You are on page 1of 35

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/318588707

A Glossary of Zooarchaeological Methods

Chapter · March 2017

CITATIONS READS

0 2,210

2 authors:

Mauro Rizzetto Umberto Albarella


The University of Sheffield The University of Sheffield
27 PUBLICATIONS   16 CITATIONS    163 PUBLICATIONS   2,684 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Investigating changes in animal husbandry between the late- and the post-medieval periods in the Basque Country through the analysis of faunal remains View project

The Zooarchaeology of the Northern Iberian Peninsula during the Middle Ages View project

All content following this page was uploaded by Mauro Rizzetto on 21 February 2018.

The user has requested enhancement of the downloaded file.


A Glossary of Zooarchaeological Methods

Oxford Handbooks Online


A Glossary of Zooarchaeological Methods  
The Oxford Handbook of Zooarchaeology
Edited by Umberto Albarella, Mauro Rizzetto, Hannah Russ, Kim Vickers, and Sarah Viner-
Daniels

Print Publication Date: Mar 2017 Subject: Archaeology Online Publication Date: Apr 2017

(p. 757) A Glossary of Zooarchaeological Methods

Note

This methodological glossary presents brief explanations of the main analytical methods
employed by zooarchaeologists and makes reference to those chapters in the Handbook
that provide examples of their applications. The aim is to provide non-expert readers with
a basic understanding of how the evidence presented in this volume has been obtained.
The limitations and biases of each type of analysis are also briefly outlined, as their
acknowledgement represents an essential prerequisite to research on archaeological
materials. The definitions always refer to zooarchaeological applications of the term,
although many of them may be employed in other disciplines with similar or different
meanings. Each term makes reference to the chapters where it is mentioned, with the
exception of approaches that are invariably adopted (e.g. taxonomic identification,
quantification).

Ageing
(see Chapters 2, 5–21, 23, 26–31, 37, 39, 42, 43, 45)
The determination of the age-at-death of animals from their archaeological remains
can provide insights into culling, hunting, and foraging strategies. Kill-off patterns can
be reconstructed when a statistically reliable group of aged remains from the same
taxon is considered. For domestic animals, such patterns can be interpreted as the
intentional attempt of herders to focus animal exploitation on one or more services/
outputs, since each of them requires the animal to be killed at more or less specific
age-stages in order to optimize production. Hunters could also prefer (or be forced to
choose) specific age-stages, for optimizing the results of hunting activities, maintaining

Page 1 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

a stable animal population, or for other non-economic reasons. Kill-off patterns are
usually represented as bar charts or mortality curves, often in combination with an
equivalent survival curve.
Most of the ageing methods devised by zooarchaeologists are based on tooth eruption
and wear, as well as the epiphyseal fusion of postcranial bones. The former method
provides more detailed and reliable reconstructions of culling strategies, since the
gradual eruption and wear of teeth allows the identification of specific age-stages (e.g.
Ewbank et al., 1964; Payne, 1973; Grant, 1982; O’Connor, 1988). The different timing
at which diaphyses and epiphyses of long (p. 758) bones fuse allows grouping skeletal
elements into broader fusion age-stages (e.g. Silver, 1969), providing alternative or
complementary ageing information.
Other more sophisticated (though expensive and time-consuming) techniques are also
used. Growth rings on teeth and periosteal surface of bones, and the relative
frequency or area of osteons observed on bone thin sections can provide ageing
information with a variable degree of accuracy (Dammers, 2006). Counts of annual
layers of cementum from tooth sections can also be used for determining the age-at-
death of an animal and, in some cases, the season in which it died (Stallibrass, 1982).
Information on the overall age-at-death of an archaeological animal population can be
provided by those postcranial skeletal measurements that are more susceptible to
growth with age. Joints or parts of the bone that are less constricted by other bones or
ligaments tend to continue growing even after skeletal maturity; examples include the
smallest breadth of the diaphysis of long bones (SD) and the smallest length of the
scapular neck (SLC). Variability in such measurements may help in identifying age
groups, particularly for very young animals (Rowley-Conwy, 2001).
Attempts have been made to age cattle using size, shape, and texture of horn-cores
(Armitage, 1982); the size and complexity of cervid antlers are also correlated to age
(Brown, 1983), although other variables are at play (Clutton-Brock, 1981). The
incremental structures observed in many mollusc shells and in otoliths (fish ear bones)
represent specific growth periods and can be used for ageing (Jones, 1983; Secor et
al., 1995).
Allometry
This is the study of morphometrical relations between body parts. See Biometry.
Analytical error
(see Chapters 12, 38)
Any error resulting from the mistaken use of primary data can be defined as such.
Analytical errors can be determined by the misuse of analytical methods in relation to
the research question or to the nature of the material, by errors made by the
researcher when processing data, or by overlooking methodological pitfalls. Analytical
errors are distinguished from ‘input errors’, which are instead the result of recovery
biases and other taphonomic processes.
Archaeoentomology
(see Chapter 15)

Page 2 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

This is the study of insect remains from archaeological sites. Insects may represent
reliable palaeoenvironmental indicators and can provide information about the
occurrence of animal parasitic diseases (see Palaeopathology).
Archaeomalacology
(see Chapters 17, 22, 32, 39, 41, 43–7)
This is the study of mollusc remains from archaeological sites. Molluscs (mainly
marine) have been extensively exploited as a source of food and raw materials, and
their shells (mainly terrestrial) can also be used in palaeoenvironmental
reconstructions and seasonality studies (see Seasonality).
Artistic sources.
See Documentary and artistic sources.
Biometry
(see Chapters 2, 5–9, 11–13, 15, 16, 18–21, 25, 26, 28, 30, 34, 42, 43)
The size and shape of past animals can be studied through biometrical analysis of their
remains. Measurements can be taken in many different ways and this is useful in order
to address different research questions, but some form of standardization is valuable
to allow direct comparison of different datasets. Many different systems of taking
measurements for specific taxa or across different taxa have been proposed, but the
most universally used manual for mammals and birds is that of von den Driesch (1976).
(p. 759)

Biometry is regularly employed to distinguish between taxa that cannot be easily


separated on morphological grounds, such as sheep and goat (Payne, 1969), pig and
wild boar (Payne and Bull, 1988), equids (Davis, 1982; Johnstone, 2005), camelids
(Wapnish, 1984), felids (O’Connor, 2007), and fish (Libois and Hallet-Libois, 1988). It
can be particularly useful for investigating the appearance of domesticates, as these
are usually smaller than their wild ancestors (e.g. Albarella et al., 2006; Zeder, 2006;
Dobney et al., 2007). Biometrical differences can also be used to separate breeds of
the same species (e.g. Harcourt, 1974; De Cupere et al., 2005; MacKinnon, 2010), or
identify successful attempts of ‘improvement’ of domesticates (e.g. Albarella et al.,
2008; Davis, 2008), as well as reverse processes of size decrease (e.g. Teichert, 1984;
Forest and Rodet-Belarbi, 1998). Biometrical differences can also be the consequence
of changing environmental conditions (e.g. Davis, 1981; Hill et al., 2008; Albarella et
al., 2009) (see Palaeoenvironmental reconstruction).
Genuine size and shape differences between two animal populations, however, can be
masked or over-emphasized by the existence of different sex ratios, age patterns, or
pathological incidences. Sexual dimorphism affects certain elements more than others
(e.g. distal metapodials, distal humerus); these can be ruled out to avoid the
interference of sexual dimorphism, or purposely selected when investigating changes
in the sex ratio or the relative proportions of male, female, and castrate individuals.
The same procedure can be applied for those measurements that are more affected by
ageing (e.g. the smallest breadth of long bone diaphyses) (see Ageing). Teeth are
relatively unaffected by sexual dimorphism and age and tend to react slower to size
changes, unless new genetic stock is introduced into a population; this latter property

Page 3 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

of tooth biometry can be employed, for example, to distinguish between local


improvements and introduction of new breeds (e.g. Albarella et al., 2008).
Biometrical analyses are usually visualized in histograms or scatter plots, and results
are often supported by statistical tests. The analyses can be applied to anatomical
elements individually or by merging the available data through a scaling index
technique (see Scaling index technique). Single measurements can also be used to
estimate the withers height of an animal.
Butchery
(see Chapters 2, 3, 8, 10, 11, 13–15, 21, 26, 28, 30, 31, 34, 36–9, 43, 45)
Butchery marks generally refer to any traces left by humans when processing animal
carcasses; such activities include culling, skinning, splitting of the carcass, and
removal or extraction of flesh, fat, ligaments, and marrow. A basic classification of the
types of butchery marks may distinguish between cuts, chops, saw marks, and
percussion marks, although much more detailed classifications have been produced on
the basis of size, morphology, and location of marks (e.g. Lauwerier, 1988). Attempts
are sometimes also made to distinguish between fractures occurring on fresh and dry
bones (e.g. Binford, 1981), which has obvious consequences for their interpretation.
Butchery evidence tends to be recorded more intensely on large species (e.g. cattle),
due to the multiple steps required by the processing of larger carcasses and the
reduction of larger muscle masses into sizeable portions. For these animals, it is
usually possible to detect and distinguish between primary and secondary butchery;
the former consists of the initial removal of inedible or less-meaty appendices (e.g.
lower limb bones, head) and a first dismemberment of the carcass, while the latter is
much more variable and includes further sub-division into smaller units of meat,
careful cutting or more energetic chopping through the joints, filleting, marrow
extraction, and evidence for curing. In addition, the inedible wastes of carcass-
processing can represent an abundant and valuable source of raw materials, and have
been variably exploited for the production of functional, ritual, and ornamental items
(Choyke and O’Connor, 2013). A number of manufacture traditions even developed into
specialized (p. 760) industries, as in the case of leather production, horn-working, or
comb-making (e.g. Albarella, 2003; Rijkelijkhuizen, 2011).
A better understanding of the nature and purpose of carcass-processing marks can be
achieved through experimental studies (e.g. Jones, 1980; Seetah, 2008; Lloveras et al.,
2009).
The analysis of butchery marks, along with the distribution of body parts, the material
culture recovered from the site, and documentary sources, can provide information on
the types of tools employed for carcass-processing and their purpose (e.g. Seetah,
2006; Maltby, 2007).
Comparative collection.
See Reference collection.
Coprolites (fossilized dung).
See Dung analysis.
Dating

