SlideShare a Scribd company logo
1 of 91
Journal of Archaeological Science 32 (2005) 1408e1416
http://www.elsevier.com/locate/jas
Why domesticate food animals? Some zoo-archaeological
evidence from the Levant
Simon J.M. Davis *
Instituto Português de Arqueologia, Avenida da Índia 136, P-
1300-300 Lisbon, Portugal
Received 4 February 2005; received in revised form 10 March
2005
Abstract
Zoo-archaeological remains from the southern Levant indicate
two shifts in the pattern of animal exploitation from
Palaeolithic
to Pre-Pottery Neolithic times. These shifts were especially
marked towards the end of this time span. One is the increased
consumption of small animals and the other shift is an increased
hunting of juvenile gazelles compared to adults. Both are
interpreted in terms of an increased intensity of exploitation of
environmental resources due, it is suggested, to population
increase,
which subsequently forced people to husband animals.
� 2005 Elsevier Ltd. All rights reserved.
Keywords: Domestication; Population; Neolithic; Demography;
Levant; Near East
‘‘.the true cause that set in motion the great tide of
northern emigration, and that continued to propel it till
it rolled at different periods against China, Persia, Italy,
and even Egypt, was a scarcity of food, a population
extended beyond the means of supporting it.’’ [33]
1. Introduction
Why did our ancestors domesticate food animals?
For a long time, it was believed that the transition from
hunting to husbandry was a move from a precarious
existence to one providing greater security. The notion
that hunting requires high expenditure of energy and
that a major saving of effort can be gained by switching
to farming is probably wrong. Pioneering work by
* Tel.: C351 21 361 6557; fax: C351 01 361 6559.
E-mail address: [email protected]
0305-4403/$ - see front matter � 2005 Elsevier Ltd. All rights
reserved.
doi:10.1016/j.jas.2005.03.018
Richard Lee [30], who studied modern hunteregatherers
in southern Africa, discovered that they enjoy a plentiful
and balanced diet and spend a mere 2e3 days per week
in their quest for food. Husbanding animals is, it would
seem, more arduous than simply going out and hunting
them. The view taken here is a more gradualist one in
which a slow shift in the balance between people and
their source of food has occurred. Clearly, in order to
try and understand the background to domestication
in the archaeological record, we need to clarify what
happened during the millennia that preceded this
change. One of the first studies that attempted to answer
the question why people began domesticating plants and
animals was Mark Cohen’s The Food Crisis in Prehistory
[6]. Cohen suggested that the period prior to domes-
tication was characterized by increased strain on the
environment due to rising demographic pressure.
Amongst the evidence he cited is an increase in the
occurrences of pathological conditions in human skel-
etal remains from archaeological sites e probably
reflecting increasingly poor nutrition. He suggested that
a rise in the human population at that time ‘‘forced’’
mailto:[email protected]
http://www.elseiver.com/locate/jas
1409S.J.M. Davis / Journal of Archaeological Science 32 (2005)
1408e1416
people to change their relationship with the natural
world and to assume a greater degree of control over it.
Others too have linked domestication indirectly
or directly to demographic pressure and the carrying
capacity of the environment [21,43], and these factors
combined with a trigger e deteriorating climate at the
end of the last Ice Age [17]. But not all agree that
population growth was a factor (see, for example, [41]),
indeed it is quite likely that the causes varied depending
upon which part of the world is considered. Hayden [25]
has suggested a ‘‘socio-economic’’ explanation for the
origin of domestication; domesticates were initially
used as prestige commodities. These served to ‘‘enhance
individual or corporate group power’’ in complex
hunteregatherer communities. Others like Isaac [27]
and Cauvin [5] have sought a more religious/psycholog-
ical explanation for the beginnings of domestication.
Before addressing the question of why animals were
domesticated, we need to consider briefly where and
when animal husbandry began. We now know of some
nine independent areas where food production began.
The ‘‘Fertile Crescent’’ of the Near East was un-
doubtedly the earliest Old World hearth of domesti-
cation of both animals and plants, although it is still
unclear exactly where within this area it began
[2,17,23,24,49,50]. Evidence from Cyprus and the
southern Levant now indicates that bovids (cattle, sheep
and goats) and pigs were first domesticated towards the
end of the 9th millennium cal BC in the northern part of
the Near East, and that during the 8th millennium BC,
these animals were domesticated (or introduced as
domestic stock) in the southern Levant [8,22]. Clearly,
the 9th and 8th millennia, during the so-called Pre-
Pottery Neolithic, were crucial times for the advent of
husbandry; hence, we need to examine the relations
between people and animals in the period before those
millennia in order to answer why people began
husbanding animals.
A possible zoo-archaeological answer to this major
question first became evident to me when studying the
fauna from the Natufian e Aceramic Neolithic site of
Hatoula, in Israel, in the mid-1980s. It is especially
gratifying to note that the last two decades have seen
other zoo-archaeologists interpret their data in a similar
way e measurements of bones and shells, and consid-
erations of species’ frequencies can reflect pressure on
the environment in turn linked to demographic pressure.
Most noteworthy for the Levant is the new study of the
fauna from Hayonim and nearby sites by Natalie
Munro [36,37]. She too suggests that population in-
creased quite markedly in the Natufian. This paper, like
earlier reports [11,12,14,15], considers other studies, as
well as several from other regions. They have been and
still are interpreted in a similar way. By including the
faunal evidence from subsequent periods I hope to
present a broader view of this major event.
I shall present two lines of evidence from the zoo-
archaeological record of the Mousterian to the Pre-
Pottery Neolithic in the Levant which corroborate the
demographic pressure thesis. The first is a shift of
people’s prey in the course of time from big game to
smaller mammals and ultimately fish and birds e which
Flannery [19] recognised as a shift from a narrow
spectrum of environmental resources to a broader one.
Indeed, this ‘‘spectrum shift’’ is well attested in many
regions where people ‘‘were forced to become even more
eclectic in their food gathering, to eat more and more
unpalatable foods, and in particular to concentrate on
foods of low trophic level and high density’’ [6]. The
second line of evidence has not, to my knowledge, been
considered very much by zoo-archaeologists and is an
increase in the proportion of juvenile prey e here gazelle
e in these faunal assemblages. This I term the ‘‘age
shift’’. (Legge [31] had observed an increase in numbers
of juvenile gazelles at Mount Carmel. He suggested that
this may signify their domestic status, a proposal he
subsequently rejected; pers. comm. The gazelle is not
generally considered to be amenable to domestication.)
Both these changes occurred over a long time span from
the Mousterian, some 50,000 years ago, to the Pre-
Pottery Neolithic A of the 9th millennium BC.
Moreover, these changes were even more rapid during
the two or three millennia before domestic animals first
made their appearance in the 8th millennium BC [11,15].
And, as I shall argue, the interpretation of these changes
helps to explain why people had, to a large extent, to
abandon hunting in favour of husbandry.
2. Material and methods
This article discusses two kinds of zoo-archaeological
data from the Upper Pleistocene and early Holocene.
Most come from Israel. They are counts of:
(a) Bones of different species of animals and
(b) Numbers of young versus adult gazelles.
The data presented here all derive from sites whose
faunal assemblages were recovered by sieving (Fig. 1).
Hence, biases against smaller bones such as unfused
epiphyses that belonged to young animals are probably
less serious than is often the case with hand-collected
material [38]. Another problem sometimes encountered
in zoo-archaeology is observer variation. Zoo-archae-
ologists may count bones and deduce the age-at-death of
animals in different ways. One person studied all the
assemblages discussed here using the same methods [15].
In brief, mandibles and a restricted suite of ‘‘parts of the
skeleton always recorded ’’ were counted. They consist of
a predetermined suite of articular ends/epiphyses of
girdle, limb and foot bones. This in part explains why the
1410 S.J.M. Davis / Journal of Archaeological Science 32
(2005) 1408e1416
Fig. 1. Map of the southern Levant to show places mentioned in
the text. The dotted line represents the present-day 200 mm
isohyet.
sample sizes in many cases appear to be small. Since birds
no longer have teeth and fish lack limbs, and in order to
compare frequencies of mammals with those of birds and
fish at one of the key sites considered here (Hatoula), the
numbers of mammalian mandibles were compared with
the minimum numbers of individuals of fish and birds.
(This is admittedly not entirely satisfactory and has
probably led to an under-representation of fish.)
The growing end of a mammal limb-bone is its
epiphysis, which, at maturity, fuses to its respective
shaft. For large assemblages of a particular species, an
approximate estimate of the proportion of juveniles
(that is osteologically immature) culled may be calcu-
lated by counting separately the unfused and fused
epiphyses. A similar approach was adopted for man-
dibles. Premolar teeth replace their milk predecessors at
a given age. In the case of the gazelle this occurs around
12e13 months, which is also the age when the third
molar tooth erupts, and this is also very approximately
the age when many of the limb-bones fuse [7]. Hence, it
is possible to calculate the proportion of juveniles via the
numbers of mandibles with milk teeth or unerupted
third molars.
3. The ‘‘Spectrum Shift’’ (Figs. 2 and 3)
At the sequence of open-air sites, which extends from
the Mousterian on the Golan Heights (Biqat Quneitra)
to the late Natufian on the eastern side of the Sea of
Galilee (Nahal Ein Gev II), we [14] observed a change
from big game to gazelles, hares and foxes. Many of the
animals hunted at Biqat Quneitra were very large and
included species like rhinoceros, aurochs, red deer and
horses. As time progressed, remains of these very large
beasts became rare on archaeological sites and fallow
deer and gazelle were clearly the most important animals
exploited in Aurignacian and Kebaran times (i.e.
1411S.J.M. Davis / Journal of Archaeological Science 32 (2005)
1408e1416
Fig. 2. The spectrum shift as seen in the series of open-air sites
from the Mousterian at Biqat Quneitra to the Natufian at Nahal
Ein Gev. Note the
gradual shift with time from large game to gazelles. Key: BQ,
Biqat Quneitra (Mousterian c. 54,000 bp); NEG I, Nahal Ein
Gev I (Late Upper
Palaeolithic c. 25e20,000 bp); EG I, Ein Gev I (Kebaran c.
16,000 bp); EG III, Ein Gev III (Geometric Kebaran c.
16,500e14,500 bp); EG IV, Ein
Gev IV (Late Geometric Kebaran c. 14,500e12,750 bp); NEG II,
Nahal Ein Gev II (Late Natufian/Khiamian c. 10,500e10,000
bp). The numbers in
parentheses are the counts of identified and recorded bones.
between approximately 30,000 and 15,000 BC). Fallow
deer appear to have become even scarcer as we approach
the late Natufian when the most common medium-sized
mammal represented is gazelle. At about this time,
during the Epipalaeolithic, increasing numbers of small
mammals such as fox and hare are also represented. Let
us look now at what happened at a site in another part
of this region e Hatoula, in central Israel e with its
1412 S.J.M. Davis / Journal of Archaeological Science 32
(2005) 1408e1416
Fig. 3. The spectrum shift as seen in the NatufianePPNA
(Khiamian and Sultanian) sequence at Hatoula. Percentages of
the different groups of
animals, computed from mandible counts for the mammals, and
minimum numbers of individuals for the fish and birds. Small
mammals include
hedgehog, hare, polecat, marten, badger, fox and cat only. Note
the increase of small mammals, fish and birds in the PPNA.
sequence from Natufian to PPNA (the latter subdivided
into Khiamian and Sultanian levels). Hatoula is located
on the western slopes of the Judean Hills roughly
midway between Jerusalem and Tel Aviv. The samples
of animal bones recovered at this site are quite large,
with 2540 bones recorded in the Natufian levels, and
1016 in the PPNA levels. Here the continued ‘‘down-
sizing’’ of animals represented is quite remarkable. Note
the shift between the Natufian and PPNA levels, from
gazelle to small mammals, fish and birds. It is tempting
to regard this shift at Hatoula as the culmination of
a long and more gradual move to smaller taxa, which
had been happening since Mousterian times.
Whether this shift reflects local habitat degradation
or a change in hunting techniques (or both) is difficult to
determine. However, it is worth noting that although
Fig. 4. The age shift of gazelles culled in Israel between
Mousterian and PPN times. This is a plot of the percentages of
unfused limb-bone epiphyses
(distal radius, distal metacarpal, distal metatarsal, distal femur,
distal tibia, and calcaneum-tuber calcis) plotted against time.
Total includes these
unfused epiphyses plus specimens of the same bone-parts with
fused epiphyses. Data come from the following sites, from left
to right: Mousterian:
Kebara cave; Mousterian: Hayonim cave level E; Upper
Palaeolithic: Kebara cave; Aurignacian: Hayonim cave level D;
Kebaran: Ein Gev I;
Kebaran: Hayonim cave level C; Natufian: Hayonim cave level
B; Natufian: Hayonim terrace; Natufian: Hatoula; Khiamian
(PPNA): Hatoula;
Sultanian (PPNA): Hatoula. Data for Hatoula are in refs.
[10,15]. Data for other sites are in ref. [9].
1413S.J.M. Davis / Journal of Archaeological Science 32 (2005)
1408e1416
increasingly scarce in zoo-archaeological assemblages,
the continuing presence of many very large animals in
the Levant well after their apparent demise at Ein Gev
makes it seem unlikely that the environment changed to
any very great extent. For example, rhinoceros and red
deer are known in the Kebaran (at Hayonim), hippo-
potamus is known in the Iron Age (at Tel Qassile), and
the aurochs survived into the Early Bronze Age (at Tel
Yarmouth; [8,13]). A shift in emphasis from very large
animals to small mammals, birds and fish must have
required the invention and development of new hunting
techniques, but this was surely brought about by the
increasing need to rely upon such smaller sources of
sustenance.
Cohen [6], like Flannery [19], has amassed a wide
range of evidence, which shows that in many regions just
before the advent of animal and plant husbandry people
began to expand their resource base. This included
many smaller species of animals whose capture required
a higher expenditure of energy for a given unit of flesh
procured. On the limestone steppe in Jordan, Martin
[34] studied a succession of small open-air sites, where
she noted that hare increased dramatically in the early
and middle PPNB levels (at Jilat). In the Jordan valley,
Tchernov [48] too reported large quantities of hare and
bird bones at the PPNA site of Netiv Hagdud.
Elsewhere in the Mediterranean, others have observed a
shift to smaller species of animals. For example at the
caves of Franchthi in Greece and Nerja in Andalusia,
Payne [39] and Morales et al. [35] found that fishing
began in the Mesolithic and Magdalenian periods,
respectively. Morales et al. [35] call this shift to ex-
ploiting marine resources the ‘‘Tardiglacial paradigm’’.
The increased exploitation of marine resources and
occurrences of shell middens in the archaeological
record may to some extent be due to the encroachment
of the seashore as sea levels rose after the Pleistocene [1].
Besides small game, birds and marine resources, Lubell
[32] has recently pointed out that remains of land snails
too become more abundant just prior to the advent of
agriculture in the circum Mediterranean region. The
Table 1
The age-at-death of gazelles from late PleistoceneeHolocene
Israel
deduced from tooth eruption
Period Site Mandibles
with M3
‘‘Unerupted’’
Mandibles
with M3
‘‘Erupted’’
%
Juvenile
PPNA Hatoula 14 27 34
Late Natufian Hatoula 8 41 16
Natufian Hayonim terrace 27 135 17
Aurignacian Hayonim cave D 13 115 10
Mousterian Kebara cave 7 251 3
Numbers of mandibles with unerupted third molars (i.e. less
than c. 13
months of age) and erupted M3s from various large zoo-
archaeological
assemblages (see refs. [9,15]).
most spectacular examples are the Capsian ‘‘escar-
gotières’’ of Algeria and Tunisia. He suggests ‘‘the
presence of abundant land snails represents part of
a signature for the broad spectrum revolution’’. Stiner
[44] has provided further evidence for a broadening of
the spectrum of exploited species in the Mediterranean
basin. Presumably big game had become scarce, but
why? One explanation, and the one offered here, is
simply that there were now many more mouths to feed;
people were forced to exploit small game, fowl and fish
in order to survive.
4. The ‘‘Age Shift’’ (Fig. 4 and Table 1)
The numbers of unfused (juvenile) limb-bone epiph-
yses compared to the fused (adult) ones in the sequence
of Israeli zoo-archaeological assemblages indicate that
some 20e25% of the gazelles hunted before the Neo-
lithic were osteologically immature. Subsequently, in the
PPNA, this percentage increased markedly to around
35e40%. The counts of mandibles with unerupted and
erupted third molars indicate a similar trend in the
course of time (Table 1). At Hatoula, the presence of
milk or permanent teeth in the anterior part of the
gazelle mandibles shows that the proportion of juveniles
increased from 39% in the Natufian (13 have dP4s and
20 have P4s) to 59% in the PPNA (19 have dP4s but only
13 have P4s). In sum then, both limb-bones and
mandibles indicate a substantial chronological trend
towards increased culling of juvenile gazelles.
There are several interesting parallel examples from
zoo-archaeological successions elsewhere. Perhaps of
some relevance here, and certainly the inspiration for my
interpretation of the Israeli data, is Elder’s [18] study of
deer mandibles from three ‘prehistoric’ (i.e. pre-Euro-
pean settlement) and two ‘historic’ Indian sites in
Missouri, USA. At his ‘prehistoric’ sites, the age
distributions of the culled deer showed moderate
numbers of old and senile animals as well as young.
Elder suggested that this pattern is similar to that
observed in a stable population of ungulates. However,
most of the mandibles from the two ‘historic’ or ‘post-
contact’ sites dated between AD 1725 and AD 1780,
derived from young individuals, with very few old and
senile ones represented. This kind of age distribution is
typical of a population undergoing rapid turnover
caused by heavy predation. Elder suggested that
a lucrative venison trade and more efficient hunting
with firearms and horses (both the result of the arrival of
European settlers) were the two main factors which
caused this change in the age-profiles of Missouri deer.
Three other studies are worth quoting. Between Middle
and Late Stone Ages in South Africa, the limpets
collected by people became smaller (i.e. younger), which
Klein and colleagues [28,29] correlates with increasingly
1414 S.J.M. Davis / Journal of Archaeological Science 32
(2005) 1408e1416
heavy foraging. In northern Israel, Stiner et al. [45]
recorded a size reduction of tortoises’ limb-bone shafts
(the girth of the shaft in reptile bones continues to
increase with age) in the Epipalaeolithic at Hayonim,
western Galilee, which they suggest, was due to an in-
crease in the predation pressure on tortoises, in turn
a reflection of human population increase. Surovell [46]
simulated what can happen to fast reproducing species
of small game like hares and partridges, and slow ones
like the tortoise, when predation pressure rises due to
increasing occupation density and population. He found
that the faster reproducing species will increase in
number in the archaeological record while slower
reproducers will decrease and even become extinct.
5. Conclusion
What then do the changes in the zoo-archaeological
record of the Levant mean? My interpretation of the
gazelle age shift is similar to that which Elder proposed
to explain his age shift in Missouri deer. What we
observe in the Israeli sequence is a gradual increase in
the intensity with which gazelles were hunted. This
hunting intensity increased at an even greater rate in the
PPNA as can be seen at Hatoula, and appears to have
occurred alongside a shift towards exploiting smaller
animals, and in the Natufian and PPNA, more extensive
exploitation of marine and avian resources. I suggest
that it was the same factor, an increasing imbalance
between predator, that is man-the-hunter, and prey,
which was responsible for both of these trends. Natalie
Munro has restudied the Natufian fauna of Hayonim
and two other more recently excavated Natufian sites.
It is gratifying to note that she comes to similar
conclusions to the ones suggested here. In concluding
her thesis, Munro [36] writes ‘‘.in the Early Natufian
phase, the southern Levant likely supported the densest
and.probably the highest gross population sizes the
region had seen to this point.’’
While it is possible to view population levels as being
a result of food resource availability, Cohen [6] and
Rowley-Conwy [42] see population levels as the driving
force (at least in the Neolithic) behind the need to
change from hunting to husbanding. In my opinion it
was this factor e the ratio of human beings to wildlife e
that caused Neolithic people to adopt husbandry. The
human population of the Levant had undoubtedly been
rising steadily since the Palaeolithic. With the adoption
of sedentism in the Natufian as first suggested by Perrot
[40], population increase must have become much more
rapid. We know from modern examples of newly settled
nomads, that sedentism leads to rapid population
increase [47]. Goring-Morris’s [20] archaeological survey
of the Negev desert has shown a decrease in the number
of PPNA sites there as compared with the Natufian. He
attributes this ‘‘virtual abandonment’’ of the Negev to
aridification at that time, and suggests that people
moved northwards into the more fertile Mediterranean
zone which may have exacerbated an already precarious
balance between people and their resources.
Henry [26] cites archaeological data from that region
such as site size and artefact densities that point to
possible Natufian population increase. In their studies
on the Natufian and PPNA, Belfer-Cohen and Bar-
Yosef [4] note certain features on human skeletal size
such as a reduction in the difference in male/female
stature, as well as increasing site size that point to stress
and demographic pressure at this time e just prior to the
adoption of animal husbandry. Population density may
have increased from 0.25 persons per km
2
in the
Natufian to 1 person per km
2
in the PPNA and then 4
persons per km
2
in the PPNB [3].
In the southern Levant, probably during the PPNB,
human population levels continued to rise rapidly and
reached some critical threshold. The carrying capacity of
the environment was exceeded e the ever-increasing
number of human mouths could no longer be adequately
fed. Gazelle and big-game stocks had become too scarce
and fishing and fowling did not suffice either. It was then,
during the PPNB, that people had to exert greater
control over environmental resources and begin husband-
ing those species that were amenable to this kind of
treatment such as cattle, sheep, goats and pigs e
a strategy in which a given area can sustain many more
herders and farmers than hunteregatherers. (The switch
to husbanding bovids and pigs in the course of the PPNB
raises another question. Were these species, all present
but admittedly not very common on Natufian and PPNA
sites in the Levant, domesticated locally or introduced
from other parts of the Near East? Given the imprecision
of dates available it is probably still too early to answer
this question with any degree of security. Their in-
troduction from beyond the Levant as already managed
stock is of course a strong possibility, and could help
explain why sheep and goat were not hunted in increasing
numbers in the Natufian and PPNA.) Perhaps it was the
climate changes of the end of the Pleistocene, as
Diamond [17] suggests, that acted as the last straw that
‘‘broke the camel’s back’’.
People could no longer survive by hunting alone e
they were now banished from the Garden of Eden:
animal husbandry is hard work and domesticated
animals require daily care and supervision. But with
these four-legged resources under human control, people
were able to become ever more ‘‘fruitful and multiply’’.
The human population in the Near East rose even more
rapidly. The stage was now set for further economic
change, the spread of farming peoples across the world
and all kinds of developments both welcome and
unwelcome [16].
1415S.J.M. Davis / Journal of Archaeological Science 32 (2005)
1408e1416
Acknowledgements
I am very grateful to Ofer Bar Yosef, the late Eitan
Tchernov, Naama Goren-Inbar, Donald Henry, Jean
Perrot, Monique Lechevallier and Avraham Ronen
who at various times have invited me to study zoo-
archaeological remains from their sites. It was a study
of the fauna of Hatoula that led me to the ideas put
forward here. I have had numerous useful discussions
with colleagues both in Israel and England, and both
Cathy Douzil and Umberto Albarella have encouraged
me to write this paper and offered useful comments. I
thank Nigel Goring-Morris, Anna Belfer-Cohen and
Ofer Bar Yosef for answering various queries and
Juliet Clutton-Brock, who first drew my attention to
William Elder’s study of deer mandibles in Missouri. I
am grateful to Arturo Morales as well as several
anonymous referees who also read and commented
upon an earlier version of this text.
References
[1] G.N. Bailey, A.S. Craighead, Late Pleistocene and Holocene
coastal palaeoeconomies: a reconsideration of the molluscan
evidence from northern Spain, Geoarchaeology: An
International
Journal 18 (2003) 175e204.
[2] O. Bar-Yosef, Earliest food producers e pre Pottery
Neolithic
(8000e5500), in: T.E. Levy (Ed.), The Archaeology of Society
in the Holy Land, Leicester University Press, London, 1995,
pp. 190e204.
[3] O. Bar-Yosef, R. Meadow, The origin of agriculture in the
Near
East, in: A.B. Gebauer, D. Price (Eds.), Last Hunters, First
Farmers, School of American Research, Santa Fe, 1995, pp.
39e94.
[4] A. Belfer-Cohen, O. Bar-Yosef, Early sedentism in the Near
East:
a bumpy ride to village life, in: I. Kuijt (Ed.), Life in Neolithic
Farming Communities: Social Organization, Identity, and
Differ-
entiation, Kluwer Academic/Plenum, New York, 2000, pp.
19e37.
[5] J. Cauvin, Naissance des divinités Naissance de
l’agriculture,
Flammarion, Paris, 1998.
[6] M.N. Cohen, The Food Crisis in Prehistory, Yale University
Press, New Haven, 1977.
[7] S.J.M. Davis, A note on the dental and skeletal ontogeny of
Gazella, Israel Journal of Zoology 29 (1980) 129e134.
[8] S.J.M. Davis, Climatic change and the advent of
domestication:
the succession of ruminant artiodactyls in the late Pleistocenee
Holocene in the Israel region, Paléorient 8 (1982) 5e15.
[9] S.J.M. Davis, The age profiles of gazelles predated by
ancient man
in Israel: possible evidence for a shift from seasonality to
sedentism in the Natufian, Paléorient 9 (1983) 55e62.
[10] S.J.M. Davis, A preliminary report of the fauna from
Hatoula:
a NatufianeKhiamian (PPNA) site near Latroun, Israel, in:
M. Lechevallier, A. Ronen (Eds.), Le site NatoufieneKhiamien
de Hatoula prés de Latroun, Israël; fouilles 1980e1982 rapport
préliminaire, vol. 1, Centre de Recherches français de
Jérusalem,
1985, pp. 71e98.
[11] S.J.M. Davis, Why did prehistoric people domesticate food
animals? in: O. BarYosef, B. Vandermeersch (Eds.), Investiga-
tions in South Levantine Prehistory, British Archaeological
Reports International Series, vol. 497, pp. 43e59 (Oxford).
[12] S.J.M. Davis, When and why did prehistoric people
domesticate
animals? Some evidence from Israel and Cyprus, in: O. Bar-
Yosef,
F.R. Valla (Eds.), The Natufian Culture in the Levant,
Michigan,
InternationalMonographsinPrehistory,pp.381e390(AnnArbor).
[13] S.J.M.Davis, H. Hemmer, More animal bones from Tel
Yarmouth:
an Early Bronze Age site in Israel, in: P. de Miroschedji (Ed.),
Tel Yarmouth II.
[14] S.J.M. Davis, R. Rabinovich, N. Goren-Inbar, Quaternary
extinctions and population increase in western Asia: the animal
remains from Biq’at Quneitra, Paléorient 14 (1988) 95e105.
[15] S.J.M. Davis, O. Lernau, J. Pichon, The animal remains:
new
light on the origin of animal husbandry, in: M. Lechevallier,
A. Ronen (Eds.), Le gisement de Hatoula en Judée occidentale,
Israël, Mémoires et Travaux du Centre de Recherche français de
Jérusalem 8, Association Paléorient, Paris, 1994, pp. 83e100.
[16] J. Diamond, The Rise and Fall of the Third Chimpanzee,
Radius,
London, 1991.
[17] J. Diamond, Evolution, consequences and future of plant
and
animal domestication, Nature 418 (2002) 700e707.
[18] W.H. Elder, Primaeval deer hunting pressures revealed by
remains from American Indian middens, Journal of Wildlife
Management 29 (1965) 366e370.
[19] K. Flannery, Origins and ecological effects of early
domestication
in Iran and the Near East, in: P.J. Ucko, G.W. Dimbleby (Eds.),
The Domestication and Exploitation of Plants and Animals,
Duckworth, London, 1969, pp. 73e100.
[20] N.A. Goring-Morris, Developments in terminal Pleistocene
hunteregatherer socio-cultural systems: a perspective from the
Negev and Sinai deserts, in: I. Hershkowitz (Ed.), Proceedings
of
the Second Symposium on the Upper Palaeolithic, Mesolithic,
and Neolithic Populations around the Mediterranean Basin,
BAR
International Series, vol. 508, pp. 7e28 (Oxford).
[21] L. Groube, The impact of diseases upon the emergence of
agriculture, in: D.R. Harris (Ed.), The Origins and Spread of
Agriculture and Pastoralism in Eurasia, University College
Press,
London, 1996, pp. 101e129.
[22] J. Guilaine, F. Briois, J.-D. Vigne, I. Cerrère, Découverte
d’un
Néolithique précéramique ancien chypriote (fin 9
e
, début 8
e
millénaires cal. BC), apparenté au PPNB ancien/moyen du
Levant
nord, Comptes Rendu de l’Académie de Science, Paris, Sciences
de la Terre et des planètes 330 (2000) 75e82.
[23] E. Hahn, Die Haustiere und ihre Beziehungen zur
Wirtschaft des
Menschen, Eine geographische Studie, Duncker and Humblot,
Leipzig, 1896.
[24] J.R. Harlan, D. Zohary, Distribution of wild wheat and
barley,
Science 153 (1966) 1075e1080.
[25] B. Hayden, A new overview of domestication, in: T.
Douglas Price,
A.B. Gebauer (Eds.), Last Hunters e First Farmers, School of
AmericanResearchPress,SantaFe,NewMexico,1995,pp.273e299.
[26] D.O. Henry, Models of agricultural origins and proxy
measures
of prehistoric demographics, in: R.T.J. Cappers, S. Bottema
(Eds.), The Dawn of Farming in the Near East, Studies in Early
Near Eastern Production, Subsistence, and Environment, vol. 6,
ex oriente, Berlin, 1999, pp. 15e25 (2002).
[27] E. Isaac, Geography of Domestication, Prentice-Hall,
Englewood
Cliffs, New Jersey, 1970.
[28] R.G. Klein, Stone Age exploitation of animals in southern
Africa,
American Scientist 67 (1979) 151e160.
[29] R.G. Klein, G. Avery, K. Cruz-Uribe, D. Halkett,
J.E. Parkington, T. Steele, T.P. Volman, R. Yates, The
Ysterfontein 1 Middle Stone Age site, South Africa, and early
human exploitation of coastal resources, Proceedings of the
National Academy of Science, USA 101 (2004) 5708e5715.
[30] R.B. Lee, What hunters do for a living, or, how to make
out on
scarce resources, in: R.B. Lee, I. DeVore (Eds.), Man the
Hunter,
Aldine, Chicago, 1968, pp. 30e48.
[31] A.J. Legge, Prehistoric exploitation of the gazelle in
Palestine, in:
E.S. Higgs (Ed.), Papers in Economic Prehistory, Cambridge
University Press, Cambridge, 1972, pp. 119e124.
1416 S.J.M. Davis / Journal of Archaeological Science 32
(2005) 1408e1416
[32] D. Lubell, Are land snails a signature for the Mesolithice
Neolithic transition? Documenta Praehistorica 31 (2004) 1e24.
[33] T.R. Malthus, An Essay on the Principle of Population, As
It
Affects the Future Improvement of Society; with Remarks on
the
Speculations of Mr. Godwin, M. Condorcet and Other Writers,
Johnson, London, 1798 (1970, Penguin, Harmondsworth).
[34] L. Martin, Mammal remains from the eastern Jordanian
