up per palaeo lithic an i mal ex ploi ta tion at klissoura cave 1 in

Transcription

up per palaeo lithic an i mal ex ploi ta tion at klissoura cave 1 in
Eurasian Prehistory, 7 (2): 107–132.
UPPER PALAEOLITHIC ANIMAL EXPLOITATION
AT KLISSOURA CAVE 1 IN SOUTHERN GREECE:
DIETARY TRENDS AND MAMMAL TAPHONOMY
Britt M. Starkovich* and Mary C. Stiner
School of Anthropology, University of Arizona, 1009 E. South Campus Drive, Tucson, AZ 85721-0030 USA;
(*) [email protected]
Abstract
The faunal remains from the Upper Paleolithic (UP) through Mesolithic layers at Klissoura Cave 1 (Prosymna) in
Peloponnese, Greece, were examined to understand changes in hominid diets over the course of the sequence, as well as
the human and non-human taphonomic processes that affected the assemblages. The range of hunted species varied with
time in response to a combination of environmental factors and human hunting pressures. Evidence for the latter includes a
shift in small game hunting from mainly high-ranking slow moving species (tortoises) in the Early UP to greater use of low
ranking fast-moving species (hares and birds) and the eventual inclusion of land snails in later Paleolithic and Mesolithic
diets. Fallow deer dominate the ungulate remains throughout the sequence. The ungulate faunas are particularly diverse,
however, in the earliest Aurignacian layers and again in the Epigravettian and Mesolithic layers. The evidence for bone
modification is uniformly anthropogenic, with rare if any indication of non-human taphonomic processes. Body part analyses of the fallow deer remains reveal a paucity of axial elements below the neck, which probably were discarded at kill
sites. Hare body part profiles consistently lack foot and vertebral elements, which may also reflect field processing or spatially discrete (unexcavated) processing areas on-site. The highly varied prey spectra of the Upper Paleolithic–Mesolithic
occupations show that a wide range of economic activities generally took place at this site.
Key words: Zooarchaeology, vertebrate taphonomy, Aurignacian, Uluzzian, Mesolithic, faunal turnover, small game
hunting.
INTRODUCTION
The faunal remains from Klissoura Cave 1
provide unique information on Upper Paleolithic
diets and ecology in southern Greece. The site is
situated at the interface of the hilly Berbati Valley
and the Argive Plain, which provided diverse
habitats for small and large game animals. This
paper presents findings on the Early Upper Paleolithic through Mesolithic faunas, relating to trends
in human subsistence, regional ecology and
taphonomic processes at the site. Earlier studies
of small samples of the Klissoura 1 faunas, reported in Koumouzelis et al. (2001) and Tomek
and Bocheñski (2002), have provided critical
starting points for the research discussed here.
The present study represents a near-full analysis
of the Early Upper Paleolithic and later layers
excavated at Klissoura Cave 1 to date, including
vertebrate and land snail exploitation. Of special
interest in this presentation are changes in prey
choice potentially as a function of rising human
population density versus changes in natural biotic diversity in eastern Peloponnese. Ungulate
body part representation and taphonomic observations are also examined for evidence of site
function and variation in site occupation intensity.
CONCEPTUAL BACKGROUND
Two major issues that have emerged in recent
studies of Upper Paleolithic subsistence in the
Mediterranean region rely on observations about
108
B. M. Starkovich & M. C. Stiner
the proportionality (“evenness”) of key prey items
in the diet. One of these issues concerns changes
in small game use and can be related in some instances to hunting pressure and increases in human population densities (Stiner et al., 2000;
Stiner, 2001; Stiner and Munro, 2002). The other
issue concerns the dominance of one or just a few
large game species in some faunal assemblages,
and whether this reflects natural variation in biotic diversity or specialized hunting and habitatspecific patterns of land use by humans (Straus,
1987; Gamble, 1997; Phoca-Cosmetatou, 2003b,
2005). With regard to the first issue, small game
animals clearly were important sources of supplementary meat at Klissoura Cave 1. As for the second issue, the ungulate assemblages are overwhelmingly dominated by fallow deer (Dama
dama) in some periods, and considerably less so
in others. Both of these topics are pertinent to understanding prehistoric hunting adaptations and
their ecological contexts at Klissoura 1, although
their implications for human behavior and ecology are different. This presentation attempts to
distinguish among these hypothesized effects in
the early Upper Paleolithic through Mesolithic
faunas at this site.
Stiner and colleagues (Stiner et al., 2000;
Stiner, 2001; Stiner and Munro, 2002) previously
have demonstrated a marked change in small game
use between the Middle and Upper Paleolithic in
other Mediterranean regions, the first of a series of
stages in dietary expansion that occurred among
Late Pleistocene foragers. Meat diets continued to
broaden from the Upper Paleolithic to Epipaleolithic, though ungulates remained the main source
of meat in all periods. In the Middle Paleolithic,
only slow-moving or sessile small animals, such
as tortoises and shellfish, were collected in significant quantities to fill gaps in the availability of
large game. These slow small animals continued
to be exploited in the early Upper Paleolithic and
later periods, but a variety of quick small mammals and game birds became ever more important.
By the Epipaleolithic, quick small prey were a
major part of the meat diet at many Mediterranean
sites. This phenomenon is also known from
higher latitude regions of Europe, but the trend
often is not clearly expressed until the Magdalenian and later (Berke, 1984; Jochim, 1998;
Costamagno et al., 2008).
An increasing reliance on quick small game is
behaviorally significant because the costs of capture (i.e. technological investments) are higher
than for slow-moving or stationary small prey of
roughly the same body size. Ethnographic accounts show that, while large animals are difficult
to capture, the potential yields are also very high
(e.g., Kelly, 1995). Slow-moving small game animals are only somewhat less attractive than large
game, because the cost of procurement is so low.
Among recent foragers, capture costs are normally are reduced technologically, and the foraging equipment requires significant effort to produce and maintain.
Prey choice models provide a valuable framework for testing hypotheses about the circumstances in which diets may constrict or expand
(Stephens and Krebs, 1986). Considerable flexibility is expected within the adaptation of any human forager, but evolutionary changes are implied by long-term trends the diet breadth. A broadened diet tends to occur when high-ranked
game animals become less abundant in the environment. Lower-ranked resources are exploited to
compensate for the decline in the optimal (sensu
highest yield) prey types. Evidence of expanding
diet breadth does not require that new species be
added to the diet, although this may also occur.
Rather, the most important criterion is greater
proportions of low-ranked prey species alongside
high-ranked species (Colson, 1979; Stephens and
Krebs, 1986; Kelly, 1995). In addition to prey
composition, intensified processing of prey carcasses may also indicate rising pressures on the
supply of animal foods in the environment (Stiner,
2002a).
Another question about the Upper Paleolithic
record that historically has received much attention concerns specialized forms of large game
hunting. It is argued by some archaeologists that
specialized hunting should be evidenced by a narrow focus on just one or two ungulate species.
Hypotheses about specialized hunting are usually
accompanied by interpretations of site function,
focusing on hunting camps and use of highland
areas. Tests of this hypothesis therefore require
information on natural species diversity in the region and the habitats that could be exploited by
the foragers. Some archaeologists propose that
hunting specialization is a hallmark of the Upper
Upper Palaeolithic animal exploitation at Klissoura Cave 1
Paleolithic or fully modern human behavior (Mellars, 1973, 1989). However, recent zooarchaeological studies have shown that (1) mono- specific
assemblages also occur in some Middle Paleolithic sites throughout Europe and southwest Asia
(see review by Gaudzinski, 1995), and (2) though
one species often dominates an assemblage, uneven species representation is not necessarily evidence of foraging specialization. Certain gregarious species were targeted during seasonal migrations in some regions, whereas a wider variety of
taxa were exploited during rest of the year by the
same cultures (e.g. in southern France, Surmely et
al., 2003; Costamagno, 2004; Costamagno et al.,
2006). Moreover, mammal species diversity naturally declines with latitude, and climate change
produces a similar effect through time for any
given region of the world.
Straus (1987) has examined ibex hunting as a
unique kind of specialization at sixteen sites in
Cantabrian Spain and the French Pyrenees. He
finds that ibex-dominated faunas often occur at
high altitudes and during the Late Glacial, and
they tend to associate with evidence for limited
on-site activities. At Klithi Cave in northern
Greece, Gamble (1999) found evidence for short
occupation spans, low prey species diversity, a
lack of carnivores, and biases favoring high-utility animal body portions, not unlike the situation
at six of the Iberian sites reviewed by Straus
(1987). In Italy, Phoca-Cosmetatou (2003a,
2004a, 2005) finds that, while there do appear to
be specific tactics involved in ibex hunting, there
also is considerable variation in the contents of
ibex-dominated sites. She argues that the ecological association between ibex and high-altitude
habitats is a recent phenomenon; in the past they
also could be found on craggy ground at low altitudes (Phoca-Cosmetatou, 2004b), based on the
presence of their remains in low-lying archaeological sites. In nearly all of the Italian assemblages she examined, ibex are either the dominant
species or they are virtually absent, suggesting
that exploitation of ibex noted by Straus (1987)
and Gamble (1997) also occurred in Italy (PhocaCosmetatou, 2004a).
The faunas from Klissoura Cave 1 contain a
variety of ungulate species, each with somewhat
distinct habitat requirements. Some of the species,
such as the extinct European wild ass (Equus
109
hydruntinus), would have preferred lower grassland areas, whereas ibex (Capra cf. ibex) and
chamois (Rupicapra rupicapra) tend to inhabit
craggy terrain. The position of the site at the interface of lowland and rugged hills must have contributed to the patterns of diversity in the
Klissoura 1 prey assemblages. The duration of the
occupation might also have affected the diversity
of large mammals in the faunal assemblages.
Klissoura 1 served as a kind of residential camp
during most or all of the Upper Paleolithic occupations, based on the presence of diverse artifact
forms and features, but the duration of occupations may have varied a great deal through time.
Layer IV represents a particularly intense occupation, whereas Layer V is more ephemeral (see
Koumouzelis et al., 2001; Karkanas, this issue;
Stiner, this issue).
SAMPLES AND METHODS
About 11,000 identified vertebrate specimens
(NISP) were analyzed from the Upper Paleolithic
through Mesolithic layers of Klissoura Cave 1.
Nearly all of the excavated bones were studied,
except from layers IIIe, IIIg and III” where large
subsamples were examined. This study excludes
the small samples reported in Koumouzelis et al.
(2001) and the avian fauna studied by Tomek and
Bocheñski (2002), due to differences in the criteria for sample selection. Our goal here is to determine the full prey spectrum and the relative proportions of the many classes of prey. Discrepancies between layer totals in Table 1 and later tables are from the exclusion of shattered dental
specimens that potentially inflate species frequencies. Taphonomic data on the mammalian remains
are presented to establish their associations with
human activities in the cave. Apart from burning
frequencies, taphonomic observations for the bird
remains are reported only in Bocheñski and
Tomek (this issue). Though no full count of land
snail shells is available, changes in relative frequencies of land snails in squares AA4-BB4 of
the stratigraphic column and modification data
are also provided below.
The largest faunal assemblages come from
layer IV and parts of the layer III group. Layer
definitions follow the sequences defined by Karkanas (this issue) combined with technological
B. M. Starkovich & M. C. Stiner
110
Table 1
NISP and MNI of all species in the Upper Paleolithic through Mesolithic sequences at Klissoura
Mesolithic
Epigravettian
(3-5a)
Med.
