items is not higher in Level IV-A, at least if we take into account the average lengths of flakes reported. This would suggest that the use of knapping methods aimed at producing small flakes does not cause an increase of these artifacts in the lithic assemblage. I think that Dibble and McPherron should explain this apparent paradox. Level IVA shows the highest percentage of small cores, but this does not necessarily mean that the production of small flakes was

796

intensified in this level. Two of the techniques dominant in Level IV-A (Kombewa and truncated-faceted) are characterized by their low productivity; the number of flakes that can be obtained from such methods is always very limited. The evidence from other Middle Paleolithic sites indicates that the small-flake-to-core ratio can be considerably increased using reduction methods such as discoid technology. What should be explained is therefore why such low-productivity methods are used more in Level IV-A than in other levels. A large number of small cores does not imply a stronger emphasis on the production of small flakes. Other assemblages with many small flakes can present this same emphasis, although because of the use of more productive techniques the number of small cores is smaller. According to Dibble and McPherron, the final objective of this production sequence was the use of these small flakes. As they point out, this conclusion seems logical because of intentional nature of the small-flake production, and I agree that there is no obstacle to the use of these artifacts. Usewear data would be essential to proving that the small flakes were effectively used or used more in Level IV-A than in the other levels, but the preliminary microwear data cited do not seem conclusive. In any case, there is no argument that small flakes were produced for specific functions. Given that the form/function relationship has been rejected by several usewear studies, is a size/function relationship being proposed? I think that a conclusive corpus of microwear data would be needed to test such a hypothesis. It should be stressed that, from the point of view of their potential use, small flakes simply offer cutting edges just as large artifacts do. Perhaps the differences between large and small items are related not to function but to other factors such as use length. There is evidence that large and small items are treated differentially from the point of view of mobility and transport (Morin 2004; Morrow 1996; Otte, Bonjean, and Patou-Mathis 2001; Texier et al. 1998). Small flakes tend to correspond to reduction sequences carried out on the site, while a significant proportion of large artifacts are introduced from elsewhere as single items. For instance, in Level N of the Abric Romanı´ all the knapping episodes carried out on the spot were aimed at producing small flakes, whereas all the large artifacts were transported as such from elsewhere. It seems, therefore, that small flakes were produced to meet immediate needs. These factors related to mobility and transport could be a more promising line of research for understanding the size-related differences in Middle Paleolithic assemblages.

Reply First, we thank all of those represented here for their thoughtful comments on our article. We are especially pleased

Current Anthropology Volume 47, Number 5, October 2006

to see the generally high level of support for the results we presented and will reply both to issues raised by individual commentators and to particular themes that are echoed by several among them. Several comments stress the need for additional use-wear studies to confirm or reject the use of small flakes. As we stated, this is something that we tried to do, but so far it has not been successful. However, we are not convinced that use wear itself is sufficiently unambiguous that it can serve to determine whether particular objects were used. It is clear that current use-wear techniques depend on sufficient use of an object to leave clear traces on the edges; if a flake or tool was used only briefly, then no such traces may be visible. This could very well be the case for the small flakes discussed here, and in this case the absence of use-wear traces would be expected and could not be used as an argument against the objects’ themselves having been used. Moreover, the argument presented here concerns the purposeful production of a particular class of objects, not simply whether small flakes, perhaps chosen from among the by-products lying around the occupation surface, were used. The latter would be an entirely different kind of argument and might have to depend exclusively on evidence of their having been used. A final problem, and one emphasized by Tostevin and Churchill, is that it is difficult to identify the flake products themselves. We are lucky in the case of the Asinipodian of Pech de l’Aze´ IV that some of these products (the Levallois and Kombewa flakes) are identifiable and can be subjected to use-wear analysis. Flakes removed from truncated-faceted pieces will be impossible to identify, and certainly we can imagine other techniques of small-flake production that would be equally impossible to distinguish from other small by-products. In cases such as these and what we have tried to do here, the demonstration must be made not on the basis of the actual products but rather on the basis of those elements that reflect their production. This argument is somewhat analogous to one that identifies particular products that were removed from a site by showing that are missing from an otherwise complete refitted ensemble (Marks and Volkman 1983). A number of commentators (e.g., Ashton, Shea, Doronichev and Golovanova, Eren) appear to remain unconvinced that the production of small flakes at Pech IV is not somehow related to the intensity of core reduction and/ or represents other aspects (including size and shape) of the available raw material. Without having a sample of the actual nodules used at the time, this always remains a possibility that is difficult to test beyond what we have already presented in terms of blank-to-core and tool-to-flake ratios (both of which do reflect intensity of production). Eren, however, proposes a set of competing hypotheses, relating to the presence/absence of other, larger cores, that can be tested. In fact, the Asinipodian Level IV-A does have larger cores (see table 4), even though (and as we had already shown in figure 3) this level has the highest proportion of very small ones. Thus, as Eren suggests, the presence of these large cores in

