Anda di halaman 1dari 63

Journal of ArchaeologicalMethod and Theory, VoL 3, No.

3, 1996

Skeletal Correlates of Human Behavior in the


Americas
D o n n a C. Boyd 1

A review of recent anthropological research deriving behavioral inferences from


analyses of human skeletal remains focuses on interpretations of diet, social
organization, population structure and migration, activities, activity levels and
occupation, and warfare, violence, and death in the prehistoric and historic
Americas. Critical evaluation of these inferences shows that some of them
suffer from inadequate supporting evidence, lack of consideration of alternative
hypotheses, or absence of clear methods for deriving and testing these
inferences. Overall, however, the potential for behavioral reconstruction from
skeletal analysis is great. Behaviors leaving more direct effects on bone such
as dietary preference, warfare, and some activities produced the most secure
inferences when derived from populational samples and supported by
supplementary evidence. Human skeletal remains represent important resources
for the reconstruction of behavior.
KEY WORDS: bioeultural; behavior; skeletal biology; bioarchaeology;skeletal morphology;
paleopathology;bone chemistry.

INTRODUCTION

Recently, there has been much questioning of the future direction of


physical anthropology as a discipline (Cartmill, 1994; Kelso, 1995; Wienker,
1995). This introspection has centered, in part, around the perceived rela-
tionship between physical anthropology and other anthropology subdisci-
plines. An increasing separation (for some, alienation) between the various
branches of anthropology has been felt, especially among physical anthro-
pologists. Cartmill (1994) attributes much of this isolation to postmod-

1Departmentof Sociologyand Anthropology,Radford University,Radford Virginia 24142.

189
1072-5369/96/0900~189509-~0/0O 1996 Plenum Publishing Corporation
190 Boyd

ernism in cultural anthropology, while Kelso (1995) places the blame firmly
with physical anthropologists. Taking a myopic natural science approach has
prevented some physical anthropologists from seeing the broader, behav-
ioral implications and applications of their findings. Extending our research
focus into the biobehavioral realm has been recommended by both authors
as one avenue for reuniting and revitalizing this discipline.
Some physical anthropologists, particularly skeletal biologists, have
been and are currently conducting just such research. Human skeletal
analysis offers a unique opportunity to explore correlates of human behav-
ior. Because bone and teeth react sensitively to the environment, they often
serve as an indelible record of a population's diet, health, social organiza-
tion, activity patterns, and many other behavioral phenomena (Ubelaker,
1995). Use of a biocultural approach in skeletal biology in the last two
decades has led to the delineation of many of these behavioral signatures.
This approach focuses on the interaction between biology and culture in
evolutionary adaptation (Relethford, 1994). As Larsen (1987, p. 340) points
out, "with this approach, we are closer to breathing life into human remains
in order to view them as representative of functioning, living populations."
Identifying these skeletal correlates of human behavior is the primary
focus of this paper. In the realm of behavioral archaeology, Schiffer (1995,
p. 36) has defined correlates as "statements [which] relate behavioral vari-
ables to variables of material objects or spatial relations." When applied
to human skeletal research, these correlates can be described as statements
relating behavioral variables to skeletal morphology, pathology, or bone
chemistry. Over the past 25 years, many physical anthropologists have
either formally or informally attempted to identify these skeletal correlates
of human behavior, with varying degrees of success. Skeletal biologists have
made some interesting inferences regarding such diverse topics as the origin
and development of agriculture (Buikstra and Milner, 1991; Rose et aL,
1991), the causes of social collapse (Eisenberg, 1991b; Saul and Saul, 1989),
the nature of social systems (Milner, 1992), and the complexities of social
status (Cybulski, 1992; Hatch and Geidel, 1985; Steegrnann, 1991).
A review of recent anthropological research deriving behavioral infer-
ences such as these from human skeletal analyses is the first objective of
this paper. Skeletal research reviewed is confined primarily to those studies
involving hard bony and dental tissue from historic and prehistoric residents
of the Americas, many of them Native Americans. Choice of these skeletal
samples for review reflects not only my own research focus, but the need
for a critical reassessment of our current knowledge of Native American
skeletal biology before permanent repatriation of these remains.
As shown in this literature review (which is by no means exhaustive),
behavioral inferences drawn from human skeletal morphology are of many
Skeletal Correlates of Human Behavior 191

types. Some of these inferences are derived directly from human skeletal
or dental tissue, while others are extrapolated through more indirect means.
There is a concomitant wide range of confidence we can place in the be-
havioral inferences produced by these varying methods. Critical evaluation
of the strength of and justification for these inferences is a secondary goal
of this paper.
Finally, this review illustrates the complexities of the interaction of bio-
logical and cultural processes on the human skeleton. In so doing, I hope
to underscore the relevance, importance, and potential value of human
skeletal analysis and physical anthropology in general in reconstructing and
understanding humans as biocultural beings.

METHODS FOR DERIVING AND EVALUATING


BEHAVIORAL INFERENCES FROM BONE

Detailed considerations of major methodological approaches in skeletal


biology have been presented elsewhere (e.g., Iscan and Kennedy, 1989;
Saunders and Katzenberg, 1992). In fact, specific and standardized proce-
dures have recently been recommended by Buikstra and Ubelaker (1994)
for the analysis of prehistoric and historic Native American bone threatened
with repatriation. It is not my intent here to repeat these methodological
descriptions, only to briefly identify and discuss the importance of the major
approaches used in skeletal biology to make behavioral inferences.
Inasmuch as the complex behaviors in a living human overlap and re-
sist categorization, so too do many of the behavioral dimensions noted in
this review. Nevertheless, behavioral categories examined include diet, so-
cial organization, population structure and migration, activities, activity lev-
els and occupation, and violence, warfare, and death. Table I summarizes
the most common analytical methods and observations skeletal biologists
use to make these behavioral inferences.
Many of these inferences are derived from a macroscopic or micro-
scopic inspection of skeletal morphology and pathology. For example, non-
specific stress indicators such as dental enamel hypoplasias have often been
correlated with diet and health. Enamel hypoplasias are defects in tooth
enamel caused by disruption of amelogenesis or enamel formation (Clark-
son, 1989; Cutress and Suckling, 1982; Goodman and Rose, 1990; Suckling,
1989). Although over 100 biological stresses, including diseases, have been
attributed to their expression, Neiberger (1990) claims that significant cor-
relations between enamel hypoplasias and a specific stressor have been
largely unsuccessful. There is, however, a predictable relationship between
hypoplasias and nutritional status (Goodman, 1994). And, Suckling et al.
192 Boyd

Table I. Observations and Methods Commonly Used in Skeletal Biology to Make


Behavioral Inferences
Method Observation Behavioral inferencea
Paleopathology Enamel hypoplasias D, SO, PS
Porotic hyperostosis D, SO, PS
Nonspecific infection PS, SO, V
Dental disease D, A
Arthritis A
Tranma V, A
Paleodemography Age, sex, mortality PS, SO, V
Skeletal morphology Nonmetric variation SO, PS
Metric variation PS, SO, A
Bone chemistry Trace elements D, SO, PS
Stable isotopes D, SO, PS
Mitochondrial DNA PS, SO
~D, diet; SO, social organization; PS, population structure and migration; A, activities, activity
levels, and occupation; V,, violence, warfare, and death.

(1986) observed hypoplasias after clinically inducing a parasitic disease in


experimental sheep. Enamel hypoplasia frequencies have generally been
used to establish the existence of "stress" in archaeological populations
(Goodman, 1991; Goodman and Rose, 1990; Goodman et al., 1988). Be-
cause enamel formation occurs at a known rate, the timing of this stress
in an individual's life can be estimated (Berti and Mahaney, 1992; Sciulli,
1992). Often, researchers infer that aspects of a population's diet, including
weaning stress and overexploitation of particular resources, were contrib-
uting factors toward high enamel hypoplasia frequencies.
Porotic hyperostosis, including cribra orbitalia, is manifest on the hu-
man skeleton as cranial or orbital lesions believed to reflect iron-deficiency
anemia (Stuart-Macadam, 1991, 1992). It has traditionally been seen as evi-
dence for dependence on nutritionally poor maize agriculture in Native
American populations, although this condition has been recorded in the
absence of maize exploitation. Kent (1986) and Stuart-Macadam (1991,
1992) have shown that cultural correlates of this condition may be just as
important as biological ones. Behaviors accompanying the development and
dependence on agriculture or the rise of sociopolitical complexity may in-
elude population aggregation into large and sedentary settlements with
poor sanitation and waste disposal. These factors increase the frequency
of diarrheal infections and thus iron deficiency anemia. For this reason,
porotic hyperostosis frequencies have been used to make inferences about
diet as well as population structure and social organization.
Skeletal Correlates of Human Behavior 193

Likewise, infectious diseases have both a biological and a behavioral


component. Microbes responsible for acute crowd infections depend upon
a large and sedentary population for transmission. Thus, population struc-
ture is often inferred from the diagnosis of nonspecific infections on bone.
Also inferred are the social consequences of such diseases or other patho-
logical conditions--How was the afflicted perceived and treated by others
(Turkel, 1989)?
As might be expected, dental attrition, caries, abscesses, and microfrac-
tures have been especially useful in inferring diet (Hartnady and Rose,
1991; Teaford, 1991). They also may reflect activities such as tool use or
food preparation, which involve morphological alteration of the dentition
(Milner and Larsen, 1991).
Osteoarthritic patterns in joints and skeletal hypertrophy of major
muscle attachments have allowed some researchers to reconstruct past ac-
tivities and perhaps even occupations (Kennedy, 1989; Merbs, 1983).
Trauma to bone, including perimortem cutmarks, fractures, embedded
weapons, and mutilations, suggests important behavioral aspects of vio-
lence, warfare, and circumstances of death (Milner, 1995). Postmortem cut-
marks, fractures, burning, polishing, and animal damage to human bone
have been especially important data in support of inferences surrounding
mortuary treatment and even cannibalism (White, 1992).
Finally, paleodemographic analysis of skeletal populations can reveal
significant age and sex patterning with regard to mortality, longevity, and
fertility (Jackes, 1992; Roth, 1992; Wood et al., 1992). Aspects of population
structure and migration as well as social organization are often inferred
from inconsistencies in these vital statistics within and between populations.
For example, the absence of infants and small children in mortuary contexts
sometimes leads to a conclusion of differential burial of them elsewhere
(Buikstra, 1981).
Nonmetric skeletal traits are those features which are either present
or absent on a human skeleton and, in turn, are believed to mirror genetic
relatedness between and within skeletal samples (Ossenberg, 1976; Saun-
ders, 1989). Skeletal variation documented from nonmetric trait frequencies
has been used to infer social organization, specifically residence patterns,
by some researchers (e.g., Lane and Sublett, 1972; Konigbserg and Buikstra,
1995).
Metric examination of human skeletal remains has produced behav-
ioral inferences of two different types. First, multivariate craniometric com-
parisons of prehistoric and protohistoric Native Americans have identified
important similarities and differences among and between groups and, by
inference, population migration and gene flow across time and space (Key,
1983; Jantz, 1994). Postcranial metric studies have recently focused on long
194 Boyd

bone geometry and biomechanical function. Wolff's law forms the basis for
much of this research--bone reacts to the mechanical stresses placed on
it during life by strengthening and remodeling (Wolff, 1892). These remod-
eling processes are investigated through techniques borrowed from civil and
mechanical engineering. A mechanical beam model works well for analyz-
ing bone strength and rigidity. If the diaphysis of a limb bone is visualized
as a beam, then its cross-sectional area and second moment of area are
two important dimensions representing the strength of this limb (Ruff and
Larsen, 1990). Changes in long bone diaphyseal strength as well as size
and shape mirror changes in activity levels and biomechanical demands
placed on the pectoral and pelvic skeleton.
Finally, chemical analysis of bone is the most recent and perhaps most
promising method used by skeletal biologists in behavioral reconstructions.
Trace element studies of magnesium, strontium, lead, zinc, calcium, and
other minerals were widely applied in the late 1970s and 1980s in diet re-
construction (Aufderheide, 1989; Sandford, 1992). Relative proportions of
plant and animal sources in prehistoric and historic diets were inferred
through various trace element ratios, owing to the fact that these animals
and plants exhibit varying concentrations of these elements in relation to
their trophic position in the food web (Sandford, 1992). Strontium, mag-
nesium, and calcium occur in higher levels in plants than animals, for ex-
ample. But inconsistencies in relationships between these trace elements
and trophic levels, as well as poorly understood processes of diagenesis
(Burton and Wright, 1995; Ezzo, 1994; Harritt and Radosevich, 1992; Ra-
dosevich, 1993; Sillen et al., 1989), have led to the virtual replacement of
this technique by stable isotope analysis.
Stable carbon and nitrogen isotopes derived from bone collagen show
enormous potential for understanding prehistoric diet. These stable isotope
ratios have allowed researchers to identify and distinguish C3 (temperate
region plants and shrubs) and C4 (tropical grasses such as maize) plants
in the diet as well as marine versus terrestrial food resources (Katzenberg,
1992a). Explanations of differences in these dietary dimensions both within
and among populations and individuals have often involved inferences per-
taining to social organization and population structure and migration. More
specific methodological descriptions of stable isotope analysis are given by
Katzenberg (1992a, b), Keegan (1989), Pate (1994), Sandford (1994),
Schoeninger and Moore (1992), and Schwarcz and Schoeninger (1991).
DNA studies derived from ancient bone hold great promise for elu-
cidating individual as well as population relationships and, by inference,
population structure, migration, and social organization. To date, much of
the focus of this research has been on extracting, analyzing, and interpreting
Skeletal Correlates of Human Behavior 195

mitochondrial DNA. Hagelberg (1993) and Herrmann and Hummel (1993)


have reviewed methodological considerations in DNA analysis.
Evaluation of the strengths of behavioral inferences produced through
these various methods is multifaceted. First, the type of method used to
generate the behavioral inference is important--some methods are more
appropriate and reliable than others in defining particular skeletal corre-
lates of human behavior. Second, sample size is another dimension that
can affect the confidence placed in an inference. Although in some cases
a powerful behavioral inference can be derived from just one individual,
population-based conclusions obtained from large skeletal samples usually
produce more secure inferences than those derived from individual or idi-
osyncratic representatives. Third, do the data derived from a skeletal analy-
sis or other source dearly support the inference made? As with behavioral
inferences drawn in archaeology (Schiffer, 1995, p. 251), a variety of evi-
dence may be available to support such inferences, including experimenta-
tion, actualistic studies, and the prehistoric, historic, and ethnohistoric
bioarchaeological record. Fourth, can the inference be tested? Ultimately,
behavioral inferences should be seen as hypotheses that must be testable;
this "testability" can become a measure of inference strength. Formal test-
ing of alternative hypotheses should also be conducted. All of these factors
will be important considerations in the evaluation of these behavioral in-
ferences.

RECONSTRUCTION OF DIET

A major focus of recent bioarchaeological research has been the docu-


mentation of the effects of subsistence and subsistence change on the hu-
man skeleton (Huss-Ashmore et al., 1982; Larsen 1987, 1994, 1995). A
by-product of this focus has been the accumulation of skeletal and dental
data that have been used to infer the nature of the diet consumed by a
population as well as changes in diet through time. In addition, dietary
inferences regarding timing of weaning have also been made.

Inferences of Dietary Dependence and Diachronic Change

Bone Chemistry

Chemical analysis of human bone has enormous potential for directly


assessing the types of foods a population depended upon, as well as the
dietary variability within and among populations. For example, carbon and
196 Boyd

nitrogen stable isotope values have been derived from human bone collagen
samples from the protohistoric (A.D. 1650-1733) Arikara site of Sully in
South Dakota to test the importance of bison at this Plains site (Tuross
and Fogel, 1994). Climatic simulation models for the northern Plains (Eddy
and Cooter, 1978) suggested that inhabitants of the Sully site may have
suffered through a severe drought. These dry conditions would have re-
duced agricultural productivity, resulting in dependence on alternative food
sources such as bison. However, a previous faunal study by DiUehay (1974)
recorded variability in the presence of bison bones in southern Plains ar-
chaeological sites, calling the climatic simulation model prediction into
question. Carbon and nitrogen isotope values from the 36 individuals ex-
amined at Sully were consistent with heavy utilization of this animal. These
results lend credence to other archaeological and historical evidence for
the importance of bison in Plains Indian subsistence. Similar reconstruc-
tions of prehistoric diets through isotopic analyses are numerous and have
been reviewed by Chisholm (1989), Pate (1994), Schoeninger and Moore
(1992), and Schwarcz and Schoeninger (1991).
Stable carbon isotope analyses have been seminal to the documenta-
tion of the origin and development of maize intensification in the Americas.
Much variability across regions in the timing and nature of this intensifi-
cation has been noted. For example, Buikstra et al. (1988) reported on sta-
ble carbon isotope values from six prehistoric Nashville Basin sites ranging
in age from the Terminal Archaic (2000 B.C.) to the Late Mississippian
(post-A.D. 1200). When these values were compared to those from south-
ern Ontario and the central Mississippi valley, the Nashville Basin isotope
values were among the highest (most positive) for any North American
skeletal sample. These authors concluded that "...agricultural intensification
was more rapid and extreme in the Nashville Basin" compared to these
other areas (Buikstra et al., 1988, p. 248) and that maize intensification
and "Mississippianization" coincide temporally. Eisenberg (1986, 1991a, b)
has recorded paleopathology data reflecting significant amounts of maize-
related nutritional and disease stress at the Nashville Basin site of Averbuch
(A.D. 1275-1400) that support these conclusions.
No contemporaneity between maize dependency and "Mississippiani-
zation," however, was found in the lower Mississippi valley. Rose et al.
(1991) recorded stable carbon isotopic values reflecting the absence of
heavy reliance on maize in Early Mississippian (A.D. 700-1000) times. They
theorized that maize dependency came after increased population size, den-
sity, and sedentism placed stresses on localized food resources. They believe
that a dependence on indigenous starchy seeds such as knotweed (Poly-
gonum erectum), maygrass (Phalaris caroliniana), and goosefoot (Chenopo-
dium berlandieri) preceded the focus on maize in this region.
Skeletal Correlates of Human Behavior 197

More recently, Katzenberg et al. (1995) have recorded a general trend


toward increasing ~13C values through time (A.D. 500-1300) in eastern
North America, mirroring an increased importance of maize in the prehis-
toric aboriginal diet. But northern areas such as Illinois showed higher
~13C values earlier than the southern regions of Tennessee, Arkansas, and
Missouri. They explained this delayed intensification of maize in the south
as possibly due to reliance on previously established indigenous plants.
This documentation of variability among regions in the origin and in-
tensification of maize agriculture has led Buikstra and Milner (1991) to
conclude that an assumption of increasing dependence on maize through
time for all areas in the North American interior is unwise without addi-
tional scientific documentation. And these stable isotope studies have
shown the folly in inferring the existence of a single, uniform "Mississip-
pian" type agricultural or preagricultural diet across the Americas. Archae-
ological evidence in the form of floral and faunal analyses has supported
the existence of such dietary diversity.
Stable carbon and nitrogen isotopic analyses have also been used to
infer changes in foods other than, or in addition to, maize. These studies
have been especially common along the west and east coasts of North
America and have documented the diachronic use of marine resources in
these areas. As with the isotopic evidence for maize, much regional and
temporal variation in use of these marine resources has been noted. For
example, Walker and colleagues (Lambert and Walker, 1991; Walker and
DeNiro, 1986; Walker and Erlandson, 1986) have interpreted changing
carbon and nitrogen isotope values from the Northern Channel Islands in
California as reflecting a gradual increase in dependence on marine re-
sources through time (4000 to 400 years B.P.). As expected, residents of
these islands manifested greater marine dependence compared to those liv-
ing on the mainland coast and mainland interior. On the Georgia coast,
Schoeninger et al. (1990) recorded continued dependence on sea resources
with the adoption of maize agriculture by A.D. 1000. Carbon and nitrogen
isotope values suggest the abandonment of marine foods by the time of
Spanish control.
How confident can we be about these reconstructions of diet and die-
tary change derived from stable isotope analyses? Laboratory and field
studies of experimental animals in the late 1970s and early 1980s under-
going controlled feeding confirmed correlations between dietary variability
and isotopic composition (Bender et aL, 1981; DeNiro and Epstein, 1978,
1981). And isotopic inferences regarding the nature of the diet consumed
and change in diet through time generally agree with archaeological and
other sources of dietary information [but see Little and Schoeninger (1995)
and Walker and Erlandson (1986) for exceptions]. While there are still
198 Boyd

problems to be worked out in the interpretation of stable isotope values,


this method, when applied appropriately, represents one of the most direct
and reliable indications of past diet available to us. These inferences be-
come even more secure when coordinated with faunal, paleobotanical, and
other paleodietary indicators.
Because of aforementioned problems in the interpretation of trace ele-
ment analyses, dietary inferences produced by these methods have recently
become more circumspect. This is especially true when researchers have
gone beyond the goal of reconstructing prehistoric diets toward inferring
personal motives behind dietary behaviors. Blakely (1989) has attempted
to do just that in explaining fluctuations in female strontium and calcium
trace element ratios at the late prehistoric King (16th century A.D.) and
Etowah (A.D. 1000-1400) sites in Georgia. Higher strontium levels for re-
productive-age females compared to postmenopausal females and males
were interpreted by Blakely as reflecting pregnancy, lactation, and food
preferences: pregnant and lactating women were selectively utilizing protein
and carbohydrate-rich foods such as nuts and corn. This is an example of
a behavioral inference derived initially from a chemical analysis of bone,
but that extrapolates beyond the data at hand and is ultimately untestable.