Page 4 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

(most chapters)
Faunal assemblages are organized into chronological phases according to the dating of
archaeological contexts. Often this is achieved through a typological analysis of
associated artefacts (relative dating), but direct dating of archaeological materials,
including fauna, is also sometimes possible (absolute dating). A number of dating
techniques based on the radioactive decay of naturally occurring elements (e.g. 14C,
Uranium-series, Potassium-Argon, and thermoluminescence) or the magnetic
properties of some minerals can be used.
Because of their organic component and abundance, faunal remains are among the
best archaeological materials for 14C (radiocarbon) dating, when they are less than
approximately fifty thousand years old. For this to be effective it is important that the
association between faunal material and stratigraphic context is secure; articulated
bones in primary deposit are therefore ideal for radiocarbon dating.
The identification of allochthonous species can be used to establish a terminus post
quem for the formation of the deposit under consideration, should the date of arrival of
such species be known. This is the case, for example, of the deliberate introduction of
domesticates or wild animals in some areas, or of the accidental spread of commensal
species (see examples in O’Connor and Sykes, 2010). For more ancient periods and
larger timespans, the rapid speciation undergone by some micromammals can be used
in relative dating; the evolution of three sub-species of water vole with different dental
conformations, for example, allowed the reconstruction of a chronological sequence for
the European Pleistocene (Wilkinson and Stevens, 2008; though see Martin, 2014 for a
critique of the method). The succession of distinctive combinations of mammal species
(biostratigraphy) during the Pliocene and Pleistocene, as a result of consistent climatic
changes, allowed identification of specific biochrones, which can be used for dating
archaeological and palaeontological contexts (e.g. Currant and Jacobi, 2001; Kahlke et
al., 2011).
Density
(see Chapters 5, 6, 41)
Density represents the quantification of animal remains (either wholly or for specific
taxa) per volume unit. Density analysis can provide information on taphonomic
processes, differential use of archaeological contexts, duration, and nature of site
occupation, intensity of animal exploitation, and formation of archaeological contexts.
Diagenesis
This term describes the physical, chemical, and biological processes of decay
undergone by organic remains after burial. It refers to all taphonomic processes
occurring after burial and before excavation (see Taphonomy).

Page 5 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Diagnostic zone method


This is a recording method that identifies a pre-determined set of anatomic zones as
those requiring consistent recording and counting. These ‘zones’ are known (p. 761) as
‘diagnostic’ because they are generally chosen on the basis of their degree of
identifiability at the anatomic and taxonomic level. Diagnostic zones are also selected
on the basis of their likely high chance of preservation in the archaeological context
and the amount of information that they can provide. The method was first introduced
by Watson (1979), but has then been adopted and developed by many other
researchers (e.g. Dobney and Rielly, 1988; Serjeantson, 1991; Davis, 1992; Albarella
and Davis, 1994).
The diagnostic zone method allows researchers to overcome many of the biases related
to the nature of archaeological faunal assemblages and the different conditions under
which research is conducted. The degree of preservation of faunal assemblages is
highly variable, depending on a wide range of largely unpredictable taphonomic
processes (see Taphonomy); in addition, researchers with different experience and
access to different facilities will tend to produce results of variable detail. All these
variables must be taken into account, with the result that, in most cases, different
studies conducted in different research environments cannot be reliably compared.
The selection of diagnostic zones, however, mitigates problems deriving from
unpredictable taphonomic variables, subjectivity, and other research biases; at the
same time, the method counteracts over-representation of larger species, over-
counting of fragments from the same element, and, to some extent, differences in
skeletal complexity between taxa.
In order to avoid potential loss of information, a diagnostic zone method can be
integrated with the recording of ‘non-countable’ specimens, which will provide
qualitative rather than quantitative information. The number of chosen diagnostic
zones can vary substantially, but, in case of highly fragmented assemblages, the
number of chosen zones will need to be large enough to capture sufficient information.
Distribution of anatomical elements
(see Chapters 2, 8, 10, 13–15, 22–4, 27–32, 36, 39, 41, 45, 46)
The distribution of anatomical elements is the result of human choices of carcass
processing and disposal as well as post-depositional taphonomic processes. These
latter may generate biases such as the preferential destruction of less robust elements
(e.g. Brain, 1967), the partial recovery of smaller bones and teeth when sieving is not
undertaken, and the further selection imposed by the recording process. Once the
extent of such biases has been acknowledged, information on carcass-processing and
disposal can be extrapolated. For food mammals, a distinction is commonly made
between meat-bearing (e.g. upper limb bones) and non-meat-bearing elements (e.g.
lower limb bones, cranial elements). A higher incidence of the former tends to suggest
the prevalence of waste from consumption and secondary butchery, whereas primary
butchery will generally result in a predominance of non-meat-bearing elements (see
Butchery). Alternatively, an even distribution of body parts can be interpreted as the
result of unspecialized butchery and/or waste disposal practices. Butchery practices

Page 6 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

can, however, be determined by the need to maximize energy optimization, as well as


by cultural and religious traditions (e.g. Milne and Crabtree, 2001).
Ritual practices, bone-working, and the presence of specific animal products (e.g.
hides, furs) can produce distinctive body part distributions (e.g. Albarella, 2003).
The analysis of body part distribution can be undertaken using a variety of different
methods, but these will need to consider taphonomic biases as well potentially uneven
distribution of anatomical elements in different taxa.
DNA.
See Genetics.
Documentary and artistic sources
(see Chapters 3, 8–16, 21, 26, 29, 31, 36, 40, 41) (p. 762)
This is the study of textual evidence, and iconographic and other artistic reproductions
for the investigation of the economic, religious, and symbolic role of animals in past
societies. These studies can complement the results of zooarchaeological analysis (e.g.
MacKinnon, 2001; Soderberg, 2004).
Dung analysis
(see Chapters 10, 16)
This is the analysis of animal palaeofeces recovered from archaeological contexts. The
morphology and content of dung is distinctive for each species and can signal an
animal’s presence even when its physical remains are not recovered. The content of
animal dung can provide information on diet, health, and palaeoenvironmental
conditions (e.g. Maldre, 2006); dung could have also been used as fuel or building
material (Lancelotti and Madella, 2012).
Epiphyseal fusion.
See Ageing.
Ethnozooarchaeology
(see Chapters 10–12, 20, 21, 27, 36, 41, 44, 46)
This is a zooarchaeological research strategy (David and Kramer, 2001) that employs
ethnographic evidence as an aid to the interpretation of faunal assemblages from
archaeological sites (e.g. Binford, 1978; Bartosiewicz and Greenfield, 1999; Albarella
and Trentacoste, 2011; Halstead, 2014).
Experimental zooarchaeology
Zooarchaeologists can learn from the performance of experiments involving animals,
animal remains, or animal products in order to formulate or verify theories and
interpretations of past animal exploitation practices.
Experiments have been classified as imitative (when reproducing activities performed
in the past), reconstructive (when aiming at conserving or recreating rare or extinct
species and breeds), or as reproductions and observations of taphonomic processes
(van Wijngaarden-Bakker, 1987). This latter approach represents a prominent line of
research in zooarchaeology, due to the important impact that taphonomic biases have
on the nature and preservation of faunal assemblages. Experiments focusing on
butchery and cooking practices, as well as diagenetic processes and recovery biases,
have improved our ability to evaluate taphonomic biases (e.g. Payne, 1975; Nagaoka,

Page 7 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

1994; Nicholson, 1996; Fernández-Jalvo and Andrews, 2003). In addition, they have
also helped the identification of potential cooking patterns (e.g. Lupo and Schmitt,
1997; Vacca, 2008).
Most ageing and sexing methods used by zooarchaeologists are also based on
observations of modern animal populations (e.g. Payne, 1973; Munro et al., 2011;
Popkin et al., 2012).
Fragmentation
(see Chapters 8, 24, 37, 45)
This is the degree of breakage of animal remains. It can provide information on the
nature and impact of taphonomic processes, including specific butchery activities and
recovery biases (e.g. Outram, 2001; Maltby, 2007) (see Taphonomy).
Frequency of species
(all chapters)
This is the relative abundance of identified taxa within a faunal assemblage. It relies
on what has been recorded and the adopted quantification methods, and can be
visualized in charts, ternary plots, or tables. The changing frequency of one or more
taxa in relation to others can be investigated through the use of indices (e.g. the cattle
index used by Badenhorst, this volume).
Genetics
(see Chapters 7, 12–14, 16, 18–21, 25, 32, 42, 43, 47)
This is the study of genetic information from modern and ancient animal specimens to
aid zooarchaeological research; it is mainly based on nuclear and mitochondrial DNA
(mtDNA). (p. 763) Indirect genetic evidence can also be obtained from the amino-acid
sequences of proteins, which are encoded in the DNA (Weiner, 2010).
DNA analysis on archaeological specimens can provide a range of data, including
taxonomic identification, sex, and inherited pathological conditions, as well as genetic
affinity to a known haplotype. This latter line of research has contributed to the
identification of the wild ancestors of many domestic species. In particular, the
combination of studies on modern and ancient DNA (aDNA) is employed to reveal the
geographical area of origin of domestication and patterns of diffusion of domesticates
into other areas (e.g. Larson et al., 2005; Beja-Pereira et al., 2006; Cai et al., 2011).
Comparisons between mtDNA (maternally inherited) and Y-chromosome (only present
in male individuals) studies allow investigation of differences in genetic contributions
on the maternal and paternal lines (e.g. Götherström et al., 2005).
Preservation biases and high costs restrict the use of DNA to specific case studies,
which can nonetheless complement and verify the results from other types of analysis
(e.g. Davis et al., 2012; Telldahl et al., 2012).
Geometric morphometrics
This is the study of the shape of animal remains based on the location of a determined
set of features or landmarks. Variations on the position of landmarks in relation to
each other on different specimens are analysed using statistics; the morphological
differences highlighted by such methods can be used in taxonomic identification or in

Page 8 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

the analysis of morphometrical changes within an animal species or other taxonomic


group (e.g. Bignon et al., 2005; Ottoni et al., 2013).
Iconography.
See Documentary and artistic sources.
Identification.
See Taxonomic identification.
Input error.
See Analytical error.
Isotope analysis.
See Stable isotope analysis.
Kill-off pattern.
See Ageing.
MAU (Minimum number of Animal Units)
A quantification method calculated from the MNE, whereby this latter is adjusted
according to the frequency of each element within the skeleton of a species (Binford,
1984; Lyman, 2008). In order to allow comparisons between different samples, values
can be expressed as a percentage of the highest MAU value.
Microwear.
See Tooth microwear.
MNE (Minimum Number of anatomical Elements)
(see Chapter 5)
This is a quantification method that identifies the minimum number of anatomical
elements necessary to account for the remains observed for each taxon.
The nature of the basic unit count for this method—the ‘element’—has been widely
debated. According to different researchers, it can refer to complete anatomical
elements, incomplete bones (whose integrity can be estimated as a percentage),
anatomical parts ‘reconstructed’ by overlapping landmark features, or skeletal
portions (e.g. lower fore limb, thoracic girdle); in addition, the MNE can be calculated
by considering or disregarding differences in age, sex, and size (Grayson, 1984; Reitz
and Wing, 1999; for an overview of different approaches, see Lyman, 2008). Paired
elements can be distinguished between left and right.
The MNE provides the basis for calculations of the MAU.

