You are on page 1of 56

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 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, usewear 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
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, artifact


classification, even though relatively little work has been accomplished within the
last decade in developing this field. A major issue that typologists must contin-
ually confront is the consistency with which their systems are applied, but this
issue also has received little attention. In perhaps the only relevant recent study
of this type, Whittaker et al. (1998), using Sinagua pottery typologies from the
American Southwest, evaluate the consistency with which typologies are learned
and disseminated. They found that learning is usually accomplished by means of
a master who passes the knowledge along through generations, a process that can
result in relatively homogeneous typologies.
Classification is the subject of an important monograph involving a fraternal
collaboration of an archaeologist and a philosopher (Adams and Adams, 1991).
The authors approach is practical rather than theoretical: a typology should be
constructed for a specific purpose, selection is involved in all typology making,
and types should possess the essential properties of identity and meaning. These
principles are applied to pottery from Medieval Nubia, where the senior author
has worked for many years. This is the most exhaustive treatise on the subject of
classification to appear for several years.
Other typological systems more directly related to the lithic data base also
have been advanced recently, one of which was offered within the rubric of a
lithic manual (Andrefsky, 1998). Although this book is only partly concerned with
classification, the author presents a typological system that is clearly meant to be
a workable model for ready employment. Rozoy (1991) also provides a specific
typological construct for the French Epipaleolithic period. Through tight dating and
type definition, the author aimed to ascertain the origin and direction of diffusion
of various traits.
The foregoing classification systems are essentially paradigmatic in structure,
a quality that is not sufficiently precise for Read and Russell (1996). They prefer
using clustering algorithms on measurements of flake tools with obvious wear
under low-power magnification. Although their system is certainly more objective
than structures currently being used, it is not without problems. For example, the
system, which was applied to unretouched flakes, purports to show how tool shape
affects use, but shape is only one of a number of variables that determine how a
tool is held and utilized. And because it is doubtful that their level of usewear
determination 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 the
positive side, their quantitative system is capable of teasing out associations that
are not readily visible to intuitive typologists.
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 disciplines naive and innocent youth, the conventional
wisdom was that prehistoric tools were fashioned to follow the artisans mental
template, and this form was preserved until discovery of the tool in modern times
(Dibble, 1995a; Thomas, 1981, p. 15). However, several lines of evidence suggest
that this has seldom been the case. For instance, ongoing taphonomic evaluation,
a field initiated several decades ago, indicates that artifacts may undergo changes
of condition or stratigraphic placement between discard and discovery (Rowlett
and 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 often
served as portions of larger implements, and their forms and working edges were
occasionally sharpened and shaped during their use-lives. These modifications
altered the final form of tools found in the prehistoric record and may have been of
sufficient magnitude to change their archaeological classification. The notion that
this process may have the potential to alter archaeological topologies blossomed
in the 1980s and continued into the 1990s. Battles are currently being fought over
whether prehistoric tool modification renders our classificatory systems invalid,
or whether those changes are really inconsequential and our typologies are usable
after all.

Great Basin Projectile Points

One battleground for this issue has been projectile point types in the American
Great Basin. Early antagonists were Flenniken and Raymond (1986), who argued
that modification processes were of such magnitude as to alter archaeological
classification, and Thomas (1986), who disagreed. In more recent experiments
simulating hunting situations with 92 corner- and side-notched points, Flenniken
and Wilke (1989) found that breakage and subsequent reworking led to changes
in 32% of the type designations. Unfortunately, these assertions are unverifiable
as (1) modern knappers, not prehistoric people, did the reworking and (2) the
same researchers who reworked the points made the type assignments. Additional
arguments were employed by Bettinger et al. (1991), who tested the assumption
that, if sharpening events were sufficient to cause typological change, then certain
specific point types were likely, upon sharpening, to change into certain other
types. They depicted archetypal types that, upon sharpening, should be made
into rejuvenated forms. Their thesis is that, if Flenniken and Wilke were correct,
then archetypal forms discovered on archaeological sites should be larger, on the
whole, than rejuvenated forms. Weighing and measuring 6000 points from several
localities, they found that the supposed rejuvenated forms were as large as, or
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 Wilkes argument
(but see the latters rebuttal: Wilke and Flenniken, 1991). A study of three Great
Basin projectile assemblages by OConnell and Inoway (1994) also supports the
short chronology and refutes Flenniken and Wilkes model.
Rondeau (1996) recently attempted to resolve the dilemma by conducting
a detailed technological examination of Elko corner-notched points from one
Great Basin site. His results suggest substantial evidence for rejuvenation but,
contrary to Flenniken and Wilkes position, usually of a magnitude insufficient
to necessitate changes in type designation. He argues that his data derive from
only one site and that more research is needed to resolve the issue fully. I suggest
that one element that has been missing from this debate all along is indepen-
dent rejuvenation and typing of the same experimental collection by disinterested
parties.

Middle Paleolithic Assemblages

The other major battleground concerns Middle Paleolithic assemblages, par-


ticularly scrapers. Although the discussion has not been as rancorous as the one
over Great Basin projectile points, the principles being debated are similar. The
main 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 types
change 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 these
types were variablethough, on a larger scale, he also has indicated that Middle
Paleolithic typological similarities between southern France and the Near East are
caused by similarities in blank form, which influence the ultimate shape of tools
(Dibble, 1991). Other studies support the scenario that nonnormative factors such
as intensity of utilization, differential rates of tool reuse, and economizing behavior
are driving forces influencing Middle Paleolithic assemblage variability (Dibble,
1995b; Holdaway et 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 into
transverse scrapers at Grotta di SantAgostino, Italy. He agrees that tool form is
responsive 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 cave
of Grotta Breuil, Grimaldi and Lemorini (1993) examined flat-retouched Middle
Paleolithic tools. They concluded that it was not sharpening that determined final
tool form and trajectory of use, but the small dimensions of the nodules with which
the people started.
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 ago


in Israel to test the assertion that certain types, particularly Mousterian points,
show continuous variation. He found that, although these points may appear to
have been on the same reduction trajectory as convergent or double scrapers, they
are much too small for the amount of retouch on them. In other words, blank
selection for scrapers was completely different from that for points, supporting
the argument that the Mousterian point was conceived as an independent tool type
with a separate reduction trajectory. The choice of blank was more selective for
both points and scrapers than it was for notched pieces.
The use of microliths and microburins to define culture-stratigraphic units
in the Epi-Paleolithic of the Near East also has been challenged, for many of the
same reasons, by Neeley and Barton (1994). If accurate, their criticisms would
have a devastating effect on the culture history of the region, which is based on
these distinctions. However, their rationale does not appear very strong, and their
ideas have been hotly contested (Fellner, 1995; Kaufman, 1995).
It appears that for Middle and Epi-Paleolithic assemblages, as with Great
Basin projectile points, modification of a tool for purposes of sharpening and
shaping changes the form of that tool and sometimes its classification. The extent
of the typological changes, the exact situations in which they occur, and the role
of blank selection in the process are still being worked out.

Style

Researchers of artifact style agree that this is a devilishly difficult quality


to put ones finger on. Freeman (1992) was interested in applying the concept to
a Middle Paleolithic context, but concluded that the Bordesian typology then in
common usage excluded those attributes most relevant in depicting style in a series
of artifacts. But even if the typological construct were not a problem, the nature of
Middle Paleolithic assemblages would be, because people simply did not encode
much stylistic variability on their stone tools during this period. In a study that is
as pessimistic as Freemans, Barton (1997) defines style both as stochastic varia-
tion (like Dunnell) and as a conveyor of information, for example, for purposes of
boundary maintenance (like Wobst). Placing his analysis in an evolutionary frame-
work, he tested stochastic processes on small and large populations and concluded
that his theories dont worka refreshingly honest appraisal that underscores the
difficulty of recognizing passive style in the archaeological record.
Rick (1996) applies the concept of style to long sequences of projectile points
from two rock shelters in central Peru by breaking the typology into four hierarchi-
cal stylistic constructs that measure different degrees of inclusiveness. He was able
to discover strong social discontinuities, although such an analysis did not help
him understand the degree of social differentiation or the scale of the groups being
monitored. Like Freemans experience with the Bordesian typology, the amount of
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 Deacons (Deacon, 1992) study of Bushman arrows suggests
that 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 of
artifact will carry style with social overtones. The level at which archaeologists
comprehend the belief systems of most of the people they study is so low that they
have trouble understanding not only the messages being conveyed, but also the
types of artifacts conveying them.

FUNCTIONAL ANALYSIS

UseWear Analysis: General Considerations

The functional analysis of stone tools has developed in several directions, of


which the most avidly researched has been the study of traces of wear from uti-
lization, or usewear analysis. From somewhat contentious beginnings during
the 1970s and 1980s, the field has stabilized into positions that are more mutually
supportive. 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 of all 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 UseWear
Conference, Grace (1993a, p. 385) was able to state, methodologically a con-
sensus emerged so that different approaches (high and low power) were not seen
as competing techniques but alternative strategies dependent on the specific ar-
chaeological problem. Although I missed that consensus at Uppsala, Graces
evaluation is, on the whole, accurate.
It is becoming more common to read that both high- and low-magnification
methods were employed in an analysis; after all, they can complement each other
nicely. A recent analysis of Italian Mousterian assemblages illustrates this point:
these industries presented problems with the conservation of microtraces (edge
rounding, polish, and striations) which resulted in a lack of functional inference
this was partially compensated for by the analysis of macrotraces (microflake
scars) (Grimaldi and Lemorini, 1995, p. 146). It is disheartening to read of failed
blind tests like those reported in Fredericksen and Sewell (1991), in which the
analysts reported a use duration of only 515 min/toola duration that had already
generally been agreed on as too short to produce developed and interpretable wear
traces (Bamforth, 1988; Lewenstein, 1993; Moss, 1987). In any case, the critical
atmosphere of the past two decades has led to stronger empirical support for the
techniques promoted than would have been the case had less rigorous testing
methods been applied.
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 usewear 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 of
usewear on obsidian, a material with properties substantially different from the
flints and cherts that typically form the subject of such investigations.
Tests of commonly used analytical variables have also begun to appear. For
example, Lewenstein (1991) evaluated the hypothesis, proposed by Wilmsen in
the late 1960s, that certain tasks were associated with tools possessing specific
edge angles. Using data from the Belizian Maya habitation site of Cerros, she
concluded that edge angle overlaps activities so drastically that one cannot employ
edge angle to infer function, even when morphological class is considered. And in
a holistic evaluation of commonly employed descriptive variables, van den Dries
and van Gijn (1997) quantified 301 experiments to examine possible correlations
between tool motion/worked material and edge rounding, fracturing, and polish.
Their results demonstrate a large amount of functional overlap among attributes of
both polish and fracture. On the positive side, they established definite relationships
between certain fracture attributes and both motion and worked material, and
between certain polish attributes and worked material. This is a solid start in
quantifying usewear variables, a kind of study the field needs very much.

UseWear Issues

Genesis of Polish Formation

Clearly, the most compelling issue among high-magnification usewear prac-


titioners these days is the genesis of polish. This debate has been raging for years
and 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 flints
acquire polished surfaces through utilization the same way glass does, or whether
other 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 by
a substance that she interpreted to be silica gel (see also Del Bene, 1979). This
interpretation 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 exclusively
an abrasive phenomenon. In recent years Levi Sala (1993, 1996) has emphasized
the 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 not
essential to this process, but promotes polish formation. Experiments by Yamada
(1993), concentrating on one tiny location on a piece of progressively utilized
siliceous shale, have supported this position. He argued that, if the silica gel theory
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 as


polish develops; but if the abrasive theory is viable, then these features should be
gradually smoothed without changing their position. Yamada observed no depo-
sition of material, and the surficial pits that he was using as markers retained their
integrity through 3350 strokes, supporting the tribological theory. Grace (1993a,
1996) has agreed with these arguments, stating categorically that the silica gel
theory has been proven untenable (Grace, 1996, p. 211).
Just as we thought the issue was resolved, however, up pops somebody with
evidence to the contrary. Using a particle accelerator and a scanning electron micro-
scope (SEM) connected to an energy-dispersive X-ray spectrometer, Christensen
(1998; Christensen et al., 1992) indicated that, during use, tiny bits of worked
material accumulate in spaces on the surface of the flint and within the lepisphere
lattice. Elements in these bits can be analyzed by a spectrometer, providing an idea
of their composition. This scenario was tested on tools known to have been used
on bone, showing deposition of calcium and phosphorous; and on ivory, which is
known to have a large magnesium component, an element that shows up promi-
nently on the spectrometer. Not only does this add support to the depositional
theory, but it offers a direct method for ascertaining materials on which flint tools
were workedprovided, of course, that one can be certain which little bits belong
to the erstwhile worked material. In addition, Christensen (1998) implanted a tiny
piece of copper on the surface of a tool, which she did not succeed in rubbing
away, 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 mechanical
abrasion, 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 through
experimentation to correlate amount of silica with kind and amount of polish,
demonstrated that even small amounts of amorphous silica may play a significant
role in polishing, and that different polishes develop at different rates.
The complexity of the situation and the possibility that both mechanical and
chemical 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 by
the 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 as
Rabinowiczs molecular theory, which involves mechanical removal of material
and subsequent chemical attack, enhanced by water and abrasives, resulting in a
thin surface film.
A principal goal in all this research is to correlate discrete polish types
with specific worked materials. Doubters exist. For example, Rees and his col-
leagues investigated the fractal properties of flint, which they defined as spatial
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
(Rees et al., 1991, p. 630). They wanted to ascertain whether or not microwear
polishes are fractal and if different contact substances produce fractally different
polishes. They found that both polished and unpolished surfaces are fractal and
can be distinguished from one another, but there is no correlation between frac-
tal dimensions and specific worked materials. Grace (1993a) arrived at the same
conclusion 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 polish
attributes and found clear correspondence between worked material type and all
10 attributes. And researchers at Tohoku University identified 11 basic types of
polish on shale that are principally the result of the material worked (Aoyama,
1995, p. 131). Further experimentation by Aoyama (1995) confirmed that these
polish types can also be applied to chalcedony and agate.

Prehensile Wear

The recognition of prehensile traces has been slow in coming. Analysts such
as Collin and Jardon-Giner (1993) have reported difficulty in interpreting hafting
wear on stone tools. Having replicated more than 300 hafted hide scrapers, they
were able to distinguish no definitive hafting wear types, only general trends. When
they hafted the scrapers with resin and wax, they observed no hafting traces at all.
Other analysts, however, have reported a considerably greater occurrence of
prehensile damage, perhaps caused by different hafting practices or raw material.
Owen and Unrath (1989) conducted a series of blind tests to determine their ability
to discriminate prehensile wear from damage on an active tool edge. They asserted
that prehensile damage was produced frequently, it was sometimes mistaken for
soft 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 of
prehistory in the North American Midwest and found that hafting traces increased
through time, whereas damage from manual prehension decreased. Increased haft-
ing practices appear to be related to changing mobility strategies and increased
sedentism, for which hafting represented an attempt to produce fail-safe equipment
that 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
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 the
greatest problems with implements employed on soft materials. These experiments
established that trampling damage can be incorrectly interpreted as usewear.
McBrearty et al. (1998) tested whether the effects of trampling were negligi-
ble or whether they mimic retouch. The experimenters found that many of the tools
sustained substantial edge damage, several of them qualifying as pseudotools such
as notched pieces and denticulates. Thus trampling damage is a potentially con-
founding influence, not only for functional but even for typological interpretations.