Dental Pathology and Wear

It has long been axiomatic that dental caries frequencies are particu-
larly informative in inferring the type of diet consumed by a population.
Turner (1979) was one of the first researchers to catalogue worldwide av-
erage caries frequencies and correlate them with subsistence type. Although
there was variation within subsistence groups, foragers exhibited a com-
bined (incisor + canine + premolar + molar) caries rate of 1.72%, while
the rate for agriculturalists was 8.6%. Caries rates between 4.5 and 43.4%
were documented by Milner (1984) for a sample of eastern North American
agriculturalists, while hunters and gatherers had a much lower (0.4-7.8%)
caries rate. Larsen et al. (1991), in a comparison of 75 eastern North Ameri-
can archaeological skeletal samples, found additional support for this gra-
dient. Agriculturalists consistently manifested caries frequencies greater
than 7%, while hunters and gatherers' caries rates fell below this figure.
In contrast, dental attrition has generally been observed to be less in
agricultural populations compared to foragers (Boyd, 1988; HiUson, 1990;
Larsen, 1995; PoweU, 1985; Smith, 1982; Walker, 1978). As Larsen (1995,
p. 195) notes, this pattern reflects an adoption and reliance upon increas-
ingly softer and less abrasive agricultural foods. Heavy dental attrition has
Skeletal Correlates of Human Behavior 199

often been used as an indication of use of plant remains of a more abrasive


nature (HiUson, 1990; PoweU, 1985).
It is tempting to use these correlations derived from populations of
known (or suspected) subsistence type to predict the dietary orientation of
groups for which limited subsistence information may be present. This, in
fact, does commonly occur and is mostly used to differentiate preagricul-
tural from agricultural groups. One of the first examples of such an ap-
proach was Turner's (1979) inference of agricultural dependence among
the prehistoric Jomon of central Japan circa 1000 B.C. because of their
high (8.6%) crown caries frequency.
In the New World, this approach has been described and utilized by
Larsen et al. (1991), Milner (1984), Powell (1985), and Walker and Erland-
son (1986), among others. Rose et al. (1991) used a threshold of two carious
lesions per person to differentiate low- versus high-carbohydrate diets. The
presence of substantial frequencies of dental caries a n d attrition for Early
Mississippian (A.D. 700-1000) skeletal samples from the lower Mississippi
valley led Rose et al. (1991) to conclude a mixed subsistence economy for
this area. Dental caries were presumed to indicate only occasional maize
consumption; dental attrition reflected a dependence on indigenous plants.
Paleobotanlcal as well as isotopic data supported this assessment.
Powell and Steele (1994; Steele and Powell, 1993) have used dental
data to evaluate competing archaeological reconstructions of Paleoindian
diet. Traditionally, this diet was thought to have emphasized big game, but
more recent interpretations have stressed a more Archaic-like use of a
broad spectrum of plants and animals. Their comparisons of Paleoindian
dentitions to those of Archaic and Upper Paleolithic ones in terms of dental
caries, attrition, and abscesses supported the latter hypothesis--Paleoindian
teeth were "indistinguishable" from Archaic ones. All individuals examined
(n = 10) showed at least one abscess, while only one individual manifested
a carious lesion. The Paleoindian pattern of dental wear was also similar
overall to the Archaic one, although the anteroposterior wear pattern was
more similar to the Upper Paleolithic. Dental microwear defect compari-
sons between these groups offered the strongest support for these conclu-
sions. Five Paleoindian teeth from five individuals undergoing scanning
electron microscopy wear analysis showed occlusal enamel polishing,
scratching, and pitting similar to more recent hunters and gatherers relying
substantially on plant remains. They concluded that Paleoindians were con-
suming coarse, fibrous, vegetable materials as early as 9700 years B.P.~ sug-
gesting that "...Paleoindians and more recent American Indians shared a
common manner of living" (Steele and Powell, 1993, p. 144). While their
reconstructions of Paleoindian diet may, in fact, be accurate, the discrimi-
nation of different types of foraging diets from dental caries or abscess
200 lloyd

frequencies is uncertain, especially given the sample sizes involved. Studies


involving dental comparisons of modern foragers varying in terms of the
amount of protein and vegetable materials in their diet are needed to more
fully investigate the strength of this dental and dietary correlate.
Diachronic change in diet has also been identified through considera-
tion of dental variability. Once again, increasing dental caries frequencies
combined with decreasing dental attrition have been interpreted within a
framework of increasing dependence upon agricultural cultigens, especially
maize (Larsen et al., 1990, 1992). These inferences have been made even
when they contradict other dietary indicators. For example, archaeological
evidence in the form of dense shell and sea mammal middens suggested
an early reliance on marine resources in the Northern Channel Islands of
California. But Walker and Erlandson (1986) challenged this dietary as-
sessment based primarily on the high caries frequency of early inhabitants
of Santa Rosa Island. They considered this caries data to be good evidence
for early reliance on roots and tubers. A reduction in dental caries fre-
quencies through time (4000-400 B.P.) did corroborate stable isotope values
suggesting an increased dependence upon marine resources. These re-
searchers concluded that the tendency of archaeologists to base dietary re-
constructions on faunal and artifactual evidence alone can lead to a
distorted view of prehistoric diets.
These reconstructions of diet from an analysis of human dentitions
are not without their problems. Much caution must be exercised in in-
ferring diet strictly from dental caries or attrition rates since many vari-
ables may affect their expression. For example, use of grinding stones in
agricultural populations for food preparation can result in greater attri-
tion rates than those seen typically for foragers (Larsen, 1995). Dental
caries often occur as a result of periodontal disease, particularly in a high-
attrition environment (Smith, 1982). This may explain the presence of
high frequencies of dental caries and attrition in some prehistoric popu-
lations. Hunters and gatherers relying on high-carbohydrate indigenous
plants show a caries rate similar to agriculturalists (Sobolik, 1994). And
Larsen (1983; 1995) has noted significantly higher caries frequencies for
females compared to males for many agricultural populations, which he
attributes to "widespread gender-based differences in preparation and
consumption of food" (Larsen, 1995, p. 189). Differences in biological
susceptibility and expression of these dental conditions may also be im-
portant. These biases in the correlation between diet and dental disease
and wear make this type of behavioral inference less certain and secure
unless accompanied by supporting archaeological or other documentary
information.
Skeletal Correlates of Human Behavior 201

Skeletal Morphology and Pathology

Many researchers have used historic and prehistoric bone morphology


and pathology to document changes in diet through time, particularly in
relation to the adoption and intensification of maize agriculture (e.g., Boyd,
1988; Boyd and Boyd, 1989a; Buikstra and Milner, 1991; Buikstra et al.,
1988; Goodman et al., 1980; Hall et al., 1986; Hutchinson and Larsen, 1988;
Katzenberg, 1992b; Katzenberg et al., 1995; Lallo and Rose, 1979; Larsen,
1980, 1990, I995; Larsen et al., 1990, 1991, 1992; Rose et al., 1991). In-
creased pathology frequencies in the form of dental enamel hypoplasias,
porotic hyperostosis, and infectious diseases accompany the shift from for-
aging to agriculture in many regions, reflecting a concomitant decline in
health with this subsistence change (Larsen, 1995). Much regional variabil-
ity in this pattern exists, however (Cohen and Armelagos, 1984).
Changes in cranial and postcranial form have also correlated with the
shift from hunting and gathering to agriculture. Carlson and Van Gerven's
(1977) classic documentation of gracilization in Nubian cranial dimensions
with the adoption of agriculture represents one of the first attempts to cor-
relate such changes in diet and cranial morphology. Several such studies
have also been conducted in the Americas. For example, in a multidimen-
sional comparison of 66 craniofacial and dental measurements from a skele-
tal temporal series (6000 B.C.-A.D. 1600) in Tennessee, Boyd (1988; Boyd
and Boyd, 1989b) noted several functional changes in size and shape of
the masticatory complex across this sequence. These temporal changes in-
cluded reduced masticatory muscle size and robusticity, reduced lower and
lateral midfacial projection, and more posterior points of masticatory mus-
cle attachments relative to the temporomandibular joint, reflecting the
technological shift from a hard-textured bunting and gathering dietary com-
position to a more soft-textured maize one. Reductions and reorientations
in biomechanical demand on the masticatory complex correlated well with
the subsistence shift.
Similar studies documenting postcranial morphological change with the
adoption of agriculture have also been conducted. Because they have been
informative concerning prehistoric activities, these studies are discussed in
a later section.
The central question here, it seems, is What is the predictive power
of such observed correlations among skeletal morphology, pathology, and
diet? In other words, can high frequencies of skeletal pathologies such as
porotic hyperostosis or infectious disease be used to infer dietary orienta-
tion of a prehistoric group? Likewise, can cranial and postcranial metric
dimensions be used to make reliable dietary inferences? Such an approach
has not been commonly conducted by skeletal biologists as of yet, undoubt-
202 Boyd

edly because of the myriad of contributing factors beyond diet which are
involved in disease expression and cranial and postcranial size and shape.
Dietary inferences obtained from this indirect skeletal evidence are much
less secure without the support of more direct dietary indicators.

Inferences About Weaning

The duration of breast-feeding, timing of weaning, and nature of the


weanling diet have been inferred through two main methods--bone chem-
istry analysis and measurement of the frequency, duration, and peak of
enamel defects.

Bone Chemistry

Many researchers have recorded a "trophic effect" in stable nitrogen


isotopes wherein infants have nitrogen values roughly 3% higher than those
of adults. Reduction in these values is believed to correspond with weaning
age. Observation of isotope values of living and historic human groups has
supported this finding. Tuross et al. (1989) recorded such a nitrogen isotope
value difference between modem nursing mothers and their infants, which
disappeared shortly after weaning. Historic skeletal cemetery remains of a
mother and her infant also have shown this trophic level difference
(Katzenberg, 1991b).
This isotopic relationship has important implications for behavioral re-
constructions of weaning. For example, during the Coner ceramic phase
(A.D. 700-1250) of the Maya site of Copan, Reed (1994) determined a
weaning age for the population from 3.5 to 4.5 years of age. A young child
with a ~15N value 2.8% greater than the adult mean was, according to Reed,
still nursing. At the protohistoric (A.D. 1650-1733) Arikara site of Sully in
South Dakota, Tuross and Fogel (1994) estimated that women breastfed
their infants exclusively (without supplementation) for 1 year. Minimal
maize was consumed by infants below the age of 2. Falling nitrogen values
from 2 to 5 years of age suggested that most weanlings as well as small
children ate more fruits and vegetables, but less meat, than adults. And,
at the late prehistoric (A.D. 1530-1580) Ontario site of MacPherson,
Katzenberg et al. (1993) even identified infants who did not nurse. They
hypothesized that the death of their mothers occurred at childbirth or
shortly thereafter. Nitrogen isotope inferences about weaning represent an
exciting research focus with much potential for elucidating patterns of
health and disease in the Americas (Katzenberg, 1992a).
Skeletal Correlates of Human Behavior 203

Enamel Defects

Correlations between weaning age and enamel defects have been in-
vestigated by Blakey and Armelagos (1985), Blakey et al. (1994), Cook
(1981), Lanphear (1990), Reed (1994), and Skinner and Hung (1989). For
example, at the prehistoric Dickson Mounds in Illinois, Goodman et al.
(1984) compared enamel hypoplasia frequencies across a temporal se-
quence (A.D. 950-1300) representing a proposed intensification of agricul-
ture. They found slightly earlier enamel hypoplasia peak frequencies
(2.5-3.0 years) for the fully agricultural time period compared to the prea-
gricultural period range of 3.0-3.5 years. These peak enamel hypoplasia
frequency differences were believed to reflect earlier weaning stress for the
fully agricultural population related to earlier and more intensive use of a
weanling diet. Similarly, Lanphear (1990) correlated an enamel hypoplasia
peak stress frequency of 2.5-3.0 years for many members of a 19th century
poorhouse population from Rochester, New York, in part with early wean-
ing stress brought on by many factors related to the beginning of industri-
alization. When these historic poorhouse values were compared with
similarly derived recorded values from other historic and prehistoric sam-
pies, a pattern of earlier peak stresses (by inference, weaning stress)
through time was noted. Mean weaning ages of 3 to 6 years for many hunt-
ers and gatherers contrasted with prehistoric and historic agricultural
groups' averages of 2 to 6 years and modern populations' averages of 0-3
years.
These reconstructions of weaning behavior from enamel hypoplasia fre-
quencies have been met with criticism by several researchers. At the Middle
Woodland (50 B.C.-A.D. 400) Klunk Mound series in Illinois, Cook (1981)
recorded peak enamel defects ranging from 6 to 24 months. She concluded
it is unlikely that this wide variation reflects only weaning stress. Blakey et
aL (1994) also questioned and tested this association. They found a discrep-
ancy in enamel hypoplasia-derived weaning times and those documented in
the ethnohistoric record for 27 enslaved African-Americans from several
Maryland and Virginia historic sites dating to the 18th and 19th centuries.
The enamel hypoplasia evidence indicated a weaning age of 1.5 to 4.0 years,
whereas the average weaning age documented for historic slaves for this
time period was 9 months to 1 year. Blakey et al. (1994) rejected the "wean-
ing hypothesis" accordingly, although they admitted that postweaning stress
is probably at least indirectly related to enamel hypoplasia patterns. Good-
man and Armelagos'(1985) documentation of variation in enamel hypoplasia
formation with the type as well as developmental stage of teeth suggests
they may be more closely tied to internal biological events rather than ex-
204 Boyd

ternal (weaning) ones. All of these studies reinforce the need for caution
in making inferences about weaning from enamel defects.
In sum, bone chemistry, dental pathology and wear, paleopathology,
and skeletal morphology are valuable methods for the reconstruction of
dietary behavior of prehistoric populations. More direct measures of pa-
leodiet such as stable isotope analysis provided the most reliable inferences,
especially when combined with archaeological, paleobotanical, faunal, or
ethnohistorical dietary information. Researchers (e.g., Hall et at, 1986)
have cautioned that no single method should be relied upon exclusively in
the reconstruction of past diets.

SOCIAL ORGANIZATION

Information derived from human skeletal analysis has often been used
to make statements about the nature of kinship and social groups, social
status, political and economic systems, social change, and health care and
social systems. These behavioral inferences have been made using a variety
of methods and evidence, resulting in a concomitant wide range of reli-
ability and credibility.

Inferences About Kinship and Social Groups

Bone Chemistry

One of the most promising areas for elucidating kinship information


from human bone is that of ancient DNA analysis. Stone and Stoneking
(1993) illustrated the great potential for this technique in their comparison
of mitochondrial DNA samples derived from rib fragments of 50 prehistoric
(A.D. 1300) Norris Farm cemetery skeletons from central Illinois. Of the
four main mitochondrial DNA lineage types identified through their analy-
sis, one was shared by five males only. From this evidence, they inferred
"...immigration of maternally related males into the community" (Stone and
Stoneking, 1993, p. 467). A random distribution of mitochondrial DNA was
noted overall for the cemetery except for two males from the same grave
with identical genetic markers and lineage types. It is possible that they
represent brothers, although there is a 39% chance that the association is
coincidental. While studies like these offer enormous opportunities for di-
rect inferences about kinship relations from human bone, this technique is
still in its infancy. More research is needed to address such issues as DNA
preservation in bone, sampling techniques, contamination of DNA samples,
Skeletal Correlates of Human Behavior 205

and stochastic variation in DNA material (Hagelberg, 1993; Herrrnann and


Hummel, 1993) before such inferences can he accepted with confidence.
Stable isotope analyses have also been used to infer kinship informa-
tion. Walker and DeNiro (1986) recorded significant carbon and nitrogen
isotopic variation among individuals from two mainland coastal sites in the
Northern Channel islands of California. One site (SBa-119) dated to the
Early-Middle period (<5000 B.C.-A.D. 1150), while the other (Ven-11)
dated to the Middle-Late period (1400 B.C.-A.D. 1804). One interpreta-
tion offered to explain the existence of both high and low stable carbon
and nitrogen isotope values for different individuals within these sites was
exogamy between populations of the mainland coast and the interior. Ma-
trilocal postmarital residence was also inferred: "Men from the interior who
are likely to have consumed a predominantly terrestrial diet while growing
up would, upon marriage, move to the coast to the residence of their wives"
(Walker and De Niro, 1986: 60). Small samples prevented the testing of
this inference, but baptismal evidence showing the predominance of ma-
trilocality in early historic mainland coast populations provided support.
Trace element composition of dental enamel has been used by
Schneider and Blakeslee (1990) to establish residence patterning among
four Plains Village 2 and Protohistoric (A.D. 1250-1700) northern and cen-
tral Plains Arikara sites. Males from all of these sites showed greater "dis-
persal" in their enamel composition comparisons, suggesting out-migration
(matrilocality) for these individuals. Females and subadults showed a more
tightly clustered pattern within each site. Although Schneider and Blakeslee
rejected possible biases, they acknowledged that this difference in dental
enamel composition can be influenced by geographic or environmental fac-
tors or those relating to dietary specialization or gender differences in ele-
mental uptake.

Skeletal and Dental Morphology

Examination of morphological attributes of the prehistoric and historic


human skeleton and dentition in the Americas has resulted in behavioral
inferences regarding the identification of social groups. Multivariate com-
parison of metric dimensions of human crania has long been recognized
as a powerful tool in delineating these genetic relationships (Jantz, 1994),
although factors other than genetic ones may influence human cranial mor-
phology. Correlation of multivariate cranial data with linguistic, archae-
ological, and ethnographic information from the historic (A.D. 1804-1832)
Plains Leavenworth site in South Dakota allowed Byrd and Jantz (1994)
to differentiate two distinct Arikara groups. While they were not capable
206 Boyd

of identifying distinct bands within this site, a stepwise discriminant analysis


comparing 15 cranial measurements did indicate the presence of two dif-
ferent groups. That these two groups were probably descendants of the
Bad River and Le Beau Phase populations was suggested by archaeological
comparisons of artifacts from these groups. Linguistic study of m o d e m and
historic Arikara was also used to support their argument--two distinct Ari-
kara dialects have been recorded. Byrd and Jantz (1994, p. 206) concluded
that Leavenworth represented "...an amalgamation of formerly heteroge-
neous breeding populations. Variation among social groups apparently per-
sisted in spite of the likelihood of intermarriage at Leavenworth."
Konigsberg and Buikstra (1995) used nonmetric skeletal trait comparisons
to similarly identify within-site group differentiation as well as between-site
boundary definitions for a prehistoric skeletal series from west--central Il-
linois.
Residence patterning has also been inferred from a comparison of non-
metric or metric trait frequencies within skeletal populations (Konigsberg,
1988; Konigsberg and Buikstra, 1995; Lane and Sublett, 1972; Spence,
1974). Nonmetric cranial trait comparisons between male and female his-
toric (A.D. 1850-1930) Seneca cemetery skeletons by Lane and Sublett
(1972) showed considerable heterogeneity between the sexes within each
specific cemetery. Between-cemetery comparisons showed heterogeneity
between males but homogeneity in females. Lane and Sublett concluded
that a male-male genetic relationship existed within these cemeteries.
While they acknowledged that residence patterning could not be directly
assessed using this method, their evidence suggested patrilocality for this
group. This represents a significant departure from the ethnohistorically
documented matrilocality for the 17th- and 18th-century Iroquois. Lane and
Sublett's research has been criticized by Konigsberg (1988) as well as others
because a formal population genetic framework for deriving testable mod-
els for residence pattern hypotheses is not used.
More recently, through a combination of population genetics and
bioarchaeology, Konigsberg (1988) and Konigsberg and Buikstra (1995)
were able to infer residence patterning for a prehistoric west-central Illinois
skeletal series ranging from the Middle Woodland (A.D. 50-400) to the
Mississippian period (A.D. 1000-1400). Comparisons of male and female
covariance ratios derived from five nonmetric skeletal traits suggested pa-
trilocatity due to increased Middle and Late Woodland female mobility
compared to male. Male migration, assumed to represent matrilocality, was
seen in the Late Mississippian sample, paralleling the intensification of ag-
riculture.
The basic assumption underlying these nonmetric trait comparisons is
that variation in trait frequency reflects genetic differences. Saunders
Skeletal Correlates of Human Behavior 207

(1989) and Van Vark and Schaafsma (1992) point out that such genetic
inferences from nonmetric traits must be approached with caution. Some
nonmetric traits used by skeletal biologists to differentiate skeletal groups
reflect pathological or functional aspects of the human skeleton rather than
genetic. For those that do have a genetic component, often this association
is a weak one. Only low to moderate heritabilities have been documented
for many nonmetric cranial traits through experimentation with mice (Self
and Leamy, 1978) and rhesus macaques (Chevrud and Buikstra, 1981a, b).
When the appropriate nonmetric traits are chosen, however, this technique
does hold much potential for elucidating kin relationships.

Paleopathology and Paleodemography

Evidence derived from skeletal and dental pathologies has been used
by several researchers to infer mating and residence patterns. For example,
the most common reason for the skeletally derived behavioral inference of
inbreeding is the identification of genetic abnormalities within a population
(Turkel, 1989). Suggestions of genetic isolation have been made following
diagnoses of prehistoric dwarfism (Langdon et al., 1993), transposed teeth
(Nelson, 1992), Klippel-Feil syndrome (Danforth et aI., 1994), and vertebral
anomalies (Bennett, 1973). A small population size combined with relative
geographic isolation may have been contributing factors in some of these
cases (e.g., Nelson, 1992). The underlying assumptions here are that pre-
historic frequencies for these conditions mirror modern ones and that uni-
formity across modern groups exists for these genetic abnormality
frequencies. Neither of these assumptions may be correct.
Dental pathology has been interpreted by Cybulski (1992, 1994) to re-
flect changes in descent systems and residence patterning in the Late De-
velopment stage (1500 B.P.-present) on the prehistoric Northwest coast.
He correlated the presence of distinctive tooth scars on prehistoric British
Columbia Native American females with the wearing of labrets. Dentitions
were unaccompanied by labrets, however, leading Cybulski to propose an
"heirloom hypothesis"--labrets were reflections of matrilineal family status
and wealth that were removed and passed on to other female family mem-
bers upon the death of their loved one. The association of labret scars with
male or both male and female dentitions from earlier (5200-1500 B.P.) ar-
chaeological contexts on the Northwest coast reflected patrilineal and bi-
lateral kinship systems, respectively, for these populations. Supporting
evidence in the form of ethnographic descriptions of the importance of
labrets in Northwest Coast social status and wealth supported Cybulski's
claims, although alternative hypotheses for the correlation of labret scars
208 Boyd

and demographic data are not fully explored. More direct archaeological
and cultural documentation of the correlation between labret scars and so-
cial status would add weight to Cybulski's inferences.
Elevated infant and childhood mortality and porotic hyperostosis fre-
quencies at the prehistoric (A.D. 1300-1370) Arroyo Hondo Pueblo in New
Mexico prompted Wetterstrom (1994) to conjecture a possible shift in resi-
dence patterns for this group as a means of surviving this biological stress.
She theorizes that a bilateral kinship system would have led to more flexible
rules concerning marriage, status, responsibilities, and privileges and would
have increased the potential kinsmen available for help in a stressful en-
vironment. This is an example of a behavioral inference which was more
of an afterthought than a testable hypothesis. The association between bio-
logical stress and residence patterning is a weak one, with little evidence
to support it.

