Sie sind auf Seite 1von 65

ZOOARCHAEOLOGY: AN INTRODUCTION

ACKNOWLEDGEMENTS
DEDICATION
SECTION ONE: AN ORIENTATION TO
ZOOARCHAEOLOGY
Section 1 Introduction
Chapter 1: Introduction
 Part 1
 Part 2
 Part 3
Chapter 2: The Development of Zooarchaeology
 Part 1
 Part 2
 Part 3
Chapter 3: A Perspective on Zooarchaeological Analysis
 Part 1
 Part 2
 Part 3
References [for entire book manuscript]
about course | lecture schedule | prerequisites | course work | required texts |
E-Reserves readings | Zooarchaeology: An Introduction | optional readings at McHenry
Reserves |
illustrative materials | home

CHAPTER ONE
Part One
INTRODUCTION
For over two centuries, bones, teeth, shells, and other animal remains have been
central evidence in research on the earth's past, and on the history of humans and their
ancestors. Physical properties of such animal remains have remained uniform over the
span of hominid existence, and they can thus equally well shed light on carnivorous
habits of Pliocene hominids at Olduvai Gorge or inequalities of diet among the
inhabitants of Thomas Jefferson's plantation at Monticello. Despite faunal remains' time-
honored role in investigating the human past, the last thirty years have seen an
remarkable expansion of the range of issues that faunal materials, especially bones and
teeth, have been used to address. The literature in zooarchaeology is growing at such a
rate that even specialists have difficulty keeping pace. Zooarchaeologists are currently
debating what kinds of information faunal remains can provide, how much, and how
reliably, as well as which analytic methods are appropriate for analyzing them. Many of
the most polarized of recent discussions in human origins research, as well as in the
archaeology of later peoples, has been over the methods applied faunal analyses,
rather than simply over the conclusions drawn from them.
This recent surge of productivity in zooarchaeology results from archaeologists literally
taking faunal analysis into their own hands. It was not until the late 1960s that people
trained in archaeology acquired the skills necessary to work with fauna themselves,
replacing specialists in zoology and paleontology as the primary analysts of
archaeological faunal samples. As they did, they redefined the ways bones, teeth, and
other faunal remains, placing greater emphasis on what they could say about ancient
humans and their activities.
Some of the breakthroughs in zooarchaeological research have been technological. For
example, stable isotope analyses of human and nonhuman bones (Price 1989, Price, et
al. 1985, Sillen and Kavanagh 1982, Van der Merwe and Vogel 1978) have permitted
new refinements in the study of diet and resource use. Scanning electron microscopy,
another major technological advance in faunal analysis another major technological
advance in faunal analysis, has defined "signatures" of non-human versus human bone
processors (Potts and Shipman 1981, Shipman 1981a). However, many of the most
exciting developments in zooarchaeology have not stemmed from innovative
technological applications. Instead, they were assertions that evidence long perceptible
in faunal assemblages can testify to aspects of human behavior hitherto not deemed
accessible.
It is true that studying faunal remains requires knowledge drawn from zoology and
paleontology, but using them to capture evidence of past human choices and actions
calls for a unique range of supplemental strategies. Like paleoethnobotany (Pearsall
1989, Piperno 1988), archaeological faunal analysis combines natural historic methods
with approaches drawn from anthropology and other social sciences. The term
"zooarchaeology" (Olsen 1971) is thus an apt description for faunal analysis to address
archaeological questions.
For Whom is this Book Intended?
This book is intended both for students and for a broad range of interested professional
archaeologists, both within and outside of academic and museum settings. In the early
1990s, I conducted a survey of subscribers to the Zooarchaeological Research News,
(ZRN ) a highly informative newsletter reporting on new studies, publications, and
conferences (Gifford-Gonzalez 1993, Gifford-Gonzalez 1994). At the time, ZRN was
taken by 146 non-institutional subscribers in North America, and an impressive eighty-
five percent of those contacted responded to the survey. About a third of subscribers
worked in government or cultural resource management, and the balance were based in
outside universities and museums. Some were students, but many were earning a living
at least in part by analyzing animal remains and subscribed to keep up with the
literature.
Not surprisingly, the vast majority of subscribers based in the U. S. and Canada
described themselves as North Americanists, dealing with either prehistoric Native
American sites or historic sites created by a diversity of immigrants and native peoples.
Only five percent of ZRN respondents characterized themselves as working on pre-
modern hominids. The latter finding gave me cause for reflection on a seeming
disjunction between those who do most of the mainstream publication and those who do
most of the practice. If one were to read only the Journals of Archaeological Science,
Anthropological Archaeology, and zooarchaeological books written since 1970, one
might conclude that most zooarchaeologists were interested in the earliest African
hominids, in archaic Homo sapiens , and in late palaeolithic hunters of Africa, Eurasia,
and the Americas. Researchers with these interests have been most frequently
represented in print and have contributed immensely to zooarchaeological theory and
method. Some of these authors may not subscribe to ZRN. However, the
preponderance of zooarchaeological researchers in North America, the English-
speaking world, are not working on paleoanthropological or even Paleoindian topics.
These facts led me to ask to three questions. First, why has the cutting edge of
zooarchaeological method and theory been nearly always in the hands of the
paleoanthropological and Paleoindian researchers? Second, are methodological
advances made these areas transferable to studies of anatomically modern peoples,
including in the historic period? If so, has this information transfer actually taken place?
I think the answer to the first question is that researchers working with the sparse
archaeological evidence typical of earlier Pleistocene, and even Paleoindian, sites tend
to try to wrest every last scrap of information from their materials, often ingeniously,
sometimes over-ambitiously.
In response to the second question, whether these findings be useful in other, later
settings, I can vouch from personal experience that it is so. I have spent the last 30
years researching the subsistence choices and environmental context of Africans who
lived only a few millennia ago. I have found the paleoanthropological debates and
findings on bone modification and taphonomy to be invaluable aids to my research.
Now, as I begin working in the U.S. Southwest on historic period materials, I find this to
be even more true, and producing intriguing results, of which I will discuss later in this
book.
Finally, with regard to the final question, has there generally been enough crossing-over
of these findings into research on later periods, I think this has not been the case to
anywhere near the degree that it should. There is a need to link the recent literature to
the research agendas of people who deal with later prehistoric or historic faunas. Why
this has been the case might take another book to explain. The esoteric nature of some
of the literature, and the mutual parochial attitudes of all parties -- an unwillingness to
read across the literature produced by other people working on other periods and
problems -- surely has contributed.
This book is attempt to bridge between recent developments in zooarchaeology of older
time spans and that practiced by the majority of archaeological faunal analysts. It seeks
to condense and describe the last three decades' work in zooarchaeological and
taphonomic research and to show its points of relevance to the archaeology of recent as
well as very ancient humans. This book is thus intended for people who may or may not
be academically placed, and who want to be informed of the overall features of a swiftly
changing field. It is also hopefully useful to archaeology students interested in what one
can learn from animal remains, especially, bones and teeth, and who may dream of
even further expanding the scope of zooarchaeology in their lifetimes.

CHAPTER ONE
Part Two
Scope and Focus of the Book
The uses to which animal remains from archaeological sites are put today may be
grouped as follows:
1. To establish age of deposits (chronology)
2. To reconstruct paleoenvironment and paleoecological relations.
3. To reconstruct human diet and behavior.
These three areas were in fact all under investigation in the 18th and 19th century
faunal analyses, some in ways that critically influenced opinions about human antiquity
and gave rise to archaeology as a discrete discipline. Since methods for reconstructing
human diet and behavior have undergone the greatest growth over the last thirty years,
most of this text will deal with the third area
This book deal with what I know best: vertebrate zooarchaeology, and within that,
analysis of mammalian bones and teeth. A literature on both identification and analysis
of fish from archaeological sites exists (Brinkhuizen and Clason. 1986, Casteel 1976)
(Wheeler and Jones 1989), as does that for birds (Carey 1982, Dawson 1969, Howard
1929). Cheryl Claassen's work (Claassen 1998) book on analysis of molluscan remains
offers a fine theoretical overview.
This book's main focus is archaeological faunal analysis as practiced in Canada and the
United States, with some emphasis on its practice in Britain and other parts of the
English-speaking world. This linguistic grouping forms a logical unit, not only because of
ease of communication in a common language but also because of the shared
perspectives noted earlier. From my own experience as a non-Americanist researching
and teaching overseas, I am aware of the productive work of and linkages among
zooarchaeologists in the Americas and their counterparts in Europe, Africa, and Asia.
These will be mentioned briefly as relevant.
This text is not a guide to identification. Many visual guides to faunal identification exist,
including those of (Brown and Gustafson 1979, Gilbert 1973, 1980, Gilbert, et al. 1981,
Lawrence 1951, Olsen 1960, 1968, 1972a, 1972b, 1973, Pacheco Torres, et al. 1986,
Parmalee 1985, Schmid 1972, Walker 1985) Additionally, textbooks of veterinary
anatomy (Barone 1976, Sisson and Grossman 1975) can provide valuable information
on the relation of soft tissues to bone, as well as details of domestic animal osteology.
Neither is this book a "how-to" manual. Instead it focuses on reviews the considerations
that underlie decisions about identifying, recording, and quantifying various classes of
zooarchaeological data. Most of us know that as no two sites are alike, and digging
them according to a rote formula rather than in consideration of their special aspects
can spell irredeemable loss of data. So, too, no two faunal assemblages are identical,
nor are the conditions under which a zooarchaeologist must do a specific analysis.
Thus, a knowledge base that permits one to formulate a systematic but appropriate
research plan is more useful than any formally prescribed "program."
Moreover, zooarchaeology is currently in a state of rapid theoretical and methodological
development, some of the central issues discussed in this text may be modified or
dispensed by new research within a few years. Because of these factors, I deem it more
useful to review factors that must be taken into account at various stages of analysis,
rather than pushing my own or anyone else's detailed formula for archaeological faunal
analysis. This is not to say that we don't need standardization of classification and
scrupulous attention to conservation and documentation. I simply mean that, at higher
levels of research design, there must be the flexibility that no one-size-fits-all approach
can give. If pressed for the proper procedure for working with fauna, I'd say,
"thoughtfully."
Persons familiar with the zooarchaeological literature may well ask how this book differs
from my colleague R. Lee Lyman's (Lyman 1994) recent book, Vertebrate Taphonomy.
This is a fair question, especially since Lyman and I have similar theoretical and
methodological perspectives. I believe the fundamental difference between our books is
not orientation but focus. I agree with Lyman (1994:33) that hominid modifications to the
remains of animals fall under the larger rubric of taphonomic effects, and that there is no
a priori reason to set human effects apart from the impacts of all the other agents that
can leave their marks on bone, shell, or other organic remains. This book, narrows the
focus from taphonomy as a whole to the practice of zooarchaeology because most
archaeologists, including those who work with animal remains, ultimately want to study
human behavior and its context. From this perspective, the main questions to be
addressed in this book are: what can animal remains from archaeological contexts tell
us about the people who manipulated them or about the environmental context in which
those people lived?
This does not mean that the effects of other agents -- carnivores, rodents, wind, rain,
floods, sunlight, falling rocks -- will be ignored. The zooarchaeologist's job is to draw as
much information as possible from faunal materials. Part of this process involves
distinguishing evidence of the action of various forces, including non-human ones, on
animal remains. Specifying the effects of non-human agents on a sample does not stem
from the analytic aim of correcting for their effects, or "unbiasing" the sample. As will be
discussed in Chapter Three, this is an unrealistic goal. Rather, it stems from the view
that all evidence of pre- and post-mortem agency in archaeological faunal samples tells
us something about the past. A substantial part of this book will therefore deal with
recognizing the traces of the action of non-human bone-modifiers, and the implications
of their effects for zooarchaeological analysis.
Likewise, it is fair to ask how this book differs from my colleagues Betsy Reitz's and
Elizabeth Wing's recent compendium entitled, Zooarchaeology (Reitz and Wing 1999).
In that book, as in this one , the focus is on humans and their interactions with animals,
as testified to by zooarchaeological materials. I believe my approach overlaps
considerably with that of Reitz and Wing, because we work with similar methods and
techniques and approach similar problems. If I were to enumerate some differences, I
would say that I will devote much less time to invertebrates, relatively less time to laying
out ecological theory and basic comparative anatomy, and relatively more to presenting
an epistemological framework in which to place and coordinate different kinds of
zooarchaeological knowledge and practice. Thus, I believe that our respective books
are largely complementary.
This is a methodologically focused book. It discusses the logic of zooarchaeological
method, including uniformitarianism, the potential and limits of analogy, how one's
analytic categories, sorting choices, quantification and statistical analysis, and other
analytic choices can ultimately affect inferences. Methods, however, are developed
within theoretical perspectives, explicit or implicit, and this text will also explore
theoretical perspectives that stand behind the methods examined.
It is fair to ask from what theoretical position I have written a methodologically focused
book. The answer to that question is complex, breaking down into at least two parts,
one linked to my view of the archaeological project in general, the other involving the
specific theoretical viewpoints I bring to my own research in zooarchaeology. Let me
deal with each in turn.
First, much of my published work could be classed as "methodological" rather than
"theoretical," if one believes that conceptual life in archaeology can be divided into
distinct realms of between "high" or general theory, lower (middle-range or interpretive
theory), methods on a lower rung, and, finally, techniques (Kluckhohn 1940). I have
chosen to spend much time working through archaeological reasoning, analogy, and
uniformitarian assumptions. These are concerns that David Clarke (Clarke 1979) called
"archaeological metaphysics," that is, how do archaeologists reason, why do we feel
that certain steps in reasoning are sound? Why and how are some interpretations or
scenarios more plausible than others? How do we check ourselves? Such issues, I
believe, have been at the core of a number of the debates in the last thirty years'
zooarchaeological literature. Chapter 3 of this book deals with some of these issues
more fully.
With relation to theory, it is clear when one studies past and present arguments in
archaeology that, despite the paradigm shifts that seem to come in English-speaking
circles every three decades, some kinds of knowledge are cumulative and enduring.
This lasting knowledge entails both theory and practice and is so fundamental that we
archaeologists usually don't even attend to its unique qualities. This type of knowledge
differs somewhat in emphasis from the archaeological "craft" knowledge discussed by
Shanks and McGuire (Shanks and McGuire 1996), but these fields overlap. For
example, the principles and practices of stratigraphy, passed down to us from 18th
century field geologists, and the more recent archaeological applications of radiometric
dating, are such enduring components of archaeological knowledge and practice.
These lasting knowledge fields appear to stem from specifiable uniformitarian relations
and to link to other areas of scientific research that continue to contribute new and
useful refinements over time (e. g. AMS versus conventional radiocarbon dating). They
build upon earlier understandings in a logical way, and they are cornerstones of
archaeological reasoning, as both examples noted above aid ordering objects and sites
in time, a fundamental archaeological operation. They survive from one archaeological
paradigm shift to the next and are accepted as valid and may be used by people with
disparate theoretical agendas. For example, prehistorians continue to debate how best
to understand the role of Stonehenge among the prehistoric groups who built and used
it, but all will include radiometric dates and stratigraphic reasoning as unambiguous
components in their arguments.
Zooarchaeology is not yet on equal footing with the examples cited above. Yet I believe
it is similar to them and in the process of developing. Like them, it focuses on materials
with many uniformitarian properties, both in circumstances under which humans,
animals, plants, and geological processes interact with those materials and in the range
of evidence produced by those interactions. Zooarchaeology is linked to anatomy,
physiology, zoology, ecology, paleontology, veterinary and nutritional science, in
ongoing and fruitful ways. Based on this view of zooarchaeology as an emerging field of
theory and practice, I seek to lay out here some of its fundamental building blocks as it
has developed in the English-speaking tradition of North America and in some other
parts of the world. Thus, I hope this book will, like a handbook on stratigraphic
documentation and analysis, prove useful to students and researchers engaged in a
range of investigations from a variety of theoretical perspectives.
To take up the second question about theory mentioned before, what "general theory"
(following Binford 1977) do I use when thinking about my zooarchaeological data? For
about twenty-five years, I have been analyzing faunas from East African sites with some
of the earliest pastoral livestock in the region (Gifford-Gonzalez 1998a, Gifford-
Gonzalez 1998b, Gifford-Gonzalez 2000). I have read widely about and spent time in
the field with modern pastoralists. This literature is written from a variety of theoretical
perspectives, which I have assessed for their usefulness and fit with the way I think the
world works. I have also read a good deal of ecology and evolutionary biology.
I currently find two theoretical perspectives useful in thinking about the issues with
which I am most concerned: evolutionary-ecological (including behavioral ecological)
theory, which sees human behavior over the long term as making sense in evolutionary
terms, and structural marxist theory, which focuses on humans' relations of resource
control and acquisition, of production and distribution on a shorter-term time scale. My
own experience forces me to view pastoral stock owners in the context of regional
ecosystems and the non-negotiable demands that weather and the welfare of herd
animals make on them. Likewise, my personal experience compels me to see
pastoralists as participating in social, economic, and ideological systems that both
mediate and clash with environmental trends, and which affect their choices in
managing their livestock, households, and social relations.
Sometimes these rather disparate theoretical worlds are remarkably compatible, since
behavioral ecology and marxist paradigms share a fundamentally economic approach.
Both have an concern with the costs and benefits of effort exerted by humans in
achieving goals within a social milieu -- but in very different ways and with somewhat
different currencies. Sometimes the two paradigms grate against each other in
troublesome but interesting ways. This book will not pursue these aspects of theoretical
consonance and dissonance. However, readers will note that my perspective is one in
which "production" be both an ecological and a social concept, that people and animals
fit into both conceptual worlds, and that animal remains might "make sense" in both as
well.
Moreover, readers should be aware that I do not consider contradictions emerging from
juxtaposition of these two frames of reference (Binford 1987) are just cause for
discarding one in favor of the other. In fact, such contradictions are, at least for now, all
the more motivation to continue working with both paradigms. As regards humans,
theory is not something we archaeologists, with our unique time perspective and
samples, should expect to get "off the shelf" from another discipline. To treat any
imported body of theory as completely functional from the outset or, worse, as divine
revelation, does a disservice to the distinctiveness of our own discipline and excuses us
from the work of critically evaluating the terms and uses of theory in the context of our
own practices.
It may be helpful to elaborate a bit on the concept of a framework noted earlier in
discussing how this book differs from that of Reitz and Wing, and why I have found it
important. In 1991, I wrote an article which I hoped would clarify what I saw as some
then-murky issues in assigning agency to zooarchaeological traces (Gifford-Gonzalez
1991a). I found it helpful to me in my own thinking to move systematically from the
concrete individual specimen with its condition, surface modifications, and so forth out
to the fullest kind of contextual reconstruction one might hope to make of past human
life, including social relations and ecological relations. I argued that bones were not
enough to get us to the higher levels of inferences, that, for a variety of philosophically
grounded reasons one always had to play off the significant information derived from
faunal assemblages against other types of archaeological and contextual data. In the
process, I proposed a schematic framework (Gifford-Gonzalez 1991: Figure 2) with
which to approach zooarchaeological materials and the inferences we wish to draw from
them. That orientation organizes this book and will be discussed in detail in Chapter
Three.
What is relevant here is that I believe the strength of such an approach is that it allows
us to fit new information &endash; whether we gather it ourselves or derive it from
another researcher's efforts, into an intelligible and enduring intellectual structure. I
believe that novel discoveries meriting detailed attention today soon become folded into
the methodology of scientific practice as "givens." Likewise, many of specific
controversies current in zooarchaeology will be passé even by the time this book is
actually published. Therefore, the most important contribution a book like this can make
is to provide a structure for organizing knowledge, both old and new, both in terms of
enabling an understanding of the practical significance of a newly discovered finding
and in terms of facilitating systematic research using it and other resources in a
coordinated way. My hope is that this book will provide students of zooarchaeology with
such a framework, that will be helpful to them long after most of the hot topics of the text
are old news.

CHAPTER ONE
Part Three
Some Basic Definitions
Zooarchaeology, defined as the study of animal remains to elucidate archaeological
questions (Olsen 1971), is one of several disciplines that study faunal remains.
Zoologists study living organisms, but they sometimes use shells, skins, teeth, and
bones to assess the age, sex, health status, and taxonomic relationships of individuals,
species and regional populations. Paleontologists study the preserved remains of
ancient animals to learn more about their evolution, systematics, and ancient life
relationships. Zooarchaeologists study faunal remains from deposits created by humans
at some time in the past, prehistoric or historic. In fact, archaeologists working with
animal remains have labeled themselves with a variety of names, including
archaeological faunal analysts, archaeozoologists.
When I began to study animal remains from the standpoint of an archaeologist in the
early 1970s, we called this kind of work "faunal analysis" or "archaeological faunal
analysis." It is still an accurate term. I have chosen to use the term zooarchaeology
throughout this book for a two of reasons. It lends itself more gracefully to use in
adjective and adverb forms, and it expresses well the specific type of faunal research
we do: archaeology using animal remains.
Persons preferring the label "zooarchaeologist" tend to be concerned primarily with
what animal remains say about humans' and other agents' interactions with them. They
may spend considerable time describing patterns of bone breakage, cut marks, and
other such modifications and comparing these to remains recovered from other sites.
Many are less concerned with details of the systematics of the animal species in
archaeological deposits and see themselves primarily as archaeologists who happen to
use animal remains as a way of researching human adaptation and history. They
probably have an academic training in archaeology and may have supplemental
background in zoology or paleontology. Most of the researchers cited in subsequent
chapters of this text fall into this category.