Application of UseWear Analysis

The field of usewear analysis is still new enough that unique or unknown
phenomena are frequently reported. In one such study of tools from Egyptian
Nubia, Becker and Wendorf (1993) detected a type of polish, probably related
to a soft substance, that they were not able to replicate. In a study with a more
satisfying (though not necessarily more correct) outcome, Gassin (1993) noticed
brightly polished flint tools from a Neolithic Chasseen site in France. Replicative
experimentation established a visual correlation between this wear and wear from
cutting 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 range
of tasks in which implements of Olduwan and Acheulean people may have been
engaged. Concentrating on Mousterian tools, Kazaryan (1993) determined that ob-
sidian flakes and convergent scrapers from two sites in Armenia were specialized
butchery 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 probably
caused by the small cobble flint that these people used for their tools. Functionally
varied, the Grotta Breuil industry constitutes an attempt to reconcile the pursuit of
large preforms with the optimal exploitation of large cobbles. Applications of these
techniques to Holocene assemblages include an analysis of a Neolithic Michels-
berg village (Schreurs, 1992) and a functional comparison of a Mesolithic hunting
camp with a Neolithic village (Pawlik, 1995).
Several New World applications of usewear analysis have emphasized the
working of wood. At the Mayan settlement of Cerros in Belize, Lewenstein (1993)
found that woodworking with a wide variety of tool formssome hafted, some
notwas the most common activity at the site. Similarly, a study of key-shaped
unifaces excavated from the Interior Plateau of British Columbia established that
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. Hudlers (Hudler,
1997) analysis of Clear Fork gouges established that their principal function was
for 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 conducted
with gravers and beaked scrapers at the Fisher site in Ontario. In another study, a
cache 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 showed
hide-scraping wear, a tool kit that probably served as insurance gear for occasional
logistical forays (Ballenger, 1996). Examining tools from the area around Modoc
Rockshelter in Illinois, Ahler (1998) established major differences between Early
and Middle Archaic assemblages on the basis of functional diversity. On the other
hand, Bradbury (1998) found very little use at all on an assemblage from southern
Kentucky, supporting his interpretation of the place as a location in which the only
discernible activity was biface reduction.

Recent Advances in UseWear Analysis

Advances in this field have been made on several fronts. High-magnification


analyses 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) appear
to be resolvable. Bienenfeld (1995) has shown that, despite problems of bubbling,
possible melting with heat, and being time-consuming, epoxy casts replicate polish
well enough for use in microwear analysis, even using the scanning electron mi-
croscope. On the sticky issue of artifact preparation, Coffey (1994) was unable to
duplicate the degree of chemical erosion from alkali solutions reported by Plisson
and Mauger (1988). He concluded that the previous researchers must have used
heated solutions.
Advances in low-magnification analysis have been slower in coming, be-
cause fewer people have been working on improving the technique. An exception
is Tomenchuk (1997), who has developed a parametric usewear method using
engineering principles of fracture mechanics on edge scarring. He recently applied
these principles to two artifact concentrations at the early Paleoindian Fisher site
in southern Ontario, concluding that both areas contained the same suite of pre-
historic activities. He also found that beaked, single-spurred, and double-spurred
scrapers all had a multiplicity of uses.
Two other research programs have the potential to advance the discipline. One
involves topographic measuring techniques, for which Kimball et al. (1995) have
employed the atomic force microscope. This equipment is capable of producing
quantitative measures of surface attributes such as polish and three-dimensional
digital mapping of these features. Anderson et al. (1998) also reported research with
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 bearing


area, average valley, and so forth. At this point, the authors can only report that
bearing area perimeter increases with use and that tool raw material influences the
results.
The other innovation is the development of expert systems of analysis that
computerize a large range of usewear attributes derived from low- and high-
magnification microscopes. Although Grace (1993b) has touted its accuracy and
speed (he even calls it the FAST system), it nevertheless remains quite complex,
considerably slower than some alternatives, and applicable only to fine-grained
flint. Yet it constitutes a real advance for certain kinds of applications and illustrates
the rapid changes that this field has undergone during the past decade. A similar
computerized descriptive system for functional analysis has been proposed by
Lohse (1996).

Residue Analysis

In residue analysis, the researcher isolates substances that adhere to the surface
of a stone tool. The preferred isolate is, of course, a substance that was intimately
associated with that tool during its use-life. Particularly informative are parts of
the 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 rats
with their everready rat knives? Herein lies one of the difficulties with residue
analysis: substances adhering to the surface of a tool are usually assumed to be
associated with that tools use, a logical leap that is not without danger. Many other
problems with residue analysis also exist. Because there is abundant evidence that
prehistoric people processed both plants and animals with stone tools, I consider
each separately.

Blood Residues: Positive Results

When an animal is butchered, some of the blood from that animal may stay on
the butcher knife. Tom Loy first discovered this eventuality in the mid-1980s and
has been the principal advocate of blood residue analysis ever since. Borrowing
from medical labs, he has typically employed several techniques for detecting
blood.
Loy recently discovered several bifaces eroding out of a roadbed in British
Columbia. On these tools he detected bison hair, though bison have not inhabited
this region for many years. His protein analysis consisted of a Hemastix test to
detect haemoglobin (Hb) and myoglobin (Mb), an immunological dotblot screen-
ing test for mammalian immunoglobin type G (IgG), and isoelectric focusing to
identify haemoglobin and serum albumin (SA). He employed four methods to de-
termine species of origin: radioimmunoassay, isoelectric focusing, haemoglobin
crystallization of proteins, and DNA analysis. The presence of bison blood was
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 purified
sample of the blood yielded a date of 2180 160 bp (Loy, 1993). In similar types
of analysis, Loy and his colleagues have identified human and bovine blood on a
large stone altar in the Skull Building at Neolithic Cayonu Tepesi, Turkey (Loy
and 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 blood
residue extraction. In an analog to Loys bison butchering bifaces, Kooyman et al.
(1992) tested implements from the Head-Smashed-In bison kill in Alberta, where
they claimed that blood residues had remained on the tools for 5600 years. Further
testing at that site (Newman et al., 1996) yielded positive reactions to bison and
elk antisera for 9 of 31 stone tools and 6 of 16 soil samples tested. And like Loys
study of fluted points, Hyland et al. (1990) tested tools from the Shoop Paleoindian
site in Pennsylvania. Of the 45 artifacts evaluated, 13 tested positive for deer or
caribou (these species are cross-reactive).
Grinding or pounding tools also have been evaluated by this technique. Yohe
et al. (1991) established through ethnographic analysis that animal material was
commonly pulverized. Their immunological analysis of milling equipment from
southern California returned positive indications that three manos, a mortar, and a
pestle were employed prehistorically to grind or pulverize small rodents.

Blood Residues: Persistent Problems

Despite the lofty claims of its proponents, however, blood residue analysis
has experienced a rising tide of dissatisfaction. A basic lack of agreement exists
between the results of blood residue analysis and those of many other kinds of
studies, outlined in stark detail in Fiedel (1996). From bovine blood in Ontario,
where prehistoric bison bones have never been recovered, to chickens in the
Archaic of Oregon, to a lack of trout in Trout LakeFiedel recounted a litany
of inconsistencies between the results of blood residue analyses and what we
thought 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. Downs
and 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 the
archaeological specimens (Downs and Lowenstein, 1995, p. 14). In a final phase
of the blind tests, residues on six Clovis blades were analyzed. Again, agreement
was reached on the control samples, but absolutely no agreement was achieved
with the archaeological material. These results are quite frustrating, because no
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. Wallis
and OConnor (1998) experienced similar inconsistencies between Hemastix and
immunological techniques when analyzing projectile points from two rock shelters
in northwest Australia. These points had retained very little blood but a lot of plant
residues, a finding that does not jibe with ethnographic accounts of aboriginal use
of these kinds of points in the region.
Also frustrating are cases of disagreement between the results of blood residue
and usewear 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 region
of northern Virginia (Petraglia et al., 1996). Twenty tested positive for blood
residues, 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 positive
with both techniques, a rather low rate of correspondence. Utilization traces were
poorly developed on these tools, a probable result of expedient use and chemical
degradation of polish. Just as significantly, The fact that 80 artifacts did not test
positively for immunological results also tends to support evidence for degradation
processes (Petraglia et 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 that
difference has something to do with differential preservation of blood over time.
Experiments have been devised to investigate this situation. In one series, Gurfinkel
and Franklin (1988) buried glass slides with and without blood in soil for varying
lengths of time. They detected degradation in some of the samples and concluded
that 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 the
presence of blood; protein must be preserved for immunological reactions to oc-
cur, thereby enabling specific identification (Fiedel, 1996, p. 145; Kooyman et al.,
1992, p. 265).
In a similar series of experiments, Cattaneo et al. (1993) buried 12 caches
of 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 on
meat and bone, but only for 2 min apiece; additional tools were stained with
blood and retained in the laboratory. Using a sensitive process called enzyme-
linked immunosorbent assay (ELISA), the researchers again obtained wonderful
success with the laboratory pieces, but more negative than positive reactions for
the buried artifacts. The positive identification of albumin on a scraper buried
for 1 year but negative reactions on other similarly buried artifacts illustrates
the erratic survival of blood under these conditions. Eisele et al. (1995) reported
similarly disappointing results from both archaeological samples and actualistic
experiments coating artifacts with blood and burying them in different environ-
ments for varying lengths of time. And in experiments in which freshly knapped
tools were used to cut and scrape one of 10 different species of animals and then
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 Mauldin et al., 1995) reported correct identification of only 37% of the
specimens.
Just as disturbing as the false negatives on pieces that used to have blood on
them are the false positives obtained from soils. Gurfinkel and Franklin (1988)
observed that haemoglobin binds chemically to clay and other soil particles, a
process that is difficult to recover by liquid extraction. However, Newman et 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 blood
residues, Custer et al. (1988) found that false positive reactions can be caused by
manganese oxide in the soil.
Doubts that haemoglobin molecules survive in their original state suitable for
crystallization have existed for years (Smith and Wilson, 1992). Recently Garling
(1998) conducted another series of experiments on modern and ancient samples
from the 30,000-year old Cuddie Springs site in Australia. Like researchers dis-
cussed previously, she noted inconsistencies in the three techniques used to identify
blood residues. More disturbing yet, she was unable to identify any of the residues
on these tools to the species level, as their crystal morphologies were not distin-
guishable using criteria that Loy had established for species identification: this
study shows that the similarity of crystal morphology extends across species that
are definitely unrelated (Garling, 1998, p. 42). She also noted the high potential
for confusing concentration-specific with species-specific traits, as well as the cur-
rent lack of knowledge of the ligand form of haemoglobin preserved in ancient
blood residues and bone samples. She concluded that positive results in either or
both Hemastix and immunoblots were not a reliable indicator of a samples ability
to produce consistent or regular haeomoglobin crystals, or in fact of its ability
to produce crystals at all. Indeed, negative results appeared to be a more reliable
indicator (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 likely
to be preserved on a tool. In an experiment to investigate this parameter, Tuross
et al. (1996) butchered a goat with seven tools and concluded that, in general, very
little blood is preserved on a tool following its utilization. Archaeologists should
therefore define the minimum amount of information needed from a particular
analysis, as sufficient residue may not exist to conduct all the assays one might
wish.
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, Fiedels (Fiedel, 1996) critical appraisal
of this field called for further blind testing with less ambiguous samples, just as
we have seen in usewear analysis. If these prove unsatisfactory, it will be time to
suspend use of the technique.
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. Various
kinds of plant residues, for example, starch grains, resins, and phytoliths, also have
been observed on prehistoric implements of considerable antiquity. Research on
plant residues has been occurring in pockets all over the world, especially in
Australia.
A preliminary question is whether starch grains recovered from a particular
soil or stratigraphic level are in primary context or were washed down through
overlying sediments. This issue was broached by Therin (1998), who conducted a
series of experiments to determine the rate of movement of starch grains through
sediments, using sands of different composition. He found that, the greater the size
of the grains, (1) the less their chance of becoming mobile, (2) the slower they
move, 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. He
also established a relationship with sediment size, that is, the smaller the particle
size of the sediment, the fewer starch grains move through it, though irrigation
levels do have a positive effect on grain movement. His study is encouraging, for
it suggests that, under most conditions, starch grains are likely to remain in the
location at which they were deposited.
In considering starch grains on prehistoric stone tools, it is important to know
whether the grains observed have stayed with the tool since its utilization or were
part of the sediment that adhered to the tool. This question was recently assessed by
Fullagar et al. (1998), who were analyzing obsidian implements from Papua New
Guinea. Although usewear analysis on the tools suggested the processing of tubers
such as yam or taro, ethnographic studies indicated that people in that region were
inclined to employ shell for these activities, not obsidian. The authors conducted
tuber-processing experiments and tested the sediments in which these experiments
were performed. They concluded that starch grain density was substantially greater
in the experimental sediments than in sediments in which the obsidian tools had
been excavated, bolstering their argument that the starch grains on the tools came
from tuber processing. Starch grains from processing tubers such as manioc, yam,
and arrowroot, along with maize, legumes, and palm, have also recently been
reported on early Holocene stone implements from four sites in Panama (Piperno
and Holst, 1998).
In another pilot study of a rockshelter in northern Australia, Atchison and
Fullagar (1998) observed starch grains on each of three stone pounding tools.
Tests indicated that the tools contained greater densities of starch grains than the
surrounding sediments did and that large differences existed in the sizes and shapes
of grains between the spits and the tools. These tests rendered it unlikely that the
implements were contaminated by contact with sediments and substantiated the
contention that the starch originated with processing activities. A similar argu-
ment was employed by Barton et al. (1998), who discovered significantly greater
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 other
tools from Middle and Late Archaic levels at Hinds Cave, Texas. The most fre-
quent residues on these implements were plant phytoliths, followed by plant fiber
and animal hair. Plants identified were sotol, yucca, agave, and grass. No correla-
tion between type of residue and type of usewear was observed, the tools being
very unspecialized and used for a variety of tasks. But in Australia, Kealhofer
et al. (1999) conducted a pilot study in which they compared usewear and phy-
tolith studies of the same group of artifacts. The correspondence between the two
techniques was excellent, providing some assurance that they will be regarded as
complementary and useful approaches to similar problems.
In general, plant residue analysis has experienced neither the problems nor
the depth of soul-searching that has characterized blood residue analysis. This is
primarily 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 blood
residues require immunological and other indirect tests to detect their presence. So
problems of agreement between assays has simply not occurred because, in most
cases, only one testobservationwas 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 issue
of differential preservation of plant residues has not been studied.
One negative finding about plant residues has been reported, and this study
was conducted not through direct observation, but using crossover immunoelec-
trophoresis (CIEP). Leach (1998) employed 19 flakes, manos, and hammers to
cut, scrape, grind, and pulverize desert-adapted plants such as yucca, agave, corn,
squash, and mesquite. He then boiled the vegetal material in a plain brownware
pot, dried the tools in sunlight, and refrigerated them for 12 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 beans
ground by a mano were identified as mesquite, were correct results obtained from an artifact
coated with an experimental residue that had an antiserum developed for that plant.

