ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior....

56
Journal of Archaeological Research, Vol. 9, No. 1, 2001 Stone Tool Research at the End of the Millennium: Classification, Function, and Behavior George H. Odell 1 This is the second of the two papers that review the literature of archaeological lithic analysis over the last decade. This paper concentrates on aspects of stone tool research that are not directly related to the production or procurement of the tools themselves. It is divided into classification, functional analyses, behavioral processes, and approaches to the subject currently popular among analysts. As with the previous paper, an attempt has been made to be as comprehensive as is reasonable, though availability of sources has resulted in an emphasis on North American literature. KEY WORDS: lithic analysis, functional analysis, use–wear analysis, residue analysis, technological organization, lithic classification. INTRODUCTION In a previous paper in this journal, I reviewed developments in the last decade within the broad purview of stone tool production, dividing the topic into raw materials and procurement, flake experimentation, technology, and research on specific tool types. In this paper I review recent literature on tool classification, functional analyses, behavioral processes, and popular conceptual approaches as they relate to archaeological lithic analysis. The divisions among these categories are by no means clear-cut, and frequent overlaps exist. Of particular relevance is the fact that, in expounding on several of the subjects considered in the previous paper, I found it most parsimonious for the sake of continuity to discuss functional issues together with technological ones. Therefore, many functional issues that, strictly speaking, should have been discussed in this paper have already been covered in the last one; likewise, items of a technological nature creep into this one. 1 Department of Anthropology, University of Tulsa, Tulsa, Oklahoma 74104. 45 1059-0161/01/0300-0045$19.50/0 C 2001 Plenum Publishing Corporation

description

Stone tool resaarch

Transcript of ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior....

Page 1: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Journal of Archaeological Research, Vol. 9, No. 1, 2001

Stone Tool Research at the End of the Millennium:Classification, Function, and Behavior

George H. Odell1

This is the second of the two papers that review the literature of archaeologicallithic analysis over the last decade. This paper concentrates on aspects of stonetool research that are not directly related to the production or procurement of thetools themselves. It is divided into classification, functional analyses, behavioralprocesses, and approaches to the subject currently popular among analysts. Aswith the previous paper, an attempt has been made to be as comprehensive as isreasonable, though availability of sources has resulted in an emphasis on NorthAmerican literature.

KEY WORDS: lithic analysis, functional analysis, use–wear analysis, residue analysis, technologicalorganization, lithic classification.

INTRODUCTION

In a previous paper in this journal, I reviewed developments in the last decadewithin the broad purview of stone tool production, dividing the topic into rawmaterials and procurement, flake experimentation, technology, and research onspecific tool types. In this paper I review recent literature on tool classification,functional analyses, behavioral processes, and popular conceptual approaches asthey relate to archaeological lithic analysis. The divisions among these categoriesare by no means clear-cut, and frequent overlaps exist. Of particular relevance is thefact that, in expounding on several of the subjects considered in the previous paper,I found it most parsimonious for the sake of continuity to discuss functional issuestogether with technological ones. Therefore, many functional issues that, strictlyspeaking, should have been discussed in this paper have already been covered inthe last one; likewise, items of a technological nature creep into this one.

1Department of Anthropology, University of Tulsa, Tulsa, Oklahoma 74104.

45

1059-0161/01/0300-0045$19.50/0C© 2001 Plenum Publishing Corporation

Page 2: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

46 Odell

CLASSIFICATION

General Classification

I feel compelled to start with that favorite archaeological pastime, artifactclassification, even though relatively little work has been accomplished within thelast decade in developing this field. A major issue that typologists must contin-ually confront is the consistency with which their systems are applied, but thisissue also has received little attention. In perhaps the only relevant recent studyof this type, Whittakeret al. (1998), using Sinagua pottery typologies from theAmerican Southwest, evaluate the consistency with which typologies are learnedand disseminated. They found that learning is usually accomplished by means ofa master who passes the knowledge along through generations, a process that canresult in relatively homogeneous typologies.

Classification is the subject of an important monograph involving a fraternalcollaboration of an archaeologist and a philosopher (Adams and Adams, 1991).The authors’ approach is practical rather than theoretical: a typology should beconstructed for a specific purpose, selection is involved in all typology making,and types should possess the essential properties of identity and meaning. Theseprinciples are applied to pottery from Medieval Nubia, where the senior authorhas worked for many years. This is the most exhaustive treatise on the subject ofclassification to appear for several years.

Other typological systems more directly related to the lithic data base alsohave been advanced recently, one of which was offered within the rubric of alithic manual (Andrefsky, 1998). Although this book is only partly concerned withclassification, the author presents a typological system that is clearly meant to bea workable model for ready employment. Rozoy (1991) also provides a specifictypological construct for the French Epipaleolithic period. Through tight dating andtype definition, the author aimed to ascertain the origin and direction of diffusionof various traits.

The foregoing classification systems are essentially paradigmatic in structure,a quality that is not sufficiently precise for Read and Russell (1996). They preferusing clustering algorithms on measurements of flake tools with obvious wearunder low-power magnification. Although their system is certainly more objectivethan structures currently being used, it is not without problems. For example, thesystem, which was applied to unretouched flakes, purports to show how tool shapeaffects use, but shape is only one of a number of variables that determine how atool is held and utilized. And because it is doubtful that their level of use–weardetermination enabled them to distinguish several common formal/functional cat-egories such as projectiles and tools employing projections (e.g., gravers, burins),their resulting typology assuredly does not do what they think it does. On thepositive side, their quantitative system is capable of teasing out associations thatare not readily visible to intuitive typologists.

Page 3: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 47

Form as a Dynamic Process

Back in the days of our discipline’s naive and innocent youth, the conventionalwisdom was that prehistoric tools were fashioned to follow the artisan’s mentaltemplate, and this form was preserved until discovery of the tool in modern times(Dibble, 1995a; Thomas, 1981, p. 15). However, several lines of evidence suggestthat this has seldom been the case. For instance, ongoing taphonomic evaluation,a field initiated several decades ago, indicates that artifacts may undergo changesof condition or stratigraphic placement between discard and discovery (Rowlettand Robbins, 1982; Shea, 1999; Villa, 1982). This issue has injected some much-needed caution into archaeological interpretations.

Also largely unappreciated is the fact that tool manufacture was a dynamic,not static, part of prehistoric cultural systems. The artifacts that we discover oftenserved as portions of larger implements, and their forms and working edges wereoccasionally sharpened and shaped during their use-lives. These modificationsaltered the final form of tools found in the prehistoric record and may have been ofsufficient magnitude to change their archaeological classification. The notion thatthis process may have the potential to alter archaeological topologies blossomedin the 1980s and continued into the 1990s. Battles are currently being fought overwhether prehistoric tool modification renders our classificatory systems invalid,or whether those changes are really inconsequential and our typologies are usableafter all.

Great Basin Projectile Points

One battleground for this issue has been projectile point types in the AmericanGreat Basin. Early antagonists were Flenniken and Raymond (1986), who arguedthat modification processes were of such magnitude as to alter archaeologicalclassification, and Thomas (1986), who disagreed. In more recent experimentssimulating hunting situations with 92 corner- and side-notched points, Flennikenand Wilke (1989) found that breakage and subsequent reworking led to changesin 32% of the type designations. Unfortunately, these assertions are unverifiableas (1) modern knappers, not prehistoric people, did the reworking and (2) thesame researchers who reworked the points made the type assignments. Additionalarguments were employed by Bettingeret al. (1991), who tested the assumptionthat, if sharpening events were sufficient to cause typological change, then certainspecific point types were likely, upon sharpening, to change into certain othertypes. They depicted “archetypal” types that, upon sharpening, should be madeinto “rejuvenated” forms. Their thesis is that, if Flenniken and Wilke were correct,then archetypal forms discovered on archaeological sites should be larger, on thewhole, than rejuvenated forms. Weighing and measuring 6000 points from severallocalities, they found that the supposed rejuvenated forms were as large as, or

Page 4: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

48 Odell

larger than, the archetypal forms, thus refuting Flenniken and Wilke’s argument(but see the latters’ rebuttal: Wilke and Flenniken, 1991). A study of three GreatBasin projectile assemblages by O’Connell and Inoway (1994) also supports the“short chronology” and refutes Flenniken and Wilke’s model.

Rondeau (1996) recently attempted to resolve the dilemma by conductinga detailed technological examination of Elko corner-notched points from oneGreat Basin site. His results suggest substantial evidence for rejuvenation but,contrary to Flenniken and Wilke’s position, usually of a magnitude insufficientto necessitate changes in type designation. He argues that his data derive fromonly one site and that more research is needed to resolve the issue fully. I suggestthat one element that has been missing from this debate all along is indepen-dent rejuvenation and typing of the same experimental collection by disinterestedparties.

Middle Paleolithic Assemblages

The other major battleground concerns Middle Paleolithic assemblages, par-ticularly scrapers. Although the discussion has not been as rancorous as the oneover Great Basin projectile points, the principles being debated are similar. Themain protagonist in this controversy, Harold Dibble, has argued (like Flenniken)that certain types should not be considered discrete, but as points along a contin-uum of variability, caused by frequent sharpening and other forms of modification.In a recent paper, Dibble (1995a) proposed and tested two models by which typeschange form in this way. Supporting this assertion are studies in which no func-tional consistency was found within a scraper type, and blank forms within thesetypes were variable—though, on a larger scale, he also has indicated that MiddlePaleolithic typological similarities between southern France and the Near East arecaused by similarities in blank form, which influence the ultimate shape of tools(Dibble, 1991). Other studies support the scenario that nonnormative factors suchas intensity of utilization, differential rates of tool reuse, and economizing behaviorare driving forces influencing Middle Paleolithic assemblage variability (Dibble,1995b; Holdawayet al., 1996).

Other scholars, however, also have tested these principles on Middle Pale-olithic assemblages and have obtained different results. For instance, Kuhn (1992a)found no support for the assertion that simple scrapers were transformed intotransverse scrapers at Grotta di Sant’Agostino, Italy. He agrees that tool form isresponsive to blank form, but maintains that the shapes of blanks were most di-rectly affected by techniques of core reduction, not retouching. At the nearby caveof Grotta Breuil, Grimaldi and Lemorini (1993) examined flat-retouched MiddlePaleolithic tools. They concluded that it was not sharpening that determined finaltool form and trajectory of use, but the small dimensions of the nodules with whichthe people started.

Page 5: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 49

Likewise, Gordon (1993) examined a collection excavated several years agoin Israel to test the assertion that certain types, particularly Mousterian points,show continuous variation. He found that, although these points may appear tohave been on the same reduction trajectory as convergent or double scrapers, theyare much too small for the amount of retouch on them. In other words, blankselection for scrapers was completely different from that for points, supportingthe argument that the Mousterian point was conceived as an independent tool typewith a separate reduction trajectory. The choice of blank was more selective forboth points and scrapers than it was for notched pieces.

The use of microliths and microburins to define culture-stratigraphic unitsin the Epi-Paleolithic of the Near East also has been challenged, for many of thesame reasons, by Neeley and Barton (1994). If accurate, their criticisms wouldhave a devastating effect on the culture history of the region, which is based onthese distinctions. However, their rationale does not appear very strong, and theirideas have been hotly contested (Fellner, 1995; Kaufman, 1995).

It appears that for Middle and Epi-Paleolithic assemblages, as with GreatBasin projectile points, modification of a tool for purposes of sharpening andshaping changes the form of that tool and sometimes its classification. The extentof the typological changes, the exact situations in which they occur, and the roleof blank selection in the process are still being worked out.

Style

Researchers of artifact style agree that this is a devilishly difficult qualityto put one’s finger on. Freeman (1992) was interested in applying the concept toa Middle Paleolithic context, but concluded that the Bordesian typology then incommon usage excluded those attributes most relevant in depicting style in a seriesof artifacts. But even if the typological construct were not a problem, the nature ofMiddle Paleolithic assemblages would be, because people simply did not encodemuch stylistic variability on their stone tools during this period. In a study that isas pessimistic as Freeman’s, Barton (1997) defines style both as stochastic varia-tion (like Dunnell) and as a conveyor of information, for example, for purposes ofboundary maintenance (like Wobst). Placing his analysis in an evolutionary frame-work, he tested stochastic processes on small and large populations and concludedthat his theories don’t work—a refreshingly honest appraisal that underscores thedifficulty of recognizing passive style in the archaeological record.

Rick (1996) applies the concept of style to long sequences of projectile pointsfrom two rock shelters in central Peru by breaking the typology into four hierarchi-cal stylistic constructs that measure different degrees of inclusiveness. He was ableto discover strong social discontinuities, although such an analysis did not helphim understand the degree of social differentiation or the scale of the groups beingmonitored. Like Freeman’s experience with the Bordesian typology, the amount of

Page 6: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

50 Odell

social information being encoded in Peruvian stone points may have been minimal.In any case, Janette Deacon’s (Deacon, 1992) study of Bushman arrows suggeststhat ritual and belief systems have major effects on hunting practices (and pre-sumably on hunting equipment), and that it is difficult to predict which type ofartifact will carry style with social overtones. The level at which archaeologistscomprehend the belief systems of most of the people they study is so low that theyhave trouble understanding not only the messages being conveyed, but also thetypes of artifacts conveying them.

FUNCTIONAL ANALYSIS

Use–Wear Analysis: General Considerations

The functional analysis of stone tools has developed in several directions, ofwhich the most avidly researched has been the study of traces of wear from uti-lization, or “use–wear analysis.” From somewhat contentious beginnings duringthe 1970s and 1980s, the field has stabilized into positions that are more mutuallysupportive. For awhile it appeared that analysts would be forever labeled as be-longing to what I referred to earlier as the “low-power” or “high-power” schools(Odell and Odell-Vereecken, 1980).

Analysts now advocate the use ofall available clues for functional interpre-tation, not just one or two favored types of traces (Grace, 1996, p. 217; LeMoine,1997, p. 15; Unger-Hamilton, 1989). Following the 1989 Uppsala Use–WearConference, Grace (1993a, p. 385) was able to state, “methodologically a con-sensus emerged so that different approaches (high and low power) were not seenas competing techniques but alternative strategies dependent on the specific ar-chaeological problem.” Although I missed that consensus at Uppsala, Grace’sevaluation is, on the whole, accurate.

It is becoming more common to read that both high- and low-magnificationmethods were employed in an analysis; after all, they can complement each othernicely. A recent analysis of Italian Mousterian assemblages illustrates this point:“these industries presented problems with the conservation of microtraces (edgerounding, polish, and striations) which resulted in a lack of functional inference—this was partially compensated for by the analysis of macrotraces (microflakescars)” (Grimaldi and Lemorini, 1995, p. 146). It is disheartening to read of failedblind tests like those reported in Fredericksen and Sewell (1991), in which theanalysts reported a use duration of only 5–15 min/tool—a duration that had alreadygenerally been agreed on as too short to produce developed and interpretable weartraces (Bamforth, 1988; Lewenstein, 1993; Moss, 1987). In any case, the criticalatmosphere of the past two decades has led to stronger empirical support for thetechniques promoted than would have been the case had less rigorous testingmethods been applied.

Page 7: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 51

Reviews of recent use–wear literature have been published by Olausson(1990), Shea (1992), Yerkes and Kardulias (1993), Pawlik (1995), and Grace(1996). A glossary of descriptive terms for microscopic fracturing has been of-fered by Prost (1993). And Hurcombe (1992) published a general consideration ofuse–wear on obsidian, a material with properties substantially different from theflints and cherts that typically form the subject of such investigations.

Tests of commonly used analytical variables have also begun to appear. Forexample, Lewenstein (1991) evaluated the hypothesis, proposed by Wilmsen inthe late 1960s, that certain tasks were associated with tools possessing specificedge angles. Using data from the Belizian Maya habitation site of Cerros, sheconcluded that edge angle overlaps activities so drastically that one cannot employedge angle to infer function, even when morphological class is considered. And ina holistic evaluation of commonly employed descriptive variables, van den Driesand van Gijn (1997) quantified 301 experiments to examine possible correlationsbetween tool motion/worked material and edge rounding, fracturing, and polish.Their results demonstrate a large amount of functional overlap among attributes ofboth polish and fracture. On the positive side, they established definite relationshipsbetween certain fracture attributes and both motion and worked material, andbetween certain polish attributes and worked material. This is a solid start inquantifying use–wear variables, a kind of study the field needs very much.

Use–Wear Issues

Genesis of Polish Formation

Clearly, the most compelling issue among high-magnification use–wear prac-titioners these days is the genesis of polish. This debate has been raging for yearsand is still not resolved. Glass polishing is an abrasive process that has been re-searched and practiced for a long time, but it has never been clear whether flintsacquire polished surfaces through utilization the same way glass does, or whetherother processes are involved. Patricia Anderson-Gerfaud (1980) threw a monkey-wrench into the exclusively tribological (abrasive) theory when she noted that,during utilization, plant phytoliths seemed to be trapped on the surface of flint bya substance that she interpreted to be silica gel (see also Del Bene, 1979). Thisinterpretation accelerated the pace of research into the genesis of polish formation.

On the other side of the issue are scholars who declare that polish is exclusivelyan abrasive phenomenon. In recent years Levi Sala (1993, 1996) has emphasizedthe mechanical removal of surface asperities, which slide over the flint and pol-ish it, also frequently causing comet-shaped pitting of the surface. Water is notessential to this process, but promotes polish formation. Experiments by Yamada(1993), concentrating on one tiny location on a piece of progressively utilizedsiliceous shale, have supported this position. He argued that, if the silica gel theory

Page 8: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

52 Odell

is operable, then the distribution of microfeatures at this locale should change aspolish develops; but if the abrasive theory is viable, then these features should begradually smoothed without changing their position. Yamada observed no depo-sition of material, and the surficial pits that he was using as markers retained theirintegrity through 3350 strokes, supporting the tribological theory. Grace (1993a,1996) has agreed with these arguments, stating categorically that the silica geltheory has been proven untenable (Grace, 1996, p. 211).

Just as we thought the issue was resolved, however, up pops somebody withevidence to the contrary. Using a particle accelerator and a scanning electron micro-scope (SEM) connected to an energy-dispersive X-ray spectrometer, Christensen(1998; Christensenet al., 1992) indicated that, during use, tiny bits of workedmaterial accumulate in spaces on the surface of the flint and within the lepispherelattice. Elements in these bits can be analyzed by a spectrometer, providing an ideaof their composition. This scenario was tested on tools known to have been usedon bone, showing deposition of calcium and phosphorous; and on ivory, which isknown to have a large magnesium component, an element that shows up promi-nently on the spectrometer. Not only does this add support to the depositionaltheory, but it offers a direct method for ascertaining materials on which flint toolswere worked—provided, of course, that one can be certain which little bits belongto the erstwhile worked material. In addition, Christensen (1998) implanted a tinypiece of copper on the surface of a tool, which she did not succeed in rubbingaway, suggesting that abrasion was not a major factor in surface wear on tools.

But perhaps the situation is more complex than this. Perhaps, during uti-lization, flint surfaces are affected by both depositional factors and mechanicalabrasion, or perhaps another model is needed to incorporate additional elements.One of these factors is likely to be amorphous silica from either the worked ma-terial (e.g., plants) or from the tool itself. Fullagar (1991), attempting throughexperimentation to correlate amount of silica with kind and amount of polish,demonstrated that even small amounts of amorphous silica may play a significantrole in polishing, and that different polishes develop at different rates.

The complexity of the situation and the possibility that both mechanical andchemical factors may be at work have stimulated models that incorporate both.In one of these, Hurcombe (1997) experimented with abrasives and three chem-ical additives on obsidian tools. The results showed alterations of striations bythe chemicals, especially by hydrochloric acid. She theorized that physical abra-sion weakens the surface, which is then further attacked through chemical ac-tion. Mansur (1997) has postulated the presence of a similar process known asRabinowicz’s molecular theory, which involves mechanical removal of materialand subsequent chemical attack, enhanced by water and abrasives, resulting in athin surface film.