UP
backed-bladelet (non-Aurigncian)
(IIa-d)
(III')
Aurignacian
(upper)
(III")
(IIIb-d)
NISP
MNI
NISP
MNI
NISP
MNI
NISP
MNI
NISP
MNI
Small ungulate
0
NA
1
NA
5
NA
8
NA
0
NA
Roe deer (Capreolus capreolus)
1
1
0
0
0
0
2
1
0
0
Chamois (Rupicapra rupicapra)
0
0
0
0
0
0
0
0
0
0
Medium ungulate
47
NA
10
NA
506
NA
884
NA
462
NA
Fallow deer (Dama dama)
23
1
6
1
284
4
511
7
363
4
Ibex (Capra cf. ibex)
13
1
2
1
26
1
14
1
6
1
Large ungulate
9
NA
3
NA
46
NA
71
NA
31
NA
Red deer (Cervus elaphus)
9
1
3
1
10
1
18
1
2
1
Wild pig (Sus scrofa)
0
0
1
1
3
1
0
0
0
0
European ass (Equus hydruntinus)
11
1
0
0
19
1
16
1
8
1
Extra-large ungulate
1
NA
0
NA
2
NA
1
NA
1
NA
Aurochs (Bos primigenius)
6
1
1
1
3
1
7
1
1
1
Small carnivore
0
NA
0
NA
0
NA
0
NA
0
NA
Stone/Pine marten (Martes foina, M.
martes)
0
0
0
0
0
0
0
0
0
0
Wild cat (Felis silvestris)
0
0
0
0
5
1
0
0
3
1
Red fox (Vulpes vulpes)
0
0
0
0
1
1
1
1
7
1
Eurasian lynx (Lynx lynx)
0
0
0
0
0
0
0
0
0
0
Large carnivore
0
NA
0
NA
0
NA
0
NA
0
NA
Wolf (Canis lupus)
0
0
0
0
0
0
0
0
0
0
Hyena sp. (Hyaena sp.) (coprolite)
0
0
0
0
0
0
1
0
0
0
Leopard (Panthera pardus)
1
1
0
0
1
1
1
1
0
0
Eastern hedgehog (Erinaceus sp.)
0
0
0
0
0
0
0
0
0
0
Caucasian squirrel (Sciurus anomalus)
0
0
0
0
0
0
0
0
0
0
123
5
34
2
289
6
75
2
169
3
Indetermined snake
0
0
0
0
0
0
0
0
0
0
Tortoise (Testudo sp.)
0
0
0
0
3
1
1
1
3
1
Small birds
0
NA
0
NA
0
NA
0
NA
0
NA
Medium birds
2
NA
4
NA
9
NA
1
NA
3
NA
Rock partridge (Alectoris graeca)
10
1
2
1
102
7
5
1
74
3
Large birds
0
NA
0
NA
5
NA
0
NA
4
NA
Eurasian eagle owl (Bubo bubo)
0
0
0
0
0
0
0
0
0
0
Very large birds
0
NA
0
NA
1
NA
0
NA
0
NA
Ungulate
Carnivore
Small mammals
European hare (Lepus europaeus)
Reptile
Birds
Great bustard (Otis tarda)
Total
0
0
0
0
39
1
0
0
21
1
256
13
67
8
1359
27
1617
18
1158
18
Upper Palaeolithic animal exploitation at Klissoura Cave 1
111
Table 1 continued
Aurignacian
(middle)
Aurignacian
(lower)
III(e-g)
Early UP (Uluzzian)
(IV)
Total
(V)
NISP
MNI
NISP
MNI
NISP
MNI
NISP
MNI
Small ungulate
10
NA
46
NA
3
NA
73
NA
Roe deer (Capreolus capreolus)
0
0
5
1
0
0
8
3
Chamois (Rupicapra rupicapra)
0
0
14
1
0
0
14
1
Medium ungulate
766
NA
973
NA
60
NA
3708
NA
Fallow deer (Dama dama)
567
6
318
3
56
1
2128
27
Ibex (Capra cf. ibex)
22
1
236
2
4
1
323
9
Large ungulate
72
NA
507
NA
15
NA
754
NA
Red deer (Cervus elaphus)
24
1
304
3
6
1
376
10
Wild pig (Sus scrofa)
1
1
19
1
1
1
25
5
European ass (Equus hydruntinus)
28
1
179
1
11
1
272
7
Extra-large ungulate
0
NA
18
NA
0
NA
23
NA
Aurochs (Bos primigenius)
16
1
43
1
0
0
77
7
Small carnivore
1
NA
10
NA
0
NA
11
NA
Stone/Pine marten (Martes foina, M.
martes)
0
0
1
1
0
0
1
1
Wild cat (Felis silvestris)
9
1
10
1
0
0
27
4
Red fox (Vulpes vulpes)
7
1
4
1
0
0
20
5
Eurasian lynx (Lynx lynx)
1
1
0
0
0
0
1
1
Large carnivore
1
NA
2
NA
0
NA
3
NA
Wolf (Canis lupus)
1
1
0
0
0
0
1
1
Hyena sp. (Hyaena sp.) (coprolite)
0
0
0
0
0
0
1
0
Leopard (Panthera pardus)
6
1
10
1
0
0
19
5
Eastern hedgehog (Erinaceus sp.)
6
1
0
0
0
0
6
1
Caucasian squirrel (Sciurus anomalus)
1
1
1
1
0
0
2
2
137
5
353
10
26
1
1206
34
Indetermined snake
2
1
2
1
0
0
4
2
Tortoise (Testudo sp.)
12
1
17
1
38
1
74
6
Small birds
0
NA
1
NA
0
NA
1
NA
Medium birds
7
NA
15
NA
0
NA
41
NA
Rock partridge (Alectoris graeca)
2
1
5
1
2
1
202
16
Large birds
0
NA
9
NA
0
NA
18
NA
Eurasian eagle owl (Bubo bubo)
0
0
1
1
0
0
1
1
Very large birds
0
NA
5
NA
0
NA
6
NA
Ungulate
Carnivore
Small mammals
European hare (Lepus europaeus)
Reptile
Birds
Great bustard (Otis tarda)
Total
0
0
2
1
0
0
62
3
1699
26
3110
33
222
8
9488
151
112
B. M. Starkovich & M. C. Stiner
Table 2
Taxa included in each body class category and
their respective weight ranges. Masses from
Nowack (1991) and Silva and Downing (1995)
Weight range
(kg)
Small, fast-moving
Rock partridge (Alectoris graeca)
0.51–0.68
European hare (Lepus europaeus)
1.3–7
Great bustard (Otis tarda)
10–16
Small, slow-moving
Tortoise (Testudo sp.)
1–2+
Small ungulate
Roe deer (Capreolus capreolus)
15–50
Chamois (Rupicapra rupicapra)
24–50
Medium ungulate
Ibex (Capra cf. ibex)
35–150
Fallow deer (Dama dama)
40–120+
Large ungulate
Wild pig (Sus scrofa)
50–350
Red deer (Cervus elaphus)
75–340
European wild ass (Equus hydruntinus)
160–240
Very large ungulate
Aurochs (Bos primigenius)
500–1000
Small carnivore
Stone/Pine marten (Martes foina/M. martes)
0.8–2.3
Wild cat (Felis silvestris)
3–8
Red fox (Vulpes vulpes)
8–10
Eurasian lynx (Lynx lynx)
8–38
Large carnivore
Wolf (Canis lupus)
25–38
Leopard (Panthera pardus)
28–90
differences identified by Kaczanowska and others
(this issue). The eight major layer groupings,
from youngest to oldest, are as follows: Mesolithic (3-5a), Epigravettian (IIa-d), Mediterranean
backed-bladelet industry (III’), Upper Paleolithic
industry )non-Aurignacian) (III”), Aurignacian
(upper) (IIIb-d), Aurignacian (middle) (IIIe-g),
Aurignacian (lower) (IV), and the Early Upper
Paleolithic or Uluzzian (V).
The faunal remains were identified using the
skeletal reference collection of the Wiener Laboratory, American School of Classical Studies at
Athens, and unpublished electronic faunal manuals created by one of the authors (M.C.S.). Skele-
tal specimens were identified to species, genus or
body size categories (Table 2) and to the anatomical part of the skeleton. Terminology for counting
units follows Grayson (1984) and Lyman (1994),
and the coding of elements, portions-of-elements,
age criteria, and taphonomic variables follows
Stiner (Stiner, 1994, 2002b, 2004). NISP (number
of identified specimens) is the basic counting unit,
from which MNE (minimum number of elements)
and MNI (minimum number of individuals) were
derived. MNE is calculated using unique skeletal
landmarks on each element, and element representation was tabulated relative to a complete
skeletal model. MNE is important to comparisons
of prey body part representation. Portion-of-element representation (a subset of MNE) is used to
test for the differential preservation of structurally
delicate versus dense skeletal parts against independent reference data based on photon densitometry for similar mammal species (Lyman, 1994).
MNI (=MAU) is derived from the highest count
of the most commonly occurring element, divided
by the number of times it occurs in the body.
Other observations were recorded for each specimen, including fusion state in the case of bones,
wear stages for mandibular teeth, presence of burning damage and burning intensity stages (Stiner
et al., 1995), and surface damage from tool marks,
weathering, gnawing animals, and plant roots
(Fisher, 1995).
RESULTS
Species abundance and diversity
Paleolithic groups hunted a wide variety of
ungulate species in the Klissoura Gorge, upper
Berbati Valley, and Argive Plain. The main ungulate species represented in the faunas are European fallow deer (Dama dama), ibex (Capra cf.
ibex), European wild ass (Equus hydruntinus), red
deer (Cervus elaphus), and wild pig (Sus scrofa)
(Fig. 1, Table 1). Two small ungulate species, roe
deer (Capreolus capreolus) and chamois (Rupicapra rupicapra), also occur in very low numbers
in layer IV, and roe deer is present in the Mesolithic. Small game animals such as European hare
(Lepus europaeus) and large ground birds
(mainly rock partridge and great bustard; see
Bocheñski and Tomek, this issue) were also important food sources in the Upper Paleolithic
Upper Palaeolithic animal exploitation at Klissoura Cave 1
Fig. 1.
ble 1
113
NISP of the main economic species in the Upper Paleolithic sequences at Klissoura Cave 1, data from ta-
through Mesolithic. Tortoises (Testudo sp.) were
important only during the formation of layer V.
Carnivore remains are present at low frequencies in every assemblage. These are mainly small
fur-bearing carnivores, especially red fox (Vulpes
vulpes), wild cat (Felis silvestris) and Eurasian
lynx (Lynx lynx), along with one or two species of
marten (Martes martes, M. foina). Rare large carnivore remains are mainly from leopard (Panthera pardus), along with scant remains of wolf
(Canis lupus) in some layers. The general scarcity
of carnivore remains in Klissoura Cave 1, very
low frequencies of gnawing damage (Table 3 and
see below), and the presence of burning damage
and transverse fractures on some of the carnivore
bones indicate that the carnivores were human
prey. One coprolite in layer III” indicates that hyenas visited the cave, but their presence was
ephemeral. There are absolutely no indications of
carnivore dens in the site.