Leach attributed the failure of CIEP to produce accurate results to heating in


the 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 immunological
tests such as CIEP have seldom been applied to plants, questions of preservation
have seldom been asked in this way. Researchers directly observing residues such
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 is
missing. In other words, they are not inclined to inquire whether Tool B, on which
no such residues were observed, became that way because the tool was not used in
the first place, or because the residues originally present have degraded. If the latter
is the case, it implies that all those analyses that have been conducted through direct
observational methods may be usable only to show the prehistoric presence of a
particular plant. They cannot be employed to provide a representative measure of
the 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 blood
residue 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. Usewear
analysis may be helpful as a test with which to compare plant residue results.
This problem also applies to the analysis of ancient DNA. Although Hardy
et al. (1997) have extracted DNA from stone implements from the French
Mousterian site of La Quina, they reported that DNA from these ancient sources
is present only in small fragments. Data from their modern experimental series
suggested that DNA breakdown occurs rapidly after death. Thus although the ex-
traction of ancient DNA from soils or stone tools appears promising, research has
only begun to uncover the complexity of the situation.

BEHAVIORAL PROCESSES

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

Technological Organization

Establishing Mobility Parameters

In the late 1970s and early 1980s, Lewis Binford presented his now-famous
foragercollector model of huntergatherer mobility organization. This model has
had tremendous influence, at least in North America, on archaeological perceptions
of how huntergatherer 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 alternative
solutions not adequately expressed in either (Chatters, 1987; Nelson, 1991). In
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, Kellys (Kelly, 1983) cross-cultural study indicated that the mobility
strategies of even one group might differ from one year to the next.
But despite its shortcomings, Binfords model remains a useful framework
for structuring archaeological data. Some of its tenets were refined by Shott (1986)
who, breaking residential mobility into frequency and magnitude, determined that
technological 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 between
various forms of mobility? To answer this question, it is useful to consider some
archaeological implications of mobility, that is, kinds of preservable objects that
highly mobile people would be likely to carry around with them. The ultimate
question to a forager, of course, is how to perform those tasks necessary to survive
with the least amount of transport cost. Ever since Kellys (Kelly, 1988) discourse
on 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 been
supported through studies such as the one by Morrow (1997), which established
that bifacial faceting on midwestern Paleoindian end scrapers increased with dis-
tance from the source of tool raw material. But if this relationship is universally
true, then why do we not find this strategy in the European Upper Paleolithic, or
among Australian Aborigines?
Proceeding from different assumptions, Kuhn (1994) described a mobile
toolkit that would optimize potential usefulness with respect to weight (portabil-
ity). Perceiving that there exists a point at which large increments in weight bring
minimal gains in utility, Kuhn illustrated this as the highest point of a utility:mass
ratio graphed against multiple of minimum length. The optimal ratio results, on
average, 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 highly
mobile foragers to carry several smaller tools than an equivalent weight in larger
tools.
Morrow (1996) has criticized Kuhns model as being less efficient than logical
alternatives in certain circumstances (e.g., scrapers only 1.5 minimum usefulness
would quickly be reduced to an unusable state without the possibility of creating
larger, more usable tools). Perhaps a judicious assortment of smaller flakes and
larger cores, which themselves could be used for a variety of tasks, would provide
an optimal solution. In any case, the association of mobility with different techno-
logical strategies suggests that more than one solution was possible and that the
choice depended as much on historic trajectory as on ultimate efficiency.
To investigate potential solutions, it is often necessary to approach the problem
from 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 the
manufacture of those tools. Concentrating on small sites in northwestern New York,
he determined from their debitage that the technological strategies employed by
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 were
distinctly different.

Risk and Stress

Once a satisfactory level of tool efficiency has been achieved by a particular


group of huntergatherers, the successful strategy tends to continue until essential
parameters of the system change, rendering the strategy less adaptive. A useful way
to perceive this is through the concept of subsistence risk, defined as the probability
of resource loss and involving the cost of failure and the availability of alternative
sources of food. A comprehensive discussion of this concept has been provided
by Bamforth and Bleed (1997) who, on the basis of data accumulated by Oswalt
(1976), have shown worldwide patterns in resource availability and probability of
risk, and have derived a heuristic model of failure costs.
Concepts of risk and stress on a finite resource base have been employed to
explain changes in stone tools. Proceeding from diet-breadth models derived from
optimal 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-line
resource, red deer, was apparently depleted and replaced by a greater variety of
less desirable resources, including fish. Stress was reflected in the lithic data by a
greater standardization of weaponry, that is, a switch, in the Late Mesolithic, to a
uniform, geometric-based microlithic technology manufactured by the microburin
technique. This strategy may have constituted an attempt to produce more reliable,
possibly composite, tools of which these microliths formed a part. This trend was
accompanied by an increase in the number of different tool functions and the
greater utilization of debris, as people were forced to exploit a greater variety of
species to survive (Vierra, 1995).
McDonald (1991) applied Binfordian models of embedded procurement and
carrying costs to a series of early Holocene sites at the Dakhleh Oasis, Egypt. She
was able to distinguish three types of sites based on diversity of tool types, range
of lithic raw materials, and several other factors. Radiocarbon dates associated
with a dry phase suggest that aggregation in this case was a response to worsening
climatic 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 more
mobile residence strategies, indicated by higher biface:core ratios, frequencies of
formal (bifacial) tools, and prepared striking platforms. Her thesis was criticized
by Sullivan (1994), who proposed factors other than mobility to account for the
changes 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
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 costly
but reliable, parallel-flaked, Angostura points with collector strategies. Amick and
Carr (1996) also revealed this relationship among Paleoindian groups of the North
American Southeast. By the Woodland period, Indians in this region were stockpil-
ing lithic raw material and using expedient technologies containing fewer bifaces
than 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 that
correlates in a general way with increased sedentism and the development of plant
domestication. Although trends in most prehistoric activities appear to have been
stochastic in nature, usewear traces attributed to hafting increased throughout
the Holocene period, whereas those from manual prehension decreased. These
trends are also associated with the development of sedentary lifestyles and plant
domestication. Because an increasingly greater proportion of an individual tool
was committed to the haft proper, the average frequency of functional units per
tool and of different activities and worked materials per tool declined during this
period (Odell, 1994b). The need for implements maintained through sharpening
and reshaping also declined over time, though a slight increase in average uti-
lized polar coordinates per tool suggests that people used their implements more
intensely (Odell, 1994a, 1996b).

Distinguishing Subsistence Strategies

Specific subsistence strategies are not well understood for remote periods of
prehistory. From the earliest hominids through the Middle Paleolithic, it is often
unclear 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 coast
has provided some enlightenment on this situation as it pertains to the Middle
Paleolithic. Most of the interpretations of scavenging or hunting have been made
on faunal indicators, for example, scavenging being distinguished on the basis
of large proportion of head parts, small proportion of limb bones, and relatively
incomplete 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 exploitation
of blanks, and frequent nonlocal raw material. Hunting assemblages exhibit more
efficient parallel cores, smaller and narrower flakes, a relatively large quantity of
flakes taken off each core, less tool modification, and dominance of local raw ma-
terial. It is apparent that scavenging populations moved more frequently, had larger
territories, and maintained more widely transported toolkits. Hunting populations
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 for


making the technology work. If those resources are not readily available, procure-
ment becomes a scheduling problem. The severity of the problem is related to the
nature and proximity of the resource and methods of extraction. Human groups
have several ways to procure resources, by sending out task forces, long-distance
trading, exchange between neighbors, or embedding procurement in a regularly
scheduled foray such as, for huntergatherers, a seasonal round.
In considering mobility and the procurement of exotic raw materials,
MacDonald (1999) has supplemented the foregoing considerations by stressing
the importance of mating relations. Having found that individuals living in lower
density population aggregates travel farther to locate mates than those in more
densely populated aggregates (MacDonald and Hewlett, 1999), he applied these
principles to Folsom groups of the Northern Plains. Low percentages of exotic
materials on these sites were more supportive of long-distance movement of in-
dividuals than of entire social groups, a specification that allowed him to provide
mean 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 through
long-distance trade. Noting a negative correlation between distance from source
and amount of cortex present on the artifacts, he deduced that the procurement of
obsidian among huntergatherer groups was an embedded strategy. Analyzing ob-
sidian from more recent Classic Hohokam sites through X-ray diffraction, Mitchell
and Shackley (1995) came to a similar conclusion. They found an abundance of
obsidian on sites very close to the obsidian sources and a rapid fall-off rate for sites
further 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 in
the St. Lawrence Basin of Canada, found Gaspe silica on Paleoindian sites 800 km
upstream of its source, suggesting either long-distance trade or very high mobility.
Seemans (Seeman, 1994) procurement problem was different. Working on
the Nobles Pond Paleoindian camp, he wanted to know whether the materials in
11 clusters accumulated at one time or sequentially. Despite the availability of
local lithic materials, the dominant ones at the site came from two distant quarries.
He reasoned that, if disparate groups practiced embedded procurement, then the
clusters should be characterized by different raw material spectra, as these groups
would have gone through different areas. But if groups that kept returning to the
site practiced a more systematic strategy, then the clusters should possess similar
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 Pond
supports the second scenario, that is, people across the site were related in some
way and possibly contemporaneous.
For most continental regions, acquisition of raw materials from distant sources
has commonly been attributed to substantial residential mobility. This assump-
tion dominated Andrefskys (Andrefsky, 1995) interpretations of survey data from
the Lower Snake River in Washington. Instead of the common assumption that
people in this region became more sedentary through time, he found that recent
groups employed relatively large quantities of nonlocal raw material, suggesting
that they were quite residentially mobile. These rarer, nonlocal substances were
more likely to be made into formal tools, whereas the more common, local ones
were more likely to be employed as expedient tools (Andrefsky, 1994). In this way
the availability of different raw materials has an important impact on a peoples
technological organization.
Using similar principles, Feblot-Augustins (1993) compared late Middle
Paleolithic assemblages from Acquitaine, France, with contemporary assemblages
from central Europe. In both cases, most raw material procurement was stable and
local (within 5 km), but central European assemblages contained more material that
originated at least 20 km away. Greater long-distance transport probably equates
with greater seasonal mobility in central Europe, with larger subsistence and ex-
ploitation territories. In addition, it supports the exercise of planning and foresight
among Middle Paleolithic peoples.
Raw material procurement and organization of technology are mutually code-
pendent. Although previous examples have illustrated the role of procurement in
technological organization, the latter plays an equally important role in determining
the composition of specific raw materials in a tool kit. This message was conveyed
nicely by Ingbar (1994), who postulated a simple simulation. A hypothetical hu-
man group visits each of the three source areas in turn, replenishing a certain
percentage of its tool kit at each area. The amount of material gathered at a source
and carried away to become part of a future assemblage usually depends largely on
the organizational needs of the group, which is factored into the simulation. That
is, some kinds of tools will be replenished and others maintained, affecting the
composition of the assemblage removed from each source area and assuring that
the percentages of each of the materials used will be different depending on where
in this cycle one samples the assemblage. Even in this simplified representation,
it is obvious that the reconstruction of a groups movements is not easy and is at
least 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 up
on this issue, because some of the pieces in a typical archaeological assemblage
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 important
consideration that our distant archaeological ancestors usually retained only the
modified pieces, which they called tools; unretouched flakes became waste,
as though they were unlikely ever to have been utilized. Today we know better, but
most archaeologists, for good reasons, separate their assemblages into categories
such as tools or type collections, and debris.
A logical ramification of the modification of stone is that such pieces appear
to have been treated differently from those that were not retouched or ground. The
very fact of their modification suggests that more attention was paid to them, either
through sharpening, shaping into a specific tool form, or altering specific portions
in order to hold the tool better. Thus the idea that some tools were curated and
others expedient fell on fertile ground, as archaeologists had secretly known this
all along.
The problem was Binfords (Binford, 1973) use of the term curation, which
was not very specific or exclusive in the first instance and became even less so
with use. This situation has occasioned extensive criticism of archaeologists use
of the term by Hayden (1975) and more recently by Odell (1996a) and Nash
(1996). Odell evaluated Bamforths (Bamforth, 1986) five categories of curation,
using lithic assemblages from the Illinois Valley. He found some inapplicable and
two of them contradictory, and concluded that the term should never be employed
without defining it strictly at the outset. Nash was equally critical. Using excavated
Middle Paleolithic assemblages and finding that the term means different things
to different people, Nash (1996, p. 96) stated,
Given that the concept is now embedded in the literature, a standardized lexicon should be
negotiated 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 degree
of use or utility extracted, that is, the difference between the potential utility the
tool starts with versus the amount left on discard. He also perceives the term to be
applicable 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 an
ochre-stained spot in a bell-shaped pit located in the uplands near the American
Bottom. Given a lack of evidence for burial at this locale, they suggested that these
implements had been produced in advance of use and cached there. Because the
majority were end scrapers and ethnographically women are more firmly associated
with scraping activities than men are, the authors speculated that this was a curated
womens cache.
Close (1996) employed the curational concept of transporting implements
from one locale to another. Prehistoric occupants of sand ripple sites in the Sahara
apparently carried flakes and cores from one ripple to the next in a circuitous
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. Tool
recycling and maintenance also seems to have occurred at the German Middle
Paleolithic Wallertheim D site (Conard and Adler, 1997). Artifacts were not pro-
duced on site, but arrived there in already finished form and were sharpened and
maintained there, suggesting that this was a retooling center. Also working in
the Middle Paleolithic, Kuhn (1992b) noted the transportation of retouched tools
and tool blanks for significant distances; however, he differed with Conard and
Adler in the meaning of this behavior. Instead of conscious preplanning, it might
have been only a temporary hedge against unforeseen difficulties, thus reflecting
opportunistic patterns of land use.