Page 9 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

MNI (Minimum Number of Individuals)


(see Chapters 13, 27, 28, 31, 36, 45–7) (p. 764)
This is a quantification method that identifies the minimum number of animals
necessary to account for the number of remains observed for each taxon. The MNI can
be regarded to be either the highest MAU or MNE for each taxon.
Since the unit of analysis (and interpretation) is the entire animal, differences in
skeletal complexity between taxa do not bias this method, which is also less affected
than others by recovery bias, as it relies on the most common element, regardless of
its size. At the same time, however, it may lead to an over-estimation of rarer species
within an assemblage (especially when samples are small) and may vary substantially
with different aggregations of archaeological contexts (Grayson, 1984; Lyman, 2008).
In addition, adopting single individuals as the unit of analysis may be misleading due
to the fact that meat exploitation may have focused on carcass portions rather than
entire animals, which would have been processed and disposed of accordingly
(Binford, 1978).
It is important to consider that the main aim of the MNI is not to calculate, or even
approximate, the actual number of animals present on site at a given time—very rarely
possible for archaeological assemblages—but rather provide another measure of
relative taxonomic frequency.
Morphometry.
See Biometry.
NISP (Number of Identified Specimens)
(most chapters)
This is a quantification method that considers every countable element for each taxon,
as defined by the counting method employed. Although it overcomes the problem of
aggregation and over-estimation of rarer species (see MNI), it does not take into
account the different size and skeletal composition of different taxa, leading to an
over-estimation of larger species. For this reason, some authors provide adjusted
measures of NISP (e.g. Albarella and Davis, 1996). In addition, the NISP introduces the
problem of interdependence, whereby more fragments from the same animal may
contribute to the relative frequency of a given species (Grayson, 1984; Lyman, 2008).
The basic unit count for this method—the ‘specimen’—can be variably interpreted,
either explicitly (see Diagnostic zone method) or implicitly.
Organic residue analysis
(see Chapters 6, 8, 14, 17, 20)
Ancient organic materials can sometimes be detected on tools and containers
recovered from archaeological sites. Organic materials that are less susceptible to
diagenetic processes (e.g. fats and resins) can be identified through chemical analysis,
mass spectrometry, or gas liquid chromatography (Weiner, 2010). Organic residue
analysis has been contributing to many important archaeological questions, such as
food preferences, cuisine and the origins of domestication and dairying (e.g.
Spangenberg et al., 2006; Evershed et al., 2008; Mukherjee et al., 2008; Salque, 2012).
Osteometry.

Page 10 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

See Biometry.
Palaeoenvironmental reconstruction
(see Chapters 6, 16, 19, 39–41, 45)
The reconstruction of past environmental conditions represents an important aspect of
zooarchaeological research, though some taxa are better than others for this purpose
(Evans and O’Connor, 1999; Wilkinson and Stevens, 2008). Micromammals, land
molluscs, and insects tend to be sensitive to environmental conditions, and can provide
valuable information about vegetation cover, climate, moisture level, and many other
palaeoenvironmental features (e.g. Montuire, 1999; Davies, 2008; Vickers et al., 2011).
(p. 765)

Larger mammals are usually more flexible in terms of environmental adaptability and
often move or migrate over large distances; however, their presence or absence can
still provide a broad indication of regional environmental conditions (e.g. Jousse and
Escarguel, 2006; Kahlke et al., 2011). In addition, biometrical changes can be the
result of important climatic developments, since body size and temperature tend to be
inversely related in large mammal species (e.g. Davis, 1981; Bell and Walker, 2005).
Stable isotope analysis can also be relevant to environmental analysis (see Stable
isotope analysis).
Palaeogenetics.
See Genetics.

Page 11 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Palaeopathology
(see Chapters 6–8, 19, 20, 23, 29, 38)
In zooarchaeology, palaeopathological studies try to understand how animal disease
may be related to, or can affect, the interaction between humans and animals.
As in humans, pathological manifestations in animals can be the result of
developmental deficiencies or anomalies, ageing (including arthropathies), traumatic
lesions, inflammatory diseases (infections), parasitic activities, and other
environmental stresses, as well as inherited (genetic) disorders. The pathological
evidence detected on faunal remains is often non-specific and its interpretation is
hindered by equifinality, whereby different pathological conditions would result in
similar reactions and manifestations (Bartosiewicz with Gál, 2013). Since
palaeopathological analyses on archaeological materials are limited to observations of
the normally preserved hard tissues (e.g. bones, teeth, and shells) those diseases
leaving no traces on such tissues will not be detected. In the absence of soft tissues,
indirect evidence for these pathological conditions can be provided by, for example,
the recovery of parasites or carriers of parasites, the occurrence of animal mass
graves suggestive of epidemics, documentary sources, and studies combining
environmental conditions and animal biology (Baker and Brothwell, 1980).
Particular attention has been paid by zooarchaeologists to those pathological
insurgences resulting from human exploitation of animals. Animal domestication led to
a number of anatomical, biological, and behavioural changes that resulted in an
increase in developmental anomalies and pathological conditions. Overcrowding and
intensive husbandry practices could also facilitate the insurgence and spread of
diseases, while protracted inbreeding would have increased the incidence of genetic
anomalies. A number of bone and dental pathologies are the direct result of the use of
animals as a workforce. Overworking in draught cattle, for example, could lead to
arthritic conditions in the shoulder and pelvic girdle as well as on the lower limb
bones, while horseback riding often resulted in spondylotic fusion and ankyloses of
vertebral elements (Bartosiewicz et al., 1997).
Quantification
(all chapters)
Through quantitative methods zooarchaeologists make estimates of the relative
frequencies of different categories, such as age, sex, and size groups, but, most
typically, the term ‘quantification’ is associated with an evaluation of body part and
taxonomic frequencies. Quantitative methods can rely on raw fragment counts (e.g.
NISP), adjustments of raw fragment counts based on animal anatomy (e.g. MNI, MNE,
MAU), or meat weight estimations. More than one counting method can be included in
zooarchaeological analyses, in order to enhance comparability with other studies and
evaluate the biases generated by different methods.
Recording
(all chapters)
This is the initial process of selection and data collection performed by
zooarchaeologists on faunal assemblages. The material is recorded in order to provide

Page 12 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

quantitative and qualitative information to be used in data analysis and interpretation.


(p. 766)

The set of recorded data varies according to the type and degree of preservation of
each specimen, as well as to research questions, experience of the researcher, and
time and facilities available.
Recovery bias
(see Chapters 5, 6, 10–12, 14, 15, 17, 20, 22, 32, 33, 40, 46)
The retrieval of faunal remains from archaeological excavations is almost invariably
incomplete, and this will generate a bias. Typically, the bias consists of an under-
representation of smaller species and smaller anatomical elements. The extent of
recovery bias is determined by excavation and collection strategies, as well as by a
range of case-specific variables.
When only hand collecting is implemented, the quality of recovery relies on variable
circumstances such as the physical characteristics of the soil, work and climatic
conditions, excavation techniques (e.g. shovelling, trowelling), and the excavators’
experience and tiredness.
The impact of such variables is considerably reduced by sieving (especially water-
sieving) and flotation. The size of the sieve mesh plays an important role in rectifying
the recovery bias produced by hand collection; small meshes are obviously more
efficient but require longer time to process samples, and choices must be made
according to research questions and time/money constraints. An alternative solution is
to devise a systematic strategy for the sieving of sub-samples. Due to the qualitative
and quantitative differences of sieved and unsieved samples, it is important to keep
sieved material separate from the rest; mixing them would simply blur the recovery
bias of hand-collected remains and compromise the information added by sieved
materials. There will be loss of faunal material even through the sieving mesh, but at
least sieving will allow monitoring of the degree of loss, as indicated by the mesh size.
The importance of sieving (or acknowledging the stronger recovery bias of unsieved
assemblages) has been the focus of a number of experimental studies employing
modern and archaeological materials, which highlighted how partial recovery can
result in a biased frequency of species, lower taxonomic abundance, and misleading
distribution of anatomical elements (e.g. Payne, 1975; Nagaoka, 1994; Zohar and
Belmaker, 2005).
The extent of the recovery bias can be estimated in order to verify the reliability of
zooarchaeological analyses. Methods of bias estimation may rely on comparisons of the
frequency of a small element in relation to a large one it articulates with (e.g. astragali
and distal tibiae, first phalanges and distal metapodials, loose teeth and jaws), with the
assumption that adjoining elements are likely to be deposited in roughly equal
quantities; such frequencies are then compared between species of different size.

Page 13 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Reference collection
This is a collection of known animal specimens used for taxonomic and anatomic
identification of unknown faunal remains. Reference collections often specialize in
specific geographical areas and/or tend to focus on some classes of animals, unless
owned by large museums, and therefore not primarily focused on their
zooarchaeological potential. They can be organized by individual or anatomical
element and mainly consist of modern material, although ancient specimens can also
be included (e.g. extinct breeds or species). For ease of use bones are generally kept
loose rather than as mounted skeletons, as the latter are more appropriate for
anatomical studies and public displays.
Scaling index technique
(see Chapters 12, 42)
This is a biometrical method that compares different measurements taken on
archaeological material to a standard set of the same measurements, obtained from
either modern or ancient material. The absolute values of different measurements are
thus converted into relative values, which can be plotted on the same scale. This can
result in a substantial (p. 767) amplification of the size of the sample being analysed;
however, some measurements are affected to different extents by different biological
variables, compromising the readability of the results (i.e. potential patterns or
changes are either masked or over-emphasized) (Meadow, 1999; Albarella, 2002). For
this reason, measurements should be selected according to the research question. In
addition, length, breadth, and depth measurements should ideally be considered
separately, as dimensions lying on different axes can react differently and their
analysis may reveal changes in shape (Davis, 1996; Meadow, 1999). Even more
importantly, tooth and bone measurements, which respond differently to
environmental and genetic factors, need to be analysed separately. A very important
added value of this method is that, even when measurements are plotted separately,
they can be compared directly as they will be using the same scale. This can allow
inferences regarding shape and not just size.
Different scaling techniques have been used by zooarchaeologists since the
introduction of the size index (SI) in the late 1960s (Ducos, 1968); Uerpmann (1979)
re-elaborated the same concept by calculating the relative size index (RSI) and the
variability size index (VSI); Meadow (1981) sought to simplify Uerpmann’s indices by
adopting Simpson et al.’s (1960) log size index (LSI) (Meadow, 1999). The LSI
calculates the decimal logarithm of the ratio between each measurement and the
standard (i.e. of the SI); this latter method is currently the most widely used in
zooarchaeology, since it is easy to calculate and provides a better visual representation
of results (Albarella, 2002; Albarella and Payne, 2005).
Sclerochronological analysis
(see Chapter 43)
This is the study of incremental growth structures in the hard tissues of some
invertebrates (e.g. mollusc shells). It provides information on ageing,