A principal goal in all this research is to correlate discrete polish typeswith specific worked materials. Doubters exist. For example, Rees and his col-leagues investigated the fractal properties of flint, which they defined as “spatial

Page 9: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 53

distributions or patterns which possess self-similarity so that there exists a statisti-cal equivalence between small-scale and large-scale fluctuations in these patterns”(Reeset al., 1991, p. 630). They wanted to ascertain whether or not microwearpolishes are fractal and if different contact substances produce fractally differentpolishes. They found that both polished and unpolished surfaces are fractal andcan be distinguished from one another, but there is no correlation between frac-tal dimensions and specific worked materials. Grace (1993a) arrived at the sameconclusion from another perspective, that is, if polish is an abrasive phenomenon,then logically it should not be able to be associated with any worked material.

On the positive side, Yamada and Sawada (1993) defined a series of polishattributes and found clear correspondence between worked material type and all10 attributes. And researchers at Tohoku University “identified 11 basic types ofpolish on shale that are principally the result of the material worked” (Aoyama,1995, p. 131). Further experimentation by Aoyama (1995) confirmed that thesepolish types can also be applied to chalcedony and agate.

Prehensile Wear

The recognition of prehensile traces has been slow in coming. Analysts suchas Collin and Jardon-Giner (1993) have reported difficulty in interpreting haftingwear on stone tools. Having replicated more than 300 hafted hide scrapers, theywere able to distinguish no definitive hafting wear types, only general trends. Whenthey hafted the scrapers with resin and wax, they observed no hafting traces at all.

Other analysts, however, have reported a considerably greater occurrence ofprehensile damage, perhaps caused by different hafting practices or raw material.Owen and Unrath (1989) conducted a series of blind tests to determine their abilityto discriminate prehensile wear from damage on an active tool edge. They assertedthat prehensile damage was produced frequently, it was sometimes mistaken forsoft material wear (which, in fact, it is), and manual prehension sometimes pro-duces traces that look like butchery wear.

If this kind of wear can be detected accurately, it can be a very useful at-tribute. For example, Odell (1994b) monitored prehension through 7500 years ofprehistory in the North American Midwest and found that hafting traces increasedthrough time, whereas damage from manual prehension decreased. Increased haft-ing practices appear to be related to changing mobility strategies and increasedsedentism, for which hafting represented an attempt to produce fail-safe equipmentthat would be more effective and less subject to breakdown (also, see Odell, 1998).

Effects of Trampling

The effects of prehistoric trampling, which can potentially complicate func-tional interpretation, have been the focus of recent experimentation. In a series of

Page 10: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

54 Odell

blind tests (Shea and Klenck, 1993), four sets of tools, utilized for 20 min apiece,were either left unaltered (1 set) or trampled for varying lengths of time (3 sets).The analyst depicted use parameters for the untrampled sample quite accurately,but correctly interpreted only 40% of utilized edges in trampled sets, having thegreatest problems with implements employed on soft materials. These experimentsestablished that trampling damage can be incorrectly interpreted as use–wear.

McBreartyet al.(1998) tested whether the effects of trampling were negligi-ble or whether they mimic retouch. The experimenters found that many of the toolssustained substantial edge damage, several of them qualifying as pseudotools suchas notched pieces and denticulates. Thus trampling damage is a potentially con-founding influence, not only for functional but even for typological interpretations.

Application of Use–Wear Analysis

The field of use–wear analysis is still new enough that unique or unknownphenomena are frequently reported. In one such study of tools from EgyptianNubia, Becker and Wendorf (1993) detected a type of polish, probably relatedto a soft substance, that they were not able to replicate. In a study with a moresatisfying (though not necessarily more correct) outcome, Gassin (1993) noticedbrightly polished flint tools from a Neolithic Chasseen site in France. Replicativeexperimentation established a visual correlation between this wear and wear fromcutting and shaping the surfaces of clay pots.

Functional analyses have recently been applied to a large variety of situa-tions and chronological periods in the Old World. For instance, Schick and Toth(1993) have conducted an extensive series of experiments to replicate the rangeof tasks in which implements of Olduwan and Acheulean people may have beenengaged. Concentrating on Mousterian tools, Kazaryan (1993) determined that ob-sidian flakes and convergent scrapers from two sites in Armenia were specializedbutchery tools. The extreme wear on their edges suggested economizing behavior.And at Mousterian Grotta Breuil in Central Italy, Grimaldi and Lemorini (1995)encountered a lithic industry whose non-Levallois characteristics were probablycaused by the small cobble flint that these people used for their tools. Functionallyvaried, the Grotta Breuil industry constitutes an attempt to reconcile the pursuit oflarge preforms with the optimal exploitation of large cobbles. Applications of thesetechniques to Holocene assemblages include an analysis of a Neolithic Michels-berg village (Schreurs, 1992) and a functional comparison of a Mesolithic huntingcamp with a Neolithic village (Pawlik, 1995).

Several New World applications of use–wear analysis have emphasized theworking of wood. At the Mayan settlement of Cerros in Belize, Lewenstein (1993)found that woodworking with a wide variety of tool forms—some hafted, somenot—was the most common activity at the site. Similarly, a study of key-shapedunifaces excavated from the Interior Plateau of British Columbia established that

Page 11: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 55

their principal task was woodworking (Rousseau, 1992). Other studies show a dom-inance of woodworking among a multiplicity of other activities. Hudler’s (Hudler,1997) analysis of Clear Fork gouges established that their principal function wasfor working wood but that these tools were used for other tasks as well.

Other applications demonstrate that woodworking was not always dominant.Storck (1997) found that woodworking was only one of several tasks conductedwith gravers and beaked scrapers at the Fisher site in Ontario. In another study, acache of flakes and blades found in north Texas was very uniform in both morphol-ogy and function. Most of these preforms were made into endscrapers and showedhide-scraping wear, a tool kit that probably served as insurance gear for occasionallogistical forays (Ballenger, 1996). Examining tools from the area around ModocRockshelter in Illinois, Ahler (1998) established major differences between Earlyand Middle Archaic assemblages on the basis of functional diversity. On the otherhand, Bradbury (1998) found very little use at all on an assemblage from southernKentucky, supporting his interpretation of the place as a location in which the onlydiscernible activity was biface reduction.

Recent Advances in Use–Wear Analysis

Advances in this field have been made on several fronts. High-magnificationanalyses involving metallurgical microscopes have benefited from the develop-ment of Nomarski optics (Kay, 1996), and difficulties of artifact size (i.e., observ-ing artifacts too big to fit on the stage or under the lens of the microscope) appearto be resolvable. Bienenfeld (1995) has shown that, despite problems of bubbling,possible melting with heat, and being time-consuming, epoxy casts replicate polishwell enough for use in microwear analysis, even using the scanning electron mi-croscope. On the sticky issue of artifact preparation, Coffey (1994) was unable toduplicate the degree of chemical erosion from alkali solutions reported by Plissonand Mauger (1988). He concluded that the previous researchers must have usedheated solutions.

Advances in low-magnification analysis have been slower in coming, be-cause fewer people have been working on improving the technique. An exceptionis Tomenchuk (1997), who has developed a parametric use–wear method usingengineering principles of fracture mechanics on edge scarring. He recently appliedthese principles to two artifact concentrations at the early Paleoindian Fisher sitein southern Ontario, concluding that both areas contained the same suite of pre-historic activities. He also found that beaked, single-spurred, and double-spurredscrapers all had a multiplicity of uses.

Two other research programs have the potential to advance the discipline. Oneinvolves topographic measuring techniques, for which Kimballet al. (1995) haveemployed the atomic force microscope. This equipment is capable of producingquantitative measures of surface attributes such as polish and three-dimensionaldigital mapping of these features. Andersonet al.(1998) also reported research with

Page 12: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

56 Odell

topographic analysis of flint surfaces to derive quantitative measures of bearingarea, average valley, and so forth. At this point, the authors can only report thatbearing area perimeter increases with use and that tool raw material influences theresults.

The other innovation is the development of “expert systems” of analysis thatcomputerize a large range of use–wear attributes derived from low- and high-magnification microscopes. Although Grace (1993b) has touted its accuracy andspeed (he even calls it the “FAST” system), it nevertheless remains quite complex,considerably slower than some alternatives, and applicable only to fine-grainedflint. Yet it constitutes a real advance for certain kinds of applications and illustratesthe rapid changes that this field has undergone during the past decade. A similarcomputerized descriptive system for functional analysis has been proposed byLohse (1996).

Residue Analysis

In residue analysis, the researcher isolates substances that adhere to the surfaceof a stone tool. The preferred isolate is, of course, a substance that was intimatelyassociated with that tool during its use-life. Particularly informative are parts ofthe material that were contacted by the tool such as a starch grain, rodent hair,. . .

Uh, rodent hair? Does this indicate that our prehistoric forebears were fileting ratswith their everready rat knives? Herein lies one of the difficulties with residueanalysis: substances adhering to the surface of a tool are usually assumed to beassociated with that tool’s use, a logical leap that is not without danger. Many otherproblems with residue analysis also exist. Because there is abundant evidence thatprehistoric people processed both plants and animals with stone tools, I considereach separately.

Blood Residues: Positive Results

When an animal is butchered, some of the blood from that animal may stay onthe butcher knife. Tom Loy first discovered this eventuality in the mid-1980s andhas been the principal advocate of blood residue analysis ever since. Borrowingfrom medical labs, he has typically employed several techniques for detectingblood.

Loy recently discovered several bifaces eroding out of a roadbed in BritishColumbia. On these tools he detected bison hair, though bison have not inhabitedthis region for many years. His protein analysis consisted of a Hemastix test todetect haemoglobin (Hb) and myoglobin (Mb), an immunological dot–blot screen-ing test for mammalian immunoglobin type G (IgG), and isoelectric focusing toidentify haemoglobin and serum albumin (SA). He employed four methods to de-termine species of origin: radioimmunoassay, isoelectric focusing, haemoglobincrystallization of proteins, and DNA analysis. The presence of bison blood was

Page 13: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 57

confirmed in the Hb crystallization and DNA tests, and AMS dating of a purifiedsample of the blood yielded a date of 2180± 160 bp (Loy, 1993). In similar typesof analysis, Loy and his colleagues have identified human and bovine blood on alarge stone “altar” in the Skull Building at Neolithic Cayonu Tepesi, Turkey (Loyand Wood, 1989), as well as blood from several mammals (including mammoth)on fluted points from eastern Beringia (Loy and Dixon, 1998).

Other researchers also have reported success using techniques for bloodresidue extraction. In an analog to Loy’s bison butchering bifaces, Kooymanet al.(1992) tested implements from the Head-Smashed-In bison kill in Alberta, wherethey claimed that blood residues had remained on the tools for 5600 years. Furthertesting at that site (Newmanet al., 1996) yielded positive reactions to bison andelk antisera for 9 of 31 stone tools and 6 of 16 soil samples tested. And like Loy’sstudy of fluted points, Hylandet al.(1990) tested tools from the Shoop Paleoindiansite in Pennsylvania. Of the 45 artifacts evaluated, 13 tested positive for deer orcaribou (these species are cross-reactive).

Grinding or pounding tools also have been evaluated by this technique. Yoheet al. (1991) established through ethnographic analysis that animal material wascommonly pulverized. Their immunological analysis of milling equipment fromsouthern California returned positive indications that three manos, a mortar, and apestle were employed prehistorically to grind or pulverize small rodents.

Blood Residues: Persistent Problems

Despite the lofty claims of its proponents, however, blood residue analysishas experienced a rising tide of dissatisfaction. A basic lack of agreement existsbetween the results of blood residue analysis and those of many other kinds ofstudies, outlined in stark detail in Fiedel (1996). From “bovine blood” in Ontario,where prehistoric bison bones have never been recovered, to chickens in theArchaic of Oregon, to a lack of trout in Trout Lake—Fiedel recounted a litanyof inconsistencies between the results of blood residue analyses and what wethought we knew about the prehistoric record.

Another troubling disagreement is internal: results of the various tests em-ployed to detect blood residues frequently do not agree with one another. Downsand Lowenstein (1995) illustrated this problem with a series of blind tests of sam-ples of unknown archaeological residues from North, Central, and South America,and known modern control samples. Although the participants reached total agree-ment on the controls, there was “almost total disagreement and lack of compara-bility among the three techniques in the few specific identifications reported on thearchaeological specimens” (Downs and Lowenstein, 1995, p. 14). In a final phaseof the blind tests, residues on six Clovis blades were analyzed. Again, agreementwas reached on the control samples, but absolutely no agreement was achievedwith the archaeological material. These results are quite frustrating, because no

Page 14: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

58 Odell

one will ever know which, if any, of these assays might have been correct. Wallisand O’Connor (1998) experienced similar inconsistencies between Hemastix andimmunological techniques when analyzing projectile points from two rock sheltersin northwest Australia. These points had retained very little blood but a lot of plantresidues, a finding that does not jibe with ethnographic accounts of aboriginal useof these kinds of points in the region.

Also frustrating are cases of disagreement between the results of blood residueand use–wear analysis, where both have been attempted on the same assemblage.A case in point is a study of 100 tools from four sites in the Piedmont regionof northern Virginia (Petragliaet al., 1996). Twenty tested positive for bloodresidues, and 16 had microwear traces. The authors failed to compare these results,but an inspection of their Table 4 indicates that only four tools tested positivewith both techniques, a rather low rate of correspondence. Utilization traces werepoorly developed on these tools, a probable result of expedient use and chemicaldegradation of polish. Just as significantly, “The fact that 80 artifacts did not testpositively for immunological results also tends to support evidence for degradationprocesses” (Petragliaet al., 1996, p. 134).

Thus researchers have discovered a profound difference in accuracy in de-tecting blood residues between modern lab samples and ancient tools, and thatdifference has something to do with differential preservation of blood over time.Experiments have been devised to investigate this situation. In one series, Gurfinkeland Franklin (1988) buried glass slides with and without blood in soil for varyinglengths of time. They detected degradation in some of the samples and concludedthat the haem portion of blood was more stable over time than the protein portion.Unfortunately, the haem portion itself is not sufficient to positively identify thepresence of blood; protein must be preserved for immunological reactions to oc-cur, thereby enabling specific identification (Fiedel, 1996, p. 145; Kooymanet al.,1992, p. 265).

In a similar series of experiments, Cattaneoet al. (1993) buried 12 cachesof bloodstained and unstained flint tools, along with human and animal bone,in a pit in a garden in Sheffield, England. Some of the tools were utilized onmeat and bone, but only for 2 min apiece; additional tools were stained withblood and retained in the laboratory. Using a sensitive process called enzyme-linked immunosorbent assay (ELISA), the researchers again obtained wonderfulsuccess with the laboratory pieces, but more negative than positive reactions forthe buried artifacts. The positive identification of albumin on a scraper buriedfor 1 year but negative reactions on other similarly buried artifacts illustratesthe erratic survival of blood under these conditions. Eiseleet al. (1995) reportedsimilarly disappointing results from both archaeological samples and actualisticexperiments coating artifacts with blood and burying them in different environ-ments for varying lengths of time. And in experiments in which freshly knappedtools were used to cut and scrape one of 10 different species of animals and then

Page 15: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 59

submitted directly to a lab for analysis without burial, Leach and Mauldin (1995;see also Mauldinet al., 1995) reported correct identification of only 37% of thespecimens.

Just as disturbing as the false negatives on pieces that used to have blood onthem are the false positives obtained from soils. Gurfinkel and Franklin (1988)observed that haemoglobin binds chemically to clay and other soil particles, aprocess that is difficult to recover by liquid extraction. However, Newmanet al.(1993) were successful in recovering blood protein from soil adhering to an artifact.This may be a false positive reaction because, using chemstrip tests for bloodresidues, Custeret al. (1988) found that false positive reactions can be caused bymanganese oxide in the soil.

Doubts that haemoglobin molecules survive in their original state suitable forcrystallization have existed for years (Smith and Wilson, 1992). Recently Garling(1998) conducted another series of experiments on modern and ancient samplesfrom the 30,000-year old Cuddie Springs site in Australia. Like researchers dis-cussed previously, she noted inconsistencies in the three techniques used to identifyblood residues. More disturbing yet, she was unable to identify any of the residueson these tools to the species level, as their crystal morphologies were not distin-guishable using criteria that Loy had established for species identification: “thisstudy shows that the similarity of crystal morphology extends across species thatare definitely unrelated” (Garling, 1998, p. 42). She also noted the high potentialfor confusing concentration-specific with species-specific traits, as well as the cur-rent lack of knowledge of the ligand form of haemoglobin preserved in ancientblood residues and bone samples. She concluded that “positive results in either orboth Hemastix and immunoblots were not a reliable indicator of a sample’s abilityto produce consistent or regular haeomoglobin crystals, or in fact of its abilityto produce crystals at all. Indeed, negative results appeared to be a more reliableindicator” (Garling, 1998, p. 43).

If researchers are ambivalent about haemoglobin crystallization as a diagnos-tic process, there is also some concern about the amount of blood that is likelyto be preserved on a tool. In an experiment to investigate this parameter, Turosset al.(1996) butchered a goat with seven tools and concluded that, in general, verylittle blood is preserved on a tool following its utilization. Archaeologists shouldtherefore define the minimum amount of information needed from a particularanalysis, as sufficient residue may not exist to conduct all the assays one mightwish.

It is obvious that blood residue analysis still has massive problems. Grace(1996) observed that the development of this field is comparable to that of use–wear analysis about 10 years ago. Indeed, Fiedel’s (Fiedel, 1996) critical appraisalof this field called for further blind testing with less ambiguous samples, just aswe have seen in use–wear analysis. If these prove unsatisfactory, it will be time tosuspend use of the technique.

Page 16: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

60 Odell

Plant Residues: Starch Grains

Animal blood is not the only surviving type of residue on a stone tool. Variouskinds of plant residues, for example, starch grains, resins, and phytoliths, also havebeen observed on prehistoric implements of considerable antiquity. Research onplant residues has been occurring in pockets all over the world, especially inAustralia.

A preliminary question is whether starch grains recovered from a particularsoil or stratigraphic level are in primary context or were washed down throughoverlying sediments. This issue was broached by Therin (1998), who conducted aseries of experiments to determine the rate of movement of starch grains throughsediments, using sands of different composition. He found that, the greater the sizeof the grains, (1) the less their chance of becoming mobile, (2) the slower theymove, and (3) the less chance of their being trapped once mobile. More importantly,he found that a very low percentage of starch grains of any size becomes mobile. Healso established a relationship with sediment size, that is, the smaller the particlesize of the sediment, the fewer starch grains move through it, though irrigationlevels do have a positive effect on grain movement. His study is encouraging, forit suggests that, under most conditions, starch grains are likely to remain in thelocation at which they were deposited.

In considering starch grains on prehistoric stone tools, it is important to knowwhether the grains observed have stayed with the tool since its utilization or werepart of the sediment that adhered to the tool. This question was recently assessed byFullagaret al.(1998), who were analyzing obsidian implements from Papua NewGuinea. Although use–wear analysis on the tools suggested the processing of tuberssuch as yam or taro, ethnographic studies indicated that people in that region wereinclined to employ shell for these activities, not obsidian. The authors conductedtuber-processing experiments and tested the sediments in which these experimentswere performed. They concluded that starch grain density was substantially greaterin the experimental sediments than in sediments in which the obsidian tools hadbeen excavated, bolstering their argument that the starch grains on the tools camefrom tuber processing. Starch grains from processing tubers such as manioc, yam,and arrowroot, along with maize, legumes, and palm, have also recently beenreported on early Holocene stone implements from four sites in Panama (Pipernoand Holst, 1998).

In another pilot study of a rockshelter in northern Australia, Atchison andFullagar (1998) observed starch grains on each of three stone pounding tools.Tests indicated that the tools contained greater densities of starch grains than thesurrounding sediments did and that large differences existed in the sizes and shapesof grains between the spits and the tools. These tests rendered it unlikely that theimplements were contaminated by contact with sediments and substantiated thecontention that the starch originated with processing activities. A similar argu-ment was employed by Bartonet al. (1998), who discovered significantly greater

Page 17: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 61

quantities of starch grains on utilized tools than on unutilized tools or in the sedi-ments in which the tools had been unearthed.