Neo-
lithic, and the nature of Caprine herding in the steppe,
Paléorient
25 (1999) 87e104.
[35] A. Morales, E. Roselló, F. Hernández, Late Upper
Paleolithic
subsistence strategies in southern Iberia: tardiglacial faunas
from
Cueva de Nerja (Málaga, Spain), European Journal of Archae-
ology 1 (1998) 9e50.
[36] N.D. Munro, A prelude to agriculture: game use and
occupation
intensity during the Natufian period in the southern Levant,
PhD
dissertation, Department of Anthropology, University of Arizo-
na, Tucson, 2001.
[37] N.D. Munro, Zooarchaeological measures of hunting
pressure
and occupation intensity in the Natufian, Current Anthropology
45 (Suppl.) (2004) S5eS33.
[38] S. Payne, Partial recovery and sample bias: the results of
some sieving
experiments, in: E.S. Higgs (Ed.), Papers in Economic
Prehistory,
Cambridge University Press, Cambridge, 1972, pp. 49e64.
[39] S. Payne, Faunal change at Franchthi cave from 20,000 BC
to
3000 BC, in: A.T. Clason (Ed.), Archaeozoological Studies,
North
Holland, Amsterdam, New York, 1975, pp. 120e131.
[40] J. Perrot, Le gisement natoufien de Mallaha (Eynan),
Israël,
L’Anthropologie 70 (1963) 437e483.
[41] P.J. Richerson, R. Boyd, R.L. Bettinger, Was agriculture
impossible during the Pleistocene but mandatory during the
Holocene? A climate change hypothesis, American Antiquity 66
(2001) 387e411.
[42] P. Rowley-Conwy, The laziness of the short-distance
hunter: the
origins of agriculture in western Denmark, Journal of Anthropo-
logical Archaeology 3 (1984) 300e324.
[43] A. Sherratt, Plate tectonics and imaginary prehistories:
structure
and contingency in agricultural origins, in: D.R. Harris (Ed.),
The
Origins and Spread of Agriculture and Pastoralism in Eurasia,
University College Press, London, 1996, pp. 130e140.
[44] M.C. Stiner, Thirty years on the ‘‘Broad Spectrum
Revolution’’
and paleolithic demography, Proceedings of the National
Academy of Sciences, USA 98 (2001) 6993e6996.
[45] M.C. Stiner, N.D. Munro, T.A. Surovell, E. Tchernov, O.
Bar-
Yosef, Paleolithic population growth pulses evidenced by small
animal exploitation, Science 283 (1999) 190e194.
[46] T.A. Surovell, Modeling occupation intensity and small
game use
in the Levant, in: J.C. Driver (Ed.), Zooarchaeology of the
Pleistocene/Holocene Boundary, Proceedings of a Symposium
held at the Eighth Congress of the International Council for
Archaeozoology (ICAZ) Victoria, British Columbia, Canada
August 1998. BAR International Series, vol. 800, pp. 31e36
(Oxford).
[47] R.M. Sussman, Child transport, family size and increase in
human population during the Neolithic, Current Anthropology
13 (1972) 258e259.
[48] E. Tchernov, An early Neolithic village in the Jordan
Valley. Part
II: the fauna of Netiv Hagdud, Peabody Museum of Archaeology
and Ethnology, Harvard University, American School of Pre-
historic Research. Bulletin 44 (1994).
[49] H.-P. Uerpmann, Probleme der Neolithisierung des
Mittelmeer-
raums, in: Tübinger Atlas des Vorderen Orients, B, vol. 28,
Ludwig Reichert, Wiesbaden, 1979.
[50] F.E. Zeuner, A History of Domesticated Animals,
Hutchinson,
London, 1963.
Why domesticate food animals? Some zoo-archaeological
evidence from the LevantIntroductionMaterial and methodsThe
"Spectrum Shift " (Figs. 2 and 3)The "Age Shift "(Fig.4 and
Table 1)ConclusionAcknowledgementsReferences
ARTICLES
Central Questions in the Domestication of Plants
and Animals
MELINDA A. ZEDER
DEFINING DOMESTICATION
All approaches to defining domesti-
cation in both plants and animals rec-
ognize that domestication involves a
relationship between humans and tar-
get plant or animal populations. There
are, however, distinct and often dis-
cordant perspectives taken regarding
the balance of power in this relation-
ship and its central defining features
(Fig. 1). Many approaches to defining
domestication, especially those focus-
ing on animals, emphasize the domi-
nant role humans play in assuming
“mastery” over all aspects of the re-
production, movement, distribution,
nourishment, and protection of do-
mesticates.2–5 Integral to definitions
that place humans in control of the
process is the notion of intentionality,
that humans with foresight and delib-
erate intent intervened in the life cycle
of target plant and animal popula-
tions and assumed responsibility for
their care to meet specific and well-
defined objectives serving human
needs. Also often associated with this
emphasis on the human dimension is
the notion that domestication in-
volves a fundamental change in socio-
economic organization in which suc-
cessive generations of domesticates
become integrated into human societ-
ies as objects of ownership.3,6
Other researchers object to “anthro-
pocentric” approaches to defining do-
mestication that portray domesticates
as passive pawns in the process, point-
ing out that domesticates also reap
benefits through vastly enhanced re-
productive fitness and expanded rang-
es.7 Those operating within an evolu-
tionary biology perspective, in
particular, maintain that the relation-
ship between humans and domesti-
cates is no different from other mutu-
alistic relationships in the “natural
world” that bring together species like
ants and aphids in partnerships of in-
creasing co-dependency.8 Moreover,
as one moves further along the spec-
trum, from a relatively balanced mu-
tualistic perspective to ones that focus
on the domesticate, the role of delib-
erate human intent declines. The
more extreme positions at this end of
the spectrum tip the balance in favor
of the domesticate, which is seen as
manipulating its human partners for
its own evolutionary advantage, en-
snaring humans in a relationship that
may have actually reduced human fit-
ness.9
Another axis of variation in defini-
tional approaches to domestication is
the relative primary given to genetic
and associated morphological change.
An emphasis on genetic change and
its phenotypic expression is particu-
larly common among researchers fo-
cusing on plant domestication, espe-
cially the domestication of large-
seeded annuals, where human
intervention results in fairly rapid ge-
netic changes with easily observed
phenotypic expressions.9,10 Some re-
searchers focusing on animals also
see genetic isolation and subsequent
quick-onset morphological change as
essential attributes of domestica-
tion.11
The requirement that domesticates
show evidence of morphological or
even genetic change, however, is not
universally accepted. Nor is the basic
premise underlying this requirement:
that the process of domestication is
contingent on reproductive isolation
and resultant genetically driven mor-
Melinda A. Zeder is Director of the Ar-
chaeobiology Program of the Smithso-
nian Institution’s National Museum of Nat-
ural History. Her research focuses on
questions of domestication, origins of ag-
riculture, and the environmental and so-
cial impacts of early agricultural econo-
mies in the ancient Near East. She is the
lead editor of the volume, Documenting
Domestication: New Genetic and Archae-
ological Paradigms, with co-editors Eve
Emshwiller, Daniel G. Bradley, and Bruce
D. Smith, published in spring of 2006 by
the University of California Press.
Key words: domestication; origins of agriculture;
plants; animals; climate change; demographic
pressure; social forces; Southwest Asia
© 2006 Wiley-Liss, Inc.
DOI 10.1002/evan.20101
Published online in Wiley InterScience
(www.interscience.wiley.com).
Along with symbolic communication, tool use, and bipedalism,
the domestica-
tion of plants and animals, together with the associated
emergence of agriculture,
stands as one of the pivotal thresholds in human evolution. For
more than a
hundred years researchers have wrestled with the questions of
what domestica-
tion is, how it is detected, and why it happened. The past decade
in particular has
witnessed a remarkable acceleration of interest in
domestication, thanks to ad-
vances in our ability to detect the context, timing, and process
of domestication in
a wide array of different plant and animal species around the
world.1 This review
focuses on overarching issues of defining, documenting, and
explaining the do-
mestication of plants and animals, tracing a path through often
discordant view-
points to offer some new perspectives.
Evolutionary Anthropology 15:105–117 (2006)
phological change. This is particularly
true for animals, where morphologi-
cal change, when it occurs at all, is
often both delayed and difficult to tie
directly to domestication.12 As a re-
sult, many researchers define animal
domestication not in terms of ob-
served genetic or morphological
change, but in terms of causal human
behavior. According to this view, do-
mestication falls along a continuum
of increasing human intervention
ranging from predation to genetic en-
gineering13 in which there are varying
degrees of investment in altering an
animal’s natural behavior (its move-
ment, breeding schedule, or popula-
tion structure) to suit human
needs.6,14 A similar view is becoming
increasingly common in consider-
ations of plant domestication espe-
cially perennial plants such as root
crops propagated through vegetative
cloning or very long-lived tree crops in
which genetic and morphological
change may be less automatic and
more subtle than in annual seed
crops.15 Smith,16 for example, main-
tains that for both plants and animals
the central defining feature of domes-
tication and the creation of domesti-
cates is the nature of the “ongoing re-
lationship of intervention initiated
and sustained by humans.” This em-
phasis on the evolving relationship be-
tween humans and plant or animal
populations turns attention away
from a range of secondary conse-
quences of domestication, such as ge-
netic and morphological change or so-
cial notions of property, and properly
returns it to a consideration of the
new partnership that humans create
with target populations.
Domestication does indeed have
many features in common with other
mutualistic relationships among
plants and animals. Both partners in
the relationship of domestication
clearly derive benefits. Plant and ani-
mal partners benefit in increased re-
productive fitness and range expan-
sion. Human partners gain increased
security and predictability in their ac-
cess to resources of interest. Both
partners respond to this relationship
in ways that enhance respective pay-
offs and further deepen their mutual
investment in its continuation. But
the mutualism that lies at the heart of
the domestication process involving
human societies and target plant and
animal populations varies in signifi-
cant ways from other similar relation-
ships found in nature.
In a recent overview of the homolo-
gies between human agriculturists
and fungus-growing ants, Schultz and
coworkers17 highlight the many paral-
lels between these convergent forms
of mutualism while also underscoring
key qualitative differences between at-
tine and human agriculture. The co-
dependent relationships between
farmer ants and domesticated fungi
are the result of a gradual co-evolu-
tionary process based on mutation-in-
duced behavioral and morphological
change in both partners. Humans, on
the other hand, are capable of modi-
fying their behavioral repertories
through “trial and error, observation,
and imitation.”17 That ability enables
humans to rapidly develop behavioral
strategies aimed at meeting con-
sciously recognized needs. The highly
developed human ability for cultural
transmission of learned behavior,
Shultz and coworkers argue, greatly
accelerated the adaptive modification
of human behavior, shifting the bal-
ance of power in the emergent mutu-
alism. Humans quickly assume a
dominant role because they are free to
choose among genetic variants in the
partner population, to manipulate the
behavior and life history of symbionts
(even to their own detriment), or to
terminate the relationship with one
partner symbiont and choose another.
This is where intentionality comes
into the picture. It is true that humans
Figure 1. Definitions of domestication tend to fall somewhere
along three axes of variation. Definitions that award the balance
of power
in the domestic relationship to humans tend to stress human
intentionality and the social and economic impacts of
domestication.
Definitions that tip the balance of power in favor of the
domesticate tend to discount the role of human intentionality in
the process and
stress its biological impacts on the domesticate.
106 Zeder ARTICLES
could not have foreseen the adaptive
responses by plant and animal part-
ners to the new selective factors
brought into play by the relationship
of domestication. Nor did humans
likely appreciate the long-term bene-
fits (or the negative consequences)
that might accrue from domestication
and the subsequent development of
agricultural economies. However, rul-
ing out this kind of prescience on the
part of humans does not take inten-
tionality out of the picture. While they
might not have understood the princi-
ples of genetic engineering, humans
could appreciate the fact that tending,
nurturing, and intervening in the life
cycle of certain plants and animals
yielded various immediate benefits.
On the basis of these returns they
could then consciously and deliber-
ately decide to continue to engage in
these behaviors, and to elaborate on
them, instead of engaging in other
strategies.
Intentionality, then, becomes the
key factor that distinguishes domesti-
cation from other similar mutualistic
relationships in nature. The deliberate
role humans take in actively pursuing
the domestic partnership also distin-
guishes it from other biological rela-
tionships between humans and plants
and animal species, such as commen-
sal relationships with mice, sparrows,
or weeds that take advantage of new
niches created by human habitation.
Definitions that try to pigeon-hole do-
mestication as either a cultural or bi-
ological process are bound to come up
short. Clearly, domestication has a bi-
ological component as a mutualistic
relationship between humans and
plant or animal symbionts. Just as
clearly, however, human intentional-
ity sets domestication apart from
other forms of mutualism. The
uniqueness of the relationship comes
from its cultural component and the
dominant role humans play in con-
sciously and deliberately perpetuating
it to their own advantage.
If the process of domestication is
best viewed as a form of mutualism
that is asymmetrically enhanced by
the human ability to culturally trans-
mit learned behavior, then at what
point along this developmental trajec-
tory does the plant or animal partner
become a domesticate? Is there a
threshold that, once crossed, sepa-
rates the “wild” from the “domestic”?
If so, what does this threshold look
like? To some extent, it remains a
matter of personal preference to de-
cide just when a domestic subsection
of a plant or animal species has been
created. Threshold criteria that re-
quire total genetic isolation and emer-
gent speciation or complete depen-
dence on humans for survival set a
very high bar that many, if not most,
widely accepted domesticates would
fail to clear. Even somewhat looser
standards that involve a lesser degree
of genetic modification in the target
plant or animal population, or a cer-
tain level of human investment in
propagating, nurturing, or owning the
resource, run the risk of constructing
artificial boundaries along what was
really a more seamless incremental
process.
Ducking the issue by adopting the
term “proto-domesticate” also does
not help much. This term implies that,
if just given enough time and perhaps
a little more investment by either part-
ner, full domestic status would be
achieved. The actual trajectory of do-
mestication, however, is highly con-
tingent on a wide range of factors,
including the ability of the plant or
animal to take advantage of the rela-
tionship, the strategies and accompa-
nying technologies humans develop to
manage the resource, and its chang-
ing value vis-à-vis available alternative
resources. In some plant and animal
species, genetic modification and
more focused human investment in
the resource may quickly follow. In
others there may be a long and very
stable relationship involving fairly
minimal commitment by either part-
ner. Further, it appears that budding
domestic relationships sometimes fail
altogether, never moving beyond an
initial courtship phase.
It is best to step back and not focus
too closely or obsessively on defining
the exact demarcation between do-
mestic and wild, and to turn, instead,
to a consideration of the full span of
the evolving nature of domestic rela-
tionships. Different stages in the evo-
lution of this relationship might be
characterized by the degree of invest-
ment by both partners (Fig. 2). For the
plant or animal, this would involve the
extent of genetic modification made in
response to new selective pressures,
the degree of its genetic isolation from
populations not involved in the part-
nership, the nature of subsequent
morphological or behavioral change,
and its increasing co-dependency on
humans. For humans, this might be
the level of investment in the produc-
tion of the resource; that is, in tilling,
watering, burning, and land clear-
ance, sowing, and transplanting
plants, or in taming, protecting, herd-
ing, culling, and selectively breeding
animals. It might also include the de-
gree of incorporation of domesticates
within the socio-economic organiza-
tion of the human groups investing in
its production.
By expanding the scope of inquiry
to encompass the vast “middle
ground” between foraging and farm-
ing, hunting and herding,16 we can
approach a deeper, more comprehen-
sive, and ultimately more informative
appreciation of the range of possibili-
ties open to humans and their plant
and animal partners. This expanded
territory of investigation includes the
stable, long-lived systems of low-level
food production involving a mix of
both morphologically altered and
nonaltered domesticates, as well as
“wild” resources, featured in recent
books on indigenous resource man-
agement in California and the North-
west Coast.18,19 At the other end of the
spectrum are highly structured agri-
cultural economies with complete de-
pendence on domesticates and total
investment in their production. Try-
ing to understand the full richness of
the various ways in which the domes-
tic partnership may manifest itself in
different contexts, we, in turn, stand a
much better chance of being able to
document and explain domestication.
DOCUMENTING
DOMESTICATION
In both plants and animals, this ef-
fort requires identifying clear-cut
markers that can be explicitly linked
to a specific aspect or stage of the un-
folding domestication process.1,20 Dif-
ferent markers may be more effective
in detecting different stages of this
process. Markers will also vary de-
pending on the biology of the domes-
ARTICLES Domestication of Plants and Animals 107
ticate and its relationship with hu-
mans. There are, in particular,
fundamental differences in the selec-
tive pressures on plants and animals
undergoing domestication, and, as a
result, in the corresponding markers
used to document plant and animal
domestication.
Selective pressures on plants, espe-
cially annuals, tend to operate directly
on morphological traits that can, in
turn, be used as unambiguous mark-
ers of domestication.15 Morphological
impacts of the domestication of annu-
als may come about as largely auto-
matic responses to human planting
and harvesting that result in such
changes as increased seed size, thin-
ner seed coats, reconfiguration of seed
head architecture, or the development
of indehiscent seed pods.15,21,22 Inten-
tional selection for specific morpho-
logical attributes in annual plants,
such as larger fruit size, appear to
happen later in the developing rela-
tionship of domestication.15
Perennial plants sustained by trans-
planting root fragments, on the other
hand, are not subjected to the same
seed-bed pressure and human har-
vesting selective pressures that result
in the morphological markers used to
document domestication in annual
seed plants.15 At the same time, how-
ever, because there may be more of an
opportunity for humans selectively to
replant root fragments with desired
traits, these plants may respond fairly
quickly to deliberate human selection
in the development of larger fruits, the
loss of chemical defenses against her-
bivory, or changes in sugars and
starches.23 While many of these crops
were grown in tropical areas with
poor preservation of plant macro-fos-
sils, the development of breakthrough
techniques for the recovery and iden-
tification of plant micro-fossils (that
is, phytoliths and starch grains) has
made it possible to detect these do-
mestication-induced morphological
changes in root and other crop plants
(Fig. 3).24,25
Recent years have seen an increase
the use of nonmorphological markers
of the intensification of human-plant
interactions that may precede clear-
cut evidence of morphological change
in plants. Evidence of land clearance,
modification of natural drainage sys-
tems, intentional burning, and
changes in the composition of weedy
plants in archeological assemblages
have all been effectively used to track
human modification of landscapes
and plant communities as part of the
domestication process.26 –28 The oc-
currence of plant macro- or micro fos-
sils in areas thought to be far outside
their natural range has also been in-
terpreted as evidence of human trans-
port and tending of plants.26,29,30
There are special challenges to find-
ing markers of animal domestication.
This is because the leading-edge pres-
sures on animals undergoing domes-
tication are likely to focus on behav-
Figure 2. Domestication is best viewed as an evolving of
mutualism between humans and populations of plants or
animals. The relationship
can be characterized along various scales of investment by
either the human or the plant or animal partners. All of these
scales usually are
involved in the process of domestication, though they often
operate independently of one another. The degree of change
along each
scale is contingent on the biology of the species involved, as
well as the ecological and cultural circumstances of the human
partners.
Attempting to distinguish just where and along what scale
domestication occurs is not only difficult, but may not be very
useful.
108 Zeder ARTICLES
ioral attributes rather than on
morphological traits.31 There are a va-
riety of behaviors that probably made
certain animal species better candi-
dates for domestication; among them
tolerance of penning, a social struc-
ture based on dominance hierarchies,
sexual precocity, weak alarm systems
and, above all, reduced wariness and
aggression.32 Behavioral responses to
domestication in animals elaborated
on these initial preselection qualifying
attributes and include a general re-
duction in responsiveness to environ-
mental stimuli, reduced activity lev-
els, increased social compatibility,
and intensified sexual behavior.4,33
Many morphological traits com-
monly seen in domestic animals are
thought to be linked to these behav-
ioral changes. These attributes in-
clude piebald coats, lop ears and, of
special importance here, reduced
brain size and an overall juveniliza-
tion of cranial form.4,33 This latter fea-
ture may result in a shortened muzzle,
tooth crowding, and reduction in
tooth size, traits frequently seen as
leading-edge markers of domestica-
tion in dogs and pigs.34,35 Selection
for these behavioral traits and their
associated morphological effects,
however, may not be uniquely re-
stricted to domestication. Similar be-
havioral traits, such as reduced wari-
ness and timidity, are also selected for
in animals such as rats and sparrows
that develop commensal relationships
with humans. These animals also
show changes in pelage coloration
and brain size.4 It is possible, then,
that the initial modifications in tooth
size and cranial form in pigs and dogs,
widely seen as markers of domestica-
tion, may in fact be attributable to an
early commensal relationship be-
tween humans and such omnivore
species that began their association
with humans as camp-follower scav-
engers.31
Other genetically driven morpho-
logical changes in animals undergo-
ing domestication come about when
humans begin deliberately selecting
breeding partners. Changes in the size
and shape of horns in animals like
goats and sheep, for example, are di-
rectly tied to the relaxation of selective
pressures for and, quite likely, active
selection against large horns once hu-
mans assume control over breeding.36
Other morphological changes may en-
sue when animals are moved into new
territories, either through founder ef-
fects, random genetic drift, or di-
rected adaptation to new environmen-
tal conditions. Later still, and
probably much later in animals than
in plants, deliberate human selection
for attributes that enhance meat, fi-
ber, milk yields, or labor potential
may result in still other morphological
markers that might be used to detect
intensification in the human-animal
relationship.
Domestication has also been sug-
gested to have resulted in a marked
and rapid reduction in body size,
which, until recently, has been widely
held to be a definitive marker of initial
domestication.37 This proposed mor-
phological response has been vari-
ously attributed to a plastic response
to nutritional deficiencies, an adaptive
advantage of smaller bodies for ani-
mals subjected to impoverished con-
ditions, or deliberate human selection
for more tractable individuals.38 – 40
But body size in animals is also
known to be affected by well-docu-
mented factors such as sex, environ-
ment, climate, and age, which may be
entirely unrelated to domestication
and may mask or be mistaken for
changes in body size induced by do-
mestication.12
Given the looser connection be-
tween domestication and morpholog-
ical change in animals, it is not sur-
prising that considerable attention
has been devoted to identifying mark-
ers of domestication that do not rely
on genetically driven morphological
change, but that, instead, reflect hu-
man actions directed at managing an-
imals. Demographic markers aimed at
detecting the different harvest strate-
gies of hunters and herders were
among the first nonmorphological
markers used to detect animal domes-
tication.41 Largely abandoned in the
1980s and 1990s, when most archeo-
zoologists embraced size reduction as
a leading-edge marker of animal do-
mestication, demographic markers
are seeing a resurgence, thanks in part
to the development of methods for
constructing high-resolution sex-spe-
cific harvest profiles.12,42 Applying
these methods to archeological as-
semblages has shown that what was
once interpreted as evidence of do-
mestication-induced body size reduc-
tion in goats (and likely other live-
stock species) is, instead, a reflection
of a change in the demographics of
the adult portion of managed herds
dominated by females (Fig. 4).12 Un-
ambiguous changes in morphological
traits such as body size or horn form
seem to postdate human management
of herd animals by hundreds of years
and represent later phases in the do-
mestication process.12,43
Both the presence of animals out-
side their presumed natural habitat
and a sudden dramatic increase in a
previously little-exploited animal have
also been used as markers of animal
domestication.41,43,44 But the use of
these markers (in both animals and
Figure 3. Starch grains from wild and domes-
tic yams, Dioscorea sp. A. Granules from
modern domestic yams (D. trifida). B. starch
grains from a wild yam (D. cymosula) from
Panama. Starch grains from wild yam spe-
cies studied thus far from the Neotropics are
distinct in morphology as compared with
domestic varieties. As a rule, wild forms are
also highly variable within a single tuber,
whereas domesticated species have a sin-
gle morphological type of starch, which
may be a result of human selection.
ARTICLES Domestication of Plants and Animals 109
plants) needs to be tempered by ac-
knowledgment of our generally poor
understanding of the geographic
range of biotic communities in the
past and of the paleo-environmental
conditions that shaped these ranges. A
rapid increase in the abundance of a
plant or animal resource in an arche-
ological assemblage might simply sig-
nal the intensification of hunting and
gathering strategies, not the begin-
ning of food production.
Markers of animal domestication
may also be found in plastic, nonge-
netically driven responses such as
bone and tooth pathologies, evidence
of pandemic disease, or chemical
changes in the composition of bone
and tooth enamel used to track
changes in nutrition and the seasonal
movement of managed animals.45– 48
The presence of corrals, pens, or other
traces of animals, such as manure or
hoof prints, in human settlements,
changes in human settlement pat-
terns, artifacts related to the exploita-
tion of domestic animals (bits or milk
churns and storage vessels), and even
changes in food distribution patterns
have been used with varying effect to
build cases for animal domestica-
tion.49 The application of these later
plastic responses and cultural mark-
ers needs to be tempered by the real-
ization that they may not be mani-
fested in all instances of animal
domestication or may result from
other pressures unrelated to domesti-
cation. Application of such markers is
most effective when many of them are
brought together to build strong cir-
cumstantial cases for domestica-
tion.49
Advances in methods for extracting
and amplifying both modern and an-
cient DNA is recent years have pro-
vided an exciting new window on the
genetic changes associated with the
domestication of plants and ani-
mals.1,20,50,51 Some of this work has
focused on identifying the genes or
gene complexes that are specifically
selected for or against in the process
of domestication, especially of crop
plants.52 However, most genetic stud-
ies of domestication look to largely
neutral noncoding loci and organellar
genomes. These procedures have
proven useful in tracing the diver-
gence of domesticates from their wild
progenitors; identifying the number
and geographic location of domestica-
tion events, which now appear to have
been multiple for most animal domes-
ticates and many plant crops; and in
tracking the dispersal of domesticates
and their human partners out of cen-
ters of origin.
There are critical differences in the
relative rates of evolution in the dif-
ferent genomes of plants and animals
that play a major role in the genetic
markers used. The relatively rapid
rate of evolution in mitochondrial
DNA (mtDNA) in animals makes
mtDNA particularly well suited to
tracking the relatively shallow time
depth of divergence between domesti-
cates and their wild progenitors
(�10,000 years). This is why most
studies of animal domesticates focus
on this genome.51 While it is less vari-
able than mtDNA and evolves much
more slowly, low-copy Y-chromo-
some nuclear DNA provides a window
into patrilineal inheritance, which in
many animal domesticates is quite
different from matrilineal history.53,54
Variation in noncoding nuclear mic-
rosatellite DNA, contributed by both
parents, has also proven useful in
tracking the divergence of different
breeds of animals.55,56
While not approaching the rate of
evolution of animal mtDNA, loci in
the nuculear genome of plants evolve
at about the same rate as nDNA loci in
mammals. They also evolve about
four times faster than loci in the
choloroplast genome and twelve times
faster than plant mtDNA. Conse-
quently, genetic studies of plant do-
mestication tend to focus on nDNA,
especially on highly polymorphic mi-
crosatellites that provide sufficient in-
traspecific variation to document the
domestication process.50,57–59 The nu-
clear genome in plants has proven es-
pecially useful in tracking down the
various ancestral genomes contribut-
ing to the complicated genetic heri-
tage of hybrids and polyploid crop
plants. These common conditions in
plant crops generally are not found
among animal domesticates.50,60
Most genetic approaches to docu-
menting domestication are based on
modern domesticates and likely wild
progenitor species. But the window
they provide on the origin and early
dispersal of domesticates is unavoid-
ably clouded by thousands of years of
selective breeding, hybridization, and
introgression between wild and do-
mestic populations. Ancient DNA
(aDNA), on the other hand, has the
potential to shed more direct light on
the process of genetic divergence of
domesticates. Due to the greater pres-
ervation of DNA encased in animal
bone and the suitability of high-copy
mtDNA in animals for tracking shal-
low time depth divergences, aDNA
studies of animal domesticates have
been particularly successful, espe-
cially those tracing the more recent
dispersal of domestic animals through