Page 14 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

palaeoenvironmental conditions, and on seasonality of human exploitation of these


resources (see Ageing, Palaeoenvironmental reconstruction, Seasonality).
Seasonality
(see Chapters 10, 13, 32, 34, 37, 40, 43, 45, 46)
Zooarchaeologists are interested in the season when an animal died as this can
provide important information about past human behaviour. Seasonality is here
intended in terms of warm/dry and cold/wet seasons, whose alternation affects
population dynamics of wild species and decision-making of herders.
Patterns of seasonality are determined by a number of variables, including the
biological characteristics of animal species, environmental conditions, residence
patterns, technology, and the level of success of agricultural and husbandry practices,
as well as the economic and socio-political choices of a community (Reitz and Wing,
1999). Seasonal periodicity can be investigated through the presence and incidence of
some animal species (e.g. migratory birds, anadromous fish from inland sites), stable
isotope analysis (e.g. the 18O/16O ratio from the last growth ring on mollusc shells,
reflecting a warm or cold season), observation of incremental growth structures (on
mammal teeth, mollusc shells, bird medullary bone, on some fish bones and otoliths) or
seasonal anatomic features (e.g. antlers in cervids, when the pedicle is present),
ageing (e.g. evidence for seasonal culling of juvenile mammals), biometry, and disposal
patterns (for some examples, see Lefèvre, 1997; Cerón-Carrasco, 1998; Uchiyama,
1999; Higham and Horn, 2000; Pike-Tay and Cosgrove, 2002; Balasse et al., 2012;
Colonese et al., 2012).
Zooarchaeologists may in some cases detect the seasonality of a certain activity (e.g.
hunting or foraging patterns, culling of domesticates) but this does not necessarily
coincide with the seasonal occupation of a site.
Sexing
(see Chapters 8, 12, 13, 16, 20, 23, 26, 30, 31, 37, 43) (p. 768)
The sexual identification of animal remains relies on morphological, physiological, and
biometrical characters, as well as genetics. Most sexing methods based on
morphological and physiological characters are conceived for specific taxa. A study of
the shape and size of cattle horn-cores from modern animals has been used as an aid
to sex remains from archaeological sites (Armitage, 1982). Attempts have been made
using other morphometric traits of the skull and a number of postcranial bones
(Grigson, 1982a; 1982b); in particular, the conformation of the pelvis proved to be
useful in distinguishing sexes in bovids (Grigson, 1982a; Prummel and Frisch, 1986),
and the same method has been extended to other ungulates (Greenfield, 2006). In
other species, some specific anatomical elements are distinctive of male or female
individuals.
Antlers, for example, are only grown by male deer, with the exception of reindeer/
caribou (Rangifer tarandus) (Schmid, 1972). Similarly, a number of animal species
(including some primates, rodents, insectivores, and carnivores) have a penis bone
(baculum) in males and a clitoral bone (baubellum) in females.

Page 15 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Male birds grow a spur near the distal end of the tarsometatarsus; its presence or
absence is often used to estimate sex-ratios in faunal assemblages, especially within
large samples of domestic fowl remains (Sadler, 1990). Limitations to this method
include the timing of appearance of the spur (immature specimens cannot be sexed)
and the effects of caponization (i.e. castration), which results in a more or less visible
scar on the bone (West, 1982). Birds can also be sexed by verifying the presence or
absence of the medullary bone, which is deposited by laying birds inside their long
bones (Driver, 1982).
Male suids develop clearly distinctive canines, which can be easily distinguished from
female teeth.
Biometrical analysis of sexually dimorphic elements rarely allows sex identification on
individual specimens but can provide estimates of the sex-ratio at the population level
(see Biometry).
The analysis of aDNA allows sexing even small fragments with a certain degree of
confidence, although it remains bounded to preservation biases, methodological
limitations, and financial constraints (Weiner, 2010) (see Genetics). For this reason, it
is performed only occasionally and in specific circumstances; DNA sex identification on
sub-samples, for example, can verify the validity of more widely applicable sexing
methods, such as biometry (e.g. Davis et al., 2012; Telldahl et al., 2012).
Sieving.
See Recovery bias.
Site catchment analysis
(see Chapters 10, 11, 14, 16, 32, 33, 40, 41, 44–6)
This is the study of the area of procurement of natural resources ordinarily used by the
members of a community for their basic needs (Higgs and Jarman, 1975). Procurement
strategies within the catchment area evaluate costs, risks, and advantages; these, in
turn, depend on a number of variables, such as resource availability, technology,
residence pattern, and a range of economic and socio-political choices (Reitz and
Wing, 1999).
Information on a site’s catchment area can be provided by frequency of animal species,
taxonomic abundance, density of remains, and biometry. Zooarchaeological analyses
can highlight evidence for over-exploitation of faunal resources, extent and nature of
the territory exploited, environmental changes, and changes in procurement strategies
(i.e. choices and technology). Evaluation of the roles of all these variables, however,
must be supported by the analysis of complementary contextual information.
Spatial distribution
(see Chapters 8, 14, 23, 32, 40, 41) (p. 769)
Patterns of spatial distribution can provide information on the nature and scale of
animal exploitation, site function, area use, and waste disposal practices (e.g. Enloe et
al., 1994; Navas et al., 2008). Information from faunal remains should be
complemented by other lines of archaeological evidence (e.g. structures and features,
other materials), and can be biased by the partial excavation of a site.
Species frequency.

Page 16 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

See Frequency of species.


Stable isotope analysis
(see Chapters 4, 10, 16, 17, 19, 21, 25, 26, 28, 32, 34, 39, 42, 47)
The analysis of the relative frequencies of stable isotopes of specific elements can be
useful for the investigation of animal diet, management, and provenance, as well as the
reconstruction of palaeoenvironmental conditions.
Each chemical element is represented in nature by a number of isotopes, which have
the same number of protons (i.e. the same chemical properties) but a different number
of neutrons (i.e. atomic weight, determined by the sum of protons and neutrons).
Isotopes can be unstable (or radioactive), when the excess of neutrons causes
instability and gradual release of energy (e.g. 14C, which decays into 12C), or stable.
The proportion of stable isotopes in the biosphere varies according to various chemical
processes and can be analysed to investigate a number of issues (Price et al., 1985). In
zooarchaeology, samples for isotopic analysis can be extracted from both the organic
(collagen) and inorganic (hydroxyapatite) components of bones, from tooth enamel and
from mollusc shells. The different isotopic compositions of carbon (14C/13C), nitrogen
(15N/14N), oxygen (18O/16O), sulphur (34S/32S), and strontium (87Sr/86Sr) measured on
samples of animal tissues are a reflection of the food intake, climatic conditions, and
local isotope geochemistry, and therefore can provide important dietary and
environmental information (e.g. Bocherens and Drucker, 2003; Stevens et al., 2008;
Rawlings and Driver, 2010). Since tooth enamel is not remodelled once formed, intra-
tooth variability in isotopic composition can be analysed to detect changes in an
animal’s diet and residence (Balasse, 2002), providing useful insights in animal
management and movement (e.g. Viner et al., 2010; Balasse et al., 2012).

Page 17 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Taphonomy
(see Chapters 2, 4–6, 11, 14, 24, 29, 31, 32, 36, 37, 40, 41, 43, 46)
Taphonomy was defined by Efremov (1940, 85) as ‘the study of the transition (in all of
its details) of animal remains from the biosphere to the lithosphere’. Post-mortem
taphonomic processes in particular are the focus of zooarchaeological analyses, as
they usually include most of the modifications undergone by animal remains and, more
importantly, most of those resulting from human animal exploitation, such as butchery,
selection of body parts, and burning. Non-human modifications include gnawing and
partial digestion by other animals, as well as exposure to weathering agents (e.g.
water, wind). The incorporation of animal remains into archaeological deposits (either
as deliberate burial or through sedimentation) results in further modifications and
selection of the remains. The composition, acidity, and humidity of sediments
surrounding animal remains can affect their preservation by initiating, accelerating, or
arresting physical and/or chemical processes; the type and extent of microbial attack
also depends on such variables. Finally, excavation, recovery, analysis, and storage can
also affect the preservation of animal remains (Lyman, 1994).
The nature and extent of taphonomic alterations can be estimated by analysing the
taxa and anatomical elements present, the degree of fragmentation, and the type and
degree of macroscopic and microscopic diagenetic (i.e. after burial) surface
modifications (Lyman, 1994; Child, 1995; Weiner, 2010). Such analyses and the ability
to recognize taphonomic biases have been improved by a number of experimental
studies (e.g. Shipman et al., 1984; Nicholson, 1996; (p. 770) Fernández-Jalvo and
Andrews, 2003). Aside from estimating effects on preservation, the analysis of
taphonomic processes provides useful information on practices of animal exploitation,
as well as on palaeoenvironmental and sedimentological conditions (e.g. Tappen, 1994;
Jans et al., 2004; Turner-Walker and Jans, 2008; Vacca, 2008).
Taxonomic abundance
(see Chapters 3, 11, 19, 22, 32, 33, 38, 45, 46)
This method aims to measure the degree of taxonomic variability within an
assemblage. Low levels of taxonomic abundance are typical of specialized practices of
animal exploitation (e.g. agricultural communities focusing on few domesticates); on
the contrary, high taxonomic variability characterizes broad-spectrum economies, as in
the case of the Epipalaeolithic ‘broad-spectrum revolution’ (Flannery, 1969).
Taxonomic abundance, however, is usually biased (sometimes to a considerable extent)
by taphonomic biases; differential preservation and incomplete recovery can
exponentially reduce the number of taxa recorded within an assemblage, especially
when sieving is not implemented (e.g. Payne, 1975; Zohar and Belmaker, 2005) (see
Recovery bias).
Taxonomic identification
(all chapters)
The attribution of an animal specimen to a taxon relies on morphological and
biometrical observations of taxon-specific traits, based on the researcher’s experience

