ARTICLES
https://doi.org/10.1038/s41562-021-01083-y
Evidence for early dispersal of domestic sheep
into Central Asia
William T. T. Taylor 1,2 ✉, Mélanie Pruvost 3, Cosimo Posth 4,5, William Rendu 3,6,
Maciej T. Krajcarz7, Aida Abdykanova 8, Greta Brancaleoni 7, Robert Spengler 2, Taylor Hermes 4,
Stéphanie Schiavinato9, Gregory Hodgins 10, Raphaela Stahl4, Jina Min11, Saltanat Alisher kyzy 12,13,
Stanisław Fedorowicz14, Ludovic Orlando8, Katerina Douka 2, Andrey Krivoshapkin12,13,
Choongwon Jeong 11, Christina Warinner 4,15 and Svetlana Shnaider 6,12 ✉
The development and dispersal of agropastoralism transformed the cultural and ecological landscapes of the Old World,
but little is known about when or how this process first impacted Central Asia. Here, we present archaeological and biomolecular evidence from Obishir V in southern Kyrgyzstan, establishing the presence of domesticated sheep by ca. 6,000 BCE.
Zooarchaeological and collagen peptide mass fingerprinting show exploitation of Ovis and Capra, while cementum analysis of
intact teeth implicates possible pastoral slaughter during the fall season. Most significantly, ancient DNA reveals these directly
dated specimens as the domestic O. aries, within the genetic diversity of domesticated sheep lineages. Together, these results
provide the earliest evidence for the use of livestock in the mountains of the Ferghana Valley, predating previous evidence by
3,000 years and suggesting that domestic animal economies reached the mountains of interior Central Asia far earlier than
previously recognized.
T
he early Holocene domestication of crops and livestock in the
Fertile Crescent is among the earliest in the world, with the
first traits of domestication appearing in plants and animals
by 7,500 BCE1–6. The food-producing economy of this region, based
on sheep, goats, cows, cereals and legumes, launched humanity’s
first agricultural demographic transition6,7, which would eventually
reshape human populations, both genetically and culturally, across
the ancient world8. Recent human genomic studies have clearly
illustrated that the Neolithization of western Eurasia involved a
demic wave of expansion as peoples of southwest Asian ancestry
expanded and admixed with local populations in Europe and West
Asia9–11. Understanding the dynamics of early Neolithic dispersals
informs the processes that shaped the trajectory of human societies
in Eurasia.
Archaeozoological evidence suggests that the first progenitors
of domesticated goats and sheep came under sustained, multigenerational human control by ca. 11,000–9,000 years ago in a region
stretching from eastern Anatolia to the Zagros Mountains of Iran
and Iraq12–14. Following their initial domestication, animal livestock
(including sheep and goats, as well as cattle) and plant crops from
this region dispersed across Eurasia and Africa in one of the most
important globalization processes in human prehistory15. While
these three livestock species sometimes moved together into new
regions as part of a ‘Neolithic package’, domesticated sheep and
goats became particularly widespread in the ancient world, reaching areas across Europe and the Mediterranean by ca. 6,000 BCE
and North Africa by 5,000 BCE15.
Although the impact of Western Asian livestock dispersals on
the ancient cultural landscape of western Eurasia and Africa has
been substantial, the spread of animal domesticates to the Eurasian
interior is poorly understood16,17. Domesticated sheep and goat are
not found in the archaeological record of the eastern steppe region
of Inner Asia until the arrival of the pastoralist Afanasievo culture ca. 3,000 BCE18,19, and agropastoralism has only been traced
to the early third millennium BCE in Central Asia20–22. However,
Soviet archaeologists have long hypothesized a much earlier arrival
of agropastoralism from southwest Asia between the seventh
and fourth millennia BCE associated with the Kelteminar culture (Fig. 1)23–27. This poorly understood culture, concentrated in
the Khoresm region beyond the southern edge of the Aral Sea, is
characterized by their microlith technology, distinctive arrowheads
with a prominent notch cut out at their base, handmade coarseware
pottery with pointed bases and geometric incisions and a mixed
pastoralist–hunting–fishing economy.26,28–31 Cattle and ovicaprid
(sheep/goat) remains dating to at least the fifth millennium BCE
have been reported from Kelteminar sites, but they have not been
subjected to detailed zooarchaeological analysis27,32, and thus it is
not clear whether they represent wild or domesticated populations.
Museum of Natural History, University of Colorado-Boulder, Boulder, CO, USA. 2Department of Archaeology, Max Planck Institute for the Science of
Human History, Jena, Germany. 3De la Préhistoire à l’Actuel: Culture, Environnement et Anthropologie (PACEA), Université de Bordeaux, Pessac, France.
4
Department of Archaeogenetics, Max Planck Institute for the Science of Human History, Jena, Germany. 5Institute for Archaeological Sciences, University
of Tübingen, Tübingen, Germany. 6ArchaeoZOOlogy in Siberia and Central Asia – ZooSCAn, CNRS – IAET SB RAS International Research Laboratory, IRL
2013, Institute of Archaeology SB RAS, Novosibirsk, Russia. 7Institute of Geological Sciences, Polish Academy of Sciences, Warszawa, Poland. 8American
University of Central Asia, Bishkek, Kyrgyzstan. 9Faculté de Médecine Purpan, Université Paul Sabatier, Toulouse, France. 10Accelerator Mass Spectrometry
Laboratory, University of Arizona, Tucson, AZ, USA. 11School of Biological Sciences, Seoul National University, Seoul, Republic of Korea. 12Institute of
Archaeology and Ethnography SB RAS, Novosibirsk, Russia. 13Novosibirsk State University, Novosibirsk, Russia. 14Department of Geomorphology and
Quaternary Geology, University of Gdańsk, Gdańsk, Poland. 15Department of Anthropology, Harvard University, Cambridge, MA, USA.
✉e-mail: william.taylor@colorado.edu; sveta.shnayder@gmail.com
1
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
Kazakhstan
Kyrgyzstan
Uzbekistan
Obishir V
Tajikistan
China
Afghanistan
Pakistan
Kelteminar
culture
C
S E
A
A
B L A C K
S
P
I A
N
A
S E
Ca. 8,500 BCE
KOPE
T D
AG
M T
NS
Ca. 6,500 BCE
M
E
D
I T
E R
R A
N E A
N
Obishir
Djeitun
Hissar
culture
SE A
Before ca. 7,000 BCE
Fig. 1 | Modeled dispersal of domestic animals into Central Asia. Proposed centre of sheep and goat domestication (dark red) and early Holocene
dispersals out of western Eurasia (information from Vigne16 and Murphy90) alongside relevant archaeological cultures and newly hypothesized dispersal
event(s) (red dashed arrow). Inset map of Obishir V and the Ferghana Valley situated within the mountainous zone of Central Asia.
More recently, extensive research at the Djeitun site in southern
Turkmenistan (Fig. 1) has revealed a complex agropastoral system,
relying on harvesting tools, grinding stones, irrigation and a mixed
crop assemblage, including glume wheats and barley, dated to ca.
6500 BCE. Based on the predominance of Ovis and Capra remains
at Djeitun, along with their small size and some distinct morphological traits, domesticated sheep and goat are also thought to have
been an important part of the subsistence economy at Djeitun33–36.
Soviet scholars have also hypothesized the presence of another
Neolithic culture known as the Hissar in the Hissaro-Alay
Mountains of Tajikistan (Fig. 1). In its classic formulation, the
Hissar culture is dated to between the sixth and second millennia
BCE and is characterized by a pebble and microblade lithic technology, which includes geometric microliths37. Importantly, scholars
have also argued that the Hissar economy included goat and sheep
pastoralism37. However, as at Djeitun, the inference of domestic animals at Hissar sites has been based largely on the high frequency of
Ovis and Capra remains and indirect lines of evidence such as site
location, rather than direct evidence of human management.