Dibble and McPherron The Missing Mousterian

797

the Asinipodian supports the other results suggesting that intensity of use is not behind the production of small flakes. Several commentators, including Hiscock, Doronichev and Golovanova, Churchill, and Ashton, argue that all of the pieces we examined here—small cores, Kombewa cores, and truncated-faceted pieces—could have served functions other than the production of small flakes and that especially the latter two could reflect the thinning of the pieces for hafting. For at least the high frequency of truncated-faceted pieces in the Zagros Mousterian, this explanation has already been rejected (Dibble 1995a); Dibble and Holdaway 1990), and in fact that industry resembles the Asinipodian at Pech de l’Aze´ IV in many features, including many very small cores. Nonetheless, this represents an alternative hypothesis that could and should be tested. It is certainly possible that the same technology served one purpose at, for instance, Pech de l’Aze´ IV and an entirely different purpose elsewhere. However, on the basis of our experience with the Pech de l’Aze´ IV assemblage as well as several others we have studied from the same perspective, we are skeptical of claims for, for instance, hafting modifications that do not consider the role of the modification technology in the entire assemblage rather than for hand-picked examples. One of the fundamental points that we have tried to make in this paper is that such demonstrations should be based on empirical and perhaps experimental data. In our view, it is simply not enough to interpret Middle Paleolithic artifacts on the basis of our own twenty-first-century views of what they could have been or to build interpretations around a few examples while overlooking the range of variability that may exist in the assemblage as a whole. We would not go as far as to characterize these pieces as microliths as Churchill suggests. First, microliths in more recent industries exhibit retouch that shaped them for hafting, and we do not know whether these small, unretouched flakes were hafted. In fact, our presentation of the research of Niewoehner was intended to show that it is not absolutely necessary that they were. Proposing a much more complex technology based on composite tools involving these small flakes would go well beyond any data that we have in hand. Second, the production of microliths in much later contexts Table 4. Frequency Distribution of Core Sizes in Selected Levels of Pech de l’Aze´ IV Maximum Length (mm) 10–20 20–30 30–40 40–50 50–60 160 Total

II-C

I-B

IV-A

V

N

%

N

%

N

%

N

%

0 7 50 71 43 37 208

0.0 3.4 24.0 34.1 20.7 17.8

1 2 16 63 37 45 164

0.6 1.2 9.8 38.4 22.6 27.4

1 75 323 184 50 31 664

0.2 11.3 48.6 27.7 7.5 4.7

1 2 35 43 34 10 125

0.8 1.6 28.0 34.4 27.2 8.0

presumably represents adaptations that are fundamentally different from those recognized in the Mousterian, and carrying over that term to industries with purposeful production of small flakes—which bear no resemblance to true microliths—would only be more confusing. There appears to be some confusion among the comments regarding whether we think that the small flakes (and the various pieces from which they were derived) represent distinct, formal types. The answer is very simple: we do not think that these pieces should be put forward as new types. What we have tried to do here is to show that in the Middle Paleolithic of Pech de l’Aze´ IV there was the deliberate production of flakes that have otherwise largely been assumed to be by-products of other processes (either core reduction or tool production). It appears that most of the respondents agree with this conclusion. As both Vaquero and Churchill point out, it could be that these flakes were intended to be used for tasks that could not be undertaken with larger flakes, but it is also possible that they represent just one end of a continuum of flake sizes that were universally useful. Deciding among these alternatives is difficult, although the fact that there is such a spike in their production during the Asinipodian suggests that, for whatever reason, their usefulness must have been particularly pronounced during that time. As Bouzouggar suggests, it could be related to the way the site was used at the time, which itself could be related to differences in climate. Even if that is the case, as we stated above, we do not think that it is currently possible to recognize unambiguously all such flakes and classify them as a distinct type. A related issue, one that we discussed and that many commentators mention, is the larger question of types and their relevance to archaeological analysis. Two comments (those of Hiscock and Toma´sˇkova´) take exception to our statement that “it is irrelevant how certain things were made unless we know what things they were trying to make,” though they question it for different reasons. Hiscock says that we should not be concerned with the knappers, thoughts, and we could not agree more. Contra Van Peer, we are not making the claim that small flakes represent an emic type, and we would argue (as we have in numerous publications) that it is almost impossible to demonstrate native classifications; even if we could find them, we could question their relevance for much of what we, as archaeologists, want to learn about past behavior. The Ford-Spaulding debate took place nearly a half-century ago, and most archaeologists who adhere to the so-called reductionist school of lithic analysis stopped worrying about emic categories a long time ago. However, like Van Peer, we are interested in identifying units of analysis that inform us about past behavior and adaptation. Toma´sˇkova´’s complaint reflects exactly the point of view that our statement was trying to address. She appears to express a Europeanist chaine ope´ratoire approach that sees types and typology as somehow bad for archaeological inquiry (see also the comment by Shea). She says, for example, that