temperate environments.61– 64 Al-
though it is more difficult to extract
enough low-copy nDNA from un-
charred archeological plant remains
to provide meaningful results, some
stunning results have recently been
Figure 4. Goat (Capra hircus) first phalanges
from Ganj Dareh ca. 10,000 cal B.P. The
small bone on the left falls within the size
range of such bones from female goats,
while the larger specimen on the right is
likely from a male. The female phalanx is
fully fused, indicating the animal was
slaughtered after this bone fused, which in
goats is about 13 months of age. The male
specimen is unfused, indicated that this an-
imal was younger than 13 months of age
when slaughtered. The selective slaughter
of young male animals with prolonged sur-
vivorship of female animals is symptomatic
of human management. (Photo by Carl
Hansen)
110 Zeder ARTICLES
obtained in the use of aDNA to track
the origin and dispersal of domestic
plants.65,66
In the excitement over the possible
contributions of genetic analysis to
the documentation of plant and ani-
mal domestication, it is important
not to lose sight of the fact that there
is more to domestication than ge-
netic change. The real power of
these new tools for tracking the tra-
jectory of domestication can be real-
ized only when genetic analyses are
more fully integrated into broader
archeological analyses. Genetic
studies represent one, albeit very
powerful, line among many parallel
and mutually illuminating lines of
evidence, which, when considered
together, provide a fine-grained view
of unfolding domestic partnerships
(Box 1).
EXPLAINING DOMESTICATION
Efforts to explain domestication
and the origins of agriculture tend to
cycle among a relatively limited num-
ber of forcing factors championed as
primary and often universal levers of
change. These forcing factors can be
generally grouped under three banners:
environmental change, demographi-
cally induced resource pressure, and
changes in social organization and ide-
ology.
Explanations focusing on environ-
mental change can be traced to V.
Gordon Childe, who credited post-
Pleistocene aridification with bring-
ing humans together with plants and
animals around water sources in po-
sitions of “enforced juxtaposition”
that promoted “symbiosis between
man and beast,” resulting in domesti-
cation.67 While climate models were
largely out of favor in the 1960s
through 1980s, recent advances in
paleo-environmental reconstruction
have resulted in its rehabilitation as a
primary player in agricultural origins.
In particular, the now well-docu-
mented brief return to Ice Age condi-
tions at about 13,000 cal. B.P., known
as the Younger Dryas, is increasingly
featured as having played a key role in
agricultural origins,68 with domestica-
tion coming about either during this
climatic downturn as a way of coping
with environmental degradation69,70
or after it as a response to the follow-
ing climatic amelioration and stabili-
zation.71 One recent climate based
model of agricultural origins even
maintains that agriculture was a
“compulsory” development of cli-
matic stabilization and rise in ambi-
ent CO2 following the final pulse of Ice
Age climate in the Younger Dryas.72
Explanatory frameworks founded
on notions of population dynamics
and resultant resource pressures can
be traced back to Binford’s Edge-Zone
Hypothesis of the late 1960s.73 Ac-
cording to this theory, agricultural or-
igins were the result of resource pres-
sure in marginal zones caused by
emigration from more optimal zones
experiencing high rates of population
growth. Mark Cohen’s74 subsequent
“food-crisis” model held that runaway
population growth worldwide, not
just in marginal zones, forced people
around the planet to abandon more
nutritious hunting and gathering
strategies and assume the burden of
tending to domesticated plants and
animals. Binford’s75 recent return to
the topic down-plays the role of pop-
ulation growth, resource pressure,
and emigration, and instead high-
lights population packing, in particu-
lar a threshold limit of 9.098 people
per 100 km2 as the “universal condi-
tioner of change in . . . subsistence
strategies” and the deus ex machina of
agricultural origins.
Although explicitly rejecting univer-
sal normative explanations of domes-
tication and agricultural origins,
those operating under the general ru-
bric of human behavioral ecology
(HBE) also feature resource pressure
as a primary causal component.76 A
basic axiom of HBE models is that
humans will always opt for strategies
that optimize immediate returns from
high-ranking resources, with the rela-
tive rank of a resource determined by
kilocalorie return over pursuit and
handling costs.77 Food production vi-
olates these central principals. First of
all, farming involves a rise to promi-
nence of low-rank plant resources,
while herding requires deferring the
rewards from high-rank animal re-
sources.78,79 Moreover, farming and
herding are often seen as carrying
higher production and processing
costs than do hunting and gather-
ing.80 A recent review highlights dif-
ferent paradigms within HBE that
might be helpful in understanding the
transition from foraging to farming,
such as concepts of constrained opti-
mization, marginal value, opportunity
costs, discounting, and risk-sensitivi-
ty.76 Diet-breadth models that include
avenues for plant resources of lower
profitability to elevate their rank
through changes in density or extrac-
tion return hold particular promise
here. However, even allowing for such
advances up the resource-rank ladder,
HBE models generally predict that
plant resources will largely be ignored
so long as animal protein, a more
highly ranked resources, is sufficiently
abundant.78 By necessity, then, strict
adherence to the basic axiomatic te-
nets of HBE casts the transition to
food production in terms of rather
bleak cost-benefit trade-offs. Follow-
ing HBE principles, agriculture comes
about when people, faced with pres-
sure on resources, whether caused by
fluctuating climates, population
growth, or packing, are forced to fo-
cus on resources and extraction strat-
egies that under other circumstances
would be considered far from opti-
mal.
Other theorists deny that external
factors like climate change, popula-
tion growth, or resource pressure
have any causative role in this transi-
tion. Instead, they look to forces
within the human character, espe-
cially a supposedly innate compulsion
for self-aggrandizement, as primary
forcing factors in the domestication of
plants and animals and the transition
to agriculture.81 Perhaps the best
known and most influential of these
socially based explanatory ap-
proaches is that promulgated by
Brian Hayden.82,83 In contrast to a
backdrop of scarcity and stress cen-
tral to other models, Hayden conjures
up a more Eden-like setting for the
origin of plant and animal domestica-
tion, claiming that agriculture only
develops in contexts of plenty, where
an abundance of resources ignites a
basic human predisposition for acqui-
sition and social competition. In such
settings certain particularly success-
ful “accumulators” were able to mar-
shal high-prestige food items and en-
hance their own social advantage
through such mechanisms as compet-
ARTICLES Domestication of Plants and Animals 111
itive feasting. Domestic resources
were especially attractive since they
were more amenable to ownership
and their supply could be both manip-
ulated and appropriated. This model
predicts, then, that domesticates were
not, at least initially, ubiquitous di-
etary staples, but were, instead, more
likely to be rare and desirable limited-
access delicacies, of little nutritional
value, used by avaricious accumula-
tors to both display and enhance their
social power.
Another stress-free theory, this one
championed by Jacques Cauvin,84
Box 1. Gourds, Dogs, and the Peopling of the Americas
Location of archeological sites with directly dated bottle gourd
fragments studied by
Erickson and coworkers.66
The recent integration of archeo-
logical and genetic research on both
the bottle gourd (Lagenaria siceraria)
and the dog (Canis familiaris) has
substantially clarified our understand-
ing of the initial domestication of
these two very early domesticates, as
well as their late Pleistocene radiation
from Asia across Beringia into the
New World.20
The bottle gourd, indigenous to Af-
rica and long valued as a container
crop rather than as a food source, has
consistently been recovered in arche-
ological association with the earliest
evidence of New World domesticated
plants in many regions of the Ameri-
cas. The prevalent consensus has
been that L. siceraria was carried by
ocean currents as a wild plant from
Africa to South America. However, a
morphological analysis comparing ar-
cheological rind fragments from sites
in the Americas (Fig. 1) with recently
described wild L. siceraria fruits from
Zimbabwe showed that the much
thicker archeological rinds represent
domesticated plants.66 Direct dates
of these fragments indicate that do-
mestic bottle gourd was present in
the Americas by at least 10,000 years
ago. Moreover, DNA recovered from
nine archeological rind fragments
predating the arrival of Europeans
were identical to the modern Asian
reference group, indicating that do-
mesticated bottle gourds were car-
ried to the New World during the late
Pleistocene from Asia, not Africa. Al-
though it is possible that the bottle
gourd could have been carried from
Asia to the Americas by the north Pa-
cific current, it is more likely this early
domesticate accompanied Paleoin-
dian colonists as they crossed Ber-
ingia into the New World.
A parallel genetic study points to
an Old World origin of the domestic
dog and suggests that at least five
founding dog lineages invaded
North America with humans as they
colonized the New World.63 The ear-
liest archeological evidence of do-
mesticated dogs in the Old World,
dating to ca. 13,000 –17,000 B.P.,
comes from widely dispersed sites
extending from the Near East
across Eastern Europe. Although
the earliest domesticated bottle
gourd in the Old World dates to
8,000 –9,000 B.P. in China and Ja-
pan, it is reasonable, given its arrival
in central Mexico by 10,000 B.P., to
estimate that it was initially domes-
ticated in the same general time
frame as the dog, ca 13,000 B.P. or
earlier. Together, these studies indi-
cate that Paleoindians entered the
New World with the world’s two ear-
liest domestic species, dogs and
bottle gourds, and that initial do-
mesticates served utilitarian func-
tions, but not as sources of food.
112 Zeder ARTICLES
also sets the stage for agricultural or-
igins in a time of plenty and denies
economic necessity a primary cata-
lytic role. But Cauvin looks even more
inward into the human psyche for the
root cause of this transition. Accord-
ing to Cauvin, domestication is a di-
rect consequence of a conceptual shift
in mankind’s mental template from
one that saw humans as part of nature
to one that cast humans in a dominant
position, now free to manipulate and
transform nature to their liking. This
profound and irreversible transforma-
tion in the way that humans saw
themselves in relation to nature, cod-
ified in religious ideology, found ex-
pression in concrete ways: in art,
household and community structure,
and the domestication of plants and
animals. A similar notion can be
found in Hodder’s85 emphasis on the
role of symbols and metaphors of hu-
man dominance over nature made
concrete in the form of the house, as
the domus of domestication and the
crucible of community.85
Increasing sophistication in ap-
proaches to defining and document-
ing domestication make it ever more
difficult to support explanatory frame-
works based on any single forcing fac-
tor. This is particularly true for the
Near East, where we have, arguably,
the most complete record of the initial
domestication of many plant and ani-
mal species.
Climate clearly played its part in ag-
ricultural origins here. Increases in
rainfall and temperature following the
Last Glacial Maximum at about
15,000 cal. B.P. undoubtedly contrib-
uted to the adoption of increasingly
less mobile, more territorially focused
strategies centered on intensive ex-
ploitation of plant resources rebound-
ing out of glacial refugia,70 where they
had been used by humans since the
Upper Paleolithic.86,87 A subsequent
pulse of cold, arid Ice Age conditions
during the Younger Dryas, ca. 13,000
to 11,600 cal. B.P., was met with more
mobile strategies in the Southern Le-
vant88,89 and, quite possibly, the initial
domestication of cereal crops and
pulses in well-watered oasis localities
in the Northern Levant.70 Ameliorat-
ing climates following the Younger
Dryas saw the domestication of other
crop plants in the Southern Levant.86
Animal domestication seems to have
come about sometime later on in the
Central and Eastern Fertile Cres-
cent.12 A brief warming and drying
climatic pulse at about 9,000 cal. B.P.
coincided with the collapse of early
agricultural communities in more
arid parts of the Southern Levant and
their proliferation throughout the rest
of the Near East.88 So while climate
change played a role in domestication
and agricultural origins in this region,
it did not do so in the simple stimulus-
response way implied by many mod-
els that award environment prime-
mover status. Instead, climate change
alternately helped push and pull peo-
ple along a pathway toward domesti-
cation and agriculture, providing both
opportunities and challenges that
people across the region met in vari-
ous ways depending on highly local-
ized circumstances.
The transition from foraging to
farming in the Near East clearly saw
significant changes in mobility, popu-
lation growth, and nucleation. But the
record from the region does not sup-
port the exponential population
growth and attendant impoverish-
ment of natural resources called for in
Cohen’s90 doomsday model. In the ab-
sence of settlement-pattern data ro-
bust enough to demonstrate that his
population-packing Rubicon (9.098
people per 100 km2) had been crossed,
Binford pointed to the clear reduction
in mobility, intensification in plant re-
source use and development of stor-
age technology during the postglacial
era in the Near East as proxy evidence
of packing.75 This troubling circular-
ity of using the proposed results of
packing as evidence of packing is
found in many such demographic
models.91 While population clearly in-
creased, the admittedly incomplete
settlement-pattern data do not sup-
port the level of population packing
that Binford grants exclusive causality
for increasing sedentism, intensifica-
tion of resource use and, ultimately,
domestication. It is also hard to make
a case that the initial focus on re-
bounding populations of cereals,
pulses, and nut trees after the Glacial
maximum was caused by widespread
depletion in higher ranking animal re-
sources, as is required by most HBE
models.89
Rather than being forced to settle
down and focus on less desirable re-
sources, it seems more likely that peo-
ple took advantage of newly abundant
high-yield plant resources and associ-
ated herbivores in ways that enabled
them to increase the size and duration
of community nucleation beyond that
possible under Ice Age conditions. It
is also possible that when people were
faced with localized pressures on re-
sources resulting from more seden-
tary ways of living, an interest in pre-
serving the bonds of community
provided an important incentive for
the development of strategies that
helped promote the yield and predict-
ability of these resources. Moreover,
these same social considerations also
probably helped guide the subsequent
responses to region-wide pressures
caused by the climatic squeeze of the
Younger Dryas and the stabilization
of climate that followed.
Yet while it is possible to award so-
cial factors a more active role in the
origins of agriculture in the Near East,
the record from the region clearly
does not support Brian Hayden’s82,83
competitive feasting model. Even
those with the loosest tether to reality
would have a hard time seeing cereals,
the earliest and most important do-
mesticates in the region, as anything
but widely available staple resources.
And while there is some evidence for
feasting on large numbers of ani-
mals,92 no credible case can be made
for Hayden’s83 blanket assertion that
meat was consumed only within con-
trolled ritual contexts. Moreover, all
of the indicators of status differentia-
tion and unequal appropriation of so-
cial and economic prestige that Hay-
den sees as material manifestations of
his greedy accumulators have since
been more convincingly cast by
Kuijt93 and others as evidence of
mechanisms for maintaining an
equalitarian status quo in the face of
mounting social tensions incurred
when larger groups stay together for
longer periods of time.93
Kuijt maintains, however, that do-
mestication played little role in these
social developments, either as a cause
or a consequence.94 He bases this con-
clusion on the fact that morphologi-
cally altered domesticates appear in
the archeological record of the Levant
ARTICLES Domestication of Plants and Animals 113
several thousand years after the first
signs of leveling mechanisms for pro-
moting social cohesion in the Natu-
fian and at least a thousand years be-
fore their ultimate collapse and the
emergence of social inequality in the
Late Pre-Pottery Neolithic. Like oth-
ers who use the appearance of mor-
phological domesticates as a thresh-
old moment in their causal
scenarios,84 Kuijt makes the common
mistake of conflating morphological
change with domestication. Once the
artificial boundary crossing of mor-
phological change is removed and fo-
cus is more properly placed on the
evolving relationship between hu-
mans and plants and animals, then a
more complex and ultimately more
satisfying picture emerges of the syn-
chronous and mutual reinforcing so-
cial and economic forces that shaped
the Neolithic Near East.
The partnership between humans
and resurgent plant and animal re-
sources that began the domestication
process after the Last Glacial Maxi-
mum provided the measure of re-
source security and predictability
needed to establish nucleated seden-
tary communities bound together by
an ethos of balanced reciprocity and
increasingly elaborate rules of social
order and altered world views. By the
same token, the goal of maintaining
community was probably a factor
contributing to the subsequent inten-
sification of evolving domestication
relationships, with their attendant im-
pacts on both the human and the
plant and animal partners. The do-
mestication process, in turn, created
resources that were more amenable to
ownership and restricted access, ulti-
mately contributing to the overthrow
of the egalitarian communities they
initially helped foster and maintain.
Thus, rather than a single forcing
mechanism, it seems more likely
that the trajectory of plant and ani-
mal domestication in the Near East
and the emergence of agriculture
was shaped by various broad-scale
factors, such as climate change, eco-
nomic goals, and social opportuni-
ties and constraints, interacting with
highly local, contingent factors, such
as the density and diversity of avail-
able resources, the history of human
occupation, and the agency of indi-
viduals coping with their environ-
ment, each other, and their universe.
While the entire region was engaged
in this process, the pace and the di-
rection it took varied depending on
the distinctive mix of these factors in
the Southern and Northern Levant,
the Central and Eastern Fertile Cres-
cent.
A similar range of factors operat-
ing at the same time in other places
on the planet took very different
courses (Fig. 5). In Mexico, for ex-
ample, squash, corn, and beans, ap-
parently domesticated in different
places and at different times, were
minor components of mobile forag-
ing strategies for millennia before
the adoption of sedentism and the
development of agricultural econo-
mies.95 In both eastern North Amer-
ica and Japan, small-seeded locally
domesticated annuals were blended
into the diverse economic round of
stable, sedentary low-level food pro-
ducers for thousands of years before
the introduction of domestic crop
plants, maize in eastern North Amer-
ica and rice in Japan. Previously
used in small numbers, these plants
formed the foundation for the emer-
gence of more fully agricultural
economies and increasingly strati-
fied societies.95,96
Thus, the story of domestication
and agricultural origins consists of a
Figure 5. Boxes indicate the general location of centers of
independent plant or animal domestication. Currently, at least
ten such centers
are recognized around the world, making research on the origins
of agriculture a truly world-wide enterprise. Future research
will, no doubt,
discover others.
114 Zeder ARTICLES
series of complex regional puzzles
shaped in unique ways by a dynamic
multi-scalar range of macro- and mi-
cro-forces. Attempts at explanation
that champion any one of these fac-
tors and deny the importance of oth-
ers will not, in the long run, contrib-
ute to understanding agricultural
origins either as a general process or
as it played out in particular in-
stances.
CONCLUSIONS
As this review demonstrates, central
questions about the definition, docu-
mentation, and explanation of domes-
tication are not easily answered. Do-
mestication cannot be simply defined
as either a biological or a cultural phe-
nomenon, but rather needs to be seen
as a form of biological mutualism
transformed by the highly developed
human capacity to effect behavioral
change through learning and cultural
transmission. Definitional approaches
to domestication are most effective,
then, when they focus on the evolving
relationship between humans and tar-
get plant or animal populations as a
nexus between biology and culture,
not on the manifestations or conse-
quences of such relationships. Ge-
netic and related morphological
changes in domesticates are not defin-
ing features of domestication, but are
instead artifacts of evolving relation-
ships that vary in their intensity and
pace of development. Notions of own-
ership and restructuring of social re-
lations are similarly best viewed as
possible results of domestication, not
as central to its definition. Nor are the
clear-cut thresholds that define when
wild resources become domesticated
ones. Rather than looking for defini-
tive either-or boundary conditions in
defining domesticates, it is much
more profitable, if more challenging,
to look at the whole span of evolving
domestic relationships as they operate
over various scales of investment on
the part of both human and plant or
animal partners.
There are also no easy, universally
applicable ways to document domes-
tication. Instead, documenting do-
mestication requires a clear under-
standing of the species-specific
linkage between a proposed marker of
domestication and the stage of the un-
folding domestication process it is
held to mark. It also requires recog-
nizing that markers vary depending
on the biology of the species involved
and the cultural context of human
populations engaged in the domesti-
cation process. Above all, effective
documentation means not letting the
availability of new scientific tech-
niques lead the search for new mark-
ers without first thinking about how
the process of domestication might
manifest itself in whatever these tech-
niques are designed to measure.
So, too, causal scenarios that nar-
rowly focus on single, universally ap-
plicable prime-mover levers of change
will never provide satisfying answers
to the critical “why” questions about
the origins of domestication and sub-
sequent agricultural emergence. It is
easy when dealing with complex and
extended processes to draw artificial
thresholds that help make the case for
the primacy of whatever variable one
is sponsoring as the cause of events
that follow, whether it be climate
change, population increase, or social
and ideological transformations. Yet
we have gained too sophisticated an
understanding of the process of do-
mestication and the means of detect-
ing it to support the kind of drive-by
theorizing that selectively chooses ac-
commodating bits of information
from individual regional scenarios to
support a favored epistemology de
jour. Advances in answering “why”
questions about domestication and
agricultural origins can only be
gained through close-grained under-
standing of complex multi-scalar re-
gional puzzles and assessment of the
commonality and the differences in
the way the pieces of these different
puzzles fit together.16,97
There are, then, no easy answers to
central questions about domestica-
tion and agricultural origins. It is no
wonder that for more than 100 years
this area of inquiry has held the atten-
tion of archeologists working world-
wide and representing all of archeolo-
gy’s many and rapidly increasing
subdisciplines. It is a research domain
that carries broad currency with
scholars based in biological and phys-
ical sciences, social sciences, and hu-
manities. It is a topic that captures the
imagination of a public interested in
how the familiar world around them
came to be. It is a problem that truly
matters. With an enhanced under-
standing of the nature of the problem
and an expanding array of powerful
tools for studying it, there has never
been a time of greater promise for
pursuing challenging questions about
the origin and diffusion of domesti-
cates and agricultural economies in
virtually all areas of the globe.
REFERENCES
1 Zeder MA, Bradley DG, Emshwiller E, Smith
BD, editors. 2006. Documenting domestication:
new genetic and archaeological paradigms.
Berkeley: University of California Press.
2 Clutton-Brock J. 1994. The unnatural world:
behavioural aspects of humans and animals in
the process of domestication. In: Manning A, Ser-
pell JA, editors. Animals and human society:
changing perspectives. London: Routledge. p 23–
35.
3 Ducos P. 1978. “Domestication” defined and
methodological approaches to its recognition in
faunal assemblages. In: Meadow RH, Zeder MA,
editors. Approaches to faunal analysis in the
Middle East. Peabody Museum Bulletin 2. Cam-
bridge, MA: Harvard University. p 49 –52.
4 Hemmer H. 1990. Domestication: the decline
of environmental appreciation. Cambridge: Cam-
bridge University Press.
5 Bökönyi S. 1989. Definitions of domestication.
In: Clutton-Brock J, editor. The walking larder:
patterns of domestication, pastoralism, and pre-
dation. Cambridge: Unwin. p 1– 4.
6 Ingold T. 1996. Growing plants and raising an-
imals: an anthropological perspective on domes-
tication. In: Harris D, editor. The origins and
spread of agriculture and pastoralism in Eurasia.
Washington, DC: Smithsonian Institution Press.
p 12–24.
7 O’Connor TP. 1997. Working at relationships:
another look at animal domestication. Antiquity
71:149 –156.
8 Moray D. 1994. The early evolution of the do-
mestic dog. Am Sci 82:336 –347.
9 Rindos D. 1984. The origins of agriculture: an
evolutionary perspective. Orlando: Academic
Press.
10 Harris DR. 1996. Introduction: themes and
concepts in the study of early agriculture. In:
Harris DR, editor. The origins and spread of ag-
riculture and pastoralism in Eurasia. Washing-
ton: Smithsonian Institution Press. p 1–11.
11 Uerpmann HP. 1996. Animal domestication:
accident or intention? In: Harris DR, editor. The
origins and spread of agriculture and pastoral-
ism in Eurasia. Washington: Smithsonian Insti-
tution Press. p 227–237.
12 Zeder MA. 2006. A critical examination of
markers of initial domestication in goats (Capra
hircus). In: Zeder MA, Bradley DG, Emshwiller
E, Smith BD, editors. Documenting domestica-
tion: new genetic and archeological paradigms.
Berkeley: University of California Press. p 181–
208.
13 Jarman MR. 1976. Early animal husbandry.
Philos Trans R Soc London B 275:85–97.
14 Hecker H. 1982. Domestication revisited: its
implications for faunal analysis. J Field Archeol
9:217–236.
15 Smith BD. 2006. Documenting domestic
ARTICLES Domestication of Plants and Animals 115
plants in the archaeological record. In: Zeder
MA, Bradley DG, Emshwiller E, Smith BD, edi-
tors. Documenting domestication: new genetic
and archaeological paradigms. Berkeley: Univer-
sity of California Press. p 15–24.
16 Smith BD. 2001. Low level food production. J
Archeol Res 9:1–43.
17 Schultz TR, Mueller UG, Currie CR, Rehner
AH. 2005. Reciprocal illumination: a comparison
of agriculture in humans and in fungus-growing
ants. In: Vega F, Balckwell M, editors. Ecological
and evolutionary advances in insect-fungal asso-
ciations. New York: Oxford University Press. p
149 –190.
18 Anderson MK. 2005. Tending the wild: Native
American knowledge and the management of
California’s natural resources. Berkeley: Univer-
sity of California Press.
19 Turner NJ, Duer D. 2005. Keeping it living.
Seattle: University of Washington Press.
20 Zeder MA, Emshwiller E, Smith BD, Bradley
DG. 2006. Documenting domestication: the in-
tersection of genetics and archaeology. Trends
Genet 22:134 –155.
21 Blumler M, Bryne R. 1991. The ecological
genetics of domestication and the origins of ag-
riculture. Curr Anthropol 32:23–54.
22 Harlan JJ. 1973. Comparative evolution of ce-
reals. Evolution 27:311–325.
23 Heiser C. 1988. Aspects of unconscious selec-
tion and the evolution of domesticated plants.
Euphytica 37:77–85.
24 Piperno DR, Pearsall DM. 1998. The origins of
agriculture in the lowland neotropics. San Diego:
Academic Press.
25 Piperno DR. 2006. Phytolith analysis in ar-
chaeology and environmental history. Walnut
Creek: Altamira Press.
26 Denham TP, Haberle SG, Lentfer C, Fullagar
R, Field J, Therin M, Porch N, Winsborough B.
2003. Origins of agriculture at Kuk Swamp in the
highlands of New Guinea. Science 301:189 –193.
27 Piperno DR. 1993. Phytolith and charcoal
records from deep lake cores in the American
tropics. In: Pearsall DM, Piperno DR, editors.
Current research in phytolith analysis: applica-
tions in archaeology and paleoecology. MASCA
Research Papers in Science and Archaeology Vol
10. Philadelphia: The University Museum of Ar-
chaeology and Anthropology, University of Penn-
sylvania. p 58 –71.
28 Colledge S. 2002. Identifying pre-domestica-
tion cultivation using multivariate analysis: pre-
senting the case for quantification. In: Cappers
RTJ, Bottema S, editors. The dawn of farming in
the Near East. Studies in Near Eastern Produc-
tion, Subsistence, and Environment 6. Berlin: Ex
Oriente. p 141–152.
29 Smith BD. 2002. Rivers of change. Washing-
ton, DC: Smithsonian Institution Press.
30 Mdiba C, De Langhe E Vrydaghs L, Doutrele-
pont H, Swennen R, Van Neer W, de Maret P.
2006. Phytolith evidence for the early presence of
domesticated banana. In: Zeder MA, Bradley DG,
Emshwiller E, Smith BD, editors. Documenting
domestication: new genetic and archaeological
paradigms. Berkeley: University of California
Press. p 68 – 81.
31 Zeder MA. 2006. Archaeological approaches
to documenting animal domestication. In: Zeder
MA, Bradley DG, Emshwiller E, Smith BD, edi-
tors. Documenting domestication: new genetic
and archaeological paradigms. Berkeley: Univer-
sity of California Press. p 171–180.
32 Clutton-Brock J. 1999. Domesticated animals,
2nd ed. London: British Museum of Natural His-
tory.
33 Kruska D. 1988. Mammalian domestication
and its effect on brain structure and behavior. In:
Jerison HJ, Jerison I, editors. Intelligence and
evolutionary biology. New York: Springer-Ver-
lag. p 211–250.
34 Moray D. 1992. Size, shape, and development
in the evolution of the domestic dog. J Archaeol
Sci 19:181–204.
35 Ervynck A, Dobney K, Hongo H, Meadow RH.
2002. Born free!: new evidence for the status of
pigs from Çayönü Tepesi, Eastern Anatolia. Pa-
léorient 27:47–73.
36 Shaffer VM, Reed CA. 1972. The co-evolution
of social behavior and cranial morphology in
sheep and goats (Bovidae, Caprini). Fieldiana
Zool 61. Chicago: Field Museum of Natural His-
tory.
37 Meadow RH. 1989. Osteological evidence for
the process of animal domestication. In: Clutton-
Brock J, editor. The walking larder: patterns of
domestication, pastoralism, and predation. Cam-
bridge: Unwin. p 80 –90.
38 Noddle B. 1974. Age of epiphyseal closure in
feral and domestic goats and ages of dental erup-
tion. J Archaeol Sci 1:195–204.
39 Uerpmann H-P. 1978. Metrical analysis of
faunal remains from the Middle East. In:
Meadow RH, Zeder MA, editors. Approaches to
faunal analysis in the Middle East. Peabody Mu-
seum Bulletin 2. Cambridge, MA: Harvard Uni-
versity. p 41– 45.
40 Zohary D, Tchernov E, Horwitz LK. 1998. The
role of unconscious selection in the domestica-
tion of sheep and goats. J Zoo 245:129 –135.
41 Hole F, Flannery KV, Neely JA. 1969. Prehis-
tory and human ecology on the Deh Luran Plain.
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx
Journal of Archaeological Science 32 (2005) 1408e1416http.docx