Taken together, these archaeological finds provide evidence for
the presence of Neolithic material culture and domestic plants, along
with tentative evidence for domestic animals, along the margins of
Central Asia during the seventh millennium BCE. However, most of
the relevant scholarly research on these cultures was performed prior
to the widespread availability of scientific methods, including radiocarbon dating via accelerator mass spectrometry. Moreover, no direct
archaeobotanical or zooarchaeological evidence for food production
has been found east of the Kopet Dag piedmont of Turkmenistan
prior to ca. 3,500 BCE22. With the advent of powerful new methods
in archaeological science over the past decade, it is now possible to
revisit these Soviet-era cultural designations and Neolithic dispersal
models and to test these hypotheses in a robust scientific framework.
Here we investigate animal husbandry at the early Holocene
site of Obishir V in southern Kyrgyzstan. Located in the heart of
Central Asia within the Inner Asian Mountain Corridor, Obishir
V is situated along the southern margin of the Ferghana Valley, a
historically significant crossroads for the exchange of people and
animals across eastern and western Eurasia21. Through new excavations, we identified stone tools and faunal remains suggestive of
grain processing and livestock pastoralism, dating to the late seventh millennium BCE. Zooarchaeological analysis and collagen fingerprinting reveal a faunal assemblage consisting primarily of Ovis
and Capra, which based on cementum annulation appear to have
been killed during the autumn season. Based on DNA sequences
from five well-preserved specimens, we further identify at least four
of these animals as O. aries within the diversity of domestic sheep,
and showing genetic affinity to modern Anatolian and South Asian
breeds. The presence of these domestic sheep at Obishir V ca. 6,000
BCE suggests that pastoral livestock species reached the foothills of
the Pamiro-Alay mountains millennia earlier than previously recognized. These findings warrant a re-evaluation of the timing and
routes of the earliest eastward agropastoral expansions and the role
of domestic plants and animals in the development of social complexity in this crossroads region.
Results
Site excavation. The site of Obishir V (39° 57′ N, 71° 16′ E) is located
near the Town of Aidarkyen, in Batken Prefecture of Osh Province,
along the northern edge of the Pamiro-Alay range in the southern
Ferghana Basin in Kyrgyzstan (Fig. 1). The area has an arid, montane
climate with cold winters and hot summers, situated at an elevation
of roughly 1,700 m above sea level. The site itself consists of a small
rockshelter, originally identified by U.I. Islamov in 1965 and excavated throughout the 1960s and 1970s38. These initial excavations
explored an area of 141 m2. Renewed excavations began at the site
in 2015, revealing a stratigraphic sequence more than 4 m in depth,
with five primary layers (Fig. 2). The sedimentary sequence was
formed largely through colluvial rockfall and scree accumulation
from Palaeozoic limestone and shale, which form the steep hillside.
The lowermost layer, layer 5, consists of an aeolian loess-like deposit,
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
S
N
X (m)
8
9
10
11
13
12
14
15
–2
0
0
2.1
1
2.1
2.1
2.2
1
–3
2.3
2.3
4
2.4
3
Z (m)
–4
4
5
–5
5
–6
Legend:
Plan of 2015–2019 excavation area
N
Debris (limestone and shale angular clasts)
X
Riverine rounded pebbles
Western profile
Bedrock (carboniferous shale)
Layer numbers
Y
1
Layer boundaries
Intra-layer sedimentary structures
1m
Fig. 2 | Stratigraphic profile of Obishir v. The key layers are layer 0 (modern topsoil), layer 1 (Bronze through Middle Ages), layer 2 (early Holocene strata,
including occupation surfaces and palaeosol), layers 3 and 4 (underlying colluvium containing contemporaneous and earlier cultural material) and layer 5
(a sterile final late Pleistocene/initial early Holocene loess deposit).
Table 1 | Bayesian radiocarbon model (uniform prior) for
Obishir v, by stratigraphic series, produced in OxCal using the
iNTCAL20 calibration curve65
Boundary
Modelled date (1σ)
Modelled date (2σ)
Layer 1 end
447–1372 cal CE
424 cal CE–present
Layer 1 start
1,783–939 cal BCE
2,499–902 cal BCE
Layer 2–3 end
2,897–2,342 cal BCE
2,914–1,624 cal BCE
Layer 2–3 start
7,349–6,876 cal BCE
7,781–6,817 cal BCE
Layer 4 end
8,129–7,555 cal BCE
8,160–7,161 cal BCE
Layer 4 start
8,608–7,875 cal BCE
9,827–7,819 cal BCE
Layer 5 end
10,124–8,276 cal BCE
11,581–7,993 cal BCE
Layer 5 start
11,958–8,905 cal BCE
14,200–8,300 cal BCE
For radiocarbon dates and model distributions, see Supplementary Information.
which passes upward into layer 4 with a weakly developed palaeosol.
Layer 5 was dated to approximately 10,000 BCE using thermoluminescence (Table 1 and Supplementary Information). Layers 2, 3 and
4 contain cultural deposits formed through complex sedimentary
processes, involving some downslope relocation. One radiocarbon
date on unidentified animal bone suggests possible cultural activity
in layer 4 beginning as early as ca. 8,198–7,820 cal BCE. Layer 2 is a
dark organic palaeosol and occupation surface with cultural activity
clustering around ca. 6,000 cal BCE, underlain by several metres of
colluvium that contain roughly contemporaneous cultural material,
including from layer 3 (Supplementary Information). Dates from
animal bones and charcoal recovered from layer 1 suggest that this
layer formed much later and dates from the late Bronze Age through
the Early Middle Ages (Supplementary Information).
Comparing the lithic assemblage of Obishir V with earlier materials from Central Asia reveals an important economic transition
characterized by a decline in the use of projectiles and a concomitant increased use of microblades for carcass processing, as well
as an increased emphasis on grinding stones for food production.
This lithic transition is reflected across much of mid-Holocene
Central Asia39. Stone tools from layers 2–4 reveal a lithic industry
based on pressure knapping of microblades, with important similarities to the Hissar culture, including a technological emphasis on
retouched bladelets, trapezoids, end-scrapers and grinding stones,
and evidence of processing of wild or domestic grain33. In addition,
as seen in the Kelteminar culture23, the Obishirian industry also utilized pressure knapping from prismatic and cis-prismatic cores, but
only yielded a handful of bullet cores and lacked other diagnostic
features of the Kelteminar stone tool assemblage, such as scalene
triangles, sickle blades and ‘Kelteminar points’. Finally, excavations
at Obishir V recovered ornamental stone pendants and grinding
stones, perhaps associated with food production, that have close
parallels in both Hissar and Kelteminar cultures31,40.
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
a
b
Cervidae
287
1
Non-mammalian
4
250
Cervidae
Bovidae
4
5
Deer
Gazella sp.
Saiga sp.
Ovinae
Ovis sp.
Saiga sp.
1
200
8
Bovidae
Count
Deer
Gazella sp.
Saiga sp.
Ovis sp.
150
35
100
Insufficient
collagen
for
identification
88
7
Ovis sp.
Capra sp.
Pantholops sp.
6
50
Capra sp.
22
1
0
1
1
d.
ul
U
ni
Canis sp.
Vulpes sp.
1
Lagomorph
(unknown sp.)
1
Rodentia
(unknown sp.)
U
ng
ul
at
at
e
e
(M
(L
)
)
pr
a
U
ng
O
vi
s/
ca
ra
C
ap
O
vi
s
sp
.
sp
.
1
Taxonomic category
Fig. 3 | identification of animal remains at Obishir v by ZooMS. a, Morphology-based identifications of animal remains from layers 2–4 at Obishir V, by
number of identified specimens (NISP) and taxonomic category. b, Collagen peptide-based identifications for layers 2–4, square R8 at Obishir V, grouped
by general phylogenetic relationships. Each colored circle represents a different identified taxon. Specimens with a grey circle are missing necessary peptide
markers for more detailed identification, and silhouettes depict a range of possible taxa from which specimens in that category may derive. Icons created or
modified by Hans Sell. Unid., unidentified.
In addition to lithics, Obishir V also yielded a small archaeofaunal assemblage, which has allowed an investigation of subsistence
changes underlying the technological shift towards microblades.