798

“following the ‘how’ to its behavioral implications of decision making, divisions of labor, learning processes, or cognitive abilities” makes the “what” irrelevant, citing her own work as showing that the “formal category of burin in [her] study was a stage in tool production or modification as much as an end in itself.” Now, over the past 20 years both of us have written numerous articles on the fact that many of the classic Lower and Middle Paleolithic types of scrapers (Dibble 1995a), bifaces (McPherron 2003), and notched tools (Holdaway, Roth, and McPherron 1996) simply reflect morphological changes brought about by reworking of the same piece, and so we agree completely that one “type” of tool may simply be a stage in a reduction continuum that would encompass several formal types. While we cannot access her cited work on burins (we are writing this reply in the field), we suspect that her conclusions are analogous both to our work and perhaps to others’ (such as Frison 1968 and Cahen, Keeley, and van Noten 1979). If so, then we are in agreement with her on this point. We disagree, however, on whether it is a necessary and logical first step to know what the goal is before one can truly understand how past flintknappers achieved it. First, we are probably all agreed that during the European Middle Paleolithic flintknappers were goal-directed—that they did not strike stone randomly without the goal of producing particular products. Rather, there must have been goals, whether it was to achieve flakes with particular qualities (size, shape) or to produce particular kinds of working edges. Second, the emphasis on process, which lies at the heart of both reductionist and chaıˆne ope´ratoire approaches, is designed to examine the techniques and other factors that led either to end products that reflected the realization of those goals and/or to forms that outlived their usefulness and were discarded. It should be clear that if we misinterpret these goals, then we will likely misinterpret the processes that led to them. This is one of the important points stressed by Holdaway. For example, if Oldowan choppers and chopping tools are indeed tools—the desired goal in and of themselves—then our reconstruction of the processes underlying their manufacture will focus on how the useful qualities of those pieces were formed. If, however, they are simply cores for the manufacture of flakes, then our approach will likely focus on the processes that led to the kinds of flakes produced from them. These are very different things, and clearly knowing the intended products is of fundamental importance to carrying out the analysis. In a similar vein, assuming that a particular object represents a desired end product when, in fact, it represents a point beyond its use life may deceive experimental flint knappers into replicating what is essentially garbage rather than objects that were useful in the past. Likewise, reconstructing the chaıˆne ope´ratoire of the Tayacian of Fonte´chevade before it was shown to be due to natural agencies (Dibble et al. 2006) would have been somewhat less than satisfying. There is, then, a huge difference between trying to identify emic categories—native classifications—and trying to develop

Current Anthropology Volume 47, Number 5, October 2006

units of analysis that are behaviorally meaningful or otherwise useful for archaeological interpretation. Could we refer to such units of analysis as types? Of course, since, as Van Peer points out, there are many different senses of that word. But we have tried to avoid that term simply because of the current polemic, very strongly expressed among those in the chaıˆne ope´ratoire school, that a typological approach is to be avoided. This is, however, a misplaced criticism, and it has been pointed out before (Bar-Yosef and Dibble 1995) that even socalled technology studies still rely on developing typologies, taxonomies, units of analysis, or whatever other term is used. Indeed, the identification of a particular process that is distinct from another one is essentially identifying two “types” of processes. We do believe that the definition of appropriate units of analysis is one of the most fundamental tasks facing Paleolithic archaeologists. But such definitions are only the first step in what is a two-step process. Once defined, the next and equally important step is to uncover what external factors have governed both the expression of those units and variability within them. It is on that basis that true explanations of Middle Paleolithic behavior and adaptation can be put forward and tested. We do agree completely with Holdaway when he asks what else is left unobserved in current Middle Paleolithic systematics because they have not yet been identified as useful units of analysis. It would be a very naı¨ve archaeologist who assumed that we already know and can recognize all of the things that were important to Neandertals; it is far more likely that we recognize only the proverbial tip of the iceberg. —Harold L. Dibble and Shannon P. McPherron

References Cited Ambrose, S. H. 2002. Small things remembered: Origins of early microlithic industries in Sub-Saharan Africa. In Thinking small: Global perspectives on mircolithization, ed. R. G. Elston and S. L. Kuhn, 9–30. Washington, D.C.: American Anthropological Association. [JJS] Amick, D., and R. P. Mauldin. 1989. Comments on Sullivan and Rozen’s “debitage analysis and archaeological interpretation.” American Antiquity 54:166–68. [ST] Anderson-Gerfaud, P. 1990. Aspects of behaviour in the Middle Paleolithic: Functional analysis of stone tools from south-west France. In The emergence of modern humans, ed. Paul Mellars, 389–418. Ithaca: Cornell University Press. [BLH] Andrefsky, W., Jr. 1994. Raw-material availability and the organization of technology. American Antiquity 59:21–34. [GBT] ———, ed. 2001. Lithic debitage: Context, form, and meaning. Salt Lake City: University of Utah Press. [ST]