Plant Residues: Phytoliths and Research Needs

Plant residues other than starch grains on stone tools have also been stud-ied. For instance, Sobolik (1996) inspected 55 chert flakes, scrapers, and othertools from Middle and Late Archaic levels at Hinds Cave, Texas. The most fre-quent residues on these implements were plant phytoliths, followed by plant fiberand animal hair. Plants identified were sotol, yucca, agave, and grass. No correla-tion between type of residue and type of use–wear was observed, the tools beingvery unspecialized and used for a variety of tasks. But in Australia, Kealhoferet al. (1999) conducted a pilot study in which they compared use–wear and phy-tolith studies of the same group of artifacts. The correspondence between the twotechniques was excellent, providing some assurance that they will be regarded ascomplementary and useful approaches to similar problems.

In general, plant residue analysis has experienced neither the problems northe depth of soul-searching that has characterized blood residue analysis. This isprimarily a result of the nature of the residues, that is, plant residues such as phy-toliths and starch grains can be individually observed and identified, whereas bloodresidues require immunological and other indirect tests to detect their presence. Soproblems of agreement between assays has simply not occurred because, in mostcases, only one test—observation—was applied. If plant traces are observable,they are usually identifiable to species or genus level; if they are not observable,that fact may be reported, but is rarely pursued further. To my knowledge, the issueof differential preservation of plant residues has not been studied.

One negative finding about plant residues has been reported, and this studywas conducted not through direct observation, but using crossover immunoelec-trophoresis (CIEP). Leach (1998) employed 19 flakes, manos, and hammers tocut, scrape, grind, and pulverize desert-adapted plants such as yucca, agave, corn,squash, and mesquite. He then boiled the vegetal material in a plain brownwarepot, dried the tools in sunlight, and refrigerated them for 1–2 months before testing.In the words of the author (Leach, 1998, p. 173),

The results of the blind-test were disappointing. In only one case, where mesquite beansground by a mano were identified as mesquite, were correct results obtained from an artifactcoated with an experimental residue that had an antiserum developed for that plant.

Leach attributed the failure of CIEP to produce accurate results to heating inthe ceramic vessel, which may render the residues immunologically undetectable.Alternatively, I suggest that the same problem encountered with blood may be oper-ative with plant residues, that is, differential preservation. Because immunologicaltests such as CIEP have seldom been applied to plants, questions of preservationhave seldom been asked in this way. Researchers directly observing residues such

Page 18: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

62 Odell

as phytoliths or starch grains are inclined to consider what they see, not what ismissing. In other words, they are not inclined to inquire whether Tool B, on whichno such residues were observed, became that way because the tool was not used inthe first place, or because the residues originally present have degraded. If the latteris the case, it implies that all those analyses that have been conducted through directobservational methods may be usable only to show the prehistoric presence of aparticular plant. They cannot be employed to provide a representative measure ofthe activities that were practiced in antiquity, because, without a lot more research,we will never know which residues have degraded and which have not. Therefore,even though it may appear that plant residue analysis is more accurate than bloodresidue analysis, this is a false impression, as insufficient research has been con-ducted so far to be able to assess the issue of differential preservation. Use–wearanalysis may be helpful as a test with which to compare plant residue results.

This problem also applies to the analysis of ancient DNA. Although Hardyet al. (1997) have extracted DNA from stone implements from the FrenchMousterian site of La Quina, they reported that DNA from these ancient sourcesis present only in small fragments. Data from their modern experimental seriessuggested that DNA breakdown occurs rapidly after death. Thus although the ex-traction of ancient DNA from soils or stone tools appears promising, research hasonly begun to uncover the complexity of the situation.

BEHAVIORAL PROCESSES

Evidence from stone tools, both individually and corporately, can providevaluable information concerning the lifeways of ancient people. They can informon industrial production, subsistence, even transport of materials, sometimes overlong distances. But another conceptual plane exists, one that involves behavioralprocesses and articulates with higher levels of societal organization. On this levelthe lithic database has proven quite robust, providing information on processes asdiverse as mobility organization, gender, and cultural complexity.

Technological Organization

Establishing Mobility Parameters

In the late 1970s and early 1980s, Lewis Binford presented his now-famousforager–collector model of hunter–gatherer mobility organization. This model hashad tremendous influence, at least in North America, on archaeological perceptionsof how hunter–gatherer populations managed their seasonal movements. However,the model has been criticized for simplifying foraging decisions into a dichotomy,when most groups probably employed a mixture of the two systems or alternativesolutions not adequately expressed in either (Chatters, 1987; Nelson, 1991). In

Page 19: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 63

addition, Kelly’s (Kelly, 1983) cross-cultural study indicated that the mobilitystrategies of even one group might differ from one year to the next.

But despite its shortcomings, Binford’s model remains a useful frameworkfor structuring archaeological data. Some of its tenets were refined by Shott (1986)who, breaking residential mobility into frequency and magnitude, determined thattechnological diversity is related more closely to the former than to the latter.Several of the studies mentioned here have taken these relationships even further.

The lithic dataset is perceived to be helpful in understanding prehistoric mo-bility organization, but what characteristics of tools can help discriminate betweenvarious forms of mobility? To answer this question, it is useful to consider somearchaeological implications of mobility, that is, kinds of preservable objects thathighly mobile people would be likely to carry around with them. The ultimatequestion to a forager, of course, is how to perform those tasks necessary to survivewith the least amount of transport cost. Ever since Kelly’s (Kelly, 1988) discourseon biface utility, North American archaeologists have tended to think that the ulti-mate solution to this problem was the use of bifaces as cores. This position has beensupported through studies such as the one by Morrow (1997), which establishedthat bifacial faceting on midwestern Paleoindian end scrapers increased with dis-tance from the source of tool raw material. But if this relationship is universallytrue, then why do we not find this strategy in the European Upper Paleolithic, oramong Australian Aborigines?

Proceeding from different assumptions, Kuhn (1994) described a “mobiletoolkit” that would optimize potential usefulness with respect to weight (portabil-ity). Perceiving that there exists a point at which large increments in weight bringminimal gains in utility, Kuhn illustrated this as the highest point of a utility:massratio graphed against multiple of minimum length. The optimal ratio results, onaverage, in quite a small tool, whose length is only 1.5 times its minimum us-able length. According to this scenario, it would be more efficient for highlymobile foragers to carry several smaller tools than an equivalent weight in largertools.

Morrow (1996) has criticized Kuhn’s model as being less efficient than logicalalternatives in certain circumstances (e.g., scrapers only 1.5×minimum usefulnesswould quickly be reduced to an unusable state without the possibility of creatinglarger, more usable tools). Perhaps a judicious assortment of smaller flakes andlarger cores, which themselves could be used for a variety of tasks, would providean optimal solution. In any case, the association of mobility with different techno-logical strategies suggests that more than one solution was possible and that thechoice depended as much on historic trajectory as on ultimate efficiency.

To investigate potential solutions, it is often necessary to approach the problemfrom a different angle, as Kuhn did. Cowan (1999) also took a different tack,arguing not from the vantage of finished tools but from the debris left from themanufacture of those tools. Concentrating on small sites in northwestern New York,he determined from their debitage that the technological strategies employed by

Page 20: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

64 Odell

people living in the Archaic, Early Woodland, and Late Woodland periods weredistinctly different.

Risk and Stress

Once a satisfactory level of tool efficiency has been achieved by a particulargroup of hunter–gatherers, the successful strategy tends to continue until essentialparameters of the system change, rendering the strategy less adaptive. A useful wayto perceive this is through the concept of subsistence risk, defined as the probabilityof resource loss and involving the cost of failure and the availability of alternativesources of food. A comprehensive discussion of this concept has been providedby Bamforth and Bleed (1997) who, on the basis of data accumulated by Oswalt(1976), have shown worldwide patterns in resource availability and probability ofrisk, and have derived a heuristic model of failure costs.

Concepts of risk and stress on a finite resource base have been employed toexplain changes in stone tools. Proceeding from diet-breadth models derived fromoptimal foraging theory, resource stress should be detectable in the archaeologi-cal record through changes in emphasis from preferred to second-line resources.Such a case has been made in the Portuguese Mesolithic, in which a first-lineresource, red deer, was apparently depleted and replaced by a greater variety ofless desirable resources, including fish. Stress was reflected in the lithic data by agreater standardization of weaponry, that is, a switch, in the Late Mesolithic, to auniform, geometric-based microlithic technology manufactured by the microburintechnique. This strategy may have constituted an attempt to produce more reliable,possibly composite, tools of which these microliths formed a part. This trend wasaccompanied by an increase in the number of different tool functions and thegreater utilization of debris, as people were forced to exploit a greater variety ofspecies to survive (Vierra, 1995).

McDonald (1991) applied Binfordian models of embedded procurement andcarrying costs to a series of early Holocene sites at the Dakhleh Oasis, Egypt. Shewas able to distinguish three types of sites based on diversity of tool types, rangeof lithic raw materials, and several other factors. Radiocarbon dates associatedwith a dry phase suggest that aggregation in this case was a response to worseningclimatic conditions in the desert, where dispersed groups spent most of their time.Similar arguments were advanced by Young (1994) for the American Southwest.In this region, she reasoned that, in times of stress, people would revert to moremobile residence strategies, indicated by higher biface:core ratios, frequencies offormal (bifacial) tools, and prepared striking platforms. Her thesis was criticizedby Sullivan (1994), who proposed factors other than mobility to account for thechanges in the lithic data.

The equation of formal (usually bifacial) tools with greater residential mo-bility has pervaded the literature of lithic technological organization. Bousman

Page 21: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 65

(1993) has employed this model to contrast late Paleoindian technological strate-gies involving projectile points. He associated frequently sharpened, maintained,and reused Plainview points with residentially mobile foragers and the more costlybut reliable, parallel-flaked, Angostura points with collector strategies. Amick andCarr (1996) also revealed this relationship among Paleoindian groups of the NorthAmerican Southeast. By the Woodland period, Indians in this region were stockpil-ing lithic raw material and using expedient technologies containing fewer bifacesthan in the Archaic.

Odell (1996b, 1998) also noted the decline in bifacial technologies through-out the prehistoric cultural sequence of the Lower Illinois Valley, a tendency thatcorrelates in a general way with increased sedentism and the development of plantdomestication. Although trends in most prehistoric activities appear to have beenstochastic in nature, use–wear traces attributed to hafting increased throughoutthe Holocene period, whereas those from manual prehension decreased. Thesetrends are also associated with the development of sedentary lifestyles and plantdomestication. Because an increasingly greater proportion of an individual toolwas committed to the haft proper, the average frequency of functional units pertool and of different activities and worked materials per tool declined during thisperiod (Odell, 1994b). The need for implements maintained through sharpeningand reshaping also declined over time, though a slight increase in average uti-lized polar coordinates per tool suggests that people used their implements moreintensely (Odell, 1994a, 1996b).

Distinguishing Subsistence Strategies

Specific subsistence strategies are not well understood for remote periods ofprehistory. From the earliest hominids through the Middle Paleolithic, it is oftenunclear whether folks hunted for their meat or scavenged the kills of other species.Combined research on the lithic and faunal databases of the central Italian coasthas provided some enlightenment on this situation as it pertains to the MiddlePaleolithic. Most of the interpretations of scavenging or hunting have been madeon faunal indicators, for example, scavenging being distinguished on the basisof large proportion of head parts, small proportion of limb bones, and relativelyincomplete faunal assemblages (Stiner and Kuhn, 1992).

Lithic data correlate well with faunal indices at these Mousterian sites. Scav-enging assemblages are characterized by disc-core or radial Levallois techniques,heavy modification of tools, low frequency of large flakes, intensive exploitationof blanks, and frequent nonlocal raw material. Hunting assemblages exhibit moreefficient parallel cores, smaller and narrower flakes, a relatively large quantity offlakes taken off each core, less tool modification, and dominance of local raw ma-terial. It is apparent that scavenging populations moved more frequently, had largerterritories, and maintained more widely transported toolkits. Hunting populations

Page 22: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

66 Odell

provisioned a home base, that is, brought carcasses back home, and were conse-quently less mobile and more efficient in manufacturing their implements (Kuhn,1993, 1995).

Procurement Considerations

Technology is dependent on the procurement of resources appropriate formaking the technology work. If those resources are not readily available, procure-ment becomes a scheduling problem. The severity of the problem is related to thenature and proximity of the resource and methods of extraction. Human groupshave several ways to procure resources, by sending out task forces, long-distancetrading, exchange between neighbors, or embedding procurement in a regularlyscheduled foray such as, for hunter–gatherers, a seasonal round.

In considering mobility and the procurement of exotic raw materials,MacDonald (1999) has supplemented the foregoing considerations by stressingthe importance of mating relations. Having found that individuals living in lowerdensity population aggregates travel farther to locate mates than those in moredensely populated aggregates (MacDonald and Hewlett, 1999), he applied theseprinciples to Folsom groups of the Northern Plains. Low percentages of exoticmaterials on these sites were more supportive of long-distance movement of in-dividuals than of entire social groups, a specification that allowed him to providemean mating distances for Folsom groups.

Shackley (1990) approached the procurement scheduling problem for oc-cupants of Paleoindian and Archaic sites in Arizona through obsidian sourcing,establishing that obsidian in this region was taken locally rather than throughlong-distance trade. Noting a negative correlation between distance from sourceand amount of cortex present on the artifacts, he deduced that the procurement ofobsidian among hunter–gatherer groups was an embedded strategy. Analyzing ob-sidian from more recent Classic Hohokam sites through X-ray diffraction, Mitchelland Shackley (1995) came to a similar conclusion. They found an abundance ofobsidian on sites very close to the obsidian sources and a rapid fall-off rate for sitesfurther away, indicating that obsidian was not traded long distances, but was em-bedded in a larger pattern of mobility. In contrast, J. V. Wright (1994), working inthe St. Lawrence Basin of Canada, found Gaspe silica on Paleoindian sites 800 kmupstream of its source, suggesting either long-distance trade or very high mobility.

Seeman’s (Seeman, 1994) procurement problem was different. Working onthe Nobles Pond Paleoindian camp, he wanted to know whether the materials in11 clusters accumulated at one time or sequentially. Despite the availability oflocal lithic materials, the dominant ones at the site came from two distant quarries.He reasoned that, if disparate groups practiced embedded procurement, then theclusters should be characterized by different raw material spectra, as these groupswould have gone through different areas. But if groups that kept returning to thesite practiced a more systematic strategy, then the clusters should possess similar

Page 23: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 67

spectra of lithic substances. Raw material similarity among clusters at Nobles Pondsupports the second scenario, that is, people across the site were related in someway and possibly contemporaneous.

For most continental regions, acquisition of raw materials from distant sourceshas commonly been attributed to substantial residential mobility. This assump-tion dominated Andrefsky’s (Andrefsky, 1995) interpretations of survey data fromthe Lower Snake River in Washington. Instead of the common assumption thatpeople in this region became more sedentary through time, he found that recentgroups employed relatively large quantities of nonlocal raw material, suggestingthat they were quite residentially mobile. These rarer, nonlocal substances weremore likely to be made into formal tools, whereas the more common, local oneswere more likely to be employed as expedient tools (Andrefsky, 1994). In this waythe availability of different raw materials has an important impact on a people’stechnological organization.

Using similar principles, Feblot-Augustins (1993) compared late MiddlePaleolithic assemblages from Acquitaine, France, with contemporary assemblagesfrom central Europe. In both cases, most raw material procurement was stable andlocal (within 5 km), but central European assemblages contained more material thatoriginated at least 20 km away. Greater long-distance transport probably equateswith greater seasonal mobility in central Europe, with larger subsistence and ex-ploitation territories. In addition, it supports the exercise of planning and foresightamong Middle Paleolithic peoples.

Raw material procurement and organization of technology are mutually code-pendent. Although previous examples have illustrated the role of procurement intechnological organization, the latter plays an equally important role in determiningthe composition of specific raw materials in a tool kit. This message was conveyednicely by Ingbar (1994), who postulated a simple simulation. A hypothetical hu-man group visits each of the three source areas in turn, replenishing a certainpercentage of its tool kit at each area. The amount of material gathered at a sourceand carried away to become part of a future assemblage usually depends largely onthe organizational needs of the group, which is factored into the simulation. Thatis, some kinds of tools will be replenished and others maintained, affecting thecomposition of the assemblage removed from each source area and assuring thatthe percentages of each of the materials used will be different depending on wherein this cycle one samples the assemblage. Even in this simplified representation,it is obvious that the reconstruction of a group’s movements is not easy and is atleast partly dependent on its technological organization.

Curation

Technological organization is difficult to derive from stone tools, but archae-ologists specializing in lithic materials have always thought that they had a leg upon this issue, because some of the pieces in a typical archaeological assemblage

Page 24: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

68 Odell

are usually modified, whereas others are not. This has been such an importantconsideration that our distant archaeological ancestors usually retained only themodified pieces, which they called “tools”; unretouched flakes became “waste,”as though they were unlikely ever to have been utilized. Today we know better, butmost archaeologists, for good reasons, separate their assemblages into categoriessuch as “tools” or “type collections,” and “debris.”

A logical ramification of the modification of stone is that such pieces appearto have been treated differently from those that were not retouched or ground. Thevery fact of their modification suggests that more attention was paid to them, eitherthrough sharpening, shaping into a specific tool form, or altering specific portionsin order to hold the tool better. Thus the idea that some tools were “curated” andothers “expedient” fell on fertile ground, as archaeologists had secretly known thisall along.

The problem was Binford’s (Binford, 1973) use of the term “curation,” whichwas not very specific or exclusive in the first instance and became even less sowith use. This situation has occasioned extensive criticism of archaeologists’ useof the term by Hayden (1975) and more recently by Odell (1996a) and Nash(1996). Odell evaluated Bamforth’s (Bamforth, 1986) five categories of curation,using lithic assemblages from the Illinois Valley. He found some inapplicable andtwo of them contradictory, and concluded that the term should never be employedwithout defining it strictly at the outset. Nash was equally critical. Using excavatedMiddle Paleolithic assemblages and finding that the term means different thingsto different people, Nash (1996, p. 96) stated,

Given that the concept is now embedded in the literature, a standardized lexicon should benegotiated to facilitate some clarity of communication. In the absence of such standardiza-tion, we should drop the term from the archaeological literature altogether.

Given the popularity of the concept, a small but understandable backlash,led by Mike Shott (1996), has occurred. Shott has defined the term as the degreeof use or utility extracted, that is, the difference between the potential utility thetool starts with versus the amount left on discard. He also perceives the term to beapplicable to individual tools rather than entire assemblages.

Several aspects of the curation concept have been applied during the 1990s.For instance, Walthall and Holley (1997) reported a cache of 10 Dalton tools in anochre-stained spot in a bell-shaped pit located in the uplands near the AmericanBottom. Given a lack of evidence for burial at this locale, they suggested that theseimplements had been produced in advance of use and cached there. Because themajority were end scrapers and ethnographically women are more firmly associatedwith scraping activities than men are, the authors speculated that this was a curatedwomen’s cache.

Close (1996) employed the curational concept of transporting implementsfrom one locale to another. Prehistoric occupants of sand ripple sites in the Saharaapparently carried flakes and cores from one ripple to the next in a circuitous

Page 25: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 69

pattern. They also recycled these tools and stockpiled them for the future. Toolrecycling and maintenance also seems to have occurred at the German MiddlePaleolithic Wallertheim D site (Conard and Adler, 1997). Artifacts were not pro-duced on site, but arrived there in already finished form and were sharpened andmaintained there, suggesting that this was a retooling center. Also working inthe Middle Paleolithic, Kuhn (1992b) noted the transportation of retouched toolsand tool blanks for significant distances; however, he differed with Conard andAdler in the meaning of this behavior. Instead of conscious preplanning, it mighthave been only a temporary hedge against unforeseen difficulties, thus reflectingopportunistic patterns of land use.

Cultural Complexity

Lithic artifacts have traditionally been associated with more technologicallyprimitive stages of cultural development. Recently, however, we have startedto understand and appreciate the processes by which stone tools contributed tomovements toward cultural complexity, as well as how they were replaced bymetal. Stafford’s (Stafford, 1999) study of lithic assemblages spanning the lateMesolithic/early Neolithic transition in Scandinavia, for example, demonstratedthat the process there was a gradual transformation of indigenous societies ratherthan sudden migrations of Neolithic people from elsewhere. Also partly on thebasis of lithic analysis, Stafford offered a redefinition of the Neolithic—from do-mesticated plants and animals to the importation of certain status objects such aspolished stone axes, a process that represented a symbolic transformation of northEuropean culture. Accompanying economic and social complexity were differentritual and ceremonial practices than those that had existed among noncomplexsocieties, practices that are discussed later.