More Related Content

Similar to Journal of Archaeological Science 32 (2005) 1408e1416http.docx

Caves & Castles news 2015: archaeology
Caves & Castles news 2015: archaeologyCaves & Castles news 2015: archaeology
Caves & Castles news 2015: archaeologyBluestone Heights
 
The Paleolithic Age
The Paleolithic AgeThe Paleolithic Age
The Paleolithic Agecckanth
 
Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)
Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)
Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)Shaina Mavreen Villaroza
 
The Worst Mistake In The History Of The Human Raceby Jared.docx
The Worst Mistake In The History Of The Human Raceby Jared.docxThe Worst Mistake In The History Of The Human Raceby Jared.docx
The Worst Mistake In The History Of The Human Raceby Jared.docxchristalgrieg
 
Paleontological evidence of evolution
Paleontological evidence of evolutionPaleontological evidence of evolution
Paleontological evidence of evolutionAftab Badshah
 
EarthAndLife.pdf
EarthAndLife.pdfEarthAndLife.pdf
EarthAndLife.pdfForPrivate
 
1.0 IntroductionAs a strategically positioned country with its e.docx
1.0 IntroductionAs a strategically positioned country with its e.docx1.0 IntroductionAs a strategically positioned country with its e.docx
1.0 IntroductionAs a strategically positioned country with its e.docxjackiewalcutt
 
Evidence For Evolution
Evidence For EvolutionEvidence For Evolution
Evidence For EvolutionMark McGinley
 
11. fossils and creation
11. fossils  and creation11. fossils  and creation
11. fossils and creationAriel Roth
 
Megaliths & Monuments news, 2016
Megaliths & Monuments news, 2016Megaliths & Monuments news, 2016
Megaliths & Monuments news, 2016Bluestone Heights
 
Pearsonarticle
PearsonarticlePearsonarticle
Pearsonarticleruggiejr17
 
Unit 5 prehistory theory
Unit 5 prehistory theoryUnit 5 prehistory theory
Unit 5 prehistory theorysergio.historia
 
Saturn: Roman God of Agriculture & Civilization
Saturn: Roman God of Agriculture & CivilizationSaturn: Roman God of Agriculture & Civilization
Saturn: Roman God of Agriculture & CivilizationVapula
 
PaceOfEvolution.pptx
PaceOfEvolution.pptxPaceOfEvolution.pptx
PaceOfEvolution.pptxTrickAquino
 
Unit 7: Jericho: Endurance in the Fertile Crescent
Unit 7: Jericho: Endurance in the Fertile CrescentUnit 7: Jericho: Endurance in the Fertile Crescent
Unit 7: Jericho: Endurance in the Fertile CrescentBig History Project
 

Similar to Journal of Archaeological Science 32 (2005) 1408e1416http.docx (16)

Caves & Castles news 2015: archaeology
Caves & Castles news 2015: archaeologyCaves & Castles news 2015: archaeology
Caves & Castles news 2015: archaeology
 
The Paleolithic Age
The Paleolithic AgeThe Paleolithic Age
The Paleolithic Age
 
Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)
Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)
Evo lutionary Biology (Bio 110) - INTRODUCTION (What is Evolution, History)
 
The Worst Mistake In The History Of The Human Raceby Jared.docx
The Worst Mistake In The History Of The Human Raceby Jared.docxThe Worst Mistake In The History Of The Human Raceby Jared.docx
The Worst Mistake In The History Of The Human Raceby Jared.docx
 
Paleontological evidence of evolution
Paleontological evidence of evolutionPaleontological evidence of evolution
Paleontological evidence of evolution
 
EarthAndLife.pdf
EarthAndLife.pdfEarthAndLife.pdf
EarthAndLife.pdf
 
1.0 IntroductionAs a strategically positioned country with its e.docx
1.0 IntroductionAs a strategically positioned country with its e.docx1.0 IntroductionAs a strategically positioned country with its e.docx
1.0 IntroductionAs a strategically positioned country with its e.docx
 
Evidence For Evolution
Evidence For EvolutionEvidence For Evolution
Evidence For Evolution
 
11. fossils and creation
11. fossils  and creation11. fossils  and creation
11. fossils and creation
 
Megaliths & Monuments news, 2016
Megaliths & Monuments news, 2016Megaliths & Monuments news, 2016
Megaliths & Monuments news, 2016
 
Pearsonarticle
PearsonarticlePearsonarticle
Pearsonarticle
 
Unit 5 prehistory theory
Unit 5 prehistory theoryUnit 5 prehistory theory
Unit 5 prehistory theory
 
Saturn: Roman God of Agriculture & Civilization
Saturn: Roman God of Agriculture & CivilizationSaturn: Roman God of Agriculture & Civilization
Saturn: Roman God of Agriculture & Civilization
 
2013.SHIS142
2013.SHIS1422013.SHIS142
2013.SHIS142
 
PaceOfEvolution.pptx
PaceOfEvolution.pptxPaceOfEvolution.pptx
PaceOfEvolution.pptx
 
Unit 7: Jericho: Endurance in the Fertile Crescent
Unit 7: Jericho: Endurance in the Fertile CrescentUnit 7: Jericho: Endurance in the Fertile Crescent
Unit 7: Jericho: Endurance in the Fertile Crescent
 

More from priestmanmable

9©iStockphotoThinkstockPlanning for Material and Reso.docx
9©iStockphotoThinkstockPlanning for Material and Reso.docx9©iStockphotoThinkstockPlanning for Material and Reso.docx
9©iStockphotoThinkstockPlanning for Material and Reso.docxpriestmanmable
 
a 12 page paper on how individuals of color would be a more dominant.docx
a 12 page paper on how individuals of color would be a more dominant.docxa 12 page paper on how individuals of color would be a more dominant.docx
a 12 page paper on how individuals of color would be a more dominant.docxpriestmanmable
 
978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx
978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx
978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docxpriestmanmable
 
92 Academic Journal Article Critique  Help with Journal Ar.docx
92 Academic Journal Article Critique  Help with Journal Ar.docx92 Academic Journal Article Critique  Help with Journal Ar.docx
92 Academic Journal Article Critique  Help with Journal Ar.docxpriestmanmable
 
A ) Society perspective90 year old female, Mrs. Ruth, from h.docx
A ) Society perspective90 year old female, Mrs. Ruth, from h.docxA ) Society perspective90 year old female, Mrs. Ruth, from h.docx
A ) Society perspective90 year old female, Mrs. Ruth, from h.docxpriestmanmable
 
9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx
9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx
9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docxpriestmanmable
 
9 AssignmentAssignment Typologies of Sexual AssaultsT.docx
9 AssignmentAssignment Typologies of Sexual AssaultsT.docx9 AssignmentAssignment Typologies of Sexual AssaultsT.docx
9 AssignmentAssignment Typologies of Sexual AssaultsT.docxpriestmanmable
 
9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx
9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx
9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docxpriestmanmable
 
900 BritishJournalofNursing,2013,Vol22,No15©2.docx
900 BritishJournalofNursing,2013,Vol22,No15©2.docx900 BritishJournalofNursing,2013,Vol22,No15©2.docx
900 BritishJournalofNursing,2013,Vol22,No15©2.docxpriestmanmable
 
9 Augustine Confessions (selections) Augustine of Hi.docx
9 Augustine Confessions (selections) Augustine of Hi.docx9 Augustine Confessions (selections) Augustine of Hi.docx
9 Augustine Confessions (selections) Augustine of Hi.docxpriestmanmable
 
8.3 Intercultural CommunicationLearning Objectives1. Define in.docx
8.3 Intercultural CommunicationLearning Objectives1. Define in.docx8.3 Intercultural CommunicationLearning Objectives1. Define in.docx
8.3 Intercultural CommunicationLearning Objectives1. Define in.docxpriestmanmable
 
8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx
8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx
8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docxpriestmanmable
 
8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx
8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx
8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docxpriestmanmable
 
800 Words 42-year-old man presents to ED with 2-day history .docx
800 Words 42-year-old man presents to ED with 2-day history .docx800 Words 42-year-old man presents to ED with 2-day history .docx
800 Words 42-year-old man presents to ED with 2-day history .docxpriestmanmable
 