Dibble and McPherron The Missing Mousterian

Antoine, M. 1938. Notes de pre´histoire marocaine. 14. Un coˆne de re´surgence du Pale´olithique moyen a` Tit Mellil, pre`s de Casablanca. Bulletin de la Socie´te´ Pre´historique du Maroc 12:3–95. [AB] Ashton, N. M., J. Cook, S. G. Lewis, and J. Rose. 1992. High Lodge: Excavations by G. de G. Sieveking 1962–68 and J. Cook 1988. London: British Museum Press. [NA] Bar-Yosef, O., and H. L. Dibble. 1995. Preface. In The definition and interpretation of Levallois technology, ed. H. L. Dibble and O. Bar-Yosef, ix–xiii. Madison: Prehistory Press. [GBT] ———. 2005. Preface. In The definition and interpretation of Levallois variability. Madison: Prehistory Press. Baumler, M. F., and L. B. Davis, 2000. Upon closer examination: Paleoindian behavioral inferences from a Folsom feature lithic assemblage at the Indian Creek occupation site, west-central Montana Rockies. Archaeology in Montana 41(2):17–62. [GBT] ———. 2004. The role of small-sized debitage in aggregate lithic analysis. In Aggregate analysis in chipped stone, ed. C. Hall and M. L. Larson, 45–64. Salt Lake City: University of Utah Press. [GBT] Baxter, J. 2005. The archaeology of childhood: Children, gender, and material culture. Lanham, Md.: AltaMira Press. [JJS] Belfer-Cohen, A., and N. Goren-Inbar. 1994. Cognition and communication in the Levantine Lower Palaeolithic. World Archaeology 26:144–57. [ST] Beyries, S. 1988. Industries lithiques: Traceologie et technologie. 2 vols. British Archaeological Reports International Series 411. [BLH] Binford, L. R. 1981. Bones: Ancient men and modern myths. New York: Academic Press. [PV] Binford, L., and S. Binford. 1966. A preliminary analysis of functional variability in the Mousterian of Levallois facies. American Anthropologist 68:238–95. Bisson, M. S. 2000. Nineteenth-century tools for twenty-firstcentury archaeology? Why the Middle Paleolithic typology of Franc¸ois Bordes must be replaced. Journal of Archaeological Method and Theory 7:1–48. Boe¨da, E. 1986. Approche technologique du concept Levallois et evaluation de son champ d’application. Ph.D. diss., Universite´ de Paris. ———. 1994. Le concept Levallois: Variabilite´ des me´thodes. Monographie du CRA 9. Paris: Centre National de la Recherche Scientifique. [GBT] Boe¨da, E., J. Connan, D. Dessort, S. Muhesen, N. Mercier, H. Valladas, and N. Tisnerat. 1996. Bitumen as a hafting material on Middle Paleolithic artefacts. Science 380: 336–38. [BLH] Bordes, F. 1961a. Typologie du Pale´olithique ancien et moyen. Paris: Centre National de la Recherche Scientifique. ———. 1961b. Mousterian cultures in France. Science 134. ———. 1975. Le gisement de Pech de l’Aze´ IV: Note pre´liminaire. Bulletin de la Socie´te´ Pre´historique Franc¸aise 72: 293–308.

799

Bordes, F., and M. Bourgon. 1951. Le complexe mousterien: Mousterien, Levalloisien, et Tayacien. L’Anthropologie 55: 1–23. Bourguignon, L., 1997. Le Mouste´rien de type Quina: Nouvelle de´finition d’une technique, 2 vols. Ph.D. diss., Universite´ de Paris X-Nanterre. [AT] Bourguignon, L., J.-Ph. Faivre, and A. Turq. 2004. Ramification des chaıˆnes ope´ratoires: Une spe´cificite´ du Mouste´rien? Pale´o no. 15. [AT] Bourguignon, L., and A. Turq. 2003. Une chaıˆne ope´ratoire de de´bitage discoı¨de sur e´clat du Mouste´rien a` denticule´s aquitain: Les exemples de Champs de Bossuet et de CombeGrenal c. 14. In Discoid lithic technology: Advances and implications, ed. Marco Peresani, 131–52. British Archaeological Reports International Series 1120. [AT] Bouzouggar, A. 1997a, Matie`res premie`res, processus de fabrication et de gestion d’outils dans la se´quence ate´rienne de la grotte d’El Mnasra I (grotte des Contrebandiers) a` Te´mara (Maroc). Ph.D. diss., Universite´ Bordeaux I. [AB] ———. 1997b. E´conomie des matie`res premie`res et gestion du de´bitage dans la se´quence ate´rienne de la grotte des Contrebandiers a` Te´mara (Maroc). Pre´histoire et Anthropologie Me´diterrane´ennes 6:35–51. [AB] Bouzouggar, A., J. Kozlowski, and M. Otte. 2002. E´tude des industries lithiques ate´riennes de la grotte El Aliya a` Tanger (Maroc). L’Anthropologie 106:207–48. [AB] Bre´zillon, M. 1968. La de´nomination des objets de pierre taille´e: Mate´riaux pour un vocabulaire des pre´historiens de langue franc¸aise. Gallia Pre´histoire suppl. 7. Cahen, Daniel, Lawrence H. Keeley, and F. Van Noten. 1979. Stone tools, toolkits, and human behavior in prehistory. Current Anthropology 20:661–86. Churchill, S. E. 2001. Hand morphology, manipulation, and tool use in Neanderthals and early modern humans of the Near East. Proceedings of the National Academy of Sciences, U.S.A. 98:2953–55. Clark, J. G. D. 1968. World prehistory: A new outline. 2d ed. Cambridge: Cambridge University Press. [SEC] Clarke, D. L. 1986. Analytical archaeology. London: Methuen. [PV] Coon, C. 1951. Cave explorations in Iran. Philadelphia: University Museum Press. Crew, H. L. 1976. The Mousterian site of Rosh Ein Mor. In Prehistory and paleoenvironments in the Central Negev, Israel, vol. 1, The Avdat/Aqev area, pt. 1, ed. A. E. Marks, 75–112. Dallas: Southern Methodist University Press. Dauvois, M. 1981. De la simultaneite´ des concepts Kombewa et Levallois dans l’Acheule´en de Maghreb et du Sahara nord-occidental. In Pre´histoire africain, ed. C. Roubet, H. Hugot, and G. Souville, 313–21. Paris: E´ditions A.D.P.F. Davidson, I., and W. Noble. 1993. Tools and language in human evolution. In Tools, language, and cognition in human evolution, ed. K. Gibson and T. Ingold, 363–88. Cambridge: Cambridge University Press. Debe´nath, A. 1988. Recent thoughts on the Riss and Early