Other researchers who work with archaeological faunas are in fact happier with the term
archaeozoologist. Archaeozoologists tend to be more interested in the evolutionary and
ecological status of animals included in archaeological sites, and generally less
concerned with what bones can tell us about details of human behavior and social
relations. They focus on reading the history of certain species, such as wild cattle, from
their remains in Pleistocene and Holocene sites, the morphological transitions from wild
to domestic forms, and the regional variability of ancient domesticates.
Persons more comfortable with the title "archaeozoologist" often received their primary
training in the life sciences and only secondarily in archaeology. They tend to
emphasize archaeological assemblages as local samples of species, to be compared to
other archaeological samples through detailed metrical and morphological analyses.
Human modifications to bones as butchery marks, breakage, and evidence of cooking
are not heavily emphasized (Bökönyi 1984, von den Driesch and Boessneck 1975), nor
are detailed reconstructions of human behavior. Thus, archaeozoologists aim more
toward constructing the zoology of ancient faunas.
Although it oversimplifies a complex situation, most North Americans dealing with
archaeological faunas would probably, if pressed, call themselves out as
zooarchaeologists, most continental Europeans would probably call themselves
archaeozoologists, and British researchers might be evenly divided between the two
labels. (Many, preferring to leave detailed taxonomies for their objects of study, are
content with the label "faunal analyst."). The journal Archaeozoologia published in
Grenoble, France, reflecting the dominant perspectives in Continental Europe. Pioneer
archaeozoological researchers in Continental European obtained their training in
zoology, paleontology, and veterinary sciences, began their archaeological faunal
analyses around the end of World War II, and trained students into the 1980s. They
include the late Joachim Boessneck (Boessneck 1969, von den Driesch and Boessneck
1975), the late Sandor Bökönyi (Bökönyi 1984); the late Elisabeth Schmid (Schmid
1972), Anneke Clason (Clason 1991, Clason and van Es 1993, Lasker 1976), the late
Hans-Peter Uerpmann (Uerpmann 1973, 1978), Pierre Ducos (Ducos 1968), and
others.
Notable exceptions to the foregoing dichotomies exist, since some U. S. researchers
have engage in both kinds of work, including George Frison and his students (Frison
1974, 1978, Frison and Todd 1987), Melinda Zeder (Zeder 1984, 1991, Zeder and Arter
1994), Jane Wheeler(Wheeler 1982), and Donald Grayson (Grayson 1988), and some
European faunal analysts focus more on reconstructing behavior (Grayson and Delpech
1994), often in collaboration with North American researchers. Others have questioned
the underlying assumptions of "animals-first" analyses (Legge 1978). Moreover, with
time, younger researchers on all continents are converging more in their interests, and I
expect that the dichotomies true in the late 20th century will not hold up so well in the
early 21st.
Returning to terms that will be commonly used in this book, it is sometimes useful to
refer to all studies of animal remains, regardless of goals or the disciplinary grounding of
the practitioners. In this book, I will use faunal studies or faunal analysis, to signify any
research with animal remains, whether undertaken by zoologists, paleontologists,
physical anthropologists, or archaeologists, regardless of aims. Mainly to have a word
can readily be used as an adjective, I will use the term archaeofauna to refer to the
faunal remains recovered from archaeological sites. This use follows Grayson (Grayson
1984) and is less cumbersome than phrases like "archaeological faunal remains." I will
use this term for historic archaeological faunas as well as more ancient prehistoric
samples, even though it may not sit so comfortably in a semantic sense at Monticello as
it does as Olduvai.
The term taphonomy has already been used in this chapter and requires some further
definition. This began as a subfield of paleontology. This term is derived from the Greek
words for burial (taphos) and rules or system (nomos). It was coined by Soviet
paleontologist I. A. Efremov (Efremov 1940) to describe studying animal remains to
elucidate their circumstances of deposition, or better define the agencies that modified
them before deposition. Over the same span as Efremov, other paleontological
researchers were also studying the transition of animal remains from death to
depositional context (Weigelt 1989), but the term coined by Efremov subsumes their
work. The field became much more salient to paleontological practice in the 1950s and
1960s, as researchers developed more ambitious goals of paleoecological
reconstructions (Clark and Kietzke 1967, Olson 1966, Olson 1971, Olson 1980,
Voorhies 1969). For a review of the taphonomic literature, see Gifford (Gifford 1981),
Behrensmeyer and Kidwell (Behrensmeyer and Kidwell 1984), and Lyman (1994:12-40).
Considerable overlap exists among concepts and analytic methods in vertebrate
taphonomy and contemporary zooarchaeology. In fact, researchers in each area
regularly communicate and engage in research that blurs the boundaries of the fields.
As noted earlier, from a paleontological taphonomist's point of view, human
modifications to animal remains, and human actions that influence their burial, are just
another array of forces affecting them. From the point of view of an archaeologist,
taphonomic analyses are essential to distinguish between traces of human action and
those of other creatures or natural forces that can affect animal remains. Taphonomic
research will therefore be prominently featured in this book.
The next chapter considers the history of zooarchaeology, focusing especially on North
America, to gain a better understanding of the different strands of interest and activities
that make up current debates. Chapter Three presents a perspective on
zooarchaeological analysis, discussing uniformitarianism and zooarchaeology, exploring
in detail the reasons why faunal remains can be used so widely in time and space to
infer climate, season, age of an animal, its sex and physiological condition, and why we
can so readily diagnose cut marks, the gnaw marks of carnivores and rodents, and
other impacts on bones. It will deal with the limits as well as the extent of uniformitarian
assumptions.
Before turning to a single-minded examination of faunal remains, it is necessary to
acknowledge the importance of other types of evidence for human subsistence and
behavior. Animals remains are only one line of evidence on human environment and
subsistence. The last twenty-five years of anthropological research on people who
gather and hunt have shown the importance of plant foods even in their diets, to say
nothing of in the lives of subsistence farmers or members of horizontally and vertically
integrated complex societies. Thus, paleoethnobotany has equal importance as
zooarchaeology in research on human diet and resource use. Isotopic analyses of
human bone can provide a wealth of specific dietary information to complement faunal
and botanical evidence from archaeological contexts. Ultimately, all biological data are
of limited use if not juxtaposed with artifactual, architectural, and settlement information,
within a well-reasoned analytic framework, a point discussed further in Chapter Three.

CHAPTER TWO
Part One
THE DEVELOPMENT OF ZOOARCHAEOLOGY
Faunal Analysis and the History of Archaeology
Two cases from the beginnings of archaeology show the central role that animal
remains have played in defining questions and providing answers about ancient
humans, as noted in Chapter One. In 1797, deep in the clay deposits dug for making
bricks near the town of Hoxne, in Suffolk, English antiquarian John Frere encountered
stones he believed showed human working, lying below sea shells and bones of
animals which had never lived during documented times in England. The sea shells and
other "marine substances" convinced him that the sea had once encroached upon this
place, after the stone implements had been deposited. The bones compelled Frere to
speculate about human existence beyond the span conventionally allotted to human
history in his day. Referring to their depth and to their position below "extraordinary
bones," including a "jawbone of enormous size," and said that the disposition of the
stone tools tempt us, "o refer them to a very remote period indeed; even beyond that of
the present world" (Frere 1797 in Heizer 1962:70-71). Frere's find was one of many in
the late 18th and early 19th centuries in England, France, Belgium, and Germany that
challenged commonly held models of human origins and antiquity (Daniel 1976;
Grayson 1983).
The bones encountered in such sites raised problems because, by the end of the 18th
century, they and their stratigraphic situation testified so compellingly to interested
scholars. By this time, it was widely accepted among antiquarians that Europe had once
been populated by animal species unknown in documented times. By the time Charles
Lyell published his synthesis of earth history in the early 1830s (Lyell 1830), many also
accepted that the Pleistocene epoch was one of glacial ages, populated by cold-loving
mammals. However, humans were thought to have only appeared in the ensuing
Holocene epoch. This scenario sat comfortably with those who wished to retain the
Judaeo-Christian story of the special creation of humans, as well as with scholars less
inclined to take the biblical story so literally. Human artifacts -- and in some cases,
human bones -- in association with glacial faunas clearly challenged this model
(Grayson 1983).
By the late 18th century as well, it seemed natural to scholars to interpret the
stratigraphic context of the finds as evidence of age and ancient environment and the
associated bones as reflections of ancient environments. By then, these were standard
and uncontroversial geological and paleontological inferences and understandings of
objects in the earth.
The fundamental question came down to whether humans coexisted with other species
in the Pleistocene epoch. Bones and stratigraphy, which had given rise to the
controversy, were also seen as key evidence in settling it. In 1857, the Royal Society of
Great Britain commissioned a formal attempt to determine whether stone tools and
bones of extinct animals were truly to be understood as contemporaneously deposited.
A group of prominent geologists and antiquaries supervised a controlled excavation of a
limestone cave at Windmill Hill, overlooking the harbor town of Torquay, in southwestern
England (Daniel 1976). The cave was chosen because a thick limestone deposit overlay
and effectively sealed older deposits. Because of this, it was thought unlikely that more
recently made stone artifacts could have penetrated into deeper, bone-bearing layers.
The limestone layer was breached, and artifacts we now refer to the Upper Palaeolithic
were found together with bones of hyenas, reindeer, and other Ice Age fauna. The
combination of stratigraphic context and extinct fauna were considered by the
researchers to be conclusive proof that humans lived in the Pleistocene (Heizer 1962).
Other carefully controlled checks of sites in France and elsewhere produced similarly
compelling evidence for the Pleistocene existence of humans (Grayson 1983) These
two cases reveal that fauna can both serve to present interpretive challenges to
investigators of the human past, and function to resolve the issues they and other
archaeological materials raise.
In the 1860s, these findings and the paradigmatic changes caused by acceptance of
Darwin's and Wallace's theory of organic evolution gave rise to the new and rapidly
professionalizing practice of prehistoric archaeology. Excavations were explicitly aimed
at recovering evidence of very early -- pre-historic -- humans, and vertebrate remains
played a major role in defining the chronology and environmental setting of Pleistocene
sites. In fact, the first chronological ordering of such ancient sites was explicitly faunal,
with Gabriel Lartet's "Age of Cave Bears," "Age of Rhinoceros," "Age of Reindeer"
sequence for southwestern France (Lartet and Christy 1865-1875).
Regional Traditions in Archaeological Faunal Analysis
The different approaches of North American, British, and continental European
researchers to faunal analysis noted in Chapter One are grounded in divergent
academic backgrounds and perspectives. In the United States, Canada, and Australia,
prehistoric archaeology is usually housed within departments of anthropology. In Britain
and on the European continent, as well as in Asia and Africa, prehistoric archaeology is
not normally allied with anthropology. This distinction has strongly influenced the course
and tone of archaeological faunal studies in these regions.
In North America and Australia, the connection between ethnography and the
archaeological study of the continent was apparent to early academic researchers.
Aboriginal peoples still existed when academic departments dealing with regional
archaeological materials were established in governments and universities, and
researchers considered these aboriginal peoples to be descendants of those who
created the archaeological sites. In some places and at some times, establishing the
link between living aboriginal peoples and archaeological remains became part of a
political agenda of anthropology (e.g. Boas 1940, Trigger 1989). Similar linkages
between ethnology and archaeology were also forged in other, Spanish-speaking
nations of the Americas such as Mexico, Guatemala, Peru, and Chile. In many such
contexts, however, the academic curriculum incorporated more historical analysis than
did anthropological archaeology in English-speaking countries (Lorenzo 1981).
Since the 1920s, North American archaeology has been practiced within departments of
anthropology. Even when North American archaeologists have worked beyond their
national borders, they have taken with them a perspective colored by their position in
anthropology. Until the late 1960s, archaeology was seen as a dependent, theory-
receiving subdiscipline of ethnology, or cultural anthropology (Willey and Sabloff 1983).
As the result of their training, North American archaeologists have been more inclined to
look for "the Indian behind the arrowhead," that is, to look for what archaeological
remains can say about ancient lifeways, including social relations, than their British or
Continental peers. Because old-style ethnographies in the North American tradition
emphasized the role of material culture in the everyday and ritual lives of aboriginal
peoples, archaeologists could find a reasonable link between the Boasian view of
cultural practice and their own work. This could lead archaeologists to believe that
social and cultural information was implicit in archaeological remains, if only it could be
accessed. This theme was first explicitly articulated by Walter Taylor (Taylor 1948) and
then restated with much conviction by the "new archaeologists" of the 1960s (Binford
1962). However, it is implicit in earlier discussions of Southwestern pottery traditions
(Kidder 1917, 1927).
At some point in their training, North American archaeologists also had to grapple with
the term "culture" as something more than simply a classificatory term. They have had
to sort through the contesting definitions and connections of the word in anthropological
theory. Some may even have made a considered decision about which of these many
definitions suits archaeological practice better than others. Until recently, British and
European archaeologists had no occasion to read the anthropological literature on the
relation of the individual to cultural norms, social action, ethnic identity, the dynamics of
cultural and social change, and other aspects of cultural theory. Although rare
exceptions like Grahame Clark (Clark 1953) existed, archaeological training in Britain
did not lead students toward ethnography and ethnology.
In Britain and on the Continent, the link between ethnology and prehistoric archaeology
seemed less "natural." The ethnographic literature was being generated overseas in
colonized areas. Archaeology was at first locally focused on Europe, actually on
charting the course of development of modern European nations through a narrative
grounded in the dim prehistoric past (Trigger 1989). In the late 19th and early 20th
centuries, as archaeological research became institutionalized, a distinction was often
drawn between the histories of less developed, "backwards" colonized peoples and that
of the "advanced" cultures of Europe. Some early popularizers of archaeology did cite
examples from ethnographic cases of "savages" to flesh out the stages of human
progress, as read from the European archaeological record (Avebury 1865, Mortillet
1897). However, neither then nor later was a systematic study of ethnography and
ethnological theory developed from anthropological cases, brought to bear in
interpreting archaeological materials.
In France and Central Europe, prehistoric archaeology has until quite recently been
more closely allied to geology and paleontology than to history or the social sciences, a
situation that has influenced archaeological interpretations down to the present
(Audouze and Leroi-Gourhan 1979; Sackett 1968, 1978). Faunal remains from
archaeological deposits were seen primarily as chronological and environmental
indicators. Ancient economic and social relations were not considered to be readily
accessible to study through archaeological materials. Concepts about the mental and
social capacities of ancient hominids, and even of earlier members of our species are
either not articulated at all or emphasize stasis and predictable levels of complexity,
depending upon the evolutionary stage of development.
Some notable exceptions to the general Continental pattern do exist. In Scandinavia, a
more practically-oriented school of archaeological interpretation developed early in the
19th century. Artifacts and sites were studied for what they could tell about ancient
environment and human behavior, rather than simply as representatives of a stage of
progressive development (Gråslund 1981, Trigger 1989). Long before English-speaking
archaeologists began to feature such work, Scandinavian researchers emphasized
experimental replication and use of archaeological materials, understanding ancient
sites and artifacts through modern peasant analogues, and limited use of overseas
ethnographies. Early interpretations of Mesolithic "kitchen-midden" sites as habitations
with trash deposits exemplify the Scandinavian approach (Trigger 1989).
English prehistoric archaeology initially developed in much the same pattern as it did in
France; faunal and artifactual remains were used mainly as to date sites and infer
environmental context. However, by the 1930s, a number of British researchers at both
Oxford and Cambridge began to experiment with putting ancient humans into the
ancient British landscape. In the 1930s, Harry Godwin at Cambridge began to develop
palynology into a systematic reconstruction of prehistoric vegetation from Ice Age to
historic times, following pioneering studies by Scandinavian researchers. Godwin and
others were careful to delineate the influence of humans in modifying vegetation
patterns in the agricultural colonization of Britain. In concert with such geographical
research, British archaeologists discussed "man-land relationships" (Burkitt 1933; Clark
1938) prefiguring what would later be called ecological archaeology. As articulated by its
leading archaeological exponent, Grahame (J. G. D.) Clark, this approach emphasized
reconstructing the ancient environment in which humans lived. Faunal and other
biological remains were seen as essential to both environmental reconstruction and to
understanding human resource use, or, as the Cambridge school phrased it, "economy,"
within that environmental context (Clark 1952).
British archaeologists' inclination toward environmental and economic reconstruction
was enhanced after the Second World War by new dating methods and analytic
techniques for biological remains. Faunal studies were themselves given a tremendous
boost in the 1950s, when the British government funded a major collaborative research
project on the origins of agriculture, involving several major departments of archaeology.
Based in the Department of Archaeology at the University of Cambridge, the Ancient
Agriculture Project incorporated analysis of archaeofaunas as one of its key strategies.
Eric Higgs (Clark 1952; Higgs 1962, 1972; Higgs and Jarman 1975) and his students
(Jarman, et al. 1982; Sturdy 1975) worked on animal use, both hunting and husbandry,
in a variety of modern and archaeological case studies. They presented techniques for
diagnosing herd management and offtake from faunal assemblages. The focus on new
ways of studying animal utilization also influenced other Cambridge students interested
in earlier phases of human history, such as Africanists Glynn Isaac and J. Desmond
Clark, who in turn emphasized zooarchaeological analysis among their students (Isaac
1971).
Despite this strong interest in "palaeoeconomy," British faunal analysts had little formal
training in anthropology. It was not until the late 1970s that the concept of "culture," in
any of its many articulations, became a subject of discussion and research in British
archaeology (Hodder 1982a, 1982b; Hodder and Orton 1976). Reconstructions normally
focused on the immediate human-animal or human-plant interface, with little
consideration of wider questions about human social relations, cultural variations in
practices, and so forth.
Do the Regional Traditions Communicate?
Given their disparities in training and goals, it might seem that these different regional
traditions in archaeological faunal analysis have little to say to one another. Indeed,
although an international organization, the International Congress of Archaeozoologists
(ICAZ) exists and has members from all regions, and meets every four years, my
experience is that session papers from one regional tradition are often of limited interest
to researchers in other regional traditions.
However, trends encouraging greater communication and commonality of goals have
emerged in the last decade. First, some individuals have managed to work across one
or more traditions and are able to communicate with a variety of researchers in the
North American and European traditions. Second, the ICAZ meetings themselves have
brought researchers together in face-to-face interactions, increasing the likelihood that
they might develop an interest in one another's work. Third, some continental
researchers have turned more toward analysis of foraging patterns and other issues
dear to the hearts of Anglo-American zooarchaeologists. For example, the research of
Françoise Audouze and colleagues on late Magdalenian sites in the basin of the Seine
River (Audouze 1987, 1988; Audouze, et al. 1981) features many research questions
and approaches of interest in North American archaeologists. Fourth, certain areas of
research, especially studies of early hominids, have served to bring together previously
unacquainted researchers, often early in their careers. The last twenty years of
zooarchaeological research is thus worth examining in detail, to understand the
backgrounds, commonalties, and divergences of interests of its practitioners.
Streams of North American Research Converge in the 1970s
Much of the convergence of research on prehistoric bones stems from the common
interests of scholars in different disciplines in vertebrate taphonomy. Efremov (Efremov
1940, 1953) and other early workers thought of taphonomy as a subfield of paleontology
(Olson 1980). However, vertebrate taphonomy came to be a focus of great interest for
paleoanthropologists and archaeologists investigating ancient bones as evidence of
early hominid activities, and many recent advances in taphonomy in fact have been
produced by nonpaleontologists. This is partly because paleoanthropological research
in the U. S. and elsewhere received much greater national and private funding than has
paleontological research which doesn't study hominids or ancient primates. Greater
resources have permitted bigger research projects and faster progress. Results from
taphonomic work by any disciplinary group have been useful to the research programs
of the others.
In the English-speaking community, communication across disciplinary boundaries has
been intense and relatively free. Reasons for this are theoretical, practical, and
personal. Many recognized a common interest in the role of bones in ancient biological
and ecological systems, and in the loss of information about such systems due to
postmortem processes acting on bones. As zooarchaeology emerged in the 1970s, so
few people were working on taphonomy and paleoecology that they had to cross
disciplinary boundaries to find colleagues with whom to talk. Because interest in
taphonomy became strong in the late 1960s and early 1970s, many of the first
researchers to do taphonomic research were of roughly the same generation. This
enhanced swift communication, since relations among researchers have been informal
and usually relaxed.
Many taphonomic researchers actually worked together on field teams, despite
attending different universities, concentrating in different disciplines, and even residing
in different countries. Such fieldwork contacts were often within large, well-funded
research projects on early hominids or origins of agriculture. The 1960s were
economically prosperous times for the U.S. and Western Europe, and funding for large-
scale, interdisciplinary research on human origins, the emergence of food production,
and the development of complex societies was at its height. Over this period and into
the 1970s, the U. S. National Science Foundation and private foundations gave millions
of dollars to such large interdisciplinary research projects, which normally included
faunal analysts. These projects brought together students from varied institutions,
disciplines, and regional traditions. For example, Andrew Hill, a student of English
geologist W. W. Bishop at the University of London, Kay Behrensmeyer, a graduate
student in paleontology from Harvard University, and myself, a student of Glynn Isaac at
the University of California, Berkeley, all spent time together at Lake Turkana, Kenya,
associated in different ways with the Koobi Fora Research Project.