8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx
8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx
8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docxpriestmanmable
 
8.0 RESEARCH METHODS These guidelines address postgr.docx
8.0  RESEARCH METHODS  These guidelines address postgr.docx8.0  RESEARCH METHODS  These guidelines address postgr.docx
8.0 RESEARCH METHODS These guidelines address postgr.docxpriestmanmable
 
95People of AppalachianHeritageChapter 5KATHLEEN.docx
95People of AppalachianHeritageChapter 5KATHLEEN.docx95People of AppalachianHeritageChapter 5KATHLEEN.docx
95People of AppalachianHeritageChapter 5KATHLEEN.docxpriestmanmable
 
9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx
9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx
9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docxpriestmanmable
 
8-10 slide Powerpoint The example company is Tesla.Instructions.docx
8-10 slide Powerpoint The example company is Tesla.Instructions.docx8-10 slide Powerpoint The example company is Tesla.Instructions.docx
8-10 slide Powerpoint The example company is Tesla.Instructions.docxpriestmanmable
 
8Network Security April 2020FEATUREAre your IT staf.docx
8Network Security  April 2020FEATUREAre your IT staf.docx8Network Security  April 2020FEATUREAre your IT staf.docx
8Network Security April 2020FEATUREAre your IT staf.docxpriestmanmable
 

More from priestmanmable (20)

9©iStockphotoThinkstockPlanning for Material and Reso.docx
9©iStockphotoThinkstockPlanning for Material and Reso.docx9©iStockphotoThinkstockPlanning for Material and Reso.docx
9©iStockphotoThinkstockPlanning for Material and Reso.docx
 
a 12 page paper on how individuals of color would be a more dominant.docx
a 12 page paper on how individuals of color would be a more dominant.docxa 12 page paper on how individuals of color would be a more dominant.docx
a 12 page paper on how individuals of color would be a more dominant.docx
 
978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx
978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx
978-1-5386-6589-318$31.00 ©2018 IEEE COSO Framework for .docx
 
92 Academic Journal Article Critique  Help with Journal Ar.docx
92 Academic Journal Article Critique  Help with Journal Ar.docx92 Academic Journal Article Critique  Help with Journal Ar.docx
92 Academic Journal Article Critique  Help with Journal Ar.docx
 
A ) Society perspective90 year old female, Mrs. Ruth, from h.docx
A ) Society perspective90 year old female, Mrs. Ruth, from h.docxA ) Society perspective90 year old female, Mrs. Ruth, from h.docx
A ) Society perspective90 year old female, Mrs. Ruth, from h.docx
 
9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx
9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx
9 dissuasion question Bartol, C. R., & Bartol, A. M. (2017)..docx
 
9 AssignmentAssignment Typologies of Sexual AssaultsT.docx
9 AssignmentAssignment Typologies of Sexual AssaultsT.docx9 AssignmentAssignment Typologies of Sexual AssaultsT.docx
9 AssignmentAssignment Typologies of Sexual AssaultsT.docx
 
9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx
9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx
9 0 0 0 09 7 8 0 1 3 4 4 7 7 4 0 4ISBN-13 978-0-13-44.docx
 
900 BritishJournalofNursing,2013,Vol22,No15©2.docx
900 BritishJournalofNursing,2013,Vol22,No15©2.docx900 BritishJournalofNursing,2013,Vol22,No15©2.docx
900 BritishJournalofNursing,2013,Vol22,No15©2.docx
 
9 Augustine Confessions (selections) Augustine of Hi.docx
9 Augustine Confessions (selections) Augustine of Hi.docx9 Augustine Confessions (selections) Augustine of Hi.docx
9 Augustine Confessions (selections) Augustine of Hi.docx
 
8.3 Intercultural CommunicationLearning Objectives1. Define in.docx
8.3 Intercultural CommunicationLearning Objectives1. Define in.docx8.3 Intercultural CommunicationLearning Objectives1. Define in.docx
8.3 Intercultural CommunicationLearning Objectives1. Define in.docx
 
8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx
8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx
8413 906 AMLife in a Toxic Country - NYTimes.comPage 1 .docx
 
8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx
8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx
8. A 2 x 2 Experimental Design - Quality and Economy (x1 and x2.docx
 
800 Words 42-year-old man presents to ED with 2-day history .docx
800 Words 42-year-old man presents to ED with 2-day history .docx800 Words 42-year-old man presents to ED with 2-day history .docx
800 Words 42-year-old man presents to ED with 2-day history .docx
 
8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx
8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx
8.1 What Is Corporate StrategyLO 8-1Define corporate strategy.docx
 
8.0 RESEARCH METHODS These guidelines address postgr.docx
8.0  RESEARCH METHODS  These guidelines address postgr.docx8.0  RESEARCH METHODS  These guidelines address postgr.docx
8.0 RESEARCH METHODS These guidelines address postgr.docx
 
95People of AppalachianHeritageChapter 5KATHLEEN.docx
95People of AppalachianHeritageChapter 5KATHLEEN.docx95People of AppalachianHeritageChapter 5KATHLEEN.docx
95People of AppalachianHeritageChapter 5KATHLEEN.docx
 
9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx
9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx
9 781292 041452ISBN 978-1-29204-145-2Forensic Science.docx
 
8-10 slide Powerpoint The example company is Tesla.Instructions.docx
8-10 slide Powerpoint The example company is Tesla.Instructions.docx8-10 slide Powerpoint The example company is Tesla.Instructions.docx
8-10 slide Powerpoint The example company is Tesla.Instructions.docx
 
8Network Security April 2020FEATUREAre your IT staf.docx
8Network Security  April 2020FEATUREAre your IT staf.docx8Network Security  April 2020FEATUREAre your IT staf.docx
8Network Security April 2020FEATUREAre your IT staf.docx
 

Recently uploaded

SOCIAL AND HISTORICAL CONTEXT - LFTVD.pptx
SOCIAL AND HISTORICAL CONTEXT - LFTVD.pptxSOCIAL AND HISTORICAL CONTEXT - LFTVD.pptx
SOCIAL AND HISTORICAL CONTEXT - LFTVD.pptxiammrhaywood
 
Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...
Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...
Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...fonyou31
 
Introduction to Nonprofit Accounting: The Basics
Introduction to Nonprofit Accounting: The BasicsIntroduction to Nonprofit Accounting: The Basics
Introduction to Nonprofit Accounting: The BasicsTechSoup
 
Accessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactAccessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactdawncurless
 
Interactive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communicationInteractive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communicationnomboosow
 
Arihant handbook biology for class 11 .pdf
Arihant handbook biology for class 11 .pdfArihant handbook biology for class 11 .pdf
Arihant handbook biology for class 11 .pdfchloefrazer622
 
CARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxCARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxGaneshChakor2
 
mini mental status format.docx
mini    mental       status     format.docxmini    mental       status     format.docx
mini mental status format.docxPoojaSen20
 
Call Girls in Dwarka Mor Delhi Contact Us 9654467111
Call Girls in Dwarka Mor Delhi Contact Us 9654467111Call Girls in Dwarka Mor Delhi Contact Us 9654467111
Call Girls in Dwarka Mor Delhi Contact Us 9654467111Sapana Sha
 
Software Engineering Methodologies (overview)
Software Engineering Methodologies (overview)Software Engineering Methodologies (overview)
Software Engineering Methodologies (overview)eniolaolutunde
 
Web & Social Media Analytics Previous Year Question Paper.pdf
Web & Social Media Analytics Previous Year Question Paper.pdfWeb & Social Media Analytics Previous Year Question Paper.pdf
Web & Social Media Analytics Previous Year Question Paper.pdfJayanti Pande
 
Sports & Fitness Value Added Course FY..
Sports & Fitness Value Added Course FY..Sports & Fitness Value Added Course FY..
Sports & Fitness Value Added Course FY..Disha Kariya
 
The byproduct of sericulture in different industries.pptx
The byproduct of sericulture in different industries.pptxThe byproduct of sericulture in different industries.pptx
The byproduct of sericulture in different industries.pptxShobhayan Kirtania
 
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...EduSkills OECD
 
APM Welcome, APM North West Network Conference, Synergies Across Sectors
APM Welcome, APM North West Network Conference, Synergies Across SectorsAPM Welcome, APM North West Network Conference, Synergies Across Sectors
APM Welcome, APM North West Network Conference, Synergies Across SectorsAssociation for Project Management
 
BASLIQ CURRENT LOOKBOOK LOOKBOOK(1) (1).pdf
BASLIQ CURRENT LOOKBOOK  LOOKBOOK(1) (1).pdfBASLIQ CURRENT LOOKBOOK  LOOKBOOK(1) (1).pdf
BASLIQ CURRENT LOOKBOOK LOOKBOOK(1) (1).pdfSoniaTolstoy
 
Student login on Anyboli platform.helpin
Student login on Anyboli platform.helpinStudent login on Anyboli platform.helpin
Student login on Anyboli platform.helpinRaunakKeshri1
 
The basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptxThe basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptxheathfieldcps1
 

Recently uploaded (20)

SOCIAL AND HISTORICAL CONTEXT - LFTVD.pptx
SOCIAL AND HISTORICAL CONTEXT - LFTVD.pptxSOCIAL AND HISTORICAL CONTEXT - LFTVD.pptx
SOCIAL AND HISTORICAL CONTEXT - LFTVD.pptx
 
Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...
Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...
Ecosystem Interactions Class Discussion Presentation in Blue Green Lined Styl...
 
Introduction to Nonprofit Accounting: The Basics
Introduction to Nonprofit Accounting: The BasicsIntroduction to Nonprofit Accounting: The Basics
Introduction to Nonprofit Accounting: The Basics
 
Accessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impactAccessible design: Minimum effort, maximum impact
Accessible design: Minimum effort, maximum impact
 
Interactive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communicationInteractive Powerpoint_How to Master effective communication
Interactive Powerpoint_How to Master effective communication
 
Arihant handbook biology for class 11 .pdf
Arihant handbook biology for class 11 .pdfArihant handbook biology for class 11 .pdf
Arihant handbook biology for class 11 .pdf
 
CARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptxCARE OF CHILD IN INCUBATOR..........pptx
CARE OF CHILD IN INCUBATOR..........pptx
 
Mattingly "AI & Prompt Design: Structured Data, Assistants, & RAG"
Mattingly "AI & Prompt Design: Structured Data, Assistants, & RAG"Mattingly "AI & Prompt Design: Structured Data, Assistants, & RAG"
Mattingly "AI & Prompt Design: Structured Data, Assistants, & RAG"
 
mini mental status format.docx
mini    mental       status     format.docxmini    mental       status     format.docx
mini mental status format.docx
 
Call Girls in Dwarka Mor Delhi Contact Us 9654467111
Call Girls in Dwarka Mor Delhi Contact Us 9654467111Call Girls in Dwarka Mor Delhi Contact Us 9654467111
Call Girls in Dwarka Mor Delhi Contact Us 9654467111
 
Software Engineering Methodologies (overview)
Software Engineering Methodologies (overview)Software Engineering Methodologies (overview)
Software Engineering Methodologies (overview)
 
Web & Social Media Analytics Previous Year Question Paper.pdf
Web & Social Media Analytics Previous Year Question Paper.pdfWeb & Social Media Analytics Previous Year Question Paper.pdf
Web & Social Media Analytics Previous Year Question Paper.pdf
 
Sports & Fitness Value Added Course FY..
Sports & Fitness Value Added Course FY..Sports & Fitness Value Added Course FY..
Sports & Fitness Value Added Course FY..
 
The byproduct of sericulture in different industries.pptx
The byproduct of sericulture in different industries.pptxThe byproduct of sericulture in different industries.pptx
The byproduct of sericulture in different industries.pptx
 
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
Presentation by Andreas Schleicher Tackling the School Absenteeism Crisis 30 ...
 
APM Welcome, APM North West Network Conference, Synergies Across Sectors
APM Welcome, APM North West Network Conference, Synergies Across SectorsAPM Welcome, APM North West Network Conference, Synergies Across Sectors
APM Welcome, APM North West Network Conference, Synergies Across Sectors
 
BASLIQ CURRENT LOOKBOOK LOOKBOOK(1) (1).pdf
BASLIQ CURRENT LOOKBOOK  LOOKBOOK(1) (1).pdfBASLIQ CURRENT LOOKBOOK  LOOKBOOK(1) (1).pdf
BASLIQ CURRENT LOOKBOOK LOOKBOOK(1) (1).pdf
 
Student login on Anyboli platform.helpin
Student login on Anyboli platform.helpinStudent login on Anyboli platform.helpin
Student login on Anyboli platform.helpin
 
The basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptxThe basics of sentences session 2pptx copy.pptx
The basics of sentences session 2pptx copy.pptx
 
Mattingly "AI & Prompt Design: The Basics of Prompt Design"
Mattingly "AI & Prompt Design: The Basics of Prompt Design"Mattingly "AI & Prompt Design: The Basics of Prompt Design"
Mattingly "AI & Prompt Design: The Basics of Prompt Design"
 