800

Wu¨rm lithic assemblages of La Chaise de Vouthon (Charente, France). In Upper Pleistocene prehistory of western Eurasia, ed. H. L. Dibble and A. Montet-White, 85–89. Philadelphia: University Museum, University of Pennsylvania. Debe´nath, A., and H. L. Dibble. 1994. The handbook of Paleolithic typology. Vol. 1. The Lower and Middle Paleolithic of Europe. Philadelphia: University Museum, University of Pennsylvania. Delagnes, A., 1992. L’organisation de la production lithique au Pale´olithique moyen: Approche technologique a` partir de l’e´tude des industries de la Chaise-de-Vouthon (Charente). 2 vols. Paris: Universite´ de Paris X–Nanterre. [AT] Dibble, H. L. 1984. The Mousterian industry from Bisitun Cave (Iran). Pale´orient 10:23–34. ———. 1985. Raw material variability in Levallois flake manufacture. Current Anthropology 26:391–93. [AB] ———. 1986. The interpretation of Middle Paleolithic scraper morphology. American Antiquity 52:109–17. [PV] ———. 1988a. Typological aspects of reduction and intensity of utilization of lithic resources in the French Mousterian. In Upper Pleistocene prehistory of western Eurasia, ed. H. L. Dibble and A. Montet-White, 181–87. University Museum, University of Pennsylvania. ———. 1988b. The interpretation of Middle Paleolithic scraper reduction patterns. In L’homme de Ne´andertal, vol. 4, La technique, ed. L. Binford and J. Ph. Rigaud, 49–59. Lie`ge: Universite´ de Lie`ge. [PV] ———. 1989. The implications of stone tool types for the presence of language during the Middle Paleolithic. In The human revolution: Behavioural and biological perspectives on the origins of modern humans, ed. P. Mellars and C. B. Stringer, 415–32. Edinburgh: Edinburgh University Press. ———. 1991. Local raw material exploitation and its effects on Lower and Middle Paleolithic assemblage variability. In Raw material economies among prehistoric hunter-gatherers, ed. Anta Montet-White and Steve Holen, 33–48. University of Kansas Publications in Anthropology 19. [AB] ———. 1995a. Middle Paleolithic scraper reduction: Background, clarification, and review of evidence to date. Journal of Archaeological Method and Theory 2:299–368. ———. 1995b. Biache-Saint-Vaast, Level Ia: A comparison of approaches. In The definition and interpretation of Levallois variability, ed. H. L. Dibble and O. Bar-Yosef, 93–116. Madison: Prehistory Press. ———. 1995c. Raw material availability, intensity of utilization, and Middle Paleolithic assemblage variability. In The Middle Paleolithic site of Combe-Capelle Bas (France), ed. H. Dibble and M. Lenoir, 289–315. Philadelphia: University Museum Press. Dibble, H. L., and O. Bar-Yosef, eds. 1995. The definition and interpretation of Levallois variability. Madison: Prehistory Press. Dibble, H. L., and S. J. Holdaway. 1990. Le Pale´olithique moyen de l’abri sous roche de Warwasi et ses relations avec le Mouste´rien du Levant. L’Anthropologie 94: 619–42.