Craft Specialization

One of the hallmarks of complex society is the emergence of craft specializa-tion. But as Milliken (1998) has documented, there existed at least partial craft spe-cialization in the Upper Paleolithic, that is, in Aurignacian ivory beads, Solutreanfoliate points, Magdalenian spear throwers, and the products of knappers at placessuch as Etiolles and Corbiac. Given that specialization is a continuous variable,there has been a lot of discussion about what constitutes true craft specializationand how to determine it from the archaeological record.

At least moderate craft specialization appears to have existed at the LateNeolithic/Chalcolithic village of El Malagon, Spain, located close to the La Ventaflint mine (Ramos-Millan, 1997). People at the village specialized in the produc-tion of bladelets, which they appear to have exchanged for other commodities.Investigations of the eight huts that existed in the later period indicated spatial

Page 26: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

70 Odell

differences in lithic production among them, suggesting that the fundamental pro-duction node was the domestic unit. Ramos-Millan concluded that specializationand political control increased through time, reflected in greater quantities of bothlocal and exotic mined stone. Professional craft specialization also appears to haveexisted in the Bulgarian Eneolithic, evidenced in the extremely long flint blades as-sociated with high-status burials in cemeteries such as Varna (Manolakakis, 1996).This scenario differs from one depicted by Rosen (1997a) at the Camel site, anEarly Bronze Age pastoral camp in the Negev highlands. There he found not truecraft specialization in bead production, but cottage industries that produced excessgoods for exchange in a higher-level economic system.

Shell bead production, usually involving stone microdrills, is also a commonspecialization among coastal peoples. A major project in the Channel Islands offthe California coast has established a considerable trade in this commodity by theChumash and their predecessors (Arnold, 1992; Arnold and Munns, 1994). Theexact nature of the bead trade was a little hard to define, but appears to have in-volved peripherally attached specialists whose product distribution was controlledby highly ranked persons on the mainland. Likewise, shell bead manufacture onMotupore Island, Papua New Guinea, constituted an important source of incomethroughout the history of that island. There production was at a village level anddid not necessarily imply distinct social ranking (Allenet al., 1997).

Most New World work on the topic of specialization has been accomplishedamong the Maya, where conclusions have not always been based on the firmestdata sets. As a case in point, Aoyama (1995) compared areas in and around twostructures at Copan, employing high-magnification use–wear techniques basedexclusively on polish indicators for flint, and polish and striations for obsidian. Hisconclusion that specialized production of marine shell ornaments occurred in frontof structure 10L-16 contains an array of problems, including very small samplesizes, activity and worked material totals that do not match, use of a combinedbone/shell/antler category instead of just shell, and actual dominance of meat/hide,not shell, working.

Fortunately, other contenders for lithic craft specialization in the Maya regionare supported by stronger arguments. For instance, Clark and Bryant (1991) ana-lyzed the production of chert projectile points from a Maya site in Chiapas. Fromthe high error rate and large variability within one point style, they concluded thatthis was not a specialized craft at this locale, but a part-time domestic activitypracticed at the household level. Likewise, use–wear work by Lewenstein (1993)at Cerros in Belize uncovered no evidence of specialized production of stone tools.And further north, the testing of three areas at the central Mexican urban center ofXochicalco by Hirth (1995) established that this town may have been a regionalsupplier of finished obsidian tools, but little evidence existed for large-scale craftspecialization as at Teotihuacan.

But such specialization surely existed in the Maya area, as it did in otherregions. A good candidate is the flint workshop site of Colha, where oval bifaces

Page 27: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 71

and tranchet axes were produced for export elsewhere from the Middle Preclassicto the Terminal Classic period (Shafer and Hester, 1991). Another is the Michoacanarea of western Mexico, where Darras (1994) has described a series of obsidianmines, workshops, and associated settlements. In both cases, the volume of debrisat the workshops suggests substantially more production than would have beennecessary for use within the associated village. Likewise, lapidary production atthe eastern edge of Otumba in the Valley of Mexico, described by Otis Charlton(1993), probably represents specialized jewelry production at this locale.

Trade and Exchange

An important element of increasing cultural complexity among later prehis-toric societies was the proliferation of trade and exchange relationships. Long-distance trade can be very difficult to distinguish from down-the-line exchange(White, 1996, p. 203), or from the primary acquisition of a resource or trade itemthrough embedding it in normal hunter–gatherer seasonal rounds (Wright, 1994).With respect to lithic materials, interpretation involves an intimate knowledge ofthe resources in a region. A case in point is Gibson’s focus on Poverty Point ex-change (Gibson, 1994). Here in the Lower Mississippi Valley, a principal sourcefor toolstone was secondarily deposited local gravels, which are recognizable bytheir brown cortexual rind. However, the gravels probably did not provide all of thepredominantly gray varieties of stone that appear in Poverty Point assemblages,and it is difficult to detect any commodity moving northward in exchange for thematerials that were probably entering the area. These problems affect many regionsand will require considerable effort to resolve.

Several examples of trade or exchange in lithic items have been offered duringthe past decade. For example, Bourque (1994) described the ebbs and flows of tradeon the Maritime Peninsula of eastern North America, primarily in exotic stone. Ofparticular interest were participants in trading networks who evolved novel strate-gies for taking advantage of changing cultural climates. Termed “Tarrentines,”these were native traders who adopted European-style commercial practices. Fur-ther south, the Basin of Mexico also possessed independent merchants who wereresponsible for a high volume of trade during the Late Postclassic period. Despitetheir existence under the strong Aztec state, trade developed through market sys-tems and was never subjected to coercive political control (Smith, 1990). Across theocean, Takacs-Biro (1997) has provided a comprehensive analysis of the distri-bution of lithic trade items in the Late Neolithic of Hungary, including northernvarieties of flint that became important in the later part of this period.

Several items of stone moved through these exchange systems: One commonproduct was obsidian, and that too not just in Central America. Obsidian alsoconstituted a key component of exchange systems in the American Northwest,peaking in intensity at different times in the Plateau (Galm, 1994) and in British

Page 28: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

72 Odell

Columbia (Carlson, 1994). Specific artifact forms were also traded extensively.In the northeastern United States, for example, a trading network was establishedduring the Middle Woodland period featuring Jack’s Reef corner-notched points.Strauss (1992) speculated that a prehistoric coastal trade network involved rawjasper, with finished points being transported up major rivers to the interior.

Raw materials also filled this niche in other parts of the world. For instance,neither chert nor obsidian is indigenous to the Reefs/Santa Cruz Islands of Oceania;both were shipped in from hundreds of kilometers away during Lapita (earlypottery) times (Sheppard, 1993, 1996; White, 1996). Despite the distances, stonefor tools was not transported from the nearest sources of these materials, nor didtool users appear to be optimizing material, as discarded cores were not exhausted.Tool users appear to have responded more to social relationships such as giftingthan to strict utilitarian economics. In another example, the green Pachuca obsidianfrom Teotihuacan has been found on Maya sites in ritual areas and was made intospecialized tool forms, suggesting that its overall function was more symbolic orceremonial than economic (Spence, 1996).

The processes by which material was transferred from a producer to a con-sumer society have been elucidated by research at the Maya lithic workshop site ofColha and its client polities. Two standardized tools—the tranchet bit axe and ovalbiface—were produced in vast quantities at Colha, which was situated on largeoutcrops of high-quality chert nodules. These tools were shipped to consumer sitessuch as Cuello, Pulltrouser Swamp, and Cerros, at which archaeologists have foundnumerous sharpening flakes, but few cores, of Colha chert (Dockall, 1994; Dockalland Shafer, 1993; McSwain, 1991). Distribution of Colha chert occurred generallynorthwest of the producer site in the Late Preclassic period. By the Late Classic, itoccurred in all directions from Colha, as chert was shipped freely throughout theriver systems (Santone, 1997).

Igneous and metamorphic rock were also objects of long-distance trade. Thiswas certainly true of Neolithic western Europe, where axes, adzes, and chiselsconstituted important constituents of the tool kit. Working in the western Alpineregion, Ricq-de Bouard and Fedele (1993) found that only certain kinds of sub-stances were employed for these tools and that specific rock sources were tradedfrom east to west across the Alps. Pre-Dynastic grinding stones from at leastone site in Upper Egypt were manufactured from volcanic rocks from perhaps asmuch as 150 km away, attesting to the existence of a widespread material exchangenetwork during this period (Mahmoud and Bard, 1993). And X-ray fluorescenceanalyses of Late Bronze Age-to-Roman period millstones and querns from Cyprusshowed that materials for their manufacture often originated in distant places, suchas the Levant or the Aegean Islands (Williams-Thorpeet al., 1991).

Finally, the persistence of cultural norms in the face of changing exchangerelationships has been documented by Johnson and Hayes (1995), who comparedtwo noncontemporaneous Middle Woodland sites in the northern Yazoo drainage,Mississippi. Snyders-style points and blade techniques were introduced into the

Page 29: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 73

earlier Fant assemblage, the majority of the tools from which were nonlocal anddominated by Illinois cherts such as Cobden and Burlington. These styles andtechniques were continued in the later Oak Grove assemblage, despite the fact thatmost of the artifacts were made of Arkansas novaculite. In other words, despitea change in procurement or trade relationships from Illinois to Arkansas, artifactstyles and techniques persisted.

Sociopolitical Control

A question related to craft specialization involves the kind of regime un-der which these crafts were structured, that is, were production or distributionof lithic materials strictly controlled by a central polity or social class, or didthey remain under the authority of smaller, household- or kinship-based units?In Mesopotamia, Pope and Pollock (1995) compared the Uruk Mound with twoother tells of the Early Dynastic period. All sites received tool stone in the shapeof already preformed cores, from which toolmakers fashioned blades and otheritems. The authors concluded that, at all three sites, production and use were notlimited to a small number of households, but were widespread within the villages.They found little evidence for either centralization or administrative control of theresource.

A few recent studies have uncovered the possibility of sociopolitical controlamong North American prehistoric groups. For instance, Bayman (1995) notedthat obsidian production and consumption in the Marana Hohokam community inthe American Southwest were concentrated on one large platform mound, whereasother mounds on the site contained very little obsidian. This evidence demonstratedcentralized redistribution of obsidian and suggested the existence of an emergingpower elite. Similarly, in the Plum Bayou culture of central Arkansas, Nassaney(1996) detected concentrations of quartz crystal at Coy Mound and in an area eastof the plaza at the Toltec site. Quartz crystal is most likely a prestige item, andthese concentrations may indicate the presence of social ranking.

In no instance, however, have I run across evidence, uncovered within the pastdecade, of strong sociopolitical control over a lithic resource in the New World, apoint exemplified by Olivella shell bead production in California’s Channel Islands(Arnold and Munns, 1994). The authors isolated several production centers onthe islands but none on the mainland, though it is apparent that political powerwas centered on the mainland. The most likely scenario is that mainland elitescontrolled bead distribution, but left production in the hands of loosely attachedspecialists on the islands.

Even at a highly specialized lithic workshop such as Colha in the Mayalowlands, the most likely authority was lineage based rather than elite control-led (Shafer and Hester, 1991). In addition, the lack of chert workshops at largeMaya centers such as Tikal and Yaxha suggests that production of chert tools was

Page 30: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

74 Odell

decentralized (Fedick, 1991). On the other hand, Fedick noted that redistributionfrom major centers did occur, for the greater prevalence of early reduction stagedebitage at households located away from major centers is consistent with closerhouseholds having greater access to already preformed tools.

Johnson (1996) constructed a similar argument in comparing lithic produc-tion among the Mayans at Colha with that of Middle Archaic Benton people ofnortheastern Mississippi. The discovery of several Benton period caches of finelymade bifaces and of discrete biface production loci on Benton sites suggests a highlevel of production, probably for exchange (Sassaman, 1994b). Johnson’s pointwas, if a Benton period site were found in Belize and attributed to the Classicperiod, it would be presented as evidence for craft specialization controlled by anelite, whereas in reality, whatever specialization there was, was probably looselycontrolled and household based. The primary specialization among the Mayans,and the only one controlled by the elite, concerned activities of a ritual nature.Thus obsidian employed for bloodletting was produced for ritual specialists, withsurplus raw material going to the general population along lines of kinship.

Studies cited above demonstrate a general lack of sociopolitical control overlithic resources among prehistoric occupants of both the Old and New Worlds. Thisimpression has been strengthened by Pope (1994), who compared evidence fromtwo very different geographical and historical contexts: the Early Dynastic Urukpeople in Mesopotamia and the Emergent Mississippian Black Warrior Valleyculture of the North American Southeast. In the latter region specialized beadmaking occurred, but the remains of this activity were widely dispersed within sites,suggesting that production was organized by household or kin group. Similarlyat Uruk Mound, chipped stone tool production was not a centralized activity, andneither techniques nor raw materials were restricted once they were imported tothe site.

Material Replacement

Eventually stone tools were replaced by metal in most areas of the world.Conventional wisdom suggests that this replacement occurred rapidly throughoutall tool types, but until this decade, little research had been done on the processesof the replacement. Several studies demonstrate that the process was complex andinvolved specific tool types at different times.

Working in mid-late Holocene Britain, Edmonds (1995) found that metal bat-tle axes and axe hammers made their first appearance in the Early Bronze Age.Stone was still in general use at this time, but knapping control and tool formalityhad declined. This process continued in the Middle Bronze Age, and knappers in-creasingly employed local raw materials and simple core-flake techniques. Metalsbecame increasingly important in the context of trade, as perceptions of the use ofmaterials changed and sociopolitical forces created new demands and desires.

Page 31: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 75

The most definitive work on material replacement in the Near East has beenproduced by Rosen (1996, 1997b), whose interest spans the Chalcolithic, Bronze,and Iron Ages. In this region the process of replacement took 3000 years andproceeded in stages, first involving stone arrowheads, then burins, then axes anddrills, then ad hoc tools, and finally sickle blades. Reasons for replacement by metalwere different in each case. For example, established trade routes were important inpromoting copper axes, whereas the relative cheapness of flint and well-establishedspecialization in this material ensured that stone sickle blades would continue tobe used long after other lithic tool types had disappeared. Clearly, the replacementof stone by metal was not a linear process, but a highly complex one involvingconsiderably more than simple utility.

Similar complexity accompanied replacement in regions outside Europe andthe Near East. Among the Pawnee of the North American Plains, for instance,metal knives quickly replaced stone ones, but stone scrapers and abrading toolscontinued in use longer because they were less likely to have superior metal coun-terparts. As their utilitarian desirability declined, stone tools eventually assumeda sacred character, as in implements for curative bleeding, etc. (Hudson, 1993).Likewise, among Aborigines of Australia’s northwestern Northern Territory, stonewas rapidly replaced for utilitarian items such as chisels, knives, and spear points.However, stone continued to be valued in the form of prestige objects such as largepoints used as gifts and for exchange (Head and Fullagar, 1997).

Symbolism and Ritual

Stone in Myth and Ritual

The use of stone tools for nonutilitarian purposes has generated an impressivebody of literature in recent years, partly because archaeologists are more aware,through postprocessualist and other currents of thought, of the important role thatsocial considerations play in technology (Dobres and Hoffman, 1994). Practices ofAustralian Aborigines have attracted study because, among many of these groups,stone is regarded in mythical terms. That is, myths related to stone and quarrieshave emerged over the years, as artifacts have been embedded with social and reli-gious value. This point was presented effectively in Robert Paton’s (Paton, 1994)study from northern Australia of leilira blades of cobble quartzite, which weremade for trade and whose use was accompanied by a strict code of conduct. Butwhen inspecting habitation sites at varying distances from the quarry, he foundnary a blade, despite evidence of quartzite debris from their manufacture. Withoutknowledge of Aboriginal behavior generated from elsewhere, one would neverknow what had been manufactured at these sites. McBryde (1997) also made thispoint for the area around Lake Eyre in east-central Australia, a region of exten-sive Aboriginal connecting trails (“dreaming tracks”). She employed petrographic

Page 32: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

76 Odell

techniques to link specific tools to grindstone quarries in the basin and was thenable to associate these material distributions to exchange systems portrayed in oraltradition.

Stone was also imbued with mythical characteristics among some NorthAmerican Indian tribes such as the Iroquois, for whom “Flint” was one of thecreator twins and “Stone Giants” figured prominently in another set of legends(Moulton and Abler, 1991). Characteristics of a similar nature can be read intocertain Maya glyphs associated with stone implements such as axes and hatchets.“Cauac” signs signifying stone, when accompanied with certain warrior figures,seem to have commemorated a particular type of military engagement known as“lightning wars” or “axe-wars” (Thompson 1996). Stone was, in these instances,a highly emotionally charged medium carrying symbolic meanings far beyond itsutilitarian purposes.

Further back in North American prehistory, there is evidence that a few widelytraded objects possessed symbolic significance. One of these was the turkey-tailpoint, around which a Late Archaic/Early Woodland trading network was estab-lished in the Great Lakes region. Usually made of Wyandotte or Dongola chertand often found in association with copper beads, this distinctive object may haveconstituted a form of trade regulator, like the shell necklaces and armbands in theKula Ring of the Trobriand Islands. The fact that some of the points were sharp-ened supports their occasional employment as utilitarian objects, and the existenceof caches containing well-matched points suggests that they were made by part-time specialists (Krakker, 1997). Another such item, produced for exchange andceremonial use in the American Southeast, may have been the Early Archaic Kirkcorner-notched point (Sassaman, 1994a).

Objects other than projectile points also probably served a special, symboli-cally imbued purpose among North American cultures. The simultaneous appear-ance of pottery and bannerstones in the Stallings culture of the American Southeastat about 4000 bp appears to have represented a risk abatement strategy among atleast some of the groups at this time. Sassaman (1998) has theorized that marginaluplands peoples may have produced and traded bannerstones to more dominantpottery-making bands in order to cement alliances. Another possible ritual item,traded in the Dalton (late Paleoindian) culture in the central Mississippi Valley, wasthe well-made Sloan biface. Possibly associated with male hunting ritual, it mayalso have been used to strengthen ties with allied peoples (Walthall and Koldehoff,1998).

Trading relations constitute important evidence for postulating social or ritualsignificance for certain artifact forms, but other types of evidence have also beenused. For example, at the early Paleoindian Parkhill site in Ontario, Ellis (1994)discovered seven miniature artifacts—mostly tiny points on channel flakes, butalso one scraper. He invoked their discard location, clustered close together in twoareas of the site, to establish that these could be neither utilitarian implements norchildren’s toys. From this evidence he postulated that they were ideotechnic items

Page 33: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 77

such as parts of medicine bundles, charms, or elements of witchcraft, discarded inareas of intense activity.

Sievert (1994) also chose an unconventional conduit for studying ritual arti-facts at the Spiro Caddoan ceremonial site. After inspecting a subset of Spiro toolsfor breakage, use–wear, and residues, she established that these techniques wereviable and compiled a list of characteristics that were common among ceremonialimplements. She conducted a similar analysis of a ritually specialized tool assem-blage from the sacred cenote at Postclassic Chich´en Itza on the Yucat´an Peninsula(Sievert, 1992), in which she rendered a functional profile of what an assemblageof ceremonial stone tools might look like. Interestingly, she detected considerableimpact damage from projectile use and depicted a specialized ritual signature thatprobably involved bloodletting, sacrifice, and the use of copal.

Tools in the Acquisition of Symbols and Language

Language has frequently been envisioned as a type of symboling system re-quiring cognitive capabilities similar to those necessary for other symboling sys-tems. Scholars have been increasingly interested in investigating the role of stonetool technique in the development of symboling and language. However, ThomasWynn (1991, 1993) has formulated a strong argument that prehistoric tool behavioris learned in substantially different ways than languages. Wynn is a Chomskyanwho argues for linguistic learning through innate cognitive structures—a proce-dure that is radically different from learning stone tool techniques through roteserial memorization in an apprenticeship format. He concludes that the learning ofprehistoric lithic techniques is unlikely to tell us very much about the origin andevolution of grammar. He does postulate, however, that the Acheulean stage ofbiface production (though not the Olduwan) requires a constellation of responsesthat is a cognitive stage above chimp behavior.