Although the faunal material (Supplementary Data A) is highly
fragmented, layers 2–4 contained a total of 400 animal bone fragments. Of these, 24 could be morphologically identified as Ovis
(n = 1), Capra (n = 1) or Ovis/Capra (n = 22). An additional 89
were identified as ungulate (88 medium-sized ungulates and 1 large
ungulate), and 287 fragments were not identifiable to a taxonomic
class (Fig. 3a). While the assemblage displayed many specimens
with burn marks (n = 36), some with cut marks (n = 5) and others
with spiral fracture (n = 3), only a small number of these anthropogenically modified specimens were taxonomically identifiable
(Ovis, n = 1; ovicaprid, n = 1). A single human tooth was identified
in the assemblage (a deciduous incisor shed from a child of approximately 6 years of age); however, this specimen was not recovered
in stratigraphic context but was found during screening of sediments at the contact between layer 1 and 2. Radiocarbon date estimates for this tooth suggest that it belongs to the late Holocene
cultural level 1 (Supplementary Information). Despite the small
sample size and heavy fragmentation of the skeletal assemblage, the
archaeofaunal material suggests that early Holocene animal economies at Obishir incorporated meaningful exploitation of sheep and
goat. We next applied biomolecular methods to test whether these
animals represent wild or domestic individuals, and to assess their
antiquity directly.
Archaeofaunal remains. Zooarchaeology by mass spectrometry
(ZooMS) analysis41 of 74 archaeofaunal skeletal fragments from
square R8, layers 2–4 (Supplementary Data B) confirms the presence of Ovis and Capra in the early Holocene layers at Obishir V,
and among identifiable specimens suggests a nearly complete economic emphasis on these two taxa. Although nearly half of the
assemblage from square R8 no longer had identifiable collagen
(n = 35), the largest number of identifiable specimens were
assigned as sheep (n = 8) or goat (n = 6), or ovicaprid lacking the
necessary peptide marker to distinguish between sheep and goat
(n = 7; Fig. 3b). We also identified one deciduous tooth belonging
to a canine (Canis, likely a domestic dog or fox), one rodent bone,
one lagomorph bone and one bone broadly assigned to deer/Saiga/
gazelle. On the basis of observed peptide markers, four specimens
appeared non-mammalian, four were assigned only to Cervidae/
Bovidae and six additional specimens were consistent with either
sheep or deer/Saiga/gazelle. Comparing the faunal assemblages
of the Neolithic layers 2–4 at Obishir V with an Early Bronze Age
assemblage from the nearby Alay Valley21 shows that an emphasis
on Ovis, augmented by Capra, is characteristic of local Neolithic/
Bronze Age pastoral subsistence strategies. Later, during the first
millennium BCE and first millennium CE, the Obishir V fauna also
includes additional domesticated taxa, such as Bos sp. and Equus sp.
(identified through DNA analysis as a male domesticated donkey,
E. asinus; Supplementary Information). Dietary focus on domestic
sheep, supplemented by goat and other taxa, characterizes modern
herding lifeways in the region42, and may reflect cultural preferences
or herd compositions optimized to the dry montane environment.
Cementum analysis and dental eruption/wear. Cementum analysis of intact tooth specimens can provide insights into the age structure and management of animal herds, including the season during
which animals are culled. Cementum is a connective calcified tissue
that is deposited on the outer surface of the dental root, linking it
with the fibres of the periodontal ligament. Consisting of a collagen
matrix within which hydroxyapatite crystals form, its growth is continuous throughout an animal’s life43–45 and follows an annual cycle
that typically produces light-coloured bands of low-mineral-density
growth during the late spring to fall followed by dark banding of
high mineral density during the non-growth, winter season46–48. On
the basis of these patterns, the season of death for the animal can
be established by identifying the growth status of the final mineral
deposit44,49–51.
Four faunal teeth (OB20-01, OB20-06, OB20-07 and OB20-09)
were sufficiently intact to allow cementum analysis, and it was possible to identify at least one region of interest (ROI) per specimen
based on annulation patterns (Supplementary Information). The
four teeth likely represent four distinct individuals: three Ovis (as
indicated by ZooMS and ancient DNA (aDNA)), and one Ovis/Capra
(identified only as sheep/cervid by ZooMS from poor collagen
preservation, but Ovis/Capra based on morphology). Across
these four specimens, the cementum was globally well preserved.
Limited recrystallizations were identified, but not concerning the
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
Table 2 | Osteological information and wear, age (cementum) and sex (DNA) estimates for teeth recovered from early Holocene
cultural layers at Obishir v, along with taxonomic identifications from peptide fingerprinting
Layer
iD number
Species iD
(ZooMS)
DNA sex/ Element
species
Side
Wear
stage
(Payne
1973)
1–2 contact DA-OBI-1518-20.4 Ovis
O.aries,
male
Lower incisor
(permanent)
R
—
2
DA-OBI-1518-20.7 Ovis
—
Lower M1
(permanent)
R
C-I
5 years
Late
cementogenenesis
(late fall?)
2
DA-OBI-1518-20.6 Ovis
O.aries,
female
Lower I3
(deciduous)
L
—
2 years
Late
cementogenenesis
(late fall?)
2
DA-OBI-1518-20.5 Ovis
—
Lower P4
(deciduous)
L
B-C
2.1
DA-OBI-1518-20.9 Deer/Saiga/
sheep/gazelle
—
Upper M1
(permanent)
R
B
1 year
Late
cementogenenesis
(late fall?)
2.2
DA-OBI-1518-20.1
O.aries,
female
Upper M3
(permanent)
L
G-I
5 years
Late
cementogenenesis
(late fall?)
2.2
DA-OBI-1518-20.2 Deer/Saiga/
sheep/gazelle
—
Upper P2
(permanent)
R
—
2.3
DA-OBI-1518-20.8 Capra
—
Upper P3
(permanent)
R
—
Ovis/Capra
Age
(cementum)
Estimated season of Notes
death (cementum)
Direct
radiocarbon
date
4,260 ± 36
7,012 ± 45
Deciduous (<18
months)
Hyper-cementosis 6,989 ± 45
advanced wear
Grey shading indicates DNA-based identification as domestic O. aries.
b
a
m
ntu
me
Ce
Epoxy
Dentine
Dentine
m
tu
en
m
Ce
Epoxy
50 μm
200 μm
Fig. 4 | Cementum analysis of sheep and goat remains from Obishir v. a, OB20-01. A thick deposit of mixed cementum is followed by incremental
acellular cementum with extrinsic fibrils. Within it, four pairs of growth zones + annuli can be seen. The last increment is a complete growth zone,
marking the end of the seasonal record. The cementum is well preserved and is not affected by post mortem modification in the ROI. b, OB20-07. Annuli
are underlain by a line of cementocytes, showing five complete growth zones (the last one constituting the outermost increment). Observations were
conducted in cellular cementum because no extrinsic fibrils of acellular cementum were present on the tooth. For both teeth, observations were conducted
under polarized light with the insertion of a lambda plate under high magnification (500×). Dots show the location of annuli (winter bands); black arrow
indicates the direction of cementum growth.
outermost deposits, and no microbial damage was identified.
Localized weathering influenced the cementum exterior on two
specimens, but in both cases at least one tooth root was sufficiently
preserved to allow reliable observation. Two teeth exhibited evidence of an older age-at-death (4 and 5 years, respectively), while
the remaining two belonged to juvenile animals (1 and 2 years;
Table 2). For all four teeth, the final increment of growth was nearly
equivalent in width to the previous increment, suggesting the animals were at the end of their growth phase (fall or early winter) at
the time of death (Fig. 4). Late-fall livestock culling patterns are a
typical feature of pastoralist herd management52.
Animal DNA. Five tooth fragments from Obishir V, square R8,
layer 2 were selected for genetic analysis, on the basis of preservation and identification as Ovis (n = 4) or Capra (n = 1) through
ZooMS (Table 3). Ancient DNA was successfully extracted from all
five fragments, built into Illumina libraries and sequenced using a
shotgun strategy. We observed high variability in endogenous DNA
preservation across the fragments, ranging from 0.2% to 8.3%, and
all five specimens exhibited patterns of post mortem DNA damage consistent with ancient DNA (Table 3 and Supplementary
Information). We next compared genome-wide sequences of
the Obishir V specimens with published genomes of modern
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
Table 3 | Whole mitogenome DNA results for specimens analysed in this study
Specimen
iD
Species
Reads
sequenced
Endogenous
DNA
No. of
genomewide SNPs
Damage
First
base 3′
Second
base 3′
First
base 5′
Sex
Second
base 5′
Avg frag.
length
No. of mt
reads
mt genome mt haplogroup
coverage
OB20-01
O. aries
18,506,454
8.32%
494,366
0.1242
0.0372
0.1628
0.0309
59.7
XX
718
1.9×
A
OB20-06
O. aries
17,858,515
6.25%
316,536
0.0797
0.0291
0.0787
0.0157
55.86
XX
6,846
13.6×
A
OB21-06
O. aries
25,946,522
0.23%
16,784
0.0758
0.0372
0.0688
0.0165
54.49
XX
399
0.69×
n.d.