Cultural Complexity

Lithic artifacts have traditionally been associated with more technologically


primitive stages of cultural development. Recently, however, we have started
to understand and appreciate the processes by which stone tools contributed to
movements toward cultural complexity, as well as how they were replaced by
metal. Staffords (Stafford, 1999) study of lithic assemblages spanning the late
Mesolithic/early Neolithic transition in Scandinavia, for example, demonstrated
that the process there was a gradual transformation of indigenous societies rather
than sudden migrations of Neolithic people from elsewhere. Also partly on the
basis of lithic analysis, Stafford offered a redefinition of the Neolithicfrom do-
mesticated plants and animals to the importation of certain status objects such as
polished stone axes, a process that represented a symbolic transformation of north
European culture. Accompanying economic and social complexity were different
ritual and ceremonial practices than those that had existed among noncomplex
societies, 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, Solutrean
foliate points, Magdalenian spear throwers, and the products of knappers at places
such as Etiolles and Corbiac. Given that specialization is a continuous variable,
there has been a lot of discussion about what constitutes true craft specialization
and how to determine it from the archaeological record.
At least moderate craft specialization appears to have existed at the Late
Neolithic/Chalcolithic village of El Malagon, Spain, located close to the La Venta
flint 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
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 specialization
and political control increased through time, reflected in greater quantities of both
local and exotic mined stone. Professional craft specialization also appears to have
existed 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, an
Early Bronze Age pastoral camp in the Negev highlands. There he found not true
craft specialization in bead production, but cottage industries that produced excess
goods for exchange in a higher-level economic system.
Shell bead production, usually involving stone microdrills, is also a common
specialization among coastal peoples. A major project in the Channel Islands off
the California coast has established a considerable trade in this commodity by the
Chumash and their predecessors (Arnold, 1992; Arnold and Munns, 1994). The
exact nature of the bead trade was a little hard to define, but appears to have in-
volved peripherally attached specialists whose product distribution was controlled
by highly ranked persons on the mainland. Likewise, shell bead manufacture on
Motupore Island, Papua New Guinea, constituted an important source of income
throughout the history of that island. There production was at a village level and
did not necessarily imply distinct social ranking (Allen et al., 1997).
Most New World work on the topic of specialization has been accomplished
among the Maya, where conclusions have not always been based on the firmest
data sets. As a case in point, Aoyama (1995) compared areas in and around two
structures at Copan, employing high-magnification usewear techniques based
exclusively on polish indicators for flint, and polish and striations for obsidian. His
conclusion that specialized production of marine shell ornaments occurred in front
of structure 10L-16 contains an array of problems, including very small sample
sizes, activity and worked material totals that do not match, use of a combined
bone/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 region
are supported by stronger arguments. For instance, Clark and Bryant (1991) ana-
lyzed the production of chert projectile points from a Maya site in Chiapas. From
the high error rate and large variability within one point style, they concluded that
this was not a specialized craft at this locale, but a part-time domestic activity
practiced at the household level. Likewise, usewear 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 of
Xochicalco by Hirth (1995) established that this town may have been a regional
supplier of finished obsidian tools, but little evidence existed for large-scale craft
specialization as at Teotihuacan.
But such specialization surely existed in the Maya area, as it did in other
regions. A good candidate is the flint workshop site of Colha, where oval bifaces
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 Preclassic
to the Terminal Classic period (Shafer and Hester, 1991). Another is the Michoacan
area of western Mexico, where Darras (1994) has described a series of obsidian
mines, workshops, and associated settlements. In both cases, the volume of debris
at the workshops suggests substantially more production than would have been
necessary for use within the associated village. Likewise, lapidary production at
the 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 item
through embedding it in normal huntergatherer seasonal rounds (Wright, 1994).
With respect to lithic materials, interpretation involves an intimate knowledge of
the resources in a region. A case in point is Gibsons focus on Poverty Point ex-
change (Gibson, 1994). Here in the Lower Mississippi Valley, a principal source
for toolstone was secondarily deposited local gravels, which are recognizable by
their brown cortexual rind. However, the gravels probably did not provide all of the
predominantly gray varieties of stone that appear in Poverty Point assemblages,
and it is difficult to detect any commodity moving northward in exchange for the
materials that were probably entering the area. These problems affect many regions
and will require considerable effort to resolve.
Several examples of trade or exchange in lithic items have been offered during
the past decade. For example, Bourque (1994) described the ebbs and flows of trade
on the Maritime Peninsula of eastern North America, primarily in exotic stone. Of
particular 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 were
responsible for a high volume of trade during the Late Postclassic period. Despite
their existence under the strong Aztec state, trade developed through market sys-
tems and was never subjected to coercive political control (Smith, 1990). Across the
ocean, Takacs-Biro (1997) has provided a comprehensive analysis of the distri-
bution of lithic trade items in the Late Neolithic of Hungary, including northern
varieties of flint that became important in the later part of this period.
Several items of stone moved through these exchange systems: One common
product was obsidian, and that too not just in Central America. Obsidian also
constituted a key component of exchange systems in the American Northwest,
peaking in intensity at different times in the Plateau (Galm, 1994) and in British
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 established
during the Middle Woodland period featuring Jacks Reef corner-notched points.
Strauss (1992) speculated that a prehistoric coastal trade network involved raw
jasper, 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 (early
pottery) times (Sheppard, 1993, 1996; White, 1996). Despite the distances, stone
for tools was not transported from the nearest sources of these materials, nor did
tool users appear to be optimizing material, as discarded cores were not exhausted.
Tool users appear to have responded more to social relationships such as gifting
than to strict utilitarian economics. In another example, the green Pachuca obsidian
from Teotihuacan has been found on Maya sites in ritual areas and was made into
specialized tool forms, suggesting that its overall function was more symbolic or
ceremonial 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 of
Colha and its client polities. Two standardized toolsthe tranchet bit axe and oval
bifacewere produced in vast quantities at Colha, which was situated on large
outcrops of high-quality chert nodules. These tools were shipped to consumer sites
such as Cuello, Pulltrouser Swamp, and Cerros, at which archaeologists have found
numerous sharpening flakes, but few cores, of Colha chert (Dockall, 1994; Dockall
and Shafer, 1993; McSwain, 1991). Distribution of Colha chert occurred generally
northwest of the producer site in the Late Preclassic period. By the Late Classic, it
occurred in all directions from Colha, as chert was shipped freely throughout the
river systems (Santone, 1997).
Igneous and metamorphic rock were also objects of long-distance trade. This
was certainly true of Neolithic western Europe, where axes, adzes, and chisels
constituted important constituents of the tool kit. Working in the western Alpine
region, 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 traded
from east to west across the Alps. Pre-Dynastic grinding stones from at least
one site in Upper Egypt were manufactured from volcanic rocks from perhaps as
much as 150 km away, attesting to the existence of a widespread material exchange
network during this period (Mahmoud and Bard, 1993). And X-ray fluorescence
analyses of Late Bronze Age-to-Roman period millstones and querns from Cyprus
showed that materials for their manufacture often originated in distant places, such
as the Levant or the Aegean Islands (Williams-Thorpe et al., 1991).
Finally, the persistence of cultural norms in the face of changing exchange
relationships has been documented by Johnson and Hayes (1995), who compared
two noncontemporaneous Middle Woodland sites in the northern Yazoo drainage,
Mississippi. Snyders-style points and blade techniques were introduced into the
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 and
dominated by Illinois cherts such as Cobden and Burlington. These styles and
techniques were continued in the later Oak Grove assemblage, despite the fact that
most of the artifacts were made of Arkansas novaculite. In other words, despite
a change in procurement or trade relationships from Illinois to Arkansas, artifact
styles 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 distribution
of lithic materials strictly controlled by a central polity or social class, or did
they remain under the authority of smaller, household- or kinship-based units?
In Mesopotamia, Pope and Pollock (1995) compared the Uruk Mound with two
other tells of the Early Dynastic period. All sites received tool stone in the shape
of already preformed cores, from which toolmakers fashioned blades and other
items. The authors concluded that, at all three sites, production and use were not
limited to a small number of households, but were widespread within the villages.
They found little evidence for either centralization or administrative control of the
resource.
A few recent studies have uncovered the possibility of sociopolitical control
among North American prehistoric groups. For instance, Bayman (1995) noted
that obsidian production and consumption in the Marana Hohokam community in
the American Southwest were concentrated on one large platform mound, whereas
other mounds on the site contained very little obsidian. This evidence demonstrated
centralized redistribution of obsidian and suggested the existence of an emerging
power elite. Similarly, in the Plum Bayou culture of central Arkansas, Nassaney
(1996) detected concentrations of quartz crystal at Coy Mound and in an area east
of the plaza at the Toltec site. Quartz crystal is most likely a prestige item, and
these concentrations may indicate the presence of social ranking.
In no instance, however, have I run across evidence, uncovered within the past
decade, of strong sociopolitical control over a lithic resource in the New World, a
point exemplified by Olivella shell bead production in Californias Channel Islands
(Arnold and Munns, 1994). The authors isolated several production centers on
the islands but none on the mainland, though it is apparent that political power
was centered on the mainland. The most likely scenario is that mainland elites
controlled bead distribution, but left production in the hands of loosely attached
specialists on the islands.
Even at a highly specialized lithic workshop such as Colha in the Maya
lowlands, the most likely authority was lineage based rather than elite control-
led (Shafer and Hester, 1991). In addition, the lack of chert workshops at large
Maya centers such as Tikal and Yaxha suggests that production of chert tools was
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 redistribution
from major centers did occur, for the greater prevalence of early reduction stage
debitage at households located away from major centers is consistent with closer
households 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 of
northeastern Mississippi. The discovery of several Benton period caches of finely
made bifaces and of discrete biface production loci on Benton sites suggests a high
level of production, probably for exchange (Sassaman, 1994b). Johnsons point
was, if a Benton period site were found in Belize and attributed to the Classic
period, it would be presented as evidence for craft specialization controlled by an
elite, whereas in reality, whatever specialization there was, was probably loosely
controlled 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, with
surplus raw material going to the general population along lines of kinship.
Studies cited above demonstrate a general lack of sociopolitical control over
lithic resources among prehistoric occupants of both the Old and New Worlds. This
impression has been strengthened by Pope (1994), who compared evidence from
two very different geographical and historical contexts: the Early Dynastic Uruk
people in Mesopotamia and the Emergent Mississippian Black Warrior Valley
culture of the North American Southeast. In the latter region specialized bead
making occurred, but the remains of this activity were widely dispersed within sites,
suggesting that production was organized by household or kin group. Similarly
at Uruk Mound, chipped stone tool production was not a centralized activity, and
neither techniques nor raw materials were restricted once they were imported to
the site.

Material Replacement

Eventually stone tools were replaced by metal in most areas of the world.
Conventional wisdom suggests that this replacement occurred rapidly throughout
all tool types, but until this decade, little research had been done on the processes
of the replacement. Several studies demonstrate that the process was complex and
involved 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 formality
had declined. This process continued in the Middle Bronze Age, and knappers in-
creasingly employed local raw materials and simple core-flake techniques. Metals
became increasingly important in the context of trade, as perceptions of the use of
materials changed and sociopolitical forces created new demands and desires.
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 been
produced by Rosen (1996, 1997b), whose interest spans the Chalcolithic, Bronze,
and Iron Ages. In this region the process of replacement took 3000 years and
proceeded in stages, first involving stone arrowheads, then burins, then axes and
drills, then ad hoc tools, and finally sickle blades. Reasons for replacement by metal
were different in each case. For example, established trade routes were important in
promoting copper axes, whereas the relative cheapness of flint and well-established
specialization in this material ensured that stone sickle blades would continue to
be used long after other lithic tool types had disappeared. Clearly, the replacement
of stone by metal was not a linear process, but a highly complex one involving
considerably more than simple utility.
Similar complexity accompanied replacement in regions outside Europe and
the Near East. Among the Pawnee of the North American Plains, for instance,
metal knives quickly replaced stone ones, but stone scrapers and abrading tools
continued in use longer because they were less likely to have superior metal coun-
terparts. As their utilitarian desirability declined, stone tools eventually assumed
a sacred character, as in implements for curative bleeding, etc. (Hudson, 1993).
Likewise, among Aborigines of Australias northwestern Northern Territory, stone
was 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 large
points 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 impressive
body of literature in recent years, partly because archaeologists are more aware,
through postprocessualist and other currents of thought, of the important role that
social considerations play in technology (Dobres and Hoffman, 1994). Practices of
Australian Aborigines have attracted study because, among many of these groups,
stone is regarded in mythical terms. That is, myths related to stone and quarries
have emerged over the years, as artifacts have been embedded with social and reli-
gious value. This point was presented effectively in Robert Patons (Paton, 1994)
study from northern Australia of leilira blades of cobble quartzite, which were
made for trade and whose use was accompanied by a strict code of conduct. But
when inspecting habitation sites at varying distances from the quarry, he found
nary a blade, despite evidence of quartzite debris from their manufacture. Without
knowledge of Aboriginal behavior generated from elsewhere, one would never
know what had been manufactured at these sites. McBryde (1997) also made this
point for the area around Lake Eyre in east-central Australia, a region of exten-
sive Aboriginal connecting trails (dreaming tracks). She employed petrographic
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 then
able to associate these material distributions to exchange systems portrayed in oral
tradition.
Stone was also imbued with mythical characteristics among some North
American Indian tribes such as the Iroquois, for whom Flint was one of the
creator twins and Stone Giants figured prominently in another set of legends
(Moulton and Abler, 1991). Characteristics of a similar nature can be read into
certain 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 its
utilitarian purposes.
Further back in North American prehistory, there is evidence that a few widely
traded objects possessed symbolic significance. One of these was the turkey-tail
point, around which a Late Archaic/Early Woodland trading network was estab-
lished in the Great Lakes region. Usually made of Wyandotte or Dongola chert
and often found in association with copper beads, this distinctive object may have
constituted a form of trade regulator, like the shell necklaces and armbands in the
Kula Ring of the Trobriand Islands. The fact that some of the points were sharp-
ened supports their occasional employment as utilitarian objects, and the existence
of caches containing well-matched points suggests that they were made by part-
time specialists (Krakker, 1997). Another such item, produced for exchange and
ceremonial use in the American Southeast, may have been the Early Archaic Kirk
corner-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 Southeast
at about 4000 bp appears to have represented a risk abatement strategy among at
least some of the groups at this time. Sassaman (1998) has theorized that marginal
uplands peoples may have produced and traded bannerstones to more dominant
pottery-making bands in order to cement alliances. Another possible ritual item,
traded in the Dalton (late Paleoindian) culture in the central Mississippi Valley, was
the well-made Sloan biface. Possibly associated with male hunting ritual, it may
also have been used to strengthen ties with allied peoples (Walthall and Koldehoff,
1998).
Trading relations constitute important evidence for postulating social or ritual
significance for certain artifact forms, but other types of evidence have also been
used. For example, at the early Paleoindian Parkhill site in Ontario, Ellis (1994)
discovered seven miniature artifactsmostly tiny points on channel flakes, but
also one scraper. He invoked their discard location, clustered close together in two
areas of the site, to establish that these could be neither utilitarian implements nor
childrens toys. From this evidence he postulated that they were ideotechnic items
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 in


areas 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 tools
for breakage, usewear, and residues, she established that these techniques were
viable and compiled a list of characteristics that were common among ceremonial
implements. She conducted a similar analysis of a ritually specialized tool assem-
blage from the sacred cenote at Postclassic Chichen Itza on the Yucatan Peninsula
(Sievert, 1992), in which she rendered a functional profile of what an assemblage
of ceremonial stone tools might look like. Interestingly, she detected considerable
impact damage from projectile use and depicted a specialized ritual signature that
probably 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 stone
tool technique in the development of symboling and language. However, Thomas
Wynn (1991, 1993) has formulated a strong argument that prehistoric tool behavior
is learned in substantially different ways than languages. Wynn is a Chomskyan
who argues for linguistic learning through innate cognitive structuresa proce-
dure that is radically different from learning stone tool techniques through rote
serial memorization in an apprenticeship format. He concludes that the learning of
prehistoric lithic techniques is unlikely to tell us very much about the origin and
evolution of grammar. He does postulate, however, that the Acheulean stage of
biface production (though not the Olduwan) requires a constellation of responses
that is a cognitive stage above chimp behavior.
Although tool behavior may not be very helpful in studying the origin of lan-
guage, it can possibly be used to inform on human cognitive abilities for symboling
behavior (and therefore, language acquisition). Several scholars have pointed to the
Middle/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 content
by the Upper Paleolithic period. Likewise, Davidson and Noble (1993) denigrated
the idea that early hominid tool makers possessed the cognitive abilities to form a
template of a specific tool form and then carry out a plan necessary to manufacture
it, calling this the finished artifact fallacy. In their view, language arises from
the discovery that meaning is conveyed by signs (1993, p. 382; italics in original),
and the earliest indications of such heightened consciousness are the colonization
of Australia and the creation of sculptures and bas reliefs in Europeboth about
the time of the Middle/Upper Paleolithic transition.
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 engraved
lines, a large number of beads and pendants in assemblages from Traditions firmly
rooted in the Middle Paleolithic (1995, p. 614), and bones with incised lines from
sites such as Bilzingsleben and Stranska Skala. About this controversy, one thing
is certain: the evidence for any position is so tenuous that there is unlikely to be a
clear winner any time soon.