Current Anthropology Volume 47, Number 5, October 2006

Dibble, H. L., and S. McPherron. 2006. Truncated-faceted pieces: Hafting modification, retouch, or cores. In Tools versus cores: Alternative approaches to stone tool analysis, ed. S. McPherron, 50–71. Newcastle: Cambridge Scholars Press. Dibble, H. L., S. McPherron, P. G. Chase, W. Farrand, and A. Debe´nath. 2006. Anthropology, taphonomy, and the concept of Paleolithic cultures: A study of the Tayacian from Fonte´chevade (Charente, France). PaleoAnthropology 2006:1–21. Dibble, H. L., T. Raczek, and S. McPherron. 2005. Excavator bias at the site of Pech de l’Aze´ IV, France. Journal of Field Archaeology 30:317–28. Dibble, H. L., and N. Rolland. 1992. On assemblage variability in the Middle Paleolithic of Western Europe: History, perspectives, and a new synthesis. In The Middle Paleolithic: Adaptation, behavior, and variability, ed. H. L. Dibble and P. Mellars, 1–20. Philadelphia: University Museum Press. Dunnell, R. C. 1971. Systematics in prehistory. New York: Free Press. [PV] Ensor, H. B., and E. Roemer Jr. 1989. Comments on Sullivan and Rozen’s debitage analysis and archaeological interpretation. American Antiquity 54:175–78. [ST] Fonton, M., V. Lhomme, and M. Christensen. 1991. Un cas de “re´duction” et de transformation d’un outil au Pale´olithique moyen: Un racloir de´jete´ de la grotte du Coustal a` Noailles, Corre`ze. Pale´o no. 3, 43–47. [AT] Frison, G. 1968. A functional analysis of certain chipped stone tools. American Antiquity 33:149–55. Geneste, J-M. 1985. Analyse lithique d’industries mouste´riennes du Pe´rigord: Une approche technologique du comportement des groupes humains au Pale´olithique moyen. Ph.D. diss., Universite´ de Bordeaux 1. Geneste, J-M., and H. Plisson. 1996. Production et utilisation de l’outillage lithique dans le Mouste´rien du Sud-ouest de la France: Les Tares a` Sourzac, valle´e de l’lsle, Dordogne. In Reduction processes (Chaıˆnes ope´ratoires) for the European Mousterian: Proceedings of the International Round Table, Rome, May 26–28. 1995. ed. A. Bietti and S. Grimaldi, 343–68. Quaternaria Nova 6 [AT] Gibson, Kathleen R., and Tim Ingold, eds. 1993. Tools, language, and cognition in human evolution. Cambridge: Cambridge University Press. Gordon, D. 1993. Mousterian tool selection, reduction, and discard at Ghar, Israel. Journal of Field Archaeology 20: 105–218. Goren-Inbar, N. 1988. Tool small to be true? Reevaluation of cores on flakes in Levantine Mousterian assemblages. Lithic Technology 17:37–44. Goren-Inbar, N., and I. Saragusti. 1996. An Acheulian biface assemblage from Gesher Benot Ya’aqov, Israel: Indications of African affinities. Journal of Field Archaeology 23:15–30. [ST] Gowlett, J. A. J. 1984. Mental abilities of early man: A look at some hard evidence. In Hominid evolution and com-

Dibble and McPherron The Missing Mousterian

munity ecology: Prehistoric human adaptation in biological perspective, ed. R. A. Foley, 167–92. London: Academic Press. Hardy, B. L., M. Kay, A. E. Marks, and K. Monigal. 2001. Stone tool function at the Paleolithic sites of Starosele and Buran Kaya III, Crimea: Behavioural implications. Proceedings of the National Academy of Sciences, U.S.A. 98: 10972–77. [BLH] Henri-Martin, G. 1957. La Grotte de Fontechevade. Pt. 1. Historique, fouilles, stratigraphie, arche´ologie. Archives de l’Institut de Pale´ontologie Humaine 28. Hiscock, P. 1996. Mobility and technology in the Kakadu coastal wetlands. Bulletin of the Indo-Pacific Prehistory Association 15:151–57. [PH] ———. 2004. Slippery and Billy: Intention, selection, and equifinality in lithic artefacts. Cambridge Archaeological Journal 14:71–77. [PH] ———. 2006. Blunt and to the point: Changing technological strategies in Holocene Australia. In Archaeology in Oceania: Australia and the Pacific Islands, ed. I. Lilley, 69–95. Oxford: Blackwell. [PH] Hiscock, P., and C. Clarkson, n.d. Retouched notches at Combe Grenal (France) and the reduction hypothesis. American Antiquity. In press. [PH] Holdaway, Simon, Barb Roth, and Shannon P. McPherron. 1996. Notched tool reuse and raw material variability in French Middle and Lower Paleolithic sites. American Antiquity 61:377–87. Holloway, R. L. 1969. Culture: A human domain. Current Anthropology 10:395–412. ———. 1981. Culture, symbols, and human brain evolution. Dialectical Anthropology 5:287–303. Hovers, E. 2006. The many faces of cores-on-flakes: A perspective from the Levantine Mousterian. In Tools versus cores: Alternative approaches to stone tool analysis, ed. S. McPherron, 29–69. Newcastle: Cambridge Scholars Press. Kardulias, N., and R. Yerkes. 1996. Microwear and metric analysis of threshing sledge flints from Greece and Cyprus. Journal of Archaeological Science 23:657–66. [BLH] Kohn, M., and S. Mithen. 1999. Handaxes: Products of sexual selection? Antiquity 73:518–26. Kuhn, S. L. 1992. Blank form and reduction as determinants of Mousterian scraper morphology. American Antiquity 57: 115–28. [WAN] ———. 1994. A formal approach to the design and assembly of mobile toolkits. American Antiquity 59:426–42. [GBT] ———. 1995. Mousterian lithic technology. Princeton: Princeton University Press. [SEC, WAN] Leakey, L. S. B. 1931. The Stone Age cultures of Kenya Colony. Cambridge: Cambridge University Press. McBrearty, S. L. Bishop, T. Plummer, R. Dewar, and N. J. Conard. 1998. Tools underfoot: Human trampling as an agent of lithic artifact edge modification. American Antiquity 63:108–29. McBrearty, S., and A. S. Brooks. 2000. The revolution that