OB20-04
O. aries
23,830,365
7.61%
473,283
0.1621
0.0352
0.1683
0.0212
48.97
XY
1589
3.6×
A
OB21-04
Capra sp.
27,141,767
0.27%
19,313
0.0975
0.0419
0.1155
0.0316
59.35
XY
37
0.3×
n.d.
Notes n.d., not determined Genome-wide SNPs are those from the ISGC SNP chip. HVRI and MT-CYTB numbering for sheep is relative to the sheep reference genome NC001941; numbering for goat is
relative to the goat reference genome KR059154.
O. orientalis
0.2
0.1
O. aries, N Europe
O. aries, C Europe
O. aries, SW Europe
PC 3 (2.24 %)
PC 2 (6.09 %)
0.1
C. hircus
Obishir
Obishir 21-4
0.0
O. aries
(modern)
O. orientalis
Obishir 21–4
O. canadensis
O. dalli
0.0
O. aries, Africa
C. hircus
O. aries, SW Asia
–0.1
Obishir
O. aries, Tibet
–0.1
O. aries, Indonesia
O. canadensis
O. dalli
–0.2
O. aries, S Asia
–0.2
0.0
0.1
0.2
0.3
0.0
0.1
0.2
0.3
PC 1 (11.49 %)
PC 1 (11.49 %)
Fig. 5 | PCA results of Obishir v sheep and goat samples. a,b The results for principal component (PC) 1 versus PC 2 (a) and PC 3 (b), projected onto a
reference database of wild sheep taxa as well as domestic sheep and goat specimens from across Eurasia and Africa.
domestic sheep and goats from Eurasia and Africa, as well as wild
Asiatic mouflon using principal component analysis (PCA). Four of
the Obishir V specimens (OB20-01, OB20-06, OB21-06 and OB20-04)
clustered together with present-day domesticated sheep breeds
(O. aries) while being clearly separated from wild sheep species
(Fig. 5). Based on the autosomal-to-sex chromosome ratio in the
samples, three sheep (OB20-01, OB20-06 and OB21-06) were identified as female (Table 3 and Supplementary Information) and one
sheep (OB20-04) was identified as male. Specimens OB20-01 and
OB20-06 were directly dated through ultrafiltration to 6,989 ± 45
and 7,012 ± 45 14C years BP, or between 5,983 and 5,751, and 5,991
and 5,771 cal BCE, respectively (2σ calibrated range). Both specimens showed good measures of quality control (fraction of modern carbon 0.4189 ± 0.00231 (1σ) for OB20-1 and 0.4178 ± 0.0023
(1σ) for OB20-06). No material remained for direct dating of the
third Ovis specimen with poor coverage (OB21-06), while specimen
OB20-04 was dated to the Early Bronze Age (ca. 3,003–2,701 cal
BCE, 2σ). The single Capra tooth (OB21-04), recovered from the
contact between layers 1 and 2, was revealed to be Late Bronze Age
in origin (Supplementary Information).
In PCA space, the Obishir V sheep genomes fall close to modern southwest and Central Asian sheep breeds (Fig. 5). In addition,
the Obishir V sheep carry hypervariable region I (HVRI) polymorphisms that are shared with domestic sheep breeds but that are not
found in modern wild sheep (Supplementary Information). For
three of the four Obishir V sheep (OB20-01, OB20-04 and OB20-06),
we obtained sufficient mitogenome coverage (1.9–13.6x) to perform phylogenetic analysis comparing the Obishir V samples with
160 modern mitogenomes (Supplementary Data C) of O. aries
(domestic sheep), O. ammon (wild argali), O. vignei (wild urial),
O. orientalis ophion (wild Cyprus mouflon) and O. aries musimon
(wild European mouflon). We found that Neolithic Obishir V
sheep, as well as the single Early Bronze Age specimen, all clustered
within O. aries haplogroup A, a major mitochondrial DNA subclade
of domestic sheep (Fig. 6 and Table 3).
Discussion
Our identification of O. aries belonging to mitochondrial haplogroup A at Obishir V ca. 6,000 BCE has critical implications for
our understanding of transcontinental connections and domestic
animal dispersals. Although our assemblage is small and badly
fragmented, ZooMS analysis enabled us to confirm inferences
based on morphological comparisons that the Obishir V economy
included both sheep and goat (Ovis and Capra sp.). Wild Ovis and
Capra in this region include the large argali sheep (O. ammon)
and Siberian ibex (C. sibirica), but a ratio of roughly 3:1 between
Ovis and Capra has characterized pastoralist archaeological
assemblages in Central and Inner Asia since the Bronze Age22,52,
and is also typical of other early western Eurasian pastoral assemblages53. Based on the ZooMS results, Ovis and Capra appear to
account for the majority of the unidentifiable fragments in the
layer 2 faunal assemblage.
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
KF938359
KF938349
KU681202
B1a12
KF938329
KF938339
86
KU6811
1183
KU68
1185
KU68 76
811
KU6
90
811
KU6 192
681
KU
182
681 80
KU
1
a7
681
5
B1
KU 8117
5
6
9
KU
11
68
55
KU 024 4
3
5
KF
24 3
a9
30 22
KF 681 14
B1
2
KU 681 224
KU 681 344
3
KU 938 34
8
KF 93
KF
European mouflon
O. orientalis musimon
HM
HM 23
2 6
KF 36 185
1
KF 302 84
4
3
KF 024 51
3
5
0
0
KU
2
68 449
12
KR
B1
04
8
a6
KF 6867
B2
93
83 8
KU
681 54
2
KF
938 03
348
MF
004
246
KF9
383
50
KF9
3835
7
KF93
8352
KF938
335
KF9383
B1b
33
KF3024
47
KU681200
KU681199
KU681196
KP702285
KF938328
KU681212
NC_001941
.1
AY858379
B1a5
KF9383
56
KF938
355
EF49
0451
KF9
7784
6
KF3
024
53
KF3
0
KF9 2452
383
KF
51
3
KF 02448
9
MF 3834
EF 0042 0
4
4
KF 904 5
KF 302 55
4
KU 302 61
KU 681 4 6 2
KF 68 217
KF 93 121
97 834 3
78
7
45
Haplogroup B
76
61 7
23 617
HM 23 341
M
H 938 346
KF 938 358
KF 938 52
KF 904 6
4
2
5
a
EF 904
B1
0
4
EF 0246
3
KF 38353
9
KF
456
2
0
3
KF
57
024
KF3
458
302
F
K
8
59
024
B1a
3
F
K
1379
8
9
KP
1216
KU68
05
KU6812
EF490453
B1a10
KU681210
EF490454
KR059146
HM236189
HM236186
HM236187
HM236188
NC_02065
6_1
HM2361
88_2
HM23
6188_1
NC_0
2065
6_2
KT7
8168
9_2
KT7
8
KP9 1689_2
813
KP
8
0
98
HM 1378
2
HM 3618
KF 2361 2
83
31
HM 22
38
HM 236
1
KF 236 78
1
9
7
KF 38
9
KP 938 318
KT 99 327
14 847
89 1
68
KF977847
KF938336
KF938332
98
KU6811
94
KU6811
1193
KU68
84
11
8
6
KU
7
8119
KU6
81
811
KU6
11
812
6
U
K
215
681 79
KU
1
8
6 1
7
KU 8117
9
6
KU 8120
6
22
U
K
12
68 207
U
K
81 74
6
KU 2361 21
2
HM 681 206
9
KU 681 21
1
KU 68
KU
b
A1
O. vignei
O. ammon
East Asian
O. aries
Obishir 20-4
KF938331
KU
KU 68
1
KF 681 22
0
KF 938 218
3
9
KF 38 24
3
9
KF 383 21
9
2
KF 383 2
2
9
KF 3831 3
9
9
KF 3833
93
83 4
25
KF3
A2
024
44
KF3
024
4
3
KF3
024
42
KF30
2441
KF30
2440
KF3024
45
KU6812
08
KF302446
KF938345
KF938326
Haplogroup E
87
11 8
68 118 9
KU 68 118
KU 68 191
KU 681 201
KU 681 20
83 0
KU
3
9
18
KF 36 81
2
HM 2361
HM
243
004
MF
472
998
KP 8337
3
KF9 6175
3
HM2
0
3833
KF9
4
0424
MF0
242
MF004
73
KP9984
KP998470
KU575248
KF938342
KF938338
Haplogroup A
a
A1
C. hircus
Cypriot mouflon
O. orientalis ophion
Haplogroup C
Haplogroup D
Obishir 20-1
Obishir 20-6
Fig. 6 | Phylogenetic tree produced using maximum parsimony. The tress shows the whole mitochondrial genome relationships of the Obishir V
specimens (stars) versus modern reference genomes of the domestic sheep O. aries (haplogroups A–E), the European mouflon (O. aries musimon), the
Cypriot mouflon (O. orientalis ophion), the wild urial (O. vignei) and the wild argali (O. ammon), with C. hircus as an outgroup. Major haplogroups A–E are
labelled following previous published studies91. For full details on tree generation, see Supplementary Information.