The faunal samples from some layers are
comparatively small (NISP < 300, Table 1), par-
ticularly those for the Mesolithic (3-5a), Epigravettian (IIa-d) and Early Upper Paleolithic or
Uluzzian (V). The sample from layer IV is exceptionally large by contrast. Sample size differences
certainly account for some of the variation in the
number of species (richness) among the stratigraphic layers. The Reciprocal of Simpson’s Index,
(1/D) or 1/S(ri)2, where r is the proportion of
each prey type for the array i in an assemblage
(Simpson, 1949; Levins, 1968), evaluates species
richness and evenness (i.e. diversity) and corrects
for sample size differences among the faunal assemblages. There are no clear trends in overall
species diversity (Table 5, rs = 0.548, p = 0.160,
n = 8), ungulate species diversity (rs = –0.095, p =
0.823, n = 8) or small game species diversity (rs =
0.371, p = 0.365, n = 8). Significant variation exists within the ungulate assemblage, however,
with high diversity in Aurignacian layer IV and
again in the Epigravettian (IIa-d) and Mesolithic
(3-5a) (index = 3.31–4.41). Ungulate species diversity is consistently lower in the intermediate
B. M. Starkovich & M. C. Stiner
114
Table 3
Non-human taphonomic alterations at Klissoura Cave 1
Mesolithic
Epigravettian
(3-5a)
(IIa-d)
Med. backed-bladelet UP (non-Aurignacian)
(III')
(III")
n
%
n
%
n
%
n
%
Gnawing
none
287
99.7
77
100
1447
99.7
1791
99.9
carnivore
1
0.3
0
0
4
0.3
1
0.1
rodent
0
0
0
0
0
0
0
0
Total
288
100
77
100
1451
100
1792
100
Weathering
none
288
100
77
100
1449
99.9
1791
99.9
many cracks, most "open"
0
0
0
0
0
0
0
0
some exfoliation
0
0
0
0
0
0
0
0
advanced exfoliation
0
0
0
0
2
0.1
0
0
chemical weathering
0
0
0
0
0
0
1
0.1
288
100
77
100
1451
100
1792
100
Total
Fig. 2.
Prey group NISP by sequence
layers and in the Early Upper Paleolithic, where
fallow deer dominates to a great degree (index =
1.09–1.84).
Figure 2 presents NISP values according to
more general prey groups by layer (see Table 2
for the species included in each category). From
this perspective, a major shift occurs between
layer III’ and the Epigravettian (IIa-d); small
quick animals occur at high frequencies in the
later periods and medium ungulates dominate the
earlier periods.
A variety of small game animals were hunted
throughout the Upper Paleolithic and Mesolithic
periods. Hares were exploited in all periods but
their importance increased with time, whereas tortoises were a significant food item only in the
very beginning of the UP sequence (Fig. 1).
Ground birds such as rock partridge (Alectoris
graeca) and the great bustard (Otis tarda) gained
importance in the upper Aurignacian layers
(IIIb-d) and the Mediterranean backed-bladelet
industry (III’) (see also Bocheñski and Tomek,
this issue). The frequencies of partridges and bustards rise and fall together across the layers (Fig.
3). Both species prefer more open and grassy habitats (Vavalekas et al., 1993; Handrinos and
Akriotis, 1997), and changes in their importance
in the faunas may indicate a shift in plant communities near the Klissoura Gorge.
The relative frequencies of the common small
game species are plotted by period in Figure 3.
The four species – hares, partridges, bustards, and
Upper Palaeolithic animal exploitation at Klissoura Cave 1
115
Table 3 continued
Aurignacian
(upper)
Early UP
(middle)
(IIIb-d)
(lower)
(IIIe-g)
(Uluzzian)
(IV)
(V)
Total
n
%
n
%
n
%
n
%
n
%
Gnawing
none
1195
99.9
1893
99.8
4134
100
355
99.7
11179
99.9
carnivore
1
0.1
4
0.2
1
0
1
0.3
13
0.1
rodent
0
0
0
0
0
0
0
0
0
0
Total
1196
100
1897
100
4135
100
356
100
11192
100
1196
100
1895
99.9
4110
99.4
356
100
11162
99.7
many cracks, most "open"
0
0
0
0
1
0
0
0
1
0
some exfoliation
0
0
2
0.1
6
0.1
0
0
8
0.1
advanced exfoliation
0
0
0
0
5
0.1
0
0
7
0.1
chemical weathering
0
0
0
0
13
0.3
0
0
14
0.1
1196
100
1897
100
4135
100
356
100
11192
100
Weathering
none
Total
Fig. 3.
Small game by sequence
tortoises – have very different life history characteristics, habitat requirements, and flight responses, and the bustard is further distinguished
by its exceptionally large size (Table 2). The most
productive small prey, hares and partridges, were
used heavily by Upper Paleolithic and Mesolithic
hunters in all periods except the Early UP (layer
V). Tortoises are slow-moving and for this reason
have high return rates. Tortoises are slow to develop, however, so a population is easily impacted by human overexploitation (Lambert, 1982;
Blasco et al., 1986-87; Hailey et al., 1988; Stiner
et al., 2000). Climate variation would never have
been great enough to extirpate Mediterranean tortoises from southern Greece, and in fact they were
important throughout much of the Middle and
Late Pleistocene in the eastern Mediterranean and
Middle East (Stiner, 1994, 2005; Speth and
Tchernov, 2002; Reynaud Savioz and Morel,
2005). The disappearance of tortoises in the later
UP at Klissoura Cave 1 therefore may be a result
of the suppression of these populations by human
foragers. Great bustard provides a higher return
than hare or partridge because of their large size.
These large birds were never common in the diet,
however, and heavy exploitation would not have
been sustainable for long. Bustards have low reproductive and development rates, in marked contrast to the high reproductive and development
rates of hares and partridges (Stiner et al., 2000).
B. M. Starkovich & M. C. Stiner
116
Table 4
Burn frequency and degree by species category at Klissoura Cave
Medium
ungulate
Large
ungulate
n
%
n
%
n
%
n
%
n
%
n
%
n
%
Unburned
74
89.2
28
96.6
0
NA
107
87
12
100
1
100
222
89.5
< 1/2 carbonized
(black)
2
2.4
1
3.4
0
NA
1
0.8
0
0
0
0
4
1.6
>1/2 carbonized
2
2.4
0
0
0
NA
1
0.8
0
0
0
0
3
1.2
fully carbonized
0
0
0
0
0
NA
13
10.6
0
0
0
0
13
5.2
<1/2 calcined (white)
1
1.2
0
0
0
NA
1
0.8
0
0
0
0
2
0.8
>1/2 calcined
2
2.4
0
0
0
NA
0
0
0
0
0
0
2
0.8
Tortoise sp.
European
hare
Birds
Carnivores
Total
Mesolithic (3-5a)
fully calcined
2
2.4
0
0
0
NA
0
0
0
0
0
0
2
0.8
Total
83
100
29
100
0
NA
123
100
12
100
1
100
248
100
Unburned
17
94.4
6
85.7
0
NA
32
94.1
5
83.3
0
NA
60
92.3
< 1/2 carbonized
(black)
1
5.6
0
0
0
NA
1
2.9
1
16.7
0
NA
3
4.6
>1/2 carbonized
0
0
0
0
0
NA
1
2.9
0
0
0
NA
1
1.5
fully carbonized
0
0
0
0
0
NA
0
0
0
0
0
NA
0
0
<1/2 calcined (white)
0
0
1
14.3
0
NA
0
0
0
0
0
NA
1
1.5
>1/2 calcined
0
0
0
0
0
NA
0
0
0
0
0
NA
0
0
fully calcined
0
0
0
0
0
NA
0
0
0
0
0
NA
0
0
Total
18
100
7
100
0
NA
34
100
6
100
0
NA
65
100
Epigravettian (IIa-d)
Med. backed-bladelet (III')
Unburned
681
83.4
66
85.7
4
100
256
88.6
150
96.2
7
100
1164
86.2
< 1/2 carbonized
(black)
16
2
2
2.6
0
0
2
0.7
1
0.6
0
0
21
1.6
>1/2 carbonized
23
2.8
2
2.6
0
0
6
2.1
0
0
0
0
31
2.3
fully carbonized
53
6.5
4
5.2
0
0
20
6.9
3
1.9
0
0
80
5.9
<1/2 calcined (white)
16
2
3
3.9
0
0
2
0.7
1
0.6
0
0
22
1.6
>1/2 calcined
11
1.3
0
0
0
0
3
1
0
0
0
0
14
1
fully calcined
17
2.1
0
0
0
0
0
0
1
0.6
0
0
18
1.3
Total
817
100
77
100
4
100
289
100
156
100
7
100
1350
100
1092
77.5
83
79
0
NA
52
69.3
6
100
0
NA
1233
77.3
58
4.1
3
2.9
0
NA
2
2.7
0
0
0
NA
63
3.9
>1/2 carbonized
82
5.8
6
5.7
0
NA
10
13.3
0
0
0
NA
98
6.1
fully carbonized
101
7.2
10
9.5
0
NA
5
6.7
0
0
0
NA
116
7.3
<1/2 calcined (white)
31
2.2
2
1.9
0
NA
2
2.7
0
0
0
NA
35
2.2
>1/2 calcined
23
1.6
1
1
0
NA
1
1.3
0
0
0
NA
25
1.6
fully calcined
22
1.6
0
0
0
NA
3
4
0
0
0
NA
25
1.6
1409
100
105
100
0
NA
75
100
6
100
0
NA
1595
100
UO (non-Aurignacian) (III")
Unburned
< 1/2 carbonized
(black)
Total
Upper Palaeolithic animal exploitation at Klissoura Cave 1
117
Table 4 continued
Medium
ungulate
Large
ungulate
n
%
n
%
n
%
n
%
n
%
n
%
n
%
Unburned
724
87.1
36
87.8
2
66.7
152
89.9
100
98
9
90
1023
88.5
< 1/2 carbonized
(black)
24
2.9
2
4.9
1
33.3
4
2.4
2
2
0
0
33
2.9
>1/2 carbonized
30
3.6
2
4.9
0
0
2
1.2
0
0
1
10
35
3
fully carbonized
25
3
1
2.4
0
0
2
1.2
0
0
0
0
28
2.4
<1/2 calcined (white)
13
1.6
0
0
0
0
3
1.8
0
0
0
0
16
1.4
>1/2 calcined
7
0.8
0
0
0
0
3
1.8
0
0
0
0
10
0.9
Tortoise sp.