801

wasn’t: A new interpretation of the origin of modern human behavior. Journal of Human Evolution 39:453–563. [SEC] McDonald, M. 1991. Systematic reworking of lithics from earlier cultures in the Early Holocene of Dakhleh Oasis, Egypt. Journal of Field Archaeology 18:269–73. McPherron, S. P. 1994. A reduction model for variability in Acheulean biface morphology. Ph.D. diss., University of Pennsylvania. ———. 1995. A re-examination of the British biface data. Lithics 16:47–63. ———. 2000. Handaxes as a measure of mental capabilities of early hominids. Journal of Archaeological Science 27: 655–63. [GBT] ———. 2003. Typological and technological variability in the bifaces from Tabun Cave, Israel. In From prehistoric bifaces to human behavior: Multiple approaches to the study of bifacial technology, ed. Marie Soressi and Harold Dibble. Philadelphia: University of Pennsylvania Museum Press. McPherron, S. P., and H. L. Dibble. 2000. The lithic assemblages of Pech de l’Aze´ IV (Dordogne, France). Pre´histoire Europe´enne 15:9–43. McPherron, S. P., M. Soressi, and H. L. Dibble. 2001. Deux nouveaux projets de recherche a` Pech de l’Aze´ (Dordogne, France). Pre´histoire du Sud-Ouest 8:11–30. Marks, A. E., and P. Volkman. 1983. Changing core reduction strategies: A technological shift from the Middle to the Upper Paleolithic in the southern Levant. In The Mousterian legacy, ed. E. Trinkaus, 13–34. British Archaeological Reports International Series 164. Meignen, L. 1988. Un exemple de comportement technologique diffe´rentiel selon les matie`res premie`res: Marillac, couches 9 et 10. In L’homme de Neandertal, vol. 4, La technique, ed. L. Binford and J.-P. Rigaud, 71–80. Lie`ge: ERAUL. Mellars, P. 1969. The chronology of Mousterian industries in the Pe´rigord region of south-west France. Proceedings of the Prehistoric Society 35:134–71. ———. 1996. The Neanderthal legacy: An archaeological perspective from Western Europe. Princeton: Princeton University Press. Moncel, M-H. 2003. Tata (Hongrie): Un assemblage microlithique du de´but du Ple´istoce`ne supe´rieur en Europe Centrale. L’Anthropologie 107:117–51. Mora, R., I. De La Torre, and J. M. Moreno. 2004. Middle Palaeolithic mobility and land use in the Southwestern Pyrenees: The example of level 10 in La Roca Dels Bous (Noguera, Catalunya, Northeast Spain). In Settlement dynamics of the Middle Paleolithic and Middle Stone Age, vol. 2, ed. N. J. Conard, 415–35. Tu¨bingen: Kerns Verlag. [AB] Morin, E. 2004. An experimental model for interpreting Levallois patterning and toolkit transport. In Settlement dynamics of the Middle Paleolithic and Middle Stone Age, vol. 2, ed. N. J. Conard, 317–41. Tu¨bingen: Kerns Verlag. [MV] Morrow, T. A. 1996. Bigger is better: Comments on Kuhn’s

802

formal approach to mobile toolkits. American Antiquity 61: 581–90. [MV] Moss, E. 1986. Aspects of site comparison: Debitage samples, technology, and function. World Archaeology 18:116–33. [ST] Newcomer, M. H., and F. Hivernel-Guerre. 1974. Nucle´us sur e´clat: Technologie et utilisation par diffe´rentes cultures pre´historiques. Bulletin de la Socie´te´ Pre´historique Franc¸aise 71: 119–28. [AT] Nielson, A. 1991. Trampling the archaeological record: An experimental study. American Antiquity 56:483–503. Niewoehner, W. A. 2001. Behavioral inferences from the Skhul/Qafzeh early modern human hand remains. Proceedings of the National Academy of Sciences, U.S.A. 98: 2979–84. ———. n.d. Neanderthal hands in their proper perspective. In Neanderthals revisited: New approaches and perspectives, ed. T. Harrison and K. Harvati. New York: Plenum. [WAN] Niewoehner, W. A., A. H. Weaver, and E. Trinkaus. 1997. Neandertal capitate-metacarpal articular morphology. American Journal of Physical Anthropology 103:219–33. Nishiaki, Y. 1985. Truncated-faceted flakes from Levantine Mousterian assemblages. University of Tokyo Bulletin 4: 215–26. Noble, W., and I. Davidson, 1996. Human evolution, language, and mind: A psychological and archaeological inquiry. Cambridge: Cambridge University Press. [PV] Odell, G. H. 2001. Stone tool research at the end of the millennium: Classification, function, and behavior. Journal of Anthropological Research 9:45–100. [JJS] Otte, M., D. Bonjean, and M. Patou-Mathis, 2001. Contractions temporelles au Pale´olithique de Sclayn: L’utilisation de diffe´rents paysages. In Settlement dynamics of the Middle Paleolithic and Middle Stone Age, ed. N. J. Conard, 267–91. Tu¨bingen: Kerns Verlag. [MV] Owen, W. E. 1938. The Kombewa culture, Kenya Colony. Man 38:203–5. Ramenofsky, A. S., and A. Steffen, eds. 1998. Unit issues in archaeology: Measuring time, space, and style. Sall Lake City: University of Utah Press. Roberts, M. B., and S. A. Parfitt, eds. Boxgrove: A Middle Pleistocene hominid site at Eartham Quarry, Boxgrove, West Sussex. London: Heritage. Rolland, N. 1981. The interpretation of Middle Paleolithic variability. Man 16:15–42. Rolland, N., and H. L. Dibble. 1990. A new synthesis of Middle Paleolithic assemblage variability. American Antiquity 55: 480–99. Roth, B., M. Lenoir, and H. Dibble. 1995. Description of the lithic assemblages. In The Middle Paleolithic site of CombeCapelle Bas (France), ed. H. Dibble and M. Lenoir, 41–159. Philadelphia: University of Pennsylvania, University Museum Press. Sackett, J. R. 1981. From de Mortillet to Bordes: A century of French Paleolithic research. In Towards a history of ar-