Further genetic analysis suggests that the Obishir V includes
domestic animals. Mitochondrial DNA data recovered from sheep
teeth confirms that these specimens fall within the present-day
genetic diversity of domestic O. aries, and at least three (two
Neolithic and one Early Bronze Age specimen) are members
of haplogroup A. This haplogroup has been hypothesized to be
linked with the earliest domestication of sheep that widely dispersed throughout Eurasia, as it is the only haplogroup found
among the local sheep populations of North Africa, Britain and
Northern Europe54,55. Significantly, haplogroup A is also found at
high frequency in early domestic sheep populations in East Asia56.
Information available for domesticated sheep suggests that domesticated sheep are monophyletic with five distinct haplogroups
(haplogroups A–E)57,58, with the sole exception of the Cypriot
mouflon (O. orientalis ophion) and the European mouflon (O. aries
musimon). These subspecies are often considered to be relict populations of early domesticated European sheep. The identification of
Obishir V specimens as members of haplogroup A does not rule
out the possibility that this haplogroup was present in a previously
undocumented wild species of Ovis exploited by ancient Central
Asian hunters. However, the fact that all Obishir V specimens with
complete mitochondrial genomes are nested within the range of
established diversity in domesticated sheep but do not form a basal
lineage or cluster with other wild sheep taxa (Fig. 6) casts doubt
on this possibility. Moreover, our PCA comparison of Obishir V
sheep places them between modern South Asian, East Asian and
European populations (Fig. 5).
Based on their co-occurence with O. aries, the Capra sp. specimens recovered at Obishir V may also represent domestic animals.
However, no haplogroup assignment could be made for the Capra
specimen due to the poor preservation of the recovered DNA, and
the only directly dated Capra specimen, recovered from the contact
between layers 1 and 2, was dated to the early first millennium BCE.
As a result, future research is necessary to assess whether the Capra
specimens at Obishir V are indeed reliably associated with the early
occupation layer and how they relate to known ancient populations
of wild and domestic animals.
Careful excavation and direct radiocarbon dating of analysed
specimens (including two teeth identified through aDNA as
O. aries: specimens OB20-01 and OB20-06) confirms the deep
antiquity of the Obishir V sheep specimens. Radiocarbon dating
of faunal remains at Obishir V provides a robust chronology, dating the initial occupation of layers 2 and 3 to before 6,000 BCE.
Bayesian model estimates for the onset of occupation in this layer
range between 7,800 and 6,800 cal BCE at the 2σ range (Table 1). The
earliest radiocarbon date in the early Holocene cultural layer comes
from unidentified wood charcoal (Obishir V, R-11, depth 352, layer
2), which may be influenced by old wood effect and may not reliably date cultural activity at the site. During the early formation of
this layer, geogenic material and archaeological material may have
been re-deposited and mixed together over a period of two or three
millennia. Thus, while human activity at the site might have begun
as early as 9,800–7,800 cal BCE with unidentified faunal materials in
layer 4, specimens identified as Ovis and Capra (as well as those of
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
similar size class but identified only to lower-resolution taxonomic
categories) produced radiocarbon dates clustering around ca. 6,000
BCE (Supplementary Information). Sample sizes are too small to
demonstrate clear patterns of management by humans, but age and
seasonality data from the Obishir V teeth provide important context
supporting the genetic findings, and strengthen the interpretation of
Obishir V sheep as domestic animals. Incremental cementum banding on four intact teeth from Obishir V shows that two animals died
between 1 and 2 years of age, a pattern consistent with early pastoral
culling practices across southwest Asia after 7,500 BCE and probably representing a focus on meat or dairy rather than wool59. The
slaughter of mature females (older than 2 years) that have failed to
reproduce or are small in body size is another hallmark of pastoral
management of a breeding herd12,59,60. Based on cementum analysis,
the female O. aries specimens identified through DNA sequencing
were 2+ years of age and 5+ years of age (Table 2). Another deciduous premolar (<18 monoths) and several unfused metapodial shafts
reinforce the impression of the regular slaughter of juvenile individuals at Obishir V. Analysis of banding patterns indicates seasonal
slaughter of Obishir V sheep and goat in the fall or early winter in
all analysed teeth. Among contemporary pastoral sheep and goat
dietary economies, it is common to cull animals that are unlikely
to survive the winter in late autumn, so as to improve winter herd
survival and store meat more easily over winter months61. The consistency in the apparent seasonality of death among all four teeth
inspected for cementum analysis implies either: (A) seasonal occupation of Obishir V during the fall months paired with hunting of
juveniles and mature females or (B) regular late-season slaughter of
domestic animals.
Although our previous research21 demonstrated that pastoral
economies had been present in the region since the early Bronze
Age, until now, no direct evidence had been found for domestic
animals in montane Central Asia during the early Holocene. Our
findings indicate that the subsistence behaviours that underpinned
these nascent interaction networks could date from at least three
millennia earlier than previously demonstrated. Assuming a single
centre of domestication in the Near East, the evidence from Obishir
V would appear to directly demonstrate the movement of domestic animals deep into the Eurasian interior, either through human
movements or cultural exchange with western Eurasia. The timing of probable domesticated animal use at Obishir V chronologically aligns with the first dispersal of domesticated bovids into
Mediterranean North Africa15 and the arrival of domestic sheep
and goat in the Kopet Dag of southern Turkmenistan34. Therefore,
Obishir V may form an extended outward branch of animals from
southwest Asia by before ca. 6,000 BCE (Fig. 1). In comparison with
the Iranian Plateau and the Levant, the relatively more challenging
ecological setting of montane Central Asia (with its high altitude,
hard frosts and cold winters) implies that such conditions were not
an impermeable ecological barrier for the dispersal of domestic animals during the mid-Holocene, and raise the possibility that similar
assemblages may exist across other areas of Central Asia.
The presence of probable domestic sheep in the Ferghana Basin
by the seventh millennium BCE may warrant reconsideration of
the timing of other components of the ‘Neolithic package’, including goat, cattle and crops such as wheat and barley, in the Inner
Asian Mountain Corridor. Our identification of O. aries dated to
the early Bronze Age (OB20-04, dating to ca. 3,003–2,701 cal BCE at
the 2σ level) from layer 2, along with the faunal remains from layer 1
(Supplementary Data A, B) point to continuous use of domestic
animals in the Ferghana region from their Neolithic introduction
throughout the Middle Ages. In later history and prehistory, the
southern Ferghana region became linked to northwest China
through networks of pastoral mobility that drove exchange across
the interior62. Our research shows that such networks could date
from at least three millennia earlier than previously demonstrated.