Although tool behavior may not be very helpful in studying theorigin of lan-guage, it can possibly be used to inform on human cognitive abilities for symbolingbehavior (and therefore, language acquisition). Several scholars have pointed to theMiddle/Upper Paleolithic transition as such a cognitive node. For instance, Byers(1994, 1999) argued that early hominids possessed nonsymbolic action cues that,through the mediation of tools as “framing devices,” had assumed symbolic contentby the Upper Paleolithic period. Likewise, Davidson and Noble (1993) denigratedthe idea that early hominid tool makers possessed the cognitive abilities to form atemplate of a specific tool form and then carry out a plan necessary to manufactureit, calling this the “finished artifact fallacy.” In their view, “language arises fromthe discoverythatmeaning is conveyed by signs” (1993, p. 382; italics in original),and the earliest indications of such heightened consciousness are the colonizationof Australia and the creation of sculptures and bas reliefs in Europe—both aboutthe time of the Middle/Upper Paleolithic transition.

Page 34: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

78 Odell

A dissenter to this view is Bednarik (1995), who sees the development of sym-boling behavior as gradual. As evidence he points to “concept-mediated marking”on Lower and Middle Paleolithic objects such as mineralized fossils with engravedlines, a large number of beads and pendants in assemblages from “Traditions firmlyrooted in the Middle Paleolithic” (1995, p. 614), and bones with incised lines fromsites such as Bilzingsleben and Stranska Skala. About this controversy, one thingis certain: the evidence for any position is so tenuous that there is unlikely to be aclear winner any time soon.

CONCEPTUAL APPROACHES

The discussion to this point has emphasized the results of archaeologicalendeavor; blueprints used to conceptualize and generate these results have beenmentioned in passing, but not dwelt upon. Common approaches chronicled aboveinclude experimental, ethnoarchaeological, and ecological, but these are not theonly conceptual models employed by lithic analysts. Here I briefly discuss fourconstructs that have helped analysts put their data in perspective. They can becategorized as one approach (gender) and three conceptual models (design theory,cultural technology, and selectionist anthropology).

Gender

To emphasize gender in lithic analysis is similar to emphasizing ecologicalor social influences—these are the influences that the analyst thinks are worthbringing to the fore within the contexts being considered. Stone tool analysishas traditionally been such a male-dominated domain, however, that an emphasison gender is an acknowledgment that serious bias of mindset has warped ourperception of the archaeological record. Because the conventional wisdom of whomade and used stone tools is based on such weak empirical and ethnographicsupport, it is high time that we question these pillars of knowledge.

The dominant male role in making and using stone tools was duly challengedby Gero (1991). Establishing through ethnography that women made and usedstone tools in some societies, she postulated that, in prehistory, women were likelyto have employed implements in household situations within habitation areas,where the stone from which tools were manufactured was predominantly localand flake tools were utilized expediently. She tested these ideas on assemblagesfrom the Huaricoto settlement in highland Peru, a site that changed from beinga ritual center to a residential village. She found that, as the site changed func-tion, biface production and amount of retouch decreased, whereas the quantityof expedient tools and amount of cortex per tool increased. Projectile points andbifaces, over-represented in earlier strata for ceremonial situations in which malesprobably dominated, were swamped numerically by expedient flake tools of localraw materials in later strata for female-dominated household situations.

Page 35: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 79

Sassaman (1992) operated from similar premises, that female use of toolswas likely to dominate at domestic habitations, whereas male tool use would dom-inate in ceremonial activities and for the procurement of certain resources (e.g.,hunting). He noted that changes in stone knapping toward expedient technologiescoincided with greater reliance on pottery, traditionally assumed to be a femaleactivity. Like Gero, Sassaman interpreted the burgeoning quantities of expedienttools through time as an increase in the contribution of women to stone tool pro-curement, reduction, and use at domestic sites. And also like Gero’s, Sassaman’sarguments are difficult to prove definitively, but they are logically constructed andthey alter our perceptions of the meaning of the archaeological record.

My perambulations through the literature appear to have missed more recentgender studies specifically related to stone tools. However, a few studies do mentiongender in passing, among which is the aforementioned discovery of a concentrationof Dalton scrapers in the American Bottom, which Walthall and Holley (1997)speculate was a curated women’s cache.

Design Theory

Design theory is a way of approaching data that “emphasizes various con-straints in solving given problems by technological means” (Haydenet al., 1996,p. 10). That is, tool design is seen as a strategy for overcoming difficulties in theextraction of resources or in other activities; similarities in design suggest simi-larities in adaptive response to these difficulties. Design theory developed withinthe fields of engineering, industry, and architecture (Jones, 1970; Pye, 1964); onlyrecently has it been applied to archaeological situations. Within the field of lithicanalysis, its application was stimulated by Bleed’s (Bleed, 1986) treatise on main-tainable and reliable tools, followed by Nelson’s (Nelson, 1991) reformulation.Bousman (1993) has discussed these concepts as they relate to hunter–gathererconsiderations such as scarcity, risk, repair strategies, and curation.

At first, Bleed’s conceptualization seemed to be a useful way to encapsulatebehaviorally essential qualities of entire assemblages. Then reality set in: (1) alltools are both maintainable and reliable; (2) expressing these qualities preciselyis very difficult; and (3) quantifying them for comparison is even more difficult.As a consequence, these concepts have been neither extensively developed norapplied.

In an interesting ethnographic study, Horsfall (1987) used a design theoreti-cal framework to analyze grinding stones from Mayan communities in Guatemala.The author found that variation in grinding functions exists in traditional contextsand that variation in function is most closely related to variation in raw mate-rial (Horsfall, 1987, p. 369). Applying concepts such as maintainability, relia-bility, flexibility, and versatility to the chipped stone Keatley Creek assemblagein interior British Columbia, Haydenet al. (1996) delineated several technolog-ical strategies that ultimately enabled them to better understand the assemblage

Page 36: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

80 Odell

organization, tool morphology, and behavioral responses of the occupants of thesettlement.

Design theory has also been applied to the British Mesolithic. Eerkens (1998)believed that climatic and resulting environmental differences between the earlierand later parts of this period would have caused different subsistence strategies tohave been pursued. He reasoned that the more open, parkland environment of theEarly Mesolithic would have induced seasonal intercept-based hunting strategies,for which reliable weapons would have been necessary. If this were true, theseweapons, including microlithic inserts, would have been highly standardized. Bythe more closed–forested Late Mesolithic, reindeer and other herd animals wouldhave disappeared, necessitating encounter-based strategies for more solitary an-imals such as deer. Maintainability of weaponry would have been more valuedunder such circumstances, and resulting microlithic inserts would have been lessstandardized. Statistical analyses of measured attributes of microliths from severalMesolithic sites bear out these relationships.

Cultural Technology

“Cultural technology” is the way in which a cultural group articulates itstechnology to other facets of its structure, or how technology “fits in” to its cul-tural persona. In reconstructing this parameter, several archaeologists in the OldWorld, particularly France, have turned to a holistic approach known as thechaineoperatoire. As originally perceived in its homeland, the term connoted “gesture,”translated as “material action.” Almost mystical in its ramifications, this concepthas connections with language, almost a syntax of action (see Graves, 1994). Oper-ationally, these ideas are not likely to get us very far, so a more practical definition(Sellet, 1993, p. 106) might be that it

aims to describe and understand all cultural transformations that a specific raw materialhad to go through. It is a chronological segmentation of the actions and mental processesrequired in the manufacture of an artifact and in its maintenance into the technical systemof a prehistoric group. The initial stage of the chain is raw material procurement, and thefinal stage is the discard of the artifact.

Thus thechaine operatoireseeks to reveal a dynamic system from its inception toits placement in archaeological context.

For some reason, thechaine operatoirehas been equated by many researcherswith “reduction sequence” (e.g., Grimaldi, 1998, p. 19). Certainly, the conceptencompasses lithic reduction sequences but it is by no means limited to this aspectof behavior. De Bie (1998, p. 91) more accurately describes thechaine operatoireas

integrating knapping methods and tooling, but also processes like raw material procurement,use, abandonment, etc. Rather than merely describing the artefacts, the goal is now toreconstruct (partly by reproduction) and to explain the behavioural processes responsiblefor the formation of the lithic record.

Page 37: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 81

American archaeologists also profess these goals, but their efforts are more oftendirected toward specific elements of the puzzle and are less consciously inclusivethan those who actually adhere to thechaine, that is, those who regard it as morethan just figuring out a reduction sequence. In fact, thechaine operatoirecomesclosest to Schiffer’s behavioral archaeology, which considers objects within theirsystemic contexts from procurement to discard (and beyond). Interestingly, de-spite efforts to bring it back, behavioral archaeology is somewhat out of vogue inAmerica these days, eclipsed by research into broad behavioral concepts such asthe organization of mobility and technology.

Geneste and Maury (1997) apply thechaine operatoirein their researchon spear throwers. These authors do not just engage in a few projectile experi-ments, they reconstruct an entire technology as it would have operated in, say,the Solutrean period. In the process, they typically address questions of produc-tion, blank form, hafting, breakage, efficiency, costs, and constraints—in short,all the factors that would have impinged on a Solutrean person making and usingthis kind of technology. This technological immersion frequently yields a depthof understanding more comprehensive than the piecemeal style of experimen-tation conducted by most experimenters. It is this immersion that distinguishesthechaine operatoirefrom other systems of analysis. Unfortunately, many of thescholars who have mouthed their affection for this system have done nothing morethan give us warmed-over lithic reduction sequences, with occasional lip service toprocurement.

Selectionist Models

Theories of transmission of cultural information, many involving some formof Darwinian evolution, are currently circulating like viruses in cyberspace. Theattraction is strong, as the humanities and social sciences lack a universal the-ory of behavior; the formulation of a theory of cultural transmission would havetremendous influence on every aspect of human endeavor. I do not pretend to beconversant with the range of models or the nuances of difference among them,but several have conspicuously employed lithic data in their formulations and aretherefore included in this review.

The endeavor is difficult. The failed attempt by Barton (1997) to depict stylein lithic artifacts by examining stochastic processes in an evolutionary frameworkhas already been noted. Shott (1997) also attempted to test theories of culturaltransmission on projectile points from the Range site in the American Bottom.Only one variable—projectile length—fit the expected pattern, and this result mayhave been caused by sharpening the pieces.

Other researchers have claimed to have demonstrated some sort of culturaltransmission, though their success has been spotty. Following Parry and Kelly’s(Parry and Kelly, 1987) seminal synthesis associating sedentism with expedient

Page 38: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

82 Odell

technologies, Abbottet al. (1996) inspected the change from biface to flake tech-nologies in five regions. All manifested the same directionality, thus renderingthe operation of stochastic processes unlikely. Because evolutionary explanationsmust reside outside the human system (i.e., they must be environmental), theauthors rejected the decrease in residential mobility as a causal factor, that is,sedentism was aproduct, not a cause. So in searching for causes, the authorssettled on maize agriculture as the agent of change. By this scenario, seden-tism and technological change were both linked, by sorting, to selective forcesfavoring maize production. Thus the authors considered maize agriculture “en-vironmental.” But these postulates make little sense. First, how can maize agri-culture be considered “environmental” (i.e., outside the human system), whileprocesses of sedentism remained within that system? Second, where is the con-firming evidence that maize agriculture is related in any way to an expedient flaketechnology? And third, in order to be causal, maize agriculture had to precedethe shift to a flake technology, but in North America, at least, the opposite istrue.

Bettinger and Eerkens (1997, 1999) have provided more believable tests ofNeo-Darwinist precepts. They noted that cultural transmission theory predicts thatsocially transmitted behavior will vary inversely with the complexity of the behav-ior, the complexity of the context in which it occurs, and the number of individualsinvolved in the transmission. Testing this theory on a collection of Great Basinprojectile points, they gauged complexity by measuring metric dispersion throughthe coefficient of variation statistic. Indeed, complex shapes showed less variabil-ity than simple ones, and arrowheads (representative of a more complex deliverysystem) showed less variability than dart points. They also applied the theory tothe spread of bow-and-arrow technology in the Great Basin. Using weight andbasal width measurements, they contrasted learning through close social contact(indirect bias) with learning through copying prevailing social models (guidedvariation). Reasoning that the former would result in more direct copying and lessexperimentation, and therefore more homogeneous traits in manufactured objects,significant differences in homogeneity of attributes of the same projectile typesshould have been caused by differences in cultural transmission. Indeed, the au-thors discovered significantly greater attribute homogeneity in arrowheads fromcentral Nevada than in those from eastern California. Studies such as these lend adegree of hope that the pursuit of cultural transmission theory will open up newavenues of insight in the future.

PERSPECTIVES

A substantial amount of effort has been spent in the areas of functional analysisand behavioral approaches during the past decade. Less has been directed to lithicclassification. Here I summarize the points made earlier.

Page 39: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 83

Classification

The methodological development of classificatory procedures has not been atthe forefront of archaeological endeavor in the past few years. The most exhaustiverecent work is more a philosophicalexegesis and practical guide to clear classifica-tory principles than an innovative breakthrough. A subset of scholars, employingfactor analysis and clustering algorithms, have branched out from common prac-tice with taxonomic methods that are designed to uncover quantitative structure informal data. However, systems such as these have been slow to gain acceptance,as most typologies remain intuitive.

Although issues regarding the classification of archaeological data have notbeen popular, issues involving the prehistoric behavioral phenomena that producedcertain types have been. A principal question here is, do types reflect the achieve-ment of formal templates in the minds of prehistoric tool makers, or are they theend result of a process of use and sharpening throughout the complex use-life of atool? This issue has been debated in two arenas: projectile points in the AmericanGreat Basin and Middle Paleolithic scrapers in Europe and the Levant. In neithercase has a definitive verdict been reached, though in the American case it doesnot appear that sharpening has significantly changed the archaeological typing ofthe points involved. It is probable that both forces influenced the final forms ofprehistoric artifacts and did so on a case-by-case basis.

Ethnically distinctive style has been extremely difficult to characterize amonglithic artifacts, but some headway has been made through classifying projectilepoints within different levels of inclusion. It appears increasingly doubtful thatadvances in stylistic studies will be made using Bordesian typologies.

Use–Wear Analysis

Use–wear analysis has become much less contentious these days (and a lotmore fun) because practitioners have concluded that they must useall avail-able wear traces in their assessments and that different techniques are appro-priate for different kinds of questions. More people are recognizing prehensilewear these days, a perceptual breakthrough that was long in coming. Perhapsthe most eagerly debated issue of the 1990s was the genesis of use-polishes,which has proven to be a complex problem. We have recently heard from advo-cates of an abrasive model, a silica gel model, and a combination of both, andthe end is not in sight. At least one partisan has stated that, if an abrasive ori-gin of use-polish proves to be dominant (which, of course, he maintains with noreservation), then the discrimination of worked material from polish formationshould be difficult or impossible. Such statements, of course, cause some concernamong those who derive their livelihood from peering exclusively at this type ofevidence.

Page 40: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

84 Odell

Other issues have proven easier to resolve. For example, a comparison ofedge angles to use–wear has shown too much overlap for edge angle data to bevery useful. In addition, experiments have shown that damage on stone tools fromtrampling can be confused with both use–wear and retouch. And the quantificationof use–wear variables has progressed somewhat, notably with Tomenchuk’s para-metric method. However, with this technique it still takes a long time to peruse asingle piece and, to my knowledge, nobody but Tomenchuk is practicing it. Somework in quantifying experimental variables has begun in the Netherlands.

Advances in use–wear analysis include the aforementioned development ofparametric methods, employed with low-magnification microscopes. Higher mag-nification analyses have been aided by the improvement of epoxy casts. Thesefeature very high resolution surfaces and, by looking at smaller sections of arti-facts in cast form, they offer the possibility of analyzing larger artifacts that do notfit on the stage of some microscopes. In addition, the atomic force microscope hasconsiderable potential for use–wear analysis through its ability to distinguish andmap out differences in topography. Finally, the development of expert systems ofanalysis, combining low- and high-magnification equipment, has proven to be anaccurate method for analyzing a lithic assemblage.

Residue Analysis

In contrast with use–wear analysis, which bounced back after some difficulttimes in the 1980s, the prognosis for certain types of residue studies is grim. Theproblem with blood residue analysis is not with its ability to distinguish recentlyutilized laboratory specimens, which it seems to do quite well. The problem is witharchaeological specimens, some of which it is capable of distinguishing to at leasta genus level, some of which it is not. This discrepancy is the most frustrating part,because there is currently no way of knowing which kinds of blood residues havesurvived and which have deteriorated. Without this information, we have no ideahow representative of the archaeological sample a particular assay is. And I amsaying all this assuming that there are no false negatives or false positives, thoughwe know that both have been detected by analysts. If these problems cannot beresolved, then support for this type of analysis will, and should, be curtailed.

Superficially, plant residue analysis appears to be in better shape, but with acurious twist. The principal difference between plant and blood residue analysis isthat the former often relies on the identification of specific particles adhering to thetool surface, such as starch grains, resins, and phytoliths. If one can be certain thatdifferential preservation of these particles has not occurred, then the sample can beassumed to be representative, the results can be used to characterize the assemblagefrom which they were taken, and we can go beyond a simple presence–absence levelof interpretation. But to my knowledge, one cannot be certain because research hasnot yet determined the preservability of various starch grains and phytoliths. In the

Page 41: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 85

one instance to which I can point, in which techniques of immunoelectrophoresiswere used in the analysis of plant residues, the technique fared no better thansimilar nondirect techniques did with blood residue analysis.

On a more positive note, research on tuber processing, starch grains, and phy-toliths has developed at an impressive rate, which lends confidence that the prob-lems may be rectified in the near future. In addition, work conducted in Australiaon the movement of starch grains in sediments has demonstrated that most grainsmove little from their original locus of deposition; thus, potential postdepositionalproblems may be minimal.

Technological Organization

Models of technological organization, which have employed concepts of mo-bility organization, maintainability or reliability of tools, curation, expedience, andso forth, have been popular throughout the 1990s. Some of these terms, however,subsume more than one meaning and have often been loosely applied. When usingterms like “curation” and “expediency,” it has been recommended that the authordefine the term tightly before using it.

Testing these models and concepts has frequently been accomplished throughlithic data. For instance, arguments of hunter–gatherer mobility in the southernEuropean Paleolithic have been framed with respect to the size and sharpeningpotential of implements in the lithic tool kit. Later in the Mesolithic, changesin hunting strategy were accompanied by changes in microlith-tipped weaponry.And in North America, decreases through time in bifacial industries correspondwith societal movements toward increased sedentism and reliance on domesticatedplants. Although most functional trends are stochastic because they depict a com-mon suite of activities pursued in all periods, certain ones do correlate with trendstoward sedentism and domestication, and with increased hafting of implementsthroughout the Holocene.

The procurement of tool stone was strongly influenced by raw material avail-ability and the organization of technology. Embedding of procurement in a hunter–gatherer seasonal round has been illustrated by a simulation demonstrating thatthe amount of a particular material in an assemblage may be dependent on wherein that seasonal round the occupation occurred. However, distinguishing whethera material arrived on a site through embedded procurement or trade can be verydifficult.

Cultural Complexity

Lithic studies in the past decade have not been constrained by exclusiveapplication to hunter–gatherer societies; they have also been applied to issues of

Page 42: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

86 Odell

social complexity. One of these issues involves the process by which stone toolswere replaced by metal. This did not happen all at once, nor were all tool typesreplaced at the same time. The process was complex and involved the relativeusefulness of either material for the intended tasks. For example, metal madesubstantially more efficient cutting and chopping implements, so stone knivesand axes were replaced quickly in most contact situations. On the other hand,stone sickle blades were almost as efficient as their metal counterparts and wereconsiderably cheaper, so they remained in the Near Eastern tool kit for a long time.In several societies, stone was replaced for utilitarian implements but continuedas ritual objects (e.g., as sacrificial knives) or as prestige items for exchange.