Journal of Archaeological Science 32 (2005) 1408e1416http.docx

  • 1. Journal of Archaeological Science 32 (2005) 1408e1416 http://www.elsevier.com/locate/jas Why domesticate food animals? Some zoo-archaeological evidence from the Levant Simon J.M. Davis * Instituto Português de Arqueologia, Avenida da Índia 136, P- 1300-300 Lisbon, Portugal Received 4 February 2005; received in revised form 10 March 2005 Abstract Zoo-archaeological remains from the southern Levant indicate two shifts in the pattern of animal exploitation from Palaeolithic to Pre-Pottery Neolithic times. These shifts were especially marked towards the end of this time span. One is the increased consumption of small animals and the other shift is an increased hunting of juvenile gazelles compared to adults. Both are interpreted in terms of an increased intensity of exploitation of environmental resources due, it is suggested, to population increase, which subsequently forced people to husband animals. � 2005 Elsevier Ltd. All rights reserved. Keywords: Domestication; Population; Neolithic; Demography; Levant; Near East
  • 2. ‘‘.the true cause that set in motion the great tide of northern emigration, and that continued to propel it till it rolled at different periods against China, Persia, Italy, and even Egypt, was a scarcity of food, a population extended beyond the means of supporting it.’’ [33] 1. Introduction Why did our ancestors domesticate food animals? For a long time, it was believed that the transition from hunting to husbandry was a move from a precarious existence to one providing greater security. The notion that hunting requires high expenditure of energy and that a major saving of effort can be gained by switching to farming is probably wrong. Pioneering work by * Tel.: C351 21 361 6557; fax: C351 01 361 6559. E-mail address: [email protected] 0305-4403/$ - see front matter � 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.jas.2005.03.018 Richard Lee [30], who studied modern hunteregatherers in southern Africa, discovered that they enjoy a plentiful and balanced diet and spend a mere 2e3 days per week in their quest for food. Husbanding animals is, it would seem, more arduous than simply going out and hunting them. The view taken here is a more gradualist one in which a slow shift in the balance between people and their source of food has occurred. Clearly, in order to try and understand the background to domestication in the archaeological record, we need to clarify what happened during the millennia that preceded this change. One of the first studies that attempted to answer the question why people began domesticating plants and animals was Mark Cohen’s The Food Crisis in Prehistory [6]. Cohen suggested that the period prior to domes-
  • 3. tication was characterized by increased strain on the environment due to rising demographic pressure. Amongst the evidence he cited is an increase in the occurrences of pathological conditions in human skel- etal remains from archaeological sites e probably reflecting increasingly poor nutrition. He suggested that a rise in the human population at that time ‘‘forced’’ mailto:[email protected] http://www.elseiver.com/locate/jas 1409S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 people to change their relationship with the natural world and to assume a greater degree of control over it. Others too have linked domestication indirectly or directly to demographic pressure and the carrying capacity of the environment [21,43], and these factors combined with a trigger e deteriorating climate at the end of the last Ice Age [17]. But not all agree that population growth was a factor (see, for example, [41]), indeed it is quite likely that the causes varied depending upon which part of the world is considered. Hayden [25] has suggested a ‘‘socio-economic’’ explanation for the origin of domestication; domesticates were initially used as prestige commodities. These served to ‘‘enhance individual or corporate group power’’ in complex hunteregatherer communities. Others like Isaac [27] and Cauvin [5] have sought a more religious/psycholog- ical explanation for the beginnings of domestication. Before addressing the question of why animals were domesticated, we need to consider briefly where and when animal husbandry began. We now know of some
  • 4. nine independent areas where food production began. The ‘‘Fertile Crescent’’ of the Near East was un- doubtedly the earliest Old World hearth of domesti- cation of both animals and plants, although it is still unclear exactly where within this area it began [2,17,23,24,49,50]. Evidence from Cyprus and the southern Levant now indicates that bovids (cattle, sheep and goats) and pigs were first domesticated towards the end of the 9th millennium cal BC in the northern part of the Near East, and that during the 8th millennium BC, these animals were domesticated (or introduced as domestic stock) in the southern Levant [8,22]. Clearly, the 9th and 8th millennia, during the so-called Pre- Pottery Neolithic, were crucial times for the advent of husbandry; hence, we need to examine the relations between people and animals in the period before those millennia in order to answer why people began husbanding animals. A possible zoo-archaeological answer to this major question first became evident to me when studying the fauna from the Natufian e Aceramic Neolithic site of Hatoula, in Israel, in the mid-1980s. It is especially gratifying to note that the last two decades have seen other zoo-archaeologists interpret their data in a similar way e measurements of bones and shells, and consid- erations of species’ frequencies can reflect pressure on the environment in turn linked to demographic pressure. Most noteworthy for the Levant is the new study of the fauna from Hayonim and nearby sites by Natalie Munro [36,37]. She too suggests that population in- creased quite markedly in the Natufian. This paper, like earlier reports [11,12,14,15], considers other studies, as well as several from other regions. They have been and still are interpreted in a similar way. By including the faunal evidence from subsequent periods I hope to
  • 5. present a broader view of this major event. I shall present two lines of evidence from the zoo- archaeological record of the Mousterian to the Pre- Pottery Neolithic in the Levant which corroborate the demographic pressure thesis. The first is a shift of people’s prey in the course of time from big game to smaller mammals and ultimately fish and birds e which Flannery [19] recognised as a shift from a narrow spectrum of environmental resources to a broader one. Indeed, this ‘‘spectrum shift’’ is well attested in many regions where people ‘‘were forced to become even more eclectic in their food gathering, to eat more and more unpalatable foods, and in particular to concentrate on foods of low trophic level and high density’’ [6]. The second line of evidence has not, to my knowledge, been considered very much by zoo-archaeologists and is an increase in the proportion of juvenile prey e here gazelle e in these faunal assemblages. This I term the ‘‘age shift’’. (Legge [31] had observed an increase in numbers of juvenile gazelles at Mount Carmel. He suggested that this may signify their domestic status, a proposal he subsequently rejected; pers. comm. The gazelle is not generally considered to be amenable to domestication.) Both these changes occurred over a long time span from the Mousterian, some 50,000 years ago, to the Pre- Pottery Neolithic A of the 9th millennium BC. Moreover, these changes were even more rapid during the two or three millennia before domestic animals first made their appearance in the 8th millennium BC [11,15]. And, as I shall argue, the interpretation of these changes helps to explain why people had, to a large extent, to abandon hunting in favour of husbandry. 2. Material and methods This article discusses two kinds of zoo-archaeological
  • 6. data from the Upper Pleistocene and early Holocene. Most come from Israel. They are counts of: (a) Bones of different species of animals and (b) Numbers of young versus adult gazelles. The data presented here all derive from sites whose faunal assemblages were recovered by sieving (Fig. 1). Hence, biases against smaller bones such as unfused epiphyses that belonged to young animals are probably less serious than is often the case with hand-collected material [38]. Another problem sometimes encountered in zoo-archaeology is observer variation. Zoo-archae- ologists may count bones and deduce the age-at-death of animals in different ways. One person studied all the assemblages discussed here using the same methods [15]. In brief, mandibles and a restricted suite of ‘‘parts of the skeleton always recorded ’’ were counted. They consist of a predetermined suite of articular ends/epiphyses of girdle, limb and foot bones. This in part explains why the 1410 S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 Fig. 1. Map of the southern Levant to show places mentioned in the text. The dotted line represents the present-day 200 mm isohyet. sample sizes in many cases appear to be small. Since birds no longer have teeth and fish lack limbs, and in order to compare frequencies of mammals with those of birds and fish at one of the key sites considered here (Hatoula), the numbers of mammalian mandibles were compared with the minimum numbers of individuals of fish and birds. (This is admittedly not entirely satisfactory and has probably led to an under-representation of fish.)
  • 7. The growing end of a mammal limb-bone is its epiphysis, which, at maturity, fuses to its respective shaft. For large assemblages of a particular species, an approximate estimate of the proportion of juveniles (that is osteologically immature) culled may be calcu- lated by counting separately the unfused and fused epiphyses. A similar approach was adopted for man- dibles. Premolar teeth replace their milk predecessors at a given age. In the case of the gazelle this occurs around 12e13 months, which is also the age when the third molar tooth erupts, and this is also very approximately the age when many of the limb-bones fuse [7]. Hence, it is possible to calculate the proportion of juveniles via the numbers of mandibles with milk teeth or unerupted third molars. 3. The ‘‘Spectrum Shift’’ (Figs. 2 and 3) At the sequence of open-air sites, which extends from the Mousterian on the Golan Heights (Biqat Quneitra) to the late Natufian on the eastern side of the Sea of Galilee (Nahal Ein Gev II), we [14] observed a change from big game to gazelles, hares and foxes. Many of the animals hunted at Biqat Quneitra were very large and included species like rhinoceros, aurochs, red deer and horses. As time progressed, remains of these very large beasts became rare on archaeological sites and fallow deer and gazelle were clearly the most important animals exploited in Aurignacian and Kebaran times (i.e. 1411S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 Fig. 2. The spectrum shift as seen in the series of open-air sites
  • 8. from the Mousterian at Biqat Quneitra to the Natufian at Nahal Ein Gev. Note the gradual shift with time from large game to gazelles. Key: BQ, Biqat Quneitra (Mousterian c. 54,000 bp); NEG I, Nahal Ein Gev I (Late Upper Palaeolithic c. 25e20,000 bp); EG I, Ein Gev I (Kebaran c. 16,000 bp); EG III, Ein Gev III (Geometric Kebaran c. 16,500e14,500 bp); EG IV, Ein Gev IV (Late Geometric Kebaran c. 14,500e12,750 bp); NEG II, Nahal Ein Gev II (Late Natufian/Khiamian c. 10,500e10,000 bp). The numbers in parentheses are the counts of identified and recorded bones. between approximately 30,000 and 15,000 BC). Fallow deer appear to have become even scarcer as we approach the late Natufian when the most common medium-sized mammal represented is gazelle. At about this time, during the Epipalaeolithic, increasing numbers of small mammals such as fox and hare are also represented. Let us look now at what happened at a site in another part of this region e Hatoula, in central Israel e with its 1412 S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 Fig. 3. The spectrum shift as seen in the NatufianePPNA (Khiamian and Sultanian) sequence at Hatoula. Percentages of the different groups of animals, computed from mandible counts for the mammals, and minimum numbers of individuals for the fish and birds. Small mammals include
  • 9. hedgehog, hare, polecat, marten, badger, fox and cat only. Note the increase of small mammals, fish and birds in the PPNA. sequence from Natufian to PPNA (the latter subdivided into Khiamian and Sultanian levels). Hatoula is located on the western slopes of the Judean Hills roughly midway between Jerusalem and Tel Aviv. The samples of animal bones recovered at this site are quite large, with 2540 bones recorded in the Natufian levels, and 1016 in the PPNA levels. Here the continued ‘‘down- sizing’’ of animals represented is quite remarkable. Note the shift between the Natufian and PPNA levels, from gazelle to small mammals, fish and birds. It is tempting to regard this shift at Hatoula as the culmination of a long and more gradual move to smaller taxa, which had been happening since Mousterian times. Whether this shift reflects local habitat degradation or a change in hunting techniques (or both) is difficult to determine. However, it is worth noting that although Fig. 4. The age shift of gazelles culled in Israel between Mousterian and PPN times. This is a plot of the percentages of unfused limb-bone epiphyses (distal radius, distal metacarpal, distal metatarsal, distal femur, distal tibia, and calcaneum-tuber calcis) plotted against time. Total includes these unfused epiphyses plus specimens of the same bone-parts with fused epiphyses. Data come from the following sites, from left to right: Mousterian: Kebara cave; Mousterian: Hayonim cave level E; Upper Palaeolithic: Kebara cave; Aurignacian: Hayonim cave level D; Kebaran: Ein Gev I; Kebaran: Hayonim cave level C; Natufian: Hayonim cave level
  • 10. B; Natufian: Hayonim terrace; Natufian: Hatoula; Khiamian (PPNA): Hatoula; Sultanian (PPNA): Hatoula. Data for Hatoula are in refs. [10,15]. Data for other sites are in ref. [9]. 1413S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 increasingly scarce in zoo-archaeological assemblages, the continuing presence of many very large animals in the Levant well after their apparent demise at Ein Gev makes it seem unlikely that the environment changed to any very great extent. For example, rhinoceros and red deer are known in the Kebaran (at Hayonim), hippo- potamus is known in the Iron Age (at Tel Qassile), and the aurochs survived into the Early Bronze Age (at Tel Yarmouth; [8,13]). A shift in emphasis from very large animals to small mammals, birds and fish must have required the invention and development of new hunting techniques, but this was surely brought about by the increasing need to rely upon such smaller sources of sustenance. Cohen [6], like Flannery [19], has amassed a wide range of evidence, which shows that in many regions just before the advent of animal and plant husbandry people began to expand their resource base. This included many smaller species of animals whose capture required a higher expenditure of energy for a given unit of flesh procured. On the limestone steppe in Jordan, Martin [34] studied a succession of small open-air sites, where she noted that hare increased dramatically in the early and middle PPNB levels (at Jilat). In the Jordan valley, Tchernov [48] too reported large quantities of hare and
  • 11. bird bones at the PPNA site of Netiv Hagdud. Elsewhere in the Mediterranean, others have observed a shift to smaller species of animals. For example at the caves of Franchthi in Greece and Nerja in Andalusia, Payne [39] and Morales et al. [35] found that fishing began in the Mesolithic and Magdalenian periods, respectively. Morales et al. [35] call this shift to ex- ploiting marine resources the ‘‘Tardiglacial paradigm’’. The increased exploitation of marine resources and occurrences of shell middens in the archaeological record may to some extent be due to the encroachment of the seashore as sea levels rose after the Pleistocene [1]. Besides small game, birds and marine resources, Lubell [32] has recently pointed out that remains of land snails too become more abundant just prior to the advent of agriculture in the circum Mediterranean region. The Table 1 The age-at-death of gazelles from late PleistoceneeHolocene Israel deduced from tooth eruption Period Site Mandibles with M3 ‘‘Unerupted’’ Mandibles with M3 ‘‘Erupted’’ %
  • 12. Juvenile PPNA Hatoula 14 27 34 Late Natufian Hatoula 8 41 16 Natufian Hayonim terrace 27 135 17 Aurignacian Hayonim cave D 13 115 10 Mousterian Kebara cave 7 251 3 Numbers of mandibles with unerupted third molars (i.e. less than c. 13 months of age) and erupted M3s from various large zoo- archaeological assemblages (see refs. [9,15]). most spectacular examples are the Capsian ‘‘escar- gotières’’ of Algeria and Tunisia. He suggests ‘‘the presence of abundant land snails represents part of a signature for the broad spectrum revolution’’. Stiner [44] has provided further evidence for a broadening of the spectrum of exploited species in the Mediterranean basin. Presumably big game had become scarce, but why? One explanation, and the one offered here, is simply that there were now many more mouths to feed; people were forced to exploit small game, fowl and fish in order to survive. 4. The ‘‘Age Shift’’ (Fig. 4 and Table 1) The numbers of unfused (juvenile) limb-bone epiph- yses compared to the fused (adult) ones in the sequence of Israeli zoo-archaeological assemblages indicate that
  • 13. some 20e25% of the gazelles hunted before the Neo- lithic were osteologically immature. Subsequently, in the PPNA, this percentage increased markedly to around 35e40%. The counts of mandibles with unerupted and erupted third molars indicate a similar trend in the course of time (Table 1). At Hatoula, the presence of milk or permanent teeth in the anterior part of the gazelle mandibles shows that the proportion of juveniles increased from 39% in the Natufian (13 have dP4s and 20 have P4s) to 59% in the PPNA (19 have dP4s but only 13 have P4s). In sum then, both limb-bones and mandibles indicate a substantial chronological trend towards increased culling of juvenile gazelles. There are several interesting parallel examples from zoo-archaeological successions elsewhere. Perhaps of some relevance here, and certainly the inspiration for my interpretation of the Israeli data, is Elder’s [18] study of deer mandibles from three ‘prehistoric’ (i.e. pre-Euro- pean settlement) and two ‘historic’ Indian sites in Missouri, USA. At his ‘prehistoric’ sites, the age distributions of the culled deer showed moderate numbers of old and senile animals as well as young. Elder suggested that this pattern is similar to that observed in a stable population of ungulates. However, most of the mandibles from the two ‘historic’ or ‘post- contact’ sites dated between AD 1725 and AD 1780, derived from young individuals, with very few old and senile ones represented. This kind of age distribution is typical of a population undergoing rapid turnover caused by heavy predation. Elder suggested that a lucrative venison trade and more efficient hunting with firearms and horses (both the result of the arrival of European settlers) were the two main factors which caused this change in the age-profiles of Missouri deer. Three other studies are worth quoting. Between Middle
  • 14. and Late Stone Ages in South Africa, the limpets collected by people became smaller (i.e. younger), which Klein and colleagues [28,29] correlates with increasingly 1414 S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 heavy foraging. In northern Israel, Stiner et al. [45] recorded a size reduction of tortoises’ limb-bone shafts (the girth of the shaft in reptile bones continues to increase with age) in the Epipalaeolithic at Hayonim, western Galilee, which they suggest, was due to an in- crease in the predation pressure on tortoises, in turn a reflection of human population increase. Surovell [46] simulated what can happen to fast reproducing species of small game like hares and partridges, and slow ones like the tortoise, when predation pressure rises due to increasing occupation density and population. He found that the faster reproducing species will increase in number in the archaeological record while slower reproducers will decrease and even become extinct. 5. Conclusion What then do the changes in the zoo-archaeological record of the Levant mean? My interpretation of the gazelle age shift is similar to that which Elder proposed to explain his age shift in Missouri deer. What we observe in the Israeli sequence is a gradual increase in the intensity with which gazelles were hunted. This hunting intensity increased at an even greater rate in the PPNA as can be seen at Hatoula, and appears to have occurred alongside a shift towards exploiting smaller animals, and in the Natufian and PPNA, more extensive exploitation of marine and avian resources. I suggest
  • 15. that it was the same factor, an increasing imbalance between predator, that is man-the-hunter, and prey, which was responsible for both of these trends. Natalie Munro has restudied the Natufian fauna of Hayonim and two other more recently excavated Natufian sites. It is gratifying to note that she comes to similar conclusions to the ones suggested here. In concluding her thesis, Munro [36] writes ‘‘.in the Early Natufian phase, the southern Levant likely supported the densest and.probably the highest gross population sizes the region had seen to this point.’’ While it is possible to view population levels as being a result of food resource availability, Cohen [6] and Rowley-Conwy [42] see population levels as the driving force (at least in the Neolithic) behind the need to change from hunting to husbanding. In my opinion it was this factor e the ratio of human beings to wildlife e that caused Neolithic people to adopt husbandry. The human population of the Levant had undoubtedly been rising steadily since the Palaeolithic. With the adoption of sedentism in the Natufian as first suggested by Perrot [40], population increase must have become much more rapid. We know from modern examples of newly settled nomads, that sedentism leads to rapid population increase [47]. Goring-Morris’s [20] archaeological survey of the Negev desert has shown a decrease in the number of PPNA sites there as compared with the Natufian. He attributes this ‘‘virtual abandonment’’ of the Negev to aridification at that time, and suggests that people moved northwards into the more fertile Mediterranean zone which may have exacerbated an already precarious balance between people and their resources. Henry [26] cites archaeological data from that region such as site size and artefact densities that point to
  • 16. possible Natufian population increase. In their studies on the Natufian and PPNA, Belfer-Cohen and Bar- Yosef [4] note certain features on human skeletal size such as a reduction in the difference in male/female stature, as well as increasing site size that point to stress and demographic pressure at this time e just prior to the adoption of animal husbandry. Population density may have increased from 0.25 persons per km 2 in the Natufian to 1 person per km 2 in the PPNA and then 4 persons per km 2 in the PPNB [3]. In the southern Levant, probably during the PPNB, human population levels continued to rise rapidly and reached some critical threshold. The carrying capacity of the environment was exceeded e the ever-increasing number of human mouths could no longer be adequately fed. Gazelle and big-game stocks had become too scarce and fishing and fowling did not suffice either. It was then, during the PPNB, that people had to exert greater control over environmental resources and begin husband- ing those species that were amenable to this kind of treatment such as cattle, sheep, goats and pigs e a strategy in which a given area can sustain many more herders and farmers than hunteregatherers. (The switch to husbanding bovids and pigs in the course of the PPNB raises another question. Were these species, all present but admittedly not very common on Natufian and PPNA
  • 17. sites in the Levant, domesticated locally or introduced from other parts of the Near East? Given the imprecision of dates available it is probably still too early to answer this question with any degree of security. Their in- troduction from beyond the Levant as already managed stock is of course a strong possibility, and could help explain why sheep and goat were not hunted in increasing numbers in the Natufian and PPNA.) Perhaps it was the climate changes of the end of the Pleistocene, as Diamond [17] suggests, that acted as the last straw that ‘‘broke the camel’s back’’. People could no longer survive by hunting alone e they were now banished from the Garden of Eden: animal husbandry is hard work and domesticated animals require daily care and supervision. But with these four-legged resources under human control, people were able to become ever more ‘‘fruitful and multiply’’. The human population in the Near East rose even more rapidly. The stage was now set for further economic change, the spread of farming peoples across the world and all kinds of developments both welcome and unwelcome [16]. 1415S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 Acknowledgements I am very grateful to Ofer Bar Yosef, the late Eitan Tchernov, Naama Goren-Inbar, Donald Henry, Jean Perrot, Monique Lechevallier and Avraham Ronen who at various times have invited me to study zoo- archaeological remains from their sites. It was a study of the fauna of Hatoula that led me to the ideas put
  • 18. forward here. I have had numerous useful discussions with colleagues both in Israel and England, and both Cathy Douzil and Umberto Albarella have encouraged me to write this paper and offered useful comments. I thank Nigel Goring-Morris, Anna Belfer-Cohen and Ofer Bar Yosef for answering various queries and Juliet Clutton-Brock, who first drew my attention to William Elder’s study of deer mandibles in Missouri. I am grateful to Arturo Morales as well as several anonymous referees who also read and commented upon an earlier version of this text. References [1] G.N. Bailey, A.S. Craighead, Late Pleistocene and Holocene coastal palaeoeconomies: a reconsideration of the molluscan evidence from northern Spain, Geoarchaeology: An International Journal 18 (2003) 175e204. [2] O. Bar-Yosef, Earliest food producers e pre Pottery Neolithic (8000e5500), in: T.E. Levy (Ed.), The Archaeology of Society in the Holy Land, Leicester University Press, London, 1995, pp. 190e204. [3] O. Bar-Yosef, R. Meadow, The origin of agriculture in the Near East, in: A.B. Gebauer, D. Price (Eds.), Last Hunters, First
  • 19. Farmers, School of American Research, Santa Fe, 1995, pp. 39e94. [4] A. Belfer-Cohen, O. Bar-Yosef, Early sedentism in the Near East: a bumpy ride to village life, in: I. Kuijt (Ed.), Life in Neolithic Farming Communities: Social Organization, Identity, and Differ- entiation, Kluwer Academic/Plenum, New York, 2000, pp. 19e37. [5] J. Cauvin, Naissance des divinités Naissance de l’agriculture, Flammarion, Paris, 1998. [6] M.N. Cohen, The Food Crisis in Prehistory, Yale University Press, New Haven, 1977. [7] S.J.M. Davis, A note on the dental and skeletal ontogeny of Gazella, Israel Journal of Zoology 29 (1980) 129e134. [8] S.J.M. Davis, Climatic change and the advent of domestication: the succession of ruminant artiodactyls in the late Pleistocenee Holocene in the Israel region, Paléorient 8 (1982) 5e15. [9] S.J.M. Davis, The age profiles of gazelles predated by ancient man
  • 20. in Israel: possible evidence for a shift from seasonality to sedentism in the Natufian, Paléorient 9 (1983) 55e62. [10] S.J.M. Davis, A preliminary report of the fauna from Hatoula: a NatufianeKhiamian (PPNA) site near Latroun, Israel, in: M. Lechevallier, A. Ronen (Eds.), Le site NatoufieneKhiamien de Hatoula prés de Latroun, Israël; fouilles 1980e1982 rapport préliminaire, vol. 1, Centre de Recherches français de Jérusalem, 1985, pp. 71e98. [11] S.J.M. Davis, Why did prehistoric people domesticate food animals? in: O. BarYosef, B. Vandermeersch (Eds.), Investiga- tions in South Levantine Prehistory, British Archaeological Reports International Series, vol. 497, pp. 43e59 (Oxford). [12] S.J.M. Davis, When and why did prehistoric people domesticate animals? Some evidence from Israel and Cyprus, in: O. Bar- Yosef, F.R. Valla (Eds.), The Natufian Culture in the Levant, Michigan, InternationalMonographsinPrehistory,pp.381e390(AnnArbor). [13] S.J.M.Davis, H. Hemmer, More animal bones from Tel Yarmouth:
  • 21. an Early Bronze Age site in Israel, in: P. de Miroschedji (Ed.), Tel Yarmouth II. [14] S.J.M. Davis, R. Rabinovich, N. Goren-Inbar, Quaternary extinctions and population increase in western Asia: the animal remains from Biq’at Quneitra, Paléorient 14 (1988) 95e105. [15] S.J.M. Davis, O. Lernau, J. Pichon, The animal remains: new light on the origin of animal husbandry, in: M. Lechevallier, A. Ronen (Eds.), Le gisement de Hatoula en Judée occidentale, Israël, Mémoires et Travaux du Centre de Recherche français de Jérusalem 8, Association Paléorient, Paris, 1994, pp. 83e100. [16] J. Diamond, The Rise and Fall of the Third Chimpanzee, Radius, London, 1991. [17] J. Diamond, Evolution, consequences and future of plant and animal domestication, Nature 418 (2002) 700e707. [18] W.H. Elder, Primaeval deer hunting pressures revealed by remains from American Indian middens, Journal of Wildlife Management 29 (1965) 366e370.
  • 22. [19] K. Flannery, Origins and ecological effects of early domestication in Iran and the Near East, in: P.J. Ucko, G.W. Dimbleby (Eds.), The Domestication and Exploitation of Plants and Animals, Duckworth, London, 1969, pp. 73e100. [20] N.A. Goring-Morris, Developments in terminal Pleistocene hunteregatherer socio-cultural systems: a perspective from the Negev and Sinai deserts, in: I. Hershkowitz (Ed.), Proceedings of the Second Symposium on the Upper Palaeolithic, Mesolithic, and Neolithic Populations around the Mediterranean Basin, BAR International Series, vol. 508, pp. 7e28 (Oxford). [21] L. Groube, The impact of diseases upon the emergence of agriculture, in: D.R. Harris (Ed.), The Origins and Spread of Agriculture and Pastoralism in Eurasia, University College Press, London, 1996, pp. 101e129. [22] J. Guilaine, F. Briois, J.-D. Vigne, I. Cerrère, Découverte d’un Néolithique précéramique ancien chypriote (fin 9 e
  • 23. , début 8 e millénaires cal. BC), apparenté au PPNB ancien/moyen du Levant nord, Comptes Rendu de l’Académie de Science, Paris, Sciences de la Terre et des planètes 330 (2000) 75e82. [23] E. Hahn, Die Haustiere und ihre Beziehungen zur Wirtschaft des Menschen, Eine geographische Studie, Duncker and Humblot, Leipzig, 1896. [24] J.R. Harlan, D. Zohary, Distribution of wild wheat and barley, Science 153 (1966) 1075e1080. [25] B. Hayden, A new overview of domestication, in: T. Douglas Price, A.B. Gebauer (Eds.), Last Hunters e First Farmers, School of AmericanResearchPress,SantaFe,NewMexico,1995,pp.273e299. [26] D.O. Henry, Models of agricultural origins and proxy measures of prehistoric demographics, in: R.T.J. Cappers, S. Bottema (Eds.), The Dawn of Farming in the Near East, Studies in Early
  • 24. Near Eastern Production, Subsistence, and Environment, vol. 6, ex oriente, Berlin, 1999, pp. 15e25 (2002). [27] E. Isaac, Geography of Domestication, Prentice-Hall, Englewood Cliffs, New Jersey, 1970. [28] R.G. Klein, Stone Age exploitation of animals in southern Africa, American Scientist 67 (1979) 151e160. [29] R.G. Klein, G. Avery, K. Cruz-Uribe, D. Halkett, J.E. Parkington, T. Steele, T.P. Volman, R. Yates, The Ysterfontein 1 Middle Stone Age site, South Africa, and early human exploitation of coastal resources, Proceedings of the National Academy of Science, USA 101 (2004) 5708e5715. [30] R.B. Lee, What hunters do for a living, or, how to make out on scarce resources, in: R.B. Lee, I. DeVore (Eds.), Man the Hunter, Aldine, Chicago, 1968, pp. 30e48. [31] A.J. Legge, Prehistoric exploitation of the gazelle in Palestine, in: E.S. Higgs (Ed.), Papers in Economic Prehistory, Cambridge
  • 25. University Press, Cambridge, 1972, pp. 119e124. 1416 S.J.M. Davis / Journal of Archaeological Science 32 (2005) 1408e1416 [32] D. Lubell, Are land snails a signature for the Mesolithice Neolithic transition? Documenta Praehistorica 31 (2004) 1e24. [33] T.R. Malthus, An Essay on the Principle of Population, As It Affects the Future Improvement of Society; with Remarks on the Speculations of Mr. Godwin, M. Condorcet and Other Writers, Johnson, London, 1798 (1970, Penguin, Harmondsworth). [34] L. Martin, Mammal remains from the eastern Jordanian Neo- lithic, and the nature of Caprine herding in the steppe, Paléorient 25 (1999) 87e104. [35] A. Morales, E. Roselló, F. Hernández, Late Upper Paleolithic subsistence strategies in southern Iberia: tardiglacial faunas from Cueva de Nerja (Málaga, Spain), European Journal of Archae- ology 1 (1998) 9e50.