Inferences About Political and Economic Systems

Prehistoric and historic involvement in political and economic spheres


has been inferred through several different types of skeletal studies. These
studies have involved primarily an analysis of trace elements and stable
isotopes, craniometrics, and paleopathology.

Bone Chemistry

Skeletal evidence for trade has been documented through analysis of


trace elements, particularly lead. Reinhard and colleagues (Reinhard and
Ghazi, 1992; Reinhard et aL, 1994; Ghazi et al., 1994) have recorded high
lead values for 18th-century Omaha Indian children and adult males. Some
of this lead bore a metallic-lead signature, apparently traded in the form
of musket balls and ornaments from the Missouri area. Correlations in
these artifact lead values with trace element values for some skeletal re-
mains supported this inference. The majority of the lead, however, was of
a "non-Mississippi River valley type," because it lacked the enriched ra-
diogenic isotopes typical of Missippian River valley lead. This non-Missis-
sippi River valley type of lead was traced to pigments applied to the face
and chest of individuals in pre- as well as postmortem situations. High levels
of lead in adult males were explained in terms of involvement in trade of
artifacts and, especially, heavier pigment application of their faces. High
frequencies for children reflected the physiologically documented increased
Skeletal Correlates of Human Behavior 209

absorption rate of lead into their bodies, probably from skin pigment ap-
plication as well.
In the American Southwest, Spielmann et al. (1990) evaluated a pre-
vious archaeological hypothesis concerning the impact of an increase in
trade between 15th- to 17th-century inhabitants of Pecos Pueblo, New Mex-
ico, and Plains hunters and gatherers. This trading behavior was believed
to have had a significant impact on Pecos Pueblo, resulting in a replacement
of mule deer by bison in food preference. This inference was tested by
these researchers using stable carbon and nitrogen isotope and strontium
trace element ratios. Spielmann and colleagues found no such evidence for
this dietary shift at Pecos Pueblo during this trading period, in spite of an
increase in bison bones in faunal assemblages from this area at this time.
Thus, it appears that this trade had an insignificant effect upon Pecos
Pueblo diet.

Skeletal Morphology

In a fascinating study by Jantz and Owsley (1994), a multivariate cra-


niometric comparison of individuals from the Coalescent tradition Plains
Indian village of Swan Creek (A.D. 1675-1725) in South Dakota also re-
vealed information about trade. Jantz and Owsley were able to differentiate
the cranium of a Caucasian, probably French, male among the historic Ari-
kara crania and concluded he was one of the earliest traders to contact
this region. This physical evidence confirmed that contact between Euro-
peans and the historic Arikara took place many years earlier than docu-
mented ethnohistoricaUy. But specific skeletaUy derived conclusions about
early trading behavior between these groups must remain conjectural with-
out the input of the ethnohistoric record.

Paleopathology and Paleodemography

Perhaps the most indirect economic and political behavioral inferences


have come from the documentation of biological stress in prehistoric popu-
lations. For example, Goodman and Martin (1992) have interpreted a de-
cline in health at the Illinois Dickson Mounds through time (A.D.
950-1350) as evidence for increased involvement in the interregional eco-
nomic system. Increased trauma and degenerative pathology frequencies
across time were explained in terms of greater emphasis on trade with Ca-
hokia--the necessity for agricultural surpluses at Dickson Mounds for trad-
ing purposes may have resulted in agricultural intensification, increased
210 Boyd

workloads, and associated trauma and biological stress. The result, accord-
ing to these researchers, may have been formal inequities in wealth.
In contrast, the relatively better health for the prehistoric (A.D. 800-
1150) Black Mesa Anasazi of Arizona is explained in terms of their more
marginal geographic location within the Anasazi sphere of influence. This
resulted in greater political and behavioral flexibility for this group (Good-
man and Martin, 1992; Goodman et aL, 1992). Pathology data in the form
of enamel hypoplasia, porotic hyperostosis, and infectious disease frequen-
cies indicated only mild to moderate endemic stress, suggesting that the
Black Mesa Anasazi may have been able to adapt successfully to their enivi-
ronment through food sharing, trade, and food storage. Martin et al. (1991)
believe that this behavioral flexibility was the key to their continued survival.
Although Goodman and Martin (1992, p. 59) assert, "Life style and health
are invariably linked to politics and economics," a clearly defined methodo-
logical basis for supporting these skeletal correlates has yet to be established.

Inferences About Social Status

A number of researchers have explored the skeletal manifestations of


social status (Dye, 1976; Little et aL, 1992; Parham and Scott, 1980; Rathbun
and Scurry, 1991; Schurr, 1992; Smith, 1993b; ~inter, 1980). Information
concerning how status might have been conferred in a society as well as
how individuals of different statuses may have been treated has often been
inferred from bone chemistry, skeletal morphology, and paleopathology.

Bone Chemistry

Intrapopulation variation in stable isotopes and trace elements is often


interpreted as status-related differential access to resources. For example,
trace element analyses have been used to infer social restrictions placed on
animal protein at Chalcatzingo, Mexico (Schoeninger, 1979), Gibson
Mounds in Illinois (Lambert et al., 1979), and the Dallas culture in East
Tennessee (Hatch and Geidel, 1985). In all these cases, archaeologically de-
fined higher status individuals showed greater consumption of animal pro-
tein. Variation in stable carbon isotope ratios at the ancient Mayan site of
Copan (Reed, 1994), the precontact site of LaFlorida in highland Ecuador
(Ubelaker et aL, 1995), and Lit'the Cypress Bayou in the central Mississippi
valley (Rose et aL, 1991) was believed to represent differential access to
maize between high- and low-status individuals from these regions.
Skeletal Correlates of Human Behavior 211

Aufderheide and coUeages (1985, 1988) have used trace element stud-
ies of historic American skeletal populations to identify socioeconomic sub-
groups and infer their treatment. Specifically, high levels of lead in Colonial
period skeletons often reflect the wealthy's use of lead-glazed ceramic ves-
sels. At the historic Catoctin Furnace in Maryland, late 18th and early 19th
century African-American iron-working slaves generally manifested low
lead values, with the exception of some African-American females. These
researchers interpreted these higher lead values as due to their consider-
able access to wealthy landowners' homes, including their cooking vessels.

Skeletal Morphology

Differences (or lack thereof) in skeletal morphology within a popula-


tion have sometimes been used to identify status relationships among the
population's members. Wilkinson and Norelli (1981) documented similari-
ties in dental and skeletal metric and nonmetric features between high-
status tomb and low-status burial interments from the prehistoric site of
Monte Alban (500 B.C.-A.D. 650). They inferred from this evidence a cer-
tain amount of flexibility in movement across social categories. The absence
of reduced stature for individuals from a 19th century New York poorhouse
cemetery prompted Steegmann (1991) to conclude adequate nutrition for
this group. He speculated that food must have been available to even the
poor and that economic stratification was less than that seen today.

Paleopathology and Paleodemography

Documentation of intrapopulation similarities and differences in


health by some researchers has resulted in implications for status relation-
ships. For example, Powell's (1988, 1991) investigation into status and
health at Moundville (A.D. 1250-1550) in Alabama revealed only very
slight differences in the health status of archaeologically defined elite and
nonelite individuals. Few status-related differences were seen in dental dis-
ease, infection, anemia, nonspecific stress, or trauma. She concluded that
the Moundville nonelite did not suffer stress-related resource deprivation
because of their inferior status and that much interregional variability exists
in the skeletal manifestations of status.
Trauma frequency comparisons across social groups have alsc~ led to
insights about the nature of status. Lahren and Berryman (1984) noted a
high frequency of fractures for late prehistoric (A.D. 1000-1600) archae-
ologically defined high status Chucalissa males from West Tennessee. Frac-
tures were most common on the cranium but also included healed and
212 Boyd

unhealed postcranial injuries. These authors concluded that "...high status


males were participating in activities that increased their susceptibility to
trauma" (Lahren and Berryman, 1984, p. 18).
Blakely (1977, 1980, 1995) has used paleopathological and demo-
graphic data from prehistoric inhabitants at Etowah, Georgia, to infer the
type of status being conferred. A comparison of periostitis, osteomyelitis,
osteoporosis, osteoarthritis, porotic hyperostosis, fracture, tumor, and peri-
odontal disease frequencies across mound and village burials revealed no
overall significant differences (Blakely, 1980). Blakely concluded that status
must have been achieved rather than ascribed, since no preferential treat-
ment of mound individuals was detected. However, more recent compari-
sons of mound and village demographic profiles showed a higher
proportion of older males in the mound, leading Blakely (1995, p. 55) to
conclude that "age-progressive inclusion of males in Mound C suggests he-
reditary succession to high rank by those eligible. It does not preclude the
possibility that positions were achieved as well."
The use of skeletal indicators of health to differentiate achieved and
ascribed status for members of a population has little theoretical or experi-
mental support. As already stated, a number of environmental, physiological,
and social factors are involved in these skeletal manifestation of health. So-
cial status is just one of these factors. Use of bone chemistry, skeletal mor-
phology, and paleopathology to infer treatment of archaeologically defined
status groups also rests on indirect behavioral correlates and is subject to
similar biases. Assumptions of behavioral intent must also be made--certain
segments of a population manifest better health than others not through
serendipitous circumstances but by design. When supported by supplemen-
tary archaeological or historic evidence, these types of behavioral inferences
may be credible and testable, but must be approached with caution.

Inferences About Social Change

A number of researchers have made statements about social processes


from the human skeleton. These have been primarily derived from stable
isotope analysis as well as consideration of paleopathology data.

Bone Chemistry

Increasing sociopolitical complexity has recently been inferred through


stable carbon and nitrogen isotope analysis of bone collagen and bone min-
eral apatite from the precontact (A.D. 100--450) site of LaFlorida, highland
Skeletal Correlates of Human Behavior 213

Ecuador. Ubelaker et aL (1995) compared isotope values from archaeologi-


cally defined high- and low-status burials. They found significant maize use
in both groups, but high-status individuals consumed more. Since etlmo-
historic and archaeological sources indicate that much maize was consumed
as chicha (beer) and manufactured in the chief's household, they speculated
about the redistribution of maize beer by chiefs at large drinking parties
in a quest for power. A sociopolitical use for maize led to:
"the establishment of a chief's household as an unrivaled nucleus of production
which could outeompete all others in the acquisition of exotic sumptuary goods
whose exclusive display or preferential redistribution clearly denoted social
inequalities" (Ubelaker et aL, 1995, p. 410).

Nitrogen isotope ratios were not significantly different in high- and low-
status individuals; thus, they inferred that animal protein was not socially
restricted.

Paleopathology

Many skeletal biologists have correlated pathological evidence for bio-


logical stress in prehistoric populations with theories about social collapse
of the societies they represent (Eisenberg, 1991a, b; Milner, 1991; Walker
and Lambert, 1989). At the heart of these theories is the inference that
decline in health was one factor that accompanied social disorder, disinte-
gration, and eventual collapse. Overreliance on maize agriculture has, in
part, been used to explain the disappearance of the Maya (Reed, 1994;
Saul, 1976; Saul and Saul, 1989), prehistoric Native Americans from the
Georgia Bight (Larsen et al., 1992), and prehistoric residents of the Aver-
buch site in Tennessee (Eisenberg, 1991a, b). Eisenberg (1991b), Milner
(1991), and Larsen et al. (1992) point out, however, that a multitude of
biological and cultural factors may be involved, including diet, population
size and density, degree of sedentism, and warfare, as well as environmental
and sociopolitical considerations. No single cause may be identified.

Inferences About Health Care, Child Care, and Social Systems

Paleopathology

Identification of pathological conditions on prehistoric and historic


skeletal tissue from the Americas has resulted in a myriad of fascinating
behavioral inferences about health and child care and social systems. For
example, HoUiday (1993) has interpreted a high frequency of occipital le-
214 Boyd

sions on infants below the age of one from a Southwest New Mexico Mim-
bres-Mogollon Pueblo skeletal sample (A.D. 700-1150) to be either directly
or indirectly associated with the cultural practice of cradleboarding. Only
infants from NAN Ranch Pueblo showed these active lesions, while older
subadults and adults manifested healed lesions in the same region of the
occipital. Holliday believed these lesions may have been the result of is-
chemic ulcers or bacterial infections owing to excessive pressure and fric-
tion on an infant's head. Ethnohistoric descriptions of Pueblo infants
strapped to cradleboards up to 20 hr a day for as much as 10 months sub-
stantiate the prevalence of this cultural practice for Puebloan groups, but
the causal relationship between cradleboarding and occipital lesions re-
mains elusive. Skeletal evidence for these lesions in other Puebloan samples
is lacking. Holliday speculated that the absence of this supporting evidence
may be due to inadequate sampling and preservation at many of these other
sites. Another possibility is that the population at NAN Ranch Pueblo, mo-
tivated by aesthetic desires or greater child-care demands, had different
cradleboarding practices compared to other groups. Perhaps they simply
placed infants on cradleboards for longer periods of time.
Evidence from the dentition and skeleton has been used to infer health
care of individuals as well as populations. The presence of caries in con-
junction with two drill holes in the enamel of a mandibular incisor from a
protohistoric (A.D. 1300-1700) Tigaran Point Hope, Alaska, individual led
Schwartz et al. (1995) to infer the first probable case of precontact New
World Arctic dentistry. In the Plains, Willey and Hofman (1994) correlated
interproximal grooves with diseased prehistoric Native American teeth from
both the Woodland and the Village periods and inferred a therapeutic func-
tion for these grooves. Ethnohistoric evidence for the use of the purple
coneflower plant in this region for toothaches supported this convincing
explanation for these grooves.
At the Snake Hill military cemetery (Pfeiffer and Williamson, 1991)
from the War of 1812, several individuals showed skeletal evidence for am-
putation of limbs. One person had undergone bilateral amputation of the
humeri (Owsley et aL, 1991). Excavation of three "medical waste" features
at the site produced eight amputated limbs. This evidence offered unique
insights into the medical and surgical care of battle casualties from this
war. Many traumas appeared to have been treated "on site," leading to
the inference that a field hospital may have been located in the near vi-
cinity. Medical records for this time period indicated that surgical opening
of the chest cavity was not performed; however, morphological alteration
of a rib fragment from one soldier suggested otherwise (Owstey et al., 1991).
The identification of a 200-year-old Plains Indian child with extreme
otitis media (middle ear infection) as well as other debilitating infectious
Skeletal Correlates of Human Behavior 215

conditions prompted Mann et al. (1994) to speculate about Omaha health


care. The prevalence of these active infections in contrast to well-healed
traumatic injuries for other individuals, they believe, reflected the ability
of the Omaha to treat trauma more effectively than infection.
The existence of deformed as well as injured individuals at several pre-
historic sites has led some researchers to reach conclusions about the care
and treatment of these individuals by other members of their group. Rath-
bun et al. (1980) interpreted partial spina bifida and severe hip degenera-
tion in two individuals from a South Carolina coastal Archaic population
(Daw's Island--38BU9) as evidence for social care of these individuals.
Langdon et al. (1993, p. 277) explained the presence of a prehistoric dwarf
at the Extended Coalescent Twins site (A.D. 1500-1600) in South Dakota
as evidence that "altruism and care for the infirm must have also played
an important role in [this] society."
One of the most famous examples of this kind of behavioral inference
comes from the Early Archaic Windover site (7500 years B.P.) in Florida
(Dickel and Doran, 1989). Here the skeleton of a 15-year-old showed se-
vere spina bifida associated with several secondary complications, including
scoliosis (curvature) of the spine. Several long bones manifested signs of
both active and healed infection and atrophy from disuse. Dickel and
Doran (1989) hypothesized that this individual was severely debilitated,
with limited limb mobility. Given the seasonal hunting and gathering round
of the culture in which he resided, they inferred he must have been cared
for by others in his group.
Dettwyler (1991) has been very critical of these conclusions and has
recommended greater caution in reconstructing behaviors such as these
only from skeletal morphology. She specifically attacked interpretations of
the Windover skeleton as well as the Romito 2 dwarf from Italy and Shani-
dar I from Iraq in her article entitled "Can paleopathology provide evi-
dence for 'compassion'?" Dettwyler answers this question with a resounding
'no'. She believes that previous researchers making these kinds of infer-
ences have incorrectly assumed that disabled persons were always incapaci-
tated and unproductive and that their survival is evidence of a group's
compassion for them. Perhaps they were simply lucky in their survival or
showed an enormous degree of resourcefulness or strength of will. Perhaps
others treated them with hatred or indifference, rather than compassion.
Turkel's (1989) discussion of the great ethnographic variability in cultural
reaction to congenitally deformed individuals supports this disassociation
of skeletal malformation and compassion. Some deformed individuals were
feared by other members of the group, some hated, and others exalted.
Dettwyler also questioned the assumption of group support of a handi-
capped individual as the compassionate thing to do, as well as the view
216 Bo,yd

that handicapped persons represent the only members of a society who may
be unproductive and dependent on others within a group. In sum,
Dettwyler (1991, p. 384) is pessimistic about reconstructions of this sort,
stating that "...these [behavioral] questions cannot be answered in archae-
ological contexts." While this attitude may be perceived by some as perhaps
too dogmatic, the healthy skepticism and caution in interpreting compas-
sion from human skeletal morphology are wise.
Conclusions about kinship, social systems, social change, economic and
political systems, and health care and social systems derived from studies
of human bone represent, in many cases, weak inferences. This is so be-
cause social organization and its behavioral manifestations leave only indi-
rect signatures on bone. Many aspects of social behavior may be
unknowable through an analysis of the human skeleton. In spite of these
limitations, skeletally derived inferences about social organization were pos-
sible in some contexts and were much more secure when supported by
other lines of evidence.

MIGRATION AND POPULATION STRUCTURE

Skeletal analyses can provide insight into the nature of prehistoric and
historic populations with respect to their size, density, settlement pattern-
ing, and degree of isolation. Evidence for migration of populations, aS well
as individuals, has also been gleaned from a variety of skeletal studies.

Skeletal Morphology

Evidence in support of or against population migration has occasion-


ally been inferred from multivariate craniometric comparisons (Boyd and
Boyd, 1991; Danforth et aL, 1994; Key, 1983). Key's multivariate compari-
son of Plains Indian crania exemplifies this approach well. His analysis re-
vealed strong evidence for biological continuity with the Middle Missouri
tradition culminating with the Mandan. In contrast, Central Plains crania
were distinct from those from the earlier Plains Woodland period, implying
an intrusion into this region perhaps from the south. Movement of Cad-
doan speakers up the Missouri River resulting in gene flow with the in-
digenous Mandan was suggested by these cranial comparisons. Although
multivariate craniometric comparisons such as these can be clouded by sev-
eral genetic as well as environmental factors, enormous potential exists for
reconstructing and understanding population relationships and mobility
(Jantz, 1994).
Skeletal Correlates of Human Behavior 217

Skeletal morphology has also been correlated with structural dimen-


sions of populations; these correlations often involve evidence for reduced
growth or poor health. Owsley and Jantz (1985), for example, observed
smaller late post-contact period (A.D. 1760-1835) Arikara infant long bone
sizes compared to the early (A.D. 1600-1753) postcontact period. These
smaller infants were explained in part by the negative effects of changing
settlement and subsistence patterns during the late postcontact period, in-
volving deterioration of the subsistence base, rapid depopulation, and large-
scale Arikara migration.
Mensforth (1985) also noted an inferior growth curve of Late Wood-
land (A.D. 800-1100) Ohio Libben preadolescents compared to Archaic
(B.C. 2655-3992) ones from the Carlston Annis Bt-5 shell midden in Ken-
tucky. But he explained these data in terms of increased population density
and sedentism in the Late Woodland period. These examples illustrate the
variability in interpretations of similar kinds of skeletal data, leading to
very different behavioral inferences.