The role of stone in the emerging process of craft specialization has beeninvestigated through several studies. It is certain that some specialization in stonetool manufacture has been practiced since the Upper Paleolithic, but at what level?Most recent analysts have been loath to attribute true craft specialization for ex-change when localized cottage industries seem more likely, though when largeindustrialized workshops are involved, a higher level of specialization is implied.

Long-distance trade has been detected in several instances though, as notedabove, this can be difficult to distinguish from embedded procurement. A varietyof lithic materials, including flint, obsidian, igneous stone, catlinite, and others,were traded prehistorically in different parts of the world. Both materials andspecific objects—for example, sandstone along the dreaming tracks of Australia,obsidian in Oceania and Mesoamerica, turkey-tail points in midcontinental NorthAmerica—may have been imbued with symbolic significance or involved in giftingrelations. In many parts of the world, stone in various forms took on symbolic ormythic meanings even as it was being replaced by metal objects.

In no recently reported case of lithic craft specialization or trade has anyscholar of prehistoric cultures detected strong sociopolitical control of the resource.Even at a specialized workshop village like Colha in Belize, tool production wasprobably household or lineage based. Emerging elites, whose power is evident inother ways at some of these sites, were more likely to control objects such as ritualitems.

Conceptual Approaches

Lithic analysis has traditionally been a male-dominated domain, so it is nosurprise that stone tools have been almost universally considered to have beenmade and used by prehistoric men. As the gender composition of lithic analystschanges, so will our perceptions of who did what in prehistory. These changeshave already started to appear in lithic analysis. A good case has been made that,at least on later habitation sites, women used, and probably also made, stone tools.In these societies, men’s tools may have been confined to the realms of huntingand ritual.

Page 43: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 87

Several conceptual models have aided the interpretation of lithic items, threeof which have been discussed in this review. Design theory, in considering how thefeatures of a particular tool type might best be employed, is able to demonstratestructural advantages and constraints that may assist in conceptualizing the generalform and purpose of lithic assemblages. Likewise, thechaine operatoire, aimed torecreate the entire system in which a technology existed, provides a useful structurefor organizing research. By immersing oneself in a larger trajectory of procurementthrough discard, the archaeologist can derive a more complete understanding ofhow the parts of that system articulate with one another.

Finally, of the Darwinian models currently being plied, cultural transmissiontheory appears to have more promise than most. At least, in its application tolithic data, its premises have been supported—a far cry from some of the otherevolutionary formulations making the rounds, which, as far as I can tell, haveremained exclusively in the theoretical realm. This theory needs more testing, butthis time we may see a payoff.

Final Comments

Like most other subdisciplines, lithic analysis has developed sporadically, aresult of a combination of communal trends and individual interests. There existsa healthy amount of information flow, and a prevailing skepticism assures that newideas will not be accepted at face value. These are good signs, which render itlikely that the field will continue to change and grow significantly through the nextdecade.

ACKNOWLEDGMENTS

These two review papers have benefited greatly from comments made onearlier drafts by Mike Shott and Mike Collins; and on later ones by Steve Kuhn,John Whittaker, Rick Yerkes, and an unidentified reviewer. I appreciate their as-sistance, as well as the encouragement and editorial assistance of Gary Feinman,Douglas Price, and Linda Nicholas. All of the opinions expressed in these tworeviews, particularly the harebrained ones, remain my responsibility.

REFERENCES CITED

Abbott, A. L., Leonard, R. D., and Jones, G. T. (1996). Explaining the change from biface to flaketechnology. In Maschner, H. (ed.),Darwinian Archaeologies, Plenum, New York, pp. 33–42.

Adams, W. Y., and Adams, E. W. (1991).Archaeological Typology and Practical Reality, CambridgeUniversity Press, Cambridge.

Ahler, S. P. (1998). Early and Middle Archaic settlement systems in the Modoc locality, southwestIllinois. Illinois Archaeology10: 1–109.

Page 44: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

88 Odell

Allen, J., Holdaway, S., and Fullagar, R. (1997). Identifying specialisation, production and exchangein the archaeological record: The case of shell bead manufacture on Motupore Island, Papua.Archaeology in Oceania32: 13–38.

Amick, D. S., and Carr, P. J. (1996). Changing strategies of lithic technological organization. InSassaman, K., and Anderson, D. (eds.),Archaeology of the Mid-Holocene Southeast, UniversityPress of Florida, Gainesville, pp. 41–56.

Anderson-Gerfaud, P. (1980). A testimony of prehistoric tasks: Diagnostic residues on stone toolworking edges.World Archaeology12: 181–194.

Anderson, P., Astruc, L., Vargiolu, R., and Zahouani, H. (1998). Contribution of quantitative analysisof surface states to a multi-method approach for characterising plant-processing traces on flinttools with gloss. InFunctional Analysis of Lithic Artefacts: Current State of the Research, XIIIInternational Congress of Prehistoric and Protohistoric Sciences, Workshops, Tome II, ABACOEdizioni, Forli, Italy, pp. 1151–1160.

Andrefsky, W., Jr. (1994). Raw-material availability and the organization of technology.AmericanAntiquity59: 21–34.

Andrefsky, W., Jr. (1995). Cascade phase lithic technology: An example from the Lower Snake River.North American Archaeologist16: 95–115.

Andrefsky, W., Jr. (1998).Lithics: Macroscopic Approaches to Analysis, Cambridge University Press,Cambridge.

Aoyama, K. (1995). Microwear analysis in the southeast Maya lowlands: Two case studies at Copan,Honduras.Latin American Antiquity6: 129–144.

Arnold, J. E. (1992). Complex hunter–gatherer–fishers of prehistoric California: Chiefs, specialists,and maritime adaptations of the Channel Islands.American Antiquity57: 60–84.

Arnold, J. E., and Munns, A. (1994). Independent or attached specialization: The origin of shell beadproduction in California.Journal of Field Archaeology21: 473–489.

Atchison, J., and Fullagar, R. (1998). Starch residues on pounding implements from Jinmium rock-shelter. In Fullagar, R. (ed.),A Closer Look, Archaeological Methods Series 6, Sydney University,Sydney, pp. 109–125.

Ballenger, J. A. M. (1996). The Southern Plains craft lithic cache.Plains Anthropologist41:297–309.

Bamforth, D. B. (1986). Technological efficiency and tool curation.American Antiquity51: 38–50.Bamforth, D. B. (1988). Investigating microwear polishes with blind tests: The Institute results in

context.Journal of Archaeological Science15: 11–23.Bamforth, D. B., and Bleed, P. (1997). Technology, flaked stone technology, and risk. In Barton, C.

M., and Clark, G. (eds.),Rediscovering Darwin: Evolutionary Theory and Archaeological Expla-nation, Archeological Papers of the American Anthropological Association, No. 7, Washington,DC, pp. 109–139.

Barton, C. M. (1997). Stone tools, style, and social identity: An evolutionary perspective on the ar-chaeological record. In Barton, C. M., and Clark, G. (eds.),Rediscovering Darwin: EvolutionaryTheory and Archaeological Explanation, Archeological Papers of the American AnthropologicalAssociation, No. 7, Washington, DC, pp. 141–156.

Barton, H., Torrence, R., and Fullagar, R. (1998). Clues to stone tool function reexamined: Comparingstarch grain frequencies on used and unused obsidian artefacts.Journal of Archaeological Science25: 1231–1238.

Bayman, J. M. (1995). Rethinking “redistribution” in the archaeological record: Obsidian exchange atthe Marana platform mound.Journal of Anthropological Research51: 37–63.

Becker, M., and Wendorf, F. (1993). A microwear study of a late Pleistocene Qadan assemblage fromsouthern Egypt.Journal of Field Archaeology20: 389–398.

Bednarik, R. G. (1995). Concept-mediated marking in the Lower Paleolithic.Current Anthropology36: 605–616.

Bettinger, R. L., and Eerkens, J. (1997). Evolutionary implications of metrical variation in Great Basinprojectile points. In Barton, C. M., and Clark, G. (eds.),Rediscovering Darwin: EvolutionaryTheory and Archaeological Explanation, Archeological Papers of the American AnthropologicalAssociation, No. 7, Washington, DC, pp. 177–191.

Bettinger, R. L., and Eerkens, J. (1999). Point typologies, cultural transmission, and the spread ofbow-and-arrow technology in the prehistoric Great Basin.American Antiquity64: 243–263.

Page 45: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 89

Bettinger, R. L., O’Connell, J. F., and Thomas, D. H. (1991). Projectile points as time markers in theGreat Basin.American Anthropologist93: 166–172.

Bienenfeld, P. (1995). Duplicating archaeological microwear polishes with epoxy casts.Lithic Tech-nology20: 29–39.

Binford, L. R. (1973). Interassemblage variability—the Mousterian and the “functional” argument. InRenfrew, C. (ed.),The Explanation of Culture Change: Models in Prehistory, Duckworth, London,pp. 227–254.

Bleed, P. (1986). The optimal design of hunting weapons: Maintainability or reliability.AmericanAntiquity51: 737–747.

Bourque, B. J. (1994). Evidence for prehistoric exchange on the Maritime Peninsula. In Baugh, T.,and Ericson, J. (eds.),Prehistoric Exchange Systems in North America, Plenum, New York,pp. 23–46.

Bousman, C. B. (1993). Hunter–gatherer adaptations, economic risk and tool design.Lithic Technology18: 59–86.

Bradbury, A. P. (1998). The examination of lithic artifacts from an Early Archaic assemblage: Strength-ening inferences through multiple lines of evidence.Midcontinental Journal of Archaeology23:263–288.

Byers, A. M. (1994). Symboling and the Middle-Upper Paleolithic transition: A theoretical and method-ological critique.Current Anthropology35: 369–381.

Byers, A. M. (1999). Communication and material culture: Pleistocene tools as action cues.CambridgeArchaeological Journal9: 23–41.

Carlson, R. L. (1994). Trade and exchange in prehistoric British Columbia. In Baugh, T., and Ericson,J. (eds.),Prehistoric Exchange Systems in North America, Plenum, New York, pp. 307–361.

Cattaneo, C., Gelsthorpe, K., Phillipos, P., and Sokol, R. J. (1993). Blood residues on stone tools:Indoor and outdoor experiments.World Archaeology25: 29–43.

Chatters, J. C. (1987). Hunter–gatherer adaptations and assemblage structure.Journal of Anthropolog-ical Archaeology6: 336–375.

Christensen, M. (1998). Processus de formation et caract´erisation physico-chimique des polisd’utilisation des outils en silex. Applications `a la technologie pr´ehistorique de l’ivoire.Bulletinde la Societe Prehistorique Franc¸aise95: 183–201.

Christensen, M., Walter, P., and Menu, M. (1992). Usewear characterisation of prehistoric flints withIBA. Nuclear Instruments and Methods in Physics Research B64: 488–493.

Clark, J. E., and Bryant, D. D. (1991). The production of chert projectile points at Yerba Buena, Chiapas,Mexico. In Hester, T., and Shafer, H. (eds.),Maya Stone Tools, Prehistory Press, Madison, WI,pp. 85–102.

Close, A. E. (1996). Carry that weight: The use and transportation of stone tools.Current Anthropology37: 545–553.

Coffey, B. P. (1994). The chemical alteration of microwear polishes: An evaluation of the Plisson andMauger findings through replicative experimentation.Lithic Technology19: 88–92.

Collin, F., and Jardon-Giner, P. (1993). Travail de la peau avec des grattoirs emmanch´es. Reflexions surdes bases exp´erimentales et ethnographiques. In Anderson, P., Beyries, S., Otte, M., and Plisson,H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 105–117.

Conard, N. J., and Adler, D. S. (1997). Lithic reduction and hominid behavior in the Middle Paleolithicof the Rhineland.Journal of Anthropological Research53: 147–175.

Cowan, F. L. (1999). Making sense of flake scatters: Lithic technological strategies and mobility.American Antiquity64: 593–607.

Custer, J. F., Ilgenfritz, J., and Doms, K. R. (1988). A cautionary note on the use of chemstrips fordetection of blood residues on prehistoric stone tools.Journal of Archaeological Science15:343–345.

Darras, V. (1994). Les mines-ateliers d’obsidienne de la region de Zinaparo-Prieto, Michoacan, Mex-ique.Bulletin de la Societe Prehistorique Franc¸aise91: 290–310.

Davidson, I., and Noble, W. (1993). Tools and language in human evolution. In Gibson, K. R., andIngold, T. (eds.),Tools, Language and Cognition in Human Evolution, Cambridge UniversityPress, Cambridge, pp. 363–388.

Deacon, J. (1992). Arrows as agents of belief amongst the/Xam Bushmen.Margaret Shaw Lecture 3,South African Museum, Cape Town.

Page 46: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

90 Odell

De Bie, M. (1998). Late Paleolithic tool production strategies: Technological evidence from Rekem(Belgium). In Milliken, S., and Peresani, M. (eds.),Lithic Technology: From Raw Material Pro-curement to Tool Production, Workshop No. 12 of the XIII International Congress of Prehistoricand Protohistoric Sciences, Forli, Italy, 1996, pp. 91–95.

Del Bene, T. A. (1979). Once upon a Striation: Current models of striation and polish formation. InHayden, B. (ed.),Lithic Use–Wear Analysis, Academic Press, New York, pp. 167–177.

Dibble, H. L. (1991). Mousterian assemblage variability on an interregional scale.Journal of Anthro-pological Research47: 239–257.

Dibble, H. L. (1995a). Middle Paleolithic scraper reduction: Background, clarification, and review ofthe evidence to date.Journal of Archaeological Method and Theory2: 299–368.

Dibble, H. L. (1995b). Raw material availability, intensity of utilization, and Middle Paleolithic as-semblage variability. In Dibble, H., and Lenoir, M. (eds.),The Middle Paleolithic Site of Combe-Capelle Bas (France), University Museum Monograph 91, University of Pennsylvania, Philadel-phia, pp. 289–315.

Dobres, M.-A., and Hoffman, C. R. (1994). Social agency and the dynamics of prehistoric technology.Journal of Archaeological Method and Theory1: 211–258.

Dockall, J. E. (1994). Oval biface celt variability during the Maya Late Preclassic.Lithic Technology19: 52–68.

Dockall, J. E., and Shafer, H. J. (1993). Testing the producer–consumer model for Santa Rita Corozal,Belize.Latin American Antiquity4: 158–179.

Downs, E. F., and Lowenstein, J. M. (1995). Identification of archaeological blood proteins: A cau-tionary note.Journal of Archaeological Science22: 11–16.

Edmonds, M. (1995).Stone Tools and Society: Working Stone in Neolithic and Bronze Age Britain, B.T. Batsford, London.

Eerkens, J. (1998). Reliable and maintainable technologies: Artifact standardization and the Early toLater Mesolithic transition in northern England.Lithic Technology23: 42–53.

Eisele, J. A., Fowler, D. D., Haynes, G., and Lewis, R. A. (1995). Survival and detection of bloodresidues on stone tools.Antiquity69: 36–46.

Ellis, C. (1994). Miniature Early Paleo-Indian stone artifacts from the Parkhill, Ontario site.NorthAmerican Archaeologist15: 253–267.

Feblot-Augustins, J. (1993). Mobility strategies in the late Middle Paleolithic of central Europe andwestern Europe: Elements of stability and variability.Journal of Anthropological Archaeology12: 211–265.

Fedick, S. (1991). Chert tool production and consumption among Classic period Maya house-holds. In Hester, T., and Shafer, H. (eds.),Maya Stone Tools, Prehistory Press, Madison, WI,pp. 103–118.

Fellner, R. (1995). Technology or typology? A response to Neeley and Barton.Antiquity69: 381–383.Fiedel, S. J. (1996). Blood from stones? Some methodological and interpretive problems in blood

residue analysis.Journal of Archaeological Science23: 139–147.Flenniken, J. J., and Raymond, A. W. (1986). Morphological projectile point typology: Replication,

experimentation, and technological analysis.American Antiquity51: 603–614.Flenniken, J. J., and Wilke, P. J. (1989). Typology, technology, and chronology of Great Basin dart

points.American Anthropologist91: 149–158.Fredericksen, C. F. K., and Sewell, B. (1991). The reliability of flaked tool function studies in New

Zealand archaeology.Archaeology of Oceania26: 123–126.Freeman, L. G. (1992). Mousterian facies in space: New data from Morin level 16. In Dibble, H.,

and Mellars, P. (eds.),The Middle Paleolithic: Adaptation, Behavior, and Variability, UniversityMuseum Symposium Series, Vol. 4, University of Pennsylvania, Philadelphia, pp. 113–125.

Fullagar, R. L. K. (1991). The role of silica in polish formation.Journal of Archaeological Science18:1–24.

Fullagar, R., Loy, T., and Cox, S. (1998). Starch grains, sediments and stone tool function: Evidencefrom Bitokara, Papua New Guinea. In Fullagar, R. (ed.),A Closer Look, Archaeological MethodsSeries 6, Sydney University, Sydney, pp. 49–60.

Galm, J. R. (1994). Prehistoric trade and exchange in the Interior Plateau of northwestern NorthAmerica. In Baugh, T., and Ericson, J. (eds.),Prehistoric Exchange Systems in North America,Plenum, New York, pp. 275–305.

Page 47: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 91

Garling, S. J. (1998). Megafauna on the menu? Haemoglobin crystallisation of blood residues fromstone artefacts at Cuddie Springs. In Fullagar, R. (ed.),A Closer Look, Archaeological MethodsSeries 6, Sydney University, Sydney, pp. 29–48.

Gassin, B., with Garidel, Y. (1993). Des outils de silex pour la fabrication de la poterie. In Anderson,P., Beyries, S., Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL,No. 50, Liege, pp. 189–203.

Geneste, J.-M., and Maury, S. (1997). Contributions of multidisciplinary experimentation to the studyof Upper Paleolithic projectile points. In Knecht, H. (ed.),Projectile Technology, Plenum, NewYork, pp. 165–189.

Gero, J. M. (1991). Genderlithics: Women’s roles in stone tool production. In Gero, J., and Conkey,M. (eds.),Engendering Archaeology, Basil Blackwell, Oxford, pp. 163–193.

Gibson, J. L. (1994). Empirical characterization of exchange systems in Lower Mississippi Valleyprehistory. In Baugh, T., and Ericson, J. (eds.),Prehistoric Exchange Systems in North America,Plenum, New York, pp. 127–175.

Gordon, D. (1993). Mousterian tool selection, reduction, and discard at Ghar, Israel.Journal of FieldArchaeology20: 205–218.

Grace, R. (1993a). New methods in use–wear analysis. In Anderson, P., Beyries, S., Otte, M., andPlisson, H. (eds.),Traces et foncion: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 385–387.

Grace, R. (1993b). The use of expert systems in lithic analysis. In Anderson, P., Beyries, S., Otte, M., andPlisson, H. (eds.),Traces et foncion: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 389–400.

Grace, R. (1996). Use–wear analysis: The state of the art.Archaeometry38: 209–229.Graves, P. (1994). My strange quest for Leroi-Gourhan: Structuralism’s unwitting hero.Antiquity68:

457–460.Grimaldi, S. (1998). Methodological problems in the reconstruction ofchaines operatoiresin Lower–

Middle Palaeolithic industries. In Milliken, S., and Peresani, M. (eds.),Lithic Technology: FromRaw Material Procurement to Tool Production, Workshop No. 12 of the XIII InternationalCongress of Prehistoric and Protohistoric Sciences, Forli, Italy, 1996, pp. 19–22.

Grimaldi, S., and Lemorini, C. (1993). Retouche specialis´ee et/ou chaine de ravivage? Les “racloirs”mousteriens de la Grotta Breuil (Monte Circeo, Italie). In Anderson, P., Beyries, S., Otte, M.,and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege,pp. 67–78.

Grimaldi, S., and Lemorini, C. (1995). Technology and microwear: Predetermined flakes from theMousterian site of Grotta Brueil (Monte Circeo, Italy). In Dibble, H., and Bar-Yosef, O. (eds.),The Definition and Interpretation of Levallois Technology, Prehistory Press, Madison, WI,pp. 143–155.