Based on the material culture links between Obishir V and other
important Neolithic horizons (such as Hissar and Kelteminar), as
well as the paucity of scientific study on highly fragmented early
faunal assemblages from the Pamiro-Alay and Tian Shan region,
we suggest that Obishir V may form part of a much broader phenomenon linked with early pastoralist or agropastoralist economies
(Fig. 1). If so, the region may have been a key node in an early
dispersal system for southwest Asian domesticated bovids and,
possibly, plants into other areas of East and South Asia. No archaeobotanical data are yet available for Obishir V. Future investigation of
this and other mid-Holocene sites in Central Asia should assess the
possibility that domestic plants were also dispersed into the region
during the mid-Holocene, and clarify whether this newly identified
Neolithic expansion was demic (accompanied by human dispersals, as in much of Eastern Europe) or took place through transfer
of organisms and technologies, as in much of non-Mediterranean
Africa63 to better understand the cultural and ecological dynamics of the Neolithic transition across the ancient world. Until now,
archaeological data have pointed to a relatively late arrival of the
agropastoral economy to interior Central Asia, around 3,500 BCE.
Results from Obishir V suggest that the chronology of Neolithization
in this region is in need of revision. Archaeological investigation
of the early Holocene deposits at the rockshelter of Obishir V have
revealed major changes in material culture, along with fragmented
faunal material potentially linked to an early pastoral economy.
Using a combination of collagen fingerprinting, ancient DNA and
thin-section cementum analysis, we propose an early dispersal of
domestic sheep into the Ferghana Valley by at least 6,000 BCE, concurrent with other out-migrations from Western Asia into the Old
World. These results support Soviet scholarship hypothesizing a
Neolithic dispersal of food-producing economies into Central Asia
through connections with Western Asia, and raise the possibility
that domesticated sheep, perhaps in association with other early
domesticated plants and animals, could have dispersed into the continental interior millennia earlier than previously recognized.
Methods
Excavation. Excavations at the Obishir V site were carried out between 2015 and
2019 by a joint Russian–Kyrgyzstani archaeological expedition. In 2015–2016,
researchers working at the site excavated an area of 8 m2 adjoining the 1968–1969
excavation area (Supplementary Information). Each artefact (lithic and faunal
remains) greater than 1 cm in size was individually piece-plotted with 1 mm
accuracy using a Leica Total Station TS02 plus in tandem with the Trimble software
EDM Mobile, and documented along with contextual information (orientation,
dip, inclination, find category, stratigraphic level, etc.). Fragments smaller than this
were collected in bulk for each 0.25 m2 area. Features (natural or anthropogenic)
were recorded with the total station using outlines, surface points and breaklines.
Large surfaces, such as walls of old excavation pits, sections and archaeological
horizons with numerous artefacts, were also documented using photogrammetry
and three-dimensional (3D) scanning. Digital records were supplemented by field
drawings and notes. All sediment obtained during the excavation was wet-screened
with mesh of size of 2.0 mm, and finds were documented according to excavated
square, layer and depth. Stratigraphic units were defined during the excavation
based on grain size composition, consistency, colour, presence of erosional surfaces
and accumulation of limestone clasts.
Radiocarbon dating. Nine radiocarbon samples of identified and unidentified
animal bone and tooth (n = 7) and charcoal (n = 2) were selected from various
depths across the stratigraphic unit of layers 2–4 and submitted for radiocarbon
dating at the Oxford Radiocarbon Accelerator Laboratory, the Center for Isotope
Research at the University of Groningen, the Golden Valley Laboratory at
Novosibirsk and the University of Arizona AMS Laboratory. Portions of domestic
sheep teeth OB20-01, OB20-06 and OB20-04 from layer 2 used in genomic analysis
were submitted to the University of Arizona Accelerator Mass Spectrometry
Laboratory. Collagen was obtained from dentin using acid–base–acid (ABA)
pretreatment, gelatinization, 0.45 micron filtration and ultrafiltration. The quality
control parameters, collagen yield of 7.1%, 6.4% and 7.2%, carbon yield of 42.9%,
36% and 36% and d13C values of −19.7, −19.3 and −19.6 ± 0.1 per mil, indicated
good preservation64. The atomic C:N ratio for OB20-01 (3.2) and OB20-04
(3.2) also fell within ranges indicating good preservation, although no material
remained from tooth OB20-06 for C:N measurement after 14C analysis. All dates
were calibrated in OxCal using the INTCAL20 calibration curve65.
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
We performed Bayesian stratigraphic phase modelling (using ordered phases and
combining layers 2 and 3 due to their apparent mixing) in the software OxCal
with a uniform prior using all available radiocarbon dates from Obishir V, and
two thermoluminescence dates from layer 5, following methods outlined by
Ramsey66. Because charcoal dates, particularly those from layer 2, suggest an old
wood effect that produced unsuccessful models in OxCal, only radiocarbon dates
from bone and teeth were included in the final chronology. The models provided
good agreement, and repeating the analysis using a general outlier model did not
significantly alter estimated parameters. The OxCal code used in the analysis,
along with the human tooth dating methods, are outlined in Supplementary
Information.
Thermoluminescence dating. During the 2017 field season, two sediment
samples were collected from the southern portion of the western excavation
profile of layer 5 at a depth of 295 and 335 cm below the surface. We measured
the deposit moisture in each sample and, after drying, determined the dose rate
(DR) using a MAZAR gamma spectrometer. Concentrations of 226Ra, 228Th and
40
K in each sample were obtained from 20 measurements lasting 2,000 s each.
We established equivalent dose (ED) on the 63–80 mm polymineral fraction,
after 10% HCl and 30% H2O2 washing and ultraviolet optical bleaching. The
samples were irradiated with 20, 30, 40, 50 and 100 Gy doses from 60Co gamma
source. Before measurement, we heated the samples at 140 °C for 3 h. A sample
pre-treated in this way was used to determine the ED by the thermoluminescence
multiple-aliquot regenerative technique67, according to the description published
by Fedorowicz et al.68. Curve registration was performed on a RA’94 (Mikrolab)
thermoluminescence reader coupled with an EMI 9789 QA photomultiplier.
Finally, we calculated thermoluminescence age according to Frechen69. A detailed
description of the preparation and the equipment used in the study is provided in
Fedorowicz et al.68.
Zooarchaeology and ZooMS. For each specimen, we conducted taphonomic
analysis and morphology-based species identification using comparative faunal
collections held by the Russian Academy of Sciences-Siberia in their field station at
Aidarkyen, Kyrgyzstan, and attempted refits for all specimens to control for issues
of fragmentation. We recorded taphonomic indicators such as rodent and carnivore
damage, root etching, and weathering70, along with evidence for anthropogenic
modification such as spiral fracturing, cut marks and burning via visual
inspection (Supplementary Data A). In addition, we used collagen fingerprinting
(ZooMS) to taxonomically identify a subset of this material, all derived from 2017
excavations. We analysed all bone specimens from square R8, layers 2.2, 3 and 4,
as well as all bone specimens from square R8, layer 1.3 using ZooMS. To do so,
we demineralized 10 mg of bone using 50 mM ammonium bicarbonate buffer
(Sigma-Aldrich) produced in UHQ water and pH adjusted to 8.0 using
ammonium hydroxide (1 M), following the protocol outlined by van Doorn et al.71.
The extracted collagen was digested into component peptides using trypsin
(Pierce), and was purified using Pierce C18 tips (Thermo Scientific), eluting in
50% acetonitrile (Sigma-Aldrich) and 0.1% trifluoroacetic acid (Sigma-Aldrich).
Samples were then spotted on Bruker AnchorChip with Bruker Peptide
Calibration Standard in calibration spots directly neighbouring the samples.
Mass spectrometric analysis was conducted using a Bruker Autoflex Speed LRF
matrix-assisted laser desorption/ionization–time of flight (MALDI-TOF) device
in the dedicated ZooMS Laboratory at the Max Planck Institute for the Science of
Human History, Jena, Germany. The acquisition used the following parameters:
4,000 laser shots at 50–60% intensity (50 shots per spot), mass range 600–3,500 Da
and reflector mode. Identifications were made using published reference spectra
from the Eurasian mammals database72 and are reported according to the level
of taxonomic specificity. In some cases (for example, Ovis versus muskox and
chamois), species that were not necessarily separable on the basis of observed
peptide markers were inferred on the basis of known habitat distribution. All
peptide marker data are provided in Supplementary Data B.