Paleopathology
Much information about the structure of populations has been pro-
duced in attempts to understand and explain differing levels of health be-
tween populations. Many researchers, for example, have correlated high
frequencies of indicators of biological stress, such as porotic hyperostosis
and periostitis, with increases in population size, density, and sedentism
(Eisenberg, 1986, 1988; Larsen, 1993; Larsen and Thomas, 1982; Milner,
1992; Reinhard, 1992; Storey, 1986, 1992; Stuart-Macadam, 1992; Ubelaker,
1992a, b; Weaver, 1981). For example, a compelling case for the association
of health and population structure has been made by Walker (1986; also
see Lambert, 1993) for the Northern Channel islands of California. He
compared porotic hyperostosis frequencies across different sites from this
region spanning 5000 years and found the highest at the small, isolated
island of San Miguel. This island had limited fresh water and terrestrial
resources. Walker inferred that water contamination resulted from prob-
lems in waste disposal and the aggregation of a large number of people
around a limited water supply. This contamination, combined with the eth-
nographically documented practice of eating raw fish and sea mammal meat
would have increased susceptibility to diarrheal infections.
The paleopathological diagnosis of infectious diseases at prehistoric
Native American sites has led some researchers to speculate about popu-
lation structure and general living conditions (e.g., Fink, 1985; Pfeiffer,
1984). Diagnosis of one prehistoric Anasazi (A.D. 900-1300) child from
218 Boyd

Tocito, New Mexico, with probable tuberculosis prompted Fink (1985) to


speculate about a wide range of behavioral and environmental factors im-
portant to the spread of this disease for the Anasazi--inhabiting and work-
ing in damp, dark, closely spaced and poorly ventilated pueblos, improper
waste disposal, keeping of domestic dogs and turkeys near habitation areas,
nonsegregation of sick from the healthy, and overall population aggregation
and sedentism. While ethnographic descriptions of historic Anasazi lifeways
lend credence to this reconstruction, the inference of population structure
from the identification of tuberculosis in one prehistoric Anasazi child is
tenuous at best.
Intensification of agriculture has often been seen as the impetus for
changes in population structure and the resulting skeletal signatures. For
example, Arriaza (1993) noted an increase in inflammatory disorders of
the autonomous system across 8500 years of prehistory in northern Chile.
These conditions included ankylosing spondylitis, Reiter's syndrome, pso-
riatric arthritis, and rheumatoid arthritis. Noninflammatory diffuse idi-
opathic skeletal hyperostosis (DISH) frequencies also increased through
time. These changes mirrored agricultural intensification in this region. Ar-
riaza suggested that, although most of these conditions have a genetic com-
ponent, increased population size, density, and sedentism concomitant with
the subsistence change triggered their appearance and spread. But the pre-
cise mechanisms for this development are not known.
In contrast, no significant health decline was correlated with agricul-
tural intensification in the valley of Oaxaca. Hodges (1987) compared bio-
logical stress indicators of periostitis, porotic hyperostosis, and enamel
hypoplasia across nonintensive agricultural (ca. 1400-400 B.C.) and inten-
sive agricultural (ca. 500 B.C.-A.D. 1400) archaeological sites. No signifi-
cant increases in these patholgies were noted. She explained these results
partially in terms of the maintenance of a smaller population size and di-
versified diet combined with the absence of extreme sedentism. As many
researchers realized before her, Hodges (1987, p. 329) concluded, "The
decline in health observed with agricultural development cannot be attrib-
uted to dietary changes alone as sedentarization and increasing population
density, which usually accompany the shift to agriculture, can also affect
health."
These studies show that, while behavioral and environmental correlates
of health are often just as important as biological ones, the use of skeletal
markers to reconstruct these phenomena can be a risky business. Few stud-
ies of the skeletal effects of overcrowding, sedentism, and poor living con-
ditions have been conducted on modem or historic populations. Until more
is known about these interrelationships, inferences of population structure
from skeletal pathology will be difficult to evaluate.
Skeletal Correlates of Human Behavior 219

Bone Chemistry

Information derived from a chemical analysis of bone has been used


to infer migration of individuals between populations. Individuals with
unique stable isotope or trace element values compared to their peers have
been identified as foreigners, slaves, war captives, or other strangers in the
population.
For example, Schurr (1992) interpreted very low stable carbon isotope
values for one adult female from the late prehistoric (A.D. 1200-1450)
Middle Missouri Angel site of Southwest Indiana as a foreigner with little
or no maize consumption. Her companions at this site showed isotope val-
ues indicative of heavy utilization of maize. This individual's mortuary con-
text was a charred secondary burial found between the legs of an articulated
older male. In addition, two other females with unique mortuary contexts
showed lower carbon isotope ratios than the majority of the site's inhabi-
tants. Based on this mortuary and dietary evidence, Schurr constructed a
scenario suggesting that females were captured, recruited, or exchanged
from other groups for the purpose of additional labor input into this civic-
ceremonial complex. Similarly, Verano and DeNiro (1993) have noted in-
trapopulation carbon and nitrogen isotopes as well as craniometric
differences between prehistoric individuals at the Middle Horizon and Late
Intermediate period Peruvian site of Pacatnamu (A.D. 600-1400). The out-
liers were seen as possible war captives who were sacrificed at this site,
although pathological evidence for this victimization is absent.
At the Snake Hill military cemetery in southern Canada, Pfieffer and
Williamson (1991) and Katzenberg (1991a) used stable isotope bone col-
lagen to identify two soldiers from the War of 1812 with distinctly different
diets compared to their companions. These individuals had stable carbon
and nitrogen isotopes indicative of a "European-like" diet rich in C3 plants,
in contrast to the nitrogen-rich diet of meat and fish for other soldiers in
this sample. Possible Native American soldiers were also identified through
their diet. Katzenberg (1991a, p. 250) concludes, "It should not be sug-
gested that an individual is British, or Native American, based on stable
isotopes alone. However, there is justification in stating that when the diets
are similar; place of residence may also be." She adds that other lines of
evidence must be used in conjunction with these findings.
Are such inferences reasonable? While it is true that individuals from
different populations often exhibit diverse diets, this type of variability can
also be seen within populations. Food preferences, medical conditions, oc-
cupations, and a number of other behavioral, environmental, and physi-
ological factors affect the type of diet consumed. Katzenberg's cautions are
well advised.
220 Boyd

ACTIVITIES, ACITVITY LEVELS, AND OCCUPATIONS

Paleopathology and postcranial metric data have been particularly use-


ful to skeletal biologists in inferring past activities, activity levels, and even
occupations. These conclusions have been derived from populational as
well as individual samples.

Inferences About Activity Levels

Postcranial Metrics

Prehistoric and historic populational levels of activity have been di-


rectly investigated through biomechanical study of the postcranial skeleton.
This has been accomplished through a metric comparison of limb size and
shape, reflected by cross-sectional cortical areas and second moments of
areas of femoral and humeral diaphyses. Principles derived from engineer-
ing form the theoretical basis for these studies. The postcranial limb bone
is seen functionally as a beam, whose geometric properties of size and
shape reflect its general strength, rigidity, and capacity for resisting stress
(Ruff, 1992). Postcranial cross-sectional size reflects generalized mechanical
stress levels on bone, while differences in cross-sectional shape mirror types
of biomechanical loads. Computer-reconstructed cross-sectional cortical ar-
eas are a measure of a bone's strength, while second moments of areas
reflect bending and torsion (twisting) of the limbs. Increased populational
activity levels are correlated with this biomechanical stress, although body
proportion, weight, and posture are examples of other factors that may also
be involved (Ruff, 1992). Ruff (1992, p. 38) states,
...Beam model properties of an archaeological specimen should reflect the
mechanical Ioadings of the specimen while it was alive, and thus the biological and
behavioral characteristics of the individual that produced these loadings.

Biomechanical analyses of geometric properties of limb bones have


been conducted by a number of researchers. A primary focus of most of
these studies has been temporal variation in activity levels in conjunction
with subsistence, environmental, and sociopolitical change. Larsen, Ruff
and colleagues have illustrated this focus well. They recorded a decrease
in postcranial size and cross-sectional area for Georgia Bight coastal popu-
lations spanning 4000 years (Larsen, 1981, 1993; Larsen and Ruff, 1991,
1994; Larsen et aL, 1990; Ruff, 1992, 1994; Ruff et aL, 1984). These reduc-
tions are interpreted as reflecting decreased activity levels with the advent
of agriculture: Georgia coast agricultural populations generally became less
Skeletal Correlates of Human Behavior 221

mobile and engaged in less long-distance travel and running compared to


preagricultural ones. Geometric bone strength indicators increased with the
onset of European contact, mirroring increased labor demands placed on
Native Americans by the Spanish. Ruff and Larsen (1990) were even able
to identify a subsample of five early Colonial period Native American males
with exceptionally large cross-sectional femoral areas. They believe that
these individuals may have been conscripted for long-distance travel as part
of the Spanish "repartimiento" labor system. Ruff (1992, p. 41) concluded
that "it is sometimes possible to discern fairly specific and relatively subtle
behavioral differences with this kind of analysis."
Similar biomechanical studies have been performed for skeletal popu-
lations in other regions, establishing a great deal of regional variability. For
example, Brock and Ruff (1988) recorded increases in postcranial geomet-
ric dimensions across the Anasazi and Mogollon Early Village (A.D. 500-
1150) and Abandonment (A.D. 1150-1300) periods of the American
Southwest, reflecting increased activity demands for these populations. This
trend reversed, however, later in the Aggregated Village period (A.D.
1300-1540), particularly for males. Cole (1994) found little evidence for
changes in activity levels through time in the Plains. A comparison of 15
linear femoral and tibial measurements across a period of 1200 years re-
vealed few differences. Cole concluded that the behavioral modifications
concomitant with the shift to agriculture may have been insufficient to leave
functional evidence on the skeleton, owing to the fact that extreme maize
dependence was not seen in this area.
Bridges (1985, 1989a, 1991b) showed that Pickwick Basin, Alabama,
agricultural populations (A.D. 1200-1500) had postcranial areas and polar
moments of inertia that were larger than those for the Archaic preagricul-
tural (6000-1000 B.C.) populations. Thus, these agriculturalists showed in-
creased bone strength due to a hypothesized increase in work-related
agricultural activity demands. She concluded that there is no single, uniform
biomechanical response to the development of agriculture. The great
amount of regional variability in these activity level changes suggests a num-
ber of variables affect their bony expression, including degree of agricul-
tural intensification, type, duration, and severity of activities engaged in as
well as nutritional and demographic features of the population.
Sexual dimorphism in populational activity levels has also been ex-
plored through biomechanical postcranial metric analysis. In the lower limb,
the ratio of anteroposterior/mediolateral (A-P/M-L) bending strength in the
femoral and tibial midshaft has exhibited gender-related temporal variabil-
ity. Ruff (1987) as well as others have noted a decrease in this bending
strength sexual dimorphism since the Middle Paleolithic. Prehistoric hunter
and gatherer males show greater bending strengths compared to females,
222 Boyd

while modem industrialized population bending strengths are more similar


between the sexes. The inference which follows from this evidence is that
modem males engage in less biomechanically demanding activities such as
running and long distance travel compared to their prehistoric counterparts.
Prehistoric agricultural populations like Pecos Pueblo were intermediate
between foragers and more modem samples, leading Ruff to suggest that
Pecos males and females engaged frequently in sex-specific behaviors re-
quiting different lower limb bending loads compared to modern popula-
tions.
For the upper limbs, changes in humeral bilateral asymmetry across
the sexes through time has been a focus of research on Georgia Bight
(Fresia et aL, 1990) and Pickwick Basin, Alabama, populations (Bridges,
1985, 1989a, 1991b). In both of these regions, bilateral asymmetry in
humerus size and strength declined more significantly in females than males
with the advent of agriculture. Bridges (1985) inferred an increased par-
ticipation of females in agricultural activities compared to hunting and
gathering ones. In contrast, it was suggested that males had similar upper
limb activity levels in the hunting and gathering and agricultural periods.
In the Georgia Bight, European contact appeared to have a greater effect
on male rather than female humeral size and strength--bilateral asymmetry
decreases in males were more significant. Fresia et al. (1990) interpreted
this finding as evidence that males increased their participation in agricul-
tural activities with the arrival of forced labor in the Spanish missions.

Inferences About Activities and Occupations

Paleopathology

Many researchers have attempted to correlate skeletal pathologies with


specific activities or even occupations (Angel et aL, 1987; Edynak, 1976;
Hill, 1994; Merbs, 1983; Stirland, 1991; Trinkaus, 1975; Walker and Holli-
man, 1989). In 1989, Kennedy compiled a list of over 140 of what he termed
"occupational markers" in his review of past research on this topic. For
example, frequent horseback riding has been correlated with a complex of
skeletal features, including superior elongation of the acetabulum, enlarged
gluteus medius and gluteus minimus attachments, and osteoarthritis of the
first metatarsal due to the frequent use of toe stirrups (Bradtmiller, 1983;
Kennedy, 1989; Reinhard et al., 1994).
As early as 1935, Hrdlicka began to suspect that auditory exostoses
were possibly correlated with behavior. Kennedy (1986) noted higher fre-
quencies of these bony ear tumors in populations engaging in swimming
Skeletal Correlates of Human Behavior 223

and diving in cold water. More recently, Munizaga (1991) has documented
hyperextension and hyperflexion of the occipitoaltoidal joint leading to su-
pernumerary condyle formation and arthritis at the base of the skull in
conjunction with this same behavior.
Skeletal evidence for activities and occupations has primarily come
from three sources--dental pathology and wear, hypertrophy of bony mus-
cle attachments, and degenerative joint disease. Research on dental pathol-
ogy and wear has often focused on habitual use of teeth as tools. Milner
and Larsen (1991) and Larsen (1987) have reviewed this evidence. The
presence of unique patterning of dental grooves, notches, and wear on and
between prehistoric and historic American teeth have often been inter-
preted as evidence for processing of plant fibers for consumption of manioc
or the manufacture and use of baskets, fishing nets, cordage, sinew, or cloth
(Blakely and Beck, 1984; Hartnady and Rose, 1991; Irish and Turner, 1987;
Larsen, 1985; Larsen and Thomas, 1982; Molnar, 1972; Rathbun et al.,
1980; Schulz, 1977). Dental microfractures in Eskimo dentitions have been
correlated with use of their teeth to crush bone (Turner, 1979).
Food preparation behaviors have also been suggested through dental
analyses of prehistoric Native Americans. Powell (1985) noted greater rates
of dental attrition for hunter-gatherers from several sites representing the
Fourche Maline Focus in southeastern Oklahoma compared to a late pre-
historic Caddoan agricultural skeletal sample from Arkansas. She corre-
lated this higher attrition with the use of grinding stones. The hunting and
gathering sample had lower caries rates as well, partly from the lower car-
bohydrate diet, but also due to the "prophylatic" cleaning effects of grit
between the teeth. At Lower Pecos, Texas (8000 B.C.-A.D. 1000), Hartnady
and Rose (1991) explained the high number of middle-aged prehistoric
edentulous maxillas and mandibles in terms of food processing techniques.
Use of limestone manos and metates and sotol baking pits introduced grit
as well as ash into the dentition, resulting in increased wear and loss of
teeth. Smith et al. (1980) correlated decreases in anterior tooth size through
time in the Tennessee valley partly with decreased attritional stress. This
in turn may reflect the more infrequent use of mortars and grindstones in
the Mississippian period. In sum, dental pathology and wear data often
directly reflect masticatory and paramasficatory activities. But greater em-
phasis on actualistic studies are needed to confirm such relationships.
Hypertrophied supinator crests and deep supinator fossae of the ulna
have been interpreted by Kennedy (1983, 1989) as the result of habitual
throwing of missile weapons like the aflatl and spear. These conditions oc-
cur very frequently in prehistoric and historic hunters and gatherers be-
lieved to have used such missiles. Evidence in support of this association
comes from the realm of sports medicine, where athletes habitually using
224 Boyd

their brachial skeleton for throwing show well developed crests and deep
fossa. Over 40% of professional baseball pitchers also manifest bony spurs
on the medial surface of the ulnar notch believed to be associated with
throwing stresses (Kennedy, 1983). Hypertrophy of the ulnar supinator crest
was observed in historic slaves from the Eastern coast and was correlated
by Kelley and Angel (1983, 1987) with elbow-extending twisting movements
similar to those informally observed by these authors in individuals using
an ax to cut trees. These examples show that, while hypertrophied bony
muscle attachments do directly reflect greater use of these muscles, the
specific behaviors attributed to them may be subjective. Many different be-
haviors can result in the same skeletal signature.
More indirect evidence for activities consists of osteoarthritis and
osteophytosis data indicative of degenerative joint disease. For example,
osteophytosis of cervical vertebrae has been correlated with extreme hy-
perextension and stress of the neck. This pathology is consistent with the
possible use of a tumpline to carry heavy burdens (Bridges, 1994; Merbs
and Euler, 1985). Bridges (1994), in fact, noted that cervical osteophytosis
is much more common in the prehistoric eastern United States than in the
West. Historic and prehistoric records of tumpline use in eastern North
America are consistent with this evidence. However, Bridges was perplexed
by the prevalence of osteophytosis in both males and females, when it is
known ethnohistorically that females were the primary burden carriers. She
concluded that males must have engaged in similar activities that stressed
the neck region.
Examination of historic period skeletal samples for degenerative joint
disease has helped substantiate interpretations of past activity patterns,
when supplemented with the historic record. The frequent degeneration of
skeletal joints at an historic (1840-1870) South Carolina plantation
(38CH778) led Rathbun (1987) to conclude that hard physical labor was
ubiquitous for this African-American population. Sexual dimorphism in the
patterning of these pathologies was evident--males showed degeneration
of the elbow and hip, while females were primarily afflicted in their knees
and shoulders. Schmorl's nodes (intervertebral disk herniations) were com-
mon for males, suggesting frequent heavy lifting. Similar evidence and con-
clusions have been noted for historic slaves from Maryland, Virginia, and
the Carolinas (Kelley and Angel, 1983, 1987), frontier soldiers from Colo-
nial period Ft. Laurens, Ohio (SciuUi and Gramly, 1989), and War of 1812
soldiers from Snake Hill in southern Canada (Owsley et al., 1991). For the
latter example, historic descriptions of the rigorous military training and
demanding lifestyles that these soldiers endured accord with the skeletal
evidence.
Skeletal Correlates of Human Behavior 225

Degenerative joint disease has also been used to infer activities related
to food processing. "Metate elbow" has been described by many re-
searchers. For example, Merbs and Euler (1985) associated elbow degen-
eration in the flexion-extension aspect with habitual corn grinding by one
Anasazi female from Bright Angel Ruin. Miller (1985) found evidence for
lateral epicondylitis in the elbow of both males and females from the pre-
historic (A.D. 1000-1400) Arizona site of Nuvakwewtaqa (Chavez Pass).
Bilateral expression of this trait led him to suggest the use of two-handed
metates by both sexes.
Occupation has even been inferred by some researchers through de-
generative joint disease studies. In a comparison of 29 urban slave skeletons
from the first New Orleans cemetery (A.D. 1725), Owsley et al. (1987) were
able to differentiate laborers from house servants. House servants, primarily
females and older males, were believed to be those exhibiting less degen-
erative joint disease and bone hypertrophy. Laborers were those exhibiting
substantial levels of degenerative arthritis and bone hypertrophy, indicative
of "high levels of physical labor and strain perhaps attributable to manual
labor on the docks of this busy shipping port, or as workers on the canals
or levees" (Owsley et aL, 1987, p. 195).
Kelley and Angel (1983, 1987) documented occupational stresses of
slaves from 25 Maryland, Virginia, and North and South Carolina historic
sites. An individual with lipping and eburnation of the shoulder joints and
vertebrae in combination with well developed phalanges was designated by
Kelley and Angel as a skilled craftsman. Another individual manifested an
arthritic elbow, possibly due to strain on the triceps in digging out ore from
banks or pounding out pig iron.
Merbs' (1983) classic study of the historic Saddlermuit Eskimo represents
the most detailed investigation of activity markers to date. He used a com-
bination of all three of these sources--dental pathology and wear, bone hy-
pertrophy, and degenerative joint disease--to reconstruct 20 activity patterns
for this group that became extinct in 1903. For example, high frequencies of
temporomandibular joint osteoarthritis in Saddlermuit females were corre-
lated with the use of their dentition, particularly the left side, to soften hides.
Osteoarthritic patterning of their wrists as well as other postcranial areas re-
flected the scraping, cutting, and sewing of these skins. Osteoarthritis in the
brachial area of males related to throwing of harpoons as well as kayak pad-
dling. And high frequencies of vertebral compression in both sexes was at-
tributed to sledding and tobogganing. While these claims were supported by
ethnographic descriptions of Saddlermuit activities and by archaeological
(toolkit) remains, Merbs acknowledged problems involved in these correla-
tions. Specific tasks such as harpoon throwing or skin scraping involve distinct
and limited joint movements and may be more easily and accurately identified
226 Boyd

compared to more general and "multijoint" tasks such as driving a sled. The
frequency of the activity engaged in is important--laboratory animal experi-
mentation has shown that activities placing normal stress on joints for ab-
normally long periods of time or abnormally heavy stress for short periods
of time are likely to result in joint degeneration. But, in spite of this estab-
lished relationship, "...the 'goodness of fit' between patterns of activity and
those of pathology usually cannot be measured with great precision" (Merbs,
1983, p. 159). And many behaviors, abnormal in neither duration nor fre-
quency, but that comprise the majority of a person's day, may not even be
discernible in terms of a skeletal signature.
Can inferences such as these be made with confidence? While some
researchers maintain that evaluation of the merit of skeletal activity or oc-
cupational markers must await further research (Kennedy, 1989), others
have been more critical. Jurmain (1990), Stirland (1991), and Waldron
(1994) have asserted that it is simply not possible to identify specific activities
or behaviors that might have resulted in a particular skeletal modification
for an individual. Degenerative joint disease, more specifically osteoarthritis,
is especially problematic. This is a multifactorial disease, influenced by such
diverse factors as age, sex, weight, race, genetic predisposition, trauma, and
type, frequency and duration of an activity (Bridges, 1989b, 1991a, 1993;
Knusel, 1993; Waldron, 1994). Thus, a one-to-one correspondence between
osteoarthritis patterns and activities or occupation should not be expected
and in fact is not generally found in clinical settings. In some cases, for
example, coal miners have high frequencies of spinal osteoarthritis, while in
other cases, they do not (Waldron, 1994). Occupational pathology may leave
little evidence on the skeleton and, conversely, several different occupations
may leave the same skeletal signature. And, osteoarthritis may occur in the
absence of any identifiable occupational stress.
Stirland (1991) noted a variety of skeletal pathologies for the 16th-
century sailors aboard the sunken English vessel Mary Rose, including spon-
dylotysis of the vertebral column, osteochondritis dessicans, and
hypertrophy of major muscle attachments. In spite of these indications of
occupational stress, Stirland (1991, p. 46) writes, "In an archaeological sam-
ple, it will never be possible to extrapolate from the general to the par-
ticular and assign an individual's occupation from a group study. In a
personal sense, I will never be able to say: 'This man was an archer.'"