Gurfinkel, D. M., and Franklin, U. M. (1988). A study of the feasibility of detecting blood residue onartifacts.Journal of Archaeological Science15: 83–97.

Hardy, B. L., Raff, R. A., and Raman, V. (1997). Recovery of mammalian DNA from Middle Paleolithicstone tools.Journal of Archaeological Science24: 601–611.

Hayden, B. (1975). Curation: Old and new. In Raymond, J., Loveseth, B., Arnold, C., and Reardon, G.(eds.),Primitive Art and Technology, University of Calgary, Alberta, Canada, pp. 47–59.

Hayden, B., Franco, N., and Spafford, J. (1996). Evaluating lithic strategies and design criteria. InOdell, G. (ed.),Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York,pp. 9–49.

Head, L., and Fullagar, R. (1997). Hunter–gatherer archaeology and pastoral contact: Perspectives fromthe northwest Northern Territory, Australia.World Archaeology28: 418–428.

Hirth, K. G. (1995). The investigation of obsidian craft production at Xochicalco, Morelos.AncientMesoamerica6: 251–258.

Holdaway, S., McPherron, S., and Roth, B. (1996). Notched tool reuse and raw material availability inFrench Middle Paleolithic sites.American Antiquity61: 377–387.

Horsfall, G. A. (1987). Design theory and grinding stones. In Hayden, B. (ed.),Lithic Studies Amongthe Contemporary Highland Maya, University of Arizona Press, Tucson, pp. 332–377.

Hudler, D. (1997).Determining Clear Fork Tool Function through Use–Wear Analysis: A Discussionof Use–Wear Methods and Clear Fork Tools, Texas Archeological Research Laboratory, Studiesin Archeology 25, University of Texas, Austin.

Hudson, L. (1993). Protohistoric Pawnee lithic economy.Plains Anthropologist38: 265–277.

Page 48: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

92 Odell

Hughes, R. E. (1994). Mosaic patterning in prehistoric California—Great Basin exchange. In Baugh,T., and Ericson, J. (eds.),Prehistoric Exchange Systems in North America, Plenum, New York,pp. 363–383.

Hurcombe, L. (1992).Use–Wear Analysis and Obsidian: Theory, Experiments and Results, Depart-ment of Archaeology and Prehistory, Archaeological Monograph 4, University of Sheffield,Sheffield.

Hurcombe, L. (1997). The contribution of obsidian use–wear analysis to understanding the formationand alteration of wear. In Ramos-Millan, A., and Bustillo, M. A. (eds.),Siliceous Rocks andCulture, Editorial Universidad de Granada, Spain, pp. 487–497.

Hyland, D. C., Tersak, J. M., Adovasio, J. M., and Siegel, M. I. (1990). Identification of the species oforigin of residual blood on lithic material.American Antiquity55: 104–112.

Ingbar, E. E. (1994). Lithic material selection and technological organization. In Carr, P. (ed.),The Or-ganization of North American Prehistoric Chipped Stone Technologies, International Monographsin Prehistory, Ann Arbor, MI, pp. 45–56.

Johnson, J. K. (1996). Lithic analysis and questions of cultural complexity: The Maya. In Odell, G.(ed.),Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York, pp. 159–179.

Johnson, J. K., and Hayes, F. L. (1995). Shifting patterns of long-distance contact during the MiddleWoodland period in the northern Yazoo Basin, Mississippi. In Nassaney, M., and Sassaman, K.(eds.),Native American Interactions, University of Tennessee Press, Knoxville, pp. 100–121.

Jones, J. C. (1970).Design Methods: Seeds of Human Futures, John Wiley and Sons, New York.Kaufman, D. (1995). Microburins and microliths of the Levantine Epipaleolithic: A comment on the

paper by Neeley and Barton.Antiquity69: 375–381.Kay, M. (1996). Microwear analysis of some Clovis and experimental chipped stone tools. In Odell,

G. (ed.),Stone Tools: Theoretical Insights into Human Prehistory, Plenum Press, New York,pp. 315–344.

Kazaryan, H. (1993). Butchery knives in the Mousterian sites of Armenia. In Anderson, P., Beyries, S.,Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege,pp. 79–85.

Kealhofer, L., Torrence, R., and Fullagar, R. (1999). Integrating phytoliths within use–wear/residuestudies of stone tools.Journal of Archaeological Science26: 527–546.

Kelly, R. L. (1983). Hunter–gatherer mobility strategies.Journal of Anthropological Research39:277–306.

Kelly, R. L. (1988). The three sides of a biface.American Antiquity53: 717–734.Kimball, L. R., Kimball, J. F., and Allen, P. E. (1995). Microwear polishes as viewed through the atomic

force microscope.Lithic Technology20: 6–28.Kooyman, B., Newman, M. E., and Ceri, H. (1992). Verifying the reliability of blood residue analysis

on archaeological tools.Journal of Archaeological Science19: 265–269.Krakker, J. J. (1997). Biface caches, exchange, and regulatory systems in the prehistoric Great Lakes

region.Midcontinental Journal of Archaeology22: 1–41.Kuhn, S. L. (1992a). Blank form and reduction as determinants of Mousterian scraper morphology.

American Antiquity57: 115–128.Kuhn, S. L. (1992b). On planning and curated technologies in the Middle Paleolithic.Journal of

Anthropological Research48: 185–214.Kuhn, S. L. (1993). Mousterian technology as adaptive response: A case study. In Peterkin, G., Bricker,

H., and Mellars, P. (eds.),Hunting and Animal Exploitation in the Later Paleolithic and Mesolithicof Eurasia, Archeological Papers of the American Anthropological Association, No. 4, Washing-ton, DC, pp. 25–31.

Kuhn, S. L. (1994). A formal approach to the design and assembly of mobile toolkits.AmericanAntiquity59: 426–442.

Kuhn, S. L. (1995).Mousterian Lithic Technology: An Ecological Perspective, Princeton UniversityPress, Princeton.

Leach, J. D. (1998). A brief comment on the immunological identification of plant residues on pre-historic stone tools and ceramics: Results of a blind test.Journal of Archaeological Science25:171–175.

Leach, J. D., and Mauldin, R. P. (1995). Additional comments on blood residue analysis in archaeology.Antiquity69: 1020–1022.

Page 49: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 93

LeMoine, G. M. (1997).Use–Wear Analysis on Bone and Antler Tools of the Mackenzie Inuit, BARInternational Series 679, Oxford.

Levi Sala, I. (1993). Use–wear traces: Processes of development and post-depositional alterations. InAnderson, P., Beyries, S., Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves,ERAUL, No. 50, Liege, pp. 401–416.

Levi Sala, I. (1996).A Study of Microscopic Polish on Flint Implements, BAR International Series 629,Oxford.

Lewenstein, S. M. (1991). Edge angles and tool function among the Maya: A meaningful relationship?In Hester T., and Shafer, H. (eds.),Maya Stone Tools, Prehistory Press, Madison, WI, pp. 207–217.

Lewenstein, S. M. (1993). Experimentation in the formation and variability of lithic use–wear traceson obsidian and chert implements. In Anderson, P., Beyries, S., Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 287–294.

Lohse, E. S. (1996). A computerized descriptive system for functional analysis of stone tools.Tebiwa26: 3–66.

Loy, T. H. (1993). The artifact as site: An example of the biomolecular analysis of organic residues onprehistoric tools.World Archaeology 25: 44–63.

Loy, T. H., and Dixon, E. J. (1998). Blood residues on fluted points from eastern Beringia.AmericanAntiquity63: 21–46.

Loy, T. R., and Wood, A. R. (1989). Blood residue analysis at Cayonu Tepesi, Turkey.Journal of FieldArchaeology16: 451–460.

MacDonald, D. H. (1999). Modeling Folsom mobility, technological organization, and mating strategiesin the Northern Plains.Plains Anthropologist44: 141–161.

MacDonald, D. H., and Hewlett, B. S. (1999). Reproductive interests and forager mobility.CurrentAnthropology40: 501–514.

Mahmoud, A.-M. A., and Bard, K. A. (1993). Sources of the Predynastic grinding stones in the Hu-Semaineh region, Upper Egypt, and their cultural context.Geoarchaeology8: 241–245.

Manolakakis, L. (1996). Production lithique et ´emergence de la hierarchie sociale: l’industrie lithiquede l’Eneolithique en Bulgarie (premi`ere moitie du IVe millenaire).Bulletin de la SocietePrehistorique Franc¸aise93: 119–123.

Mansur, M. E. (1997). Functional analysis of polished stone-tools: Some considerations about thenature of polishing. In Ramos-Millan, A., and Bustillo, M. A., (eds.),Siliceous Rocks and Culture,Editorial Universidad de Granada, Spain, pp. 465–486.

Mauldin, R. P., Leach, J. D., and Amick, D. S. (1995). On the identification of blood residues onPaleoindian artifacts.Current Research in the Pleistocene12: 85–87.

McBrearty, S., Bishop, L., Plummer, T., Dewar, R., and Conard, N. (1998). Tools underfoot: Humantrampling as an agent of lithic artifact edge modification.American Antiquity63: 108–129.

McBryde, I. (1997). “The landscape is a series of stories.” Grindstones, quarries and exchange inAboriginal Australia: A Lake Eyre case study. In Ramos-Millan, A., and Bustillo, M. A. (eds.),Siliceous Rocks and Culture, Editorial Universidad de Granada, Spain, pp. 587–607.

McDonald, M. M. A. (1991). Technological organization and sedentism in the Epipaleolithic of DakhlehOasis, Egypt.African Archaeological Review9: 81–109.

McSwain, R. (1991). A comparative evaluation of the producer-consumer model for lithic exchangein northern Belize, Central America.Latin American Antiquity2: 337–351.

Milliken, S. (1998). The ghost of Childe and the question of craft specialization in the Paleolithic.In Milliken, S., and Vidale, M. (eds.),Craft Specialization: Operational Sequences and Beyond,BAR Series 720, Oxford, pp. 1–7.

Mitchell, D. R., and Shackley, M. S. (1995). Classic period Hohokam obsidian studies in southernArizona.Journal of Field Archaeology22: 291–304.

Morrow, J. E. (1997). End scraper morphology and use-life: An approach for studying Paleoindianlithic technology and mobility.Lithic Technology22: 70–85.

Morrow, T. A. (1996). Bigger is better: Comments on Kuhn’s formal approach to mobile tool kits.American Antiquity61: 581–590.

Moss, E. H. (1987). A review of “Investigating microwear polishes with blind tests.”Journal of Ar-chaeological Science14: 473–481.

Moulton, A. L., and Abler, T. S. (1991). Lithic beings and lithic technology: References from northernIroquoian mythology.Man in the Northeast42: 1–7.

Page 50: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

94 Odell

Nash, S. E. (1996). Is curation a useful heuristic? In Odell, G. (ed.),Stone Tools: Theoretical Insightsinto Human Prehistory, Plenum, New York, pp. 81–99.

Nassaney, M. S. (1996). The role of chipped stone in the political economy of social ranking. InOdell, G. (ed.),Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York,pp. 181–224.

Neeley, M. P., and Barton, C. M. (1994). A new approach to interpreting late Pleistocene microlithindustries in Southwest Asia.Antiquity68: 275–288.

Nelson, M. C. (1991). The study of technological organization. In Schiffer, M. (ed.),ArchaeologicalMethod and Theory, University of Arizona Press, Tucson, pp. 57–100.

Newman, M. E., Ceri, H., and Kooyman, B. (1996). The use of immunological techniques in the analysisof archaeological materials—a response to Eisele; with report of studies at Head-Smashed-Inbuffalo jump.Antiquity70: 677–682.

Newman, M. E., Yohe, R. M., II, Ceri, H., and Sutton, M. Q. (1993). Immunological protein residueanalysis of non-lithic archaeological materials.Journal of Archaeological Science20: 93–100.

O’Connell, J. F., and Inoway, C. M. (1994). Surprise Valley projectile points and their chronologicalimplications.Journal of California and Great Basin Anthropology16: 162–198.

Odell, G. H. (1994a). Assessing hunter–gatherer mobility in the Illinois Valley: Exploring ambigu-ous results. In Carr, P. (ed.),The Organization of North American Prehistoric Chipped StoneTechnologies, International Monographs in Prehistory, Ann Arbor, MI, pp. 70–86.

Odell, G. H. (1994b). Prehistoric hafting and mobility in the North American Midcontinent: Examplesfrom Illinois. Journal of Anthropological Archaeology13: 51–73.

Odell, G. H. (1996a). Economizing behavior and the concept of “curation.” In Odell, G. (ed.),StoneTools: Theoretical Insights into Human Prehistory, Plenum, New York, pp. 51–80.

Odell, G. H. (1996b).Stone Tools and Mobility in the Illinois Valley: From Hunting–Gathering Campsto Agricultural Villages, International Monographs in Prehistory, Ann Arbor, MI.

Odell, G. H. (1998). Investigating correlates of sedentism and domestication in prehistoric NorthAmerica.American Antiquity63: 553–571.

Odell, G. H., and Odell-Vereecken, F. (1980). Verifying the reliability of lithic use–wear assessmentsby “blind tests:” The low-power approach.Journal of Field Archaeology7: 87–120.

Olausson, D. (1990). Edge-wear analysis in archaeology: The current state of research.LaborativArkeologi 4, Humanistisk-Samhallsvetenskapliga Forskningsradet, Stockholm.

Oswalt, W. (1976).An Anthropological Analysis of Food-Getting Technology, Wiley and Sons, NewYork.

Otis Charlton, C. L. (1993). Obsidian as jewelry: Lapidary production in Aztec Otumba, Mexico.Ancient Mesoamerica4: 231–243.

Owen, L., and Unrath, G. (1989). Microtraces d’usure dues `a la prehension.l’Anthropologie 93:673–688.

Parry, W. J., and Kelly, R. L. (1987). Expedient core technology and sedentism. In Johnson, J. K.,and Morrow, C. A. (eds.),The Organization of Core Technology, Westview Press, Boulder, CO,pp. 285–304.

Paton, R. (1994). Speaking through stones: A study from northern Australia.World Archaeology26:172–184.

Pawlik, A. (1995).Die microskopische Analyse von Steingeraten: Experimente-Auswertungsmethoden-Artefaktanalyse, Urgeschichtliche Materialhefte 10, Verlag Archaeologica Venatoria, Tubingen.

Petraglia, M., Knepper, D., Glumac, P., Newman, M., and Sussman, C. (1996). Immunological andmicrowear analysis of chipped-stone artifacts from piedmont contexts.American Antiquity61:127–135.

Piperno, D. R., and Holst, I. (1998). The presence of starch grains on prehistoric stone tools fromthe humid Neotropics: Indications of early tuber use and agriculture in Panama.Journal ofArchaeological Science25: 765–776.

Plisson, H., and Mauger, M. (1988). Chemical and mechanical alteration of micro-wear polishes: Anexperimental approach.Helinium28: 3–16.

Pope, M. K. (1994). Mississippian microtools and Uruk blades: A comparative study of chipped stoneproduction, use, and economic organization.Lithic Technology19: 128–145.

Pope, M., and Pollock, S. (1995). Trade, tools, and tasks: A study of Uruk chipped stone industries.Research in Economic Anthropology16: 227–265.

Page 51: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 95

Prost, D.-C. (1993). Nouveaux termes pour une description microscopique des retouches et autresenlevements.Bulletin de la Societe Prehistorique Franc¸aise90: 190–195.

Pye, D. (1964).The Nature of Design, Studio Vista, London.Ramos-Millan, A. (1997). Flint political economy in a tribal society. A material-culture study in the

El Malagon settlement (Iberian southeast). In Ramos-Millan, A., and Bustillo, M. A. (eds.),Siliceous Rocks and Culture, Editorial Universidad de Granada, Spain, pp. 671–711.

Read, D. W., and Russell, G. (1996). A method for taxonomic typology construction and an example:Utilized flakes.American Antiquity61: 663–684.

Rees, D., Wilkinson, G. G., Grace, R., and Orton, C. R. (1991). An investigation of the fractalproperties of flint microwear images.Journal of Archaeological Science18: 629–640.

Rick, J. W. (1996). Projectile points, style, and social process in the Preceramic of central Peru. InOdell, G. (ed.),Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York,pp. 245–278.

Ricq-de Bouard, M., and Fedele, F. G. (1993). Neolithic rock resources across the western Alps:Circulation data and models.Geoarchaeology8: 1–22.

Rondeau, M. F. (1996). When is an Elko? In Odell, G. (ed.),Stone Tools: Theoretical Insights intoHuman Prehistory, Plenum, New York, pp. 229–243.

Rosen, S. A. (1996). The decline and fall of flint. In Odell, G. (ed.),Stone Tools: Theoretical Insightsinto Human Prehistory, Plenum, New York, pp. 129–158.

Rosen, S. A. (1997a). Beyond meat and milk: Lithic evidence for economic specialization in the EarlyBronze Age pastoral periphery in the Levant.Lithic Technology22: 99–109.

Rosen, S. A. (1997b).Lithics after the Stone Age: A Handbook of Stone Tools from the Levant,AltaMira Press, Walnut Creek, CA.

Rousseau, M. K. (1992).Integrated Lithic Analysis: The Significance and Function of Key-ShapedFormed Unifaces on the Interior Plateau of Northwestern North America, Department ofArchaeology, Publication No. 20, Simon Fraser University, Burnaby, BC.

Rowlett, R. M., and Robbins, M. C. (1982). Estimating original assemblage content to adjust forpost-depositional vertical artifact movement.World Archaeology14: 73–83.

Rozoy, J.-G. (1991). Typologie et chronologie.Paleo3: 207–211.Santone, L. (1997). Transport costs, consumer demand, and patterns of intraregional exchange: A

perspective on commodity production and distribution from northern Belize.Latin AmericanAntiquity8: 71–88.

Sassaman, K. E. (1992). Lithic technology and the hunter–gatherer sexual division of labor.NorthAmerican Archaeologist13: 249–262.

Sassaman, K. E. (1994a). Changing strategies of biface production in the South Carolina coastalplain. In Carr, P. (ed.),The Organization of North American Prehistoric Chipped Stone ToolTechnologies, International Monographs in Prehistory, Ann Arbor, MI, pp. 99–117.

Sassaman, K. E. (1994b). Production for exchange in the Mid-Holocene Southeast: A Savannah RiverValley example.Lithic Technology19: 42–51.

Sassaman, K. E. (1998). Crafting cultural identity in hunter–gatherer economies. In Costin, C.,and Wright, R. (eds.),Craft and Social Identity, Archeological papers of the AmericanAnthropological Association, No. 8. Washington, DC, pp. 93–107.

Schick, K. D., and Toth, N. (1993).Making Silent Stones Speak: Human Evolution and the Dawn ofTechnology, Simon and Schuster, New York.

Schreurs, J. (1992). The Michelsberg site Maastricht-Klinkers: A functional interpretation.AnalectaPraehistorica Leidensia25: 129–171.

Seeman, M. F. (1994). Intercluster lithic patterning at Nobles Pond: A case for “disembedded”procurement among Early Paleoindian societies.American Antiquity59: 273–288.

Sellet, F. (1993). Chaine op´eratoire: The concept and its applications.Lithic Technology18: 106–112.

Shackley, M. S. (1990).Early Hunter–Gatherer Procurement Ranges in the Southwest: Evidence fromObsidian Geochemistry and Lithic Technology, Ph.D. dissertation, Department of Anthropology,Arizona State University, Tempe.

Shafer, H. J., and Hester, T. R. (1991). Lithic craft specialization and product distribution at the Mayasite of Colha, Belize.World Archaeology23: 79–97.

Shea, J. J. (1992). Lithic microwear analysis in archaeology.Evolutionary Anthropology1: 143–150.

Page 52: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

96 Odell

Shea, J. J. (1999). Artifact abrasion, fluvial processes, and “living floors” from the Early Paleolithicsite of ‘Ubeidiya (Jordan Valley, Israel).Geoarchaeology14: 191–207.

Shea, J. J., and Klenck, J. D. (1993). An experimental investigation of the effects of trampling on theresults of lithic microwear analysis.Journal of Archaeological Science20: 175–194.