Page 18 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

and comparisons with anatomical atlases and reference collections (see Reference
collection).
Biometrical analyses can be employed to distinguish between taxa that are difficult to
separate on morphological grounds (see Biometry). Species-level identification using
aDNA can be fairly precise, although it is subjected to a number of practical limitations
(see Genetics). Similarly, the analysis of peptide mass fingerprinting in bone collagen
can provide reliable taxonomic identification (Buckley, 2009).
Textual evidence.
See Documentary and artistic sources.
Tooth microwear
(see Chapters 21, 43)
Microwear and mesowear patterns on teeth are analysed for reconstructing animal
diet, which can provide information on past animal management and
palaeoenvironmental conditions (e.g. Hogue, 2006; Vanpoucke et al., 2009; Kahlke and
Kaiser, 2011). The results from faunal remains can be compared to experimental
studies on modern animals (e.g. Mainland, 1998; 2003).
Management issues addressed by microwear studies include the distinction between
grazing, free-ranging, and stall-fed animals (e.g. Mainland, 1998; Vanpoucke et al.,
2009); differential feeding practices observed on samples from the same domestic
population can be the result of the different use of animals deposited in distinctive
archaeological contexts (e.g. Mainland and Halstead, 2005). Changing microwear
patterns can also reflect the seasonality of hunting activities (e.g. Rivals and Deniaux,
2005).
Tooth wear.
See Ageing.
Weight method
(see Chapters 27, 28, 38, 41, 46)
This is a quantification method that measures relative abundance on the basis of
weight rather than counts. It relies on raw weight itself or estimations of the potential
edible output represented by faunal remains. For this latter version of the method,
calculations are made by multiplying a factor, empirically derived from modern
animals, by either the bone weight or the MNI of a specific taxon. Meat weight
quantifications aim to interpret the relative (p. 771) importance of food animals on the
basis of their potential meat output rather than numerical frequency (Reitz and Wing,
1999).
The reliability of establishing a constant relationship between bone weight/MNI and
meat weight has been widely debated (e.g. Casteel, 1978; Jackson, 1989; Vigne, 1991).
Meat quantification implies a series of methodological biases and subjective
assumptions about past practices of animal exploitation. Multiplying factors are
obtained from modern animals: the considerable improvement and fattening of modern
breeds do not make them an ideal proxy for past domestic types; at the same time,
weight is not a fixed parameter and fluctuates according to age, sex, and seasonality
(Stewart and Stahl, 1977). In addition, species-level identification is not always

Page 19 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

possible, leading to further increased error margins when dealing with higher
taxonomic groups (e.g. Caprinae—sheep/goat, Sus sp.—pig/wild boar, Bos sp.—cattle/
aurochs, Cervidae, etc.) (Vigne, 1991).
Meat weight quantification methods based on bone weight also overlook patterns of
body part distribution: different elements are associated with different amounts of
meat and, therefore, distributions skewed towards meat-bearing or non-meat-bearing
elements imply sharply different meat weights. In addition, immature bones are lighter
(not only smaller) than those of adult and elderly individuals, potentially leading to an
underestimation of meat weight from young animals.
The meat and offal weight (MOW) method proposed by Vigne (1988) tries to overcome
these methodological drawbacks by including age, sex, and biometrical information
into the calculations. However, this also results in an accumulation of analytical biases
and observer errors typical of such analyses; in addition, since it is based on the MNI,
it carries the assumption that complete carcasses were introduced and consumed at
the site (Vigne, 1991), which is not necessarily the case.
It is important to consider that the meat weight method relies on the assumption that
carcasses were fully exploited for food, and therefore this cannot be applied to non-
food animals or cases where consumption was incomplete.
Withers height
(see Chapters 2, 26)
This is the height of an animal calculated from the withers (highest point of the
shoulders) to the ground. In zooarchaeology, it is calculated by multiplying bone
lengths by a factor derived from studies of modern animals (Reitz and Wing, 1999).
Similarly to the scaling index technique (see Scaling index technique), it allows
comparing different measurements, thus potentially amplifying the sample size of
biometrical data (O’Connor, 2000); however, since complete lengths are rarely
measurable on fragmented archaeological material, this potential is rarely fulfilled. It
has been argued that the multiplying factors devised from modern animals cannot
reliably account for the variety of morphometrical differences characterizing ancient
faunal populations, especially when considering domestic breeds or geographical types
(Forest, 1998). For these reasons, more recent research has placed less emphasis on
the use of withers heights.
Zoogeography
(see Chapters 2, 4, 9–11, 16, 18, 40, 44, 45)
The study of the geographic distribution of animal species is very important in
zooarchaeology as this can change over time, as a consequence of the modification of
environmental and/or cultural contexts. Zooarchaeology can contribute to reconstruct
past ranges of certain species, both in terms of human-mediated movement of animals
and natural expansion/shrinkage of animal populations (e.g. O’Connor and Sykes,
2010; Forcina et al., 2015).

(p. 772) Acknowledgements

Page 20 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

We are grateful to Jane Ford for comments on an earlier draft.

Albarella, U. (2002) ‘ “Size matters”: how and why biometry is still important in
zooarchaeology’, in Dobney, K. and O’Connor, T. (eds) Bones and the Man: Studies in
Honour of Don Brothwell, pp. 51–62. Oxford: Oxbow Books.

Albarella, U. (2003) ‘Tanners, tawyers, horn working and the mystery of the missing goat’,
in Murphy, P. and Wiltshire, P. (eds) The Environmental Archaeology of Industry, pp. 71–
86. Oxford: Oxbow Books.

Albarella, U. and Davis, S. J. M. (1994) ‘The Saxon and Medieval Animal Bones Excavated
1985–1989 from West Cotton, Northamptonshire’. Ancient Monuments Laboratory Report
17/94.

Albarella, U. and Davis, S. J. M. (1996) ‘Mammals and birds from Launceston Castle,
Cornwall: decline in status and the rise of agriculture’, Circaea, 12(1), 1–156.

Albarella, U. and Payne, S. (2005) ‘Neolithic pigs from Durrington Walls, Wiltshire,
England: a biometrical database’, Journal of Archaeological Science, 32(4), 589–99.

Albarella, U. and Trentacoste, A. (eds) (2011) Ethnozooarchaeology: The Present and Past
of Human-Animal Relationships, Oxford: Oxbow Books.

Albarella, U., Dobney, K., and Rowley-Conwy, P. (2006) ‘The domestication of the pig (Sus
scrofa): new challenges and approaches’, in Zeder, M. A., Bradley, D. G., Emshwiller, E.,
and Smith, B. D. (eds) Documenting Domestication: New Genetic and Archaeological
Paradigms, pp. 209–27. Berkeley: University of California Press.

Albarella, U., Dobney, K., and Rowley-Conwy, P. (2009) ‘Size and shape of the Eurasian
wild boar (Sus scrofa), with a view to the reconstruction of its Holocene history’,
Environmental Archaeology, 14(2), 103–36.

Albarella, U., Johnstone, C., and Vickers, K. (2008) ‘The development of animal husbandry
from the late Iron Age to the end of the Roman period: a case study from south-east
Britain’, Journal of Archaeological Science, 35, 1828–48.

Armitage, P. (1982) ‘A system for ageing and sexing the horn cores of cattle from British
Post-Medieval sites (17th to early 18th century) with special reference to unimproved
British longhorn cattle’, in Wilson, B., Grigson, C., and Payne, S. (eds) Ageing and Sexing
Animal Bones from Archaeological Sites. BAR British Series 109, pp. 37–54. Oxford:
Archaeopress.

Baker, J. and Brothwell, D. (1980) Animal Diseases in Archaeology, London: Academic


Press.

Page 21 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Balasse, M. (2002) ‘Reconstructing dietary and environmental history from enamel


isotopic analysis: time resolution of intra-tooth sequential sampling’, International Journal
of Osteoarchaeology, 12, 155–65.

Balasse, M., Boury, L., Ughetto-Monfrin, J., and Tresset, A. (2012) ‘Stable isotope insights
(δ18O, δ13C) into cattle and sheep husbandry at Bercy (Paris, France, 4th millennium BC):
birth seasonality and winter leaf foddering’, Environmental Archaeology, 17(1), 29–44.

Bartosiewicz, L. and Greenfield, H. J. (eds) (1999) Transhumant Pastoralism in Southern


Europe: Recent Perspectives from Archaeology, History and Ethnology, Budapest:
Archaeolingua.

Bartosiewicz, L. with Gál, E. (2013) Shuffling Nags, Lame Ducks: The Archaeology of
Animal Disease, Oxford: Oxbow Books.

Bartosiewicz, L., Van Neer, W., Lentacker, A., and Fabiš, M. (1997) Draught
(p. 773)

Cattle: Their Osteological Identification and History. Annalen Zoölogische


Wetenschappen 281. Tervuren: Koninklijk Museum voor Midden-Afrika.

Beja-Pereira, A., Caramelli, D., Lalueza-Fox, C., Vernesi, C., Ferrand, N., Casoli, A.,
Goyache, F., Royo, L. J., Conti, S., Lari, M., Martini, A., Ouragh, L., Magid, A., Atash, A.,
Zsolnai, A., Boscato, P., Triantaphylidis, C., Ploumi, K., Sineo, L., Mallegni, F., Taberlet, P.,
Erhardt, G., Sampietro, L., Bertranpetit, J., Barbujani, G., Luikart, G., and Bertorelle, G.
(2006) ‘The origin of European cattle: evidence from modern and ancient DNA’,
Proceedings of the National Academy of Sciences of the United States of America,
103(21), 8113–18.

Bell, M. and Walker, M. J. C. (2005) Late Quaternary Environmental Change: Physical and
Human Perspectives, 2nd edn, Harlow: Pearson-Prentice Hall.

Bignon, O., Baylac, M., Vigne, J.-D., and Eisenmann, V. (2005) ‘Geometric morphometrics
and the population diversity of Late Glacial horses in Western Europe (Equus caballus
arcelini): phylogeographic and anthropological implications’, Journal of Archaeological
Science, 32, 375–91.

Binford, L. R. (1978) Nunamiut Ethnoarchaeology, New York: Academic Press.

Binford, L. R. (1981) Bones: Ancient Men and Modern Myths, Orlando: Academic Press.

Binford, L. R. (1984) Faunal Remains from Klasies River Mouth, New York: Academic
Press.

Bocherens, H. and Drucker, D. (2003) ‘Trophic level isotopic enrichment of carbon and
nitrogen in bone collagen: case studies from recent and ancient terrestrial ecosystems’,
International Journal of Osteoarchaeology, 13, 46–53.

Page 22 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Brain, C. (1967) ‘Hottentot food remains and their bearing on the interpretation of fossil
bone assemblages’, Scientific Papers of the Namib Desert Research Station, 32, 1–11.

Brown, R. D. (ed.) (1983) Antler Development in Cervidae, Kingsville: Caesar Klebery


Wildlife Research Institute.