VIOLENCE, WARFARE, AND DEATH

Osteological evidence for violence, warfare, and death has only re-
cently been systematically investigated (Frayer and Martin, 1996; Hutchin-
Skeletal Correlates of Human Behavior 227

son, 1990, 1991, 1993; Larsen and Huynh, 1993; Milner, 1995; Milner et
al., 1991). Skeletal documentation of these events has become more so-
phisticated (e.g., Olsen and Shipman, 1994) and once-accepted examples
of these behaviors (Blakely 1988; Blakely and Mathews, 1990) are now be-
ing reexamined and questioned (Milner et aL, 1994). The goal of some
researchers (e.g., Haas and Creamer, 1993; Merbs and Birkby, 1985) has
been simply establishing the presence of violence at an archaeological site.
Other osteologists have made a variety of behavioral inferences about in-
dividuals as well as populations from evidence of skeletal trauma. These
include such aspects as perimortem or postmortem treatment of the dead,
and the nature and variation of conflict and warfare between or within
populations. Much of this skeletal evidence has been derived from the field
of paleopathotogy.

Inferences About Mortuary Treatment of the Dead

Most information regarding perimortem or postmortem mortuary


treatment of individuals has come from archaeological sources. Some be-
havioral inferences can be gleaned, however, from morpholggical evidence
preserved on skeletal remains. Strong evidence for the existence of body
processing rituals, religious practices, and even cannibalism has been pre-
sented by many researchers.
Cutmarks even on small fragments of bone may reveal much about
how and perhaps why a body was processed after death (Blick, 1990; Boyd
and Boyd, 1992; Fenton, 1991; Owsley et al., 1994). For example, Olsen
and Shipman (1994) conducted pilot SEM studies of prehistoric cutmarks
on Native American Plains Indian bone spanning the Middle Woodland
through Disorganized Coalescent periods (A.D. 400-1832). Comparisons
of these cutmarks to experimentally induced modifications of bone from a
variety of cultural and natural sources allowed these authors to differentiate
funerary cuts from confict-related trauma. Patterning in defleshing and dis-
articulation cuts indicated that defleshing occurred at various intervals be-
fore secondary burial. In that interim, variations in bone color and texture
due to weathering suggest that bodies were exposed to sunlight, probably
through placement on scaffolds. In some cases, this exposure occurred be-
fore cleaning; in other cases, it occurred after. Olsen and Shipman even
entertained the distinct possibility that handedness could be inferred from
the orientation of funerary cutmarks. In short, they were able to reconstruct
mortuary treatment of these individuals based on actualistic studies of bone
modification processes. A similar reconstruction of the mechanical proce-
dures involved in the disarticulation and defleshing of individuals from the
228 Boyd

prehistoric (A.D. 1000-1300) southeastern Michigan site of Riviere aux


Vase has been completed by Raemsch (1993) based on surface analysis of
cutmark patterning.
Identification of carnivore damage to human bone has led to the in-
ference that individuals were left above ground for some time before burial.
Extensive evidence for postmortem animal-induced damage to many of the
skeletons from Crow Creek (A.D. 1325) in South Dakota led Willey (1990;
Willey and Emerson, 1993; Zimmerman and Bradley, 1993) to suggest that
these massacre victims were left exposed above ground for some time be-
fore their eventual collection and burial. This carnivore damage was similar
to that produced under modem experimental conditions with wolves and
also fit with carnivore damage to modem forensic cases (Snyder and Willey,
1989). Similar inferences have been made by Milner et al. (1991) and Mil-
ner and Smith (1989) for the Norris Farm cemetery in Illinois (A.D. 1300),
Owsley et al. (t977) for the Larson site in South Dakota (A.D. 1750-1785),
and Sciulli and Gramly (1989) for the Ft. Laurens, Ohio, frontier skeletal
sample. Milner et al. (1991) were able to effectively distinguish human-in-
duced trauma from animal scavenging marks and identify these scavenging
marks as canid. Once again, comparison of the Norris Farm trauma pattern
to that of canid-damaged modem forensic cases (Hagtund et al., 1988, 1989)
aided in interpretation of this evidence.
Differential burial of segments of populations has been suggested by
Buikstra (1981) and Danforth et al. (1994) through their paleopathological
research. Buikstra (1981) compared pathology and mortuary data across
skeletal samples from the Middle Archaic Koster and Gibson sites and the
Modoc Rock Shelter and found that individuals with frequent and severe
pathologies were buried in habitation areas. "Normal" adults, those exhib-
iting less frequent and severe pathologies, were buried in bluff crest ceme-
teries, while infants and young children were apparently buried in an
unknown location. She concludes that individuals' pathological status, and
their ability to perform normal tasks, may have determined their burial
location. Danforth et al. (1994) also suggested possibly differential burial
of 34 individuals from Carter Ranch Pueblo (A.D. 1100-1225) in Arizona.
These individuals manifested high pathology frequencies, including genetic
abnormalities like Klippel-Feil syndrome and premature suture closure.
They perceived this small Pueblo community as existing in relative genetic
isolation. High trauma frequencies, especially, "suggest a disability-related
access to burial at Carter Ranch Pueblo" (Danforth et aL, 1994, p. 97).
Ritual has also been inferred from skeletal pathology. Puncture marks
on human bone have been used by Torbenson and coUeages (Torbenson et
aL, 1992) as an indication of religious ritual. They believe that postmortem
perforations seen on over 150 postcranial limb bones from the Laurel cul-
Skeletal Correlates of Human Behavior 229

ture Smith Mound Four (A.D. 565) in Minnesota reflect attempts at release
of the soul. Ethnographic evidence suggesting the importance of bone as
the seat of the human soul for many Native Americans offered support for
this hypothesis.
Inferences about prehistoric ritualistic cannibalism of human remains
have also been made from skeletal pathology. In 1993, Turner reviewed
the evidence for Anasazi cannibalism. He compiled over 40 possible cases
involving at least 472 individuals. These remains exhibited the "minimal
taphonomic signature of probable cannibalism": burning, cutmarks, anvil
or hammerstone abrasions, deliberate bone breakage, and missing verte-
brae (Turner et aL, 1993, p. 83; Turner and Turner, 1992). Because of the
increased evidence for this behavior in the American Southwest, he aban-
doned his "social pathology" theory for its existence. Instead, he believes
that Anasazi cannibalism was part of a Mesoamerican-like system of insti-
tutionalized violence. As Anasazi social organization became more com-
plex, the need for cultural enforcing mechanisms such as terrorism, raids,
and cannibalism became great. He cites concurrent dates for Anasazi cul-
tural florescence and many "cannibalized sites" as support for his theory.
But dates for many other "cannibalized sites" cover at least 800 years of
prehistory, indicating a fairly great time depth for this proposed cultural
practice (White, 1992).
Perhaps the strongest evidence for prehistoric cannibalism has been
the use of experimental studies on modern faunal assemblages to identify
butchering patterns on bone. The use of a "faunal analogy," according to
White (1992), is the only way to recognize cannibalism in the archaeological
record. He documented cutmarks, chopmarks, fracture and crushing pat-
terns, as well as heat damage and pot-polish on the 12th century human
skeletal assemblage from Mancos and correlated these patterns with the
very similar damage seen on faunal remains butchered and consumed. The
pattern of damage to the Mancos skeletal assemblage was consistent with
the motive of "nutritive extraction" seen in faunal remains, rather than
mortuary or conflict-related alterations. Removal of bone grease from
spongy bone appears to have been one objective, according to White, al-
though he hesitated in attributing this behavior to an ultimate motivation
such as hunger or warfare. Greater application of forensic anthropology
crime-scene techniques are recommended as one way to aid identification
and interpretation of such assemblages in the future.
Inferences of cannibalistic behavior from human skeletal morphology
have by no means gone unchallenged. Bullock (1991) was very critical of
the faunal analogy used by White and Turner. He maintained that it is not
a valid approach because it fails to consider cultural variability, including
human motivations in behavior. In other words, analogous patterning in
230 Boyd

human and faunal remain modification is not adequate "proof" of canni-


balistic intent by prehistoric humans. As an alternative, Bullock compared
the skeletal evidence for Anasazi cannibalism to that seen on Little Big
Horn battle casualties and inferred that the Anasazi remains represent the
results of warfare-related death, corpse mutilation, and possibly even mor-
tuary (defleshing) practices. Bullock (1991, p. 12) believes that there has
been a bias amongst scientists, particularly in regard to the Anasazi, "to-
wards identifying behavior as deviate or abnormal, rather than recognizing
variability and depth within cultural manifestations."

Inferences About Warfare and Conflict

By noting not just the presence or absence of trauma to bone, but the
frequency, location, severity, and archaeological context of this trauma,
much has been inferred about aggression and warfare (Milner, 1995; Smith,
1993b, 1996). Inter- and intrapopulation as well as spatial and temporal
patterning in these aspects of trauma has allowed researchers to infer the
possible cause, nature, and duration of conflict as well as common features
of its intended victims. Examples of many of these types of behavioral in-
ferences follow.
Correlation of paleopathology frequencies with demographic informa-
tion of age and sex has allowed some researchers to identify "at-risk" seg-
ments of populations more likely to become victims of intentional trauma.
In many of these cases in the prehistoric Americas, this segment has been
gender specific. For example, all eight individuals from the Late Archaic
(3000-5000 B.P.) Ward site in western Kentucky who showed unhealed peri-
mortem cutmarks indicative of intentional trauma were male (Mensforth
and Baker, 1995). And in the Kentucky Lake Reservoir of the western Ten-
nessee valley, only males (n = 10) manifested perimortem trauma sugges-
tive of warfare in the Archaic period (Smith, 1996b).
A pattern of female-directed violence was observed by Wilkinson and
van Wagenen (1993) at the Late Woodland (A.D. 1000-1300) Riviere aux
Vase site in Michigan. Of 19 healed cranial traumas documented for this
skeletal sample of 370 individuals, 15 occurred on reproductive-aged fe-
males. These injuries were seen on all major bones of the cranial vault,
although those found on the rear and sides of the skull were most severe.
Frequent cranial, but infrequent postcranial, traumas suggested to these
authors that these injuries were not the result of accidents. In their recon-
structions of behavior leading to these traumas, Wilkinson and van
Wagenen wisely exercised caution. They speculated about such sources as
intrasexual competition among cowives, wife-beating, and abduction, tor-
Skeletal Correlates of Human Behavior 231

ture, and eventual adoption of foreign women. Ultimately, they concluded,


no single cause is knowable, but ethnographic evidence for the frequent
capture of females during raids between different Native American groups
suggested a likely source for this female violence. Similar documentation
of female-directed violence, involving parry fractures and craniofacial
trauma, has been recorded for the prehistoric Anasazi by Martin and Akins
(1994).
The direct correlation between gender and pathology frequencies may
be misleading, however. Smith (1996) interpreted the high frequency of
female parry fractures from the Late Archaic (2500-500 B.C.) Eva site in
West Tennessee to be the result of sampling bias instead of female-directed
violence. These midshaft fractures of the radius and/or ulna are often seen
as skeletal signatures of interpersonal violence. While they appeared at Eva
to affect females disproportionately, this pattern was shown through statis-
tical comparisons to be simply a reflection of the greater number of females
at this site. The absence of significant female cranial trauma supported this
conclusion. This study shows the importance of carefully analyzing paleo-
pathology data within the context of other skeletal and archaeological in-
formation to eliminate possible biases in interpretation.
Evidence suggesting the existence of warfare or inter- and intragroup
conflict can be categorized as direct or indirect. Direct evidence consists
of physical remnants of aggressive behaviors. For example, a total of 10
embedded obsidian points was found in the skeletons of nine individuals
from Ala-329, a prehistoric (A.D. 500-contact) shell mound from Central
California (Jurmain, 1991). This represents one of the highest incidences
of embedded points for any skeletal sample in North America. All but one
of the wounds at this California site occurred in the abdomen or lower
thorax, leading Jurmain to suggest that these injuries were sustained at
close range with the victims restrained. He believes that these individuals
may have been executed.
More indirect suggestions of warfare and violence are extrapolated
from skeletal documentation of decapitation, dismemberment, scalping, and
trophy taking. All of these behavioral inferences depend on the differen-
tiation of cutmarks attributable to funerary versus violent trauma activities.
Scalping results in highly diagnostic and often readily recognized circum-
ferential cranial cutmarks (Owsley, 1994; Smith, 1996b). Decapitation, dis-
memberment, and trophy-taking may be identified by the absence of a body
segment in an otherwise well-preserved and undisturbed individual, tn ad-
dition, the presence of circumferential cutmarks adjacent and restricted to
the missing segment offers additional support for this inference (Smith,
1996b). Just such an approach has been used by Smith (1995; 1996b) and
Mensforth and Baker (1995) in their documentation of warfare-related rio-
232 Boyd

lence of Archaic inhabitants of western Tennessee and western Kentucky,


respectively. Smith (1996b) was able to differentiate cutmarks related to
mortuary positioning (flexure) from those associated with scalping and tro-
phy-taking in a sample of one Middle Archaic (6000-3500 B.C.) and seven
Late Archaic (2500-500 B.C.) sites from the Kentucky Lake Reservoir in
Tennessee. She inferred a motive of prestige enhancement for the trophy-
taking. This represents one of the earliest examples of such activities in
the Southeast and has important implications for understanding the evo-
lution of societal complexity in this region. Mensforth and Baker (1995)
recorded similar evidence for projectile point injuries, stab wounds, cranial
trauma, parry fractures, and cutmarks related to scalping, dismemberment,
and decapitation at the Late Archaic Ward site in western Kentucky.
Inferences about prehistoric violence and aggression have even been
made in the absence of skeletal evidence for these behaviors. Blakely's
(1977) demographic reconstruction of prehistoric inhabitants of Etowah,
Georgia, showed a prevalence of young females and older males in Mound
C. He (1977, p. 61) concluded that females must have derived their status
through marriage and that "logic leads to the conclusion that many females
were sacrificed at the death of their husbands." Saul and Saul (1989) sug-
gested the possibility that 11 male Maya skeletons found under a ball court
at Seibal were sacrificial members of a losing ball team. In both of these
cases, no direct evidence of traumatic violence was manifest on the skele-
tons. While this may reflect the absence of permanent skeletal signatures
of these violent episodes, it may also reflect the absence of such violence
altogether. As such, this negative evidence represents a very weak basis for
behavioral inferences regarding violence and aggression. Fowler (1984) has
also inferred ritual sacrifice of a group of 33 Late Pre-Classic (ca. 100 B.C.-
A.D. 100) individuals from Chalchuapa, E1 Salvador, in the absence of di-
rect skeletal indications of trauma. But their distinctive mortuary treatment,
demographic composition (most, if not all, were male), and absence of
some skeletal elements offered supporting evidence for this scenario. Even
so, a sacrificial context for this population remains conjectural without
more direct skeletal evidence for this activity.
Spatial and temporal patterning of violent trauma evidence has al-
lowed some researchers to reach conclusions about the social and cultural
context of warfare and conflict. For example, inter- versus intragroup ag-
gression has been distinguished on the basis of the patterning of skeletal
trauma. Owsley (1994) found greater incidences of Coalescent period (A.D.
1600-1832) scalping in the Grand-Moreau region of the northern Plains
compared to the southern Plains. Since these northern Plains sites were
on the periphery of the ANkara territory and were highly fortified, he con-
cluded that most Plains warfare during this time period was intertribal.
Skeletal Correlates of Human Behavior 233

Mensforth and Baker (1995) attributed less severe and healed blunt
cranial trauma at the Ward site to intragroup "game-related" aggression,
while more serious unhealed trauma was seen as the lethal consequence
of intergroup conflict. However, scenarios for intragroup serious injury and
intergroup minor injury could also be envisioned.
Many researchers have been able to document the diachronic charac-
ter of skeletal trauma reflecting considerable time depth to human violence
in the Americas (Rathbun, 1993; Seeman, 1988; Smith, 1993a, b, 1995). At
the late prehistoric (A.D. 1300) cemetery of Norris Farms in west-central
Illinois, skeletal evidence for interpersonal violence was widespread, but
episodic (Milner et al., 1991). There were no mass graves or hurried burials.
Instead, graves were orderly aligned, presumably reflecting excavation of
pits and interment of bodies at different times over several decades. The
43 individuals (of a total 264) manifesting unhealed trauma and mutilation
varied in terms of the amount of carnivore damage to their bone as well
as the completeness and mortuary context of their remains. This variability
suggested that the interval from death to interment differed across the
cemetery. Evidence for attacks on small groups, possibly work parties, came
from six examples of multiple burials. Most of these were single-sex (male
or female) victims similarly articulated and buried synchronically. Milner
et al. (1991; Milner, 1995) concluded that interpersonal conflict at this site
was regular and violent, consistent with ethnographic descriptions of small-
scale warfare and raiding by neighboring groups. They also hypothesized
that this pattern of retaliatory attacks may have significantly disrupted vil-
lage functioning, particularly subsistence endeavors.
Similar observations of skeletal trauma and mortuary variability have
been made by Jurmain (1991) for the prehistoric California shell mound
Ala-329 and Sciulli and Gramly (1989) for the Colonial (A.D. 1779) Ft.
Laurens, Ohio, skeletal sample and have led these authors to similar con-
clusions about the nature of these warfare incidences--small-scale, inter-
mittent, but fatal. In the latter case, historic records detailing Native
American attacks on this fort over a period of 9 months, resulting in at
least 21 deaths, vindicate this depiction of frontier warfare.
Placing violence-related trauma frequencies in temporal perspective
has also allowed researchers to assess change in warfare through time. In
the Plains, considerable time depth has been established for the practice
of scalping. The existence of this practice throughout the Coalescent tra-
dition led Owsley (1994, p. 341) to conclude that "...the roots of Plains
Indian warfare are deep in the prehistoric past", not simply a result of
European contact. In the northern Plains, Olsen and Shipman (1994) re-
corded an increase in scalping, cutmarks, projectile point wounds, and other
conflict-related trauma spanning approximately 1400 years (A.D. 400-
~4 B~d

1832). They attributed this increase to greater intertxibal warfare owing to


competition for local resources instead of the interactions of Europeans,
since much of this evidence predates European contact.
Similar increases in violent trauma through time have been recorded
or noted by Turner et al. (1993) for the prehistoric Anasazi in the American
Southwest, Walker (1989; Lambert and Walker, 1991) for the prehistoric
Northern Channel islands of California, and Milner (1995) for a wide geo-
graphic region encompassing the Midcontinent and eastern North America.
In most of these cases, increased frequency of violent conflict in the late
prehistoric period has been interpreted as at least partially due to larger,
more sedentary and complex populations competing for spatially restricted
resources.
In summary, skeletal pathology can offer perhaps the most direct and
detailed behavioral inferences concerning warfare, aggression, and mortu-
ary treatment of the dead. This is not surprising given that many of these
activities leave immediate and permanent signatures on the human skele-
ton. Even so, these inferences need to be supported by actualistic studies
of the effects of pre- and postmortem alteration of human bone. For ex-
ample, taphonomic studies on human bone similar to those by Lyman
(1994) for nonhuman vertebrate skeletons can offer a wealth of compara-
tive data regarding the identification and interpretation of cutmarks, animal
modification, and other pre- and postdepositional processes affecting bone.
Studies reviewed here that utilized such comparative data produced the
strongest behavioral inferences.

SUMMARY AND CONCLUSIONS

Goodman (1991; Goodman et aL, 1988; also see Bush, 1991) has re-
cently asserted that skeletal biologists "quit too soon" in their interpretation
of skeletally derived biological data by not exploring the biocultural and
behavioral connotations of their results. Other researchers (most notably,
Dettwyler, 1991; Leone and Palkovich, 1985; Waldron, 1994) claim that we
have overstepped the boundaries of our data in our behavioral reconstruc-
tions. In light of the behavioral inferences discussed in this paper, then,
have osteologists gone too far or not far enough?
Consideration of this important question centers around evaluating the
accuracy and reliability of behavioral interpretations in a scientific context.
Behavioral inferences should be framed as hypotheses to be evaluated
through empirical testing. More often than not, however, these inferences
are attached as afterthoughts to the standard skeletal analysis and, as such,
cannot be scientifically evaluated. They are by and large ad hoc or post
Skeletal Correlates of Human Behavior 235

hoc arguments that are not clearly operationalized in terms of their em-
pirical signatures. Thus, many of these inferences are simply speculations,
and no more. As Ortner (1991, p. 11) states, "While enlightened specula-
tion may be helpful, it is very easy to careen down scientifically blind alleys
because of ignorance or because we have overextended our data."
The behavioral inferences discussed in this paper vary tremendously
in terms of the confidence we can place in them precisely because of this
variability in the use of a clear-cut scientific methodology. This can be seen
in at least four ways. First, when behavioral hypotheses are constructed,
oftentimes they are inherently tmtestable. For example, how does one em-
pirically operationalize compassion or altruism in a prehistoric population?
Such conjectures of emotional motives may not be discernible from an ex-
amination of skeletal morphology. And as Dettwyler (1991) and Leone and
Palkovich (1985) suggest, we may be incorrectly applying modern perspec-
tives to past situations.
Second, when testable hypotheses are generated, alternative hypothe-
ses are often not considered. For example, conclusions reached concerning
the existence of "foreigners" within a population from their unique diet
fail to consider other possibilities for these differences. Individual dietary
preferences, occupational or social differences, physiological factors, or
sampling bias could explain intrapopulational dietary variation. More se-
cure conclusions are reached when alternative hypotheses are considered
and tested.
Third, in many cases, the theoretical bridge between the behavioral
inference and empirical data supporting that inference is absent or weak.
For example, how does one reliably establish matrilineal descent from the
presence of labret scars on prehistoric dentitions? And how does one dif-
ferentiate ascribed from achieved status by means of intrapopulation pa-
thology differences? The skeletal data seen in these and other examples
are inadequate for assessing these behavioral inferences alone.
Finally, behavioral inferences are only as reliable as the methods used
to generate them. Recent criticisms of trace element analysis weaken in-
ferences about diet and social organization derived from this method. Stud-
ies based on nonmetric skeletal variation may or may not illuminate
differences in social organization and population structure. Paleopathology
is an inexact science. Many pathological states leave no lasting effect on
bone, and many different pathologies may leave identical signatures. Thus,
correlations between specific occupations and skeletal pathology may be
highly variable. Many activities engaged in by individuals throughout much
of their day may not be manifested in skeletal morphology at all.
The most secure conclusions about behavior were those for which hu-
man behavior left direct effects on bones and teeth. Dietary reconstructions
236 Boyd

from stable isotope ratios and dental pathologies, evidence for warfare and
violence from skeletal trauma, and levels of activity and mobility from post-
cranial biomechanical geometry offered the most scientifically replicable
and testable interpretations. In contrast, conclusions regarding social or-
ganization (including welfare systems, social status, kinship, marriage and
residence patterns) were the most tenuous and speculative, reflecting the
more indirect physiological impact of these behaviors on the human skele-
ton and requiring more supporting arguments and evidence.
An important component in the evaluation of these inferences is the
existence of supporting data from archaeology or ethnohistory. For exam-
ple, dietary reconstructions benefit from the concordance of osteological
and archaeological information. Conclusions about prehistoric warfare do
as well. Behavioral inferences derived from indirect skeletal evidence alone
need additional supporting evidence even more. In addition, population-
based inferences are often much stronger than those based on smaller sam-
ple sizes. For example, activity level and mobility reconstructions within
and across prehistoric populations are more secure and testable than spe-
cific activities or occupations of isolated individuals. In other situations,
however, small samples may offer tremendous insight into the behavioral
realm--the presence of just one scalped and mutilated victim at a prehis-
toric site has its own set of strong behavioral connotations.
Provocative interpretations of some basic components of human be-
havior, such as subsistence, violence and warfare, migration and cultural
interaction, and general activities are possible through the rigorous and sys-
tematic study of human skeletal samples. An enormous potential for the
future generation of hypotheses addressing these and additional aspects of
human behavior is apparent. But relatively little is known about the biocul-
tural history of prehistoric and historic peoples in many geographic regions.
The probability of imminent reburial of many significant skeletal popula-
tions makes it imperative that skeletal biologists not stop with the standard
descriptive report, but collect data so that biological and behavioral infer-
ences can be made and adequately tested. Only then can the ability to
bring to life past peoples through skeletal signatures of their behavior be
realized.