  • 26. [36] N.D. Munro, A prelude to agriculture: game use and occupation intensity during the Natufian period in the southern Levant, PhD dissertation, Department of Anthropology, University of Arizo- na, Tucson, 2001. [37] N.D. Munro, Zooarchaeological measures of hunting pressure and occupation intensity in the Natufian, Current Anthropology 45 (Suppl.) (2004) S5eS33. [38] S. Payne, Partial recovery and sample bias: the results of some sieving experiments, in: E.S. Higgs (Ed.), Papers in Economic Prehistory, Cambridge University Press, Cambridge, 1972, pp. 49e64. [39] S. Payne, Faunal change at Franchthi cave from 20,000 BC to 3000 BC, in: A.T. Clason (Ed.), Archaeozoological Studies, North Holland, Amsterdam, New York, 1975, pp. 120e131. [40] J. Perrot, Le gisement natoufien de Mallaha (Eynan), Israël,
  • 27. L’Anthropologie 70 (1963) 437e483. [41] P.J. Richerson, R. Boyd, R.L. Bettinger, Was agriculture impossible during the Pleistocene but mandatory during the Holocene? A climate change hypothesis, American Antiquity 66 (2001) 387e411. [42] P. Rowley-Conwy, The laziness of the short-distance hunter: the origins of agriculture in western Denmark, Journal of Anthropo- logical Archaeology 3 (1984) 300e324. [43] A. Sherratt, Plate tectonics and imaginary prehistories: structure and contingency in agricultural origins, in: D.R. Harris (Ed.), The Origins and Spread of Agriculture and Pastoralism in Eurasia, University College Press, London, 1996, pp. 130e140. [44] M.C. Stiner, Thirty years on the ‘‘Broad Spectrum Revolution’’ and paleolithic demography, Proceedings of the National Academy of Sciences, USA 98 (2001) 6993e6996. [45] M.C. Stiner, N.D. Munro, T.A. Surovell, E. Tchernov, O. Bar-
  • 28. Yosef, Paleolithic population growth pulses evidenced by small animal exploitation, Science 283 (1999) 190e194. [46] T.A. Surovell, Modeling occupation intensity and small game use in the Levant, in: J.C. Driver (Ed.), Zooarchaeology of the Pleistocene/Holocene Boundary, Proceedings of a Symposium held at the Eighth Congress of the International Council for Archaeozoology (ICAZ) Victoria, British Columbia, Canada August 1998. BAR International Series, vol. 800, pp. 31e36 (Oxford). [47] R.M. Sussman, Child transport, family size and increase in human population during the Neolithic, Current Anthropology 13 (1972) 258e259. [48] E. Tchernov, An early Neolithic village in the Jordan Valley. Part II: the fauna of Netiv Hagdud, Peabody Museum of Archaeology and Ethnology, Harvard University, American School of Pre- historic Research. Bulletin 44 (1994). [49] H.-P. Uerpmann, Probleme der Neolithisierung des Mittelmeer- raums, in: Tübinger Atlas des Vorderen Orients, B, vol. 28,
  • 29. Ludwig Reichert, Wiesbaden, 1979. [50] F.E. Zeuner, A History of Domesticated Animals, Hutchinson, London, 1963. Why domesticate food animals? Some zoo-archaeological evidence from the LevantIntroductionMaterial and methodsThe "Spectrum Shift " (Figs. 2 and 3)The "Age Shift "(Fig.4 and Table 1)ConclusionAcknowledgementsReferences ARTICLES Central Questions in the Domestication of Plants and Animals MELINDA A. ZEDER DEFINING DOMESTICATION All approaches to defining domesti- cation in both plants and animals rec- ognize that domestication involves a relationship between humans and tar- get plant or animal populations. There are, however, distinct and often dis- cordant perspectives taken regarding the balance of power in this relation- ship and its central defining features (Fig. 1). Many approaches to defining domestication, especially those focus- ing on animals, emphasize the domi- nant role humans play in assuming
  • 30. “mastery” over all aspects of the re- production, movement, distribution, nourishment, and protection of do- mesticates.2–5 Integral to definitions that place humans in control of the process is the notion of intentionality, that humans with foresight and delib- erate intent intervened in the life cycle of target plant and animal popula- tions and assumed responsibility for their care to meet specific and well- defined objectives serving human needs. Also often associated with this emphasis on the human dimension is the notion that domestication in- volves a fundamental change in socio- economic organization in which suc- cessive generations of domesticates become integrated into human societ- ies as objects of ownership.3,6 Other researchers object to “anthro- pocentric” approaches to defining do- mestication that portray domesticates as passive pawns in the process, point- ing out that domesticates also reap benefits through vastly enhanced re- productive fitness and expanded rang- es.7 Those operating within an evolu- tionary biology perspective, in particular, maintain that the relation- ship between humans and domesti- cates is no different from other mutu- alistic relationships in the “natural world” that bring together species like
  • 31. ants and aphids in partnerships of in- creasing co-dependency.8 Moreover, as one moves further along the spec- trum, from a relatively balanced mu- tualistic perspective to ones that focus on the domesticate, the role of delib- erate human intent declines. The more extreme positions at this end of the spectrum tip the balance in favor of the domesticate, which is seen as manipulating its human partners for its own evolutionary advantage, en- snaring humans in a relationship that may have actually reduced human fit- ness.9 Another axis of variation in defini- tional approaches to domestication is the relative primary given to genetic and associated morphological change. An emphasis on genetic change and its phenotypic expression is particu- larly common among researchers fo- cusing on plant domestication, espe- cially the domestication of large- seeded annuals, where human intervention results in fairly rapid ge- netic changes with easily observed phenotypic expressions.9,10 Some re- searchers focusing on animals also see genetic isolation and subsequent quick-onset morphological change as essential attributes of domestica- tion.11 The requirement that domesticates
  • 32. show evidence of morphological or even genetic change, however, is not universally accepted. Nor is the basic premise underlying this requirement: that the process of domestication is contingent on reproductive isolation and resultant genetically driven mor- Melinda A. Zeder is Director of the Ar- chaeobiology Program of the Smithso- nian Institution’s National Museum of Nat- ural History. Her research focuses on questions of domestication, origins of ag- riculture, and the environmental and so- cial impacts of early agricultural econo- mies in the ancient Near East. She is the lead editor of the volume, Documenting Domestication: New Genetic and Archae- ological Paradigms, with co-editors Eve Emshwiller, Daniel G. Bradley, and Bruce D. Smith, published in spring of 2006 by the University of California Press. Key words: domestication; origins of agriculture; plants; animals; climate change; demographic pressure; social forces; Southwest Asia © 2006 Wiley-Liss, Inc. DOI 10.1002/evan.20101 Published online in Wiley InterScience (www.interscience.wiley.com). Along with symbolic communication, tool use, and bipedalism, the domestica- tion of plants and animals, together with the associated emergence of agriculture,
  • 33. stands as one of the pivotal thresholds in human evolution. For more than a hundred years researchers have wrestled with the questions of what domestica- tion is, how it is detected, and why it happened. The past decade in particular has witnessed a remarkable acceleration of interest in domestication, thanks to ad- vances in our ability to detect the context, timing, and process of domestication in a wide array of different plant and animal species around the world.1 This review focuses on overarching issues of defining, documenting, and explaining the do- mestication of plants and animals, tracing a path through often discordant view- points to offer some new perspectives. Evolutionary Anthropology 15:105–117 (2006) phological change. This is particularly true for animals, where morphologi- cal change, when it occurs at all, is often both delayed and difficult to tie directly to domestication.12 As a re- sult, many researchers define animal domestication not in terms of ob- served genetic or morphological change, but in terms of causal human behavior. According to this view, do- mestication falls along a continuum of increasing human intervention ranging from predation to genetic en- gineering13 in which there are varying
  • 34. degrees of investment in altering an animal’s natural behavior (its move- ment, breeding schedule, or popula- tion structure) to suit human needs.6,14 A similar view is becoming increasingly common in consider- ations of plant domestication espe- cially perennial plants such as root crops propagated through vegetative cloning or very long-lived tree crops in which genetic and morphological change may be less automatic and more subtle than in annual seed crops.15 Smith,16 for example, main- tains that for both plants and animals the central defining feature of domes- tication and the creation of domesti- cates is the nature of the “ongoing re- lationship of intervention initiated and sustained by humans.” This em- phasis on the evolving relationship be- tween humans and plant or animal populations turns attention away from a range of secondary conse- quences of domestication, such as ge- netic and morphological change or so- cial notions of property, and properly returns it to a consideration of the new partnership that humans create with target populations. Domestication does indeed have many features in common with other mutualistic relationships among plants and animals. Both partners in
  • 35. the relationship of domestication clearly derive benefits. Plant and ani- mal partners benefit in increased re- productive fitness and range expan- sion. Human partners gain increased security and predictability in their ac- cess to resources of interest. Both partners respond to this relationship in ways that enhance respective pay- offs and further deepen their mutual investment in its continuation. But the mutualism that lies at the heart of the domestication process involving human societies and target plant and animal populations varies in signifi- cant ways from other similar relation- ships found in nature. In a recent overview of the homolo- gies between human agriculturists and fungus-growing ants, Schultz and coworkers17 highlight the many paral- lels between these convergent forms of mutualism while also underscoring key qualitative differences between at- tine and human agriculture. The co- dependent relationships between farmer ants and domesticated fungi are the result of a gradual co-evolu- tionary process based on mutation-in- duced behavioral and morphological change in both partners. Humans, on the other hand, are capable of modi- fying their behavioral repertories through “trial and error, observation,
  • 36. and imitation.”17 That ability enables humans to rapidly develop behavioral strategies aimed at meeting con- sciously recognized needs. The highly developed human ability for cultural transmission of learned behavior, Shultz and coworkers argue, greatly accelerated the adaptive modification of human behavior, shifting the bal- ance of power in the emergent mutu- alism. Humans quickly assume a dominant role because they are free to choose among genetic variants in the partner population, to manipulate the behavior and life history of symbionts (even to their own detriment), or to terminate the relationship with one partner symbiont and choose another. This is where intentionality comes into the picture. It is true that humans Figure 1. Definitions of domestication tend to fall somewhere along three axes of variation. Definitions that award the balance of power in the domestic relationship to humans tend to stress human intentionality and the social and economic impacts of domestication. Definitions that tip the balance of power in favor of the domesticate tend to discount the role of human intentionality in the process and stress its biological impacts on the domesticate. 106 Zeder ARTICLES
  • 37. could not have foreseen the adaptive responses by plant and animal part- ners to the new selective factors brought into play by the relationship of domestication. Nor did humans likely appreciate the long-term bene- fits (or the negative consequences) that might accrue from domestication and the subsequent development of agricultural economies. However, rul- ing out this kind of prescience on the part of humans does not take inten- tionality out of the picture. While they might not have understood the princi- ples of genetic engineering, humans could appreciate the fact that tending, nurturing, and intervening in the life cycle of certain plants and animals yielded various immediate benefits. On the basis of these returns they could then consciously and deliber- ately decide to continue to engage in these behaviors, and to elaborate on them, instead of engaging in other strategies. Intentionality, then, becomes the key factor that distinguishes domesti- cation from other similar mutualistic relationships in nature. The deliberate role humans take in actively pursuing the domestic partnership also distin- guishes it from other biological rela- tionships between humans and plants and animal species, such as commen-
  • 38. sal relationships with mice, sparrows, or weeds that take advantage of new niches created by human habitation. Definitions that try to pigeon-hole do- mestication as either a cultural or bi- ological process are bound to come up short. Clearly, domestication has a bi- ological component as a mutualistic relationship between humans and plant or animal symbionts. Just as clearly, however, human intentional- ity sets domestication apart from other forms of mutualism. The uniqueness of the relationship comes from its cultural component and the dominant role humans play in con- sciously and deliberately perpetuating it to their own advantage. If the process of domestication is best viewed as a form of mutualism that is asymmetrically enhanced by the human ability to culturally trans- mit learned behavior, then at what point along this developmental trajec- tory does the plant or animal partner become a domesticate? Is there a threshold that, once crossed, sepa- rates the “wild” from the “domestic”? If so, what does this threshold look like? To some extent, it remains a matter of personal preference to de- cide just when a domestic subsection of a plant or animal species has been created. Threshold criteria that re-
  • 39. quire total genetic isolation and emer- gent speciation or complete depen- dence on humans for survival set a very high bar that many, if not most, widely accepted domesticates would fail to clear. Even somewhat looser standards that involve a lesser degree of genetic modification in the target plant or animal population, or a cer- tain level of human investment in propagating, nurturing, or owning the resource, run the risk of constructing artificial boundaries along what was really a more seamless incremental process. Ducking the issue by adopting the term “proto-domesticate” also does not help much. This term implies that, if just given enough time and perhaps a little more investment by either part- ner, full domestic status would be achieved. The actual trajectory of do- mestication, however, is highly con- tingent on a wide range of factors, including the ability of the plant or animal to take advantage of the rela- tionship, the strategies and accompa- nying technologies humans develop to manage the resource, and its chang- ing value vis-à-vis available alternative resources. In some plant and animal species, genetic modification and more focused human investment in the resource may quickly follow. In others there may be a long and very
  • 40. stable relationship involving fairly minimal commitment by either part- ner. Further, it appears that budding domestic relationships sometimes fail altogether, never moving beyond an initial courtship phase. It is best to step back and not focus too closely or obsessively on defining the exact demarcation between do- mestic and wild, and to turn, instead, to a consideration of the full span of the evolving nature of domestic rela- tionships. Different stages in the evo- lution of this relationship might be characterized by the degree of invest- ment by both partners (Fig. 2). For the plant or animal, this would involve the extent of genetic modification made in response to new selective pressures, the degree of its genetic isolation from populations not involved in the part- nership, the nature of subsequent morphological or behavioral change, and its increasing co-dependency on humans. For humans, this might be the level of investment in the produc- tion of the resource; that is, in tilling, watering, burning, and land clear- ance, sowing, and transplanting plants, or in taming, protecting, herd- ing, culling, and selectively breeding animals. It might also include the de- gree of incorporation of domesticates within the socio-economic organiza-
  • 41. tion of the human groups investing in its production. By expanding the scope of inquiry to encompass the vast “middle ground” between foraging and farm- ing, hunting and herding,16 we can approach a deeper, more comprehen- sive, and ultimately more informative appreciation of the range of possibili- ties open to humans and their plant and animal partners. This expanded territory of investigation includes the stable, long-lived systems of low-level food production involving a mix of both morphologically altered and nonaltered domesticates, as well as “wild” resources, featured in recent books on indigenous resource man- agement in California and the North- west Coast.18,19 At the other end of the spectrum are highly structured agri- cultural economies with complete de- pendence on domesticates and total investment in their production. Try- ing to understand the full richness of the various ways in which the domes- tic partnership may manifest itself in different contexts, we, in turn, stand a much better chance of being able to document and explain domestication. DOCUMENTING DOMESTICATION In both plants and animals, this ef-
  • 42. fort requires identifying clear-cut markers that can be explicitly linked to a specific aspect or stage of the un- folding domestication process.1,20 Dif- ferent markers may be more effective in detecting different stages of this process. Markers will also vary de- pending on the biology of the domes- ARTICLES Domestication of Plants and Animals 107 ticate and its relationship with hu- mans. There are, in particular, fundamental differences in the selec- tive pressures on plants and animals undergoing domestication, and, as a result, in the corresponding markers used to document plant and animal domestication. Selective pressures on plants, espe- cially annuals, tend to operate directly on morphological traits that can, in turn, be used as unambiguous mark- ers of domestication.15 Morphological impacts of the domestication of annu- als may come about as largely auto- matic responses to human planting and harvesting that result in such changes as increased seed size, thin- ner seed coats, reconfiguration of seed head architecture, or the development of indehiscent seed pods.15,21,22 Inten- tional selection for specific morpho-
  • 43. logical attributes in annual plants, such as larger fruit size, appear to happen later in the developing rela- tionship of domestication.15 Perennial plants sustained by trans- planting root fragments, on the other hand, are not subjected to the same seed-bed pressure and human har- vesting selective pressures that result in the morphological markers used to document domestication in annual seed plants.15 At the same time, how- ever, because there may be more of an opportunity for humans selectively to replant root fragments with desired traits, these plants may respond fairly quickly to deliberate human selection in the development of larger fruits, the loss of chemical defenses against her- bivory, or changes in sugars and starches.23 While many of these crops were grown in tropical areas with poor preservation of plant macro-fos- sils, the development of breakthrough techniques for the recovery and iden- tification of plant micro-fossils (that is, phytoliths and starch grains) has made it possible to detect these do- mestication-induced morphological changes in root and other crop plants (Fig. 3).24,25 Recent years have seen an increase the use of nonmorphological markers
  • 44. of the intensification of human-plant interactions that may precede clear- cut evidence of morphological change in plants. Evidence of land clearance, modification of natural drainage sys- tems, intentional burning, and changes in the composition of weedy plants in archeological assemblages have all been effectively used to track human modification of landscapes and plant communities as part of the domestication process.26 –28 The oc- currence of plant macro- or micro fos- sils in areas thought to be far outside their natural range has also been in- terpreted as evidence of human trans- port and tending of plants.26,29,30 There are special challenges to find- ing markers of animal domestication. This is because the leading-edge pres- sures on animals undergoing domes- tication are likely to focus on behav- Figure 2. Domestication is best viewed as an evolving of mutualism between humans and populations of plants or animals. The relationship can be characterized along various scales of investment by either the human or the plant or animal partners. All of these scales usually are involved in the process of domestication, though they often operate independently of one another. The degree of change along each scale is contingent on the biology of the species involved, as well as the ecological and cultural circumstances of the human partners.
  • 45. Attempting to distinguish just where and along what scale domestication occurs is not only difficult, but may not be very useful. 108 Zeder ARTICLES ioral attributes rather than on morphological traits.31 There are a va- riety of behaviors that probably made certain animal species better candi- dates for domestication; among them tolerance of penning, a social struc- ture based on dominance hierarchies, sexual precocity, weak alarm systems and, above all, reduced wariness and aggression.32 Behavioral responses to domestication in animals elaborated on these initial preselection qualifying attributes and include a general re- duction in responsiveness to environ- mental stimuli, reduced activity lev- els, increased social compatibility, and intensified sexual behavior.4,33 Many morphological traits com- monly seen in domestic animals are thought to be linked to these behav- ioral changes. These attributes in- clude piebald coats, lop ears and, of special importance here, reduced brain size and an overall juveniliza- tion of cranial form.4,33 This latter fea- ture may result in a shortened muzzle, tooth crowding, and reduction in
  • 46. tooth size, traits frequently seen as leading-edge markers of domestica- tion in dogs and pigs.34,35 Selection for these behavioral traits and their associated morphological effects, however, may not be uniquely re- stricted to domestication. Similar be- havioral traits, such as reduced wari- ness and timidity, are also selected for in animals such as rats and sparrows that develop commensal relationships with humans. These animals also show changes in pelage coloration and brain size.4 It is possible, then, that the initial modifications in tooth size and cranial form in pigs and dogs, widely seen as markers of domestica- tion, may in fact be attributable to an early commensal relationship be- tween humans and such omnivore species that began their association with humans as camp-follower scav- engers.31 Other genetically driven morpho- logical changes in animals undergo- ing domestication come about when humans begin deliberately selecting breeding partners. Changes in the size and shape of horns in animals like goats and sheep, for example, are di- rectly tied to the relaxation of selective pressures for and, quite likely, active selection against large horns once hu- mans assume control over breeding.36
  • 47. Other morphological changes may en- sue when animals are moved into new territories, either through founder ef- fects, random genetic drift, or di- rected adaptation to new environmen- tal conditions. Later still, and probably much later in animals than in plants, deliberate human selection for attributes that enhance meat, fi- ber, milk yields, or labor potential may result in still other morphological markers that might be used to detect intensification in the human-animal relationship. Domestication has also been sug- gested to have resulted in a marked and rapid reduction in body size, which, until recently, has been widely held to be a definitive marker of initial domestication.37 This proposed mor- phological response has been vari- ously attributed to a plastic response to nutritional deficiencies, an adaptive advantage of smaller bodies for ani- mals subjected to impoverished con- ditions, or deliberate human selection for more tractable individuals.38 – 40 But body size in animals is also known to be affected by well-docu- mented factors such as sex, environ- ment, climate, and age, which may be entirely unrelated to domestication
  • 48. and may mask or be mistaken for changes in body size induced by do- mestication.12 Given the looser connection be- tween domestication and morpholog- ical change in animals, it is not sur- prising that considerable attention has been devoted to identifying mark- ers of domestication that do not rely on genetically driven morphological change, but that, instead, reflect hu- man actions directed at managing an- imals. Demographic markers aimed at detecting the different harvest strate- gies of hunters and herders were among the first nonmorphological markers used to detect animal domes- tication.41 Largely abandoned in the 1980s and 1990s, when most archeo- zoologists embraced size reduction as a leading-edge marker of animal do- mestication, demographic markers are seeing a resurgence, thanks in part to the development of methods for constructing high-resolution sex-spe- cific harvest profiles.12,42 Applying these methods to archeological as- semblages has shown that what was once interpreted as evidence of do- mestication-induced body size reduc- tion in goats (and likely other live- stock species) is, instead, a reflection of a change in the demographics of the adult portion of managed herds dominated by females (Fig. 4).12 Un-
  • 49. ambiguous changes in morphological traits such as body size or horn form seem to postdate human management of herd animals by hundreds of years and represent later phases in the do- mestication process.12,43 Both the presence of animals out- side their presumed natural habitat and a sudden dramatic increase in a previously little-exploited animal have also been used as markers of animal domestication.41,43,44 But the use of these markers (in both animals and Figure 3. Starch grains from wild and domes- tic yams, Dioscorea sp. A. Granules from modern domestic yams (D. trifida). B. starch grains from a wild yam (D. cymosula) from Panama. Starch grains from wild yam spe- cies studied thus far from the Neotropics are distinct in morphology as compared with domestic varieties. As a rule, wild forms are also highly variable within a single tuber, whereas domesticated species have a sin- gle morphological type of starch, which may be a result of human selection. ARTICLES Domestication of Plants and Animals 109 plants) needs to be tempered by ac- knowledgment of our generally poor understanding of the geographic range of biotic communities in the
  • 50. past and of the paleo-environmental conditions that shaped these ranges. A rapid increase in the abundance of a plant or animal resource in an arche- ological assemblage might simply sig- nal the intensification of hunting and gathering strategies, not the begin- ning of food production. Markers of animal domestication may also be found in plastic, nonge- netically driven responses such as bone and tooth pathologies, evidence of pandemic disease, or chemical changes in the composition of bone and tooth enamel used to track changes in nutrition and the seasonal movement of managed animals.45– 48 The presence of corrals, pens, or other traces of animals, such as manure or hoof prints, in human settlements, changes in human settlement pat- terns, artifacts related to the exploita- tion of domestic animals (bits or milk churns and storage vessels), and even changes in food distribution patterns have been used with varying effect to build cases for animal domestica- tion.49 The application of these later plastic responses and cultural mark- ers needs to be tempered by the real- ization that they may not be mani- fested in all instances of animal domestication or may result from other pressures unrelated to domesti-
  • 51. cation. Application of such markers is most effective when many of them are brought together to build strong cir- cumstantial cases for domestica- tion.49 Advances in methods for extracting and amplifying both modern and an- cient DNA is recent years have pro- vided an exciting new window on the genetic changes associated with the domestication of plants and ani- mals.1,20,50,51 Some of this work has focused on identifying the genes or gene complexes that are specifically selected for or against in the process of domestication, especially of crop plants.52 However, most genetic stud- ies of domestication look to largely neutral noncoding loci and organellar genomes. These procedures have proven useful in tracing the diver- gence of domesticates from their wild progenitors; identifying the number and geographic location of domestica- tion events, which now appear to have been multiple for most animal domes- ticates and many plant crops; and in tracking the dispersal of domesticates and their human partners out of cen- ters of origin. There are critical differences in the relative rates of evolution in the dif- ferent genomes of plants and animals
  • 52. that play a major role in the genetic markers used. The relatively rapid rate of evolution in mitochondrial DNA (mtDNA) in animals makes mtDNA particularly well suited to tracking the relatively shallow time depth of divergence between domesti- cates and their wild progenitors (�10,000 years). This is why most studies of animal domesticates focus on this genome.51 While it is less vari- able than mtDNA and evolves much more slowly, low-copy Y-chromo- some nuclear DNA provides a window into patrilineal inheritance, which in many animal domesticates is quite different from matrilineal history.53,54 Variation in noncoding nuclear mic- rosatellite DNA, contributed by both parents, has also proven useful in tracking the divergence of different breeds of animals.55,56 While not approaching the rate of evolution of animal mtDNA, loci in the nuculear genome of plants evolve at about the same rate as nDNA loci in mammals. They also evolve about four times faster than loci in the choloroplast genome and twelve times faster than plant mtDNA. Conse- quently, genetic studies of plant do- mestication tend to focus on nDNA, especially on highly polymorphic mi-
  • 53. crosatellites that provide sufficient in- traspecific variation to document the domestication process.50,57–59 The nu- clear genome in plants has proven es- pecially useful in tracking down the various ancestral genomes contribut- ing to the complicated genetic heri- tage of hybrids and polyploid crop plants. These common conditions in plant crops generally are not found among animal domesticates.50,60 Most genetic approaches to docu- menting domestication are based on modern domesticates and likely wild progenitor species. But the window they provide on the origin and early dispersal of domesticates is unavoid- ably clouded by thousands of years of selective breeding, hybridization, and introgression between wild and do- mestic populations. Ancient DNA (aDNA), on the other hand, has the potential to shed more direct light on the process of genetic divergence of domesticates. Due to the greater pres- ervation of DNA encased in animal bone and the suitability of high-copy mtDNA in animals for tracking shal- low time depth divergences, aDNA studies of animal domesticates have been particularly successful, espe- cially those tracing the more recent dispersal of domestic animals through temperate environments.61– 64 Al- though it is more difficult to extract