Over the same period, other informal networks of taphonomic researchers grew among
archaeologists who studied North American prehistory, especially those concerned with
Paleoindian sites, with the earliest habitation of the Americas. Joe Ben Wheat (Wheat
1966, 1972) produced a landmark example of a detailed exegesis of site formation with
the Olsen-Chubbuck site, a Paleoindian bison kill-butchery locale. Wheat's painstaking
reconstruction of season from multiple sources (pollen, bison age structures), of the
actual hunting tactics from disposition of the bison in their arroyo trap, and of butchery
practices from the disposition of carcass units at the site, as well as his incorporation of
ethnographic and historic analogues, stood as a model for the analysis of other such
sites. George Frison (Frison 1970, 1974, 1978) and his students (Frison and Reher
1970; Frison and Todd 1987; Frison, et al. 1976; Todd 1995) and other North
Americanist researchers investigated mass bison kills in North America, ranging in age
from Paleoindian to protohistoric. Frison's own practical background as a hunter and
cattle rancher gave him a functional outlook on hunting and carcass processing
unequaled in many contemporary works (Frison 1991)
North Americanist archaeologists from the United States and Canada researching the
earliest evidence for human occupation in the Americas have also contributed
substantially to the literature on vertebrate taphonomy. Because some of the earliest
alleged traces of human activity in North America were assemblages from the Yukon
region of Canada, composed of broken and otherwise modified bones that lacked
associated lithics. Canadian archaeologists therefore became interested in determining
the causes of bone modification (Bonnichsen 1973, 1979; Bonnichsen and Bryan 1978,
Bonnichsen and Sanger 1977, Bonnichsen and Will 1980, Frisch 1968, Wilson 1983);
Guthrie 1982, 1984; Morlan 1980, 1983, 1991). Their work included a good deal of
experimental replication of bone modification.
Other archaeologists who have contributed to discussions on faunal analysis work with
Paleoindian mammoth and bison kill sites. These include Eileen Johnson (Johnson
1978, Johnson 1982a, Johnson 1985) and Dennis Stanford (Stanford 1978; Stanford, et
al. 1981) Most of these researchers began to work with bones early in their careers, but
took a somewhat different path than those working with African, European or Asian
materials. They had less of a background in paleontology and paleoecology, and were
less concerned, at least initially, with reconstructing prehistoric subsistence than were
either Wheat, Frison, and Frison's students or, for that matter, students of "Old World"
paleoecology and paleoeconomy.
By the end of the 1970s, North Americanists began to communicate with researchers
working on African and southwest Asian prehistory, sometimes informally, sometimes at
conferences, such as the 1984 Bone Modification Conference, sponsored by the Center
for the Study of Early Man, University of Maine at Orono (Bonnichsen and Sorg 1989).
As a result of these contacts an international network today exists among English-
speaking scholars that communicates about taphonomy, both formally and informally.
The network now includes persons in Latin America (Mengoni-Goñalons 1980; Borrero
1990) and other regions. Checking recent listings of members of the International
Conference of Archaeozoologists (ICAZ), one finds names of over 100 U. S. scholars
who analyze bones from archaeological and paleontological sites.
It is informative to examine these scholars' intellectual roots in more detail, because this
reveals that the faunal research network has convergent interests but no uniform
research agenda. Researchers from different disciplines often have different questions
in mind. Paleontologists are naturally less interested in distinguishing the effects of
humans on bone assemblages than are physical anthropologists or archaeologists.
More importantly, researchers in various disciplines approach their studies with differing
theoretical and methodological perspectives. This means that, although they
communicate a lot, researchers in this network are pursuing diverse goals, using
overlapping but not identical methods. As a result of this diversity, progress takes place
through a dialectical process. Intense debates on theory and method challenge each
side to develop their research more fully. The main subdivisions are paleontologists,
physical anthropologists, and archaeologists.

CHAPTER TWO
Part Two
Persons with Backgrounds in Vertebrate Paleontology and Zoology
Vertebrate taphonomy has of course been a practice of paleontologists for many years.
Among the most influential older American workers is Everett C. Olson, whose work on
community evolution in the Permian was enhanced by a taphonomic perspective earlier
than almost any other researcher. Olson's friendship with Efremov, and his ability to
speak and read Russian, allowed him to appreciate Efremov's thinking at a time when
few U. S. researchers had contact with their Russian colleagues (Olson 1980). Olson
had direct and friendly contacts with younger paleontologists who began making
taphonomic studies in the 1970s and attended the 1975 Wenner-Gren conference on
taphonomy and paleoecology.
Paul Parmalee, who co-authored the Eschelman site study with Guilday (Guilday, et al.
1962) was very influential in turning Americanist archaeology toward the practice of
zooarchaeology . Parmalee combined his background in zoology and strong interest in
the aboriginal inhabitants of the Midwest to reconstructing the ecological context and
animal diet of prehistoric peoples (1965; Parmalee et al. 1972; see also Purdue et al.
1991 for an overview of his career). As Curator of Zoology at the Illinois State Museum,
Parmalee built a strong zooarchaeology emphasis into the Museum program,
encouraging the more complete recovery of bone from archaeological sites in the region
and stressing the need for accurate taxonomic identifications and the reference
collections that enabled them. He trained and inspired many Americanist archaeologists
there and at the University of Tennesee, Knoxville during his long career (McMillan
1991). Parmalee, along with Guilday, was probably the most influential North American
worker in turning faunal reports away from simple species lists and toward a richer
reading of ecology and human behavior.
Stanley Olsen, by training a paleontologist, began while at Harvard University to
produce valuable keys to the identification of vertebrates from archaeological sites,
continuing his publications from his later post at the Arizona State Museum (Olsen
1960, 1968, 1972a, 1972b, 1973). He was among the first to use the term
"zooarchaeology" in print (Olsen 1971). More recently he has published on the
domestication of horses in China and of dogs (Olsen 1984, 1985)J.
Arnold Shotwell, both through his mentor relations with Grayson, and his written work
on inferring community ecology from element representations in deposits (Shotwell
1955, 1958, 1963).exerted a strong influence on the first generation of
zooarchaeologists. Although Grayson (Grayson 1984) has shown that some of
Shotwell's uses of quantitative data are inappropriate, Shotwell's pioneering work
reflects an early consideration of the role of depositional processes in structuring fossil
assemblages.
C. K. Brain of the Transvaal Museum in the Republic of South Africa conducted his
taphonomic research more or less independently for many years but was helpful to
younger workers training in the early 1970s. His book, The Hunters of the Hunted? An
Introduction to South African Cave Taphonomy (Brain 1981), synthesized many years of
experimental observations and analysis of paleontological samples, in which he had
produced a number of highly influential articles (Brain 1965, Brain 1967, Brain 1969,
Brain 1974, Brain 1975, Brain 1976, Brain 1980). These contacts were formally
recognized, and the circle of researchers expanded, through the 1975 Wenner-Gren
conference on taphonomy and paleoecology, which later resulted in the book Fossils in
the Making (Behrensmeyer and Hill 1980).
Several younger North American vertebrate paleontologists made valuable contributions
to taphonomic theory and method from the 1960s onwards. Michael Voorhies, an
American paleontologist, made valuable contributions to modern vertebrate taphonomy
at the end of the 1960s. Influenced by his mentor Heinrich Toots, Voorhies studied the
taphonomy of a large bone bed in Nebraska (Voorhies 1969a, 1969b). His research
included experiments on transport by water of various vertebrate elements, resulting in
definition of "transport categories," based on bones' hydrodynamic categories. These
were influential models for taphonomic analyses of hominid and hominoid sites in Africa
(Behrensmeyer 1975b; Shipman 1977,1982). Voorhies has more recently contributed to
studies of long bone fracture patterns in fossil assemblages (Myers, et al. 1980).
Among the most productive of the generation of taphonomic paleontologists trained in
the 1970s is Anna K. Behrensmeyer, whose research has often been conducted in
localities yielding early hominoids or hominids in Africa and Southwest Asia
(Behrensmeyer 1975a, 1978a, 1982a). Her dissertation research (Behrensmeyer
1975b) in the Koobi Fora - Ileret region of East Lake Turkana concentrated on the
sedimentary microstratigraphy, paleogeography, and taphonomy of the Plio-Pleistocene
fossil localities and archaeological sites. Behrensmeyer's approach to fossil
assemblages reflects her strong background in geology, both in terms of her attention to
sedimentary context of the bones and her frequent use of modern observations to
enhance understanding of the prehistoric deposits (Behrensmeyer 1982b, 1983). She
has experimented on aqueous transport of bones (Behrensmeyer and Hanson 1975;
Boaz and Behrensmeyer 1976, Hill, et al. 1985), made observations of bone weathering
in natural settings (Behrensmeyer 1978b), and documented the relations of bone
assemblages on land surfaces to ecological factors in an East African ecosystem
(Behrensmeyer and Dechant-Boaz 1980;, Behrensmeyer, et al. 1979). Behrensmeyer's
theoretical work on fossil assemblage formation reflects the interaction of her modern
observations and analysis of fossil assemblages (Behrensmeyer 1982b; Kidwell and
Behrensmeyer 1993). Although she has seldom held a formal teaching post, she has
been informal mentor and colleague to many younger American workers and to
numerous researchers in other countries. With other collaborators at the Smithsonian
Institution, she initiated a program on the evolution of terrestrial ecosystems which
synthesizes information from many phases of earth history (Behrensmeyer 1991, 1992).
Other paleontological taphonomists often read by zooarchaeologists include Daniel
Fisher, of the University of Michigan, who has contributed to studies of microvertebrate
concentrations (Fisher 1981), investigations of elephantid taphonomy in North America
and the influence of humans on those materials, using Scanning Electron Microscope
methods to document both cut marks and age at death (Fisher 1984a, 1984b, 1988;
Fox and Fisher 1994). Catherine Badgley, whose work has focused on taphonomic and
paleoecologic analyses of Miocene faunas in Asia (Badgley 1982, 1986b; Badgley, et al.
1984; Badgley and Behrensmeyer 1980), published a contribution on quantification of
individuals in fluvial fossil assemblages ; see also (Grayson 1984, Holtzman 1979,
Turner 1983). Anthony Fiorillo, contributed valuable experimental and paleontological
information on the origins of "pseudo-cut marks," on bone (Fiorillo 1984, 1987a, 1987b,
1989).
Persons with Backgrounds in Physical Anthropology
Physical anthropologists who engage in zooarchaeological or taphonomic research
include Andrew Hill, Richard Potts, and Pat Shipman. They are persons primarily
concerned with Plio-Pleistocene hominids and, in the case of Hill, Miocene hominoids.
They have worked mainly in east Africa and southwest Asia. Hill received a largely
paleontological and geological training with the late W. W. Bishop at the University of
London. His dissertation research was a taphonomic analysis of modern East African
land surfaces and their bone assemblages (Hill 1975). Hill studied patterns of bone
modification (Hill 1983, 1984, 1989a, 1976, 1982) carcass disarticulation and dispersal
(Hill 1979a, 1980, Hill 1979b, 1989b; Hill and Behrensmeyer 1984).
Potts has worked on faunal materials from Olduvai Gorge and other early East African
sites (Potts and Shipman 1981, Potts 1982, Potts 1983, Potts 1984, Potts 1988). He
has a background in biological anthropology, having studied with Erik Trinkhaus and
Alan Walker, as well as with primatologists at Harvard. The Potts and Shipman joint
publication on cut marks reflects the informal taphonomy network noted above, since
both met while working with the Olduvai materials stored in the Kenya National
Museum.
Shipman, a student of Clifford Jolly at New York University, had a background in primate
evolution and did contemporary observations on the effects of drought and scavengers
on bones (Shipman 1975, Shipman and Phillips-Conroy 1976, Shipman and Phillips-
Conroy 1977). Her dissertation analyzed the Fort Ternan, Kenya, fauna from a
taphonomic perspective (Shipman 1977, 1981b, 1982, 1986a). Shipman's pioneering
use of Scanning Electron Microscope coupled with experimental production of
modifications are well-known (Shipman 1981a, 1981c, 1989; Shipman and Rose 1983a,
1983b), as are her discussions of early hominid paleoecology (Shipman 1986b, 1988,
Shipman and Rose 1983b).
Persons with Backgrounds in Archaeology
Old World Archaeologists
One U. S. "school" in zooarchaeology that began in the 1960s was actually transplanted
from the longstanding tradition of "economic archaeology" at the University of
Cambridge, England. In the 1960s, the University of California, Berkeley, hired three
Cambridge graduates: J. Desmond Clark, Glynn Isaac, and Robert Rodden, excavator
of the early Neolithic site of Nea Nikomedeia in Greece. The three agreed on the need
for archaeologists to develop expertise in faunal analysis to answer specifically
archaeological questions about prehistoric human behavior. Their outlook was
strengthened by the F. Clark Howell's interdisciplinary approach to paleoanthropology
and recognition of the importance of faunal studies in understanding hominid
paleoecology. In the late 1960s and early 1970s, Berkeley students were encouraged to
pursue faunal studies. Most studied with one or more of the vertebrate paleontologists
on the campus: Joseph T. Gregory, William A. Clemens, and Donald Savage, who all
supported acquisition of paleontological skills by students in physical anthropology and
archaeology. Glynn Isaac especially encouraged many of his later students to undertake
experiments and observations of contemporary humans and carnivores to shed light on
evidence from prehistoric localities.
Graduates of this program have been active participants in discussions of
zooarchaeology and taphonomy (Blumenschine 1986a, 1986b, 1987, 1988, 1995;
Blumenschine, et al. 1994, 1996; Blumenschine & Selvaggio 1991; Bunn 1981a, 1981b,
1983a, 1983b, 1989; Bunn, et al. 1988; Bunn & Ezzo 1993; Bunn & Kroll 1986; Crader
1974, 1982, 1983, 1984a, 1984b; Gifford-Gonzalez 1985, 1989a, 1989b, 1991a, 1991b;
1998a; Gifford-Gonzalez & Kimengich 1984; Gifford-Gonzalez, et al. 1999; Marean
1986, 1989, Marean, et al. 1992; Marean & Ehrhardt 1995; Marean & Frey 1997;
Marean & Gifford-Gonzalez 1991; Marean & Kim 1998; Marean & Spencer 1991;
Marshall 1986; 1990; 1994a, 1994b; Marshall & Pilgram 1991, 1993; Marshall, et al.
1984). Other Berkeley PhD. archaeologists who began by working in lithic studies but
whose recent researches have involved taphonomic issues are Nicholas Toth (Toth, et
al. 1989) and Paola Villa (Villa 1989; Villa et al. 1988; Villa and Mahieu 1991). By the
end of the 1970s, many of these students were teaching zooarchaeology at other
universities, training the first group of students who entered zooarchaeology as an
already developed subdiscipline.
From the 1960s through the 1980s, researchers at the University of Michigan put into
the field a series of interdisciplinary research projects that featured, among other types
of analysis, faunal studies. These began with Flannery's important work in the Tehuacán
Valley of Mexico and the Deh Luran Plain (Flannery 1965, 1968a, 1968b). Flannery and
other scholars at Michigan trained a number of archaeological faunal analysts at the
University of Michigan, including Brian Hesse (Hesse 1982, 1990, 1995; Hesse and
Wapnish 1985), Jane Wheeler (Wheeler 1976, 1982, 1984), and Richard Redding
(Redding 1981, Redding 1991; Redding, et al. 1975). These students had the benefit of
working with Carl Hibbard, a vertebrate paleontologist at the University of Michigan.
These researchers have been less concerned in their written work with problems of
theory and method in faunal analysis, and more with practical and theoretical problems
in animal domestication and early food producing economies in southwest Asia and the
Americas.
Harvard University produced several active researchers in zooarchaeology and
taphonomy in recent years, through different academic pathways. Richard Meadow
trained as a Middle Eastern archaeologist and has written papers on general
methodology in zooarchaeological analysis (Meadow 1976, Meadow 1980, 1984, 1989,
1990). Melinda Zeder (Zeder 1984, Zeder 1990, Zeder 1991, Zeder 1995, Zeder and
Arter 1994), who obtained her PhD. from Harvard University, trained as an
undergraduate at Michigan. She, too, has contributed articles on methods of
archaeological faunal analysis and continues work on the zooarchaeology of complex
societies from her base in the archaeozoology program of the Smithsonian Institution.
More recently Paul Rissman (Rissman 1987) has engaged in studies of offtake and
adaptation in the transition from foraging to farming in the Levant.
New World Archaeologists
Another prominent contributor to zooarchaeological research, Donald Grayson
(Grayson 1973, 1979, 1984, 1989), is based at the University of Washington. Grayson
also trained to be an archaeological faunal analyst in the late 1960s and early 1970s
during his graduate work at the University of Oregon. Although concentrating in
archaeology, Grayson had a background in physical anthropology and also studied with
the noted paleoecologist J. Arnold Shotwell. At the University of Washington, Grayson
was graduate mentor to R. Lee Lyman (Lyman 1977, 1982, 1984, 1985, 1986, 1987a,
1992, 1994), now of the University of Missouri at Columbia. Lyman had already begun
to train in archaeological faunal analysis at Washington State University, reflecting the
tendency for archaeological faunal analysis to be taught as formal courses by the
middle 1970s. Other of Grayson's students include Lee Ann Kreutzer, who contributed
to the literature on bone densities (Kreutzer 1992a,1992b), and Jack Broughton, who
has applied behavioral ecological modeling to aboriginal California faunas (Broughton
1994a, Broughton 1994b, Broughton 1994c, Broughton 1997, Broughton and Grayson
1993).

CHAPTER TWO
Part Three
Self-Trained Zooarchaeologists
A number of well-known contributors to discussions of taphonomy and zooarchaeology
are people who developed skills in faunal analysis after some time working as
archaeologists. These scholars are essentially self-taught zooarchaeologists. Best
known is Lewis Binford. After a highly visible career as one of the champions of
American "new archaeology" in the 1960s, Binford began to work with faunal remains in
the early 1970s, as the result of his reflections on archaeological methodology (Binford
1983). He saw bones as a set of study materials that acted in a uniformitarian way
(Binford 1977, 1978, 1981), and that could therefore be investigated in contemporary
settings. Binford did bone-oriented ethnoarchaeology (Binford 1977, 1978, 1981) and
analyses of prehistoric faunal materials from many localities (Binford 1984; Binford and
Stone 1986). Binford has stated the value of studying bones in contemporary settings to
enhance analysis of prehistoric materials. He has developed useful approaches to
"economic anatomy" of mammals, such as the Utility Indices (GUI, MGUI) that now are
subject of much fruitful discussion in faunal analysis (to be discussed in Section Four).
Another well-known self-trained faunal analyst is Richard Klein, who was a student of F.
Clark Howell and Sherwood Washburn at the University of Chicago. Klein's early
research did not involve first-hand work with animal bones, although his training in
paleoanthropology and paleoecology provided a strong background for his later
concentration. Beginning in the 1970s, Klein worked in South Africa on Upper
Pleistocene and Holocene faunas. His work has contributed to understanding the
paleoecology of southern African (Klein 1980) and to humans' role in ancient
environments (Klein 1983, 1989). Klein and some of his students have developed an
approach to ageing animals, reconstructing mortality profiles, and inferring predation
patterns (Klein 1978, 1981, 1982;, Klein, et al. 1981, 1983; Klein and Cruz-Uribe 1984)
which will be discussed in detail in Section Four.
John Speth of the University of Michigan also turned to faunal analysis after an earlier
career in lithic analysis. His work on seasonal needs for fat in the human diet and its
probable influence on predation patterns (Speth 1983, 1988; Speth and Davis 1976;
Speth and Spielmann 1983) has strongly influenced recent interpretations of
archaeological faunal assemblages.
Another essentially self-taught zooarchaeologist and taphonomist is Gary Haynes, who
has studied effects of postmortem processes on bison, deer, and moose in North
America and on elephants and other mammals in Zimbabwe. He supplemented
observations of natural situations with experimental feeding of captive animals (Haynes
1980a, 1980b, 1982, 1983a, 1983b, 1986, 1987a, 1987b, 1988a,1988b, 1991, 1995).
Haynes has also studied prehistoric mammoth and mastodon death sites in both North
America and the Soviet Union, using information derived from his modern observations.
Haynes developed his approach while still a doctoral student. From then on, Haynes
has acted as a very important link between North Americanist and Africanist
taphonomists and zooarchaeologists, and between Russian and English-speaking
scholars.
Is Any One Background Better?
Given that zooarchaeology by definition deals with bones from an archaeological point
of view, any effective practice must deal with human actions and the contexts of those
actions. Thus, it would seem that an exclusively zoological or paleontological
background would be less suitable than one based in anthropological archaeology.
Indeed, anthropological archaeology, as practiced in the U. S. and Canada, provides a
rich context for framing research questions and interpretations. Ethnographic studies
provide us with an idea of the range in human-animal interactions possible among
recent peoples, which sometimes can be augmented by our own ethnoarchaeological
research. Cultural ecological studies provide us with theory relevant to modeling
subsistence strategies of prehistoric peoples. Especially for those of us dealing with
archaeological traces of Homo sapiens sapiens, studies of the symbolic roles of animals
in modern human cultures is also a valuable resource.
However, a paleontological training provides a different kind of rich and valuable
background. In addition to learning about the morphology of animals in a comparative
and evolutionary perspective, contemporary paleontologists assimilate principles of
ecology. These are valuable fields of knowledge for zooarchaeologists seeking to
understand the place of hominids in ancient ecosystems. Moreover, and very
importantly, paleontologists are taught to be circumspect in interpreting their evidence.