ACKNOWLEDGMENTS

I wish to thank Cliff Boyd, K. A. R. Kennedy, Clark Larsen, Mary


PoweU, George Milner, Michael Schiffer, Maria Smith, and two anonymous
reviewers for their many helpful comments and suggestions.
Skeletal Correlates of Human Behavior 237

REFERENCES CITED

Angel, J. L., Kelley, J. O., Parfington, M., and Pinter, S. (1987). Life stressesof the free black
community as represented by the firstAfrican Baptist Church, Philadelphia, 1823-1841.
American Journal of Physical Anthropology 74: 213-229.
Arriaza, B. (1993). Seronegative spondyloarthropathies and diffuse idiopathic skeletal
hyperostosis in ancient northern Chile. American Journal of Physical Anthropology 91:
263-278.
Aufderheide, A. C. (1989). Chemical analysis of skeletal remains. In Iscan, M. Y., and
Kennedy, K. A. R. (eds.), Reconstruction of Life From the Skeleton, Alan R. Liss, New
York, pp. 237-260.
Aufderheide, A. C., Angel, J. L., Kelley, J. O., Outlaw, A. C., Outlaw, M. A., Rapp, G., Jr.,
and Wittmers, L. E., Jr. (1985). Lead in bone III. Prediction of social correlates from
skeletal lead content in four Colonial American populations (Catoctin Furnace, College
Landing, Governor's Land, and Irene Mound). American Journal of PhysicalAnthropology
66: 353-361.
Aufderheide, A. C., Wittmers, L. E., Rapp, G., Jr., and Watlgren, J. (1988). Anthropological
applications of skeletal lead analysis. American Anthropologist 90: 931-936.
Bender, M. M., Baerreis, D. A., and Steventon, A. L. (1981). Further light on carbon isotopes
and Hopewell agriculture. American Antiquity 46: 346-353.
Bennett, K. A. (1973). Lumbo-sacral malformations and spina bifida occulta in a group of
proto-historic Modoc Indians. American Journal of Physical Anthropology 36: 435-440.
Berti, P. R., and Mahaney, M. C. (1992). Quantification of the confidence intervals for linear
enamel hypoplasia chronologies. In Goodman, A. H., and Capasso, L. L. (eds.), Recent
Contributions to the Study of Enamel Developmental Defects, Journal of Paleopathology
Monograph 2, Teramo, Italy, pp. 19-30.
Blakely, R. L. (1977). Sociocultural implications of demographic data from Etowah, Georgia.
In Blakely, R. L. (ed.), Bioculturat Adaptation in PrehistoricAmerica, Proceedings of the
Southern Anthropological Society 11, University of Georgia Press, Athens, pp. 45-66.
Blakely, R. L. (1980). Sociocultural implications of pathology between the village area and
Mound C skeletal remains from Etowah, Georgia. In Willey, P. S., and Smith, E H. (eds.),
The Skeletal Biology of Aboriginal Populations in the Southeastern United States,
Miscellaneous Paper No. 5, Tennessee Anthropological Assocation, Chattanooga, pp.
28-38.
Blakely, R. L. (1988). The King Site: Continuity and Contact In Sixteenth-Century Georgia,
University of Georgia Press, Athens.
Blakely, R. L. (1989). Bone strontium in pregnant and lactating females from archaeological
samples. American Journal of Physical Anthropology 80: 173-185.
Blakely, R. L. (1995). Social organization at Etowah: A reconstruction of paleodemographic
and paleonutritional evidence. Southeastern Archaeology 14: 46-59.
Blakely, R. L., and Beck, L. (1984). Tooth-tool use versus dental mutilation: A case study
from the prehistoric southeast. Midcontinental Journal of Archaeology 9: 269-284.
Blakely, R. L., and Mathews, D. S. (1990). Bioarchaeological evidence for a Spanish-Native
American conflict in the sixteenth-century southeast. American Antiquity 55: 718-744.
Blakey, M. L., and Armelagos, G. J. (1985). Deciduous enamel defects in prehistoric
Americans from Dickson Mounds: Prenatal and postnatal stress. American Journal of
Physical Anthropology 66: 371-380.
Blakey, M. L., Leslie, "E E., and Reidy, J. P. (1994). Frequency and chronological distribution
of dental enamel hypoplasia in enslaved African Americans: A test of the weaning
hypothesis. American Journal of Physical Anthropology 95: 371-383.
Blick, J. E (1990). The Quiyoughcohannoek ossuary ritual and the Huron feast of the dead:
A case for cultural diffusion? Paper presented at the Conference on Inter-tribal Relations
in Aboriginal Virginia, Emory and Henry College, Emory, VA.
238 Boyd

Boyd, C. C., and Boyd, D. C. (1991). A multidimensional investigation of biocultural


relationships among three late prehistoric societies in Tennessee. American Antiquity 56:
75-87.
Boyd, D. C. (1988). A Functional Model for Masticatory-Related Mandibular, Dental and
Craniofacial Microevolufionary Change Derivedfrom a Selected Southeastern Indian Skeletal
Temporal Series, Ph.D. dissertation, University of Tennessee, University Microfilms, Ann
Arbor, MI.
Boyd, D. C., and Boyd, C. C. (1989a). A comparison of Tennessee Archaic and Mississippian
maximum femoral lengths and midshaft diameters: Subsistence change and postcranial
variability. Southeastern Archaeology 8: 107-116.
Boyd, D. C., and Boyd, C. C. (1989b). Effects of subsistence and technological change on
masticatory anatomy across a prehistoric skeletal sample from Tennessee. Paper presented
at the 54th Annual Meeting of the Society for American Archaeology, Atlanta.
Boyd, D. C., and Boyd, C. C. (1992). Late Woodland mortuary variability in Virginia. In
Rheinhart, T, and Hodges, M. E. (eds.), Middle and Late Woodland Research in Virginia:
A Synthesis, Special Publication No. 29, Archaeological Society of Virginia, Dietz Press,
Richmond, pp. 249-275.
BradtmiUer, B. (1983). The effect of horseback riding on Arikara arthritis patterns. Paper
presented at the 82nd Annual Meeting of the American Anthropological Association,
Chicago.
Bridges, P. S. (1985). Changes in Long Bone Structure with the Transition to Agriculture:
Implications for PrehistoricActivities, Ph.D. dissertation, University of Michigan, University
Microfilms, Ann Arbor, MI.
Bridges, P. S. (1989a). Changes in activities with the shift to agriculture in the southeastern
United States. Current Anthropology 30: 385-394.
Bridges, P. S. (1989b). Spondylolysis and its relationship to degenerative joint disease in the
prehistoric southeastern United States. American Journal of Physical Anthropology 79:
321-329.
Bridges, P. S. (1991a). Degenerative joint disease in hunter-gatherers and agriculturalists from
the southeastern United States. American Journal of Physical Anthropology 85: 379-391.
Bridges, P. S. (1991b). Skeletal evidence of changes in subsistence activities between the
Archaic and Mississippian time periods in northwestern Alabama. In Powell, M. L.,
Bridges, P. S., and Mires, A. M. W. (eds.), What Mean These Bones?: Studies in
Southeastern Bioarchaeology, University of Alabama Press, Tuscaloosa, pp. 89-101.
Bridges, P. S. (I993). Reply to Dr. Knusel. American Journal of Physical Anthropology 91:
526-527.
Bridges, P. S. (1994). Vertebral arthritis and physical activities in the prehistoric southeastern
United States. American Journal of Physical Anthropology 93: 83-93.
Brock, S. L., and Ruff, C. B. (1988). Diachronic patterns of change in structural properties
of the femur in the prehistoric American southwest. American Journal of Physical
Anthropology 75: 113-127.
Buikstra, J. E. (1981). Mortuary practices, paleodemography and paleopathology: A case study
from the Koster site (Illinois). In Chapman, R., Kinnes, I., and Randsborg, K. (eds.),
The Archaeology of Death, Cambridge University Press, Cambridge, pp. 123-132.
Buikstra, J. E., and Milner, G. R. (1991). Isotopic and archaeological interpretations of diet
in the central Mississippi valley. Journal of Archaeological Science 18: 319-329.
Buikstra, J. E., and Ubelaker, D. H. (1994). Standards For Data Collection From Human
Skeletal Remains, Arkansas Archeological Survey Research Series No. 44, Fayetteville.
Buikstra J. E., Autry, W., Breitburg, E, Eisenberg, L., and van der Merwe, N. (1988). Diet
and health in the Nashville Basin: Human adaptation and maize agriculture in Middle
Tennessee. In Kennedy, B. V., and Le Moine, G. M. (eds.), Diet and Subsistence: Current
Archaeological Perspectives, Proceedings of the 19th Annual Chacmool Conference,
University of Calgary Archaeological Association, Calgary, pp. 243--259.
Bullock, P. Y. (1991). A reappraisal of Anasazi cannibalism. Kiva 57: 5-16.
Skeletal Correlates of Human Behavior 239

Burton, J. H., and Wright, L. E. (1995). Non-linearity in the relationship between bone Sr/Ca
and diet: Paleodietary implications. American Journal of Physical Anthropology 96:
273-282.
Bush, H. (1991). Concepts of health and stress. In Bush, H., and Zvelebil, M. (eds.), Health
in Past Societies: Biocultural Interpretations of Human Skeletal Remains in Archaeological
Contexts, BAR International Series 567, Oxford, pp. 11-21.
Byrd, J. E., and Jantz, R. L. (1994). Osteologlcal evidence for distinct social groups at the
Leavenworth site. In Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great
Plains: Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press,
Washington, D.C., pp. 203-208.
Carlson, D. S., and Van Gerven, D. P. (1977). Masticatory function and post-Pleistocene
evolution in Nubia. American Journal of Physical Anthropology 46: 495-506.
Cartmill, M. (1994). Reinventing anthropology: American Association of Physical
Anthropologists annual luncheon address, April 1, 1994. Yearbook of PhysicalAnthropology
37: 1-9.
Cheverud, J. M., and Buikstra, J. E. (1981a). Quantitative genetics of skeletal nonmetric traits
in the rhesus macaques on Cayo Santiago. I. Single trait heritabilities. American Journal
of Physical Anthropology 49: 43--49.
Cheverud, J, M., and Buikstra, J. E. (1981b). Quantitative genetics of skeletal nonmetric traits
in the rhesus macaques on Cayo Santiago. II. Phenotypic, genetic and environmental
correlations between traits. American Journal of Physical Anthropology 54: 51-58.
Chisholm, B. S. (1989). Variation in diet reconstructions based on stable carbon isotopic
evidence. In Price, T. D. (ed.), The Chemistry of Prehistoric Human Bone, School of
American Research Advanced Seminar Series, Cambridge University Press, Cambridge,
pp. 10-37.
Clarkson, J. (1989). Review of terminology, classification, and indices of developmental defects
of enamel. Advances in Dental Research 3: 104-109.
Cohen, M. N., and Armelagos, G. J. (1984). Paleopathology at the Origins of Agriculture,
Academic Press, Orlando, FL.
Cole, "E M., III (1994). Size and shape of the femur and tibia in northern Plains Indians. In
Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great Plains: Migration,
Warfare, Health, and Subsistence, Smithsonian Institution Press, Washington, D.C., pp.
219-233.
Cook, D. C. (1981). Mortality, age structure and status in the interpretation of stress indicators
in prehistoric skeletons: A dental example from the lower IUinois valley. In Chapman,
R., Kinnes, I, and Randsborg, IC (eds.), The Archaeology of Death, Cambridge University
Press, Cambridge, pp. 133-144.
Cutress, "E W., and Suckling, G. W. (1982). The assessment of noncarious defects of enamel.
International Dental Journal 32: 117-122.
Cybulski, J. S. (1992). A Greenville Burial Ground: Human Remains and Mortuary Elements
in British Columbia Coast Prehistory, Archaeological Survey of Canada Mercury Series
Paper No. 146, Canadian Museum of Civilization, Quebec.
Cybulski, J. S. (1994). Culture change, demographic history, and health and disease on the
northwest coast. In Larsen, C. S., and Milner, G. R. (eds.), In The Wake of Contact,
Wiley-Liss, New York, pp. 75-85.
Danforth, M. E., Cook, D. C., and Knick, S. G., III (1994). The human remains from Carter
Ranch Pueblo, Arizona: Health in isolation. American Antiquity 59: 88-101.
DeNiro, M. J., and Epstein, S. (1978). Influence of diet on the distribution of carbon isotopes
in animals. Geochimica Cosmochimica Acta 42: 495-506.
DeNiro, M. J., and Epstein, S. (1981). Influence of diet on the distribution of nitrogen isotopes
in animals. Geochimica Cosmochimica Acta 45: 341-351.
Dettwyler, K. A. (1991). Can paleopathology provide evidence for "compassion"? American
Journal of Physical Anthropology 84: 375-384.
Dickel, D. N., and Doran, (3. H. (1989). Severe neural tube defect syndrome from the early
Archaic of Florida. American Journal of Physical Anthropology 80: 325-334.
240 Boyd

Dillehay, T D. (1974). Late Quaternary bison population changes on the southern Plains.
Plains Anthropologist 19: 180--196.
Dye, L. O. N. (1976). Status and Biology: An Osteological Analysis, M.A. thesis, Louisiana
State University, Baton Rouge.
Eddy, A., and Cooter, A. (1978). A Drought Probability Model for the USA Northern Great
Plains, University of Oklahoma Press, Norman.
Edynak, G. J. (1976). Life-styles from skeletal material: A Medieval Yugoslav example, tn
Giles, E., and Friedlaunder, J. S. (eds.), The Measures of Man, Peabody Museum Press,
Cambridge, MA, pp. 408-432.
Eisenberg, L. E. (1986). Adaptation in a "Marginal" Mississippian Population from Middle
Tennessee: Biocultural Insights from Paleopathology, Ph.D. dissertation, New York
University, University Microfilms, Ann Arbor, MI.
Eisenberg, L. E. (1988). Some preliminary observations on the assoction between late
prehistoric health and settlement pattern diversity in the Middle Cumberland region of
Tennessee. Paper presented at the 45th Annual Meeting of the Southeastern
Archaeological Conference, New Orleans.
Eisenberg, L. E. (1991a). Interpreting measures of community health during the Late
Prehistoric period in Middle Tennessee: A biocultural approach. In Bush, H., and Zvelebil,
M. (eds.), Health in Past Societies: Biocultural Interpretations of Human Skeletal Remains
in Archaeological Contexts, BAR International Series 567, Oxford, pp. 115-127.
Eisenberg, L. E. (1991b). Mississippian cultural terminations in Middle Tennessee: What the
bioarchaeotogicat evidence can tell us. In PoweU, M. L., Bridges, P S., and Mires, A. M.
W. (eds.), What Mean These Bones?: Studies in Southeastern Bioarchaeology, The University
of Alabama Press, Tuscaloosa, pp. 70-88.
Ezzo, J. A. (1994). Zinc as a paleodietary indicator: An issue of theoretical validity in
bone-chemistry analysis. American Antiquity 59: 606-621.
Fenton, J. E (1991). The Social Use of Dead People: Problems and Solutions in the Analysis of
Post Mortem Body Processing in the Archaeological Record, Ph.D. dissertation, Columbia
University, University Microfilms, Ann Arbor, MI.
Fink, T. M. (1985). Tuberculosis and anemia in a Pueblo II-III (ca. A.D. 900-1300) Anasazi
child from New Mexico. In Merbs, C. E, and Miller, R. J. (eds.), Health and Disease in
the Prehistoric Southwest, Arizona State University Anthropological Research Paper 34,
pp. 359-379.
Fowler, W. R., Jr. (1984). Late preclassic mortuary patterns and evidence for human sacrifice
at Chalchuapa, El Salvador. American Antiquity 49: 603-618.
Frayer, D., and Martin, D. (1996). Troubled 17rues: Osteological and Archaeological Evidence of
l/iotence, War and Society Series, Vol. 6, Gordon and Breach, Langhorne, PA. (in press).
Fresia, A. E., Ruff, C. B., and Larsen, C. S. (1990). Temporal decline in bilateral asymmetry
of the upper limb on the Georgia coast. In Larsen, C. S. (ed.), The Archaeology of Mission
Santa Catalina de Guale: 2. Biocultural Interpretations of a Population in Transition,
Anthropological Papers of the American Museum of Natural History 68: 121-132.
Ghazi, A. M., Reinhard, K. J., Holmes, M. A., and Durrance, E. (1994). Brief communication:
Further evidence of lead contamination of Omaha skeletons. American Journal of Physieal
Anthropology 95: 427-434.
Goodman, A. (1991). Health, adaptation and maladaptation in past societies. In Bush, H.,
and Zvelebil, M. (eds.), Health in Past Societies: Biocultural Interpretations of Human
Skeletal Remains in Archaeological Contexts, BAR International Series 567, pp. 31-38.
Goodman, A. H. (1994). Cartesian reductionism and vulgar adaptationism: Issues in the
interpretation of nutritional status in prehistory. In Sobolik, K. D. (ed.), Paleonutrition:
the Diet and Health of Prehistoric Americans, Southern Illinois University Center for
Archaeological Investigations Occasional Paper No. 22, Carbondale, pp. 163-177.
Goodman, A. H., and Armelagos, G. J. (1985) The chronological distribution of enamel
hypoplasia in human permanent incisor and canine teeth. Archives of Oral Biology 30:
503-507.
Goodman, A. H., and Martin, D. L. (1992). Health, economic change, and regional
political-economic relations: Examples from prehistory. In Huss-Ashmore, R., Schall, J.,
Skeletal Correlates of Human Behavior 241

and Hediger, M. (eds.), Health and Lifestyle Change, MASCA Research Papers in Science
and Archaeology 9, University Museum of Archaeology and Anthropology, University of
Pennsylvania, Philadelphia, pp. 51-60.
Goodman, A. H., Armelagos, G. J., and Rose, J. C. (1980). Enamel hypoplasias as indicators
of stress in three prehistoric populations from Illinois. Human Biology 52: 515-528.
Goodman, A. H., Armelagos, G. J., and Rose, J. C. (1984). The chronological distribution of
enamel hypoplasias from prehistoric Dickson Mounds populations. American Journal of
Physical Anthropology 65: 259-266.
Goodman, A. H., and Rose, J. C. (1990). Assessment of systemic physiological perturbations
from dental enamel hypoplasias and associated histological structures. Yearbookof Physical
Anthropology 33: 59-110.
Goodman, A~ H., Thomas, R. B., Swedlund, A. C., and Armelagos, G. J. (1988). Biocultural
perspectives on stress in prehistoric, historical, and contemporary population research.
Yearbook of Physical Anthropology 31: 169-202.
Goodman, A. H., Martin, D. L., Klein, C. E, Peele, M. S., Cruse, N. A., McEwen, U R.,
Saeed, A., and Robinson, B. M. (1992). Cluster bands, Wilson bands and pit patches:
Histological and enamel surface indicators of stress in the Black Mesa Anasazi population.
In Goodman, A. H., and Capasso, L. L. (eds.), Recent Contributions to the Study of Enamel
Developmental Defects, Journal of Paleopathology Monograph 2, Teramo, Italy, pp.
115-127.
Haas, J., and Creamer, W. (1993). Stress and warfare among the Kayenta Anasazi of the
thirteenth century, A.D. FieMianaAnthropology No. 21 (Publ. 1450).
Hagelberg, E. (1993). Ancient DNA studies. Evolutionary Anthropology 2: 199-207.
Haglund, W. D., Reay, D. T, and Swindler, D. R. (1988). Tooth mark artifacts and survival
of hones in animal scavenged human skeletons. Journal of Forensic Sciences 33: 985-997.
Haglund, W. D., Reay, D. T, and Swindler, D. R. (1989). Canid scavenging/disarticulation
sequence of human remains in the Pacific Northwest. Journal of Forensic Sciences 34:
587-606.
Hall, R. U, Morrow, R., and Clarke, J. H. (1986). Dental pathology of prehistoric residents
of Oregon. American Journal of Physical Anthropology 69: 325-334.
Harritt, R. K., and Radosevich, S. C. (1992). Results of instrument neutron-activation
trace-element analysis of human remains from the Naknek region, Southwest Alaska.
American Antiquity 57: 288-299.
Hartnady, P., and Rose, J. C. (1991). Abnormal tooth loss patterns among Archaic-period
inhabitants of the lower Pecos region, Texas. In Kelley, M. A., and Larsen, C. S. (eds.),
Advances in Dental Anthropology, Wiley-Liss, New York, pp. 267-278.
Hatch, J. W., and Geidel, A. A. (1985). Status-specific dietary variation in two world cultures.
Journal of Human Evolution 14: 469-476.
Herrmann, B., and Hummel, S. (1993).Ancient DNA: Recoveryand Ana~sis of Genetic Material
From Paleontological, Archaeological, Museum, Medical and Forensic Specimens,
Springer-Vedag, New York.
Hill, M. C. (1994). Weight-bearing stress: interpreting habitual behavior and skeletal biology
within a socio-political context. Southeastern Archaeological Conference Bulletin 37: 40.
Hillson, S. (1990). Teeth, Cambridge Manuals in Archaeology, Cambridge University Press,
Cambridge.
Hodges, D. C. (1987). Health and agricultural intensification in the prehistoric valley of
Oaxaca, Mexico. American Journal of Physical Anthropology 73: 323-332.
Holliday, D. Y. (1993). Occipital lesions: A possible cost of cradleboards. American Journal
of Physical Anthropology 90: 283-290.
Hrdlicka, A. (1935). Ear exostoses. Smlthsonian Miscellaneous Collections 93: 1-100.
Huss-Ashmore, R., Goodman, A. H., and Armelagos, G. J. (1982). Nutritional inference from
paleopathology. In Schiffer, M. B. (ed.), Advances in Archaeological Method and Theory,
VoL 5, Academic Press, New York, pp. 395-474.
Hutchinson, D. L (1990). Postcontaet biocultural change: Mortuary site evidence. In Thomas,
D. H. (ed.), Columbian Consequences, VoL 2. Archaeological and I-Ftstorical Perspectives
242 Boyd