European|
hare
Birds
Carnivores
Total
Aurignacian (IIIb-d)
fully calcined
Total
8
1
0
0
0
0
3
1.8
0
0
0
0
11
1
831
100
41
100
3
100
169
100
102
100
10
100
1156
100
Aurignacian (IIIe-g)
Unburned
1154
85.2
101
80.8
17
70.8
123
89.8
9
100
24
92.3
1428
85.2
< 1/2 carbonized
(black)
54
4
8
6.4
3
12.5
4
2.9
0
0
0
0
69
4.1
>1/2 carbonized
45
3.3
6
4.8
2
8.3
3
2.2
0
0
1
3.8
57
3.4
fully carbonized
36
2.7
6
4.8
2
8.3
4
2.9
0
0
0
0
48
2.9
<1/2 calcined (white)
33
2.4
3
2.4
0
0
0
0
0
0
0
0
36
2.1
>1/2 calcined
21
1.5
0
0
0
0
2
1.5
0
0
1
3.8
24
1.4
fully calcined
12
0.9
1
0.8
0
0
1
0.7
0
0
0
0
14
0.8
1355
100
125
100
24
100
137
100
9
100
26
100
1676
100
Total
Aurignacian (IV)
Unburned
1122
73.5
864
85.6
27
42.9
270
76.5
36
94.7
27
73
2346
77.5
< 1/2 carbonized
(black)
80
5.2
36
3.6
1
1.6
22
6.2
0
0
4
10.8
143
4.7
>1/2 carbonized
125
8.2
34
3.4
4
6.3
18
5.1
0
0
1
2.7
182
6
fully carbonized
136
8.9
54
5.4
21
33.3
30
8.5
2
5.3
5
13.5
248
8.2
<1/2 calcined (white)
30
2
12
1.2
5
7.9
5
1.4
0
0
0
0
52
1.7
>1/2 calcined
22
1.4
8
0.8
3
4.8
4
1.1
0
0
0
0
37
1.2
fully calcined
12
0.8
1
0.1
2
3.2
4
1.1
0
0
0
0
19
0.6
1527
100
1009
100
63
100
353
100
38
100
37
100
3027
100
Unburned
83
69.2
21
63.6
76
68.5
20
76.9
2
100
0
NA
202
69.2
< 1/2 carbonized
(black)
5
4.2
0
0
1
0.9
1
3.8
0
0
0
NA
7
2.4
>1/2 carbonized
20
16.7
5
15.2
6
5.4
2
7.7
0
0
0
NA
33
11.3
fully carbonized
7
5.8
4
12.1
12
10.8
2
7.7
0
0
0
NA
25
8.6
<1/2 calcined (white)
0
0
0
0
6
5.4
0
0
0
0
0
NA
6
2.1
>1/2 calcined
0
0
0
0
1
0.9
0
0
0
0
0
NA
1
0.3
fully calcined
5
4.2
3
9.1
9
8.1
1
3.8
0
0
0
NA
18
6.2
120
100
33
100
111
100
26
100
2
100
0
NA
292
100
Total
EUP (V)
Total
B. M. Starkovich & M. C. Stiner
118
Table 5
Comparison of inverse Simpson’s index for all species at Klissoura Cave 1, ungulates,
and small game species
Total
All
All
Ungulate
Ungulate Small game Small game
Culture
Layer
NISP
N-taxa
1/D
N-taxa
1/D
N-taxa
1/D
Mesolithic
3-5a
256
9
2.40
6
4.24
2
1.16
Epigravettian
IIa-d
67
7
1.98
5
3.31
2
1.12
III'
1359
13
3.46
6
1.45
4
1.95
Mediterranean
backed-bladelet
UP (non-Aurignacian)
III"
1617
12
1.58
6
1.23
3
1.16
Aurignacian (upper)
IIIb-d
1158
11
2.59
5
1.09
4
2.07
Aurignacian (middle)
IIIe-g
1699
17
2.06
6
1.34
3
1.21
Aurignacian (lower)
IV
3110
19
5.67
8
4.41
4
1.14
Early UP (Uluzzian)
V
222
8
3.82
5
1.84
3
2.05
Bustard populations rely disproportionately on
the reproductive success of older females (Morales et al., 2002); individuals can be long-lived
and population turnover rates are comparatively
low.
Land snails are a significant component of the
faunas from the younger archaeological layers of
Klissoura Cave 1. Helix figulina, a large edible
snail common to southeastern Europe, is the dominant species throughout the Upper Paleolithic
and Mesolithic. None of the land snail shells are
burned. Few land snail shells occur in the early
Upper Paleolithic layers, and none of the shells
from layers IV or V shows clear evidence of human modification. Shell sizes vary greatly, and
some of the shells display tiny perforations made
by a small predator, not unlike the condition of
specimens found on the ground surface today.
Land snail assemblages from the younger cultural
layers are biased to comparatively large individuals, with modal diameters of 2.3–2.4 cm. Seventythree to ninety-five percent of these shells have
broken lips (aperture rims) but are otherwise in
very good condition (Fig. 4). Snails dominate the
faunal remains of the Mesolithic period.
The number of snails in the cultural deposits
increases exponentially with time. Based on samples from squares AA4-BB4, snails are rare in the
Early Upper Paleolithic (layer V) and uncommon
in the lower and middle Aurignacian series (IV–
IIIe-g). Land snails become moderately abundant
in the upper Aurignacian (layer IIIb-d) and increase greatly through the Mediterranean backed-
Fig. 4. Helix figulina apertures damaged by humans
while extracting the animals from their shells
bladelet industry (layer III”), peaking the Mesolithic (3-5a). It is not clear just when in the cultural sequence land snails became an important
food source. Few if any land snail shells in the
Aurignacian (layers IV through IIIc) seem to have
been modified by humans. Moderate frequencies
of broken lips occur on land snails in the Mediterranean backed-bladelet industry (III’) and Upper
Paleolithic industry (non-Aurignacian) (III”) layers, but the damage to the shells is not nearly as
systematic or prevalent as in the Meso- lithic assemblages. The Epigravettian (layer IIa-d) represents a distinct situation in which snails are uncommon yet species diversity is high (including
significant presence of Rumina decollata, Lindholmiola cf. spectabilis, and Zonitidae spp.,
among other species) and best resembles recent
snail assemblages from the site vicinity.
The increasing use of quick and highly productive small prey types (e.g. hares, partridges)
after the early UP, despite the higher technological costs to acquire them, is probably the result of
mild over-hunting in the study area. The rising
importance of land snails in forager diets parallels
Upper Palaeolithic animal exploitation at Klissoura Cave 1
119
Table 6
Butchery and other damage on ungulate remains at Klissoura Cave 1 by stratigraphic unit
NISP
Cone
Fractures
Crushed/
Impact
% All Impact
Damage
% Tool
marks
% Worked
% NISP
Transverse
Mesolithic (3-5a)
72
0
0
0.0
0.0
1.4
37.5
Epigravettian (IIb-d)
13
0
0
0.0
0.0
0.0
46.2
Med. backed-bladelet (III')
786
2
7
1.1
1.1
0.1
32.3
UP (non-Aurignacian) (III")
1370
5
4
0.7
0.3
0.1
28.8
Aurignacian (upper) (IIIb-d)
796
3
6
1.1
0.1
0.0
27.3
Aurignacian (middle) (IIIe-g)
1249
5
6
0.9
0.3
0.5
23.8
Aurignacian (lower) (IV)
1374
14
5
1.4
0.6
1.3
22.6
Early UP (Uluzzian) (V)
110
0
2
1.8
0.0
0.0
20.9
Mesolithic (3-5a)
18
0
0
0.0
0.0
0.0
22.2
Epigravettian (IIb-d)
5
0
0
0.0
0.0
0.0
20.0
Med. backed-bladelet (III')
70
0
2
2.9
0.0
1.4
41.4
UP (non-Aurignacian) (III")
90
0
1
1.1
1.1
0.0
32.2
Aurignacian (upper) (IIIb-d)
41
0
1
2.4
0.0
0.0
41.5
Aurignacian (middle) (IIIe-g)
109
0
1
0.9
0.0
0.0
23.9
Aurignacian (lower) (IV)
843
31
5
4.3
0.9
1.3
16.4
Early UP (Uluzzian) (V)
28
1
0
3.6
0.0
0.0
14.3
Medium ungulates
Large ungulates
(% worked) refers to antler and bone fragments, most of which represent debris from tool manufacture or small fragments of exhausted tools
this trend; though they are not difficult to collect,
the cooking and extraction of the snails can be
fairly labor intensive. Such changes in small game
use may suggest an increase in human population
densities with time or an otherwise constrained
food supply, especially after the formation of
layer III”. Variation in the relative propor- tions
of hares and partridges in the faunal series, on the
other hand, may reflect climate-driven changes in
environmental conditions, though this is not
certain.
Bone modification, bone survivorship, and
ungulate body part representation
It is clear that the Upper Paleolithic and
Mesolithic faunas from Klissoura Cave 1 were
collected and modified principally by humans.
Other biological and geological processes exerted
only minor effects on the faunas, mainly in the
forms of occasional gnawing by scavengers and
natural deaths of certain birds that resided in or
near the cave (see Bocheñski and Tomek, this issue). The quality of bone preservation in the Up-
per Paleolithic and Mesolithic layers is generally
very good and, despite considerable fragmentation, most of the specimens are easily identified to
species or body size group and to skeletal element. Carnivore gnawing and weathering damage
occurs on less than one percent of the bones (Table 3), and fragile fetal elements of mammals are
present in some of the layers. Though tool marks
were observed on some of the remains (Tables 6
and 7), the presence of thin concretion coatings on
most of the bones obscured some of the butchery
damage, making a systematic study of cut marks
and other fine tool marks difficult. It is for this
reason that we are more interested in relative differences in cut mark frequencies than absolute
values. Large-scale removal of the concretions is
impractical given the amount of labor that would
be required, and the surfaces of many of the bones
would be damaged in the process.
Other evidence of carcass processing, including cone fractures, was readily apparent on ungulate and small animal remains. Patterns in these
data do not indicate changes in the intensity of
B. M. Starkovich & M. C. Stiner
120
Table 7
Butchery damage on small game at Klissoura cave 1 by cultural level
Tortoise sp.
European Hare
Carnivores
NISP
% Cut
NISP
% Cut
NISP
Mesolithic (3-5a)
0
NA
110
0
1
% Cut
0
Epigravettian (IIa-d)
0
NA
32
0
0
NA
Med. backed-bladelet (III')
4
0
233
0
5
0
UP (non-Aurignacian) (III")
0
NA
65
0
0
NA
Aurignacian (upper) (IIIb-d)
3
0
144
0.7
8
0
Aurignacian (middle) (IIIe-g)
24
4.2
115
0
21
4.8
Aurignacian (lower) (IV)
63
0
281
0
27
0
Early UP (V)
108
2.7
22
0
0
NA
Table 8
Transverse fractures on small animal limb bones (humerus, radius, femur and tibia) by cultural level
Tortoise sp.
European Hare
Carnivores
NISP shell
% SR fracture
NISP
% TR fracture
NISP
% TR fracture
Mesolithic (3-5a)
0
NA
36
83.3
0
NA
Epigravettian (IIa-d)
0
NA
11
63.6
0
NA
Med. backed-bladelet (III')
2
100.0
96
67.7
2
100.0
UP (non-Aurignacian) (III")
0
NA
31
58.1
0
NA
Aurignacian (upper) (IIIb-d)
2
100.0
61
72.1
1
0.0
Aurignacian (middle) (IIIe-g)
17
64.7
61
57.4
4
50.0
Aurignacian (lower) (IV)
54
75.9
108
59.3
0
NA
Early UP (V)
80
88.8
9
66.7
0
NA
carcass processing through time. Transverse fractures are extremely common on the long bones of
hares and carnivores (Table 8). Spiral fractures
occur on between 65 and 89 percent of tortoise
shell remains (Table 8), indicating that the tortoises’ shells were broken while the bone was
fresh. For large game, cone and impact fractures
from marrow processing occur on 1–2 percent of
medium ungulate bones, and 1–4.5 percent of
large ungulate remains (Table 6). Fractures transverse to the main axis of the bone are apparent on
over 20 percent of the ungulate remains in most
levels (Table 6). These fractures may relate to the
partitioning of certain sections of carcasses during
butchery. The medullary cavities of all medium
ungulate long bones were opened with the exception of some of the toe elements, which contain
the least amount of bone marrow (Table 9). The
intensity of marrow processing could not be
determined for large ungulates because the
sample is too small.