CONCEPTUAL APPROACHES

The discussion to this point has emphasized the results of archaeological


endeavor; blueprints used to conceptualize and generate these results have been
mentioned in passing, but not dwelt upon. Common approaches chronicled above
include experimental, ethnoarchaeological, and ecological, but these are not the
only conceptual models employed by lithic analysts. Here I briefly discuss four
constructs that have helped analysts put their data in perspective. They can be
categorized 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 ecological


or social influencesthese are the influences that the analyst thinks are worth
bringing to the fore within the contexts being considered. Stone tool analysis
has traditionally been such a male-dominated domain, however, that an emphasis
on gender is an acknowledgment that serious bias of mindset has warped our
perception of the archaeological record. Because the conventional wisdom of who
made and used stone tools is based on such weak empirical and ethnographic
support, it is high time that we question these pillars of knowledge.
The dominant male role in making and using stone tools was duly challenged
by Gero (1991). Establishing through ethnography that women made and used
stone tools in some societies, she postulated that, in prehistory, women were likely
to have employed implements in household situations within habitation areas,
where the stone from which tools were manufactured was predominantly local
and flake tools were utilized expediently. She tested these ideas on assemblages
from the Huaricoto settlement in highland Peru, a site that changed from being
a ritual center to a residential village. She found that, as the site changed func-
tion, biface production and amount of retouch decreased, whereas the quantity
of expedient tools and amount of cortex per tool increased. Projectile points and
bifaces, over-represented in earlier strata for ceremonial situations in which males
probably dominated, were swamped numerically by expedient flake tools of local
raw materials in later strata for female-dominated household situations.
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 tools
was 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 technologies
coincided with greater reliance on pottery, traditionally assumed to be a female
activity. Like Gero, Sassaman interpreted the burgeoning quantities of expedient
tools through time as an increase in the contribution of women to stone tool pro-
curement, reduction, and use at domestic sites. And also like Geros, Sassamans
arguments are difficult to prove definitively, but they are logically constructed and
they alter our perceptions of the meaning of the archaeological record.
My perambulations through the literature appear to have missed more recent
gender studies specifically related to stone tools. However, a few studies do mention
gender in passing, among which is the aforementioned discovery of a concentration
of Dalton scrapers in the American Bottom, which Walthall and Holley (1997)
speculate was a curated womens cache.

Design Theory

Design theory is a way of approaching data that emphasizes various con-


straints in solving given problems by technological means (Hayden et al., 1996,
p. 10). That is, tool design is seen as a strategy for overcoming difficulties in the
extraction of resources or in other activities; similarities in design suggest simi-
larities in adaptive response to these difficulties. Design theory developed within
the fields of engineering, industry, and architecture (Jones, 1970; Pye, 1964); only
recently has it been applied to archaeological situations. Within the field of lithic
analysis, its application was stimulated by Bleeds (Bleed, 1986) treatise on main-
tainable and reliable tools, followed by Nelsons (Nelson, 1991) reformulation.
Bousman (1993) has discussed these concepts as they relate to huntergatherer
considerations such as scarcity, risk, repair strategies, and curation.
At first, Bleeds conceptualization seemed to be a useful way to encapsulate
behaviorally essential qualities of entire assemblages. Then reality set in: (1) all
tools are both maintainable and reliable; (2) expressing these qualities precisely
is very difficult; and (3) quantifying them for comparison is even more difficult.
As a consequence, these concepts have been neither extensively developed nor
applied.
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 contexts
and 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 assemblage
in interior British Columbia, Hayden et al. (1996) delineated several technolog-
ical strategies that ultimately enabled them to better understand the assemblage
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 the


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

Cultural Technology

Cultural technology is the way in which a cultural group articulates its


technology to other facets of its structure, or how technology fits in to its cul-
tural persona. In reconstructing this parameter, several archaeologists in the Old
World, particularly France, have turned to a holistic approach known as the chaine
operatoire. As originally perceived in its homeland, the term connoted gesture,
translated as material action. Almost mystical in its ramifications, this concept
has 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 material
had to go through. It is a chronological segmentation of the actions and mental processes
required in the manufacture of an artifact and in its maintenance into the technical system
of a prehistoric group. The initial stage of the chain is raw material procurement, and the
final stage is the discard of the artifact.

Thus the chaine operatoire seeks to reveal a dynamic system from its inception to
its placement in archaeological context.
For some reason, the chaine operatoire has been equated by many researchers
with reduction sequence (e.g., Grimaldi, 1998, p. 19). Certainly, the concept
encompasses lithic reduction sequences but it is by no means limited to this aspect
of behavior. De Bie (1998, p. 91) more accurately describes the chaine operatoire as

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 to
reconstruct (partly by reproduction) and to explain the behavioural processes responsible
for the formation of the lithic record.
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 often
directed toward specific elements of the puzzle and are less consciously inclusive
than those who actually adhere to the chaine, that is, those who regard it as more
than just figuring out a reduction sequence. In fact, the chaine operatoire comes
closest to Schiffers behavioral archaeology, which considers objects within their
systemic contexts from procurement to discard (and beyond). Interestingly, de-
spite efforts to bring it back, behavioral archaeology is somewhat out of vogue in
America these days, eclipsed by research into broad behavioral concepts such as
the organization of mobility and technology.
Geneste and Maury (1997) apply the chaine operatoire in their research
on 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 constraintsin short,
all the factors that would have impinged on a Solutrean person making and using
this kind of technology. This technological immersion frequently yields a depth
of understanding more comprehensive than the piecemeal style of experimen-
tation conducted by most experimenters. It is this immersion that distinguishes
the chaine operatoire from other systems of analysis. Unfortunately, many of the
scholars who have mouthed their affection for this system have done nothing more
than give us warmed-over lithic reduction sequences, with occasional lip service to
procurement.

Selectionist Models

Theories of transmission of cultural information, many involving some form


of Darwinian evolution, are currently circulating like viruses in cyberspace. The
attraction is strong, as the humanities and social sciences lack a universal the-
ory of behavior; the formulation of a theory of cultural transmission would have
tremendous influence on every aspect of human endeavor. I do not pretend to be
conversant with the range of models or the nuances of difference among them,
but several have conspicuously employed lithic data in their formulations and are
therefore included in this review.
The endeavor is difficult. The failed attempt by Barton (1997) to depict style
in lithic artifacts by examining stochastic processes in an evolutionary framework
has already been noted. Shott (1997) also attempted to test theories of cultural
transmission on projectile points from the Range site in the American Bottom.
Only one variableprojectile lengthfit the expected pattern, and this result may
have been caused by sharpening the pieces.
Other researchers have claimed to have demonstrated some sort of cultural
transmission, though their success has been spotty. Following Parry and Kellys
(Parry and Kelly, 1987) seminal synthesis associating sedentism with expedient
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, Abbott et al. (1996) inspected the change from biface to flake tech-
nologies in five regions. All manifested the same directionality, thus rendering
the operation of stochastic processes unlikely. Because evolutionary explanations
must reside outside the human system (i.e., they must be environmental), the
authors rejected the decrease in residential mobility as a causal factor, that is,
sedentism was a product, not a cause. So in searching for causes, the authors
settled on maize agriculture as the agent of change. By this scenario, seden-
tism and technological change were both linked, by sorting, to selective forces
favoring 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), while
processes of sedentism remained within that system? Second, where is the con-
firming evidence that maize agriculture is related in any way to an expedient flake
technology? And third, in order to be causal, maize agriculture had to precede
the shift to a flake technology, but in North America, at least, the opposite is
true.
Bettinger and Eerkens (1997, 1999) have provided more believable tests of
Neo-Darwinist precepts. They noted that cultural transmission theory predicts that
socially 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 individuals
involved in the transmission. Testing this theory on a collection of Great Basin
projectile points, they gauged complexity by measuring metric dispersion through
the coefficient of variation statistic. Indeed, complex shapes showed less variabil-
ity than simple ones, and arrowheads (representative of a more complex delivery
system) showed less variability than dart points. They also applied the theory to
the spread of bow-and-arrow technology in the Great Basin. Using weight and
basal width measurements, they contrasted learning through close social contact
(indirect bias) with learning through copying prevailing social models (guided
variation). Reasoning that the former would result in more direct copying and less
experimentation, and therefore more homogeneous traits in manufactured objects,
significant differences in homogeneity of attributes of the same projectile types
should have been caused by differences in cultural transmission. Indeed, the au-
thors discovered significantly greater attribute homogeneity in arrowheads from
central Nevada than in those from eastern California. Studies such as these lend a
degree of hope that the pursuit of cultural transmission theory will open up new
avenues of insight in the future.

PERSPECTIVES

A substantial amount of effort has been spent in the areas of functional analysis
and behavioral approaches during the past decade. Less has been directed to lithic
classification. Here I summarize the points made earlier.
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 at


the forefront of archaeological endeavor in the past few years. The most exhaustive
recent work is more a philosophical exegesis and practical guide to clear classifica-
tory principles than an innovative breakthrough. A subset of scholars, employing
factor analysis and clustering algorithms, have branched out from common prac-
tice with taxonomic methods that are designed to uncover quantitative structure in
formal 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 not
been popular, issues involving the prehistoric behavioral phenomena that produced
certain 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 the
end result of a process of use and sharpening throughout the complex use-life of a
tool? This issue has been debated in two arenas: projectile points in the American
Great Basin and Middle Paleolithic scrapers in Europe and the Levant. In neither
case has a definitive verdict been reached, though in the American case it does
not appear that sharpening has significantly changed the archaeological typing of
the points involved. It is probable that both forces influenced the final forms of
prehistoric artifacts and did so on a case-by-case basis.
Ethnically distinctive style has been extremely difficult to characterize among
lithic artifacts, but some headway has been made through classifying projectile
points within different levels of inclusion. It appears increasingly doubtful that
advances in stylistic studies will be made using Bordesian typologies.

UseWear Analysis

Usewear analysis has become much less contentious these days (and a lot
more fun) because practitioners have concluded that they must use all avail-
able wear traces in their assessments and that different techniques are appro-
priate for different kinds of questions. More people are recognizing prehensile
wear these days, a perceptual breakthrough that was long in coming. Perhaps
the 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, and
the 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 no
reservation), then the discrimination of worked material from polish formation
should be difficult or impossible. Such statements, of course, cause some concern
among those who derive their livelihood from peering exclusively at this type of
evidence.
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 of


edge angles to usewear has shown too much overlap for edge angle data to be
very useful. In addition, experiments have shown that damage on stone tools from
trampling can be confused with both usewear and retouch. And the quantification
of usewear variables has progressed somewhat, notably with Tomenchuks para-
metric method. However, with this technique it still takes a long time to peruse a
single piece and, to my knowledge, nobody but Tomenchuk is practicing it. Some
work in quantifying experimental variables has begun in the Netherlands.
Advances in usewear analysis include the aforementioned development of
parametric methods, employed with low-magnification microscopes. Higher mag-
nification analyses have been aided by the improvement of epoxy casts. These
feature 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 not
fit on the stage of some microscopes. In addition, the atomic force microscope has
considerable potential for usewear analysis through its ability to distinguish and
map out differences in topography. Finally, the development of expert systems of
analysis, combining low- and high-magnification equipment, has proven to be an
accurate method for analyzing a lithic assemblage.

Residue Analysis

In contrast with usewear analysis, which bounced back after some difficult
times in the 1980s, the prognosis for certain types of residue studies is grim. The
problem with blood residue analysis is not with its ability to distinguish recently
utilized laboratory specimens, which it seems to do quite well. The problem is with
archaeological specimens, some of which it is capable of distinguishing to at least
a 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 have
survived and which have deteriorated. Without this information, we have no idea
how representative of the archaeological sample a particular assay is. And I am
saying all this assuming that there are no false negatives or false positives, though
we know that both have been detected by analysts. If these problems cannot be
resolved, then support for this type of analysis will, and should, be curtailed.
Superficially, plant residue analysis appears to be in better shape, but with a
curious twist. The principal difference between plant and blood residue analysis is
that the former often relies on the identification of specific particles adhering to the
tool surface, such as starch grains, resins, and phytoliths. If one can be certain that
differential preservation of these particles has not occurred, then the sample can be
assumed to be representative, the results can be used to characterize the assemblage
from which they were taken, and we can go beyond a simple presenceabsence level
of interpretation. But to my knowledge, one cannot be certain because research has
not yet determined the preservability of various starch grains and phytoliths. In the
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 immunoelectrophoresis


were used in the analysis of plant residues, the technique fared no better than
similar 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 Australia
on the movement of starch grains in sediments has demonstrated that most grains
move little from their original locus of deposition; thus, potential postdepositional
problems may be minimal.

Technological Organization

Models of technological organization, which have employed concepts of mo-


bility organization, maintainability or reliability of tools, curation, expedience, and
so forth, have been popular throughout the 1990s. Some of these terms, however,
subsume more than one meaning and have often been loosely applied. When using
terms like curation and expediency, it has been recommended that the author
define the term tightly before using it.
Testing these models and concepts has frequently been accomplished through
lithic data. For instance, arguments of huntergatherer mobility in the southern
European Paleolithic have been framed with respect to the size and sharpening
potential of implements in the lithic tool kit. Later in the Mesolithic, changes
in hunting strategy were accompanied by changes in microlith-tipped weaponry.
And in North America, decreases through time in bifacial industries correspond
with societal movements toward increased sedentism and reliance on domesticated
plants. Although most functional trends are stochastic because they depict a com-
mon suite of activities pursued in all periods, certain ones do correlate with trends
toward sedentism and domestication, and with increased hafting of implements
throughout 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 that
the amount of a particular material in an assemblage may be dependent on where
in that seasonal round the occupation occurred. However, distinguishing whether
a material arrived on a site through embedded procurement or trade can be very
difficult.