Taphonomy itself deals with the processes that intervene between the life states and the
fossil contexts of animal remains. It seeks to reveal the processes responsible for the
patterning in fossil (and, by extension, archaeological) accumulations, so that a
researcher can interpret their meaning with a better sense of whether ancient life
relations created those patterns. Another kind of circumspection comes from the classic
paleontological concern with differentiating similarities in form based on common
evolutionary history versus similarities resulting from convergent evolution. This concern
with similar outcomes of different causes, also known as equifinality, is relevant to
zooarchaeological practitioners, as is paleontological concern with taxonomy and
systematics.
Likewise zoology and ecology, with their interest in living organisms, their classification
and their relationships, provides valuable background for both explicit modeling and
preliminary sorting through more or less plausible hypotheses. Methods developed in
zoology for distinguishing closely related species on the Equally importantly, study of
the ethology and behavioral ecology of animal species can enhance zooarchaeologists'
understanding of how hominids would have had to strategize in manipulating given
species, whether in hunting them, regularly locating their fresh carcasses, or living with
them as mutualists and parasites in food producing societies.
Based on feedback in the Zooarchaeology Research News survey mentioned earlier, it
is possible now, it seems, to "major" in zooarchaeology, hewing to a rather narrow track
within anthropological archaeology. Despite being a teacher of zooarchaeology, I
believe it important to maintain contact with the other life and historical sciences that
deal with animals, as well as with specific disciplines which may be relevant to
situations, for example history and historiography, animal production and veterinary
science, meat science, and nutrition.
While a multiplicity of approaches produce results useful to zooarchaeology, some do
not. Problematic approaches are not so much related to disciplines as to faulty
background in the logic of scientific inquiry. Some earlier analyses of archaeological
materials failed to demonstrate what they alleged because of a lack of understanding of
the analogical nature of archaeological interference or inadequate grasp of the potential
for equifinal results. A number of experimental studies of bone, and archaeological
inferences derived from them, have been criticized for not eliminating alternative
explanatory hypotheses within the research design. Among the most famous of
research areas in which this kind of logical error took place is that of research on spiral
fracture of long bones, which will be taken up in Section Three.
At this stage in the development of zooarchaeology, there is no single best way to
approach archaeological bone accumulations with anthropological questions in mind. All
strategies currently in practice have yielded valuable information. Grayson, with his
strong background in paleontology, has usually asked questions of a biogeographical
and evolutionary nature, and has treated many of the samples he has analyzed as
essentially paleontological in nature. Binford, working from the position that the
archaeological record can provide detailed information on human behavior and
adaptation, has analyzed his collections with behavioral reconstruction in mind. Frison,
with a deep practical knowledge of landscape and animal behavior, has asked
increasingly complex questions about the interplay of hunters and hunted in the
Western Plains. Others have followed their own agendas in designing research and
analyses, most with productive results.
Rather than argue for a unitary approach to vertebrate remains, I believe that diversity
and even dissonance produces many theoretical advances. Researchers who find
themselves trying to span the lack of fit between two related analytical systems may
well be able to expose and explore implicit and unquestioned assumptions in both. The
process of talking past each other may, given the will to ask why such
misunderstandings exist, produce new mutual insights for researchers in overlapping
fields. Knowledge is thus gained through a dialectical process, in which initial
misunderstandings and confusion may lead to more complex and deep understandings
of our objects of study. I personally have found this to be so in my own work, as an
archaeologist trained by paleontologists, as a U. S. archaeologist trained by Cambridge
graduates, as a zooarchaeologist studying recent African pastoralists with methods
devised to study early hominids, as a woman in a field as yet dominated by men. I
therefore advocate occupying, at least occasionally, the uncomfortable spaces along
disciplinary boundaries.

CHAPTER THREE
Part One
A PERSPECTIVE ON ZOOARCHAEOLOGY:
FAUNAL REMAINS AS UNIFORMITARIAN MATERIALS
Chapter Two showed that bones, teeth, and shells have been accepted as reliable
evidence for chronology, ancient environmental setting, and human activities for over
two centuries. Faunal remains have been perceived as such credible evidence of past
conditions because they possess properties that have remained uniform over long
spans of time. Zooarchaeological writings often have discussed bone's uniformitarian
properties (Binford 1981; Bonnichsen and Will 1980; Gifford 1981; Gifford-Gonzalez
1991; Lyman 1987b, 1994, Young 1989). Recent research in zooarchaeology has often
involved experiments with bone and observations of modern non-human and human
bone modifiers with the aim of building casebooks of analogues for evidence in
archaeological samples. Because studying the present to understand the past is such a
central part of zooarchaeology, and because this has been a controversial area, it is
best to begin with an examination of these matters.
Defining Uniformitarianism
The term uniformitarianism was used by English geologist Charles Lyell in the early
1830s, in developing his explanation of the geologic history of the earth. Lyell was
constructing an argument against the "Catastrophist" school of earth historians, who
argued that the discontinuities between faunas in geological deposits and those
documented in the contemporary world were created by a succession of unique events
unknown in our present world, each of which ended a major cycle of life on earth
(Hooykaas 1970). In Principles of Geology (Lyell 1830, Simpson 1970), Lyell argued
that all geological, and by extension, paleontological, deposits on earth could be
accounted for solely by invoking the operation of processes observable in the
contemporary world, rather than events exceptional to it. This assumption is what
George Gaylord Simpson (Simpson 1970) called methodological uniformitarianism.
Simpson points out that, in the span prior to publication of Darwin's and Wallace's
theory of organic evolution, Lyell's also postulated that forms of life did not alter
significantly through earth history, a position Simpson calls substantive
uniformitarianism. Once convinced by arguments for organic evolution as the source of
new species and disappearance of older forms, as well as by the new fossil evidence
and artifact associations, Lyell abandoned the latter position in later versions of his
work. However, Lyell's methodological uniformitarianism became and continues to be a
cornerstone of the historical sciences, including geology and paleontology.
Actualism
A corollary of the uniformitarian position provides historical geology and allied sciences
with a practical research strategy for learning more about the past: one can understand
the origins and nature of deposits in the earth by studying processes that are currently
forming analogous deposits in the present-day world. In historical sciences, and more
recently in archaeology, studying present-day analogues to learn more about preserved
evidence from the past is often called actualism (Binford 1981; Gifford 1981; Herm
1972; Hermans 1977; Hooykaas 1970; Lawrence 1971). To native English speakers,
this term may be puzzling, since "actually" has the meaning of "in fact" in our language.
However, the Latin root of "actual," and the word in contemporary German and
Romance languages mean "of the present." The term, drawn into English from its
original use German paleontology, thus refers to the present as a source from which to
draw meaning to assign to evidence from past.
Reasoning by Analogy
The phrase "modern-day analogues" brings in the third component in an actualistic
research strategy: analogical reasoning. What is the relation of uniformitarian
assumptions, and reasoning by analogy, and actualism? More specifically, how does
this relate to the archaeological study of bones? Faunal remains are reliable indicators
of past processes and contexts only if one takes a uniformitarian perspective. When we
diagnose a developmental stage from a feature such as an unfused epiphysis on a
bone from an ancient site, we are assuming that this feature was produced in the past
by the same ontogenetic processes as produce them in the present. We interpret traces
of an agent of postmortem modification, such as a carnivore tooth mark, on a fossil
bone by assuming such marks were produced by a carnivore in the unobservable past.
At a more fundamental level, even calling a fossil bone a bone and not a rock assumes
that objects of that particular form and internal structure have come into existence in the
same way as modern analogues, as parts of vertebrate bodies, over a long span of
earth history.
Debates over Analogical Reasoning in Archaeology
In the 1960s, some "New Archaeologists" claimed that archaeology should move
beyond analogical reasoning to inference based on deduction (Binford 1967; Freeman
1968). However most archaeologists today accept that both our methods of inference
and even how we know what we know usually rest on such analogies (Binford 1981,
1987; Hodder 1982b). Richard Gould (1980; Gould and Watson 1982) more recently
reasserted that we could "escape" from using analogies by relying on "laws" --
statements of invariable relationships -- derived from ecology, biology, and geology.
Wylie (1982, 1985), a philosopher of science, pointed out that using law-like,
uniformitarian principles is in fact a special form of analogical reasoning, predicated on
the assumption that processes and relationships in the remote past were virtually
homologous to those observed in the present, upon which law-like statements are
based.
One might sum up a perspective on the use of analogy in archaeology in three
statements: Analogy is inevitable. Analogy can be abused. Analogy may be refined by
actualistic research. Let us examine each of these points more closely.
Analogy is Inevitable
Working with bones from the prehistoric past, we use analogy at every step, from
naming the osteological element and identifying the species to inferring details of
ancient environment or ecological interactions.
Let us take an example. When we excavate a fossil bone identical to a right femur of a
modern deer, we simply call it the femur of a deer. We call the fossil a femur based on
its resemblance to modern femora, but we actually go through a complex, although very
fast, set of evaluations of its physical traits. The bone may differ markedly from modern
ones in some features, such as its weight, color, or chemical composition, but we decide
these are not relevant to identifying the bone and the species. These traits instead
reflect processes that affected the ancient bone postmortem, and our first concern in
identification is with its earlier life context. So, in naming the bone we make an analogy
with modern femora of deer, based on criteria of similarity that we think are relevant.
Most paleontologists or archaeologists will pursue inference by analogy considerably
further. Although we did not find any other bones of the deer, we accept that the fossil
femur once existed as part of a skeleton. We also "know" that the femur had certain
specfiable muscles and ligaments attaching to it (the quadruceps femoris and not the
biceps brachii, for example), with specified functions in locomotion. We have few
problems with accepting that the deer had been a browser, with a ruminant digestive
system and various other anatomical parts characteristic of the species. We would go
even further. We'd state the ancient deer was an adult when it died, because the
epiphyses at either end of the bone were fully fused. We would probably also accept
that the bone and the entire deer's body grew from a fertilized ovum, with some of its
cells diversified into specialized bone tissues. Given a modern comparative set of male
and female deer femora, we might even conclude that the fossil bone probably came
from a male. We might also accept that this ancient male underwent rutting seasons.
By this time we have inferred a great deal about the anatomy, physiology, embryology,
feeding, and reproduction of an animal we have never seen, all based on one bone.
This is what philosophers call an ampliative inference. Yet, we feel there is nothing
fantastic or assailable about these inferences. What we have done is use a very
complex set of analogy to infer the prehistoric existence of physical traits and behaviors
that we have not actually seen. We have made these broad inferences on the basis of
one object's similarity to others we know from our experience of the present-day world.
Analogy Can Be Abused
However, we might make assertions about the prehistoric deer about which we don't
feel so secure. For example, we could say that this male deer was not reproductively
successful, or that he was eating berries when he died. We do not feel secure about
these latter inferences because we cannot support clear one-to-one linkages between
the object we have and the traits we propose to infer. We think of too many possible
exceptions to the correlation between them. This tells us something about what makes
a strong analogy. Secure analogic inferences, such as those about the functional
anatomy and embryological development of the deer femur, are based on clear
functional links between the object studied and the traits and associations of
contemporary counterparts. In reality, we feel secure in those inferences because we do
not know of any cases in documented time in which femora appeared in the world
without such biological contexts and histories. Thus, the embryological development
and functional anatomy of femora are "necessary and sufficient" causes of their
existence.
Philosophers of science call this more warranted, functionally based kind of analogy
relational analogy. Relational analogies are thought to be stronger types of analogy than
those based simply on resemblances of form, or formal analogies. Relational analogies
involve arguments about functional relationships, such as structure or causation,
between the phenomena specified (Copi 1982; Hesse 1966; Wylie 1985).
An example of the contrast is useful here. A person with little background in archaeology
presents an archaeologist with a small, flat, roughly triangular stone, saying, "Look, I
found an arrowhead. What kind is it?" The archaeologist examines it and says, "This
isn't an arrowhead, it's just an unmodified stone." The layperson, a little incensed, says,
"How can you say that? Look, it's pointed and triangular, it's shaped just like an
arrowhead!" The archaeologist replies, "But, see, it doesn't have any flake scars on it,
the signs that someone shaped it into a point by chipping away stone from around the
edges. It's just a natural rock that happens to have the same shape as some
arrowheads made by people."
Depending on the quality of the archaeologist's explanation and on the people involved,
the discussion ends with the layperson accepting the archaeologist's "expert opinion," or
rejecting the archaeologist's diagnosis as another example of the mumbo jumbo
academics use to keep common folks out of their field. In this example, the layperson is
working on the level of a formal analogy, whereas the archaeologists is reasoning from
a kind of relational analogy, in which the form is less important than the distinctive
traces of human handiwork necessary to have generated a projectile point.
We can now turn to what many writers have condemned as misuse of "ethnographic
analogies" (Ascher 1961, Binford 1967, Binford 1987, Freeman 1968, Gould and
Watson 1982, Hill 1984). This misuse, like the less secure inferences from the deer
femur, involve uncritically assigning a set of traits from an ethnographically documented
case to a prehistoric one. The basis for the inferences may be formal resemblances of
some preserved prehistoric objects to materials from the ethnographic culture.
But similar problems arise in modern archaeological interpretations. Take, for example,
the U.S. Southwest, where "direct historic" links between modern Pueblos dwellers and
prehistoric peoples have been accepted since the turn of the century. Recent writings
have stressed the potential for assuming an unwarranted amount of ethnographic
similarity between ancient and recent pueblo dwellers (Cordell and Plog 1979). For
example, can we assume that, just because Anasazi pueblos had structures similar in
their physical details to modern kivas, the activities which went on in ancient kivas and
the ideas that structured their use were the same as those of recently documented
pueblo peoples? In fact, this assumption has often been made in archaeological
analyses, with little attention to developing arguments for why this should have been so.
Some recent reconstructions of Plio-Pleistocene hominid adaptations show similar
problems in overly ampliative applications inferences from ethnographic analogy. Some
writings inferred from the association of stone tools and bones at loci (sites) in the
landscapes that these were "home bases" of the sort created by modern hunter-
gatherers, and that they reflect other behaviors that go on at base camps, such as food
sharing (Isaac 1971). Binford (Binford 1981), Hill (Hill 1984), and others have criticized
these models of early hominid life as unwarranted inferences from the preserved
materials.
Reasoning about preserved archaeological materials is thus fraught with difficulties that
involve the use of analogic reasoning. Especially because human behavior is so
mutable, it is difficult to construct arguments about the meaning of archaeological
materials that are entirely relational, at least at the level we want to be working at when
studying the past. We may be able, as we'll see in later chapters, to claim a carnivore
gnaw mark on a bone is a tooth mark with a very high level of confidence; we may even
be able to tell which species of carnivore probably made it, for example, a hyena. But
what does this fact in itself tell us about the context in which the tooth mark was made?
Did the hyena hunt the animal whose bone has the mark? Did it scavenge a body
segment with the bone in it from another carnivore's kill? From a human kill? If the latter,
were the people still in the area or had they left? How common was that kind of
behavior?

CHAPTER THREE
Part Two
Causal Agent, Effector, Actor, Behavioral and Ecological Context
The hyena example contains several distinctions worth keeping in mind through most of
this book, and to which we will return at the end. We can conceive of the contexts and
processes that generate zooarchaeological assemblages as a nested set, ranging from
the most immediate and specific conditions of force and chemistry that produce an
effect on animals remains up to the most general types of context in which these
interactions occur (see Figure 3.1).
Figure 3.1 The nested relations of a trace to its causative agencies and contexts (after Gifford-Gonzalez
1991:Figure 2, redrawn with permission of Academic Press, Inc.).
At the innermost level, we have the actual trace, something we can study as empirical
evidence. Closest to it is its causal process, in this case, the action of the tooth of a
hyena pressing into a section of bone, which fails to withstand the pressure, gives way
and produces a mark, evidence of that interaction of materials and force.
Second, we have what we could call the effector of the damage, the hyena tooth, which
can be identified by analogy with modern hyena teeth and the marks they leave on
bones. Third, we have the actor creating the evidence though the immediate causal
process, the hyena itself.
Fourth and fifth, we have the context in which the agent was acting to produce the
evidence, which can be variously defined, depending on the range of one's focus. In the
most immediate sense, there is the behavioral context in which the evidence was
produced. In the hyena's case, it could be either "predation" or "scavenging." At this
analytic level, human "hunting," "storage," or "herd management" would be behavioral
contexts. Beyond this is the social and ecological context, referring to the type of
ecosystem and web of social relations in which the actors lived.
Figure 3.1 reflects a point of view that acknowledges what biologist Ernst Mayr (Mayr
1982, Mayr 1988) has called a hierarchical view of biological systems, in which
explanations or descriptions that work for the organization of life processes at a lower
level (e. g. biochemical reactions) cannot account for the entire operation of the system
at a higher level.
Much of what interests us as archaeologists is at the analytic level of behavior, social
systems, and regional ecology. However, what we have to work with, in terms of the
immediate evidence at hand, are the products of actors and effectors. The challenge is
linking our more sophisticated ability to read causal processes, effectors, and actions to
higher-order questions about these wider contexts (Gifford-Gonzalez 1991).
To pursue why this a challenge, let us return to the hyena example. Establishing a
strong relational analogy that implicates a causal process, effectors, and actor is
relatively straightforward, but it could involve eliminating other possible agents, such as
a hyena mandible in the hands of a hominid. The problem comes when we try to use
the hyena tooth mark to infer whether the animal was hunting or scavenging when it
gnawed the bone we are analyzing.
Two zooarchaeological debates over cut marks on bones illustrate some of the
problems involved in moving from one level to another in reasoning by analogy. First,
Pat Shipman and her associates (Shipman 1981a, Shipman and Rose 1983c)
described a number of detailed morphological criteria under the Scanning Electron
Microscope which they said distinguished cut marks made with stone tools from
scratches from sharp teeth of carnivores. A few years later, paleontologists working with
bone assemblages from epochs pre-dating hominids described similar marks on their
bones (Behrensmeyer, et al. 1986, 1989, Fiorillo 1987a, 1984, 1987b, 1989). Fiorillo
ascertained through experiment that these "pseudo-cut marks" could be created on
bones trampled against a sandy substrate by hoofed animals, as would be likely around
a waterhole.
In fact, the causal process producing pseudo-cut marks and real cut marks was the
same: a sharp, angular edge of a stone. However, the actual effectors and actors were
different, and therefore inferences from the evidence about them were ambiguous.
Investigators then undertook a new generation of research to distinguish these two
traces on the basis of other criteria, including substrate type (does is contain angular
materials that can cause pseudo-cuts?) , placement of marks on the bone (are they in
anatomically "logical" zones for butchery, or more or less randomly located?), and so
forth.
The contextual levels of are the most difficult in which to establish unambiguous
relational analogies, because these are in fact complex situations, where webs of
different agents and processes operate. Nonetheless, even ecosystems have been
shown to have regular relationships between different processes, as for example,
between the amount of rainfall and standing biomass (Coe and Phillipson 1976). Many
of the components in these regularities are variable over space and time. Therefore
they are not well suited to simple, cause-effect descriptions that work so well at the level
of causal process, effector, and actor. This presents inferential problems to those of us
who want to learn more about behavioral, social, and ecological context. because clear-
cut relational analogies are the power they bring to analogical inference are rare at this
level of organization. The outcome of processes is more likely to vary, and to be best
described by probabilistic, rather than "if a, then b" statements. At these levels, the
process of analogical reasoning is not impossible, it is simply more complex and
probably has different standards for evaluating its soundness. We will revisit this issue
again later in this book.
For now, I want to emphasize is that it would be a very great error to say we should
avoid using analogies taken from modern ethnographic (or ecologic) cases. In truth, it is
hard to imagine how we can say anything about prehistoric materials without using
some kinds of analogies, in naming, describing function, or defining context. Being more
critical in using analogy is a twofold task. First, we should strive to use analogies that
are less formal than they are relational. This in turn means investigating systematic,
functional links between types of archaeological evidence and their causes and agents
in the present. Second, we can explore the qualitative differences between these
analogies and those analogical arguments we make in more complex, higher-level
systems here called context.

CHAPTER THREE
Part Three
Refining Analogy by Research with Modern Cases
The only circumstances in which we can explore the functional and causal relationships
necessary to relational analogies are present-day situations. This is because it is only in
the observable present that we can document the link between a process and its
material product, a cause and its effect. I should add that I include historically and
ethnographically documented relationships in "the present." By establishing these links,
we can return to archaeological materials with a sense that the analogies we use to
interpret them are secure.
The basic relationships in the process of studying modern analogues are outlined in
Figure 3.2. In the present-day world, we observe an archaeological object. On the basis
of pre-existing assumptions and arguments that enable us to be archaeologists, we
accept that it has come down to us from some past period. We don't fully understand
some feature of this object. We think that it if we understood this feature better, we
would gain valuable information about the object's role in the past. On the basis of our
knowledge of the present-day world, we select a modern counterpart for the
archaeological object. We verify through comparison that the two objects of study are
similar in features that we believe are relevant to our inquiry, at the same time noting
respective features which differ between the two objects (Copi 1982, Wylie 1985),
keeping in mind that analogies will always have points of difference, otherwise they
would be homologies (Young 1989).
Again on the basis of our knowledge of the contemporary world, we chose a range of
processes which might possibly have created the feature in question. Under controlled
conditions, which can range from experiment to carefully monitored natural situations,
we establish which of these processes created the feature. Sometimes we discover that
some other process is involved. With our new understanding of the cause of the feature
under investigation in the present, and because of the resemblance of the experimental
set to the archaeological set, we infer that a similar process produced the
archaeological feature at some time in the past.