Burning damage rare on bones from the two
youngest layers, but is quite common on bones
from the two earliest layers (Table 4), where
many hearth features were found. Using a very
conservative criteria for determining burning
damage (blackening or calcination, but disregarding light brown coloration), bird remains seem to
be less burned overall than the mammal remains
in the Aurignacian layers. However, fracture patterns on the bird remains indicate that many were
introduced to the site by humans (see Bocheñski
and Tomek, this issue). Tortoise specimens,
though rare in the assemblages that post-date
layer V, are considerably more burned than other
remains, particularly in layer IV (Table 4). This
may be an artifact of sample size, although similarly high rates of burning have been noted on
Upper Palaeolithic animal exploitation at Klissoura Cave 1
121
Table 9
Percent of medium ungulate limb bones from Klissoura Cave 1 not opened prior to discard,
by cultural level
MNE
% Unopened
Mesolithic (3-5a)
MNE
% Unopened
Aurignacian (upper) (IIIb-d)
Femur
1
0.0
Femur
3
0.0
Humerus
1
0.0
Humerus
7
0.0
Tibia
1
0.0
Tibia
5
0.0
Metapodials
2
0.0
Metapodials
9
0.0
Radius
1
0.0
Radius
7
0.0
Scapula
0
N/A
Scapula
2
0.0
Calcaneum
0
N/A
Calcaneum
2
0.0
1st Phalanx
2
0.0
1st Phalanx
17
0.0
2nd Phalanx
2
0.0
2nd Phalanx
14
42.9
3rd Phalanx
0
N/A
3rd Phalanx
13
61.5
Epigravettian (IIa-d)
Aurignacian (middle) (IIIe-g)
Femur
0
N/A
Femur
13
0.0
Humerus
0
N/A
Humerus
9
0.0
Tibia
0
N/A
Tibia
9
0.0
Metapodials
2
0.0
Metapodials
20
0.0
Radius
0
N/A
Radius
5
0.0
Scapula
0
N/A
Scapula
2
0.0
Calcaneum
0
N/A
Calcaneum
3
0.0
1st Phalanx
0
N/A
1st Phalanx
25
0.0
2nd Phalanx
1
0.0
2nd Phalanx
20
5.0
3rd Phalanx
1
0.0
3rd Phalanx
19
57.9
Femur
5
0.0
Femur
2
0.0
Humerus
5
0.0
Humerus
2
0.0
Tibia
5
0.0
Tibia
4
0.0
Metapodials
7
0.0
Metapodials
17
0.0
Radius
2
0.0
Radius
8
0.0
Scapula
1
0.0
Scapula
2
0.0
Calcaneum
5
20.0
Calcaneum
5
0.0
1st Phalanx
17
5.9
1st Phalanx
35
5.7
2nd Phalanx
12
41.7
2nd Phalanx
26
0.0
3rd Phalanx
13
46.2
3rd Phalanx
24
20.8
Femur
4
0.0
Femur
2
0.0
Humerus
7
0.0
Humerus
1
0.0
Tibia
9
0.0
Tibia
1
0.0
Metapodials
19
0.0
Metapodials
3
0.0
Radius
14
0.0
Radius
2
0.0
Scapula
8
0.0
Scapula
1
0.0
Calcaneum
3
0.0
Calcaneum
0
N/A
1st Phalanx
27
7.4
1st Phalanx
9
0.0
2nd Phalanx
25
16.0
2nd Phalanx
2
0.0
3rd Phalanx
28
71.4
3rd Phalanx
3
1.0
Med. backed-bladelet (III')
Aurignacian (lower) (IV)
UP (non-Aurignacian) (III")
Early UP (Uluzzian) (V)
122
B. M. Starkovich & M. C. Stiner
Middle and Upper Paleolithic tortoise remains at
Kebara and Hayonim Caves in the Levant (Speth
and Tchernov, 2002; Stiner, 2005). Tortoise carapaces may have been used in close proximity to
cooking fires. In general, variation in the frequency of burning damage through the stratigraphic sequence in Klissoura Cave 1 is probably
more closely related to site use and occupation intensity than to specific cooking behaviors. We
now turn to patterns of body part representation
and causes of bias in the archaeofaunal assemblages.
Two tests evaluate the possibility of in situ
bone attrition of the mammalian remains. The first
test compares the highest tooth-based MNE to the
bone-based MNE for skull parts in the common
ungulate species in each layer (Stiner, 1994: 99–
103). The second examines observed patterns in
ungulate bone representation to independent standards of bone tissue density values obtained by
the photon-densitometry technique (Lyman,
1984, 1994; see also Lam et al. 1998 on CT
technique).
Mammal teeth are less susceptible to most destructive processes than are bones because the
mineral density of tooth enamel greatly exceeds
that of all types of bone (Currey, 1984; Lyman,
1994). Because teeth would remain within the
skull if the head is carried by hunters to a base
camp, the number of individual animals represented by the dental elements should be generally
equivalent to the number of individual animals
represented by unique bony features of the skull
(a 1:1 ratio) in the archaeofaunal assemblages. If
significant fragmentation, gnawing or other destructive processes occurred during the processing of the heads or post-depositionally, even the
more durable diagnostic bony features of the skull
should break down more readily than teeth. Table
10 shows the proportion of ungulate tooth to
bone-based skull MNE for the Early UP through
Mesolithic faunas in Klissoura Cave 1. In order to
create a more robust sample, all ungulate taxa
were combined. In general, the layers with larger
samples (MNE>10) all have near-even ratios of
tooth to bone-based MNE, with the exception of
the lower Aurignacian layer (IV), in which the
tooth-based MNE is double the bone-based MNE
for the cranium (Table 10). We have no explanation for this anomaly, since other indications of
Table 10
Comparisons of tooth-based and cranial
bone-based MNE in layers with adequate
sample sizes
Tooth
MNE
Bone
MNE
Tooth:
Bone
MNE
Mesolithic (3-5a)
5
2
2.5
Epigravettian (IIa-d)
4
1
4.0
Mediterranean
backed-bladelet (III')
8
8
1.0
UP (non_aurignacian) (III")
9
14
0.6
Aurignacian (upper) (IIIb-d)
8
9
0.9
Aurignacian (middle) (IIIe-g)
11
10
1.1
Aurignacian (lower) (IV)
40
19
2.1
Early UP (Uluzzian) (V)
4
1
4.0
bone preservation are quite good (see below). It is
possible that body parts are distributed unevenly
in layer IV, or that crania were more heavily
processed, but this cannot be evaluated from the
available sample.
Spongy bones tissues, including soft limb
epiphyses, are more likely to be destroyed than
dense long bone shafts (Binford and Bertram,
1977; Brain, 1981; Lyman, 1994). This destruction can occur from human butchering and marrow processing, carnivore ravaging or sediment
compaction (e.g. Davis, 1987; Fisher, 1995;
Lyman, 1994). Bone survivorship is examined for
each element portion, such as the head of the femur, nutrient foramen of the humeral shaft, or the
medial portion of the distal epiphysis of the tibia,
by dividing the observed frequency in the assemblage by the expected frequency for this part in a
complete skeleton. The data are then standardized
to the most commonly represented body part in
the assemblage (Binford, 1978; Lyman, 1994).
Bone density standards for American deer (Odocoileus sp.) were applied to fallow deer from
Klissoura 1 (from Lyman, 1982, 1994), and the
standards applied to European hare were developed for Canadian snow hare (from Lyman, 1982,
1984, 1994; Pavao and Stahl, 1999). In this test,
and in the discussions of body part representation
below, hare data are combined with data for indeterminate small mammals, since the great majority of these remains are very probably from hares
as well. Similarly, remains designated as medium
Upper Palaeolithic animal exploitation at Klissoura Cave 1
Table 11
Spearman’s correlation values between
survivorship and bone mineral density
for European hare and fallow deer during
the Upper Paleolithic through Mesolithic
at Klissoura Cave 1
N
rs
rs2
p
51
*0.280
0.078
0.046
Mesolithic (3-5a)
European hare
Mediterranean backed-bladelet (III')
European hare
51
*0.401
0.161
0.004
Fallow deer
80
0.038
0.001
0.735
0.083
0.007
0.462
UP (non-Aurignacian) (III")
Fallow deer
80
Aurignacian (upper) (IIIb-d)
European hare
51
0.189
0.036
0.185
Fallow deer
80
0.035
0.001
0.759
Aurignacian (middle) (IIIe-g)
European hare
51
0.258
0.067
0.068
Fallow deer
80
*0.223
0.05
0.046
Aurignacian (lower) (IV)
European hare
51
0.209
0.044
0.141
Fallow deer
80
0.153
0.023
0.174
Only larger samples are considered. Asterisks indicate a significant correlation
ungulates were combined with the taxon-specific
fallow deer data, since fallow deer was always the
most common ungulate species in the faunas. Results of a Spearman’s rank-order correlation between bone density values and percent survivorship indicate that density mediated bone destruction was very limited for ungulate remains throughout the sequence, though density-mediated processes may have affected hare bones somewhat
more in the Mediterranean backed-bladelet industry (III’) and Mesolithic (3-5a) layers (Table 11).
The comparison of observed body part representation in the Klissoura 1 faunas to independent
bone density standards is useful for identifying biases in the bone assemblages, but it cannot determine whether humans or non-human processes
caused the biases. Ethnographically, human foragers are known to be selective about which ungulate body parts will be carried over long distances, whereas small game animals tend to be
carried to the site in whole form (e.g. Binford,
1978; Bunn et al., 1988; O’Connell et al., 1988;
Yellen, 1991; Schmitt and Lupo, 1995). All of
123
these prey items would be subject both to processing for consumption on site and subsequent postdepositional processes. Transport biases aside,
the survivorship of hare bones should agree with
that of ungulate bones if post-depositional destruction was the main cause of the anatomical biases (Munro, 2004; Manne et al., 2005). If the
hare bones show no substantive indications of
density mediated attrition, but the ungulate remains do, then the cause of the biases in ungulate
body part representation are more likely to have
arisen from selective transport or bone grease rendering on site. On the other hand, evidence for
density-mediated attrition of hare bones and a
lack thereof for deer bones may have several explanations. One is that in general hare bones are
smaller and lighter than deer bones (see values in
Lyman, 1984; Pavao and Stahl, 1999) and more
subject to trampling damage. Other potential explanations might be that hares were not brought to
the sites whole, or they were processed and consumed in different areas of the site (Cochard and
Brugal, 2004). Following Munro (2004), some of
these hypotheses can be tested specifically through
comparisons of the condition of large and small
mammals in the assemblage.
The hare remains from the Mesolithic (3-5a)
and Mediterranean backed-bladelet industry (III’)
layers display a significant positive relationship
between bone density and percent survivorship.
The rs2 values indicate, however, that density-mediated attrition has the potential to explain only
8–16 percent of the variation in skeletal survivorship for the hare bones from these levels (Table 11). For fallow deer, only the remains from
the middle Aurignacian layer (IIIe-g) indicate a
significant positive relationship, and here density-mediated attrition can explain no more than
5% of the variation in deer body part representation (Table 11). Overall, deer bone survivorship is
uncorrelated or only weakly correlated with bone
density. The cranial bone-based MNE nonetheless is much lower than expected as compared to
tooth-based MNE for the lower Aurignacian (IV);
this discrepancy cannot be explained by in situ attrition. Hare bone survivorship does correlate
with bone density in the Mediterranean backedbladelet industry (III’) layer and above, so there
may be some density-mediated processes at play
in these later layers. A discussion of body part
124
B. M. Starkovich & M. C. Stiner
representation will allow us to further evaluate
some of the potential explanations for variations
in bone survivorship.