Current Anthropology Volume 47, Number 5, October 2006

chaeology, ed. G. Daniel, 85–99. London: Thames and Hudson. Sandgathe, D. M. 2004. Alternative interpretation of the Levallois reduction technique. Lithic Technology 29:147–59. [GBT] Schlanger, N. 1991. Le fait technique total. Terrain 16:114–30. [PV] Schroeder, B. 1969. The lithic industries from Jerf Ajla and their bearing on the problem of a Middle to Upper Paleolithic transition. Ph.D. diss., Columbia University. Shea, J. J. 1988. Spear points from the Middle Paleolithic of the Levant. Journal of Field Archaeology 15:441–50. [BLH] Slimak, L., 2004. Les dernie`res expressions du Mouste´rien entre Loire et Rhoˆne. Me´moire de Doctorat de l’Universite´ de Provence. [AT] Solecki, R., and R. Solecki. 1970. A new secondary flaking technique at the Nahr Ibrahim Cave site, Lebanon. Bulletin du Muse´e de Beyrought 23:137–42. Sullivan. A. P., III, and K. C. Rozen. 1985. Debitage analysis and archaeological interpretation. American Antiquity 50: 755–79. [ST, GBT] Taschini, M. 1979. La Grotta Breuil al Monte Circeo: Per un impostazione dello studio del Pontiniano. Origino 4:45–78. [WAN] Texier, P.-J., J-P. Brugal, C. Lemorini, and L. Wilson. 1998. Fonction d’un site du Pale´olithique moyen en marge d’un territoire: L’abri de La Combette (Bonnieux, Vaucluse). In E´conomie pre´historique: Les comportements de subsistance au Pale´olithique, ed. J-P. Brugal, L. Meignen, and M. PatouMathis, 325–48. Sophia Antipolis: Editions APDCA. [MV] Tixier, J. 1958–59. Les industries lithiques d’Aı¨n Fritissa (Maroc oriental). Bulletin d’ Arche´ologie Marocaine 3: 109–244. [AB] Tixier, J., M.-L. Inizan, and H. Roche. 1980. Pre´histoire de la pierre taille´e 1: Terminologie et technologie. Valbonne: Cercle de Recherches et d’E´tudes Pre´historiques. Toma´sˇkova´, S. 2000. The nature of difference: History and lithic use-wear at two Upper Paleolithic sites in Central Europe. British Archaeological Reports International Series/Archaeopress. [ST] ———. 2005. What is a burin? Typology, technology, and interregional comparison. Journal of Archaeological Method and Theory 12:79–115. [ST] Tomka, S. A. 2001. The effect of processing requirements on reduction strategies and tool form: A new perspective. In Lithic debitage: Context, form, meaning, ed. W. Andrefsky Jr., 207–23. Salt Lake City: University of Utah Press. [GBT] Tostevin, G. 2003. Attribute analysis of the lithic technologies of Stra´nska´ ska´la II-III in their regional and inter-regional context. In Stra´nska´ ska´la: Origins of the Upper Paleolithic in the Brno Basin, ed. J. Svoboda and O. Bar-Yosef, 77–118. American School of Prehistoric Research Bulletin 47. [GBT] Trinkaus, E., and I. Villemeur. 1991. Mechanical advantages

Dibble and McPherron The Missing Mousterian

of the Neandertal thumb in flexion: A test of a hypothesis. American Journal of Physical Anthropology 84:249–60. [WAN] Turq, A. 2003. Le Pale´olithique infe´rieur et moyen entre les valle´es de la Dordogne et du Lot. Pale´o, suppl. 2. [AT] Turq, A., and J. G. Marcillaud. 1976. Les racloirs a` amincissement de type “Kostienki” de La Plane, commune de Mazeyrolles (Dordogne). Bulletin de la Socie´te´ Pre´historique Franc¸aise 73:75–79. Van Peer, P., and S. Wurz. 2006. The chaıˆne ope´ratoire concept in Middle Palaeolithic lithic studies: Some problems and solutions. Meetings of the Society for American Archaeology, San Juan, Puerto Rico. http://www.saa.org/meetings/

803

esymposium/197/11.pdf. [ST] Vaquero, M., J. Vallverdu´, J. Rosell, I. Pasto´, and E. Allue´. 2001. Neandertal behavior at the Middle Palaeolithic site of Abric Romanı´, Capellades, Spain. Journal of Field Archaeology 28:93–114. [MV] Villemeur, I. 1994. La main des Neandertaliens: Comparaison avec la main des hommes de type moderne morphologie et me´canique. Paris: Centre National de la Recherche Scientifique. Wynn, T. 1979. The intelligence of later Acheulian hominids. Man 14:371–91. ———. 1985. Piaget, stone tools, and the evolution of human intelligence. World Archaeology 17:32–42.