Buckley, M., Collins, M., Thomas-Oates, J., and Wilson, J. C. (2009) ‘Species identification
by analysis of bone collagen using matrix-assisted laser desorption/ionisation time-of-
flight mass spectrometry’, Rapid Communications in Mass Spectrometry, 23(23), 3843–
54.

Cai, D., Tang, Z., Yu, H., Han, L., Ren, X., Zhao, X., Zhu, H., and Zhou, H. (2011) ‘Early
history of Chinese domestic sheep indicated by ancient DNA analysis of Bronze Age
individuals’, Journal of Archaeological Science, 38, 896–902.

Casteel, R. W. (1978) ‘Faunal assemblages and the “Wiegemethode” or weight method’,


Journal of Field Archaeology, 5, 71–7.

Cerón-Carrasco, R. (1998) ‘Fishing: evidence for seasonality and processing of fish for
preservation in the northern isles of Scotland during the Iron Age and Norse times’,
Environmental Archaeology, 3, 73–80.

Child, A. M. (1995) ‘Towards an understanding of the microbial decomposition of


archaeological bone in the burial environment’, Journal of Archaeological Science, 22,
165–74.

Choyke, A. and O’Connor, S. (2013) From These Bare Bones: Raw Materials and the Study
of Worked Osseous Objects, Oxford: Oxbow Books.

Clutton-Brock, T. H. (1981) ‘The functions of antlers’, Behaviour, 79(2/4), 108–25.

Colonese, A. C., Verdún-Castelló, E., Álvarez, M., Briz i Godino, I., Zurro, D., and
Salvatelli, L. (2012) ‘Oxygen isotopic composition of limpet shells from the Beagle
Channel: implications for seasonal studies in shell middens of Tierra del Fuego’, Journal
of Archaeological Science, 39, 1738–48.

Currant, A. and Jacobi, R. (2001) ‘A formal mammalian biostratigraphy for the late
Pleistocene of Britain’, Quaternary Science Reviews, 20, 1707–16.

Dammers, K. (2006) ‘Using osteohistology for ageing and sexing’, in Ruscillo, D.


(p. 774)

(ed.) Recent Advances in Ageing and Sexing Animal Bones: Proceedings of the 9th
Conference of the International Council of Archaeozoology, Durham, August 2002, pp. 9–
39. Oxford: Oxbow Books.

David, N. and Kramer, C. (2001) Ethnoarchaeology in Action, Cambridge: Cambridge


University Press.

Davies, P. (2008) Snails: Archaeology and Landscape Change, Oxford: Oxbow Books.

Page 23 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Davis, S. J. M. (1981) ‘The effects of temperature change and domestication on the body
size of Late Pleistocene to Holocene mammals of Israel’, Paleobiology, 7(1), 101–14.

Davis, S. J. M. (1982) ‘A Trivariate Morphometric Method to Discriminate between First


Phalanges of Equus hydruntinus, asinus/hemionus, and caballus’, Tubingen meet 1982.

Davis, S. J. M. (1992) ‘A Rapid Method for Recording Information about Mammal Bones
from Archaeological Sites’. Ancient Monuments Laboratory Report 19/92.

Davis, S. J. M. (1996) ‘Measurements of a group of adult female Shetland sheep skeletons


from a single flock: a baseline for zooarchaeologists’, Journal of Archaeological Science,
23, 593–612.

Davis, S. J. M. (2008) ‘Zooarchaeological evidence for Moslem and Christian


improvements of sheep and cattle in Portugal’, Journal of Archaeological Science, 35,
991–1010.

Davis, S., Svensson, E., Albarella, U., Detry, C., Götherström, A., Pires, A. E., and Ginja, C.
(2012) ‘Molecular and osteometric sexing of cattle metacarpals: a case study from 15th
century AD Beja, Portugal’, Journal of Archaeological Science, 39, 1445–54.

De Cupere, B., Van Neer, W., Monchot, H., Rijmenants, E., Udrescu, M., and Waelkens, M.
(2005) ‘Ancient breeds of domestic fowl (Gallus gallus f. domestica) distinguished on the
basis of traditional observations combined with mixture analysis’, Journal of
Archaeological Science, 32, 1587–97.

Dobney, K. and Rielly, K. (1988) ‘A method for recording archaeological animal bones: the
use of diagnostic zones’, Circaea, 5(2), 79–96.

Dobney, K., Ervynck, A., Albarella, U., and Rowley-Conwy, P. (2007) ‘The transition from
wild boar to domestic pig in Eurasia, illustrated by a tooth developmental defect and
biometrical data’, in Albarella, U., Dobney, K., Ervynck, A., and Rowley-Conwy, P. (eds)
Pigs and Humans: 10,000 Years of Interaction, pp. 57–82. Oxford: Oxford University
Press.

Driesch, von den, A. (1976) A Guide to the Measurement of Animal Bones from
Archaeological Sites, Harvard: Peabody Museum.

Driver, J. C. (1982) ‘Medullary bone as an indicator of sex in bird remains from


archaeological sites’, in Wilson, B., Grigson, C., and Payne, S. (eds) Ageing and Sexing
Animal Bones from Archaeological Sites. BAR British Series 109, pp. 251–4. Oxford:
Archaeopress.

Ducos, P. (1968) L’origine des animaux domestiques en Palestine, Bordeux: Delmas.

Efremov, I. A. (1940) ‘Taphonomy: a new branch of paleontology’, Pan-American


Geologist, 74, 81–93.

Page 24 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Enloe, J. G., David, F., and Hare, T. S. (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, 105–24.

Evans, J. and O’Connor, T. (1999) Environmental Archaeology: Principles and Methods,


Stroud: Sutton Publishing.

Evershed, R. P., Payne, S., Sherratt, A. G., Copley, M. S., Coolidge, J., Urem-Kotsu, D.,
Kotsakis, K., Özdogan, M., Özdogan, A. E., Nieuwenhuyse, O., Akkermans, P. M. M. G.,
Bailey, D., Andeescu, R.-R., Campbell, S., Farid, S., Hodder, I., Yalman, N., Özbasaran, M.,
Bıçakcı, E., Garfinkel, Y., Levy, T., and Burton, M. M. (2008) ‘Earliest date for milk use in
the Near East and southeastern Europe linked to cattle herding’, Nature, 455, 528–31.

Ewbank, J. M., Phillipson, D. W., Whitehouse, R. D., and Higgs, E. S. (1964)


(p. 775)

‘Sheep in the Iron Age: a method of study’, Proceedings of the Prehistoric Society, 30,
423–6.

Fernández-Jalvo, Y. and Andrews, P. (2003) ‘Experimental effects of water abrasion on


bone fragments’, Journal of Taphonomy, 1(3), 147–63.

Flannery, K. V. (1969) ‘Origins and ecological effects of early domestication in Iran and
the Near East’, in Ucko, P. J. and Dimbleby, G. W. (eds) The Domestication and
Exploitation of Plants and Animals, pp. 73–100. London: Duckworth.

Forcina, G., Guerrini, M., van Grouw, H., Gupta, B. K., Panayides, P., Hadjigerou, P., Al-
Sheikhly, O. F., Awan, M. N., Khan, A. A., Zeder, M. A., and Barbanera, F. (2015) ‘Impacts
of biological globalization in the Mediterranean: unveiling the deep history of human-
mediated gamebird dispersal’, Proceedings of the National Academy of Sciences of the
United States of America, 112(11), 3296–301.

Forest, V. (1998) ‘De la hauteur au garrot des espèces domestiques en archéozoologie’,


Revue de Médecine Vétérinaire, 149(1), 55–60.

Forest, V. and Rodet-Belarbi, I. (1998) ‘Ostéométrie du métatarse des bovins en Gaule de


la conquête romaine à l’Antiquité Tardive’, Revue de Médecine Vétérinaire, 149(11),
1033–56.

Götherström, A., Anderung, C., Hellborg, L., Elburg, R., Smith, C., Bradley, D. J., and
Ellegren, H. (2005) ‘Cattle domestication in the Near East was followed by hybridisation
with aurochs bulls in Europe’, Proceedings: Biological Sciences, 272(1579), 2345–50.

Grant, A. (1982) ‘The use of tooth wear as a guide to the age of domestic ungulates’, in
Wilson, B., Grigson, C., and Payne, S. (eds) Ageing and Sexing Animal Bones from
Archaeological Sites. BAR British Series 109, pp. 91–108. Oxford: Archaeopress.

Page 25 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Grayson, D. K. (1984) Quantitative Zooarchaeology: Topics in the Analysis of


Archaeological Faunas, Orlando: Academic Press.

Greenfield, H. J. (2006) ‘Sexing fragmentary ungulate acetabulae’, in Ruscillo, D. (ed.)


Recent Advances in Ageing and Sexing Animal Bones: Proceedings of the 9th Conference
of the International Council of Archaeozoology, Durham, August 2002, pp. 68–86. Oxford:
Oxbow Books.

Grigson, C. (1982a) ‘Sex and age determination of some bones and teeth of domestic
cattle: a review of the literature’, in Wilson, B., Grigson, C., and Payne, S. (eds) Ageing
and Sexing Animal Bones from Archaeological Sites. BAR British Series 109, pp. 7–23.
Oxford: Archaeopress.

Grigson, C. (1982b) ‘Sexing Neolithic domestic cattle skulls and horncores’, in Wilson, B.,
Grigson, C., and Payne, S. (eds) Ageing and Sexing Animal Bones from Archaeological
Sites. BAR British Series 109, pp. 25–35. Oxford: Archaeopress.

Halstead, P. (2014) Two Oxen Ahead: Pre-Mechanized Farming in the Mediterranean,


Chichester: Wiley-Blackwell.

Harcourt, R. A. (1974) ‘The dog in prehistoric and early historic Britain’, Journal of
Archaeological Science, 1, 151–75.

Higgs, E. S. and Jarman, M. R. (1975) ‘Palaeoeconomy’, in Higgs, E. S. (ed.)


Palaeoeconomy, pp. 1–7. Cambridge: Cambridge University Press.

Higham, T. F. G. and Horn, P. L. (2000) ‘Seasonal dating using fish otoliths: results from
the Shag River Mouth site, New Zealand’, Journal of Archaeological Science, 27, 439–48.

Hill, M. E., Hill, M. G., and Widga, C. C. (2008) ‘Late Quaternary Bison diminution on the
Great Plains of North America: evaluating the role of human hunting versus climate
change’, Quaternary Science Reviews, 27, 1752–71.

Hogue, H. S. (2006) ‘Carbon isotope and microwear analysis of dog burials:


(p. 776)

evidence for maize agriculture at a small Mississippian site’, in Maltby, M. (ed.)


Integrating Zooarchaeology, pp. 125–32. Oxford: Oxbow Books.