Cultural Complexity

Lithic studies in the past decade have not been constrained by exclusive
application to huntergatherer societies; they have also been applied to issues of
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 tools
were replaced by metal. This did not happen all at once, nor were all tool types
replaced at the same time. The process was complex and involved the relative
usefulness of either material for the intended tasks. For example, metal made
substantially more efficient cutting and chopping implements, so stone knives
and axes were replaced quickly in most contact situations. On the other hand,
stone sickle blades were almost as efficient as their metal counterparts and were
considerably cheaper, so they remained in the Near Eastern tool kit for a long time.
In several societies, stone was replaced for utilitarian implements but continued
as 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 been
investigated through several studies. It is certain that some specialization in stone
tool 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 large
industrialized workshops are involved, a higher level of specialization is implied.
Long-distance trade has been detected in several instances though, as noted
above, this can be difficult to distinguish from embedded procurement. A variety
of lithic materials, including flint, obsidian, igneous stone, catlinite, and others,
were traded prehistorically in different parts of the world. Both materials and
specific objectsfor example, sandstone along the dreaming tracks of Australia,
obsidian in Oceania and Mesoamerica, turkey-tail points in midcontinental North
Americamay have been imbued with symbolic significance or involved in gifting
relations. In many parts of the world, stone in various forms took on symbolic or
mythic meanings even as it was being replaced by metal objects.
In no recently reported case of lithic craft specialization or trade has any
scholar of prehistoric cultures detected strong sociopolitical control of the resource.
Even at a specialized workshop village like Colha in Belize, tool production was
probably household or lineage based. Emerging elites, whose power is evident in
other ways at some of these sites, were more likely to control objects such as ritual
items.

Conceptual Approaches

Lithic analysis has traditionally been a male-dominated domain, so it is no


surprise that stone tools have been almost universally considered to have been
made and used by prehistoric men. As the gender composition of lithic analysts
changes, so will our perceptions of who did what in prehistory. These changes
have 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, mens tools may have been confined to the realms of hunting
and ritual.
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, three
of which have been discussed in this review. Design theory, in considering how the
features of a particular tool type might best be employed, is able to demonstrate
structural advantages and constraints that may assist in conceptualizing the general
form and purpose of lithic assemblages. Likewise, the chaine operatoire, aimed to
recreate the entire system in which a technology existed, provides a useful structure
for organizing research. By immersing oneself in a larger trajectory of procurement
through discard, the archaeologist can derive a more complete understanding of
how the parts of that system articulate with one another.
Finally, of the Darwinian models currently being plied, cultural transmission
theory appears to have more promise than most. At least, in its application to
lithic data, its premises have been supporteda far cry from some of the other
evolutionary formulations making the rounds, which, as far as I can tell, have
remained exclusively in the theoretical realm. This theory needs more testing, but
this time we may see a payoff.

Final Comments

Like most other subdisciplines, lithic analysis has developed sporadically, a


result of a combination of communal trends and individual interests. There exists
a healthy amount of information flow, and a prevailing skepticism assures that new
ideas will not be accepted at face value. These are good signs, which render it
likely that the field will continue to change and grow significantly through the next
decade.

ACKNOWLEDGMENTS

These two review papers have benefited greatly from comments made on
earlier 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 two
reviews, 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 flake
technology. In Maschner, H. (ed.), Darwinian Archaeologies, Plenum, New York, pp. 3342.
Adams, W. Y., and Adams, E. W. (1991). Archaeological Typology and Practical Reality, Cambridge
University Press, Cambridge.
Ahler, S. P. (1998). Early and Middle Archaic settlement systems in the Modoc locality, southwest
Illinois. Illinois Archaeology 10: 1109.
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 exchange
in the archaeological record: The case of shell bead manufacture on Motupore Island, Papua.
Archaeology in Oceania 32: 1338.
Amick, D. S., and Carr, P. J. (1996). Changing strategies of lithic technological organization. In
Sassaman, K., and Anderson, D. (eds.), Archaeology of the Mid-Holocene Southeast, University
Press of Florida, Gainesville, pp. 4156.
Anderson-Gerfaud, P. (1980). A testimony of prehistoric tasks: Diagnostic residues on stone tool
working edges. World Archaeology 12: 181194.
Anderson, P., Astruc, L., Vargiolu, R., and Zahouani, H. (1998). Contribution of quantitative analysis
of surface states to a multi-method approach for characterising plant-processing traces on flint
tools with gloss. In Functional Analysis of Lithic Artefacts: Current State of the Research, XIII
International Congress of Prehistoric and Protohistoric Sciences, Workshops, Tome II, ABACO
Edizioni, Forli, Italy, pp. 11511160.
Andrefsky, W., Jr. (1994). Raw-material availability and the organization of technology. American
Antiquity 59: 2134.
Andrefsky, W., Jr. (1995). Cascade phase lithic technology: An example from the Lower Snake River.
North American Archaeologist 16: 95115.
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 Antiquity 6: 129144.
Arnold, J. E. (1992). Complex huntergathererfishers of prehistoric California: Chiefs, specialists,
and maritime adaptations of the Channel Islands. American Antiquity 57: 6084.
Arnold, J. E., and Munns, A. (1994). Independent or attached specialization: The origin of shell bead
production in California. Journal of Field Archaeology 21: 473489.
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. 109125.
Ballenger, J. A. M. (1996). The Southern Plains craft lithic cache. Plains Anthropologist 41:
297309.
Bamforth, D. B. (1986). Technological efficiency and tool curation. American Antiquity 51: 3850.
Bamforth, D. B. (1988). Investigating microwear polishes with blind tests: The Institute results in
context. Journal of Archaeological Science 15: 1123.
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. 109139.
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: Evolutionary
Theory and Archaeological Explanation, Archeological Papers of the American Anthropological
Association, No. 7, Washington, DC, pp. 141156.
Barton, H., Torrence, R., and Fullagar, R. (1998). Clues to stone tool function reexamined: Comparing
starch grain frequencies on used and unused obsidian artefacts. Journal of Archaeological Science
25: 12311238.
Bayman, J. M. (1995). Rethinking redistribution in the archaeological record: Obsidian exchange at
the Marana platform mound. Journal of Anthropological Research 51: 3763.
Becker, M., and Wendorf, F. (1993). A microwear study of a late Pleistocene Qadan assemblage from
southern Egypt. Journal of Field Archaeology 20: 389398.
Bednarik, R. G. (1995). Concept-mediated marking in the Lower Paleolithic. Current Anthropology
36: 605616.
Bettinger, R. L., and Eerkens, J. (1997). Evolutionary implications of metrical variation in Great Basin
projectile points. In Barton, C. M., and Clark, G. (eds.), Rediscovering Darwin: Evolutionary
Theory and Archaeological Explanation, Archeological Papers of the American Anthropological
Association, No. 7, Washington, DC, pp. 177191.
Bettinger, R. L., and Eerkens, J. (1999). Point typologies, cultural transmission, and the spread of
bow-and-arrow technology in the prehistoric Great Basin. American Antiquity 64: 243263.
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., OConnell, J. F., and Thomas, D. H. (1991). Projectile points as time markers in the
Great Basin. American Anthropologist 93: 166172.
Bienenfeld, P. (1995). Duplicating archaeological microwear polishes with epoxy casts. Lithic Tech-
nology 20: 2939.
Binford, L. R. (1973). Interassemblage variabilitythe Mousterian and the functional argument. In
Renfrew, C. (ed.), The Explanation of Culture Change: Models in Prehistory, Duckworth, London,
pp. 227254.
Bleed, P. (1986). The optimal design of hunting weapons: Maintainability or reliability. American
Antiquity 51: 737747.
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. 2346.
Bousman, C. B. (1993). Huntergatherer adaptations, economic risk and tool design. Lithic Technology
18: 5986.
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 Archaeology 23:
263288.
Byers, A. M. (1994). Symboling and the Middle-Upper Paleolithic transition: A theoretical and method-
ological critique. Current Anthropology 35: 369381.
Byers, A. M. (1999). Communication and material culture: Pleistocene tools as action cues. Cambridge
Archaeological Journal 9: 2341.
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. 307361.
Cattaneo, C., Gelsthorpe, K., Phillipos, P., and Sokol, R. J. (1993). Blood residues on stone tools:
Indoor and outdoor experiments. World Archaeology 25: 2943.
Chatters, J. C. (1987). Huntergatherer adaptations and assemblage structure. Journal of Anthropolog-
ical Archaeology 6: 336375.
Christensen, M. (1998). Processus de formation et caracterisation physico-chimique des polis
dutilisation des outils en silex. Applications a la technologie prehistorique de livoire. Bulletin
de la Societe Prehistorique Francaise 95: 183201.
Christensen, M., Walter, P., and Menu, M. (1992). Usewear characterisation of prehistoric flints with
IBA. Nuclear Instruments and Methods in Physics Research B 64: 488493.
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. 85102.
Close, A. E. (1996). Carry that weight: The use and transportation of stone tools. Current Anthropology
37: 545553.
Coffey, B. P. (1994). The chemical alteration of microwear polishes: An evaluation of the Plisson and
Mauger findings through replicative experimentation. Lithic Technology 19: 8892.
Collin, F., and Jardon-Giner, P. (1993). Travail de la peau avec des grattoirs emmanches. Reflexions sur
des bases experimentales et ethnographiques. In Anderson, P., Beyries, S., Otte, M., and Plisson,
H. (eds.), Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 105117.
Conard, N. J., and Adler, D. S. (1997). Lithic reduction and hominid behavior in the Middle Paleolithic
of the Rhineland. Journal of Anthropological Research 53: 147175.
Cowan, F. L. (1999). Making sense of flake scatters: Lithic technological strategies and mobility.
American Antiquity 64: 593607.
Custer, J. F., Ilgenfritz, J., and Doms, K. R. (1988). A cautionary note on the use of chemstrips for
detection of blood residues on prehistoric stone tools. Journal of Archaeological Science 15:
343345.
Darras, V. (1994). Les mines-ateliers dobsidienne de la region de Zinaparo-Prieto, Michoacan, Mex-
ique. Bulletin de la Societe Prehistorique Francaise 91: 290310.
Davidson, I., and Noble, W. (1993). Tools and language in human evolution. In Gibson, K. R., and
Ingold, T. (eds.), Tools, Language and Cognition in Human Evolution, Cambridge University
Press, Cambridge, pp. 363388.
Deacon, J. (1992). Arrows as agents of belief amongst the/Xam Bushmen. Margaret Shaw Lecture 3,
South African Museum, Cape Town.
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 Prehistoric
and Protohistoric Sciences, Forli, Italy, 1996, pp. 9195.
Del Bene, T. A. (1979). Once upon a Striation: Current models of striation and polish formation. In
Hayden, B. (ed.), Lithic UseWear Analysis, Academic Press, New York, pp. 167177.
Dibble, H. L. (1991). Mousterian assemblage variability on an interregional scale. Journal of Anthro-
pological Research 47: 239257.
Dibble, H. L. (1995a). Middle Paleolithic scraper reduction: Background, clarification, and review of
the evidence to date. Journal of Archaeological Method and Theory 2: 299368.
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. 289315.
Dobres, M.-A., and Hoffman, C. R. (1994). Social agency and the dynamics of prehistoric technology.
Journal of Archaeological Method and Theory 1: 211258.
Dockall, J. E. (1994). Oval biface celt variability during the Maya Late Preclassic. Lithic Technology
19: 5268.
Dockall, J. E., and Shafer, H. J. (1993). Testing the producerconsumer model for Santa Rita Corozal,
Belize. Latin American Antiquity 4: 158179.
Downs, E. F., and Lowenstein, J. M. (1995). Identification of archaeological blood proteins: A cau-
tionary note. Journal of Archaeological Science 22: 1116.
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 to
Later Mesolithic transition in northern England. Lithic Technology 23: 4253.
Eisele, J. A., Fowler, D. D., Haynes, G., and Lewis, R. A. (1995). Survival and detection of blood
residues on stone tools. Antiquity 69: 3646.
Ellis, C. (1994). Miniature Early Paleo-Indian stone artifacts from the Parkhill, Ontario site. North
American Archaeologist 15: 253267.
Feblot-Augustins, J. (1993). Mobility strategies in the late Middle Paleolithic of central Europe and
western Europe: Elements of stability and variability. Journal of Anthropological Archaeology
12: 211265.
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. 103118.
Fellner, R. (1995). Technology or typology? A response to Neeley and Barton. Antiquity 69: 381383.
Fiedel, S. J. (1996). Blood from stones? Some methodological and interpretive problems in blood
residue analysis. Journal of Archaeological Science 23: 139147.
Flenniken, J. J., and Raymond, A. W. (1986). Morphological projectile point typology: Replication,
experimentation, and technological analysis. American Antiquity 51: 603614.
Flenniken, J. J., and Wilke, P. J. (1989). Typology, technology, and chronology of Great Basin dart
points. American Anthropologist 91: 149158.
Fredericksen, C. F. K., and Sewell, B. (1991). The reliability of flaked tool function studies in New
Zealand archaeology. Archaeology of Oceania 26: 123126.
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, University
Museum Symposium Series, Vol. 4, University of Pennsylvania, Philadelphia, pp. 113125.
Fullagar, R. L. K. (1991). The role of silica in polish formation. Journal of Archaeological Science 18:
124.
Fullagar, R., Loy, T., and Cox, S. (1998). Starch grains, sediments and stone tool function: Evidence
from Bitokara, Papua New Guinea. In Fullagar, R. (ed.), A Closer Look, Archaeological Methods
Series 6, Sydney University, Sydney, pp. 4960.
Galm, J. R. (1994). Prehistoric trade and exchange in the Interior Plateau of northwestern North
America. In Baugh, T., and Ericson, J. (eds.), Prehistoric Exchange Systems in North America,
Plenum, New York, pp. 275305.
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 from
stone artefacts at Cuddie Springs. In Fullagar, R. (ed.), A Closer Look, Archaeological Methods
Series 6, Sydney University, Sydney, pp. 2948.
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. 189203.
Geneste, J.-M., and Maury, S. (1997). Contributions of multidisciplinary experimentation to the study
of Upper Paleolithic projectile points. In Knecht, H. (ed.), Projectile Technology, Plenum, New
York, pp. 165189.
Gero, J. M. (1991). Genderlithics: Womens roles in stone tool production. In Gero, J., and Conkey,
M. (eds.), Engendering Archaeology, Basil Blackwell, Oxford, pp. 163193.
Gibson, J. L. (1994). Empirical characterization of exchange systems in Lower Mississippi Valley
prehistory. In Baugh, T., and Ericson, J. (eds.), Prehistoric Exchange Systems in North America,
Plenum, New York, pp. 127175.
Gordon, D. (1993). Mousterian tool selection, reduction, and discard at Ghar, Israel. Journal of Field
Archaeology 20: 205218.
Grace, R. (1993a). New methods in usewear analysis. In Anderson, P., Beyries, S., Otte, M., and
Plisson, H. (eds.), Traces et foncion: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 385387.
Grace, R. (1993b). The use of expert systems in lithic analysis. In Anderson, P., Beyries, S., Otte, M., and
Plisson, H. (eds.), Traces et foncion: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 389400.
Grace, R. (1996). Usewear analysis: The state of the art. Archaeometry 38: 209229.
Graves, P. (1994). My strange quest for Leroi-Gourhan: Structuralisms unwitting hero. Antiquity 68:
457460.
Grimaldi, S. (1998). Methodological problems in the reconstruction of chaines operatoires in Lower
Middle Palaeolithic industries. In Milliken, S., and Peresani, M. (eds.), Lithic Technology: From
Raw Material Procurement to Tool Production, Workshop No. 12 of the XIII International
Congress of Prehistoric and Protohistoric Sciences, Forli, Italy, 1996, pp. 1922.
Grimaldi, S., and Lemorini, C. (1993). Retouche specialisee 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. 6778.
Grimaldi, S., and Lemorini, C. (1995). Technology and microwear: Predetermined flakes from the
Mousterian 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. 143155.
Gurfinkel, D. M., and Franklin, U. M. (1988). A study of the feasibility of detecting blood residue on
artifacts. Journal of Archaeological Science 15: 8397.
Hardy, B. L., Raff, R. A., and Raman, V. (1997). Recovery of mammalian DNA from Middle Paleolithic
stone tools. Journal of Archaeological Science 24: 601611.
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. 4759.
Hayden, B., Franco, N., and Spafford, J. (1996). Evaluating lithic strategies and design criteria. In
Odell, G. (ed.), Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York,
pp. 949.
Head, L., and Fullagar, R. (1997). Huntergatherer archaeology and pastoral contact: Perspectives from
the northwest Northern Territory, Australia. World Archaeology 28: 418428.
Hirth, K. G. (1995). The investigation of obsidian craft production at Xochicalco, Morelos. Ancient
Mesoamerica 6: 251258.
Holdaway, S., McPherron, S., and Roth, B. (1996). Notched tool reuse and raw material availability in
French Middle Paleolithic sites. American Antiquity 61: 377387.
Horsfall, G. A. (1987). Design theory and grinding stones. In Hayden, B. (ed.), Lithic Studies Among
the Contemporary Highland Maya, University of Arizona Press, Tucson, pp. 332377.
Hudler, D. (1997). Determining Clear Fork Tool Function through UseWear Analysis: A Discussion
of UseWear Methods and Clear Fork Tools, Texas Archeological Research Laboratory, Studies
in Archeology 25, University of Texas, Austin.
Hudson, L. (1993). Protohistoric Pawnee lithic economy. Plains Anthropologist 38: 265277.
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 CaliforniaGreat Basin exchange. In Baugh,