Figure 3.2. A model of analogical reasoning in historical science. Shaded area indicates opportunity for
contemporary observations. The "inferred similarity" on the left may also be viewed as one or more
uniformitarian assumptions.
Inferring the cause of the prehistoric feature requires that we assume that similar
processes have produced similar traces over long stretches of time. This is the core of
methodological uniformitarianism. The practice of gaining more information about
potentially uniformitarian processes and their products through experiment and
observation of contemporary analogues is actualism.
Actualism in Zooarchaeology
Archaeologists and paleontologists have done actualistic research on bones and other
organic remains more and more frequently in the last twenty years, and much of the
balance of this book will illustrate the ways in which actualistic investigations and the
analyses of archaeological faunas work together to raise and resolve analytic issues.
Research in contemporary settings has included controlled experiments and
observations of relevant natural situations. Studies of bones in "natural" settings can be
subdivided into ethnoarchaeology and taphonomy, depending on whether or not
humans were among the agents handling and modifying bones. The increasing
frequency of these studies reflects the widespread recognition of the relationships of
analogical reasoning, uniformitarian assumptions, and actualistic research in the
historical sciences. Contemporary studies allow us to define what features of an object
are likely to endure over time, and how these features are functionally related to its life
context.
Much of this research has yielded valuable results for inferring past events through
analogical reasoning. Detailed experimental and naturalistic observations of bone
collectors and bone modifiers have defined causes of modifications and patterns of
preservation also found on prehistoric bones, such as cut marks, tooth marks,
weathering, and spiral fractures, which will be discussed in Section Four This research
has enriched our understanding of ecosystems and the role of bones in them. Greater
knowledge of structural and functional relations of bones as elements in ecosystems is
just beginning to allow us to extend the ranges of our relational analogies beyond the
agents of modification to their ecological circumstances.
Taking Product-Focused Approach
This text incorporates another perspective in analyzing zooarchaeological evidence that
might be called a product-focused approach (Gifford-Gonzalez 1998). Faunal
assemblages, both paleontological and archaeological, have complex histories, in which
multiple causal processes have acted. Some processes may leave traces of their
operation; some may obscure or remove the marks of others. For example, the flaking
of bone surfaces due to weathering can remove shallow cut marks on the original bone
surface. This sequential process of modification is sometimes called the "taphonomic
overprint"(Lawrence 1968). Some earlier taphonomic writings, emphasized the loss of
information about life context through various postmortem processes acted on organic
remains. This point of view is correct to the degree that it describes the progressive
postmortem divergence of animal remains from their original state as parts of living
organisms in ecosystems.
However, this perspective holds out a chimerical goal in taphonomic and
zooarchaeological analysis: the "unbiasing" of an archaeofaunal sample back to its
original context in a living system. As Lyman (Lyman 1994) and I (Gifford 1981) have
previously noted, this is an unrealistic aim. Lyman (Lyman 1987b) has argued, as have I
(Gifford-Gonzalez 1991), that, rather than viewing the role of taphonomic analysis as
"stripping away the overprint," it is more productive to view these effects as evidence
added to specimens by postmortem processes. In fact, taphonomic evidence is a form
of trace fossil of the action of other organisms and nonbiological processes upon
organic remains. In a different context, bone isotope analyst Andrew Sillen (Sillen, et al.
1989) put it this way: "diagenesis suffers from a bad name; we tend to see it as the mist
on the window rather than part of the view."
A product-focused approach begins with the viewpoint that each specimen has its own
individual history, some of which can be discerned from its form, composition, and
modifications (Figure 3.3). These include attributes functionally related to its
development and role during the life of an animal and others resulting from the effects of
various processes upon it after death. Analysis therefore always begins with recording
data from individual bones. Despite starting with individual bones, understanding the
dominant processes that created a bone assemblage demands that data taken from
individual bones be read as an aggregate pattern. Just as there is no typical site, there
is no typical bone.
Frequently reiterated human actions affecting bone, a customary way of cutting up a
sheep, for example, as well as repeatedly occurring non-human processes, produce
patterning in the aggregate evidence. The patterning is thus the cumulative reflection of
redundant incidents of human behavior or natural agencies that produced a certain type
of damage or element deletion. These patterns reflect some of the most common
processes affecting animal remains over the time the site assemblage formed. Not
every small event that occurred may be reflected in the evidence, but rather the
dominant impactful processes. Our task as zooarchaeologists is to try to understand
those processes which created the patterning. Like other archaeologists, faunal
analysts must also consider how to cope with the possibility that some processes
affecting archaeofaunal samples left no direct or distinctive traces. This is a situation
where what Lyman (1987) has called a "forensic" approach is most useful.
"Signatures," Equifinality, and a Forensic Approach
Zooarchaeologists have sought distinctive "signatures," that is, marks which could have
been made by only one agent (e.g. a carnivore, a hominid with a stone tool) through
actualistic research. Based on uniformitarian assumptions, experimental observations of
cause and effect in contemporary settings permit a finer-grained and more plausible
reading of certain faunal evidence. However, actualistic research has sometimes shown
that different agents produce very similar effects. The example of pseudo-cut marks
noted earlier is one such case. Lyman (1987) describes this as a problem of equifinality
("same end," or same final outcome). We thus face the problem of how we determine
which causal actor was responsible for the evidence.
To handle this problem, it is impossible to argue from the ambiguous trace or line of
evidence itself. Rather, it is necessary to use several independent lines of evidence that
point to the same actor as the most probable cause of the trace. by a case of equifinality
at the level of causal process and effector by mobilizing two related by independent
lines of evidence: sedimentary context and anatomical placement of the marks.
This approach has been termed "forensic" by Lyman (1987), emphasizing parallels to
investigations in which multiple lines of evidence, each independent of the other, are
brought to bear. The more lines of circumstantial evidence point to the same causal
process, effector, and actor, the more likely is that these were in fact responsible for the
outcomes being investigated. This approach to reducing ambiguity about causal agency
or circumstances also been called "contextual analysis," and applying independent
uniformitarian "frames of reference" Binford (1989). The crucial point is that the lines of
evidence should not depend on each other in a functional sense. Thus, the problem of
equifinality in zooarchaeology means that we may not be positive about the causal
context of given line of evidence, but actualistic research allows us to limit the range of
causal possibilities and a forensic approach allows us to feel more strongly justified in
specifying one of these possibilities, when several independent lines of evidence point
to the same causal process. The courtroom standard of "beyond a reasonable doubt"
may not be a realistic criterion for zooarchaeologists, but something close to this may
be achieved.
In this context that faunal analysts may want to look beyond their own dataset to other
lines of archaeological evidence that might point to similar causal contexts. This is
especially the case when seeking to make inferences at the level of past behavioral,
social, and ecological contexts, where analogical reasoning from any one line of
evidence may be more probabilistic in nature (see Gifford-Gonzalez 1991).
Key Types of Evidence in Zooarchaeology
Zooarchaeologists handle bones one by one, but they almost immediately abstract
information from them to make aggregate patterns of data noted above. Two
fundamental categories of information are used in zooarchaeological analysis, upon
which nearly all other abstractions of data and inferences are built.
One of these fundamental types of information is element frequencies, or counts of the
numbers of anatomical elements, relative to one another, in a faunal sample. This is the
knowledge upon which statements such as "Caribou were the most common species in
the assemblage," or "Aardvarks are rare in Late Stone Age sites," are based.
Generalizations about species abundances are thus derived from element frequency
data. Moreover, reconstructing the age structures of animals present in the sample and
thereby making inferences about hunting, herd management, or seasonality depends on
summary counts of age-diagnostic bones and teeth and reckoning of their relative
frequencies. Size variations of a species over time, often linked to climatic fluctuations
(Klein 1986), depends on counts of elements with different metrical attributes. Studying
butchery and selective body segment transport by people handling animals, likewise
depends on frequencies of elements from different vertebrate body segments, also the
data on which assessing the amount of in-place destruction of more delicate bones is
based. Thus, taxonomic abundances, age and sex structures, and body part
representation, are all different permutations of basic element frequency data.
The second major category of data zooarchaeologists use in aggregate form may be
called added modifications, or surface modifications (Fisher 1995). These can be cuts,
chops, burning tooth marks, weathering, or any impacts of humans, other biological
agents, geological, or mechanical forces on anatomical elements. Sometimes,
researchers cited in this text refer to these as "traces" or "signatures" of various agents.
Our ability to recognize and derive useful information from such modifications has
expanded tremendously since the 1970s, and much of this book will be devoted to
summarizing what is known about the causes of various bone modifications, as well as
the unresolved problems in making plausible inferences from them.
This chapter has underlined that it is the uniform properties of organic remains as they
respond to destructive or modifying forces enable zooarchaeological research. This
uniformity stems from consistencies in the development and construction of anatomical
elements in animals during their lives. In other words, it is the consistency with which
bone, teeth, shell, and other organic materials react to the demands of life that create
their uniform responses to stress after death. The next section deals with those
uniformitarian properties of living vertebrate remains.

References Cited
Acsadi, G. and J. Nemeskeri
1970 History of Human Lifespan and Mortality. Academi Kiado, Budapest.
Aiello, L. C. and P. Wheeler
1995 The expensive-tissue hypothesis: the brain and the digestive system in human
and primate evolution. Current Anthropology 36(2):199-221.
Andrews, P.
1990 Owls, Caves, and Fossils. Predation, Preservation, and Accumulation of Small
Mammal Bones in Caves, with an Analysis of the Pleistocene Cave Faunas from
Westbury-sub-Mendip, Somerset, UK. University of Chicago Press, Chicago.
Armstrong, W. G., L. B. Halstead, F. B. Reed and L. Wood
1983 Fossil proteins in vertebrate calcified tissues. Phil. Trans. R. Society of London
301:301-343.
Badgley, C. E. and A. K. Behrensmeyer
1980 Paleoecology of Middle Siwalik sediments and faunas, north Pakistan.
Palaeogeography, Palaeoclimatology, Palaeoecology 30:133-155.
Bakker, R. T.
1972 Anatomical and ecological evidence of endothermy in dinosaurs. Nature 238:81-
85.
Bar-Yosef, O. and A. Khazanov (editors)
1992 Pastoralism in the Levant: Archaeological Materials in Anthropological
Perspectives. Prehistory Press, Madison, WI.
Barker, D. J. P., C. Osmond, S. J. Simmond and G. A. Wield
1993 The relation of small head circumference and thinness at birth to death from
cardiovascular disease in adult life. British Medical Journal 306(6875):422-426.
Baron, R.
1993 1. Anatomy and ultrastructure of bones. In Primer on the Metabolic Bones
Diseases and Disorders of Mineral Metabolism., edited by M. J. Favus, pp. 3-9. 2nd ed.
Raven Press, New York.
Bartram, L. E.
1993a An ethnoarchaeological analysis of Kua San (Botswana) bone food refuse.
Ph.D., University of Wisconsin &endash; Madison.
1993b Perspectives on skeletal part profiles and utility curves from eastern Kalahari
ethnoarchaeology. In Bones to Behavior: Ethnoarchaeological and Experimental
Contributions to the Interpretation of Faunal Remains, edited by J. Hudson, pp. 115-
137. vol. Occasional Paper No. 21. Center for Archaeological Investigations, Southern
Illinois University at Carbondale, Carbondale.
Bartram, L. E., E. M. Kroll and H. T. Bunn
1991 Variability in camp structure and bone food refuse patterning at Kua San camps.
In The Interpretation of Archaeological Spatial Patterning, edited by E. M. Kroll and T. D.
Price, pp. 77-148. Plenum, New York.
Behrensmeyer, A. K.
1975 The taphonomy and paleoecology of Plio-Pleistocene vertebrate assemblages
east of Lake Rudolph, Kenya. Bulletin of the Museum of Comparative Zoology 146:473-
578.
1978 Taphonomic and ecologic information from bone weathering. Paleobiology 4:150-
162.
1982 Time resolution in fluvial vertebrate assemblages. Paleobiology 8:211-227.
Behrensmeyer, A. K., K. D. Gordon and G. T. Yanagi
1986 Trampling as a cause of bone surface damage and pseudo-cutmarks. Nature
319:768-771.
1989 Nonhuman bone modification to Miocene fossils from Pakistan. In Bone
Modification, edited by R. Bonnichsen and M. Sorg, pp. 99-120. University of Maine
Center for the Study of the First Americans, Orono, ME.
Behrensmeyer, A. K. and C. B. Hanson
1975 Bone transport experiments on the East Fork River, Wyoming. unpub. ms .
Berger, L. R. and R. J. Clarke
1995 Eagle involvement in accumulation of the Taung child fauna. Journal of Human
Evolution 29:275-299.
Binford, L. R.
1962 Archaeology as Anthropology. American Antiquity 28(2):217-225.
1977a For Theory Building in Archaeology: Essays on Faunal Remains, Aquatic
Resources, Spatial Analysis, and Systemic Modeling. Academic Press, New York.
1977b Olorgesailie deserves more than the usual book review. Journal of
Anthropological Research 33(4):493-502.
1978a Dimensional analysis of behavior and site structure: Learning from an Eskimo
hunting stand. American Antiquity 43:330-361.
1978b Nunamiut Ethnoarchaeology. Academic Press, New York.
1978c Organization and formation processes: Looking at curated technologies. Journal
of Anthropological Research 35(3):255-273.
1980 Willow smoke and dogs' tails: hunter-gatherer settlement systems and
archaeology. American Antiquity 45:4-20.
1981 Bones: Ancient Men and Modern Myths. Academic Press, New York.
1982 The archaeology of place. Journal of Anthropological Archaeology 1(1):5-31.
1983 In Pursuit of the Past: Decoding the Archaeological Record. Thames and Hudson,
New York.
1984a Butchering, sharing, and the archaeological record. Journal of Anthropological
Archaeology 3:235-257.
1984b Faunal Remains from Klasies River Mouth. Academic Press, New York.
1987 Researching ambiguity: frames of reference and site structure. In Method and
Theory for Area Research. An Ethnoarchaeological Approach, edited by S. Kent, pp.
448-512. Columbia University Press, New York.
1988 Fact and fiction about the Zinjanthropus floor: data, arguments and interpretations.
Current Anthropology 29(1):123-135.
1991 When the going gets tough, the tough get going: Nunamiut local groups, camping
patterns, and economic organization. In Ethnoarchaeological Approaches to Mobile
Campsites: Hunter&endash;Gatherer and Pastoralist Case Studies, edited by C. S.
Gamble and W. A. Boismier, pp. 25-137. Ethnoarchaeological Series 1. International
Monographs in Prehistory, Ann Arbor.
Binford, L. R. and J. Bertram
1977 Bone frequencies -- and attritional processes. In For Theory Building in
Archaeology, edited by L. R. Binford, pp. 77-153. Academic Press, New York.
Binford, L. R., M. G. L. Mills and N. M. Stone
1988 Hyena scavenging behavior and its implications for the interpretation of faunal
assemblages from FLK 22 (the Zinj floor) at Olduvai Gorge. Journal of Anthropological
Archaeology 7(2):99-135.
Blumenschine, R. J.
1986a Carcass consumption sequences and the archaeological distinction of
scavenging and hunting. Journal of Human Evolution 15:639-659.
1986b Early Hominid Scavenging Opportunities: Implications of Carcass Availability.
283. British Archaeological Reports, International Series, Oxford.
1987 Characteristics of an early hominid scavenging niche. Current Anthropology
28(4):383-407.
1988a Percussion marks on bone surfaces as a new diagnostic of hominid behavior.
Nature 333(6175):763-765.
1988b Reinstating an early hominid scavenging niche: A reply to Potts. Current
Anthropology 29(3):483-486.
1989 A landscape taphonomic model of the scale of prehistoric scavenging
opportunities. Journal of Human Evolution 18(4):345-371.
Blumenschine, R. J., J. A. Cavallo and S. D. Capaldo
1994 Competition for carcasses and early hominid behavioral ecology: A case study and
conceptual framework. Journal of Human Evolution 27(1-3):197-213.
Blumenschine, R. J. and C. W. Marean
1993 A carnivore's view of archaeological bone assemblages. In From Bones to
Behavior: Ethnoarchaeological and Experimental Contributions to the Interpretation of
Faunal Remains, edited by J. Hudson, pp. 273-300. Board of Trustees, Southern Illinois
University.
Blumenschine, R. J. and M. M. Selvaggio
1991 On the marks of marrow bone processing by hammerstones and hyenas: their
anatomical patterning and archaeological implications. In Cultural Beginnings.
Approaches to Understanding Early Hominid Life-Ways in the African Savanna, edited
by J. D. Clark, pp. 17-32. Union Internationale des Sciences Préhistoriques et
Protohistoriques 111 Congress Mainz, 31 August, - 5 September, 1987. vol. 19. Dr.
Rudolf Habelt GMBH, Bonn.
Boaz, N. T. and A. K. Behrensmeyer
1976 Hominid taphonomy: transport of human skeletal parts in an artificial fluvia.
American Journal of Physical Anthropology 45:53-60.
Bocherens, H., M. Fizet and A. Mariotti
1990 Mise en évidence du régime alimentaire végétarien de l'ours des cavernes (Ursus
spelaeus ) par la biogéochemie isotopique (13C, 15N) du collagène des Vertébrés
fossiles. Comptes rendus de l'Academie des Sciences de Paris, Série II 311:1279-1284.
Boessneck, J.
1969 Osteological differences between sheep (Ovis aries Linne) and goat (Capra hi. In
Science in Archaeology, edited by D. R. Brothwell and E. S. Higgs, pp. 331-358.
Thames & Hudson, London.
Bökönyi, S.
1984 Animal Husbandry and Hunting in Tác-Gorsium. The Vertebrate Fauna of a
Roman Town in Pannonia. Akadémiai Kiadó (Hungarian Academy of Sciences),
Budapest.
Bonnichsen, R.
1973 Some operational aspects of human and animal bone alteration. In Mammalian
Osteoarchaeology: North America, edited by B. M. Gilbert, pp. 9-24. Missouri
Archaeological Society, Columbia.
1979 Pleistocene Bone Technology in the Beringian Refugium. Archaeological Survey of
Canada Papers 89.
Bonnichsen, R. and A. L. Bryan
1978 Critical arguments for Pleistocene artifacts from the Old Crow Basin, Yukon: a
preliminary statement. Archaeological Researches International.
Bonnichsen, R. and R. T. Will
1980 Cultural modification of bone: the experimental approach in faunal analysis. In
Mammalian Osteoarchaeology, edited by B. M. Gilbert, pp. 7-30. Privately Published.
Borrero, L. A.
1990 Taphonomy of guanaco bones in Tierra del Fuego [Argentina]. Quaternary
Research 34(3):361-371.
Bourdieu, P.
1977 Structures and the habitus.
Bowen, J.
1975 Probate inventories: an evaluation from the perspective of zooarchaeology and
agricultural history at Mott Farm. Historical Archaeology 9:11-25.
1988 Seasonality: an agricultual construct. In Documentary Archaeology in the New
World, edited by M. C. Beaudry, pp. 161-216. Cambridge University Press, Cambridge.
1990 A Study of Seasonality and Subsistence: Eighteenth-Century Suffield,
Connecticut. Ph.D. Dissertation, Brown University.
Bower, J. R. F. and A. Gogan-Porter
1981 Prehistoric Cultures of the Serengeti National Park. Iowa State University Papers
in Anthropology No. 3.
Boyd, D. W. and N. D. Newell
1972 Taphonomy and diagenesis of a Permian fossil assemblage from Wyoming.
Journal of Paleontology 46:1-14.
Brain, C. K.
1965 Bone weathering and the problem of bone pseudo-tools. South African Journal of
Science 63:97-99.
1967 Hottentot food remains and their bearing on the interpretation of fossil bone
assemblages. Scientific Papers of the the Namib Desert Research Station 32:1-7.
1969 The contribution of Namib desert Hottentots to an understanding of
australopithecine bone accumulations. Scientific Papers of the Namib Desert Research
Station 39:13-22.
1974 Some suggested procedures in the analysis of bone accumulations from Southern
African Quaternary sites. Annals of the Transvaal Museum 29:1-8.
1976 Some principles in the interpretations of bone accumulations associated with man.
In Human Origins: Louis Leakey and the East African Evidence, edited by G. L. Isaac
and B. R. McCown, pp. 97-116. Benjamin, Menlo Park, CA.
1980 Some criteria for the recognition of bone-collecting agencies. In Fossils in the
Making: Vertebrate Taphonomy and Paleoecology, edited by A. K. Behrensmeyer and A.
Hill, pp. 107-130. University of Chicago Press, Chicago.
1981 The Hunters or the Hunted? An Introduction to South African Cave Taphonomy.
University of Chicago Press, Chicago.
Breuil, H.
1938 The use of bone implements in the Old Paleolithic period. Antiquity 12(56-67).
1939 Bone and antler industry of the Choukoutien Sinanthropus site. Palaeontologia
Sinica n.s. D(6).
Bromage, T. G.
1984 Interpretation of Scanning Electron Microscope Images of abraded forming bone
surfaces. American Journal of Physical Anthropology 64:161-178.
Broughton, J. M.
1994a Declines in mammalian foraging efficiency during the Late Holocene, San
Francisco Bay, California. Jounal of Anthropological Archaeology 13:371-401.
1994b Late Holocene resource intensification in the Sacramento Valley, California: the
vertebrate evidence. Journal of Archaeological Science 21(4):501-514.
Broughton, J. M.
1994c Resource depression and intensification during the late Holocene, San Francisco
Bay: Evidence from the Emeryville Shellmound Vertebrate Fauna, University of
Washington.