  • 54. enough low-copy nDNA from un- charred archeological plant remains to provide meaningful results, some stunning results have recently been Figure 4. Goat (Capra hircus) first phalanges from Ganj Dareh ca. 10,000 cal B.P. The small bone on the left falls within the size range of such bones from female goats, while the larger specimen on the right is likely from a male. The female phalanx is fully fused, indicating the animal was slaughtered after this bone fused, which in goats is about 13 months of age. The male specimen is unfused, indicated that this an- imal was younger than 13 months of age when slaughtered. The selective slaughter of young male animals with prolonged sur- vivorship of female animals is symptomatic of human management. (Photo by Carl Hansen) 110 Zeder ARTICLES obtained in the use of aDNA to track the origin and dispersal of domestic plants.65,66 In the excitement over the possible contributions of genetic analysis to the documentation of plant and ani- mal domestication, it is important not to lose sight of the fact that there is more to domestication than ge-
  • 55. netic change. The real power of these new tools for tracking the tra- jectory of domestication can be real- ized only when genetic analyses are more fully integrated into broader archeological analyses. Genetic studies represent one, albeit very powerful, line among many parallel and mutually illuminating lines of evidence, which, when considered together, provide a fine-grained view of unfolding domestic partnerships (Box 1). EXPLAINING DOMESTICATION Efforts to explain domestication and the origins of agriculture tend to cycle among a relatively limited num- ber of forcing factors championed as primary and often universal levers of change. These forcing factors can be generally grouped under three banners: environmental change, demographi- cally induced resource pressure, and changes in social organization and ide- ology. Explanations focusing on environ- mental change can be traced to V. Gordon Childe, who credited post- Pleistocene aridification with bring- ing humans together with plants and animals around water sources in po- sitions of “enforced juxtaposition” that promoted “symbiosis between
  • 56. man and beast,” resulting in domesti- cation.67 While climate models were largely out of favor in the 1960s through 1980s, recent advances in paleo-environmental reconstruction have resulted in its rehabilitation as a primary player in agricultural origins. In particular, the now well-docu- mented brief return to Ice Age condi- tions at about 13,000 cal. B.P., known as the Younger Dryas, is increasingly featured as having played a key role in agricultural origins,68 with domestica- tion coming about either during this climatic downturn as a way of coping with environmental degradation69,70 or after it as a response to the follow- ing climatic amelioration and stabili- zation.71 One recent climate based model of agricultural origins even maintains that agriculture was a “compulsory” development of cli- matic stabilization and rise in ambi- ent CO2 following the final pulse of Ice Age climate in the Younger Dryas.72 Explanatory frameworks founded on notions of population dynamics and resultant resource pressures can be traced back to Binford’s Edge-Zone Hypothesis of the late 1960s.73 Ac- cording to this theory, agricultural or- igins were the result of resource pres- sure in marginal zones caused by
  • 57. emigration from more optimal zones experiencing high rates of population growth. Mark Cohen’s74 subsequent “food-crisis” model held that runaway population growth worldwide, not just in marginal zones, forced people around the planet to abandon more nutritious hunting and gathering strategies and assume the burden of tending to domesticated plants and animals. Binford’s75 recent return to the topic down-plays the role of pop- ulation growth, resource pressure, and emigration, and instead high- lights population packing, in particu- lar a threshold limit of 9.098 people per 100 km2 as the “universal condi- tioner of change in . . . subsistence strategies” and the deus ex machina of agricultural origins. Although explicitly rejecting univer- sal normative explanations of domes- tication and agricultural origins, those operating under the general ru- bric of human behavioral ecology (HBE) also feature resource pressure as a primary causal component.76 A basic axiom of HBE models is that humans will always opt for strategies that optimize immediate returns from high-ranking resources, with the rela- tive rank of a resource determined by kilocalorie return over pursuit and handling costs.77 Food production vi- olates these central principals. First of
  • 58. all, farming involves a rise to promi- nence of low-rank plant resources, while herding requires deferring the rewards from high-rank animal re- sources.78,79 Moreover, farming and herding are often seen as carrying higher production and processing costs than do hunting and gather- ing.80 A recent review highlights dif- ferent paradigms within HBE that might be helpful in understanding the transition from foraging to farming, such as concepts of constrained opti- mization, marginal value, opportunity costs, discounting, and risk-sensitivi- ty.76 Diet-breadth models that include avenues for plant resources of lower profitability to elevate their rank through changes in density or extrac- tion return hold particular promise here. However, even allowing for such advances up the resource-rank ladder, HBE models generally predict that plant resources will largely be ignored so long as animal protein, a more highly ranked resources, is sufficiently abundant.78 By necessity, then, strict adherence to the basic axiomatic te- nets of HBE casts the transition to food production in terms of rather bleak cost-benefit trade-offs. Follow- ing HBE principles, agriculture comes about when people, faced with pres- sure on resources, whether caused by fluctuating climates, population
  • 59. growth, or packing, are forced to fo- cus on resources and extraction strat- egies that under other circumstances would be considered far from opti- mal. Other theorists deny that external factors like climate change, popula- tion growth, or resource pressure have any causative role in this transi- tion. Instead, they look to forces within the human character, espe- cially a supposedly innate compulsion for self-aggrandizement, as primary forcing factors in the domestication of plants and animals and the transition to agriculture.81 Perhaps the best known and most influential of these socially based explanatory ap- proaches is that promulgated by Brian Hayden.82,83 In contrast to a backdrop of scarcity and stress cen- tral to other models, Hayden conjures up a more Eden-like setting for the origin of plant and animal domestica- tion, claiming that agriculture only develops in contexts of plenty, where an abundance of resources ignites a basic human predisposition for acqui- sition and social competition. In such settings certain particularly success- ful “accumulators” were able to mar- shal high-prestige food items and en- hance their own social advantage through such mechanisms as compet-
  • 60. ARTICLES Domestication of Plants and Animals 111 itive feasting. Domestic resources were especially attractive since they were more amenable to ownership and their supply could be both manip- ulated and appropriated. This model predicts, then, that domesticates were not, at least initially, ubiquitous di- etary staples, but were, instead, more likely to be rare and desirable limited- access delicacies, of little nutritional value, used by avaricious accumula- tors to both display and enhance their social power. Another stress-free theory, this one championed by Jacques Cauvin,84 Box 1. Gourds, Dogs, and the Peopling of the Americas Location of archeological sites with directly dated bottle gourd fragments studied by Erickson and coworkers.66 The recent integration of archeo- logical and genetic research on both the bottle gourd (Lagenaria siceraria) and the dog (Canis familiaris) has substantially clarified our understand- ing of the initial domestication of these two very early domesticates, as
  • 61. well as their late Pleistocene radiation from Asia across Beringia into the New World.20 The bottle gourd, indigenous to Af- rica and long valued as a container crop rather than as a food source, has consistently been recovered in arche- ological association with the earliest evidence of New World domesticated plants in many regions of the Ameri- cas. The prevalent consensus has been that L. siceraria was carried by ocean currents as a wild plant from Africa to South America. However, a morphological analysis comparing ar- cheological rind fragments from sites in the Americas (Fig. 1) with recently described wild L. siceraria fruits from Zimbabwe showed that the much thicker archeological rinds represent domesticated plants.66 Direct dates of these fragments indicate that do- mestic bottle gourd was present in the Americas by at least 10,000 years ago. Moreover, DNA recovered from nine archeological rind fragments predating the arrival of Europeans were identical to the modern Asian reference group, indicating that do- mesticated bottle gourds were car- ried to the New World during the late Pleistocene from Asia, not Africa. Al- though it is possible that the bottle gourd could have been carried from Asia to the Americas by the north Pa-
  • 62. cific current, it is more likely this early domesticate accompanied Paleoin- dian colonists as they crossed Ber- ingia into the New World. A parallel genetic study points to an Old World origin of the domestic dog and suggests that at least five founding dog lineages invaded North America with humans as they colonized the New World.63 The ear- liest archeological evidence of do- mesticated dogs in the Old World, dating to ca. 13,000 –17,000 B.P., comes from widely dispersed sites extending from the Near East across Eastern Europe. Although the earliest domesticated bottle gourd in the Old World dates to 8,000 –9,000 B.P. in China and Ja- pan, it is reasonable, given its arrival in central Mexico by 10,000 B.P., to estimate that it was initially domes- ticated in the same general time frame as the dog, ca 13,000 B.P. or earlier. Together, these studies indi- cate that Paleoindians entered the New World with the world’s two ear- liest domestic species, dogs and bottle gourds, and that initial do- mesticates served utilitarian func- tions, but not as sources of food. 112 Zeder ARTICLES
  • 63. also sets the stage for agricultural or- igins in a time of plenty and denies economic necessity a primary cata- lytic role. But Cauvin looks even more inward into the human psyche for the root cause of this transition. Accord- ing to Cauvin, domestication is a di- rect consequence of a conceptual shift in mankind’s mental template from one that saw humans as part of nature to one that cast humans in a dominant position, now free to manipulate and transform nature to their liking. This profound and irreversible transforma- tion in the way that humans saw themselves in relation to nature, cod- ified in religious ideology, found ex- pression in concrete ways: in art, household and community structure, and the domestication of plants and animals. A similar notion can be found in Hodder’s85 emphasis on the role of symbols and metaphors of hu- man dominance over nature made concrete in the form of the house, as the domus of domestication and the crucible of community.85 Increasing sophistication in ap- proaches to defining and document- ing domestication make it ever more difficult to support explanatory frame- works based on any single forcing fac-
  • 64. tor. This is particularly true for the Near East, where we have, arguably, the most complete record of the initial domestication of many plant and ani- mal species. Climate clearly played its part in ag- ricultural origins here. Increases in rainfall and temperature following the Last Glacial Maximum at about 15,000 cal. B.P. undoubtedly contrib- uted to the adoption of increasingly less mobile, more territorially focused strategies centered on intensive ex- ploitation of plant resources rebound- ing out of glacial refugia,70 where they had been used by humans since the Upper Paleolithic.86,87 A subsequent pulse of cold, arid Ice Age conditions during the Younger Dryas, ca. 13,000 to 11,600 cal. B.P., was met with more mobile strategies in the Southern Le- vant88,89 and, quite possibly, the initial domestication of cereal crops and pulses in well-watered oasis localities in the Northern Levant.70 Ameliorat- ing climates following the Younger Dryas saw the domestication of other crop plants in the Southern Levant.86 Animal domestication seems to have come about sometime later on in the Central and Eastern Fertile Cres- cent.12 A brief warming and drying climatic pulse at about 9,000 cal. B.P. coincided with the collapse of early
  • 65. agricultural communities in more arid parts of the Southern Levant and their proliferation throughout the rest of the Near East.88 So while climate change played a role in domestication and agricultural origins in this region, it did not do so in the simple stimulus- response way implied by many mod- els that award environment prime- mover status. Instead, climate change alternately helped push and pull peo- ple along a pathway toward domesti- cation and agriculture, providing both opportunities and challenges that people across the region met in vari- ous ways depending on highly local- ized circumstances. The transition from foraging to farming in the Near East clearly saw significant changes in mobility, popu- lation growth, and nucleation. But the record from the region does not sup- port the exponential population growth and attendant impoverish- ment of natural resources called for in Cohen’s90 doomsday model. In the ab- sence of settlement-pattern data ro- bust enough to demonstrate that his population-packing Rubicon (9.098 people per 100 km2) had been crossed, Binford pointed to the clear reduction in mobility, intensification in plant re- source use and development of stor- age technology during the postglacial era in the Near East as proxy evidence
  • 66. of packing.75 This troubling circular- ity of using the proposed results of packing as evidence of packing is found in many such demographic models.91 While population clearly in- creased, the admittedly incomplete settlement-pattern data do not sup- port the level of population packing that Binford grants exclusive causality for increasing sedentism, intensifica- tion of resource use and, ultimately, domestication. It is also hard to make a case that the initial focus on re- bounding populations of cereals, pulses, and nut trees after the Glacial maximum was caused by widespread depletion in higher ranking animal re- sources, as is required by most HBE models.89 Rather than being forced to settle down and focus on less desirable re- sources, it seems more likely that peo- ple took advantage of newly abundant high-yield plant resources and associ- ated herbivores in ways that enabled them to increase the size and duration of community nucleation beyond that possible under Ice Age conditions. It is also possible that when people were faced with localized pressures on re- sources resulting from more seden- tary ways of living, an interest in pre- serving the bonds of community provided an important incentive for the development of strategies that
  • 67. helped promote the yield and predict- ability of these resources. Moreover, these same social considerations also probably helped guide the subsequent responses to region-wide pressures caused by the climatic squeeze of the Younger Dryas and the stabilization of climate that followed. Yet while it is possible to award so- cial factors a more active role in the origins of agriculture in the Near East, the record from the region clearly does not support Brian Hayden’s82,83 competitive feasting model. Even those with the loosest tether to reality would have a hard time seeing cereals, the earliest and most important do- mesticates in the region, as anything but widely available staple resources. And while there is some evidence for feasting on large numbers of ani- mals,92 no credible case can be made for Hayden’s83 blanket assertion that meat was consumed only within con- trolled ritual contexts. Moreover, all of the indicators of status differentia- tion and unequal appropriation of so- cial and economic prestige that Hay- den sees as material manifestations of his greedy accumulators have since been more convincingly cast by Kuijt93 and others as evidence of mechanisms for maintaining an equalitarian status quo in the face of
  • 68. mounting social tensions incurred when larger groups stay together for longer periods of time.93 Kuijt maintains, however, that do- mestication played little role in these social developments, either as a cause or a consequence.94 He bases this con- clusion on the fact that morphologi- cally altered domesticates appear in the archeological record of the Levant ARTICLES Domestication of Plants and Animals 113 several thousand years after the first signs of leveling mechanisms for pro- moting social cohesion in the Natu- fian and at least a thousand years be- fore their ultimate collapse and the emergence of social inequality in the Late Pre-Pottery Neolithic. Like oth- ers who use the appearance of mor- phological domesticates as a thresh- old moment in their causal scenarios,84 Kuijt makes the common mistake of conflating morphological change with domestication. Once the artificial boundary crossing of mor- phological change is removed and fo- cus is more properly placed on the evolving relationship between hu- mans and plants and animals, then a more complex and ultimately more satisfying picture emerges of the syn-
  • 69. chronous and mutual reinforcing so- cial and economic forces that shaped the Neolithic Near East. The partnership between humans and resurgent plant and animal re- sources that began the domestication process after the Last Glacial Maxi- mum provided the measure of re- source security and predictability needed to establish nucleated seden- tary communities bound together by an ethos of balanced reciprocity and increasingly elaborate rules of social order and altered world views. By the same token, the goal of maintaining community was probably a factor contributing to the subsequent inten- sification of evolving domestication relationships, with their attendant im- pacts on both the human and the plant and animal partners. The do- mestication process, in turn, created resources that were more amenable to ownership and restricted access, ulti- mately contributing to the overthrow of the egalitarian communities they initially helped foster and maintain. Thus, rather than a single forcing mechanism, it seems more likely that the trajectory of plant and ani- mal domestication in the Near East and the emergence of agriculture was shaped by various broad-scale
  • 70. factors, such as climate change, eco- nomic goals, and social opportuni- ties and constraints, interacting with highly local, contingent factors, such as the density and diversity of avail- able resources, the history of human occupation, and the agency of indi- viduals coping with their environ- ment, each other, and their universe. While the entire region was engaged in this process, the pace and the di- rection it took varied depending on the distinctive mix of these factors in the Southern and Northern Levant, the Central and Eastern Fertile Cres- cent. A similar range of factors operat- ing at the same time in other places on the planet took very different courses (Fig. 5). In Mexico, for ex- ample, squash, corn, and beans, ap- parently domesticated in different places and at different times, were minor components of mobile forag- ing strategies for millennia before the adoption of sedentism and the development of agricultural econo- mies.95 In both eastern North Amer- ica and Japan, small-seeded locally domesticated annuals were blended into the diverse economic round of stable, sedentary low-level food pro- ducers for thousands of years before the introduction of domestic crop
  • 71. plants, maize in eastern North Amer- ica and rice in Japan. Previously used in small numbers, these plants formed the foundation for the emer- gence of more fully agricultural economies and increasingly strati- fied societies.95,96 Thus, the story of domestication and agricultural origins consists of a Figure 5. Boxes indicate the general location of centers of independent plant or animal domestication. Currently, at least ten such centers are recognized around the world, making research on the origins of agriculture a truly world-wide enterprise. Future research will, no doubt, discover others. 114 Zeder ARTICLES series of complex regional puzzles shaped in unique ways by a dynamic multi-scalar range of macro- and mi- cro-forces. Attempts at explanation that champion any one of these fac- tors and deny the importance of oth- ers will not, in the long run, contrib- ute to understanding agricultural origins either as a general process or as it played out in particular in- stances. CONCLUSIONS
  • 72. As this review demonstrates, central questions about the definition, docu- mentation, and explanation of domes- tication are not easily answered. Do- mestication cannot be simply defined as either a biological or a cultural phe- nomenon, but rather needs to be seen as a form of biological mutualism transformed by the highly developed human capacity to effect behavioral change through learning and cultural transmission. Definitional approaches to domestication are most effective, then, when they focus on the evolving relationship between humans and tar- get plant or animal populations as a nexus between biology and culture, not on the manifestations or conse- quences of such relationships. Ge- netic and related morphological changes in domesticates are not defin- ing features of domestication, but are instead artifacts of evolving relation- ships that vary in their intensity and pace of development. Notions of own- ership and restructuring of social re- lations are similarly best viewed as possible results of domestication, not as central to its definition. Nor are the clear-cut thresholds that define when wild resources become domesticated ones. Rather than looking for defini- tive either-or boundary conditions in defining domesticates, it is much more profitable, if more challenging,
  • 73. to look at the whole span of evolving domestic relationships as they operate over various scales of investment on the part of both human and plant or animal partners. There are also no easy, universally applicable ways to document domes- tication. Instead, documenting do- mestication requires a clear under- standing of the species-specific linkage between a proposed marker of domestication and the stage of the un- folding domestication process it is held to mark. It also requires recog- nizing that markers vary depending on the biology of the species involved and the cultural context of human populations engaged in the domesti- cation process. Above all, effective documentation means not letting the availability of new scientific tech- niques lead the search for new mark- ers without first thinking about how the process of domestication might manifest itself in whatever these tech- niques are designed to measure. So, too, causal scenarios that nar- rowly focus on single, universally ap- plicable prime-mover levers of change will never provide satisfying answers to the critical “why” questions about the origins of domestication and sub- sequent agricultural emergence. It is
  • 74. easy when dealing with complex and extended processes to draw artificial thresholds that help make the case for the primacy of whatever variable one is sponsoring as the cause of events that follow, whether it be climate change, population increase, or social and ideological transformations. Yet we have gained too sophisticated an understanding of the process of do- mestication and the means of detect- ing it to support the kind of drive-by theorizing that selectively chooses ac- commodating bits of information from individual regional scenarios to support a favored epistemology de jour. Advances in answering “why” questions about domestication and agricultural origins can only be gained through close-grained under- standing of complex multi-scalar re- gional puzzles and assessment of the commonality and the differences in the way the pieces of these different puzzles fit together.16,97 There are, then, no easy answers to central questions about domestica- tion and agricultural origins. It is no wonder that for more than 100 years this area of inquiry has held the atten- tion of archeologists working world- wide and representing all of archeolo- gy’s many and rapidly increasing subdisciplines. It is a research domain that carries broad currency with
  • 75. scholars based in biological and phys- ical sciences, social sciences, and hu- manities. It is a topic that captures the imagination of a public interested in how the familiar world around them came to be. It is a problem that truly matters. With an enhanced under- standing of the nature of the problem and an expanding array of powerful tools for studying it, there has never been a time of greater promise for pursuing challenging questions about the origin and diffusion of domesti- cates and agricultural economies in virtually all areas of the globe. REFERENCES 1 Zeder MA, Bradley DG, Emshwiller E, Smith BD, editors. 2006. Documenting domestication: new genetic and archaeological paradigms. Berkeley: University of California Press. 2 Clutton-Brock J. 1994. The unnatural world: behavioural aspects of humans and animals in the process of domestication. In: Manning A, Ser- pell JA, editors. Animals and human society: changing perspectives. London: Routledge. p 23– 35. 3 Ducos P. 1978. “Domestication” defined and methodological approaches to its recognition in faunal assemblages. In: Meadow RH, Zeder MA, editors. Approaches to faunal analysis in the Middle East. Peabody Museum Bulletin 2. Cam- bridge, MA: Harvard University. p 49 –52. 4 Hemmer H. 1990. Domestication: the decline
  • 76. of environmental appreciation. Cambridge: Cam- bridge University Press. 5 Bökönyi S. 1989. Definitions of domestication. In: Clutton-Brock J, editor. The walking larder: patterns of domestication, pastoralism, and pre- dation. Cambridge: Unwin. p 1– 4. 6 Ingold T. 1996. Growing plants and raising an- imals: an anthropological perspective on domes- tication. In: Harris D, editor. The origins and spread of agriculture and pastoralism in Eurasia. Washington, DC: Smithsonian Institution Press. p 12–24. 7 O’Connor TP. 1997. Working at relationships: another look at animal domestication. Antiquity 71:149 –156. 8 Moray D. 1994. The early evolution of the do- mestic dog. Am Sci 82:336 –347. 9 Rindos D. 1984. The origins of agriculture: an evolutionary perspective. Orlando: Academic Press. 10 Harris DR. 1996. Introduction: themes and concepts in the study of early agriculture. In: Harris DR, editor. The origins and spread of ag- riculture and pastoralism in Eurasia. Washing- ton: Smithsonian Institution Press. p 1–11. 11 Uerpmann HP. 1996. Animal domestication: accident or intention? In: Harris DR, editor. The origins and spread of agriculture and pastoral- ism in Eurasia. Washington: Smithsonian Insti- tution Press. p 227–237. 12 Zeder MA. 2006. A critical examination of markers of initial domestication in goats (Capra hircus). In: Zeder MA, Bradley DG, Emshwiller E, Smith BD, editors. Documenting domestica- tion: new genetic and archeological paradigms. Berkeley: University of California Press. p 181–
  • 77. 208. 13 Jarman MR. 1976. Early animal husbandry. Philos Trans R Soc London B 275:85–97. 14 Hecker H. 1982. Domestication revisited: its implications for faunal analysis. J Field Archeol 9:217–236. 15 Smith BD. 2006. Documenting domestic ARTICLES Domestication of Plants and Animals 115 plants in the archaeological record. In: Zeder MA, Bradley DG, Emshwiller E, Smith BD, edi- tors. Documenting domestication: new genetic and archaeological paradigms. Berkeley: Univer- sity of California Press. p 15–24. 16 Smith BD. 2001. Low level food production. J Archeol Res 9:1–43. 17 Schultz TR, Mueller UG, Currie CR, Rehner AH. 2005. Reciprocal illumination: a comparison of agriculture in humans and in fungus-growing ants. In: Vega F, Balckwell M, editors. Ecological and evolutionary advances in insect-fungal asso- ciations. New York: Oxford University Press. p 149 –190. 18 Anderson MK. 2005. Tending the wild: Native American knowledge and the management of California’s natural resources. Berkeley: Univer- sity of California Press. 19 Turner NJ, Duer D. 2005. Keeping it living. Seattle: University of Washington Press.
  • 78. 20 Zeder MA, Emshwiller E, Smith BD, Bradley DG. 2006. Documenting domestication: the in- tersection of genetics and archaeology. Trends Genet 22:134 –155. 21 Blumler M, Bryne R. 1991. The ecological genetics of domestication and the origins of ag- riculture. Curr Anthropol 32:23–54. 22 Harlan JJ. 1973. Comparative evolution of ce- reals. Evolution 27:311–325. 23 Heiser C. 1988. Aspects of unconscious selec- tion and the evolution of domesticated plants. Euphytica 37:77–85. 24 Piperno DR, Pearsall DM. 1998. The origins of agriculture in the lowland neotropics. San Diego: Academic Press. 25 Piperno DR. 2006. Phytolith analysis in ar- chaeology and environmental history. Walnut Creek: Altamira Press. 26 Denham TP, Haberle SG, Lentfer C, Fullagar R, Field J, Therin M, Porch N, Winsborough B. 2003. Origins of agriculture at Kuk Swamp in the highlands of New Guinea. Science 301:189 –193. 27 Piperno DR. 1993. Phytolith and charcoal records from deep lake cores in the American tropics. In: Pearsall DM, Piperno DR, editors. Current research in phytolith analysis: applica- tions in archaeology and paleoecology. MASCA Research Papers in Science and Archaeology Vol
  • 79. 10. Philadelphia: The University Museum of Ar- chaeology and Anthropology, University of Penn- sylvania. p 58 –71. 28 Colledge S. 2002. Identifying pre-domestica- tion cultivation using multivariate analysis: pre- senting the case for quantification. In: Cappers RTJ, Bottema S, editors. The dawn of farming in the Near East. Studies in Near Eastern Produc- tion, Subsistence, and Environment 6. Berlin: Ex Oriente. p 141–152. 29 Smith BD. 2002. Rivers of change. Washing- ton, DC: Smithsonian Institution Press. 30 Mdiba C, De Langhe E Vrydaghs L, Doutrele- pont H, Swennen R, Van Neer W, de Maret P. 2006. Phytolith evidence for the early presence of domesticated banana. In: Zeder MA, Bradley DG, Emshwiller E, Smith BD, editors. Documenting domestication: new genetic and archaeological paradigms. Berkeley: University of California Press. p 68 – 81. 31 Zeder MA. 2006. Archaeological approaches to documenting animal domestication. In: Zeder MA, Bradley DG, Emshwiller E, Smith BD, edi- tors. Documenting domestication: new genetic and archaeological paradigms. Berkeley: Univer- sity of California Press. p 171–180. 32 Clutton-Brock J. 1999. Domesticated animals, 2nd ed. London: British Museum of Natural His- tory. 33 Kruska D. 1988. Mammalian domestication
  • 80. and its effect on brain structure and behavior. In: Jerison HJ, Jerison I, editors. Intelligence and evolutionary biology. New York: Springer-Ver- lag. p 211–250. 34 Moray D. 1992. Size, shape, and development in the evolution of the domestic dog. J Archaeol Sci 19:181–204. 35 Ervynck A, Dobney K, Hongo H, Meadow RH. 2002. Born free!: new evidence for the status of pigs from Çayönü Tepesi, Eastern Anatolia. Pa- léorient 27:47–73. 36 Shaffer VM, Reed CA. 1972. The co-evolution of social behavior and cranial morphology in sheep and goats (Bovidae, Caprini). Fieldiana Zool 61. Chicago: Field Museum of Natural His- tory. 37 Meadow RH. 1989. Osteological evidence for the process of animal domestication. In: Clutton- Brock J, editor. The walking larder: patterns of domestication, pastoralism, and predation. Cam- bridge: Unwin. p 80 –90. 38 Noddle B. 1974. Age of epiphyseal closure in feral and domestic goats and ages of dental erup- tion. J Archaeol Sci 1:195–204. 39 Uerpmann H-P. 1978. Metrical analysis of faunal remains from the Middle East. In: Meadow RH, Zeder MA, editors. Approaches to faunal analysis in the Middle East. Peabody Mu- seum Bulletin 2. Cambridge, MA: Harvard Uni- versity. p 41– 45. 40 Zohary D, Tchernov E, Horwitz LK. 1998. The role of unconscious selection in the domestica- tion of sheep and goats. J Zoo 245:129 –135. 41 Hole F, Flannery KV, Neely JA. 1969. Prehis- tory and human ecology on the Deh Luran Plain.