T., and Ericson, J. (eds.), Prehistoric Exchange Systems in North America, Plenum, New York,
pp. 363383.
Hurcombe, L. (1992). UseWear 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 usewear analysis to understanding the formation
and alteration of wear. In Ramos-Millan, A., and Bustillo, M. A. (eds.), Siliceous Rocks and
Culture, Editorial Universidad de Granada, Spain, pp. 487497.
Hyland, D. C., Tersak, J. M., Adovasio, J. M., and Siegel, M. I. (1990). Identification of the species of
origin of residual blood on lithic material. American Antiquity 55: 104112.
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 Monographs
in Prehistory, Ann Arbor, MI, pp. 4556.
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. 159179.
Johnson, J. K., and Hayes, F. L. (1995). Shifting patterns of long-distance contact during the Middle
Woodland period in the northern Yazoo Basin, Mississippi. In Nassaney, M., and Sassaman, K.
(eds.), Native American Interactions, University of Tennessee Press, Knoxville, pp. 100121.
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. Antiquity 69: 375381.
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. 315344.
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. 7985.
Kealhofer, L., Torrence, R., and Fullagar, R. (1999). Integrating phytoliths within usewear/residue
studies of stone tools. Journal of Archaeological Science 26: 527546.
Kelly, R. L. (1983). Huntergatherer mobility strategies. Journal of Anthropological Research 39:
277306.
Kelly, R. L. (1988). The three sides of a biface. American Antiquity 53: 717734.
Kimball, L. R., Kimball, J. F., and Allen, P. E. (1995). Microwear polishes as viewed through the atomic
force microscope. Lithic Technology 20: 628.
Kooyman, B., Newman, M. E., and Ceri, H. (1992). Verifying the reliability of blood residue analysis
on archaeological tools. Journal of Archaeological Science 19: 265269.
Krakker, J. J. (1997). Biface caches, exchange, and regulatory systems in the prehistoric Great Lakes
region. Midcontinental Journal of Archaeology 22: 141.
Kuhn, S. L. (1992a). Blank form and reduction as determinants of Mousterian scraper morphology.
American Antiquity 57: 115128.
Kuhn, S. L. (1992b). On planning and curated technologies in the Middle Paleolithic. Journal of
Anthropological Research 48: 185214.
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 Mesolithic
of Eurasia, Archeological Papers of the American Anthropological Association, No. 4, Washing-
ton, DC, pp. 2531.
Kuhn, S. L. (1994). A formal approach to the design and assembly of mobile toolkits. American
Antiquity 59: 426442.
Kuhn, S. L. (1995). Mousterian Lithic Technology: An Ecological Perspective, Princeton University
Press, 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 Science 25:
171175.
Leach, J. D., and Mauldin, R. P. (1995). Additional comments on blood residue analysis in archaeology.
Antiquity 69: 10201022.
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). UseWear Analysis on Bone and Antler Tools of the Mackenzie Inuit, BAR
International Series 679, Oxford.
Levi Sala, I. (1993). Usewear traces: Processes of development and post-depositional alterations. In
Anderson, P., Beyries, S., Otte, M., and Plisson, H. (eds.), Traces et fonction: Les gestes retrouves,
ERAUL, No. 50, Liege, pp. 401416.
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. 207217.
Lewenstein, S. M. (1993). Experimentation in the formation and variability of lithic usewear traces
on 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. 287294.
Lohse, E. S. (1996). A computerized descriptive system for functional analysis of stone tools. Tebiwa
26: 366.
Loy, T. H. (1993). The artifact as site: An example of the biomolecular analysis of organic residues on
prehistoric tools. World Archaeology 25: 4463.
Loy, T. H., and Dixon, E. J. (1998). Blood residues on fluted points from eastern Beringia. American
Antiquity 63: 2146.
Loy, T. R., and Wood, A. R. (1989). Blood residue analysis at Cayonu Tepesi, Turkey. Journal of Field
Archaeology 16: 451460.
MacDonald, D. H. (1999). Modeling Folsom mobility, technological organization, and mating strategies
in the Northern Plains. Plains Anthropologist 44: 141161.
MacDonald, D. H., and Hewlett, B. S. (1999). Reproductive interests and forager mobility. Current
Anthropology 40: 501514.
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. Geoarchaeology 8: 241245.
Manolakakis, L. (1996). Production lithique et emergence de la hierarchie sociale: lindustrie lithique
de lEneolithique en Bulgarie (premiere moitie du IVe millenaire). Bulletin de la Societe
Prehistorique Francaise 93: 119123.
Mansur, M. E. (1997). Functional analysis of polished stone-tools: Some considerations about the
nature of polishing. In Ramos-Millan, A., and Bustillo, M. A., (eds.), Siliceous Rocks and Culture,
Editorial Universidad de Granada, Spain, pp. 465486.
Mauldin, R. P., Leach, J. D., and Amick, D. S. (1995). On the identification of blood residues on
Paleoindian artifacts. Current Research in the Pleistocene 12: 8587.
McBrearty, S., Bishop, L., Plummer, T., Dewar, R., and Conard, N. (1998). Tools underfoot: Human
trampling as an agent of lithic artifact edge modification. American Antiquity 63: 108129.
McBryde, I. (1997). The landscape is a series of stories. Grindstones, quarries and exchange in
Aboriginal 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. 587607.
McDonald, M. M. A. (1991). Technological organization and sedentism in the Epipaleolithic of Dakhleh
Oasis, Egypt. African Archaeological Review 9: 81109.
McSwain, R. (1991). A comparative evaluation of the producer-consumer model for lithic exchange
in northern Belize, Central America. Latin American Antiquity 2: 337351.
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. 17.
Mitchell, D. R., and Shackley, M. S. (1995). Classic period Hohokam obsidian studies in southern
Arizona. Journal of Field Archaeology 22: 291304.
Morrow, J. E. (1997). End scraper morphology and use-life: An approach for studying Paleoindian
lithic technology and mobility. Lithic Technology 22: 7085.
Morrow, T. A. (1996). Bigger is better: Comments on Kuhns formal approach to mobile tool kits.
American Antiquity 61: 581590.
Moss, E. H. (1987). A review of Investigating microwear polishes with blind tests. Journal of Ar-
chaeological Science 14: 473481.
Moulton, A. L., and Abler, T. S. (1991). Lithic beings and lithic technology: References from northern
Iroquoian mythology. Man in the Northeast 42: 17.
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 Insights
into Human Prehistory, Plenum, New York, pp. 8199.
Nassaney, M. S. (1996). The role of chipped stone in the political economy of social ranking. In
Odell, G. (ed.), Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York,
pp. 181224.
Neeley, M. P., and Barton, C. M. (1994). A new approach to interpreting late Pleistocene microlith
industries in Southwest Asia. Antiquity 68: 275288.
Nelson, M. C. (1991). The study of technological organization. In Schiffer, M. (ed.), Archaeological
Method and Theory, University of Arizona Press, Tucson, pp. 57100.
Newman, M. E., Ceri, H., and Kooyman, B. (1996). The use of immunological techniques in the analysis
of archaeological materialsa response to Eisele; with report of studies at Head-Smashed-In
buffalo jump. Antiquity 70: 677682.
Newman, M. E., Yohe, R. M., II, Ceri, H., and Sutton, M. Q. (1993). Immunological protein residue
analysis of non-lithic archaeological materials. Journal of Archaeological Science 20: 93100.
OConnell, J. F., and Inoway, C. M. (1994). Surprise Valley projectile points and their chronological
implications. Journal of California and Great Basin Anthropology 16: 162198.
Odell, G. H. (1994a). Assessing huntergatherer mobility in the Illinois Valley: Exploring ambigu-
ous results. In Carr, P. (ed.), The Organization of North American Prehistoric Chipped Stone
Technologies, International Monographs in Prehistory, Ann Arbor, MI, pp. 7086.
Odell, G. H. (1994b). Prehistoric hafting and mobility in the North American Midcontinent: Examples
from Illinois. Journal of Anthropological Archaeology 13: 5173.
Odell, G. H. (1996a). Economizing behavior and the concept of curation. In Odell, G. (ed.), Stone
Tools: Theoretical Insights into Human Prehistory, Plenum, New York, pp. 5180.
Odell, G. H. (1996b). Stone Tools and Mobility in the Illinois Valley: From HuntingGathering Camps
to Agricultural Villages, International Monographs in Prehistory, Ann Arbor, MI.
Odell, G. H. (1998). Investigating correlates of sedentism and domestication in prehistoric North
America. American Antiquity 63: 553571.
Odell, G. H., and Odell-Vereecken, F. (1980). Verifying the reliability of lithic usewear assessments
by blind tests: The low-power approach. Journal of Field Archaeology 7: 87120.
Olausson, D. (1990). Edge-wear analysis in archaeology: The current state of research. Laborativ
Arkeologi 4, Humanistisk-Samhallsvetenskapliga Forskningsradet, Stockholm.
Oswalt, W. (1976). An Anthropological Analysis of Food-Getting Technology, Wiley and Sons, New
York.
Otis Charlton, C. L. (1993). Obsidian as jewelry: Lapidary production in Aztec Otumba, Mexico.
Ancient Mesoamerica 4: 231243.
Owen, L., and Unrath, G. (1989). Microtraces dusure dues a la prehension. lAnthropologie 93:
673688.
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. 285304.
Paton, R. (1994). Speaking through stones: A study from northern Australia. World Archaeology 26:
172184.
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 and
microwear analysis of chipped-stone artifacts from piedmont contexts. American Antiquity 61:
127135.
Piperno, D. R., and Holst, I. (1998). The presence of starch grains on prehistoric stone tools from
the humid Neotropics: Indications of early tuber use and agriculture in Panama. Journal of
Archaeological Science 25: 765776.
Plisson, H., and Mauger, M. (1988). Chemical and mechanical alteration of micro-wear polishes: An
experimental approach. Helinium 28: 316.
Pope, M. K. (1994). Mississippian microtools and Uruk blades: A comparative study of chipped stone
production, use, and economic organization. Lithic Technology 19: 128145.
Pope, M., and Pollock, S. (1995). Trade, tools, and tasks: A study of Uruk chipped stone industries.
Research in Economic Anthropology 16: 227265.
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 autres
enlevements. Bulletin de la Societe Prehistorique Francaise 90: 190195.
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. 671711.
Read, D. W., and Russell, G. (1996). A method for taxonomic typology construction and an example:
Utilized flakes. American Antiquity 61: 663684.
Rees, D., Wilkinson, G. G., Grace, R., and Orton, C. R. (1991). An investigation of the fractal
properties of flint microwear images. Journal of Archaeological Science 18: 629640.
Rick, J. W. (1996). Projectile points, style, and social process in the Preceramic of central Peru. In
Odell, G. (ed.), Stone Tools: Theoretical Insights into Human Prehistory, Plenum, New York,
pp. 245278.
Ricq-de Bouard, M., and Fedele, F. G. (1993). Neolithic rock resources across the western Alps:
Circulation data and models. Geoarchaeology 8: 122.
Rondeau, M. F. (1996). When is an Elko? In Odell, G. (ed.), Stone Tools: Theoretical Insights into
Human Prehistory, Plenum, New York, pp. 229243.
Rosen, S. A. (1996). The decline and fall of flint. In Odell, G. (ed.), Stone Tools: Theoretical Insights
into Human Prehistory, Plenum, New York, pp. 129158.
Rosen, S. A. (1997a). Beyond meat and milk: Lithic evidence for economic specialization in the Early
Bronze Age pastoral periphery in the Levant. Lithic Technology 22: 99109.
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-Shaped
Formed Unifaces on the Interior Plateau of Northwestern North America, Department of
Archaeology, Publication No. 20, Simon Fraser University, Burnaby, BC.
Rowlett, R. M., and Robbins, M. C. (1982). Estimating original assemblage content to adjust for
post-depositional vertical artifact movement. World Archaeology 14: 7383.
Rozoy, J.-G. (1991). Typologie et chronologie. Paleo 3: 207211.
Santone, L. (1997). Transport costs, consumer demand, and patterns of intraregional exchange: A
perspective on commodity production and distribution from northern Belize. Latin American
Antiquity 8: 7188.
Sassaman, K. E. (1992). Lithic technology and the huntergatherer sexual division of labor. North
American Archaeologist 13: 249262.
Sassaman, K. E. (1994a). Changing strategies of biface production in the South Carolina coastal
plain. In Carr, P. (ed.), The Organization of North American Prehistoric Chipped Stone Tool
Technologies, International Monographs in Prehistory, Ann Arbor, MI, pp. 99117.
Sassaman, K. E. (1994b). Production for exchange in the Mid-Holocene Southeast: A Savannah River
Valley example. Lithic Technology 19: 4251.
Sassaman, K. E. (1998). Crafting cultural identity in huntergatherer economies. In Costin, C.,
and Wright, R. (eds.), Craft and Social Identity, Archeological papers of the American
Anthropological Association, No. 8. Washington, DC, pp. 93107.
Schick, K. D., and Toth, N. (1993). Making Silent Stones Speak: Human Evolution and the Dawn of
Technology, Simon and Schuster, New York.
Schreurs, J. (1992). The Michelsberg site Maastricht-Klinkers: A functional interpretation. Analecta
Praehistorica Leidensia 25: 129171.
Seeman, M. F. (1994). Intercluster lithic patterning at Nobles Pond: A case for disembedded
procurement among Early Paleoindian societies. American Antiquity 59: 273288.
Sellet, F. (1993). Chaine operatoire: The concept and its applications. Lithic Technology 18: 106
112.
Shackley, M. S. (1990). Early HunterGatherer Procurement Ranges in the Southwest: Evidence from
Obsidian 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 Maya
site of Colha, Belize. World Archaeology 23: 7997.
Shea, J. J. (1992). Lithic microwear analysis in archaeology. Evolutionary Anthropology 1: 143150.
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 Paleolithic
site of Ubeidiya (Jordan Valley, Israel). Geoarchaeology 14: 191207.
Shea, J. J., and Klenck, J. D. (1993). An experimental investigation of the effects of trampling on the
results of lithic microwear analysis. Journal of Archaeological Science 20: 175194.
Sheppard, P. J. (1993). Lapita lithics: Trade/exchange and technology. A view from the Reefs/Santa
Cruz. Archaeology in Oceania 28: 121137.
Sheppard, P. J. (1996). Hard rock: Archaeological implications of chert sourcing in near and remote
Oceania. In Davidson, J., Irwin, G., Leach, B. F., Pawley, A., and Brown, D. (eds.), Oceanic
Culture History, New Zealand Journal of Archaeology Special Publication, pp. 99115.
Shott, M. J. (1986). Technological organization and settlement mobility: An ethnographic examination.
Journal of Anthropological Research 42: 1551.
Shott, M. J. (1996). An exegesis of the curation concept. Journal of Anthropological Research 52:
259280.
Shott, M. J. (1997). Stones and shafts redux: The metric discrimination of chipped-stone dart and
arrow points. American Antiquity 62: 86101.
Sievert, A. K. (1992). Maya Ceremonial Specialization: Lithic Tools from the Sacred Cenote at
Chichen Itza, Yucatan, Prehistory Press, Madison, WI.
Sievert, A. K. (1994). The detection of ritual tool use through functional analysis: Comparative
examples from the Spiro and Angel sites. Lithic Technology 19: 146156.
Smith, M. E. (1990). Long-distance trade under the Aztec empire: The archaeological evidence.
Ancient Mesoamerica 1: 153169.
Smith, P. R., and Wilson, M. T. (1992). Blood residues on ancient tool surfaces: A cautionary note.
Journal of Archaeological Science 19: 237241.
Sobolik, K. D. (1996). Lithic organic residue analysis: An example from the Southwestern Archaic.
Journal of Field Archaeology 23: 461469.
Spence, M. W. (1996). Commodity or gift: Teotihuacan obsidian in the Maya region. Latin American
Antiquity 7: 2139.
Stafford, M. (1999). From Forager to Farmer in Flint: A Lithic Analysis of the Prehistoric Transition
to Agriculture in Southern Scandinavia, Aarhus University Press, Aarhus, Denmark.
Stiner, M. C., and Kuhn, S. L. (1992). Subsistence, technology, and adaptive variation in Middle
Paleolithic Italy. American Anthropologist 94: 306339.
Storck, P. L. (1997). The Fisher Site: Archaeological, Geological and Paleobotanical Studies at
an Early Paleo-Indian Site in Southern Ontario, Museum of Anthropology, Memoirs No. 30,
University of Michigan, Ann Arbor.
Strauss, A. E. (1992). Jacks Reef corner notched points in New England: Site distribution,
raw material preference, and implications for trade. North American Archaeologist 13:
333350.
Sullivan, A. P., III (1994). Adaptive diversity and limited-activity sites versus logistical mobility and
expedient technology: Adrift in normative thought. Journal of Anthropological Research 50:
159167.
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. 639660.
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. 6172.
Thomas, D. H. (1981). How to classify the projectiles from Monitor Valley, Nevada. Journal of
California and Great Basin Anthropology 3: 743.
Thomas, D. H. (1986). Points on points: A reply to Flenniken and Raymond. American Antiquity 51:
619627.
Thompson, M. (1996). Correlation of Maya lithic and glyphic data. Lithic Technology 21: 120133.
Tomenchuk, J. (1997). A parametric usewear 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. 95161.
Tuross, N., Barnes, I., and Potts, R. (1996). Protein identification of blood residues on experimental
stone tools. Journal of Archaeological Science 23: 289296.
Unger-Hamilton, R. (1989). Analyse experimentale des microtraces dusure: Quelques controverses
actuelles. lAnthropologie 93: 659672.
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. In
Ramos-Millan, A., and Bustillo, M. A. (eds.), Siliceous Rocks and Culture, Editorial Universidad
de Granada, Spain, pp. 499513.
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 Antiquity 47: 276290.
Wallis, L., and OConnor, 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. 149178.
Walthall, J. A., and Holley, G. R. (1997). Mobility and huntergatherer toolkit design: Analysis of a
Dalton lithic cache. Southeastern Archaeology 16: 152162.
Walthall, J. A., and Koldehoff, B. (1998). Huntergatherer interaction and alliance formation: Dalton
and the Cult of the Long Blade. Plains Anthropologist 43: 257273.
White, J. P. (1996). Rocks in the head: Thinking about the distribution of obsidian in Near Oceania. In
Davidson, J., Irwin, G., Leach, B. F., Pawley, A., and Brown, D. (eds.), Oceanic Culture History,
New Zealand Journal of Archaeology Special Publication, pp. 199209.
Whittaker, J. C., Caulkins, D., and Kamp, K. A. (1998). Evaluating consistency in typology and
classification. Journal of Archaeological Method and Theory 5: 129164.
Wilke, P. J., and Flenniken, J. J. (1991). Missing the point: Rebuttal to Bettinger, OConnell, and
Thomas. American Anthropologist 93: 172173.
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 Roman
periods. Geoarchaeology 6: 2760.
Wright, J. V. (1994). The prehistoric transportation of goods in the St. Lawrence River Basin. In
Baugh, T., and Ericson, J. (eds.), Prehistoric Exchange Systems in North America, Plenum, New
York, pp. 4771.
Wynn, T. (1991). Tools, grammar and the archaeology of cognition. Cambridge Archaeological
Journal 1: 191206.
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. 389406.
Yamada, S. (1993). The formation process of usewear polishes. In Anderson, P., Beyries, S.,
Otte, M., and Plisson, H. (eds.), Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege,
pp. 433445.
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. 447457.
Yerkes, R. W., and Kardulias, P. N. (1993). Recent developments in the analysis of lithic artifacts.
Journal of Archaeological Research 1: 89119.
Yohe, R. M., II, Newman, M. E., and Schneider, J. S. (1991). Immunological identification of
small-mammal proteins on Aboriginal milling equipment. American Antiquity 56: 659666.
Young, L. C. (1994). Lithics and adaptive diversity: An examination of limited-activity sites in
northeast Arizona. Journal of Anthropological Research 50: 141154.