1997 Widening diet breadth, declining foraging efficiency, and prehistoric harvest
pressure: ichthyofaunal evidence from the Emeryville Shellmound, California. Antiquity
71(274):845-862.
Broughton, J. M. and D. K. Grayson
1993 Diet breadth, adaptive change, and the White Mountains faunas. Journal of
Archaeological Science 20(3):331-336.
Buikstra, J. E.
1981 Mortuary practices, paleodemography, and paleopathology: A case study from the
Koster Site (Illinois). In The Archeology of Death, edited by R. Chapman, I. Kinnes and
K. Randsborg, pp. 123-132. Cambridge University Press, Cambridge.
Bunn, H. T.
1981 Archaeological evidence for meat-eating by Plio-Pleistocene hominids from Koobi
Fora and Olduvai Gorge. Nature 291:574-577.
1982 Animal bones and archeological inference. Science 215:494-495.
1983 Comparative analysis of modern bone assemblages from a San hunter-gatherer
camp. In Animals and Archaeology I. Hunters and Their Prey 163, edited by J. Clutton-
Brock and C. Griggs, pp. 143-148. BAR International, Oxford.
1989 Diagnosing Plio-Pleistocene hominid activity with bone fracture evidence. In Bone
Modification, edited by R. Bonnichsen and M. Sorg, pp. 299-315. Center for the Study of
Early Man, Orono, Maine.
1993 Bone assemblages at base camps: a further consideration of carcass transport
and bone destruction by the Hadza. In From Bones to Behavior: Ethnoarchaeological
and Experimental Contributions to the Interpretation of Faunal Remains, edited by J.
Hudson, pp. 156-168. Occasional Paper No. 21. Center for Archaeological
Investigations, SIU at Carbondale, Carbondale.
Bunn, H. T., L. E. Bartram, Jr. and E. M. Kroll
1988 Variability in bone assemblage formation from Hadza hunting, scavenging, and
carcass processing. Journal of Anthropological Archaeology 7(4):412-457.
Bunn, H. T. and J. A. Ezzo
1993 Hunting and scavenging by Plio-Pleistocene hominids: nutritional constraints,
archaeological patterns, and behavioural implications. Journal of Archaeological
Science 20:365-398.
Bunn, H. T., J. W. K. Harris, G. Isaac, Z. Kaufulu, E. Kroll, K. Schick, N. Toth and A. K.
Behrensmeyer
1980 FxJj50: an early Pleistocene site in northern Kenya. World Archaeology 12:109-
136.
Bunn, H. T. and E. M. Kroll
1986 Systematic butchery by Plio-Pleistocene hominids at Olduvai Gorge, Tanzania.
Current Anthropology 27:431-452.
1988 Reply to 'Fact and fiction about the Zinjanthropus floor: data, arguments, and
interpretation', by Lewis Binford. Current Anthropology 29(1):135-149.
Butte, N. F., C. Garza, E. O. Smith and B. L. Nichols
1986 Human milk intake and growth in exclusively breast-fed infants. Journal of
Pediatrics 104(2):187-95.
Callow, P. M. and J. M. Cornford (editors)
1986 La Cotte de St Brelade 1961-1978: Excavations by C B M McBurney. Geo.
Campana, D. V. and P. J. Crabtree
1987 ANIMALS: a C language computer program for the analysis of faunal remains and
its use in the study of early Iron Age fauna from Dun Ailinne. ArchaeoZoologia 1(1):57-
67.
Carr, C. J.
1977 Pastoralism in Crisis. The Dasanetch and Their Ethiopian Lands.
Casteel, R. W.
1974 On the Remains of Fish Scales from Archaeological Sites39(4, pt. 1):557-581.
Catlin, G.
1959 Episodes of Life among the Indians, and Last Rambles. University of Oklahoma,
Norman.
Claassen, C.
1998 Shells. Cambridge Manuals in Archaeology. Cambridge University Press,
Cambridge.
Clark, J. and K. K. Kietzke
1967 Paleoecology of the lower Nodular zone, Brule Formations, in the Big Badland.
Fieldiana: Geology Memoirs 5:111-129.
Clark, J. D. and J. W. K. Harris
1985 Fire and its roles in early hominid lifeways. African Archaeological Review 3:3-27.
Clark, J. G. D.
1954 Excavations at Star Carr. Cambridge University Press, Cambridge.
Conkey, M. W. and J. D. Spector
1984 Archaeology and the study of gender. Advances in Archaeological Method and
Theory 7:1-38.
Coy, J.
1981 Animal husbandry and faunal exploitation in Hampshire. In The Archaeology of
Hampshire, edited by R. T. S. Hall and S. J. Shennan, pp. 95-103. vol. 1. Hampshire
Field Club and Archaeological Society, Hampshire.
Crabtree, P. J.
1990 Zooarchaeology and complex societies: some uses of faunal analysis for the study
of trade, social status, and ethnicity. Archaeological Method and Theory 2.:155-205.
1996 Production and consumption in an early complex society: animal use in Middle
Saxon East Anglia. World Archaeology 28(1):58-75.
Crader, D. C.
1983 Recent single-carcass bone scatters and the problem of "butchery" sites in the
archaeological record. In Animals and Archaeology I. Hunters and their Prey, edited by
J. C.-B. a. C. Griggs, pp. 107-141. BAR International Series 163, Oxford.
1984 The zooarcheology of the Storehouse and the Dry Well at Monticello. American
Antiquity 49(3):542-558.
1989 Faunal remains from slave quarter sites at Monticello, Charlottesville, Virginia.
ArchaeoZoologia 3(1-2):229-236.
Cribb, R.
1985 The analysis of ancient herding systems: an application of computer simulation in
faunal studies. In Beyond Domestication in Prehistoric Europe, edited by G. Barker and
C. Gamble, pp. 75-106. Academic Press, London.
Currey, J.
1984 The Mechanical Adaptations of Bones. Princeton University Press, Princeton.
Dahl, G. and A. Hjort
1976 Having Herds. Pastoral Herd Growth and Household Economy. Stockholm Studies
in Social Anthropology No. 2. Department of Anthropology, University of Stockholm,
Stockholm, Sweden.
Dart, R. A.
1949 The predatory implemental technique of Australopithecus. American Journal of
Physical Anthropology 7:1-17.
Dart, R. A.
1957 The Osteodontokeratic Culture of Australopithecusm prometheus. Memoir of the
Transvaal Museum Nº 10.
Dart, R. A.
1959 Further light on australopithecine humeral and femoral weapons. American
Journal of Physical Anthropology 17(1):134-143.
Davis, S. J. M.
1987 The Archaeology of Animals. Yale University Press, New Haven, Conn.
Davis, T. A., H. V. Nguyen, M. L. Fiorotto and P. J. Reeds
1993 Primate and nonprimate milks have different amino acid patterns. Federation of
American Societies for Experimental Biology Journal 7((3-4)):A158.
Deagan, K.
1973 Mestizaje in Colonial St. Augustine. Ethnohistory 20(53-65).
1974 Sex, Status, and Role in the Mestizaje of Spanish Colonial Florida, University of
Florida.
1983 Spanish St. Augustine: The Archaeology of a Colonial and Creole Community 20.
Academic Press, New York.
1985 Spanish-Indian interactions in 16th Century Florida and Hispaniola. In Cultures in
Contact, edited by W. Fitzhugh. vol. 20. Smithsonian Institution Press, Washington, D.C.
Deetz, J.
1977 In Small Things Forgotten. Anchor Books, Garden City, New York.
1988 American Historical Archeology: Methods and Results. Science 239:362-367.
1996 In Small Things Forgotten. Revised and expanded edition. Anchor
Books/Doubleday, New York.
Deevy, E. S., Jr
1947 Life tables for natural populations of animals. Quarterly Review of Biology
22(4):283-314.
DeNiro, M. J.
1985 Postmortem preservation and alteration of in vivo bone collagen isotope ratio.
Nature 317:806-809.
Deniz, E. and S. Payne
1982 Eruption and wear in the mandibular dentition as a guide to ageing Turkish Angora
goats. Paper presented at the Ageing and Sexing Animal Bones from Archaeological
Sites, Oxford.
Dennell, R. W.
1974 Botanical evidence for prehistoric crop processing activities. Journal of
Archaeological Science. 1(3):275-284.
Densmore, F.
1918 Teton Souix Music. Bureau of American Ethnology Bulletin 61.
Dietler, M.
1990 Driven by drink: the role of drinking in the political economy and the case of Early
Iron Age France. Journal of Anthropological Archaeology 9:352-406.
Dodson, P.
1971 Sedimentology and Taphonomy of the Oldman Formation (Campanian), Dinosaur
Provincial Park, Alberta (Canada). Palaeogeography, Palaeoclimatology, Palaeoecology
10:21-74.
Dodson, P. and D. Wexler
1979 Taphonomic investigations of owl pellets. Paleobiology 5:275-284.
Doran, G. H., D. N. Dickel, W. E. Ballinger, O. F. Agee, P. J. Laipsis and W. W.
Hauswirth
1986 Anatomical, cellular and molecular analysis of 8000 yr old human brain tissue from
the Windover archaeological sites. Nature 323:803-806`.
Ducos, P.
1968 L' Origine des animaux domestiques en Palestine.
Duke, G. E., O. A. Evanson and A. A. Jegers
1976 Meal to pellet intervals in 14 species of captive raptors. Comparative Biochemistry
and Physiology 53A:1-6.
Duke, G. E., A. A. Jegers, G. Loff and O. A. Evanson
1975 Gastric digestion in some raptors. Comparative Biochemistry and Physiology
50A:649-656.
Emerson, A. M.
1990 Carcass product yields in Bison bison and hunter selection. Paper presented at
the International Council for Archaeozoology, Smithsonian Institution, Washington, D. C.
1993 The role of body part utility in small-scale hunting under two strategies of carcass
recovery. In From Bones to Behavior: Ethnoarchaeological and Experimental
Contributioings to the Interpretation of Faunal Remains, edited by j. Hudson, pp. 138-
155. Southern Illinois University, Carbondale, IL.
Emerson, M. C.
1988 Decorated clay tobacco pipes from the Chesapeake, University of California.
Emerson, W. K. and M. K. Jacobson
1976 American Museum of Natural History Guide to Shells. A. Knopf, New York.
Enloe, J. G.
1992 Le partage de la nourriture partir des témoins archéologiques: une application
ethnarchéologique. In Ethnoarchéologie: Justification, Problèmes, Limites, edited by A.
Gallay, F. Audouze and V. Roux, pp. 307-323. APDCA, Juan- les-Pins.
1993 Ethnoarchaeology of marrow cracking: implications for the recognition of
prehistoric subsistence organization. In Bones to Behavior: Ethnoarchaeological
Contributions to the Interpretation of Faunal Remains, edited by J. C. Hudson, pp. 82-
97. University of Southern Illinois Press, Carbondale.
Enloe, J. G. and F. David
1989 Le remontage des os par individus: le partage du renne chez les Magdaleniens.
Bulletin de la Société Préhistorique Française 86:275-281.
1992 Food sharing in the palaeolithic: carcass refitting at Pincevent. In B.A.R.
International Series, edited by L. Hofman and J. G. Enloe, pp. 296-315. vol. 578. British
Archaeological Reports, Oxford.
Enloe, J. G., F. David and T. S. Hare
1994 Patterns of faunal processing at section 27 of Pincevent: the use of spatial
analysis and ethnoarchaeological data in the interpretation of archaeological site
structure. Journal of Anthropological Archaeology 13:105-124.
Ewan, C. R.
1985 From Spaniard to Creole: The Archaeology of Hispanic American Cultural
Formation in Puerto Real, Haiti 20. University of Alabama Press, Tuscaloosa, AL.
Fairbanks, C. H. and M.-M. S. A.
1980 How did slaves live? In Early Man, pp. 2-6. vol. 2.
Fedigan, L. M.
1997 Changing views of female life histories. In The Evolving Female: a Life History
Perspective, edited by M. E. Morbeck, A. Galloway and A. L. Zihlman. vol. 15-26.
Princeton University Press, Princeton, N.J.
Fiorillo, A. R.
1984 An introduction to the identification of trample marks. Current Research in the
Pleistocene 1:47-48.
1987 Trample marks: caution from the Cretaceous. Current Research in the Pleistocene
4:73-74.
1989 An experimental study of trampling implications for the fossil record. In Bone
Modification, edited by R. Bonnichsen and M. Sorg. University of Maine Center for the
Study of Early Man, Orono.
Fisher, D. C.
1987 Mastodon procurement by Paleo-Indians in the Great Lakes region: hunting or
scavenging? In The Evolution of Human Hunting, edited by M. H. Nitecki and D. L.
Nitecki, pp. 309-421. Plenum Press, New York.
Fisher, J., John W.
1993 Foragers and farmers: material expressions of interaction at elephant processing
sites in the Ituri Farmer, Zaire. In Bones to Behavior: Ethnoarchaeological and
Experimental Contributions to the Interpretation of Faunal Remains, edited by J.
Hudson, pp. 247-262. Southern Illinois University Press, Carbondale.
1995 Bone surface modifications in zooarchaeology. Journal of Archaeological Method
and Theory 2(1):7-68.
Flannery, K.
1986 Guila Naquitz: Archaic Foraging and Early Agriculture in Oaxaca, Mexico. Studies
in archaeology. Academic Press, Orlando.
Flannery, K. V. (editor)
1976 The Early Mesoamerican Village. Academic Press, New York.
Fleay, D. H.
1932 The rare Dasyures (Native Cats). Victorian Naturalist 1932:63-69.
Forwood, M. R., I. Owan, Y. Takano and C. H. Turner
1996 Increased bone formation in rat tibiae after a single short period of dynamic
loading in vivo. American Journal of Physiology 270(3 PART 1):E419-E423.
Frison, G. C.
1970 The Glenrock Buffalo Jump, 488CO304.
1974 The Casper Site: A Hell Gap Bison Kill on the High Plains.
1978 Prehistoric Hunters of the High Plains. Academic Press, New York.
Frison, G. C. and L. Todd
1987 The Horner Site. Studies in Archaeology. Academic Press, Orlando.
Frison, G. C., M. Wilson and D. J. Wilson
1976 Fossil bison and artifacts from an early altithermal period arroyo trap in Wyoming.
American Antiquity 41(1):28-57.
Galdikas, B.
1978 Orangutan death and scavenging by pigs. Science 200:68-70.
Galloway, A.
1997 The cost of reproduction and the evolution of postmenopausal osteoporosis. In
The Evolving Female: A Life History Perspective, edited by M. E. Morbeck, A. Galloway
and A. Zihlman. Princeton University Press, Princeton NJ.
Galloway, A., W. P. and L. Snyder
1996 Human bone mineral densities and survival of bone elements: a contemporary
sample. In Forensic Taphonomy: The Postmortem Fate of Human Remains, edited by
W. D. Haglund and M. H. Sorg, pp. 295-317. CRC Press, Boca Raton, FL.
Gargett, R.
1996 Cave Bears and Modern Human Origins. The Spatial Taphonomy of Pod Hradem
Cave, Czech Republic. University Press of America, Lanham, MD.
Gargett, R. and B. Hayden
1991 Site structure, kinship and sharing in aboriginal Australia: implications for
archaeology. In The Interpretation of Archaeological Spatial Patterning., edited by E. M.
Kroll and T. D. Price, pp. 11-32. Plenum Press, New York.
Giddens, A.
1979 Central problems in social theory: action, structure, and contradiction in social
analysis. University of California Press, Berkeley.
Gifford, D. P.
1977 Observations of Modern Human Settlements as an Aid to Archaeological
Interpretation. Doctoral dissertation, University of California.
1978 Ethnoarchaeological observations on natural processes affecting cultural
materials. In Explorations in Ethnoarchaeology, edited by R. A. Gould, pp. 77-101.
University of New Mexico Press, Albuquerque.
Gifford, D. P. and A. K. Behrensmeyer
1977 Observed formation and burial of a recent human occupation site in Kenya.
Quaternary Research 8:245-266.
Gifford, D. P. and D. C. Crader
1977 A computer coding system for archaeological faunal remains. American Antiquity
42:225-238.
Gifford, D. P., G. L. Isaac and C. M. Nelson
1980 Evidence for predation and pastoralism at Prolonged Drift, a Pastoral Neolithic site
in Kenya. Azania 15:57-108.
Gifford-Gonzalez, D.
1984 Taphonomic specimens, Lake Turkana, pp. 419-428. National Geographic
Research Reports. vol. 1976 Projects. National Geographic Society, Washington, D. C.
1989a Ethnographic analogues for interpreting modified bones: some cases from East
Africa. In Bone Modification, edited by R. Bonnichsen and M. Sorg, pp. 179-246.
Institute for Quaternary Studies, University of Maine, Orono, ME.
1989b Modern analogues: developing an interpretive framework. In Bone Modification,
edited by R. B. a. M. Sorg, pp. ? Center for the Study of the First Americans, Orono.
1991a Bones are not enough: analogues, knowledge, and interpretive strategies in
zooarchaeology. Journal of Anthropological Archaeology 10:215-254.
1991b Examining and refining the quadratic crown height method of age estimation. In
Human Predation and Prey Mortality, edited by M. Stiner, pp. 41-78. Westview Press,
Boulder.
1993 Gaps in zooarchaeological analyses of butchery. Is gender an issue? In Bones to
Behavior: Ethnoarchaeological and Experimental Contributions to the Interpretation of
Faunal Remains, edited by J. Hudson, pp. 181-199. Southern Illinois University Press,
Carbondale.
Gifford-Gonzalez, D.
1998a Early pastoralists in East Africa: ecological and social dimensions. Journal of
Anthropological Archaeology 17(2):166-200.
1998b Gender and early pastoralists in East Africa. In Gender in African Prehistory,
edited by S. Kent, pp. 115-137. Altamira Press, Walnut Creek.
2000 Animal disease challenges to the emergence of pastoralism in sub-Saharan Africa.
African Archaeological Review 18.
Gifford-Gonzalez, D. and J. Kimengich
1984 Faunal evidence for early stock-keeping in the Central Rift of Kenya: preliminary
findings. In The Development and Spread of Food-Producing Cultures in Northeastern
Africa, edited by L. Krzyzaniak, pp. 457-471. Polish Academy of Science, Poznan.
Gifford-Gonzalez, D. and B. Wright
1986 A data management and table formatting system for vertebrate remains. In Data
Base Management and Zooarchaeology, edited by L. v. Wijngaarden-Bakker. PACT. vol.
14.
Gifford-Gonzalez, D. P., D. B. Damrosch, D. R. Damrosch, J. Pryor and R. L. Thunen
1985 The third dimension in site structure: an experiment in trampling and vertical
dispersal. American Antiquity 50(4):803-818.
Godfrey, K. M., T. Forrester, D. J. P. Barket, A. a. Jackson, J. P. Landman, J. S. E. Hall,
V. Cox and C. Osmond
1994 Maternal nutritional status in pregnancy and blood pressure in childhood. British
Journal of Obstetrics and Gynaecology 101(5):398-403.
Goody, J.
1983 Cooking, Cuisine, and Class: A Study in Comparative Sociology. Cambridge
University Press, Cambridge.
Gould, R. A.
1967 Notes on hunting, butchering, and sharing of game among the Ngatatjara and
their neighbors in the west Australian Desert. Kroeber Anthropological Society Papers
36:41-66.
Gowlett, J. A. J., J. W. K. Harris, D. Walton and B. A. Wood
1981 Early archaeological sites, hominid remains, and traces of fire from Chesowanja,
Kenya. Nature 294:125-129.
Grant, A.
1983 The use of tooth wear as a guide to the age of domestic ungulates. In Ageing and
Sexing Animal Bones from Archaeological Sites, edited by W. B., C. Grigson and S.
Payne, pp. 91-108. vol. 109. British Archaeological Reports Series, Oxford.
Grantham, B.
1995 Dinner in Buqata: the symbolic nature of food animals and meal sharing in a
Druze village. In The Symbolic Role of Animals in Archaeology, edited by K. Ryan and P.
J. Crabtree, pp. 73-78. MASCA Research Papers in Science and Archaeology. vol. 12.
University of Pennsylvania Museum of Archaeology and Anthropology, Philadelphia.
Grayson, D. K.
1984 Quantitative Zooarchaeology. Topics in the Analysis of Archaeological Faunas.
Studies in Archaeological Science. Academic Press, New York.
Grayson, D. K.
1988 Danger Cave, Last Supper Cave, and Hanging Rock Shelter: The Faunas.
American Museum of Natural History Anthropological Papers, Vol. 66, part 1 66 (1).
American Museum of Natural History.
Grayson, D. K.
1989 Bone transport, bone destruction, and reverse utility curves. Journal of
Archaeological Science 16:643-652.
Greenfield, H. J.
1988 Bone consumption by pigs in a contemporary Serbian village: implications for the
interpretation of prehistoric faunal assemblages. Journal of Field Archaeology 15:473-
479.
Guilday, J. E., P. Parmalee and D. Tanner
1962 Aboriginal butchering techniques at the Eschelman Site (36LA12), Lancaster
County, Pennsylvania. Pennsylvania Archaeologist 32:59-83.
Guiler, E. R.
1970 Observations on the Tasmanian Devil, Sarcophilus harrissi (Dasyuridae:
Marsupialia) 1. Numbers, home range, movements, and food in two populations.
Australian Journal of Zoology 18(49-62).
Gust, S. M.