Following Stiner (1991), standardized MNE
is determined according to nine distinct anatomical regions that are logical packages for meat
transport: horn or antler (if present in the species),
head, neck, axial skeleton, upper front limb, lower
front limb, upper hind limb, lower hind limb, and
feet. The results for hare and fallow deer, two species consistently represented across periods at
Klissoura Cave 1, are presented in Figure 5.
In all of the layers, including the older Aurignacian, which lack evidence of density-mediated
attrition for hare bones, hare and ungulate elements of the neck, axial and feet portions are
poorly represented. This observation either calls
in to question the assumption that complete hare
carcasses were brought to the site, or alternatively
raises the issue of limited spatial sampling. Figures 6 and 7 provide a more detailed element-byelement comparison of foot and neck/axial bone
representation for hares from Klissoura 1. In both
figures, specific elements are plotted in descending order of the maximally dense part of each element; foot and axial elements run the full gambit
of different structural densities found in the hare
skeleton (see values in Pavao and Stahl, 1999).
There is no bias in the representation of foot elements based on structural density (Fig. 6), with
the possible exception of the Mesolithic layers
(3-5a). In general, however, the entire foot region
of hares is almost entirely absent throughout the
sequence. Cochard and Brugal (2004) argue that
lagomorph feet may be absent from an assemblage because they were removed at the kill site,
or because they were deposited in an area of a site
not used for cooking. Hare foot bones are missing
from Klissoura Cave 1, but not because of density-mediated processes. A final possibility for the
lack of hare foot elements is that they were not retrieved during excavation or screening. This is
unlikely as other small remains, such as shell
ornaments (see Stiner, this issue) were commonly
recovered.
Moving on to the head, neck and axial skeleton of the hares (Fig. 7), it is again apparent that
the absence of these elements is not explained by
bone density. As far as the post-cranial axial skeleton is concerned, the innominate is consistently
Fig. 5. Body part profiles for hare and fallow deer in
each of the sequences with a large sample size. The vertical axis represents the minimum animal units (MAU)
or standardized MNE described in Stiner (1991)
more common than other bodily regions in all of
the layers, even though it is not the densest element. The crania and mandibles are included in
the figure because they are not as severely underrepresented as the postcranial axial skeleton in
most of the layers (Fig. 5), even though they hypothetically could be removed during skinning or
transport. A possible explanation for the presence
of the innominate and head parts, along with an
absence of vertebral and rib elements, in the Klissoura 1 assemblages is that they simply break
down into unidentifiable fragments (Cochard and
Brugal, 2004). In fact, the general body part composition in the Upper Paleolithic through Mesolithic layers at Klissoura Cave 1 is similar to the
representation found in the “zone de préparation
culinaire” described by Cochard and Brugal
(2004), with the exception of abundant long bone
epiphyses at Klissoura 1.
Upper Palaeolithic animal exploitation at Klissoura Cave 1
125
Fig. 6. Plot of the foot region of hares from Klissoura Cave 1 by sequence, along with a standardized value based
on the MNI from each level. The elements are presented in order of descending structural density from the most
dense scan site of each element. Bone density values from Pavao and Stahl (1999)
Fig. 7. Plot of the cranial, neck and axial regions of hares from Klissoura Cave 1 by sequence, along with a standardized value based on the MNI from each level. The elements are presented in order of descending structural
density based on the densest scan site from each bone, with elements lacking structural density values placed near
similar or related elements. The head region is also plotted to show that it some layers it is present although other
axial elements are not. Bone density values from Pavao and Stahl (1999)
B. M. Starkovich & M. C. Stiner
126
Table 12
Proportions of antler fragments for the Upper
Paleolithic sequences at Klissoura Cave 1
Table 13
Age distribution for ungulates in the Klissoura
Cave 1 assemblages, based on epiphyseal fusion
and tooth eruption data
Antler
NISP
Total
NISP
Percent
Mesolithic (3-5a)
12
288
4.17
Epigravettian (IIa-d)
2
77
2.60
Mediterranean
backed-bladelet (III')
28
1451
1.93
UP (non-Aurignacian) (III")
127
1792
7.09
Med. backed-bladelet (III')
Aurignacian (upper) (IIIb-d)
21
1196
1.76
Fallow deer
1
Aurignacian (middle) (IIIe-g)
104
1897
5.48
Ibex
0
Aurignacian (lower) (IV)
394
4135
9.53
Total
1
Early UP (Uluzzian) (V)
14
356
3.93
UP (non-Aurignacian) (III")
Total
702
11192
6.27
Fallow deer
0
Red deer
0
Total
0
Sequence
Fetal/
Juvenile
neonate
Primeaged Old adult Total
adult
Epigravettian (IIa-d)
The relative representation of fallow deer
limb and foot bones and heads is generally in anatomical balance throughout the later part of the sequence, beginning with the middle Aurignacian
levels (IIIe-g). Only the axial elements below the
neck are consistently underrepresented, and this
bias must be the result of transport decisions. Antlers are present in appreciable frequencies, given
that only adult males develop antlers and these are
seasonal structures. Male fallow deer possess antlers from roughly July to April (Chapman and
Chapman, 1975), so some individuals in the
Klissoura 1 assemblages must have died in these
months, or antler was collected and curated over
longer periods for tool-making. Tools made of
antler were found in several layers at Klissoura 1
(Christidou, this issue), making it likely that many
of the antler fragments represent debitage from
tool manufacture. Worked bone (mostly antler) is
the most common in Aurignacian layers IIIe-g
and IV (Table 6). The highest proportions of antler of all sorts occur in the Mesolithic (3-5a),
though this is a small sample, layer III”, IIIe-g,
IV, and V (also a small sample) (Table 12). Not
surprisingly, the layers that contain the most antler fragments also contain the most worked antler
artifacts, suggesting some on-site production of
osseous tools.
Ungulate mortality patterns
The presence of antler may provide a seasonal indicator for the human occupations at Klis-
Fallow deer
0
0
1
0
1
0
0
0
1
1
0
0
1
1
0
0
2
0
1
0
1
1
0
0
1
1
1
0
2
1
0
4
3
4
4
13
Aurignacian (upper) (IIIb-d)
Fallow deer
1
2
Aurignacian (middle) (IIIe-g)
Fallow deer
2
Aurignacian (lower) (IV)
Fallow deer
0
1
2
0
3
Red deer
0
2
1
0
3
Aurochs
0
0
0
1
1
Ibex
1
1
1
0
3
Chamois
0
1
0
0
1
Total
1
5
4
1
11
soura Cave 1, but these data are somewhat ambiguous given the presence of osseous tool manufacture on site. More information on seasonality comes from the ungulate mortality patterns. In this
section, we consider only those Upper Paleolithic
layers with large ungulate samples.
The age structures presented in Table 13 are
based on long bone epiphyseal fusion and tooth
eruption and wear patterns. Unfortunately, fusion
schedules are not known for all of the ungulate
species in the Klissoura 1 assemblages, so in
some cases similar taxa are used as proxies. Bone
fusion data applied to ibex are taken from Noddle’s (1974) study of domestic goats, white-tailed
deer data from Purdue (1983) are used as proxies
for fallow deer, and data applied to red deer are
from Schmid (1972). Fusion data from domesticated equids (Silver, 1969) are used for wild ass.
Upper Palaeolithic animal exploitation at Klissoura Cave 1
Tooth eruption and wear schedules for whitetailed deer published by Severinghaus (1949) are
applied to fallow deer, data on red deer are from
Lowe (1967), data from Angora goats from Deniz
and Payne (1982) are used for ibex and chamois,
and developmental data from Hillson’s (2005)
discussion of domestic cattle are applied to
aurochs.
The dental ages of the ungulates were estimated from the mandibular deciduous fourth premolar (dP4) in conjunction with the mandibular
fourth premolar (P4). The use of these two teeth is
preferred in this study because the deciduous
tooth must be shed before the adult tooth comes
into wear, and the dP4 and P4 are highly diagnostic even when broken. The side of the tooth was
taken into account, and only permanent teeth that
show signs of occlusal wear were considered
among the permanent tooth specimens to avoid
double counting of individual animals. Wear
stages follow Stiner (1990), and the age cohort
data were collapsed into three broader categories:
juvenile, prime-aged adult, and old adult. The juvenile stage includes all animals that died between the time of birth and the shedding of their
dP4. The division between prime-aged adults and
old adults is set at approximately 65% of the potential life span (see Stiner 1990).
Although the available sample sizes are
small, it is significant that animals of diverse ages
were exploited. This indicates relatively non-selective hunting of ungulate age groups.
Infant and young juveniles are represented in
the death assemblages. The small size of the fetal/neonate individuals probably belonged to unborn animals, based on comparisons to specimens
in known-age collections. The presence of these
very young animals in the assemblages, except in
layers IIa-d and III”, means that the site was frequently used during the late spring (April) or
early summer. Fallow deer and ibex typically give
birth in May or June (Schaller, 1977; Spiess,
1979; Braza et al., 1988). The presence of deer
antlers may also indicate that the site was also
used in from the late summer through spring during at least some of the periods, but this is less
certain. It is possible that hunters left the area in
the height of summer.
127
CONCLUDING DISCUSSION
The richness of the archaeological sequence
and high quality of skeletal preservation in the
Upper Paleolithic through Mesolithic deposits in
Klissoura Cave 1 provides a number of important
insights into the nature of subsistence change in
southern Greece from the Upper Paleolithic
through Mesolithic periods. Trends in prey selection may indicate increasing pressure on resources, along with climate-driven changes in local biotic diversity and herd structure. Based on
the presence of abundant lithic and faunal assemblages and dozens of hearth features in some layers, Klissoura Cave 1 was used mainly as a residential site for the majority of the Upper Paleolithic occupations. The intensity of the occupations
may have been greatest, however, during the
Aurignacian (Koumouzelis et al., 2001; Karkanas, this issue; Stiner, this issue).
The trend in small game exploitation at Klissoura Cave 1 generally resembles those documented at other late Pleistocene Mediterranean sites,
namely a decrease in the proportion of small, slowmoving small game species and an increasing reliance on very productive quick types such as hares
and partridges (Stiner et al., 2000; Stiner, 2001;
Munro, 2004). Quick prey animals in the Klissoura 1 sequence were primarily hares and ground
birds, which became increasingly important with
time. In the most recent layers (3-5a and IIa-d),
the NISP counts for quick small game (mainly
hares) actually surpass those for medium sized
ungulates, though not in terms of total biomass. It
is interesting that partridges and bustards are most
common in the later layers, as these birds prefer
open grasslands. Changes in their importance relative to hares may reflect vegetation changes in
the study area, possibly at the temporary expense
of hares.
Commensurate with greater use of quick
small prey with time, there is good evidence from
Klissoura Cave 1 of land snail exploitation during
the later Paleolithic and Mesolithic, except in the
Epigravettian (layer IIa-d). The history of land
snail exploitation at Klissoura 1 parallels those for
the Mediterranean rim overall (see Lubell, 2004),
although the antiquity of the practice merits continued investigation. At Klissoura 1, snail exploitation may have begun as early as the late Aurig-
128
B. M. Starkovich & M. C. Stiner
nacian or, alternatively, by the makers of the Mediterranean blade industries.