Cementum analysis. First, all analysed teeth were 3D scanned using a MicroCT
scanner, and washed in 90% ethanol solution. A cast of the occlusal surface was
retained for later micro-wear analyses. Thin sections were prepared following
recommendations by Rendu73. Teeth were embedded in epoxy within a vacuum
chamber (at 0.2 bar pressure). They were then cut at thickness of 250 µm using an
Isomet5000 automatic saw, and ground to thickness of 100 µm with a PressiCube
grinding machine. For each tooth, four independent thin sections were prepared
and analysed independently. Observations were conducted under polarized light
both with and without the insertion of a lambda plate74, using a Leica 2500P
microscope equipped with a Leica DC 120 camera. Post-depositional modifications
were analysed systematically following Geusa et al.75 and Stutz74. Measurements of
the increments were conducted using ImageJ software following methods outlined
by Lieberman et al.76.
DNA analysis. Extraction. DNA was extracted from six teeth using silica-based
purification in a dedicated laboratory for aDNA at PACEA (UMR 5199, University
of Bordeaux). We incubated 100–250 mg of bone powder in 1 mL extraction buffer
(0.5 M EDTA, 0.25 M Na2HPO4, 50 μg mL−1 proteinase K) for 48–72 h at 37 °C.
DNA was purified from the collected supernatant using a method adapted from the
QIAquick MinElute purification kit protocol (Qiagen).
PCR amplification. Purified DNA was amplified by quantitative polymerase chain
reaction (qPCR) using a LightCycler 96 Instrument (Roche Life Science). Primer
pairs were designed to amplify two regions of the mitochondrial DNA, a 130 bp
region from cytB and 170 bp from the hypervariable region (see Supplementary
Information for a list of the primers used and product sizes). We obtained each
PCR product at least twice, conducting standard negative control to ensure
the authenticity of the results. We used 1 µL of extract in 10 µL reaction with
1× FastStart Essential DNA Green Master and 1 µM of each primer. Products were
sequenced by capillary electrophoresis at Genewiz (Leipzig, Germany). Sequences
were visually inspected and manually curated, assembled and aligned using
Geneious software suite v9.1.8 (https://www.geneious.com).
Library construction and sequencing. We used DNA extracts to build
double-stranded Illumina libraries following the protocol described by Gorge
et al.77. We sequenced whole-genome shotgun libraries for ~500,000 reads on
an Illumina NextSeq 500 at the Institut de Recherche Biomédicale des Armées
in Brétigny-sur-Orge, France, to assess DNA preservation. DNA extracts of five
samples produced in Bordeaux were sent to the dedicated cleanroom facilities of
the Max Planck Institute for the Science of Human History. There, 20 μL of the
extract was used to build double-stranded libraries with partial treatment of the
uracil-DNA glycosylase enzyme (UDG-half)78. The resulting genetic libraries were
double indexed and further amplified79. Finally, all five samples were shotgun
sequenced for ~20 M reads on an Illumina HiSeq run with a 75-cycle single-end
configuration. Sequence authenticity was estimated on libraries subjected to either
partial or no USER treatment using MapDamage80.
We trimmed adapter sequences from generated sequences, merging
overlapping paired-end reads and filtering out reads shorter than 30 bases using
Clip and Merge. We then mapped our filtered reads to both the O. aries genome
assembly (International Sheep Genome Consortium build Oarv3.1) and the
C. hircus genome assembly (ARS1) using BWA81, and against the full mitochondrial
genome of O. aries (NC_001941.1) with a duplication of its first 500 bases at the
end to ensure mapping of the reads overlapping the junction resulting from the
virtual linearization of the circular mitogenome.
PCA of genome-wide sequences. Genotypes were called for the 38.5M
single-nucleotide polymorphisms (SNPs) from the International Sheep Genomics
Consortium (ISGC) SNP project from 60 domesticated sheep along with three
specimens of North American bighorn sheep (O. canadensis) and two Dall
sheep (O. dalli). We drew a single allele at random for each position (minimum
mapping and base quality of 30) using PileupCaller (https://github.com/stschiff/
sequenceTools), rendering the individuals from the dataset homozygous for each
locus. The two files were then merged using Plink v1.982. To augment the ISGC
SNP discovery panel, we added published genome data of 16 Asiatic mouflons
(O. orientalis)83. For this, we downloaded FastQ files for the 16 individuals from the
National Center for Biotechnology Information Sequence Read Archive under the
accession number PRJNA24020 and aligned reads to the oviAri4 reference genome
using BWA-mem v0.7.1781. We removed PCR duplicates using the MarkDuplicates
module of the picardtools program v2.20.0 (https://broadinstitute.github.io/
picard/). We then retained properly paired reads with mapping quality score 30
or higher using SAMtools v1.984. For the SNPs from the ISGC SNP project, we
calculated genotype likelihoods using GATK UnifiedGenotyper (v3.8.1.0) with
‘–genotype_likelihoods_model SNP–output_mode EMIT_ALL_SITES –allSitePLs’
options85. Then, we used an in-house python script to calculate posterior genotype
probability with a non-default prior [0.4995, 0.0010, 0.4995] to reduce reference
bias, following the approach taken in the Simons Genome Diversity Project86. We
took genotypes with posterior probability 0.9 or higher and kept the remaining
ones as missing. Finally, we calculated sequencing coverage with Qualimap v2.2.1,
assigning biological sex to each individual based on the X to autosome coverage
ratio (males around 0.5, females around 1.0). Modern wild and domestic sheep
samples used from the ISGC, along with 16 Asiatic mouflons, and the three goat
genomes are indicated in Supplementary Data D. We then conducted PCA using
smartPCA implemented in EIGENSOFT, projecting ancient sequences from
Obishir onto the first three components of the PCA defined by the full dataset
(including goat, wild sheep and domestic sheep reference specimens) to visualize
diversity among sheep species.
Mitochondrial genome analysis. We produced a multiple genome alignment of
our newly reconstructed sequences (excluding OB21-04 and OB21-06 because
of low coverage) along with 160 present-day sequences from various breeds of
modern O. aries, O. ammon, O. vignei, O. orientalis ophion and O. aries musimon
(Supplementary Data C), using a modern C. hircus as outgroup, with Geneious
software suite v9.1.8. Gaps were removed from the alignment. Phylogenetic trees
were inferred using both the maximum-parsimony (MP) method (Fig. 6) and
maximum likelihood (ML) with Tamura–Nei model (Supplementary Information)
in MEGA version X87–89. We visualized and manipulated phylogenetic trees in
FIGTREE v1.4.1 (http://tree.bio.ed.ac.uk/software/figtree/).
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
Reporting summary. Further information on research design is available in the
Nature Research Reporting Summary linked to this article.
Data availability
Shotgun sequencing raw files are available at the European Nucleotide Archive
(ENA) database under accession number PRJEB41594.
Received: 7 September 2020; Accepted: 17 February 2021;
Published: xx xx xxxx
References
1. Asouti, E. & Fuller, D. Q. A contextual approach to the emergence of
agriculture in southwest Asia: reconstructing early Neolithic plant-food
production. Curr. Anthropol. 54, 299–345 (2013).
2. Conolly, J. et al. Meta-analysis of zooarchaeological data from SW Asia and
SE Europe provides insight into the origins and spread of animal husbandry.
J. Archaeol. Sci. 38, 538–545 (2011).
3. Larson, G. et al. Current perspectives and the future of domestication studies.
Proc. Natl Acad. Sci. U. S. A. 111, 6139–6146 (2014).
4. Willcox, G. Measuring grain size and identifying Near Eastern cereal
domestication: evidence from the Euphrates Valley. J. Archaeol. Sci. 31,
145–150 (2004).
5. Willcox, G. & Stordeur, D. Large-scale cereal processing before domestication
during the tenth millennium cal BC in northern Syria. Antiquity 86,
99–114 (2012).
6. Larson, G. & Fuller, D. Q. The evolution of animal domestication. Annu. Rev.
Ecol. Evol. Syst. 45, 115–136 (2014).