Jackson, H. E. (1989) ‘The trouble with transformations: effects of sample size and sample
composition on meat weight estimates based on skeletal mass allometry’, Journal of
Archaeological Science, 16, 601–10.

Jans, M. M. E., Nielsen-Marsh, C. M., Smith, C. I., Collins, M. J., and Kars, H. (2004)
‘Characterisation of microbial attack on archaeological bone’, Journal of Archaeological
Science, 31, 87–95.

Johnstone, C. (2005) ‘Those elusive mules: investigating osteometric methods for their
identification’, in Mashkour, M. (ed.) Equids in Time and Space: Proceedings of the 9th

Page 26 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Conference of the International Council of Archaeozoology, Durham, August 2002, pp.


183–91. Oxford: Oxbow Books.

Jones, D. S. (1983) ‘Sclerochronology: reading the record of the molluscan shell’,


American Scientist, 71(4), 384–91.

Jones, P. R. (1980) ‘Experimental butchery with modern stone tools and its relevance for
Palaeolithic archaeology’, World Archaeology, 12(2), 153–65.

Jousse, H. and Escarguel, G. (2006) ‘The use of Holocene bovid fossils to infer
palaeoenvironment in Africa’, Quaternary Science Reviews, 25, 763–83.

Kahlke, R.-D. and Kaiser, T. M. (2011) ‘Generalism as a subsistence strategy: advantages


and limitations of the highly flexible feeding traits of Pleistocene Stephanorhinus
hundsheimensis (Rhinocerotidae, Mammalia)’, Quaternary Science Reviews, 30(17–18),
2250–61.

Kahlke, R.-D., García, N., Kostopoulos, D. S., Lacombat, F., Lister, A. M., Mazza, P. P. A.,
Spassov, N., and Titov, V. V. (2011) ‘Western Palaearctic palaeoenvironmental conditions
during the Early and early Middle Pleistocene inferred from large mammal communities,
and implications for hominin dispersal in Europe’, Quaternary Science Reviews, 30 (11–
12), 1368–95.

Lancelotti, C. and Madella, M. (2012) ‘The ‘invisible’ product: developing markers for
identifying dung in archaeological contexts’, Journal of Archaeological Science, 39, 953–
63.

Larson, G., Dobney, K., Albarella, U., Fang, M., Matisoo-Smith, E., Robins, J., Lowden, S.,
Finlayson, H., Brand, T., Willerslev, E., Rowley-Conwy, P., Andersson, L., and Cooper, A.
(2005) ‘Worldwide phylogeography of wild boar reveals multiple centres of pig
domestication’, Science, 307, 1618–21.

Lauwerier, R. C. G. M. (1988) Animals in Roman Times in the Dutch Eastern River Area,
Amersfoort: Rijksdienst voor het Oudheidkundig Bodemonderzoek.

Lefèvre, C. (1997) ‘Sea bird fowling in southern Patagonia: a contribution to


understanding the nomadic round of the Canoeros Indians’, International Journal of
Osteoarchaeology, 7, 260–70.

Libois, R. M. and Hallet-Libois, C. (1988) Éléments pour l’identification des restes


crâniens des poissons dulçaquicoles de Belgique et du nord de la France. 2—
Cypriniformes. Fiches d’ostéologie animale pour l’archéologie, Série A: Poissons. Juan-les-
Pins: APDCA.

Lloveras, Ll., Moreno-García, M., and Nadal, J. (2009) ‘Butchery, cooking and human
consumption marks on rabbit (Oryctolagus cuniculus) bones: an experimental study’,
Journal of Taphonomy, 7(2/3), 179–201.

Page 27 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Lupo, K. D. and Schmitt, D. N. (1997) ‘Experiments in bone boiling: nutritional returns


and archaeological reflections’, Anthropozoologica, 25/26, 137–44.

Lyman, R. L. (1994) Vertebrate Taphonomy, Cambridge: Cambridge University Press.

Lyman, R. L. (2008) Quantitative Palaeozoology, Cambridge: Cambridge University Press.

MacKinnon, M. (2010) ‘Cattle ‘breed’ variation and improvement in Roman Italy:


(p. 777)

connecting the zooarchaeological and ancient textual evidence’, World Archaeology,


42(1), 55–73.

Mainland, I. and Halstead, P. (2005) ‘The diet and management of domestic sheep and
goat at Neolithic Makriyalos’, in Davies, J., Fabiš, M., Mainland, I., Richards, M., and
Thomas, R. (eds) Diet and Health in Past Animal Populations: Current Research and
Future Directions. Proceedings of the 9th ICAZ Conference, Durham 2002, pp. 104–12.
Oxford: Oxbow Books.

Mainland, I. L. (1998) ‘The lamb’s last supper: the role of dental microwear analysis in
reconstructing livestock diet in the past’, Environmental Archaeology, 1, 55–62.

Mainland, I. L. (2003) ‘Dental microwear in grazing and browsing Gotland sheep (Ovis
aries) and its implications for dietary reconstruction’, Journal of Archaeological Science,
30, 1513–27.

Maldre, L. (2006) ‘What did the Bronze Age dogs eat? Coprolithic analyses’, Dogs and
People in Social, Working, Economic or Symbolic Interaction: Proceedings of the 9th
Conference of the International Council of Archaeozoology, Durham, August 2002, pp. 44–
8. Oxford: Oxbow Books.

Maltby, M. (2007) ‘Chop and change: specialist cattle carcass processing in Roman
Britain’, in Croxford, B., Ray, N., and Roth, R. (eds) TRAC 2006: Proceedings of the 16th
Annual Theoretical Roman Archaeology Conference, pp. 59–76. Oxford: Oxbow Books.

Martin, R. A. (2014) ‘A critique of vole clocks’, Quaternary Science Reviews, 94, 1–6.

Meadow, R. H. (1981) ‘Early animal domestication in South Asia: a first report of the
faunal remains from Mehrgarh, Pakistan’, in Härtel, H. (ed.) South Asian Archaeology
1979, pp. 143–79. Berlin: Dietrich Reimer Verlag.

Meadow, R. H. (1999) ‘The use of size index scaling techniques for research on
archaeozoological collections from the Middle East’, in Becker, C., Manhart, H., Peters, J.,
and Schibler, J. (eds) Historia Animalium ex Ossibus: Festschrift für Angela von den
Driesch, pp. 285–300. Rahden: Verlag Marie Leidorf GmbH.

Milne, C. and Crabtree, P. J. (2001) ‘Prostitutes, a rabbi, and a carpenter: dinner at the
Five Points in the 1830s’, Historical Archaeology, 35(3), 31–48.

Page 28 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Montuire, S. (1999) ‘Mammalian faunas as indicators of environmental and climatic


changes in Spain during the Pliocene-Quaternary transition’, Quaternary Research, 52,
129–37.

Mukherjee, A. J., Gibson, A. M., and Evershed, R. P. (2008) ‘Trends in pig product
processing at British Neolithic Grooved Ware sites traced through organic residues in
potsherds’, Journal of Archaeological Science, 35, 2059–73.

Munro, N. D., Bar-Oz, G., and Hill, A. C. (2011) ‘An exploration of character traits and
linear measurements for sexing mountain gazelle (Gazella gazella) skeletons’, Journal of
Archaeological Science, 38, 1253–65.

Nagaoka, L. (1994) ‘Differential recovery of Pacific island fish remains: evidence from the
Moturakau rockshelter, Aitutaki, Cook Islands’, Asian Perspectives, 33(1), 1–17.

Navas, E., Esquivel, J. A., and Molina, F. (2008) ‘Butchering patterns and spatial
distribution of faunal animal remains consumed at the Los Millares Chalcolithic
settlement (Santa Fe de Mondújar, Almería, Spain)’, Oxford Journal of Archaeology, 27(4),
325–39.

Nicholson, R. A. (1996) ‘Bone degradation, burial medium and species representation:


debunking the myths, an experiment-based approach’, Journal of Archaeological Science,
23, 513–33.

O’Connor, T. (1988) Bones from the General Accident Site, Tanner Row, York: Council for
British Archaeology.

O’Connor, T. (2000) The Archaeology of Animal Bones, Stroud: Sutton Publishing.

O’Connor, T. and Sykes, N. (eds) (2010) Extinctions and Invasions: A Social History of
British Fauna, Oxford: Oxbow Books.

Ottoni, C., Flink, L. G., Evin, A., Geörg, C., De Cupere, B., Van Neer, W.,
(p. 778)

Bartosiewicz, L., Linderholm, A., Barnett, R., Peters, J., Decorte, R., Waelkens, M.,
Vanderheyden, N., Ricaut, F.-X., Hoelzel, A. R., Mashkour, M., Karimlu, A. F. M., Sheikhi
Seno, S., Daujat, J., Brock, F., Pinhasi, R., Hongo, H., Perez-Enciso, M., Rasmussen, M.,
Frantz, L., Megens, H.-J., Crooijmans, R., Groenen, M., Arbuckle, B., Benecke, N.,
Viðarsdóttir, U. S., Burger, J., Cucchi, T., Dobney, K., and Larson, G. (2013) ‘Pig
domestication and human-mediated dispersal in western Eurasia revealed through
ancient DNA and geometric morphometrics’, Molecular Biology and Evolution, 30(4),
824–32.

Outram, A. K. (2001) ‘A new approach to identifying bone marrow and grease


exploitation: why the “indeterminate” fragments should not be ignored’, Journal of
Archaeological Science, 28, 401–10.

Page 29 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Payne, S. (1969) ‘A metrical distinction between sheep and goat metacarpals’, in Ucko, P.
J. and Dimbleby, G. W. (eds) The Domestication and Exploitation of Plants and Animals, pp.
296–305. London: Gerald Duckworth & Co.

Payne, S. (1973) ‘Kill-off patterns in sheep and goats: the mandibles from Aşvan Kale’,
Anatolian Studies, 23, 281–303.

Payne, S. (1975) ‘Partial recovery and sample bias’, in Clason, A. T. (ed.)


Archaeozoological Studies, pp. 7–17. New York: North Holland.

Payne, S. and Bull, G. (1988) ‘Components of variation in measurements of pig bones and
teeth, and the use of measurements to distinguish wild from domestic pig remains’,
Archaeozoologia, 2, 27–65.

Pike-Tay, A. and Cosgrove, R. (2002) ‘From reindeer to wallaby: recovering patterns of


seasonality, mobility, and prey selection in the Palaeolithic Old World’, Journal of
Archaeological Method and Theory, 9(2), 101–46.

Popkin, P. R. W., Baker, P., Worley, F., Payne, S., and Hammon, A. (2012) ‘The Sheep
Project (1): determining skeletal growth, timing of epiphyseal fusion and morphometric
variation in unimproved Shetland sheep of known age, sex, castration status and
nutrition’, Journal of Archaeological Science, 39, 1775–92.