Another trend concerns species diversity
within the ungulate remains. The dominance of
fallow deer is extreme in the middle part of the UP
at Klissoura 1. Ungulate species evenness is
somewhat higher at the beginning of the UP
(layer IV) and again at the end of the sequence in
the Epigravettian (IIa-d) and Mesolithic (3-5a).
This is true despite significant differences in the
sizes of the faunal assemblages. Though fallow
deer were always the dominant species in the ungulate assemblages, there is no indication of specialized or highly seasonal hunting. Rather, the
data indicate opportunistic responses to changes
in the species available in the region. The ungulate assemblages became more biased to fallow
deer during drier periods when grassland expanded at the expense of moist forest.
There is also the question of where the Klissoura 1 inhabitants obtained the ibex and chamois, since the area does not include true alpine
habitats. Phoca-Cosmetatou (2004b) notes a longstanding misconception of the ibex as an exclusively high-altitude species; in fact they inhabit a
much wider variety of ecozones in protected areas
today, provided that the terrain is rugged. Klissoura Cave 1 is located at the interface of rocky hills
and a large plain, so species preferring craggy terrain would have found suitable habitats in the
area. The presence of ibex and chamois in the
assemblage also is noteworthy, as these are species that specialized hunters targeted in other
Mediterranean regions (Straus, 1987; Gamble,
1997; Phoca-Cosmetatou, 2004a). The possibility
of specialized hunting is difficult to interpret archaeologically, since authors vary on what aspects of human choice should indicate this behavior. In the case of ibex or chamois, is it that
hunters make special trips to rocky or mountainous areas to harvest the only common large animals that live in those habitats, or do hunters preferentially target only one ungulate species among
many potential prey that they encounter? Anthropological definitions have tended to emphasize
the former; the works cited above pertain mainly
to selective land use and special short-term sites
in areas that can only be occupied in warmer
seasons, or of seasonal occupations along known
migratory routes.
As for Klissoura Cave 1, the high diversity of
ungulate species in layer IV, and again at the end
of the sequence, suggests a unique environmental
situation, and perhaps also long stays at a residential camp. The consistent dominance of fallow
deer through the sequence indicates that the site
was centrally located within optimal fallow deer
habitat. Longer stays at a site would mean more
time in which to accumulate rare species, along
with archaeological traces of a greater variety of
human circumstances and activities. Archaeobotanical evidence (Albert, this issue; Ntinou, this
issue) indicate that habitat diversity was indeed
higher at the beginning and the end of the Upper
Paleolithic to Mesolithic sequence, allowing more
ungulate species to compete effectively with fallow deer. Fallow deer are flexible feeders (Feldhamer et al., 1988), but they require grasses and a
complement of browse such as oak. Greater moisture in the region changes the pattern of habitats
and vegetation and increases the degree of plant
heterogeneity. The rising frequencies of openland birds in the middle of the Klissoura 1 sequence, by contrast, suggest that expansion of
drier, open habitats, when fallow deer overwhelmingly dominate the assemblages.
Other findings on Upper Paleolithic subsistence at Klissoura Cave 1 relate to human butchery practices, hare and ungulate body part representation, and ungulate mortality patterns. The
patterns of fracturing and burning damage on the
small animal bones indicate that these species
were introduced into the site by humans in nearly
all cases (but see Bocheñski and Tomek, this issue, regarding certain bird remains). Though fine
cut marks are often obscured by concretion coatings, the pervasive distribution of cone and impact fractures on the ungulate bones indicates that
the carcasses were intensively processed for marrow throughout the UP and later periods. Minor
biases in ungulate body part representation are not
explained by in situ attrition and therefore must
reflect human transport decisions. The anatomical
regions of the body are fairly evenly represented,
except for the axial parts below the neck, which
may have been discarded at kill sites. Biases also
exist in the body part profiles of hares throughout
the Klissoura 1 sequence, with a consistent and
notable absence of foot and vertebral elements.
The biases do not relate to variations in skeletal
Upper Palaeolithic animal exploitation at Klissoura Cave 1
density and therefore are attributed to humans.
The absence of foot and vertebral bones may relate field dressing or skinning, and removal of
these elements before the hares were brought to
the site, or it may have arisen because hares were
processed and cooked in different parts of the site
(Cochard and Brugal, 2004). The areal extent of
the Klissoura Cave 1 excavations is fairly limited,
partly because the shelter is small and partly
because the area immediately surrounding the
cave is under cultivation.
Ungulate mortality patterns are rather nonselective at Klissoura Cave 1, with animals of all
ages well represented. A few fetal or neonate remains were found in each assemblage, and it is
significant that they occur throughout the sequence. These bones most likely belonged to unborn animals, indicating that a few pregnant females were hunted before or during the spring birthing season. The presence of antler in the assemblages may indicate that the site was also occupied between the late summer and early spring,
though cross-season curation of antler for toolmaking cannot be refuted with available information.
Overall, many of the results of the Upper
Paleolithic through Mesolithic faunas from
Klissoura Cave 1 are similar to trends found elsewhere in the Mediterranean, including changes in
small game exploitation, body part transport of
large animals, and mortality patterns. Other features of the Klissoura Cave 1 faunal assemblage
are unique and provide new information on Upper
Pleistocene subsistence in southern Greece and its
paleoenvironmental contexts. Interestingly, no
trends were found among layers in the pattern of
body part transport to the site, or subsequent
butchery and marrow processing. The stability of
these patterns may indicate a consistent focus on
large game hunting during the UP occupations,
and perhaps also a general consistency in overall
site function. Such consistency could be explained by the strategic position of the site on the
Peloponnesian landscape.
Acknowledgements
We are grateful to Margarita Koumouzelis and
Janusz K. Koz³owski for inviting us to study the Klissoura Cave 1 fauna. Thanks to Teresa Tomek and Zbigniew Bocheñski for valuable comments on this manu-
129
script, and to Panagiotis (Takis) Karkanas, Krzysztof
Sobczyk, Sherry Fox and Mathew Devitt for their logistical support. The preliminary study of the faunas by
Poitr Wojtal and colleagues was tremendously helpful
for designing this larger study. Starkovich’s part of the
research was supported by a National Science Foundation IGERT program fellowship, a Rieker Grant, and
William Shirley Fulton Scholarship, all through the
University of Arizona, a research associate award from
the Wiener Laboratory at the American School of Classical Studies at Athens, and a dissertation improvement
grant to Britt Starkovich and Mary Stiner (advisor)
from the National Science Foundation (BCS-0827294).
Stiner’s work was supported by a grant from the National Science Foundation (BCS-0410654). Many
thanks also to the Institute for Aegean Prehistory for
supporting the excavations at Klissoura Cave 1, without which none of this research would have been possible.
REFERENCES
BERKE H. 1984. The distributions of bones from large
mammals at Petersfels. In: H. Berke, J. Hahn, C. J.
Kind (eds) Verlag Archaeologica Venatoria. Institut
fur Urgeschichte der Universitat Tubingen, Tubingen, Germany, 103–108.
BINFORD L.R. 1978. Nunamiut Ethnoarchaeology.
Academic Press, New York.
BINFORD L.R., BERTRAM J. 1977. Bone frequencies and attritional processes. In: L.R. Binford (ed.)
For Theory Building in Archaeology. Academic
Press, New York, 77–156.
BLASCO M., CRESPILLO E., SANCHEZ J.M.
1986–87. The growth dynamics of Testudo graeca
L. (Reptilia: Testudinidae) and other data on its populations in the Iberian Peninsula. Israel Journal of
Zoology 34, 139–147.
BRAIN C.K. 1981. The Hunters or the Hunted? University of Chicago Press, Chicago.
BRAZA R., SAN JOZE C., BLOM A. 1988. Birth
measurements, paturition dates and progeny sex ratio of Dama dama in Donana, Spain. Journal of
Mammalogy 69, 607–610.
BUNN H.T., BARTRAM L.E., KROLL E.M. 1988.
Variability in Bone Assemblage Formation from
Hadza Hunting, Scavenging, and Carcass Processing. Journal of Anthropological Archaeology 7,
412–457.
CHAPMAN D.I., CHAPMAN N.G. 1975. Fallow
Deer, their history, distribution and biology. Terence Dalton Ltd., Lavenham, Suffolk.
COCHARD D., BRUGAL J.P. 2004. Importance des
fonctions de sites dans les accumulations paléolithiques de léporidés. In: J.P. Brugal, J. Desse (eds)
130
B. M. Starkovich & M. C. Stiner
Petits Animaux et Sociétés Humaines, du Complément Alimentaire aux Ressources Utilitaires.
APDCA, Antibes, 283–296.
COLSON E. 1979. The Harvey Lecture Series. In
Good Years and in Bad: Food Strategies of Self-Reliant Societies. Journal of Anthropological Research 35, 18–29.
COSTAMAGNO S. 2004. Si les Magdaléniens du sud
de la France n’étient pas des chasseurs spécialisés,
qu’étaient-ils? In: P. Bodu, C. Constantin (eds)
Approches Fonctionnelles en Préhistoire. Société
Préhistorique Français, Paris, 361–369.
COSTAMAGNO S., LILIANE M., CEDRIC B., BERNARD V., BRUNO M. 2006. Les Pradelles (Marillac-le-France, France): a Mousterian Reindeer
Hunting Camp? Journal of Anthropological Archaeology 25, 466–484.
COSTAMAGNO S., COCHARD D., FERRIE J.G.,
LAROULANDIE V., CAZALS N., LANGLAIS
M., VALDEYRON N., DACHARY M., BARBAZA M., GALOP D., MARTIN H., PHILIBERT
S. 2008. Nouveaux milieux, nouveaux gibiers, nouveaux chasseurs? Evolution des pratiques cynegetiques dans les Pyrenees du Tardiglaciaire au debut
du Postglaciaire. Bulletin de la Societe Prehistorique Francaise 105, 17–27.
CURREY J. 1984. The Mechanical Adaptations of
Bones. Princeton University Press, Princeton, New
Jersey.
DAVIS S.J. 1987. The Archaeology of Animals. Yale
University Press, London.
DENIZ E., PAYNE S. 1982. Eruption and Wear in the
Mandibular Dentition as a Guide to Aging Turkish
Angora Goats. In: B. Wilson, C. Grigson, S. Payne
(eds) Ageing and Sexing Animal Bones from Archaeological Sites. British Archaeological Reports, Oxford, 155–205.
FELDHAMER G.A., FARRIS-RENNER K.C., BARKER C.M. 1988. Dama dama. Mammalian Species
317, 1–8.
FISHER J.W. 1995. Bone surface modifications in zooarchaeology. Journal of Archaeological Method
and Theory 2, 7–68.
GAMBLE C. 1997. The Animal Bones from Klithi. In:
G.N. Bailey (ed.) Klithi: Paleolithic Settlement and
Quaternary Landscapes in Northwest Greece. Volume 1: Excavation and intra-site analysis at Klithi.
McDonald Institute for Archaeological Research,
Cambridge, 207–244.
GAMBLE C. 1999. Faunal Exploitation at Klithi. In:
G.N. Bailey, E. Adam, E. Panagopoulou, C. Perles,
K. Zachos (eds) The Palaeolithic Archaeology of
Greece and Adjacent Areas: Proceedings of the
ICOPAG Conference, Ioannina. Technical Print
Services Ltd, Nottingham, 179–187.