Sheppard, P. J. (1993). Lapita lithics: Trade/exchange and technology. A view from the Reefs/SantaCruz.Archaeology in Oceania28: 121–137.

Sheppard, P. J. (1996). Hard rock: Archaeological implications of chert sourcing in near and remoteOceania. In Davidson, J., Irwin, G., Leach, B. F., Pawley, A., and Brown, D. (eds.),OceanicCulture History, New Zealand Journal of Archaeology Special Publication, pp. 99–115.

Shott, M. J. (1986). Technological organization and settlement mobility: An ethnographic examination.Journal of Anthropological Research42: 15–51.

Shott, M. J. (1996). Anexegesis of the curation concept.Journal of Anthropological Research52:259–280.

Shott, M. J. (1997). Stones and shafts redux: The metric discrimination of chipped-stone dart andarrow points.American Antiquity62: 86–101.

Sievert, A. K. (1992).Maya Ceremonial Specialization: Lithic Tools from the Sacred Cenote atChichen Itza, Yucatan, Prehistory Press, Madison, WI.

Sievert, A. K. (1994). The detection of ritual tool use through functional analysis: Comparativeexamples from the Spiro and Angel sites.Lithic Technology19: 146–156.

Smith, M. E. (1990). Long-distance trade under the Aztec empire: The archaeological evidence.Ancient Mesoamerica1: 153–169.

Smith, P. R., and Wilson, M. T. (1992). Blood residues on ancient tool surfaces: A cautionary note.Journal of Archaeological Science19: 237–241.

Sobolik, K. D. (1996). Lithic organic residue analysis: An example from the Southwestern Archaic.Journal of Field Archaeology23: 461–469.

Spence, M. W. (1996). Commodity or gift: Teotihuacan obsidian in the Maya region.Latin AmericanAntiquity7: 21–39.

Stafford, M. (1999).From Forager to Farmer in Flint: A Lithic Analysis of the Prehistoric Transitionto Agriculture in Southern Scandinavia, Aarhus University Press, Aarhus, Denmark.

Stiner, M. C., and Kuhn, S. L. (1992). Subsistence, technology, and adaptive variation in MiddlePaleolithic Italy.American Anthropologist94: 306–339.

Storck, P. L. (1997).The Fisher Site: Archaeological, Geological and Paleobotanical Studies atan Early Paleo-Indian Site in Southern Ontario, Museum of Anthropology, Memoirs No. 30,University of Michigan, Ann Arbor.

Strauss, A. E. (1992). Jack’s Reef corner notched points in New England: Site distribution,raw material preference, and implications for trade.North American Archaeologist13:333–350.

Sullivan, A. P., III (1994). Adaptive diversity and limited-activity sites versus logistical mobility andexpedient technology: Adrift in normative thought.Journal of Anthropological Research50:159–167.

Takacs-Biro, K. (1997). Raw material economy of the Late Neolithic in Hungary. In Ramos-Millan, A.,and Bustillo, M. A. (eds.),Siliceous Rocks and Culture, Editorial Universidad de Granada, Spain,pp. 639–660.

Therin, M. (1998). The movement of starch grains in sediments. In Fullagar, R. (ed.),A Closer Look,Archaeological Methods Series 6, Sydney University, Sydney, pp. 61–72.

Thomas, D. H. (1981). How to classify the projectiles from Monitor Valley, Nevada.Journal ofCalifornia and Great Basin Anthropology3: 7–43.

Thomas, D. H. (1986). Points on points: A reply to Flenniken and Raymond.American Antiquity51:619–627.

Thompson, M. (1996). Correlation of Maya lithic and glyphic data.Lithic Technology21: 120–133.Tomenchuk, J. (1997). A parametric use–wear study of artifacts from Areas C and C-east. In Storck,

P. (ed.),The Fisher Site, Museum of Anthropology, Memoirs No. 30, University of Michigan,Ann Arbor, pp. 95–161.

Tuross, N., Barnes, I., and Potts, R. (1996). Protein identification of blood residues on experimentalstone tools.Journal of Archaeological Science23: 289–296.

Unger-Hamilton, R. (1989). Analyse exp´erimentale des microtraces d’usure: Quelques controversesactuelles.l’Anthropologie93: 659–672.

Page 53: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 97

van den Dries, M., and van Gijn, A. (1997). The representativity of experimental usewear traces. InRamos-Millan, A., and Bustillo, M. A. (eds.),Siliceous Rocks and Culture, Editorial Universidadde Granada, Spain, pp. 499–513.

Vierra, B. J. (1995).Subsistence and Stone Tool Technology: An Old World Perspective, Anthropo-logical Research Papers No. 47, Arizona State University, Tempe.

Villa, P. (1982). Conjoinable pieces and site formation processes.American Antiquity47: 276–290.Wallis, L., and O’Connor, S. (1998). Residues on a sample of stone points from the West Kimberley. In

Fullagar, R. (ed.),A Closer Look, Archaeological Methods Series 6, Sydney University, Sydney,pp. 149–178.

Walthall, J. A., and Holley, G. R. (1997). Mobility and hunter–gatherer toolkit design: Analysis of aDalton lithic cache.Southeastern Archaeology16: 152–162.

Walthall, J. A., and Koldehoff, B. (1998). Hunter–gatherer interaction and alliance formation: Daltonand the Cult of the Long Blade.Plains Anthropologist43: 257–273.

White, J. P. (1996). Rocks in the head: Thinking about the distribution of obsidian in Near Oceania. InDavidson, J., Irwin, G., Leach, B. F., Pawley, A., and Brown, D. (eds.),Oceanic Culture History,New Zealand Journal of Archaeology Special Publication, pp. 199–209.

Whittaker, J. C., Caulkins, D., and Kamp, K. A. (1998). Evaluating consistency in typology andclassification.Journal of Archaeological Method and Theory5: 129–164.

Wilke, P. J., and Flenniken, J. J. (1991). Missing the point: Rebuttal to Bettinger, O’Connell, andThomas.American Anthropologist93: 172–173.

Williams-Thorpe, O., Thorpe, R. S., Elliott, C., and Xenophontos, C. (1991). Archaeology, geochem-istry, and trade of igneous rock millstones in Cyprus during the Late Bronze Age to Romanperiods.Geoarchaeology6: 27–60.

Wright, J. V. (1994). The prehistoric transportation of goods in the St. Lawrence River Basin. InBaugh, T., and Ericson, J. (eds.),Prehistoric Exchange Systems in North America, Plenum, NewYork, pp. 47–71.

Wynn, T. (1991). Tools, grammar and the archaeology of cognition.Cambridge ArchaeologicalJournal1: 191–206.

Wynn, T. (1993). Layers of thinking in tool behavior. In Gibson, K. R., and Ingold, T. (eds.),Tools, Lan-guage and Cognition in Human Evolution, Cambridge University Press, Cambridge, pp. 389–406.

Yamada, S. (1993). The formation process of “use–wear polishes.” In Anderson, P., Beyries, S.,Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege,pp. 433–445.

Yamada, S., and Sawada, A. (1993). The method of description for polished surfaces. In Anderson,P., Beyries, S., Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL,No. 50, Liege, pp. 447–457.

Yerkes, R. W., and Kardulias, P. N. (1993). Recent developments in the analysis of lithic artifacts.Journal of Archaeological Research1: 89–119.

Yohe, R. M., II, Newman, M. E., and Schneider, J. S. (1991). Immunological identification ofsmall-mammal proteins on Aboriginal milling equipment.American Antiquity56: 659–666.

Young, L. C. (1994). Lithics and adaptive diversity: An examination of limited-activity sites innortheast Arizona.Journal of Anthropological Research50: 141–154.

BIBLIOGRAPHY OF RECENT LITERATURE

Akerman, K. (1998). A suggested function for Western Arnhem Land use-polished flakes and elouras.In Fullagar, R. (ed.),A Closer Look, Archaeological Methods Series 6, Sydney University, Sydney,pp. 179–188.

Amick, D. S. (1994). Technological organization and the structure of inference in lithic analysis:An examination of Folsom hunting behavior in the American Southwest. In Carr, P. (ed.),TheOrganization of North American Prehistoric Chipped Stone Tool Technologies, InternationalMonographs in Prehistory, Ann Arbor, MI, pp. 9–34.

Amick, D. S. (1999). Raw material variation in Folsom stone tool assemblages and the division oflabor in hunter–gatherer societies. In Amick, D. (ed.),Folsom Lithic Technology, InternationalMonographs in Prehistory, Ann Arbor, MI, pp. 169–187.

Page 54: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

98 Odell

Andrefsky, W., Jr. (1991). Inferring trends in prehistoric settlement behavior from lithic productiontechnology in the Southern Plains.North American Archaeologist12: 129–144.

Andrefsky, W., Jr. (1997). Thoughts on stone tool shape and inferred function.Journal of MiddleAtlantic Archaeology13: 125–143.

Bierwirth, S. L. (1996).Lithic Analysis in Southwestern France: Middle Paleolithic Assemblages fromthe Site of La Quina, BAR International Series 633, Oxford.

Bosquet, D., and Jardon Giner, P. (1999). Etude trac´eologique du site pal´eolithique moyen deRemicourt-En Bia FloI. Notae Praehistoricae(Namur)19: 21–28.

Boszhardt, R. F., and McCarthy, J. (1999). Oneota end scrapers and experiments in hide dressing: Ananalysis from the La Crosse locality.Midcontinental Journal of Archaeology24: 177–199.

Brooks, L., and Phillips, P. (eds.) (1989).Breaking the Stony Silence: Papers from the Sheffield LithicsConference 1988, BAR British Series, No. 213, Oxford.

Brown, J. A. (1996).The Spiro Ceremonial Center, Vol. II: The Collections, Museum of Anthropology,Memoirs No. 29, University of Michigan, Ann Arbor.

Calogero, B. L. A. (1992). Lithic misidentification.Man in the Northeast43: 87–90.Carr, P. J. (ed.) (1994).The Organization of North American Prehistoric Chipped Stone Tool Technolo-

gies, International Monographs in Prehistory, Ann Arbor, MI.Clark, J., and Parry, W. J. (1990). Craft specialization and cultural complexity.Research in Economic

Anthropology12: 289–346.Clemente-Conte, I. (1997). Thermal alterations of flint implements and the conservation of microwear

polish: Preliminary experimental observations. In Ramos-Millan, A., and Bustillo, M. A. (eds.),Siliceous Rocks and Culture, Editorial Universidad de Granada, Spain, pp. 525–535.

Cowan, F. L. (1994).Prehistoric Mobility Strategies in Western New York: A Small Sites Perspective,Ph.D. dissertation, Department of Anthropology, State University of New York, Buffalo.

Dawe, B. (1997). Tiny arrowheads: Toys in the toolkit.Plains Anthropologist42: 303-318.Duff, A. I., Clark, G. A., and Chadderdon, T. J. (1992). Symbolism in the Early Paleolithic: A conceptual

odyssey.Cambridge Archaeological Journal2: 211–229.Edmonds, M. (1990). Description, understanding and the chaine op´eratoire.Archaeological Review

from Cambridge9: 55–70.Franco, N. V. (1994). Maximizaci´on en el aprovechamiento de los recursos l´ıticos: un caso anal-

izado en el ´area interserrana Bonaerense. In Lanata, J., and Borrero, L. (eds.),Arqueologıa deCazadores-Recolectores; Lımites, Casos y Apertures, Arqueolog´ıa Contempor´anea 5, Edici´onEspecial, Buenos Aires, pp. 75–88.

Fullagar, R. L. K. (1993). Flaked stone tools and plant food production: A preliminary report onobsidian tools from Talasea, West New Britain, PNG. In Anderson, P., Beyries, S., Otte, M., andPlisson, H. (eds.),Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 331–337.

Fullagar, R. L. K. (1994a). Objectives for use–wear and residue studies: Views from an Australianmicroscope.Helinium34: 210–224.

Fullagar, R. L. K. (1994b). Traces of times past: Stone artefacts into prehistory.Australian Archaeology39: 63–73.

Fullagar, R., Furby, J., and Hardy, B. (1996). Residues on stone artefacts: State of a scientific art.Antiquity70: 740–745.

Geneste, J.-M. (1991). Syst`emes techniques de production lithique: Variations techno-´economiquesdans les processus de r´ealisation des outillages pal´eolithiques.Techniques et Culture17/18:1–35.

Gibaja, J. F., and Clemente, I. (1997). El tratamiento t´ermico del s´ılex y sus repercusiones en la determi-nacion de los rastros de uso. Algunos ejemplos del neol´ıtico en Catalu˜na.Revista d’Arqueologıade Ponent7: 153–160.

Grace, R. (1989).Interpreting the Function of Stone Tools: The Quantification and Computerisationof Microwear Analysis, BAR International Series 474, Oxford.

Graves, P. (1994). Flakes and ladders: What the archaeological record cannot tell us about the originsof language.World Archaeology26: 158–171.

Hardy, B., and Garufi, G. T. (1998). Identification of woodworking on stone tools through residue anduse–wear analyses: Experimental Results.Journal of Archaeological Science25: 177–184.

Hofman, J. L. (1991). Folsom land use: Projectile point variability as a key to mobility. In Montet-White, A., and Holen, S. (eds.),Raw Material Economies among Prehistoric Hunter–Gatherers,Publications in Anthropology, No. 19, University of Kansas, Lawrence, pp. 335–355.

Page 55: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

Stone Tool Research at the End of the Millennium 99

Hurcombe, L. (1992). L’analyse des traces d’usure sur l’obsidienne.l’Anthropologie96: 179–186.Ingold, T. (1993). Tool-use, sociality and intelligence. In Gibson, K. R., and Ingold, T. (eds.),Tools,

Language and Cognition in Human Evolution, Cambridge University Press, Cambridge, pp. 429–445.

Jardon Giner, P., and Sacchi, D. (1994). Traces d’usage et indices de r´eaffutages et d’emmanchementssur des grattoirs magdal´eniens de la Grotte Gazel `a Salleles-Cabardes (Aude-France).l’Anthropologie98: 427–446.

Kimball, L. R. (1994). Microwear analysis of Late and Terminal Archaic projectile pointsfrom the Padula site (36NM15), Pennsylvania.Journal of Middle Atlantic Archaeology10:169–179.

Korobkova, G. F. (1994). Stone tools and the beginning of agriculture in the Near East (in Russian).Archaeological News(St. Petersburg)3: 166–180.

Korobkova, G. F. (1999).Narzedzia w pradziejach: Podstawy badania funkcji metoda traseologiczna,Uniwersytet Mikolaja Kopernika, Torun.

Lewenstein, S. M. (1991). Woodworking tools at Cerros. In Hester, T., and Shafer, H. (eds.),MayaStone Tools, Prehistory Press, Madison, WI, pp. 239–249.

Lurie, R. (1989). Lithic technology and mobility strategies: The Koster site Middle Archaic. In Torrence,R. (ed.),Time, Energy and Stone Tools, Cambridge University Press, Cambridge, pp. 46–56.

Moloney, N., Raposo, L., and Santonja, M. (eds.) (1996).Non-Flint Stone Tools and the PaleolithicOccupation of the Iberian Peninsula, BAR International Series 649, Oxford.

Nelson, M. C. (1993). Grinding-tool design as conditioned by land-use pattern.American Antiquity58: 286–305.

Nielsen, A. E. (1991). Trampling the archaeological record: An experimental study.American Antiquity56: 483–503.

Odell, G. H. (1993). A North American perspective on recent archeological stone tool research.Palimpsesto(Buenos Aires)3: 109–122.

Odell, G. H. (1995). Is anybody listening to the Russians?Lithic Technology20: 40–52.Patterson, L. W. (1994). Incidental impact breakage of arrow points.La Tierra21: 30–38.Patterson, L. W. (1996). Drilling holes in shell.La Tierra23: 10–13.Patterson, L. W. (1997). Hunter–gatherer mobility: Limitations of interpretation.Houston Archeological

Society Journal117: 1–8.Pauketat, T. R. (1994).The Ascent of Chiefs: Cahokia and Mississippian Politics in Native North

America, University of Alabama Press, Tuscaloosa.Pawlik, A. F. (1996). Licht- und rasterelektronenmikroskopische Untersuchungen an geschafteten

Steingeraten aus Burgaschisee-Sud.Tubinger Monographien zur Urgeschichte11: 331–340.Pawlik, A. F. (1996). Die lichtmikroscopische Gebrauchsspurenanalyse an ausgewahlten Steinartefak-

ten von Henauhof Nord II. In Kind, C.-J. (ed.),Die letzten Wildbeuter: Henauhof Nord II und dasEndmesolithikum in Baden-Wurttemberg, Materialhefte zur Archaologie in Baden-Wurttemberg,No. 39, pp. 150–178.

Pawlik, A. F. (1998). Die mikroskopische Gebrauchsspurenanalyse der Silexwerkzeuge aus Reute-Schorrenried. In Mainberger, M. (ed.),Das Moordorf von Reute, Teraqua CAP, pp. 185–198.

Pelegrin, J. (1990). Prehistoric lithic technology: Some aspects of research.Archaeological Reviewfrom Cambridge9: 117–125.

Philibert, S. (1994). L’ochre et le traitement des peaux: R´evision d’une conception traditionnelle parl’analyse fonctionnelle des grattoirs ocr´es de la Balma Margineda (Andorre).l’Anthropologie98:447–453.

Prewitt, E. R., and Tomka, S. (1993). What do I call thee? Projectile point types and archaeologicalinterpretations: Perspectives from Texas.Lithic Technology18: 49–58.

Rovner, I., and Lewenstein, S. (1997).Maya Stone Tools of Dzibilchaltun, Yucatan, Becan and Chi-canna, Campeche, Middle American Research Institute, Publication No. 65, Tulane University,New Orleans.

Schultz, J. M. (1992). The use–wear generated by processing bison hides.Plains Anthropologist37:333–351.

Shanks, O. C., Kornfeld, M., and Hawk, D. (1999). Protein analysis of Bugas-holding tools: New trendsin immunological studies.Journal of Archaeological Science26: 1183–1191.

Shchelinskii, V. E. (1994). On the function of bifaces from the Mousterian site Zaskalnaja V, Crimea(in Russian).Archaeological News(St. Petersburg)3: 16–24.

Page 56: ODELL, G. Stone Tool Research at the End of the Millennium Classification, Function, And Behavior. 2001

P1: FPQ/GAV/FYR/FZI P2: FPX/LZR/FGG/FOP QC: FPX-FVI

Journal of Archaeological Research [jar] PP044-292955 February 16, 2001 8:56 Style file version Nov. 19th, 1999

100 Odell

Shchelinskii, V. E. (1999). Technologiya kamneobrabativayushchevo proizvodstva sredne paleolitich-eskoi ctoyanki norovo I b priazovie.Archeologicheski Almanach(Donetsk)8: 109–128.

Shea, J. J. (1993). Lithic use–wear evidence for hunting in the Levantine Middle Paleolithic. In An-derson, P., Beyries, S., Otte, M., and Plisson, H. (eds.),Traces et fonction: Les gestes retrouves,ERAUL, No. 50, Liege, pp. 21–30.

Shott, M. J. (1995). How much is a scraper? Curation, use rates, and the formation of scraper assem-blages.Lithic Technology20: 52–72.

Sievert, A. K. (1992). Root and tuber resources: Experimental plant processing and resulting microwearon chipped stone tools. In Anderson, P. (ed.),Prehistoire de l’agriculture, Monographie du CRA,No. 6, Editions du CNRS, Paris, pp. 55–66.

Skakun, N. N. (1994). Agricultural implements and the problem of spreading of agriculture in south-eastern Europe.Helinium24: 294–305.

Stapert, D., and Johansen, L. (1999). Flint and pyrite: Making fire in the Stone Age.Antiquity 73:765–777.

Tacon, P. S. C. (1991). The power of stone: Symbolic aspects of stone use and tool development inwestern Arnhem Land, Australia.Antiquity65: 192–207.

Tankersley, K. B. (1995). Seasonality of stone procurement: An Early Paleoindian example in north-western New York State.North American Archaeologist6: 1–16.

van Gijn, A. L. (1990).The Wear and Tear of Flint: Principles of Functional Analysis Applied to DutchNeolithic Assemblages, Analecta Praehistorica Leidensia 22, Leiden.