Price, T. D., Schoeninger, M. J., and Armelagos, G. J. (1985) ‘Bone chemistry and past
behaviour: an overview’, Journal of Human Evolution, 14(5), 419–48.

Prummel, W. and Frisch, H.-J. (1986) ‘A guide for the distinction of species, sex and body
side in bones of sheep and goat’, Journal of Archaeological Science, 13, 567–77.

Rawlings, T. A. and Driver, J. C. (2010) ‘Paleodiet of domestic turkey, Shields Pueblo


(5MT3807), Colorado: isotopic analysis and its implications for care of a household
domesticate’, Journal of Archaeological Science, 37, 2433–41.

Reitz, E. J. and Wing, E. S. (1999) Zooarchaeology, Cambridge: Cambridge University


Press.

Rijkelijkhuizen, M. (2011) ‘Dutch Medieval bone and antler combs’, in Baron, J. and Kufel-
Diakowska (eds) Written in Bones: Studies on Technological and Social Contexts of Past
Faunal Skeletal Remains, pp. 197–206. Wrocław: Institute of Archaeology, University of
Wrocław.

Rivals, F. and Deniaux, B. (2005) ‘Investigation of human hunting seasonality through


dental microwear analysis of two Caprinae in late Pleistocene localities in southern
France’, Journal of Archaeological Science, 32, 1603–12.

Rowley-Conwy, P. (2001) ‘Determination of the season of death in European Wild Boar


(Sus scrofa ferus): a preliminary study’, in Millard, A. (ed.) Archaeological Sciences ’97:

Page 30 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Proceedings of the Conference Held at the University of Durham 2–4 September 1997.
BAR International Series 939. Oxford: Archaeopress.

Sadler, P. (1990) ‘The use of tarsometatarsi in sexing and ageing domestic fowl (Gallus
gallus L.), and recognising five toed breeds in archaeological material’, Circaea, 8(1), 41–
8.

Salque, M. (2012) ‘Was milk processed in these ceramic pots? Organic residue
(p. 779)

analyses of European prehistoric cooking vessels’, in Feulner, F., Doorn, N. L., and
Leonardi, M. (eds) May Contain Traces of Milk: Investigating the Role of Dairy Farming
and Milk Consumption in the European Neolithic, pp. 127–41. York: The University of
York.

Schmid, E. (1972) Atlas of Animal Bones for Prehistorians, Archaeologists and Quaternary
Geologists, Amsterdam: Elsevier.

Secor, D. H., Dean, J. M., and Campana, S. E. (eds) (1995) Recent Developments in Fish
Otolith Research, Columbia: University of South Carolina Press.

Seetah, K. (2006) ‘Multidisciplinary approach to Romano-British cattle butchery’, in


Maltby, M. (ed.) Integrating Zooarchaeology, pp. 111–18. Oxford: Oxbow Books.

Seetah, K. (2008) ‘Modern analogy, cultural theory and experimental replication: a


merging point at the cutting edge of archaeology’, World Archaeology, 40(1), 135–50.

Serjeantson, D. (1991) ‘ “Rid grasse of bones”: a taphonomic study of the bones from
midden deposits at the Neolithic and Bronze Age site of Runnymede, Surrey, England’,
International Journal of Osteoarchaeology, 1(2), 73–89.

Shipman, P., Foster, G., and Schoeninger, M. (1984) ‘Burnt bones and teeth: an
experimental study of color, morphology, crystal structure and shrinkage’, Journal of
Archaeological Science, 11, 307–25.

Silver, I. A. (1969) ‘The ageing of domestic animals’, in Brothwell, D. and Higgs, E. (eds)
Science in Archaeology, pp. 283–302. London: Thames and Hudson.

Simpson, G. G., Roe, A., and Lewontin, R. C. (1960) Quantitative Zoology, revised edn,
New York: Harcourt, Brace, and World.

Soderberg, J. (2004) ‘Wild cattle: red deer in the religious texts, iconography, and
archaeology of Early Medieval Ireland’, International Journal of Historical Archaeology,
8(3), 167–83.

Spangenberg, J. E., Jacomet, S., and Schibler, J. (2006) ‘Chemical analyses of organic
residues in archaeological pottery from Arbon Bleiche 3, Switzerland—evidence for
dairying in the late Neolithic’, Journal of Archaeological Science, 33, 1–13.

Page 31 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

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 Wilson, B., Grigson, C., and Payne, S. (eds) Ageing and Sexing Animal
Bones from Archaeological Sites. BAR British Series 109, pp. 109–26. Oxford:
Archaeopress.

Stevens, R. E., Jacobi, R., Street, M., Germonpré, M., Conard, N. J., Münzel, S. C., and
Hedges, R. E. M. (2008) ‘Nitrogen isotope analyses of reindeer (Rangifer tarandus),
45,000 BP to 9,000 BP: palaeoenvironmental reconstructions’, Palaeogeography,
Palaeoclimatology, Palaeoecology, 262, 32–45.

Stewart, F. L. and Stahl, P. W. (1977) ‘Cautionary note on edible meat poundage figures’,
American Antiquity, 42(2), 267–70.

Tappen, M. (1994) ‘Bone weathering in the tropical rain forest’, Journal of Archaeological
Science, 21, 667–73.

Teichert, M. (1984) ‘Size variation in cattle from Germania Romana & Germania Libera’,
in Grigson, C. and Clutton-Brock, J. (eds) Animals and Archaeology: 4. Husbandry in
Europe. BAR International Series 227, pp. 93–103. Oxford: Archaeopress.

Telldahl, Y., Svensson, E. M., Götherström, A., and Storå, J. (2012) ‘Osteometric and
molecular sexing of cattle metapodia’, Journal of Archaeological Science, 39, 121–7.

Turner-Walker, G. and Jans, M. (2008) ‘Reconstructing taphonomic histories using


histological analysis’, Palaeogeography, Palaeoclimatology, Palaeoecology, 266, 227–35.

Uchiyama, J. (1999) ‘Seasonality and age structure in an archaeological assemblage of


Sika deer (Cervus nippon)’, International Journal of Osteoarchaeology, 9, 209–18.

Uerpmann, H.-P. (1979) Probleme der Neolithisierung des Mittelmeerraums,


(p. 780)

Wiesbaden: Dr. Ludwig Reicher Verlag.

Vacca, B. (2008) ‘Ambrosetti 1 (Sesto Fiorentino, Florence): an experimental study of


burning damage of the faunal remains’, in Baioni, M., Leonini, V., Lo Vetro, D., Martini, F.,
Poggiani Keller, R., and Sarti, L. (eds) Bell Beaker in Everyday Life: Proceedings of the
10th Meeting ‘Archéologie et Gobelets’ (Florence—Siena—Villanuova sul Clisi, May 12–
15, 2006). Millenni, Studi di Archeologia Preistorica 6, pp. 353–6. Firenze: Museo
Fiorentino di Preistoria «Paolo Graziosi».

Vanpoucke, S., Mainland, I., De Cupere, B., and Waelkens, M. (2009) ‘Dental microwear
study of pigs from the classical site of Sagalassos (SW Turkey) as an aid for the
reconstruction of husbandry practices in ancient times’, Environmental Archaeology,
14(2), 137–54.

Page 32 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Vickers, K., Erlendsson, E., Church, M. J., Edwards, K. J., and Bending, J. (2011) ‘1000
years of environmental change and human impact at Stóra-Mörk, southern Iceland: a
multiproxy study of a dynamic and vulnerable landscape’, The Holocene, 21(6), 979–95.

Vigne, J.-D. (1988) Les mammifères post-glaciaires de Corse: étude archéozoologique.


«Gallia Préhistoire» supplement 26. Paris: Éditions du CNRS.

Vigne, J.-D. (1991) ‘The meat and offal weight (MOW) method and the relative proportion
of ovicaprines in some ancient meat diets of the north-western Mediterranean’, Rivista di
Studi Liguri, 57(1–4), 21–47.

Viner, S., Evans, J., Albarella, U., and Parker Pearson, M. (2010) ‘Cattle mobility in
prehistoric Britain: strontium isotope analysis of cattle teeth from Durrington Walls
(Wiltshire, Britain)’, Journal of Archaeological Science, 37, 2812–20.

Wapnish, P. (1984) ‘The dromedary and Bactrian camel in Levantine historical settings:
the evidence from Tell Jemmeh’, in Clutton-Brock, J. and Grigson, C. (eds) Animals and
Archaeology: 3. Early Herders and Their Flocks. BAR International Series 202, pp. 171–
200. Oxford: Archaeopress.

Watson, J. P. N. (1979) ‘The estimation of the relative frequencies of mammalian species:


Khirokitia 1972’, Journal of Archaeological Science, 6, 127–37.

Weiner, S. (2010) Microarchaeology: Beyond the Visible Archaeological Record,


Cambridge: Cambridge University Press.

West, N. (1982) ‘Spur development: recognizing caponized fowl in archaeological


material’, in Wilson, B., Grigson, C., and Payne, S. (eds) Ageing and Sexing Animal Bones
from Archaeological Sites. BAR British Series 109, pp. 255–61. Oxford: Archaeopress.

Wijngaarden-Bakker, van, L. H. (1987) ‘Experimental zooarchaeology’, in Groenman-van


Waateringe, W. and van Wijngaarden-Bakker, L. H. (eds) Farm Life in a Carolingian
Village: A Model Based on Botanical and Zoological Data from an Excavated Site. Studies
in Prae- en Protohistorie 1, pp. 107–17. Assen: Van Gorcum.

Wilkinson, K. and Stevens, C. (2008) Environmental Archaeology: Approaches, Techniques


and Applications, 2nd edn, Stroud: Tempus Publishing.

Zeder, M. A. (2006) ‘A critical assessment of markers of initial domestication in goats


(Capra hircus)’, in Zeder, M. A., Bradley, D. G., Emshwiller, E., and Smith, B. D. (eds)
Documenting Domestication: New Genetic and Archaeological Paradigms, pp. 181–208.
Berkeley: University of California Press.

Zohar, I. and Belmaker, M. (2005) ‘Size does matter: methodological comments on sieve
size and species richness in fishbone assemblages’, Journal of Archaeological Science, 32,
635–41.

Page 33 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018


A Glossary of Zooarchaeological Methods

Page 34 of 34

PRINTED FROM OXFORD HANDBOOKS ONLINE (www.oxfordhandbooks.com). (c) Oxford University Press, 2015. All Rights
Reserved. Under the terms of the licence agreement, an individual user may print out a PDF of a single chapter of a title in
Oxford Handbooks Online for personal use (for details see Privacy Policy).

Subscriber: University of Sheffield; date: 05 January 2018

View publication stats

You might also like