on the Spanish Borderlands East, Smithsonian Institution Press, Washington, D.C., pp.
61-70.
Hutchinson, D. L. (1991). Postcontact Native American Health and Adaptation: Assessing the
Impact of Introduced Diseases in Sixteenth-Century Gulf Coast Florida, Ph.D. dissertation,
Um'versity of Illinois at Urbana-Champaign, University Microfilms, Ann Arbor, MI.
Hutchinson, D. L. (1993). "Ihtham mound and the evidence for Spanish and Native American
confrontation. Southeastern Archaeological Conference Bulletin 36: 23.
Hutchinson, D. L., and Larsen, C. S. (1988). Determination of stress episode duration from
linear enamel hypoplasias: A case study from St. Catherine's Island, Georgia. Human
B/o/ogy 60: 93-110.
Iscan, M. Y., and Kennedy, IC A. R. (1989). Reconstruction of Life from the Skeleton, Alan
R. Liss, New York.
Irish, J. D., and Turner, C. G., II. (1987). More lingual surface attrition of the maxillary
anterior teeth in American Indians: Prehistoric Panamanians. American Journal of Physical
Anthropology 73: 209-213.
Jackes, M. (1992). Paleodemography: Problems and techniques. In Saunders, S. R., and
Katzenberg, M. A. (eds.), Skeletal Biology of Past Peoples: Research Methods, Wiley-Liss,
New York, pp. 189-224.
Jantz, R. L. (1994). The social, historical, and functional dimensions of skeletal variation. In
Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great Plains: Migration,
Warfare, Health, and Subsistence, Smithsonian Institution Press, Washington, DC, pp.
175-178.
Jantz, R. L., and Owsley, D. W. (1994). White traders in the upper Missouri: Evidence from
the Swan Creek site. In Owsley, D. W., and Jantz, R. L (eds.), Skeletal Biology in the
Great Plains: Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press,
Washington, DC, pp. 189-201.
Jurmain, R. (1990). Paleoepidemiology of a central California prehistoric population from
CA-ALA-329. II. Degenerative disease. American Journal of Physical Anthropology 83:
83-94.
Jurmain, R. (1991). Paleoepidemiology of trauma in a prehistoric central California
population. In Ortner, D. J., and Aufderheide, A. C. (eds.), Human Paleopathology:
Current Syntheses and Future Options, Smithsonian Institution Press, Washington, DC, pp.
241-248.
Katzenberg, M. A. (1991a). Analysis of stable isotopes of carbon and nitrogen. In Pfeiffer,
S., and Williamson, R. E (eds.), Snake Hill: An Investigation of a Military Cemetery from
the War of 1812, Dundum, Toronto, pp. 247-255.
Katzenberg, M. A. (1991b). Isotopic analysis. In Saunders, S. R., and Lazenby, R. (eds.), The
Links that Bind: The Harvie Faro@ Nineteenth Century Burying Ground, Copetown Press,
Dundas.
Katzenberg, M. A. (1992a). Advances in stable isotope analysis of prehistoric bones. In
Saunders, S. R., and Katzenberg, M. A. (eds.), Skeletal Biology of Past Peoples: Research
Methods, Wiley-Liss, New York, pp. 105-119.
Katzenberg, M. A. (1992b). Changing diet and health in pre- and proto-historic Ontario. In
Huss-Ashmore, R., Sehall, J., and Hediger, M. (eds.), Health and Lifestyle Change,
MASCA Research Papers in Science and Archaeology Volume 9, University Museum of
Anthropology and Archaeology, University of Pennsylvania, Philadelphia, pp. 23-31,
Katzenberg, M. A., Saunders, S. R., and Fitzgerald, W. R. (1993). Age differences in stable
carbon and nitrogen isotope ratios in a population of prehistoric maize horticulturalists.
American Journal of Physical Anthropology 90: 267-281.
Katzenberg, M. A., Schwartz, H. E, Knyf, M., and Melbye, E J. (1995). Stable isotope evidence
for maize horticulture and paleodiet in southern Ontario, Canada. American Antiquity
60: 335-350.
Keegan, W. E (1989). Stable isotope analysis of prehistoric diet. In Isean, M. Y., and Kennedy,
K. A. R. (eds.), Reconstruction of Life From the Skeleton, Alan R. Liss, New York, pp.
223-236.
Skeletal Correlates of Human Behavior 243

Kelley, L O., and Angel, J. L (1983). Workers of Catoctin Furnace. Maryland Archaeologist
19: 2-17.
Kelley, J. O., and Angel, J. L. (1987). Life stresses of slavery. American Journal of Physical
Anthropology 74: 199-211.
Kelso, J. (1995). A place for the culture concept in biological anthropology. In Boaz, N. "I:,
and Wolfe, L. D. (eds.), B~lo~cal Anthropology: The State of the Science, International
Institute for Evolutionary Research, Oregon State University Press, Corvalis, pp. 241-250.
Kennedy, G. E. (1986). The relationship between auditory exostoses and cold water: A
latitudinal analysis. American Journal of Physical Anthropology 71: 401-415.
Kennedy, K. A. R. (1983). Morphological variations in ulnar supinator crests and fossae as
identifying markers of occupational stress. Journal of Forensic Sciences 28: 871--876.
Kennedy, IC A. R. (1989). Skeletal markers of occupational stress. In Iscan, M. Y., and
Kennedy, IC A. R. (eds.), Reconstruction of Life From the Skeleton, Alan R. Liss, New
York, pp. 129-160.
Kent, S. (1986). The influences of sedentism and aggregation on porotic hyperostosis and
anemia: A case study. Man (N.S.) 21: 605-636.
Key, P. J. (1983). Craniometric Relationships Among Plains Indians, University of Tennessee
Department of Anthropology Report of Investigations No. 34, Knoxville.
Knusel, C. J. (1993). On the biomechanical and osteoarthritic differences between
hunter-gatherers and agriculturalists. American Journal of Physical Anthropology 91:
523-527.
Konigsberg, L. W. (1988). Migration models of prehistoric post-marial residence. American
Journal of Physical Anthropology 77: 471-482.
Kouigsberg, L. W., and Buikstra, J. E. (1995). Regional approaches to the investigation of
past human biocultural structure. In Beck, L. A. (ed.), Regional Approaches to Mortuary
Analysis, Plenum Press, New York, pp. 191-219.
Lahren, C. H., and Berryman, H. E. (1984). Fracture patterns and status at Chucalissa (40SI1):
A biocultural approach. TennesseeAnthropologist 9: 15-21.
Lallo, J. W., and Rose, J. C. (1979). Patterns of stress, disease and mortality in two prehistoric
populations from North America. Journal of Human Evolution 8: 323-335.
Lambert, J. B., Szpunar, C. B., and Buikstra, J. E. (1979). Chemical analysis of excavated
human bone from Middle and Late Woodland sites. Archaeometry 21: 115-129.
Lambert, P. M. (1993). Health in prehistoric populations of the Santa Barbara Channel islands.
American Antiquity 58: 509-522.
Lambert, P. M., and Walker, P. L. (1991). Physical anthropological evidence for the evolution
of social complexity in coastal southern California. Antiquity 65: 963-973.
Lane, R. A., and Sublett, A. J. (1972). Osteology of social organization: Residence pattern.
American Antiquity 37: 186-201.
Langdon, S. P., Willey, P., and Cummins, R. W. (1993). The South Dakota reburial program
and the discovery of a possible prehistoric dwarf. Plains Anthropologist 38: 271-281,
Memoir 27.
Lanphear, IC M. (1990). Frequency and distribution of enamel hypoplasias in a historic skeletal
sample. American Journal of Physical Anthropology 81: 35-43.
Larsen, C. S. (1980). Dental caries: Experimental and biocultural evidence. In Willey, P., and
Smith, E H. (eds.), The SkeletalBiology of Aboriginal Populations in the Southeastern United
States, Tennessee Anthropological Association Miscellaneous Paper No. 5, pp. 75-80.
Larsen, C. S. (1981). Functional implications of postcranial size reduction on the prehistoric
Georgia coast, U.S.A. Journal of Human Evolution 10: 489-502.
Larsen, C. S. (1983). Behavioral implications of temporal change in eariogenesis. Journal of
Archaeological Science 10: 1-8.
Larsen, C. S. (1985). Dental modifications and tool use in the western Great Basin. American
Journal of Physical Anthropology 67: 393--402.
Larsen, C. S. (1987). Bioarchaeological interpretations of subsistence economy and behavior
from human skeletal remains. In Sehiffer, M. B. (ed.),Advances in Archaeological Method
and Theory, VoL 10, Academic Press, San Diego, pp. 339--445.
244 Boyd

Larsen, C. S. (1990). The Archaeology of Mission Santa Catalina de Guale. 2. Biocultural


Interpretations of a Population in Transition, Anthropological Papers of the American
Museum of Natural History 68.
Larsen, C. S. (1993). On the frontier of contact: Mission bioarchaeology in La Florida. In
McEwan, B. G. (ed.), The Spanish Missions of La Florida, University Press of Florida,
Gainesville, pp. 322-356.
Larsen, C. S. (1994). In the wake of Columbus: Native population biology in the postcontact
Americas. Yearbook of Physical Anthropology 37: 109-154.
Larsen, C. S. (1995). Biological changes in human populations with agriculture. Annual Review
of Anthropology 24: 185-213.
Larsen, C. S., and Huynh, H. P. (1993). Death by gunshot: Biocuttural implications of trauma
at Mission San Luis. Southeastern Archaeological Conference Bulletin 36: 26.
Larsen, C. S., and Ruff, C. B. (1991). Biomechanical adaptation and behavior on the
prehistoric Georgia coast. In Powell, M. L., Bridges, P. S., and Mires, A. M. W. (eds.),
What Mean These Bones?: Studies in Southeastern Bioarchaeology, University of Alabama
Press, Tuscaloosa, pp. I02-113.
Larsen, C. S., and Thomas, D. H. (1982). The Anthropology of St. Catherine's Islan& 4. The
St. Catherine's Period Mortuary Complex, Anthropological Papers of the American
Museum of Natural History 57: 271-341.
Larsen, C. S., and Ruff, C. B. (1994). The stresses of conquest in Spanish Florida: Structural
adaptation and change before and after contact. In Larsen, C. S., and Milner, G. R.
(eds.), In the Wake of Contact, Wiley-Liss, New York, pp. 21-34.
Larsen, C. S., Schoeninger, M. J., Hutchinson, D. L., Russell IC E, and Ruff, C. B. (1990).
Beyond demographic collapse: Biological adaptation and change in native populations of
La Florida. In Thomas, D. H. (ed.), Columbian Consequences, Vot 2. Archaeological and
Historical Perspectives on the Spanish Borderlands East, Smithsonian Institution Press,
VCashington, DC, pp. 409--428.
Larsen, C. S., Shavit, R., and Griffin, M. C. (1991). Dental caries evidence for dietary change:
An archaeological context. In Kelley, M. A~, and Larsen, C. S. (eds.), Advances in Dental
Anthropology, Wiley-Liss, New York, pp. 179-202.
Larsen, C. S., Schoeninger, M. J., van der Merwe, N. J., Moore, IC M., and Lee-Thorp, J.
A. (1992). Carbon and nitrogen stable isotope signatures of human dietary change in the
Georgia Bight. American Journal of Physical Anthropology 89: 197-214.
Leone, M. P., and Palkovich, A. M. (1985). Ethnographic inferences and analogy in analyzing
prehistoric diets. In Gilbert, R. I., and Mietke, J. H. (eds.), The Analysis of Prehistoric
D/ets, Academic Press, Orlando, FL, pp. 423--431.
Little, B. J., Lanphear, IC M., and Owsley, D. W. (1992). Mortuary display and status in a
nineteenth-century Anglo-American cemetery in Manassas, Virginia. American Antiquity
57: 397--418.
Little, E., and Schoeninger, M. J. (1995). The Late Woodland diet on Nantucket island and
the problem of maize in coastal New England. American Antiquity 60: 351-368.
Lyman, R. L. (1994). Vertebrate Taphonomy, Cambridge Manuals in Archaeology, Cambridge
University Press, Cambridge.
Mann, R. W., Owsley, D. W., and Reinhard, IC, Jr. (1994). Otitis media, mastoiditis, and
infracranial lesions in two Plains Indian children. In Owsley, D. W., and Jantz, R. L.
(eds.), Skeletal Biology in the Great Plains: Migration, Warfare, Health, and Subsistence,
Smithsonian Institution Press, "?ashington, DC, 131-146.
Martin, D. L., and Akins, N. J. (1994). Patterns of violence against women in the prehistoric
southwest. Paper presented at the Southwest Symposium, Tempe, AZ.
Martin, D. L., Goodman, A. H., Armelagos, G. J., and Magennis, A. L. (1991). Black Mesa
Anasazi Health: Reconstructing Life from Patterns of Death and Disease, Southern Illinois
University at Carbondale Center for Archaeological Investigations Occasional Paper No.
14.
Mensforth, R. R (1985). Relative tibia long bone growth in the Libben and Bt-5 prehistoric
skeletal population. American Journal of Physical Anthropology 68: 247-262.
Skeletal Correlates of Human Behavior 245

Mensforth, R. P., and Baker, G. (1995). Evidence of violent injury, dismemberment, and
trophy-taking behaviors among inhabitants of the Late Archaic Ward site (Mc-ll) from
Kentucky. Paper presented at the 2nd Annual Meeting of the Midwest Bioarchaeology
and Forensic Anthropology Association, Northern Illinois University, DeKalb.
Merbs, C. E (1983). Patterns of Activity-Induced Pathology in a Canadian Inuit Population,
National Museum of Man Mercury Series Archaeological Survey of Canada Paper No.
119, Ottawa.
Merbs, C. E, and Birkby, W. H. (1985). Evidence for prehistoric scalping at Nuvakwewtaqa
(Chavez Pass) and Grasshopper Ruin, Arizona. In Merbs, C. E, and Miller, R. J. (eds.),
Health and Disease in the Prehistoric Southwest, Arizona State University Anthropological
Research Paper No. 34, pp. 23-42.
Merbs, C. E, and Euler, R. C. (1985). Atlanto-occipital fusion and spondylolisthesis in an
Anasazi skeleton from Bright Angel Ruin, Grand Canyon National Park, Arizona.
American Journal of Physical Anthropology 67: 381-391.
Miller, R. J. (1985). Lateral epicondylitis in the prehistoric Indian population from
Nuvakwewtaqa (Chavez Pass), Arizona. In Merbs, C. E, and Miller, IL J. (eds.), Health
and Disease in the Prehistoric Southwest, Arizona State University Anthropological
Research Paper No. 34, pp. 391-400.
Milner, G. R. (1984). Dental caries in the permanent dentition of a Mississippian period
population from the American Midwest. Collegium Antropologicum 8: 77-91.
Milner, G. R. (1991). Health and cultural change in the late prehistoric American Bottom,
Illinois. In Powell, M. L, Bridges, P. S., and Mires, A. M. W. (eds.), What Mean These
Bones?: Studies in Southeastern Bioarchaeology, University of Alabama Press, Tuscaloosa,
pp. 52-69.
Milner, G. R. (1992). Disease and sociopolitical systems in late prehistoric Illinois. In Verano,
J. W., and Ubelaker, D. H. (eds.), Disease and Demography in the Americas, Smithsonian
Institution Press, Washington, DC, pp. 103-116.
Milner, G. R. (1995). An osteological perspective on prehistoric warfare. In Beck, L A. (ed.),
Regional Approaches to Mortuary Analysis, Plenum Press, New York, pp. 221-244.
Milner, G. R., and Larsen, C. S. (1991). Teeth as artifacts of human behavior: Intentional
mutilation and accidental dental modification. In Kelley, M. A., and Larsen, C. S. (eds.),
Advances in Dental Anthropology, pp. 357-378.
Milner, G. R., and Smith, V. G. (1989). Carnivore alteration of human bone from a late
prehistoric site in Illinois. American Journal of Physical Anthropology 79: 43-49.
Milner, G. R., Anderson, E., and Smith, V. G. (1991). Warfare in late prehistoric west-central
Illinois. American Antiquity 56: 581-603.
Milner, G. R., Larsen, C. S., Hutchinson, D. L., and Williamson, M. (1994). Conquistadors,
excavators, or rodents. Southeastern Archaeological Conference Bulletin 37: 51.
Molnar, S. (1972). Tooth wear and culture: A survey of tooth functions among some prehistoric
populations. CurrentAnthropology 13: 511-526.
Munizaga, J. R. (1991). Human skeletal pathology in pre-Columbian populations of northern
Chile. In Ortner, D. J., and Anfderheide, A. C. (eds.), Human Paleopathology: Current
Syntheses and Future Options, Smithsonian Institution Press, Washington, DC, pp. 145-150.
Neiburger, E. J. (1990). Enamel hypoplasias: Poor indicators of dietary stress. American
Journal of Physical Anthropology 82: 231-232.
Nelson, G. C. (1992). Maxillary canine/third premolar transposition in a prehistoric population
from Santa Cruz island, California. American Journal of PhysicalAnthropology 88: 135-144.
Olsen, S., and Shipman, P. (1994). Cut-marks and perimortem treatment of skeletal remains
on the northern Plains. In Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the
Great Plains: Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press,
Washington, DC, pp. 377-387.
Ortner, D. J. (1991). Theoretical and methodological issues in paleopathology. In Ortner, D.
J., and Aufderheide, A. C. (eds.), Human Paleopathology: Current Syntheses and Future
Options, Smithsonian Institution Press, Washington, DC, pp. 5-11.
Ossenberg, N. S. (1976). Within and between race distances in population studies based on
discrete traits of the human skull. American Journal of Physical Anthropology 45: 701-716.
246 Boyd

Owsley, D. W. (1994). Warfare in Coalescent tradition populations of the northern Plains. In


Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great Plains: Migration,
Warfare, Health, and Subsistence, Smithsonian Institution Press, Washington, DC, pp.
333--343.
Owsley, D. W., and Jantz, R. L (1985). Long bone lengths and gestational age distributions
of post-contact period Arikara Indian perinatal infant skeletons. American Journal of
Physical Anthropology 68: 321-328.
Owsley, D. W., Berryman, H. E., and Bass, W. M. (1977). Demographic and osteological
evidence for warfare at the Larson site, South Dakota. Plains Anthropologist 22: 119-131,
Memoir 13.
Owsley, D. W., Orser, C. E., Jr., Mann, R. W., Moore-Jansen, P. H., and Montgomery, R. L.
(1987). Demography and pathology of an urban slave population from New Orleans.
American Journal of Physical Anthropology 74: 185-197.
Owsley, D. W., Mann, R. W., and Murphy, S. P. (1991). Injuries, surgical care and disease. In
Pfeiffer, S., and Williamson, R. E (eds.), Snake HilL'An Investigation of a Military Cemetery
from the War of 1812, Dundurn, Toronto, pp. 198-226.
Owsley, D. W., Mann, R. W., and Baugh, T G. (1994). Culturally modified human bones from
the Edwards I site. In Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great
Plains: Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press,
Washington, DC, pp, 363-375.
Parham, IC R., and Scott, G. T (1980). Porotic hyperostosis: A study of disease and culture
at Toqua (40MR6), a Late Mississippian site in eastern Tennessee. In Willey, P., and Smith,
E H. (eds.), The Skeletal Biology of Aboriginal Populations in the Southeastern United States,
Tennessee Anthropological Association Miscellaneous Paper No. 5, pp. 39-51.
Pate, E D. (1994). Bone chemistry and paleodiet. Journal of Archaeological Method and Theory
1: 161-209.
Pfeiffer, S. (1984). Paleopathology in an Iroquoian ossuary, with special reference to
tuberculosis. American Journal of Physical Anthropology 65: 181-189.
Pfeiffer, S., and Williamson, R. E (1991). Snake Hill: An Investigation of a Military Cemetery
from the War of 1812, Dundurn, Toronto.
Powell, J. E, and Steele, D. G. (1994). Diet and health of Paleoindians: An examination of
early Holocene human dental remains. In Sobolik, K. D. (ed.), Paleonutrition: The Diet
and Health of Prehistoric Americans, Southern lllinois University Center for
Archaeological Investigations Occasional Paper No. 22, Carbondale, pp. 178-194.
Powell, M. L. (1985). The analysis of dental wear and caries for dietary reconstruction. In
Gilbert, R. I., and Mielke, J. H. (eds.), The Analysis of Prehistoric Diets, Academic Press,
Orlando, FL, pp. 307-338.
PoweU, M. L. (1988). Status and Health in Prehistory:A Case Study of the Moundville Chiefdom,
Smithsonian Institution Press, Washington, DC.
Powell, M. L. (1991). Ranked status and health in the Mississippian chiefdom at Moundville.
In Powell, M. L., Bridges, P. S., and Mires, A. M. W. (eds.), What Mean These Bones?
Studies in Southeastern Bioarchaeology, The University of Alabama Press, Tuscaloosa, pp.
22-51.
Radosevich, S. C. (1993). The six deadly sins of trace element analysis: A case of wishful
thinking in science, In Sandford, M. IC (ed.), Investigations of Ancient Human Ttssue.
Gordon and Breach, Langhorne, PA, pp. 269-332.
Raemsch, C. A. (1993). Mechanical procedures involved in bone dismemberment and
defleshing in prehistoric Michigan. Mid.continental Journal of Archaeology 18: 217-244.
Rathbun, "I: A. (1987). Health and disease at a South Carolina plantation: 1840-1870.Amedcan
Journal of Physical Anthropology 74: 239-253.
Rathbun, T. A. (1993). Trauma at a Late Archaic South Carolina coastal site: 38BU9.
Southeastern Archaeological Conference Bulletin 36: 32.
Rathbun, T A., and Scurry, J. D. (1991). Status and health in Colonial South Carolina:
BeUeview Plantation, 1738-1756. In Powell, M. L, Bridges, P. S., and Mires, A. M. W.
(eds.), What Mean These Bones? Studies in Southeastern Bioarchaeology, University of
Alabama Press, Tuscaloosa, pp. 148-164.
Skeletal Correlates of Human Behavior 247