GAUDZINSKI S. 1995. Wallertheim Revisited: a Reanalysis of the Fauna from the Middle Palaeolithic
site of Wallertheim (Rheinhessen/Germany). Journal of Archaeological Science 22, 51–66.
GRAYSON D.K. 1984. Quantitative Zooarchaeology.
Academic Press, Orlando.
HAILEY A., WRIGHT J., STEER E. 1988. Population
Ecology and Conservation of Tortoises: the Effects
of Disturbance. Herpetological Journal 1, 294–301.
HANDRINOS G., AKRIOTIS T. 1997. The Birds of
Greece. Christopher Helm Ltd., A & C Black Ltd.,
London.
HILLSON S. 2005. Teeth. Cambridge University
Press, Cambridge.
JOCHIM M. 1998. A Hunter-Gatherer Landscape:
Southwest Germany in the Late Paleolithic and
Mesolithic. Plenum Press, New York.
KELLY R.L. 1995. The Foraging Spectrum: Diversity
in Hunter-Gatherer Lifeways. Smithsonian Institution Press, Washington.
KOUMOUZELIS M., GINTER B., KOZ£OWSKI J.
K., PAWLIKOWSKI M., BAR-YOSEF O., ALBERT R. M., LITYÑSKA-ZAJ¥C M., STWORZEWICZ E., WOJTAL P., LIPECKI G., TOMEK T.,
BOCHEÑSKI Z. M., PAZDUR A. 2001. The Early
Upper Palaeolithic in Greece: The Excavations in
Klissoura Cave. Journal of Archaeological Science
28, 515–539.
LAM Y.M., CHEN X., MAREAN C.W., FREY C.J.
1998. Bone Density and Long Bone Representation
in Archaeological Faunas: Comparing Results from
CT and Photon Densitometry. Journal of Archaeological Science 25, 559–570.
LAMBERT M.R.K. 1982. Studies on the growth, structure, and abundance of the Mediterranean spurthighed tortoise, Testudo graeca, in field populations. Journal of Zoology, London 196, 165–189.
LEVINS R. 1968. Evolution in Changing Environments: Some Theoretical Explorations. Princeton
University Press, Princeton, NJ.
LOWE V.P. 1967. Teeth as indicators of age, with special reference to red deer (Cervus elaphus) of known
age from Rhum. Journal of Zoology, London 152,
137–153.
LUBELL D. 2004. Prehistoric edible land snails in the
circum-Mediterranean: the archaeological evidence.
In: J.P. Brugal, J. Desse (eds) Petits Animaux et
Sociétiés Humaines, du Complément Alimentaire
aux Resources Utilitaires. APDCA, Antibes, 77–98.
LYMAN R.L. 1982. The Taphonomy of Vertebrate
Archaeofaunas: Bone Density and Differential
Survivorship of Fossil Classes. Unpublished Ph.D.
Dissertation. University of Washington, Seattle.
LYMAN R.L. 1984. Bone Density and Differential
Survivorship of Fossil Classes. Journal of Anthro-
Upper Palaeolithic animal exploitation at Klissoura Cave 1
pological Archaeology 3, 259–299.
LYMAN R.L. 1994. Vertebrate Taphonomy. Cambridge University Press, Cambridge.
MANNE T., STINER M.C., BICHO N.F. 2005. Evidence for Bone Grease Rendering during the Upper
Paleolithic at Vale Boi (Algarve, Portugal). In: N.F.
Bicho (ed.) Proceedings of the IV Congresso de
Arqueologia Peninsular, Session 4. Centro de Estudos, Faro, 1–15.
MELLARS P. 1973. The Character of the Middle-Upper Paleolithic Transition in Southwest France. In:
A.C. Renfrew (ed.) The Explanation of Culture
Change. Duckworth, London, 255–276.
MELLARS P. 1989. Major Issues in the Emergence of
Modern Humans. Current Anthropology 30, 349–
385.
MORALES M.B., ALONSO J.C., ALONSO J. 2002.
Annual productivity and individual female reproductive success in a Great Bustard Otis tarda population. Ibis 144, 293–300.
MUNRO N. 2004. Zooarchaeological Measures of
Hunting Pressure and Occupation Intensity in the
Natufian. Current Anthropology 45, S5–S33.
NODDLE B.A. 1974. Ages of Epiphyseal Closure in
Feral and Domestic Goats and Ages of Dental Eruption. Journal of Archaeological Science 1, 195–204.
NOWACK R.L. 1991. Walker’s Mammals of the
World. Johns Hopkins University Press, Baltimore,
Maryland.
O’CONNELL J.F., HAWKES K., BLURTON JONES
N. 1988. Hadza Hunting, Butchering, and Bone
Transport and Their Archaeological Implications.
Journal of Anthropological Research 44, 113–161.
PAVAO B., STAHL P. 1999. Structural Density Assays of Leporid Skeletal Elements with Implications
for Taphonomic, Actualistic and Archaeological
Research. Journal of Archaeological Science 26,
53–66.
PHOCA-COSMETATOU N. 2003a. Ibex Exploitation:
the Case of Klithi or the Case of the Upper Palaeolithic? In: E. Kotjabopoulou, Y. Hamilakis, P. Halstead, C. Gamble, P. Elefanti (eds) Zooarchaeology
in Greece: Recent Advances. British School at Athens, 161–173.
PHOCA-COSMETATOU N. 2003b. Subsistence
Changes During the Late Glacial? The Example of
Ibex Exploitation in Southern Europe. In: M. PatouMathis, H. Bocherens (eds) Le rôle de l’environment dans les comportements des chasseurs-cueilleurs préhistoriques. BAR International Series
1105, Oxford, 39–54.
PHOCA-COSMETATOU N. 2004a. Site Function and
the ‘Ibex-Site Phenomenon’: Myth or Reality? Oxford Journal of Archaeology 23, 217–242.
PHOCA-COSMETATOU N. 2004b. A Zooarchaeolo-
131
gical Reassessment of the Habitat and Ecology of
the Ibex (Capra ibex). In: R. C. G. M. Lauwerier, I.
Plug (eds) The Future from the Past: Archaeozoology in Wildlife Conservation and Heritage
Management. Oxbow Books, 64–78.
PHOCA-COSMETATOU N. 2005. Landscape Use in
Northeast Italy During the Upper Palaeolithic. Prehistoria Alpina 41, 23–49.
PURDUE J.R. 1983. Epiphyseal Closure in WhiteTailed Deer. Journal of Wildlife Management 47,
1207–1213.
REYNAUD SAVIOZ N., MOREL P. 2005. La faune
de Nadaouiyeh AÎn Askar (Syrie centrale, PléistocÀne moyen): aperçu et perspectives. Revue de
Paléobiologie, GenÀve 10, 31–35.
SCHALLER G.B. 1977. Mountain Monarchs, Wild
Sheep and Goats of the Himalaya. University of
Chicago Press, Chicago.
SCHMID E. 1972. Atlas of Animal Bones for Prehistorians, Archaeologists, and Quaternary Geologists. Elsevier Science Publishers, Amsterdam.
SCHMITT D.N., LUPO K. 1995. On Mammalian
Taphonomy, Taxonomic Diversity, and Measuring
Subsistence Data in Zooarchaeology. American Antiquity 60, 496–514.
SEVERINGHAUS C.W. 1949. Tooth Development
and Wear as Criteria of Age in White-Tailed Deer.
Journal of Wildlife Management 13, 195–216.
SILVA M., DOWNING J.A. 1995. CRC Handbook of
Mammalian Body Masses. CRC Press, Boca Raton,
Florida.
SILVER I.A. 1969. The ageing of domestic animals.
In: D. Brothwell, E. S. Higgs (eds) Science in archaeology: a survey of progress and research.
Praeger, New York, 283–302.
SIMPSON E.H. 1949. Measurement of Diversity. Nature 163, 688.
SPETH J.D., TCHERNOV E. 2002. Middle Paleolithic
tortoise use at Kebara Cave (Israel). Journal of Archaeological Science 29, 471–483.
SPIESS A.E. 1979. Reindeer and Caribou Hunters, an
Archaeological Study. Academic Press, New York.
STEPHENS D.W., KREBS J.R. 1986. Foraging Theory. Princeton University Press, Princeton.
STINER M.C. 1990. The use of mortality patterns in
archaeological studies of hominid predatory adaptations. Journal of Anthropological Archaeology 9,
305–351.
STINER M.C. 1991. Food Procurement and Transport
by Human and Non-human Predators. Journal of
Archaeological Science 18, 455–482.
STINER M.C. 1994. Honor Among Thieves: a Zooarchaeological Study of Neandertal Ecology. Princeton University Press, Princeton, NJ.
STINER M.C. 2001. Thirty Years on the “Broad Spec-
132
B. M. Starkovich & M. C. Stiner
trum Revolution” and Paleolithic Demography.
Proceedings of the National Academy of Sciences
98, 6993–6996.
STINER M.C. 2002a. Carnivory, Coevolution, and the
Geographic Spread of the Genus Homo. Journal of
Archaeological Research 10, 1–63.
STINER M.C. 2002b. On In Situ Attrition and Vertebrate Body Part Profiles. Journal of Archaeological
Science 32, 103–117.
STINER M.C. 2004. A Comparison of Photon Densitometry and Computed Tomography Parameters of
Bone Density in Ungulate Body Part Profiles. Journal of Taphonomy 2, 117–145.
STINER M.C. 2005. The Faunas of Hayonim Cave, Israel: A 200,000 Year Record of Paleolithic Diet,
Demography, and Society. Peabody Museum of Archaeology and Ethnology, Harvard University,
Cambridge.
STINER M.C., MUNRO N. 2002. Approaches to Prehistoric Diet Breadth, Demography, and Prey Ranking Systems in Time and Space. Journal of Archaeological Method and Theory 9, 181–214.
STINER M.C., WEINER S., BAR-YOSEF O., KUHN
S.L. 1995. Differential burning, fragmentation and
preservation of archaeological bone. Journal of Archaeological Science 22, 223–237.
STINER M.C., MUNRO N.D., SUROVELL T.A.
2000. The Tortoise and the Hare: Small-Game Use,
the Broad-Spectrum Revolution, and Paleolithic
Demography. Current Anthropology 41, 39–73.
STRAUS L.G. 1987. Upper Palaeolithic Ibex Hunting
in Southwest Europe. Journal of Archaeological
Science 14, 163–178.
SURMELY F., ALIX P., COSTAMAGNO S., DANIEL P., HAYS M., MURAT R., RENARD R.,
VIRMONT J., TEXIER J.P. 2003. Découverte d’un
gisement du Gravettien ancien au Iieu-dit Ie Sire
(Mirefleurs, Puy-de-Dôme). Bulletin de la Societe
Prehistorique Francaise 100, 29–39.
TOMEK T., BOCHEÑSKI Z.M. 2002. Bird Scraps
from a Greek Table: The Case of Klissoura Cave.
Acta Zoologica Cracoviensia 45, 133–138.
VAVALEKAS K., THOMAIDES C., PAPAEVANGELLOU E., PAPAGEORGIOU N. 1993. Nesting
biology of the Rock Partridge Alectoris graeca
graeca in northern Greece. Acta Ornithologica 28,
97–101.
YELLEN J.E. 1991. Small mammals: Kung San utilization and the production of faunal assemblages.
Journal of Anthropological Archaeology 10, 1–26.