1982 Faunal analysis and butchering. In The Ontiveros Adobe: Early Rancho Life in Alta
California, edited by J. Frierman, pp. 101-180. Greenwood and Associates, Pacific
Palisades, CA.
1993 Animal Bones from Historic Urban Chinese Sites: A Comparison of Sacramento,
Woodland, Tucson, Ventura, and Lovelock. In Hidden Heritage: Historical Archaeology
of the Overseas Chinese, edited by P. Wegars, pp. 177-212. Baywood Monographs in
Archaeology Series, R. L. Schuyler, general editor. Baywood Publishing Company, Inc.,
Amityville, New York.
Hancock, R. G. V., M. D. Grynpas and K. P. H. Pritzker
1989 The abuse of bone analyses for archaeological dietary studies. Archaeometry
31:169-180.
Hayden, B.
1995 The emergence of prestige technologies and pottery. In The Emergence of
Pottery. Technology and Innovation in Ancient Societies, edited by W. K. Barnett and J.
W. Hoopes, pp. 257-265. Smithsonian Institution Press, Washington D.C.
Haynes, G.
1980 Evidence of carnivore gnawing on Pleistocene and recent mammalian bones.
Paleobiology 6:341-351.
1983a Frequencies of spiral and green-bone fractures on ungulate limb bones in
modern surface assemblages. American Antiquity 48(1):102-114.
1983b A guide for differentiating mammalian carnivore taxa responsible for gnaw
damage to herbivore limb bones. Paleobiology 9(2):164-172.
1988 Longitudinal studies of African elephant death and bone deposits. Journal of
ArchaeologicalScience 15(2):131-158.
1991 Mammoths, Mastodons, and Elephants. Biology, Behavior, and the Fossil Record.
Journal of Archaeological Science. Cambridge University Press, Cambridge.
Henderson, P., C. A. Marlow, T. I. Molleson and C. T. Williams
1983 Patterns of chemical change during bone fossilization306:358-360.
Higgs, E. S. and J. P. White
1963 Autumn killing. Antiquity 37:282-289.
Hill, A.
1975 Taphonomy of Contemporary and Late Cenozoic East African Vertebrates. Ph. D.,
University of London.
1979a Disarticulation and scattering of mammal skeletons. Paleobiology 5:261-274.
Hill, A. P.
1979b Butchery and natural disarticulation: an investigatory technique. American
Antiquity 44:739-744.
Hill, A. P. and A. K. Behrensmeyer
1984 Disarticulation patterns of some modern East African mammals. Paleobiology
10:366-376.Hockett, B. S.
1989 The concept of "carrying range": a method for determining the role played by
woodrats in contributing bones to archaeological sites. Nevada Archaeologist 7(1):28-
35.
Hockett, B. S.
1991 Toward distinguishing human and raptor patterning on leporid bones. American
Antiquity 56(4).
1993 Taphonomy of the Leporid Bones from Hogup Cave, Utah: Implications for Cultural
Continuity in the Eastern Great Basin. Ph.D., University of Nevada.
Hoffman, R.
1988a The contribution of raptorial birds to patterning in small mammal assemblages.
Paleobiology 14(1):81-90.
1988b The eastern wood rat (Neotoma floridana) as a taphonomic factor in
archaeological sites. Journal of Archaeological Science 14:325-357.
Holbrook, S. J. and J. C. Mackey
1976 Prehistoric environmental change in northern New Mexico: evidence from a
Gallina phase archaeological site. The Kiva 41(4):309-317.
Holtzman, R. C.
1979 Maximum likelihood estimation of fossil assemblage composition. Paleobiology
5:79-89.
Homewood, K. M. and W. A. Rogers
1991 Maasailand Ecology. Pastoralist Development and WIldlife Conservation in
Ngorongoro, Tanzania. Cambridge University Press, Cambridge.
Horowitz, L. K. and P. Smith
1988 The effects of striped hyena activity on human remains. Journal of Archaeological
Science 15(5):471-482.
Houston, D., C. and J. A. Copsey
1994 Bone digestion and intestinal morphology of the bearded vulture. The Journal of
Raptor Research 28(2):73-78.
Hughes, A. R.
1961 Further notes on the habits of hyaenas and bone gathering by porcupines.
Zoological Society of South Africa New Bulletin 3(1).
Ingold, T.
1980 Hunters, Pastoralists, and Ranchers. Cambridge University Press, Cambridge.
Irving, W. N., A. V. Jopling and I. Kritsch-Armstrong
1989 Studies of bone technology and taphonomy. Old Crow Basin, Yukon Territory. In
Bone Modification, edited by R. Bonnichsen and M. H. Sorg, pp. 347-380. Center for the
Study of Early Man, University of Maine, Orono, ME.
Isaac, G. L.
1971 The diet of early man: aspects of archaeological evidence from Lower and Middle
Pleistocene sites in Africa. World Archaeology 2(3):278-299.
Isaac, G. L.
1983 Bones in contention: competing explanations for the juxtaposition of early
Pleistocene artifacts and faunal remains. In Animals and Archaeology: 1. Hunters and
their Prey., edited by J. Clutton-Brock and C. Grigson, pp. 3-19. Oxford ed. British
Archaeological Reports (International Series 163). vol. 163. BAR International, Oxford.
Isaac, G. L. and D. C. Crader
1981 To what extent were early hominids carnivorous? An archaeological perspective.
In Omnivorous Primates: Gathering and Hunting in Human Evolution, edited by G.
Teleki, pp. 37-103. Columbia University Press, New York.
James, S. R.
1989 Hominid use of fire in the Lower and Middle Pleistocene. Current Anthropology
30(1):1-26.
Jarvenpa, R. and H. J. Brumbach
1983 Ethnoarcheological perspectives on an Athapaskan moose kill. Arctic 36(2):174-
184.
1995 Ethnoarchaeology and gender: Chipewyan woman as hunters. Research in
Economic Anthropology 16:39-82.
Johnson, E.
1978 Paleo-indian bison procurement and butchering patterns on the Llano Estacado.
Plains Anthropologist 23(82):98-105.
Johnson, E.
1981 A review of Late Prehistoric Bison Procurement in Southeastern New Mexico: T.
American Antiquity 46(4):954-956.
1982a Paleo-Indian bone expediency tools: Lubbock Lake and Bonfire Shelter.
Canadian Journal of Anthropology 2(2):145-157.
1983 A framework for interpretation in bone technology. Paper presented at the
Carnivores, Human Scavengers and Predators: A Question of Bone Technology,
Calgary.
1985 Current developments in bone technology. Advances in Archaeological Method
and Theory 8:157-235.
Johnson, E., L. B. Davis and M. Wilson
1978 Bison Procurement and Utilization: A Symposium Plains Anthropologist Memoir
Paleo-Indian bison procurement and butchering pattern on the LLano Estacado.
Johnson, G. A.
1982b Organizational structure and scalar stress. In Theory and Explanation in
Archaeology: The Southampton Conference, edited by C. Renfrew, M. J. Rowlands and
B. A. Seagraves, pp. 389-422, New York: Academic Press.
Jones, K.
1983 Forager archaeology: the Aché of eastern Paraguay. Paper presented at the
Carnivores, Human Scavengers, and Predators: A Question of Bone Technology,
Calgary.
Jones, T. L.
1991 Marine-resource value and the priority of coastal settlement; a California:
perspective. American Antiquity 56(3):419-443.
Kaczor, M. J.
1986 . In PACT: Data Base Management and Zooarchaeology, edited by L. v.
Wijngaarden-Bakker. vol. 14, Leiden.
Keene, A. S.
1985 Nutrition and economy: models for the study of prehistoric diet. In The Analysis of
Prehistoric Diets, edited by R. I. Gilbert, Jr and J. H. Mileke. Academic Press, Orlando.
Kent, S.
1992 Studying Variability in the Archaeological Record: An Ethnoarchaeological Model
for Distinguishing Mobility Patterns. American Antiquity 57(4):635-660.
Kent, S.
1993 Sharing in an egalitarian Kalahari community. Man 28(3):481-514.
Kitching, J. W.
1963 Bone, Tooth and Horn Tools of Palaeolithic Man: an Account of the
Osteodontokeratic Discoveries in Pinhole Cave, Derbyshire. Manchester University
Press, Manchester.
Klein, R. G.
1976 The mammalian fauna of the Klasies River Mouth site, Southern Cape Province.
South African Archaeological Bulletin 31:75-98.
1978 Stone Age predation on large African bovids. Journal of Archaeological Science
5:195-217.
1982 Age (mortality) profiles as a means of distinguishing hunted species from
scavenged ones in Stone Age archeological sites. Paleobiology 8:151-158.
Klein, R. G., K. Allwarden and C. Wolf
1983 The calculation and interpretation of ungulate age profiles from dental crown
heights. In Hunter-Gatherer Economy in Prehistory: A European Perspective, edited by
G. N. Bailey, pp. 47-57. Cambridge University Press, Cambridge.
Klein, R. G. and K. Cruz-Uribe
1984 The Analysis of Animal Bones from Archaeological Sites.
Klepinger, L. L., J. K. Kuhn and W. S. Williams
1986 An elemental analysis of archaeological bone from Sicily as a test of predictability
of diagenetic change. American Journal of Physical Anthropology 70:325-332.
Kluckhohn, C.
1940 The conceptual structure of Middle American studies. In The Maya and their
Neighbors, edited by A. M. Tozzer, pp. 4-51. D. Appleton-Century Company, New York.
Kreutzer, L. A.
1992 Bison and deer bone mineral densities: comparisons and implications for the
interpretation of archaeological faunas. Journal of Archaeological Science 19(3):271-
294.
Kruuk, H.
1972 The Spotted Hyena: A Study of Predation and Social Behavior. University of
Chicago Press, Chicago.
Kurtén, B.
1953 On the variation and population dynamics of fossil and recent mammal
populations. Acta Zoologica Fennica 76:159-178.
Kyle, J. H.
1986 Effect of post-burial contamination on the concentrations of major and minor trace
elements. Journal of Archaeological Science 13:403-416.
Lambert, J. B., S. V. Simpson, C. B. Szpunar and J. E. Buikstra
1985 Bone diagenesis and dietary analysis. Journal of Human Evolution 14:477-482.
Langley-Evans, S. C., G. J. Phillips, R. Benediktsson, D. S. Gardner, C. R. W. Edwards,
A. A. Jackson and J. R. Seckl
1996 Protein intake in pregnancy, placental glucocorticoid metabolism, and the
programming o hypertension in the rat. Placenta 17(2-3):169-172.
Lawrence, D. R.
1968 Taphonomy and information losses in fossil communities. Bulletin of the
Geological Society of America 79:1315-1330.
1971 The nature and structure of paleoecology. Journal of Paleontology 45:593-607.
1979 Fossil diagenesis - Fossildiagenese. Paper presented at the Encyclopedia of
Paleontology, Stroudsberg, Pa.
Lee, R. B.
1979 The !Kung San: Men, Women, and Work in a Foraging Society. Cambridge
University Press, Cambridge.
Lee, R. B. and I. deVore (editors)
1968 Man the hunter. Aldine, Chicago.
Leechman, D.
1951 Bone grease. American Antiquity 16:355-356.
Leonard, W. R. and M. L. Robertson
1997 Rethinking the energetics of bipedality. Current Anthropology 38(2):304 -309.
Leonard, W. R., M. L. Robertson, L. C. Aiello and P. Wheeler
1996 On diet, energy metabolism, and brain size in human evolution. Current
Anthropology 37(1):125-129.
Levy, T. E.
1983 The mergence of specialized pastoralism in the southern Levant. World
Archaeology 15(1):15-36.
Lowe, V. P. W.
1957 Teeth as indicators of age with special reference to Red Deer Cervus elaphus.
Journal of Zoology 152:137-153.
Lupo, K. D.
1994 Butchering marks and carcass acquisition strategies: distinguishing hunting from
scavenging in archaeological contexts. Joural of Archaeological science 21:827-837.
1995 Hadza bone assemblages and hyena attrition: an ethnographic example of the
Influence of cooking and mode of discard on the intensity of scavenger ravaging.
Journal of Anthropological Archaeology 14(3):288-314.
Lyman, R. L.
1984 Bone density and differential survivorship in fossil classes. Journal of
Anthropological Archaeology 3:259-299.
1985 Bone frequencies, differential transport, and the MGUI. Journal of Archaeological
Science 12:221-236.
1987a Archaeofaunas and butchery studies: a taphonomic perspective. Advances in
Archaeological Method and Theory 10:249-337.
1987b Zooarchaeology and taphonomy: a general consideration. Journal of
Ethnobiology 7:93-117.
1991 Taphonomic problems with archaeological analyses of animal carcass utilization
and transport. In Beamers, Bobwhites, and Blue-Points: Tributes to the Career of Paul
W. Parmalee, edited by J. R. Purdue, W. E. Klippel and S. B. W, pp. 135-148. Scientific
Papers. vol. 23. Illinois State Museum, Springfield, Illinois.
1993 A study of variation in the prehistoric butchery of large artiodactyls. In Ancient
Peoples and Landscapes, edited by E. Johnson. Texas Tech University, Lubbock, TX.
Lyman, R. L.
1994 Vertebrate Taphonomy. Cambridge University Press, Cambridge.
Lyman, R. L. and G. L. Fox
1989 A critical evaluation of bone weathering as an indication of bone assemblage
formation. Journal of Archaeological Science 16:293-317.
Maguire, J. M., D. Pemberton and M. H. Collett
1980 The Makapansgat limeworks grey breccia: hominids, hyaenas, hystricids, or
hillwash? Palaeontologia Africana 23:75-98.
Marean, C. W.
1989 Sabertooth cats and their relevance for early hominid diet and evolution. Journal
of Human Evolution 18:559-582.
1995 Of taphonomy and zooarcheology. Review of Vertebrate Taphonomy, by R. Lee
Lyman. Evolutionary Anthropology :64-72.
1998 A critique of the evidence for scavenging by Neadertals and early modern
humans: new data from Kobeh Cave (Zagros Mountains, Iran) and Die Kelders Cave 1
Layer 10 (South Africa). Journal of Human Evolution 35:111-136.
Marean, C. W. and C. L. Ehrhardt
1995 Paleoanthropological and paleoecological implications of the taphonomy of a
sabertooth's den. Journal of Human Evolution 29(6):515-547.
Marean, C. W. and L. Spencer
1991 Impact of carnivore ravaging on zooarchaeological measures of element
abundance. American Antiquity 56(4):645-658.
Marean, C. W., L. M. Spencer, R. J. Blumenschine and S. D. Capaldo
1992 Captive hyaena bone choice and destruction, the schlepp effect and Olduvai
archaeofaunas. Journal of Archaeological Science 19:101-121.
Marshall, B. and R. Cosgrove
1990 Tasmanian Devil (Sarcophilus harrisi) scat-bone: signature criteria and
archaeological implications. Australia and Oceania 25:102-113.
Marshall, F. B.
1986 Implications of bone modification in a Neolithic faunal assemblage for the study of
early hominid butchery and subsistence practices. Journal of Human Evolution 15:661-
672.
1990 Cattle herds and caprine flocks. In Early Pastoralists of South-Western Kenya,
edited by P. T. Robertshaw, pp. 205-260. British Institute in Eastern Africa, Nairobi.
1993 Food-sharing and the faunal record. In Bones to Behavior: Ethnoarchaeological
and Experimental Contributions to the Interpretation of Faunal Remains, edited by J.
Hudson, pp. 228-246. Center for Archaeological Investigations, Carbondale, IL.
Mayhew, D. F.
1977 Avian predators as accumulators of fossil mammal material. Boreas 6:25-31.
Mayr, E.
1982 The Growth of Biological Thought: Diversity, Evolution, and Inheritance. Belknap
Press, Cambridge.
Mbae, B. N.
1986 Aspects of Maasai ethno-archaeology: implications for Archaeological
Interpretation. M. A., University of Nairobi.
McGuire, R. H.
1983 Breaking down cultural complexity: inequality and heterogeneity. Advances in
Archaeological Method and Theory 6:91-142.
McKee, L.
1991 Food supply and plantation social order: an archaeological view. In I, Too, Am
American: Studies in African-American Archaeology, edited by T. Singleton, pp. 1-49.
University Press of Virginia, Charlottesville.
Meadow, R. H.
1978 "Bonecode"- a system of numerical coding for faunal data from Middle Eastern.
Paper presented at the Approaches to Faunal Analysis in the Middle East, Cambridge,
Massachusetts.
Mellet, J.
1974 Scatological origin of microvertebrate fossil accumulations. Science 185:349-350.
Mengoni-Goñalons, G.
1980 Notas zooarqueológicas I: Fracturas en huesos. Actas del VII Congreso Nacional
de Arqueología, Colonia del Sacramento, Urugu :87-91.
Metcalfe, D. and K. T. Jones
1988 A reconsideration of animal body-part utility indices. American Antiquity 53(3):486-
504.
Miller, C. J.
1975 A study of cuts, grooves, and other marks on recent and fossil bone, 1:
Weathering cracks, fractures, splinters and other similar natural phenomena. In Lithic
Technology: Making and Using Stone Tools, edited by E. H. Swanson, pp. 211-228.
World Anthropology. Mouton, Chicago.
Mills, M. G. L.
1978 Foraging behavior of the brown hyaena (Hyaena brunnea Thunberg, 1820) in the
Southern Kalahari48:113-141.
Mills, M. G. L. and M. E. J. Mills
1978 The diet of the brown hyaena Hyaena brunnea in the Southern Kalahari. Koedoe
21:125-149.
Minnis, P., M. E. Whalen, J. H. Kelley and J. D. Stewart
1993 Prehistoric macaw breeding in the North American Southwest. American Antiquity
58(2):270-276.
Mitchell, C. J. and P. W. Smith
1991 Reliability of techniques for determining age in southern white-tailed deer. Journal
of Tennesse Academy of Science 66(3):117-120.
Moctezuma, E. M.
1984 The Great Temple of Tenochtitlan. Science 251(2):80-89.
Monastersky, R.
1993 A question of crushers. Why animals throughout history have developed a taste for
bones. Science News 144:396-397.
Morbeck, M. E., A. Galloway and A. L. Zihlman
1997 The Evolving Female: a Life History Perspective. Princeton University press,
Princeton.
Morlan, R. E.
1983 Spiral fractures on limb bones: which ones are artificial? Carnivores, Humans
Scavengers and Predators: A Question of Bone Modification :241-269.
Mundy, P. J. and J. A. Ledger
1976 Griffon vultures, carnivores, and bones. South African Journal of Science
72(6):106-110.
Myers, T. P., M. R. Voorhies and R. G. Corner
1980 Spiral fractures and bone pseudotools at paleontological sites. American Antiquity
45:483-490.
Nelson, B., M. DeNiro, M.Schoeninger,D. DePaolo
1986 Effects of diagenesis on strontium, carbon, nitrogen and oxygen concentration and
isotopic composition of bone. Geochimica et Cosmochimica Acta 50:1941-1949.
Nelson, D. A. and N. J. Sauer
1984 An evaluation of postdepositional changes in the trace element content of human
bone. American Antiquity 49:141-147.
Netting, R. M., R. R. Wilk and E. J. Arnould (editors)
1984 Households. University of California Press, Berkeley.
O'Connell, J. F.
1979 Room to move: Contemporary Alyawara settlement patterns and their implications
for Aboriginal housing policy. A Black Reality: Aboriginal Camps and Housing in Remote
Australia. Australian Institute of Aboriginal Studies, Canberra.
O'Connell, J. F., K. Hawkes and N. Blurton-Jones
1988a Hadza hunting, butchering, and bone transport and their archaeological
implications. Journal of Anthropological Research 44(2):113-161.
1988b Hadza scavenging: implications for Plio/Pleistocene hominid subsistence.
Current Anthropology 29(2):356-363.
1990 Reanalysis of large mammal body part transport among the Hadza. Journal of
Anthropological Archaeology 17:301-316.
O'Connell, J. F. and B. Marshall
1989 Analysis of kangaroo body part transport among the Alyawara of central Australia.
Journal of Archaeological Science 16:393-405.
Oliver, J. S.
1989 Analogues and site context: bone damages from Shield Trap Cave, Carbon
County, Montana, USA. In Bone Modification, edited by R. Bonnichsen and M. Sorg, pp.
73-98. Center for the Study of Early Man, Orono, Maine.
1993 Carcass processing by the Hadza: bone breakage from butchery to consumption.
In Bones to Behavior: Ethnoarchaeological and Experimental Contributions to the
Interpretation of Faunal Remains, edited by J. Hudson, pp. 200-227. Occasional Paper
No. 21. Center for Archaeological Investigations, SIU at Carbondale, Carbondale.
Olsen, S. L.
1989 On distinguishing natural from cultural damage on archaeological antler. Journal of
Archaeological Science 16:125-135.
Olsen, S. L. and P. Shipman
1988 Surface modification on bone: trampling versus butchery. Journal of
Archaeological Science 15(5):535-554.
Olson, E. C.
1957 Size frequency distributions in samples of extinct organisms. Journal of Geology
65:309-333.
1966 Community evolution and the origin of mammals. Ecology 47:291-302.
1971 Vertebrate Paleozoology. Wiley-Interscience, New York.
Olson, E. C.
1980 Taphonomy: its history and role in community evolution. Paper presented at the
Fossils in the Making, Chicago.
Parmalee, P. W.
1985 Identification and interpretation of archaeologically derived animal remains. In
Analysis of Prehistoric Diets, edited by R. I. Gilbert and J. H. Mielke, pp. 61-95.
Academic Press, Orlando.