Rathbun, "E A., Sexton, J., and Michie, J. (1980). Disease patterns in a formative period South
Carolina coastal population. In Willey, P., and Smith, E H. (eds.), The Skeletal Biology
of Aboriginal Populations in the Southeastern United States, Tennessee Anthropological
Association Miscellaneous Paper No. 5, pp. 52-74.
Reed, D. M. (1994). Ancient Maya diet at Copan, Honduras as determined through the
analysis of stable carbon and nitrogen isotopes. In Sobolik, IC D. (ed.), Paleonutrition:
The Diet and Health of Prehistoric Americans, Southern Illinois University Center for
Archaeological Investigations Occasional Paper No. 22, Carbondale, pp. 210-221.
Reinhard, IC J. (1992). Patterns of diet, parasitism and anemia in prehistoric west North
America. In Stuart-Macadam, P., and Kent, S. (eds)., Diet, Demography and Disease:
Changing Perspectives on Anemia, Aldine de Gruyter, New York, pp. 219-258.
Reinhard, K. J., and Ghazi, A. M. (1992). Evaluation of lead concentrations in 18th-century
Omaha Indian skeletons using ICP-MS. American 1ournal of Physical Anthropology 89:
183-195.
Reinhard, IC J., Tieszen, L., Sandness, IC L., Beiningen, L. M., Miller, E., Ghazi, M., Miewald,
C. E., and Barnum, S. V. (1994). Trade, contact, and female health in northeast Nebraska.
In Larsen, C. S., and Milner, G. R. (eds.), In the Wake of Contact, Wiley-Liss, New York,
pp. 63-74.
Relethford, J. H. (1994). The Human Species: An Introduction to Biological Anthropology,
Mayfield, Mountain View, CA.
Rose, J. C., Marks, M. IC, and Tieszan, L L. (1991). Bioarchaeology and subsistence in the
central and lower portions of the Mississippi valley. In Powell, M. L., Bridges, P. S., and
Mires, A. M. W. (eds.), What Mean These Bones? Studies in Southeastern Bioarchaeology,
University of Alabama Press, Tusealoosa, pp. 7-21.
Roth, E. A. (1992). Applications of demographic models to paleodemography. In Saunders,
S. R., and Katzenberg, M. A. (eds.), Skeletal Biology of Past Peoples: Research Methods,
Wiley-Liss, New York, pp. 175-188.
Ruff, C. B. (1987). Sexual dimorphism in the human lower limb structure: Relationship to
subsistence strategy and sexual division of labor. Journal of Human Evolution 16: 391--416.
Ruff, C. B. (1992). Biomechanical analyses of archaeological human skeletal samples. In
Saunders, S. R., and Katzenberg, M. A., (eds.), Skeletal Biology of Past Peoples: Research
Methods, Wiley-Liss, New York, pp. 37-58.
Ruff, C. B. (1994). Biomechanical analysis of northern and southern Plains femora: Behavioral
interpretations. In Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great
Plains: Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press,
Washington, DC, pp. 235-245.
Ruff, C. B., and Larsen, C. S. (1990). Postcranial biomechanical adaptations to subsistence
changes on the Georgia Coast. In Larsen, (3. S. (ed.), The Archaeology of Mission Santa
Catalina de Guale: 2. Biocultural Interpretations of a Population in Transition.
Anthropological Papers of the American Museum of Natural History 68: 94-120.
Ruff, C. B., Larsen, C. S., and Hayes, W. C. (1984). Structural changes in the femur with the
transition to agriculture on the Georgia coast. American Journal of Physical Anthropology
64: 125-136.
Sandford, M, IC (1992). A reconsideration of trace element analysis in prehistoric bone. In
Saunders, S. R., and Katzenberg, M. A. (eds.), Skeletal Biology of Past Peoples: Research
Methods, Wiley-Liss, New York, pp. 79-103.
Sandford, M. K. (1994). Investigations of Ancient Human Tissue: Chemical Analyses in
Anthropology, Gordon and Breach, Langhorne, PA.
Saul, E P. (1976). Osteobiography: Life history recorded in bone. In Giles, E., and
Friedlannder, J. S. (eds.), The Measures of Man, Peabody Museum Press, Cambridge,
MA, pp. 372-382.
Saul, E P., and Saul, J. M. (1989). Osteobiography: A Maya example. In Iscan, M. Y., and
Kennedy, IC A. R. (eds.), Reconstruction of Life from the Skeleton, Alan R. Liss, New
York, pp. 287-302.
Satmders, S. R. (1989). Nonmetrie skeletal variation. In Iscan, M. Y., and Kennedy, IC A. R.
(eds.), Reconstruction of Life From the Skeleton, Alan R. Liss, New York, pp. 95-108.
248 Boyd

Saunders, S. R., and Katzenberg, M. A. (1992). Skeletal Biology of Past Peoples: Research
Methods, Wiley-Liss, New York.
Schiffer, M. B. (1995). BehavioralArchaeology: First Principles, University of Utah Press, Salt
Lake City.
Schneider, K. N., and Blakeslee, D. (1990). Evaluating residence patterns among prehistoric
populations: Clues from dental enamel composition. Human Biology 62: 71-83.
Schoeninger, M. J. (1979). Diet and status at Chalcatzingo: Some empirical and technical
aspects of strontium analysis. American Journal of Physical Anthropology 51: 295-309.
Schoeninger, M. J., and Moore, K. (1992). Bone stable isotope studies in archaeology. Journal
of Worm Prehistory 6: 247-296.
Schoeninger, M. J., van der Merwe, N. J., Moore, IC, Lee-Thorp, J., and Larsen, C. S. (1990).
Decrease in diet quality between the prehistoric and contact periods. In Larsen, C. S.
(ed.), The Archaeology of Mission Santa Catalina de Guale, 2. Biocultural Interpretations
of a Population in Transition. Anthropological Papers of the American Museum of Natural
History 60: 78-93.
Schulz, P. D. (I977). Task activity and anterior tooth grooving in prehistoric California Indians.
American Journal of Physical Anthropology 46: 87-92.
Schurr, M. R. (1992). Isotopic and mortuary variability in a Middle Mississippian population.
American Antiquity 57: 300-320.
Schwarcz, H. P., and Schoeninger, M. J. (1991). Stable isotope analysis in human nutritional
ecology. Yearbook of Physical Anthropology 34: 283-321.
Schwartz, J. H., Brauer, J., and Gordon-Larsen, P. (1995). Brief communication: Tigaran (Point
Hope, Alaska) tooth drilling. American Journal of Physical Anthropology 97: 77-82.
Sciulli, P. W. (1992). Estimating age of occurrence of enamel defects in deciduous teeth. In
Goodman, G. H., and Capasso, L. L. (eds.), Recent Contributions to the Study of Enamel
Developmental Defects, Journal of Paleopathology Monograph 2, Teramo, Italy, pp. 31-39.
Sciulli, P. W., and Gramly, R. M. (1989). Analysis of the Ft. Laurens, Ohio, skeletal sample.
American Journal of PhysicalAnthropology 80: 11-24.
Seeman, M. E (1988). Ohio Hopewell trophy-skull artifacts as evidence for competition in
Middle Woodland societies circa 50 B.C.-A.D. 350. American Antiquity 53: 565-577.
Self, S. G., and Leamy, L. (1978). Heritability of quasi-continuous skeletal traits in a
randombred population of house mice. Genetics 88: 109-120.
Sillen, A., Scaly, J. C., and van der Merwe, N. J. (1989). Chemistry and paleodietary research:
No more easy answers. American Antiquity 54: 504-512.
Skinner, M. E, and Hung, J. T. W. (1989). Social and biological correlates of localized enamel
hypoplasia of the human deciduous canine tooth. American Journal of Physical
Anthropology 79: 159-175.
Smith, E H., Smith, M. O., and Hinton, R. J. (1980). Evolution of tooth size in the prehistoric
inhabitants of the Tennessee valley. In Willey, P., and Smith, E H. (eds.), The Skeletal
Biology of Aboriginal Populations in the Southeastern United States, Tennessee
Anthropological Association Miscellaneous Paper No. 5, pp. 81-103.
Smith, M. O. (1982). Patterns of Association Between Oral Health and Subsistence: A Study of
Aboriginal Skeletal Populations From the Tennessee Valley Area, Ph.D. dissertation,
University of Tennessee, Knoxville.
Smith, M. O. (1993a). A probable case of decapitation at the Late Archaic Robinson site
(40SM4), Smith County, Tennessee. TennesseeAnthropologist 18: 131-142.
Smith, M. O. (1993b). Forearm trauma and status in the Archaic. Southeastern Archaeological
Conference Bulletin 36: 35.
Smith, M. O. (1995). Scalping in the Archaic period: Evidence fxom the western Tennessee
valley. Southeastern Archaeology 14: 60--68.
Smith, M. O. (1996a). "Parry" fractures and female-directed interpersonal violence:
Implications from the Late Archaic period of West Tennessee. International Journal of
Osteoarchaeology 6: 261.1-262.8.
Smith, M. O. (1996b), Osteological indications of warfare in the Archaic period of the western
Tennessee valley. In Frayer, D., and Martin, D. (eds.), Troubled lanes: Osteological and
Skeletal Correlates of Human Behavior 249

Archaeological Evidence of lrzolence, War and Society Series, VoL 6, Gordon and Breach,
Langhorne, PA. (in press)
Snyder, L., and Willey, E S. (1989). Canid modifications of human skeletal remains: A
comparison of archaeological materials from Crow Creek, modern forensic cases and a
controlled non-human sample. Paper presented at the Society for American Archaeology,
Atlanta.
Sobolik, K. D. (1994). Paleonutrition of the lower Pecos region of the Chihuahuan desert. In
Sobolik, IC D. (ed.), PaleonuMtion: The Diet and Health of PrehistoricAmericans, Center
for Archaeological Investigation, Southern Ilinois University, Occasional Paper 22, pp.
247-264.
Spence, M. W. (1974). Residential practices and the distribution of skeletal traits in
Teotihuacan, Mexico. Man 9: 262-273.
Spielmann, K. A., Schoeninger, M. J., and Moore, K. (1990). Plains-Pueblo interdependence
and human diet at Pecos Pueblo, New Mexico. American Antiquity 55: 745-765.
Steegmann, A. T, Jr. (1991). Stature in an early mid-19th century poorhouse population:
Highland Park, Rochester, New York. American Journal of Physical Anthropology 85:
261-268.
Steele, D. G., and Powell, J. E (1993). Paleobiology of the first Americans. Evolutionary
Anthropology 2: 138-146.
Stirland, A. (1991). Diagnosis of occupationally-related paleopathology: Can it be done? In
Ortner, D. J., and Aufderheide, A. C. (eds.), Human Paleopathology: Current Syntheses
and Future Options, Smithsonian Institution Press, Washington, DC, pp. 40-47.
Stone, A. C., and Stoneking, M. (1993). Ancient DNA from a pre-Cotumbian Amerindian
population. American Journal of Physical Anthropology 92: 463--471.
Storey, R. (1986). Perinatal mortality at Pre-Columbian Teotihuacan. American Journal of
Physical Anthropology 69: 541-548.
Storey, R. (1992). Preindustrial urban lifestyle and health. In Hnss-Ashmore, R., Schall, J.,
and Hediger, M. (eds.). Health and Lifestyle Change, MASCA Research Papers in Science
and Archaeology 9, University Museum of Archaeology and Anthropology, University of
Pennsylvania, Philadelphia, pp. 33--42.
Stuart-Macadam, P. (1991). Porotic hyperostosis: Changing interpretations. In Ortner, D. J.,
and Aufderheide, A. C. (eds.), Human Paleopathology: Current Syntheses and Future
Options, Smithsonian Institution Press, Washington, DC, pp. 36-39.
Stuart-Macadam, P. (1992). Anemia in past human populations. In Stuart-Macadam, P., and
Kent, S. (eds.), Diet, Demography, and Disease: Changing Perspectiveson Anemia, Aldine
de Gruyter, New York, pp. 151-170.
Suckling, G., Elliot, D. C., and Thurley, D. C. (1986). The macroscopic appearance and
associated histological changes in the enamel organ of hypoplastic lesions of sheep incisor
teeth resulting from induced parasitism. Archives of Oral Biology 31: 427--439.
Suckling, G. (1989). Developmental defects of enamel-historical and present day perspectives
of their pathogenesis. Advances in Dental Research 3: 87-94.
q/tinter, J. A. (1980). Behavior and status in a Middle Woodland mortuary population from
the Illinois valley. American Antiquity 45: 308-313.
Teaford, M. E (1991). Dental microwear: What can it tell us about diet and dental function?
In Kelley, M. A., and Larsen, C. S. (eds.), Advances in Dental Anthropology, Wiley-Liss,
New York, pp. 341-356.
Torbenson, M., Aufderheide, A., and Johnson, E. (1992). Punctured human bones of the
Laurel culture from Smith Mound Four, Minnesota. American Antiquity 57: 506-514.
Trinkaus, E. (1975). Squatting among the Neanderthals: A problem in the behavioral
interpretation of skeletal morphology. Journal of Archaeological Science 2: 327-35t.
Turkel, S. J. (1989). Congenital abnormalities in skeletal populations. In Iscan, M. Y., and
Kennedy, K. A. R. (eds.), Reconstruction of Life From the Skeleton, Alan R. Liss, New
York, pp. 109-127.
Turner, C. G., II (1979). Dental anthropological indications of agriculture among the Jomon
people of central Japan. American Journal of Physical Anthropology 51: 619-635.
250 Boyd

Turner, C. G., H. (1993). Cannibalism in Chaco Canyon: The charnel pit excavated in 1926
at Small House Ruin by Frank H. H. Roberts, Jr. American Journal of Physical
Anthropology 91: 421-439.
Turner, C. G., II, and Turner, J. A. (1992). The first claim for cannibalism in the southwest:
Walter Hough's 1901 discovery at Canyon Butte Ruin 3, northeastern Arizona. American
Antiquity 57: 661-682.
Turner, C. G., II., Tamer, J. A., and Green, R. C. (1993). Taphonomic analysis of Anasazi
skeletal remains from Largo-Gallina sites in northwestern New Mexico. Journal of
Anthropological Research 49: 83-110.
Tuross, N., and Fogel, M. L. (1994). Stable isotope analysis and subsistence patterns at the
Sully site. In Owsley, D. W., and Jantz, R. L (eds.), Skeletal Biology in the Great Plains:
Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press, Washington,
DC, pp. 283-289.
Tuross, N., Fogel, M. L., and Owsley, D. W. (1989). Tracing human lactation with stable
n i t r o g e n isotopes. 2. Studies with subfossil h u m a n skeletal tissue. American
Anthropological Anthropological Association Abstracts of the 88th Annual Meetin~ pp.
180-181.
Ubelaker, D. H. (1992a). Patterns of demographic change in the Americas. Human Biology
64: 361-379.
Ubelaker, D. H. (1992b). Porotic hyperostosis in prehistoric Ecuador. In Stuart-Macadam, P.,
and Kent, S. (eds.), Diet, Demography, and Disease: Changing Perspectives on Anemia,
Aldine de Gruyter, New York, pp. 201-217.
Ubelaker, D. H. (1995). Latest developments in skeletal biology and forensic anthropology.
In Boaz, N. T, and Wolfe, L. D. (eds.), Biological Anthropology: The State of the Science,
International Institute for Human Evolutionary Research, Oregon State University Press,
Corvalis, pp. 91-106.
Ubelaker, D. H., Katzenberg, M. A., and Doyon, L. G. (1995). Status and diet in precontact
highland Ecuador. American Journal of Physical Anthropology 97: 403-411.
Van Vark, G. N., and Schaafsma, W. (1992). Advances in the quantitative analysis of skeletal
morphology. In Saunders, S. R., and Katzenberg, M. A. (eds.), Skeletal Biology of Past
Peoples: Research Methods, Wiley-Liss, New York, pp. 225-257.
Verano, J. W., and DeNiro, M. J. (1993). Locals or foreigners? Morphological biometric and
isotopic approaches to the question of group affinity in human skeletal remains recovered
from unusual archeological contexts. In Sandford, M. IC (ed.), Investigations of Ancient
Human Ttssue: Chemical Analyses in Anthropology, Gordon and Breach, Langhorne, PA.
Watdron, T (1994). Counting the Dead." The Epidemiology of Skeletal Populations, John Wiley
and Sons, Chichester, England.
Walker, P. L. (1978). A quantitative analysis of dental attrition rates in the Santa Barbara
Channel area. American Journal of Physical Anthropology 48: 101-106.
Walker, P. L (1986). Porotic hyperostosis in a marine-dependent California Indian population.
American Journal of Physical Anthropology 69: 345-354.
Walker, R L (1989). Cranial injuries as evidence of violence in prehistoric southern California.
American Journal of Physical Anthropology 80: 313-323.
Walker, P. L, and DeNiro, M. J. (1986). Stable nitrogen and carbon isotope ratios in bone
collagen as indices of prehistoric dietary dependence on marine and terrestrial resources
in southern California. American Journal of Physical Anthropology 71: 51-61.
Walker, P. L, and Erlandson, J. M. (1986). Dental evidence for prehistoric dietary change on
the northern Channel islands, California. American Antiquity 51: 375-383.
Walker, P. L., and Hollimon, S. E. (1989). Changes in osteoarthritis associated with the
development of a maritime economy among southern California Indians. International
Journal of Anthropology 4: 171-183.
Walker, E L , and Lambert, P. (1989). Skeletal evidence for stress during a period of cultural
change in prehistoric California. In Capasso, L (ed.),Advances in Paleopathology, Journal
of Paleopathology Monograph 1, "Ibramo, Italy, pp. 207-212.
Weaver, D. S. (1981). An osteological test of changes in subsistence and settlement patterns
at Casas Grandes, Chihuahua, Mexico. American Antiquity 46: 361-364.
Skeletal Correlates of Human Behavior 251

Wetterstrom, W. (1994). Food, diet and population at Arroyo Hondo Pueblo. In Sobolik, K.
D. (ed.), Paleonutrition: The Diet and Health of Prehistoric Americans, Southern Illinois
University Center for Archaeological Investigations Occasional Paper No. 22, Carbondale,
pp. 280-293.
White, "1~D. (1992). Prehistoric Cannibalism at Mancos 5MTUMR-2346, Princeton University
Press, Princeton, NJ.
Wienker, C. W. (1995). Biological anthropology: The current state of the discipline. In Boaz,
N. T., and Wolfe, L. D. (eds.), Biological Anthropology: The State of the Science,
International Institute for Human Evolutionary Research, Oregon State University Press,
Corvalis, pp. 251-267.
Wilkinson, R. G., and Norelli, R. J. (1981). A biocultural analysis of social organization at
Monte Alban. American A n ~ i t y 46: 743-758.
Wilkinson, R. G., and van Wagenen, K. M. (1993). Violence against women: Prehistoric
skeletal evidence from Michigan. Midcontinental Journal of Archaeology 18: 190-216.
Willey, P. (1990). Prehistoric Warfare on the Great Plains: Skeletal Analysis of the Crow Creek
Massacre l/ictims, Garland, New York.
Willey, P., and Emerson, T E. (1993). The osteology and archaeology of the Crow Creek
massacre. Plains Anthropologist 38: 227-269, Memoir 27.
Witley, P., and Hofman, J. L. (1994). Interproximal grooves, toothaches, and purple
coneflowers. In Owsley, D. W., and Jantz, R. L. (eds.), Skeletal Biology in the Great Plains:
Migration, Warfare, Health, and Subsistence, Smithsonian Institution Press, Washington,
DC, pp. 147-157.
Wolff, L (1892). Das gesetz der Transformation der Krochen, A. Hirchwild, Berlin.
Wood, J. W., Milner, G. R., Harpending, H. C., and Weiss, K. R. (1992). The osteological
paradox: Problems of inferring prehistoric health from skeletal samples. Current
Anthropology 33: 343-370.
Zimmerman, U J., and Bradley, L. E. (1993). The Crow Creek massacre: Initial Coalescent
warfare and speculations about the genesis of Extended Coalescent. Plains Anthropologist
38: 215-216, Memoir 27.

Anda mungkin juga menyukai