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. 179188.
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.), The
Organization of North American Prehistoric Chipped Stone Tool Technologies, International
Monographs in Prehistory, Ann Arbor, MI, pp. 934.
Amick, D. S. (1999). Raw material variation in Folsom stone tool assemblages and the division of
labor in huntergatherer societies. In Amick, D. (ed.), Folsom Lithic Technology, International
Monographs in Prehistory, Ann Arbor, MI, pp. 169187.
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 production
technology in the Southern Plains. North American Archaeologist 12: 129144.
Andrefsky, W., Jr. (1997). Thoughts on stone tool shape and inferred function. Journal of Middle
Atlantic Archaeology 13: 125143.
Bierwirth, S. L. (1996). Lithic Analysis in Southwestern France: Middle Paleolithic Assemblages from
the Site of La Quina, BAR International Series 633, Oxford.
Bosquet, D., and Jardon Giner, P. (1999). Etude traceologique du site paleolithique moyen de
Remicourt-En Bia Flo I. Notae Praehistoricae (Namur) 19: 2128.
Boszhardt, R. F., and McCarthy, J. (1999). Oneota end scrapers and experiments in hide dressing: An
analysis from the La Crosse locality. Midcontinental Journal of Archaeology 24: 177199.
Brooks, L., and Phillips, P. (eds.) (1989). Breaking the Stony Silence: Papers from the Sheffield Lithics
Conference 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 Northeast 43: 8790.
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
Anthropology 12: 289346.
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. 525535.
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 Anthropologist 42: 303-318.
Duff, A. I., Clark, G. A., and Chadderdon, T. J. (1992). Symbolism in the Early Paleolithic: A conceptual
odyssey. Cambridge Archaeological Journal 2: 211229.
Edmonds, M. (1990). Description, understanding and the chaine operatoire. Archaeological Review
from Cambridge 9: 5570.
Franco, N. V. (1994). Maximizacion en el aprovechamiento de los recursos lticos: un caso anal-
izado en el area interserrana Bonaerense. In Lanata, J., and Borrero, L. (eds.), Arqueologa de
Cazadores-Recolectores; Lmites, Casos y Apertures, Arqueologa Contemporanea 5, Edicion
Especial, Buenos Aires, pp. 7588.
Fullagar, R. L. K. (1993). Flaked stone tools and plant food production: A preliminary report on
obsidian tools from Talasea, West New Britain, PNG. In Anderson, P., Beyries, S., Otte, M., and
Plisson, H. (eds.), Traces et fonction: Les gestes retrouves, ERAUL, No. 50, Liege, pp. 331337.
Fullagar, R. L. K. (1994a). Objectives for usewear and residue studies: Views from an Australian
microscope. Helinium 34: 210224.
Fullagar, R. L. K. (1994b). Traces of times past: Stone artefacts into prehistory. Australian Archaeology
39: 6373.
Fullagar, R., Furby, J., and Hardy, B. (1996). Residues on stone artefacts: State of a scientific art.
Antiquity 70: 740745.
Geneste, J.-M. (1991). Systemes techniques de production lithique: Variations techno-economiques
dans les processus de realisation des outillages paleolithiques. Techniques et Culture 17/18:
135.
Gibaja, J. F., and Clemente, I. (1997). El tratamiento termico del slex y sus repercusiones en la determi-
nacion de los rastros de uso. Algunos ejemplos del neoltico en Cataluna. Revista dArqueologa
de Ponent 7: 153160.
Grace, R. (1989). Interpreting the Function of Stone Tools: The Quantification and Computerisation
of Microwear Analysis, BAR International Series 474, Oxford.
Graves, P. (1994). Flakes and ladders: What the archaeological record cannot tell us about the origins
of language. World Archaeology 26: 158171.
Hardy, B., and Garufi, G. T. (1998). Identification of woodworking on stone tools through residue and
usewear analyses: Experimental Results. Journal of Archaeological Science 25: 177184.
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 HunterGatherers,
Publications in Anthropology, No. 19, University of Kansas, Lawrence, pp. 335355.
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). Lanalyse des traces dusure sur lobsidienne. lAnthropologie 96: 179186.
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 dusage et indices de reaffutages et demmanchements
sur des grattoirs magdaleniens de la Grotte Gazel a Salleles-Cabardes (Aude-France).
lAnthropologie 98: 427446.
Kimball, L. R. (1994). Microwear analysis of Late and Terminal Archaic projectile points
from the Padula site (36NM15), Pennsylvania. Journal of Middle Atlantic Archaeology 10:
169179.
Korobkova, G. F. (1994). Stone tools and the beginning of agriculture in the Near East (in Russian).
Archaeological News (St. Petersburg) 3: 166180.
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.), Maya
Stone Tools, Prehistory Press, Madison, WI, pp. 239249.
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. 4656.
Moloney, N., Raposo, L., and Santonja, M. (eds.) (1996). Non-Flint Stone Tools and the Paleolithic
Occupation of the Iberian Peninsula, BAR International Series 649, Oxford.
Nelson, M. C. (1993). Grinding-tool design as conditioned by land-use pattern. American Antiquity
58: 286305.
Nielsen, A. E. (1991). Trampling the archaeological record: An experimental study. American Antiquity
56: 483503.
Odell, G. H. (1993). A North American perspective on recent archeological stone tool research.
Palimpsesto (Buenos Aires) 3: 109122.
Odell, G. H. (1995). Is anybody listening to the Russians? Lithic Technology 20: 4052.
Patterson, L. W. (1994). Incidental impact breakage of arrow points. La Tierra 21: 3038.
Patterson, L. W. (1996). Drilling holes in shell. La Tierra 23: 1013.
Patterson, L. W. (1997). Huntergatherer mobility: Limitations of interpretation. Houston Archeological
Society Journal 117: 18.
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 Urgeschichte 11: 331340.
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 das
Endmesolithikum in Baden-Wurttemberg, Materialhefte zur Archaologie in Baden-Wurttemberg,
No. 39, pp. 150178.
Pawlik, A. F. (1998). Die mikroskopische Gebrauchsspurenanalyse der Silexwerkzeuge aus Reute-
Schorrenried. In Mainberger, M. (ed.), Das Moordorf von Reute, Teraqua CAP, pp. 185198.
Pelegrin, J. (1990). Prehistoric lithic technology: Some aspects of research. Archaeological Review
from Cambridge 9: 117125.
Philibert, S. (1994). Lochre et le traitement des peaux: Revision dune conception traditionnelle par
lanalyse fonctionnelle des grattoirs ocres de la Balma Margineda (Andorre). lAnthropologie 98:
447453.
Prewitt, E. R., and Tomka, S. (1993). What do I call thee? Projectile point types and archaeological
interpretations: Perspectives from Texas. Lithic Technology 18: 4958.
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 usewear generated by processing bison hides. Plains Anthropologist 37:
333351.
Shanks, O. C., Kornfeld, M., and Hawk, D. (1999). Protein analysis of Bugas-holding tools: New trends
in immunological studies. Journal of Archaeological Science 26: 11831191.
Shchelinskii, V. E. (1994). On the function of bifaces from the Mousterian site Zaskalnaja V, Crimea
(in Russian). Archaeological News (St. Petersburg) 3: 1624.
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: 109128.
Shea, J. J. (1993). Lithic usewear 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. 2130.
Shott, M. J. (1995). How much is a scraper? Curation, use rates, and the formation of scraper assem-
blages. Lithic Technology 20: 5272.
Sievert, A. K. (1992). Root and tuber resources: Experimental plant processing and resulting microwear
on chipped stone tools. In Anderson, P. (ed.), Prehistoire de lagriculture, Monographie du CRA,
No. 6, Editions du CNRS, Paris, pp. 5566.
Skakun, N. N. (1994). Agricultural implements and the problem of spreading of agriculture in south-
eastern Europe. Helinium 24: 294305.
Stapert, D., and Johansen, L. (1999). Flint and pyrite: Making fire in the Stone Age. Antiquity 73:
765777.
Tacon, P. S. C. (1991). The power of stone: Symbolic aspects of stone use and tool development in
western Arnhem Land, Australia. Antiquity 65: 192207.
Tankersley, K. B. (1995). Seasonality of stone procurement: An Early Paleoindian example in north-
western New York State. North American Archaeologist 6: 116.
van Gijn, A. L. (1990). The Wear and Tear of Flint: Principles of Functional Analysis Applied to Dutch
Neolithic Assemblages, Analecta Praehistorica Leidensia 22, Leiden.

You might also like