Payne, S.
1969 A metrical distinction between sheep and goat metacarpals. Paper presented at
the Domestication and Exploitation of Plants and Animals, Chicago.
1972 Partial recovery and sample bias - the results of some sieving experiments. In
Papers in Economic Prehistory, edited by E. S. Higgs, pp. 49-64. Cambridge University
Press, Cambridge.
1973 Kill-off patterns in sheep and goats: the mandibles from Asvan Kale. Anatolian
Studies 23:281-303.
Pearsall, D. M.
1989 Paleoethnobotany: A Handbook of Procedures. Academic Press, San Diego.
Peters, C. R. and E. M. O'Brien
1981 The early hominid plant-food niche: insights from analysis of plant exploitation by
Homo, Pan, and Papio in eastern and southern Africa. Current Anthropology 22:127-
140.
Phillipson, D. W.
1984 Aspects of early food production in northern Kenya. In Origins and Early
Development of Food-producing Cultures in Northeastern Africa, edited by L.
Kryzyzaniak, pp. 489-495. Polish Academy of Science, Poznan.
Pike-Tay, A.
1991 L'analyse du cement dentaire chez les cerfs: l'application en prehistoire. Paléo
3(December):149-165.
Plug, I.
1978 Collecting patterns of six species of vultures (Aves: Accipitridae). Annals of the
Transvaal Museum 31(6):51-63.
Polednak, A. P.
1989 Various chronic disorders and other conditions: Diabetes Mellitus. In Racial and
Ethnic Differences in Disease, edited by A. P. Polednak, pp. 231-238. Oxford University
Press, New York.
Pond, C. M.
1996 The biological origins of adipose tissues in humans. In The Evolving female: A Life
History Perspective, edited by M. E. Morbeck, A. Galloway and A. Zihlman. Princeton
University Press, Princeton NJ.
1991 Adipose tissue in human evolution. In The Aquatic Ape: Fact of Fiction, edited by
M. Roede, J. Wind, J. M. Patrick and V. Reynolds, pp. 193-220. Souvenir Press Ltd.,
London.
Potts, R. B.
1982 Lower Pleistocene Site Formation and Hominid Activities at Olduvai Gorge,
Tanzania. Ph.D., Department of Anthropology, Harvard University.
1983 Foraging for faunal resources by early hominids at Olduvai Gorge, Tanzania. In
Animals and Archaeology: 1. Hunters and their Prey., edited by J. Clutton-Brock and C.
Grigson, pp. 51-62. International Series. vol. 163. British Archaeological Reports,
Oxford.
1984 Home bases and early hominids. American Scientist 72:338-347.
1986 Temporal span of bone accumulation at Olduvai Gorge and implications for early
hominid foraging behavior12:25-31.
1988 On an early hominid scavenging niche. Current Anthropology 29(1):153-155.
Potts, R. and P. Shipman
1981 Cutmarks made by stone tools on bones from Olduvai Gorge, Tanzania. Nature
291:577-580.
Pozorski, S.
1979 Late prehistoric llama remains from the Moche Valley, Peru. Annals of the
Carnegie Museum 48:139-170.
Price, T. D. and J. A. Brown
1985 Prehistoric Hunter-Gatherers: The Emergence of Cultural Complexity. New York,
Academic Press.
Price, T. D., R. W. Swick and E. P. Chase
1986 Bone chemistry and prehistoric diet: strontium studies of laboratory rats. American
Journal of Physical Anthropology 70:365-376.
Prummel, W. and H.-J. Frisch
1986 A guide for the distinction of species, sex, and body side in bones of sheep and
goat. Journal of Archaeological Science 13:567-577.
Rabinovitch, R. and L. K. Horwitz
1994 An experimental approach to porcupine damage to bones: a gnawing issue.
Artefacts 9:97-118.
Raczynski, J. and A. L. Ruprecht
1974 The effect of digestion on the osteological composition of owl pellets. Acta
Ornithologia 14:25-38.
Redding, R.
1981 Decision making in subsistence herding of sheep and goats in the Middle East.
Doctoral Dissertation, University of Michigan.
Reddy, S. N.
1997 If the threshing floor could talk: integration of agriculture and pastoralism during
the Late Harappan in Gujarat, India. Journal of Anthropological Archaeology 16:162-
187.
Reitz, E. J. and C. M. Scarry
1985 Reconstructing Historic Subsistence, with an Examples from Sixteenth Century
Spanish Florida Nº 3. Society for Historical Archaeology Special Publication, Ann Arbor.
Reitz, E. J. and E. S. Wing (editors)
1999 Zooarchaeology. Cambridge University Press, Cambridge.
Richardson, P. R. K., P. J. Mundy and I. Plug
1986 Bone crushing carnivores and their significance to osteodystrophy in griffon
vulture chicks. Journal of Zoology 210:23-43.
Richter, J.
1986 Experimental study of heat induced morphological changes in fish bone collagen.
Journal of Archaeological Science 13:477-481.
Rissman, P.
1986 Seasonal aspects of man/cattle interaction in Bronze Age western India6(2):257-
277.
1987 The potential of annular analysis for South Asian archaeology. Man and
Environment 11:15-24.
Robinette, W. L., D. A. Jones, G. Rogers and J. S. Gashwiler
1957 Notes on tooth development and wear for Rocky Mountain deer. Journal of Wildlife
Management 21:134-153.
Rolfe, W. D. I. and D. W. Brett
1969 Fossilization processes. Paper presented at the Organic Geochemistry, New York.
Rosenzweig, M. L.
1977 Aspects of biological exploitation. The Quarterly Review of Biology 52:371-380.
Rowley-Conwy, P.
1998 Improved separation of Neolithic metapodials of sheep (Ovis ) and goats (Capra )
from Arene Candide Cave, Liguria, Italy. Journal of Archaeological Science 25:251-258.
Ryder, M. L.
1959 The animal remains found at Kirkstall Abbey. Agricultural History Review 7:1-5.
Sadek-Kooros, H.
1972 Primitive bone fracturing: a method of research. American Antiquity 37(3):369-382.
Schaller, G. B.
1972 The Serengeti Lion: A Study of Predator/Prey Relations. University of Chicago
Press, Chicago.
Schick, K.
1986 Processes of Paleolithic Site Formation: An Experimental Study. Ph. D., University
of California.
1991 On making behavioral inferences from early archaeological sites. In Cultural
Beginnings: Approaches to Understanding Early Hominid Life-ways in the African
Savanna, edited by J. D. Clark, pp. 79-107. Monographien (Romisch-Germanisches
Zentralmuseum Mainz. Forschungsinstitüt fur Vor- und Frühgeschichte. R. Habelt,
Bonn.
Schick, K. D. and N. Toth
1993 Making Silent Stones Speak: Human Evolution and the Dawn of Technology.
Simon & Schuster, New York.
Schiffer, M. B.
1976 Behavioral Archaeology. Academic Press, New York.
1987 Formation Processes of the Archaeological Record. University of New Mexico
Press, Albuquerque, NM.
Schmitt, D. N.
1994 Toward the identification of coyote scatological faunal accumulations in
archaeological contexts. Journal of Archaeological Science 21:249-262.
1995 The taphonomy of golden eagle prey accumulations ot Great Basin roosts. Journal
of Ethnobiology 15(2):237-256.
Schmitt, D. N. and C. D. Zeier
1993 Not by bones alone: exploring household composition and socioeconomic status
in an isolated historic mining community. Historical Archaeology 27(4):20-27.
Scott, K.
1986 Review of L. R. Binford's Faunal Remains from Klasies River Mouth. Journal of
Archaeological Science 13:89-100.
Scrimshaw, N. S. and V. R. Young
1976 The requirements of human nutrition. Scientific American 275:50-64.
Seilacher, A.
1973 Biostratinomy: the sedimentology of biological standardized particles. Paper
presented at the Evolving Concepts in Sedimentology, Baltimore.
Selvaggio, M. M.
1998 Evidence for a three-stage sequence of hominid and carnivore involvement with
long bones at FLK Zinjanthropus, Olduvai Gorge, Tanzania. Journal of Archaeological
Science 25(3):191-202.
Sept, J. M.
1990 Vegetation studies in the Semliki Valley, Zaire as a guide to paleoanthropological
research. Virginia Museum of Natural History Memoir 1. Virginia Museum of Natural
History, Virginia.
Sept, J. M.
1994 Beyond bones: archaeological sites, early hominid subsistence, and the costs and
benefits of exploiting wild plant foods in East African riverine landscapes. Journal of
Human Evolution London 27:295-320.
Severinghaus, C. W.
1949 Tooth development and wear as criteria of age in white-tailed deer. Journal of
Wildlife Management 13:195-216.
Sherratt, A.
1983 The secondary exploitation of animals in The Old World. World Archaeology
15(1):90-104.
Shipman, P.
1981 Applications of scanning electron microscopy to taphonomic problems. Annals of
the New York Academy of Sciences 376:357-386.
1986 Scavenging or hunting in early hominids: theoretical framework and tests.
American Anthropologist 88:27-43.
1989 Altered bones from Olduvai Gorge, Tanzania: techniqies, problems, and
implications of their recognition. In Bone Modification, edited by R. Bonnichsen and M.
Sorg, pp. 317-334. Institute for Quaternary Studies, University of Maine, Orono, ME.
Shipman, P. and J. J. Rose
1983 Early hominid hunting, butchering, and carcass processing behavior: approaches
to the fossil record. Journal of Anthropological Archaeology 2:57-98.
Sillen, A., J. C. Sealy and N. J. van der Merwe
1989 Chemistry and paleodietary research: no more easy answers. American Antiquity
54(3):504-512.
Silver, I. A.
1963 The ageing of domestic animals. Paper presented at the Science in Archaeology,
New York.
Simons, J. W.
1966 The presence of leopard and a study of the food debris in the leopard lairs of the
Mount Suswa caves, Kenya. Bull. Cave Explor. Group E. Afr. 1:50-68.
Skinner, J. D. and R. J. v. Aarde
1991 Bone collection by brown hyaenas Hyaena brunnea in the Central Namib Desert,
Namibia. Journal of Archaeological Science 18:513-523.
Sobbe, I.
1990 Devils on the Darling Downs - The tooth markrecord. Memoirs of the Queensland
Museum 27:299-322.
Solomon, S.
1985 People and other aggravations: taphonomic research in Australia. B. A. Thesis,
University of New England.
Spector, J.
1982 Male-female task differentiation among the Hidatsa: toward the development of an
archaeological approach to the study of gender. In The Hidden Half: Studies of Native
Plains Women, edited by P. Albers and B. Medicine. University Press of America,
Washington, D.C.
1991 What this awl means: toward a feminist archaeology. In Engendering Archaeology,
edited by J. M. Gero and M. W. Conkey, pp. 388-406. Oxford, London.
Speth, J. D.
1983 Bison Kills and Bone Counts: Decision Making by Ancient Hunters. Prehistoric
archeology and ecology. University of Chicago Press, Chicago.
Speth, J. D. and K. A. Spielmann
1983 Energy source, protein metabolism, and hunter-gatherer subsistence strategies.
Journal of Anthropological Archaeology 2:1-31.
Spiess, A.
1976 Determining the season of death of archaeological fauna by analysis of
teeth29(1):53-54.
Spinage, C. A.
1971 Gerantodology and horn growth of the impala (Aepyceros melampus). Journal of
Zoology 164:209-225.
Spinage, C. A.
1972 Age estimation in zebra. East African Wildlife Journal 10:273-277.
Stahl, P. W.
1992 Diversity, body size, and the archaeological recovery of mammalian faunas in the
neotropical forests. Journal of the Steward Anthropological Society 20(1 &2):209-233.
1996 The recovery and interpretation of microvertebrate bone assemblages from
archaeological contexts. Journal of Archaeological Method and Theory 3(1):31-75.
Stallibrass, S.
1982 The use of cement layers for absolute ageing of mammalian teeth: a selective
review of the literature with suggestions for further studies and alternative applications.
In Ageing and Sexing Animal Bones from Archaeological Sites, edited by B. Wilson, C.
Grigson and S. Payne, pp. 109-26. British Series, 109. British Archaeological Reports,
Oxford.
Stein, G. J.
1987 Regional economic integration in early state societies: Third Millennium B. C.
pastoral production at Gritille, southeast Turkey. Paléorient 13:101-111.
Stiner, M. C.
1990 The use of mortality patterns in archaeological studies of hominid predatory
adaptations. Journal of Anthropological Archaeology 9:305-351.
1991 Food procurement and transport by human and non-human predators. Journal of
Archaeological Science 18:455-482.
1994 Honor Among Thieves: A Zooarchaeological Study of Neandertal Ecology.
Princeton University Press, Princeton, N.J.
Stiner, M. C., S. L. Kuhn, S. Weiner and O. Bar-Yosef
1995 Differential burning, recrystallization, and fragmentation of archaeological bone.
Journal of Archaeological Science 22:223-237.
Stini, W. A.
1980 Bioavailability of nutrients in human breast milk as compared to formula. Studies
in Physical Anthropology (Warsaw) 6:3-22.
Sutcliffe, A. J.
1973 Similarity of bones and antlers gnawed by deer to human artifacts. Nature
(London) 246:428-430.
1976 Further notes on bones and antlers chewed by deer and other ungulates. Deer
4:73-82.
Tanner, J. M.
1990 Fetus into Man. Physical Growth from Conception to Maturity. Revised ed.
Cambridge, MA, Harvard University Press.
Tappen, N. C. and G. R. Peske
1970 Weathering cracks and split-line patterns in archaeological bone. American
Antiquity 35:383-386.
Taylor, R. E., P. E. Hare and T. D. White
1995 Geochemical criteria for thermal alteration of bone. Journal of Archaeological
Science 22:115-119.
Teleki, G.
1973 The omnivorous chimpanzee. Scientific American 228(1):32-42.
Termine, J. D.
1993 4. Bone matrix proteins and the mineralization process. In Primer on the Metabolic
Bones Diseases and Disorders of Mineral Metabolism., edited by M. J. Favus, pp. 21-
25. 2nd ed. Raven Press, New York.
Thomas, D. H.
1986 Refiguring Anthropology: The First Principles of Probability and Statistics.
Waveland Press Inc., Prospect Heights.
Thomas, D. H. and D. Mayer
1983 The Archaeology of Monitor Valley: 2. Gatecliff Shelter. Anthropological Papers of
the American Museum of Natural History 59(1):353-391.
Thorson, R. M. and R. D. Guthrie
1984 River ice as a taphonomic agent: an alternative hypothesis for bone "artifacts".
Quaternary Research 22:172-188.
Todd, L. C.
1983 The Horner Site: Taphonomy of an Early Holocene Bison Bonebed. Ph.D.
dissertation, University of New Mexico.
Todd, L. C. and G. Frison
1986 The Colby Site: Taphonomy and Archaeology of Clovis Kill in Northwestern
Wyoming.
Turner, A.
1983 The quantification of relative abundances in fossil and subfossil bone assem.
Annals of the Transvaal Museum 33:311-321.
Tuross, N., A. K. Behrensmeyer and E. D. Eanes
1989 Strontium increases and crystallinity changes in taphonomic and archaeological
bone. Journal of Archaeological Science 16:661-672.
Van Devender, T. R.
1977 Holocene woodlands in the Southwestern deserts. Science 198:189-192.
Van Devender, T. R., P. S. Martin, R. S. Thompson, K. L. Cole, A. J. Jull, T. Long, A.
Toolin, L. J. and D. J. Donahue
1985 Fossil packrat middens and the tandem accelerator mass spectrometer. Nature
317:610-614.
Van Valkenburg, B., M. F. Teaford and A. Walker
1990 Molar microwear and diet in large carnivores: inferences concerning diet in the
sabertooth cat, Smilodon fatalis. Journal of Zoology (London) 222(2):319-340.
Van Valkenburgh, B.
1988 Incidence of tooth breakage among large predatory mammals. American
Naturalist 131(2):291-302.
Van Valkenburgh, B. and F. Hertel
1993 Tough times at La Brea: tooth breakage in large carnivores of the Late
Pleistocene. Science 261(5120):456-459.
Villa, P.
1982 Conjoinable pieces and site formation processes. American Antiquity 47:276-290.
Villa, P. and J. Courtin
1983 The interpretation of stratified sites: a view from underground. Journal of
Archaeological Science 10:267-281.
Vlach, J. M.
1993 Back of the Big House: The Architecture of Plantation Slavery. University of North
Carolina Press, Chapel Hill.
von den Driesch, A. and J. Boessneck
1975 Schnittspuren an neolithischen Tiernochen: Ein Beitrag zur Schlachttierzerlegung
in vorgeschichtlicher Zeit53:1-23.
Voorhies, M. R.
1969 Taphonomy and population dynamics of an early Pliocene vertebrate fauna, Knox
County, Nebraska. Wyoming University Contributions in Geology, Special Paper 1 .
Walker, P. L. and J. C. Long
1977 An experimental study of the morphological characteristics of tool marks.
American Antiquity 42:605-616.
Walshe, K.
1994 A Taphonomic Analysis of the Vertebrate Materials from Allen's Cave: Implications
for Australian Arid Zone Archaeology, Australian National University.
Washburn, S. L. and C. S. Lancaster
1968 The evolution of hunting. Paper presented at the Man the Hunter, Chicago.
Wattenmaker, P. and G. J. Stein
1986 Early pastoral production in southeast Anatolia: faunal remains from Kurban
Hüyük and Grittle Hüyük. Anatolica 13:90-96.
Weiss, K. M., R. E. Ferrell and C. L. Hanis
1984 A New World syndrome of metabolic diseases with a genetic and evolutionary
basis. Yearbook of Physical Anthropology 27:153-178.
Wendorf, M.
1989 Diabetes, the Ice Free Corridor, and Paleoindian settlement of North America.
American Journal of Physical Anthropology 79(503-520).
Wheat, J. B.
1966 A Paleo-indian bison kill. Scientific American 219:44-52.
1972 The Olsen-Chubbuck Site: A Paleo-Indian Bison Kill.
1979 The Jurgens Site. Anthropologist Memoir No. 15 .
White, T. D.
1992 Prehistoric Cannibalism at Mancos 5MTUMR-2346. Princeton University Press,
Princeton.
White, T. D. and N. Toth
1989 Engis: preparation damage, not ancient cutmarks. American Journal of Physical
Anthroplogy 78:361-367.
White, T. E.
1952 Observations on the butchering technique of some aboriginal peoples: I. American
Antiquity 17(4):337-338.
1953a A method of calculating the dietary percentages of various food animals utilized
by aboriginal peoples. American Antiquity 18(4):396-398.
1953b Observations on the butchering technique of some aboriginal peoples, No. 2.
American Antiquity 19:160-164.
1954 Observations on the butchering technique of some aboriginal peoples: Nos 3, 4, 5,
and 6. American Antiquity 19(3):254-264.
White, T. E.
1955 Observations on the butchering technique of some aboriginal peoples: Nos 7, 8,
and 9. American Antiquity 21:170-178.
Wilk, R. R. (editor)
1989 Household Economy: Reconsidering the Domestic Mode of Production. Westview
Press, Boulder.
Wissler, C.
1910 The material culure of the Blackfoot Indians. American Museum of Natural History,
Anthropological Papers 5(1).
Yalden, D. W. and P. E. Yalden
1985 An experimental investigation of examining Kestral diet by pellet analysis. Bird
Study 32:50-55.
Yanagisako, S. J.
1979 Family and household: the analysis of domestic groups. Annual Review of
Anthropology 8(1):61-205.
1984 Explicating residence: a cultural analysis of changing households among
Japanese-Americans. In Households, edited by R. M. Netting, R. R. Wilk and E. J.
Arnould, pp. University of California Press. University of California Press, Berkeley.
Yellen, J. E.
1977a Archaeological Approaches to the Present: Models for Reconstructing the Past.
Studies in Archeology. Academic Press, New York.
1977b Cultural patterning in faunal remains: evidence from the !Kung Bushmen. In
Experimental Archaeology, edited by D. Ingersoll, J. E. Yellen and W. Macdonald, pp.
271-331. Columbia University Press, New York.
1986 Optimization and risk in human foraging strategies. Journal of Human Evolution
15:733-750.
1991a Small mammals: !Kung San utilization and the production of faunal assemblages.
Journal of Anthropological Archaeology 10(1):1-26.
1991b Small mammals: post-discard patterning of !Kung San faunal remains. Journal of
Anthropological Archaeology 10:152-192.
Zeder, M. A.
1984 Meat distribution in the highland Iranian urban center of Tal-e Malyan. Paper
presented at the Animals and Archaeology. Early Herders and their Flocks 3.
Zeder, M. A.
1990 Animal exploitation at Tell Halif. Bulletin of the American Schools of Oriental
Research, Supplement 26:24-30.
1991 Feeding cities: specialized animal economy in the ancient Near East. Smithsonian
Institution Press, Washington, DC.
1995 The archaeobiology of the Khabur Basin. Bulletin of the Canadian Society for
Mesopotamian Studies 29:21-32.
Zeder, M. A. and S. R. Arter
1994 Changing patterns of animal utilization at ancient Gordion. Paléorient 20:105-118.
Ziegler, A. C.
1973 Inference from prehistoric faunal remains. Addison-Wesley Module in
Anthropology 43.
Zococo, T. G. and H. Schwartz
1994 Microstructural analysis of bone of the sauropod dinosaur Seismosaurus by
transmission electron microscopy. Paleontology 37(3):493-503.

Das könnte Ihnen auch gefallen