7. Bocquet‐Appel, J. Paleoanthropological traces of a Neolithic demographic
transition. Curr. Anthropol. 43, 637–650 (2002).
8. Bellwood, P. First Farmers: the Origins of Agricultural Societies (Wiley, 2004).
9. Omrak, A. et al. Genomic evidence establishes Anatolia as the source of the
European Neolithic gene pool. Curr. Biol. 26, 270–275 (2016).
10. Olalde, I. et al. The genomic history of the Iberian Peninsula over the past
8000 years. Science 363, 1230–1234 (2019).
11. Brace, S. et al. Ancient genomes indicate population replacement in early
Neolithic Britain. Nat. Ecol. Evol. 3, 765–771 (2019).
12. Zeder, M. A. & Hesse, B. The initial domestication of goats (Capra hircus) in
the Zagros mountains 10,000 years ago. Science 287, 2254–2257 (2000).
13. Daly, K. G. et al. Ancient goat genomes reveal mosaic domestication in the
Fertile Crescent. Science 361, 85–88 (2018).
14. Alberto, F. J. et al. Convergent genomic signatures of domestication in sheep
and goats. Nat. Commun. 9, 813 (2018).
15. Zeder, M. A. in Human Dispersal and Species Movement: from Prehistory to
the Present (eds Petraglia, M.D., Crassard, R. & Boivin, N.) p. 261
(Cambridge Univ. Press, 2017).
16. Vigne, J.-D. Early domestication and farming: what should we know or do for
a better understanding? Anthropozoologica 50, 123–151 (2015).
17. Pereira, F. & Amorim, A. Encyclopedia of Life Sciences https://doi.
org/10.1002/9780470015902.a0022864 (2010).
18. Hermes, T. R. et al. Mitochondrial DNA of domesticated sheep confirms
pastoralist component of Afanasievo subsistence economy in the Altai
Mountains (3300–2900 cal BC). Archaeol. Res. Asia 24, 100232 (2020).
19. Wilkin, S. et al. Dairy pastoralism sustained eastern Eurasian steppe
populations for 5,000 years. Nat. Ecol. Evol. 4, 346–355 (2020).
20. Hermes, T. R. et al. Early integration of pastoralism and millet cultivation in
Bronze Age Eurasia. Proc. Biol. Sci. 286, 20191273 (2019).
21. Taylor, W. et al. Early pastoral economies along the Ancient Silk Road:
biomolecular evidence from the Alay Valley, Kyrgyzstan. PLoS ONE 13,
e0205646 (2018).
22. Spengler, R. N. & Willcox, G. Archaeobotanical results from Sarazm,
Tajikistan, an early Bronze Age settlement on the edge: agriculture and
exchange. Environ. Archaeol. 18, 211–221 (2013).
23. Korobkova, G. F. Tools and Economy of the Neolithic Populations in Central
Asia (Nauka, 1969).
24. Masson, V. M. The Jeitun Settlement: the Emergence of a Productive Economy
(Nauka, 1971).
25. Itina, M.A. History of Steppe Tribes of Southern Aral Sea Region (2–1 Ka BP)
(Nauka, 1977).
26. Lamberg-Karlovsky, C. C. The Bronze Age khanates of Central Asia. Antiquity
68, 398–405 (1994).
27. Brunet, F. Pour une nouvelle étude de la culture néolithique de Kel’teminar.
Ouzbékistan. Paléorient 31, 87–105 (2005).
28. Masson, V. M. The Bronze Age in Khorasan. Hist. Civiliz. Cent. Asia 1,
225 (1999).
29. Ranov, V. A. Hissar culture - Neolithic mountain regions of Central Asia. in
Stone Age of Northern, Middle and Eastern Asia 27–28 (Nauka, 1985).
30. Yablonsky, L. T. Kelteminar craniology. Intra-group analysis. Sov. Ethnogr.,
Mosc., USSR Acad. Sci. 2, 127–140 (1985).
31. Vinogradov, A. V. Drevnie okhotniki i rybolovy Sredneaziatskogo
Mezhdurechija (Former hunters and fishermen of Central Asian Mesopotamia)
(Nauka (Science), 1981a).
32. Vinogradov, A. V. in Kul ‘tura i iskusstvo drevnego Khorezma (Culture
and Art of Ancient Khoresam) (eds Itina, M. A., Raporort, J. A. et al.)
pp. 88–98 (1981b).
33. Harris, D. R., Origins of Agriculture in Western Central Asia https://doi.
org/10.9783/9781934536513 (Univ. of Pennsylvania, 2010).
34. Dolukhanov, P. M. The ecological prerequisites for early farming in southern
Turkmenia. Sov. Anthropol. Archeol. 19, 359–385 (1981).
35. Lisitsina, G. N. in The Bronze Age Civilization of Central Asia: Recent Soviet
Discoveries (ed. Kohl, P. L.) pp. 350–358 (M. E. Sharpe, 1981).
36. Larkum, M. in Origins of Agriculture in Western Central Asia: An
Environmental–Archaeological Study (ed. Harris, D.) pp. 142–149 (Univ. of
Pennsylvania Museum of Archaeology and Anthropology, 2010).
37. Ranov, V. A. & Korobkova, G. F., Tutkaul – multilayered settlement site of the
Gissar culture in southern Tajikistan. Sov. Archaeol. 133–147 (1971).
38. Islamov, U. I., Obishirian Culture (FAN, 1980).
39. Fedorchenko, A. Y. et al. Personal ornament production technology in the
early Holocene complexes of western Central Asia: insights from Obishir-5.
Archaeol., Ethnol. Anthropol. Eurasia 46, 3–15 (2018).
40. Szymczak, K. & Khudzhanazarov, M., Exploring the Neolithic of the
Kyzyl-Kums. Ayakagytma “The Site” and other collections. Swiatowit Suppl.
Ser. P: Prehistory and Middle Ages 11. Central Asia–Prehist. Stud (2006).
41. Buckley, M., Collins, M., Thomas-Oates, J. & Wilson, J. C. Species
identification by analysis of bone collagen using matrix-assisted laser
desorption/ionisation time-of-flight mass spectrometry. Rapid Commun. Mass
Spectrom. 23, 3843–3854 (2009).
42. Kerven, C., Steimann, B., Ashley, L., Dear, C. & ur Rahim, I. Pastoralism and
Farming in Central Asia’s Mountains: a Research Review (Univ. of Central
Asia, 2011).
43. Lieberman, D. E. Life history variables preserved in dental cementum
microstructure. Science 261, 1162–1164 (1993).
44. Klevezal, G. A. Recording Structures of Mammals: Determination of Age and
Reconstruction of Life History (A. A. Balkema, 1996).
45. Diekwisch, T. G. The developmental biology of cementum. Int. J. Dev. Biol.
45, 695–706 (2001).
46. Klevezal, G. & Kleinenberg, S. Age determination of mammals from annual
layers in teeth and bones of mammals. Trans. from Russian by the Israel
Program for Scientific Translations, Jerusalem (1969).
47. Grue, H. & Jensen, B., Review of the formation of incremental lines in tooth
cementum of terrestrial mammals [age determination, game animal,
variation, sex, reproductive cycle, climate, region, condition of the animal].
Dan. Rev. Game Biol. 11 (1979).
48. Stock, S. R. et al. Cementum structure in Beluga whale teeth. Acta Biomater.
48, 289–299 (2017).
49. Gordon, B. C. Of Men and Reindeer Herds in French Magdalenian Prehistory
https://doi.org/10.30861/9780860545040 (BAR Publishing, 1988).
50. Pike-Tay, A. Red Deer Hunting in the Upper Paleolithic of South-West France:
a Study in Seasonality (BAR Oxford, 1991).
51. Burke, A. & Castanet, J. Histological observations of cementum growth in
horse teeth and their application to archaeology. J. Archaeol. Sci. 22,
479–493 (1995).
52. Frachetti, M. D. Multiregional emergence of mobile pastoralism and
nonuniform institutional complexity across Eurasia. Curr. Anthropol. 53,
2–38 (2012).
53. Stiner, M. C. et al. A forager–herder trade-off, from broad-spectrum hunting
to sheep management at Aşıklı Höyük, Turkey. Proc. Natl Acad. Sci. U. S. A.
111, 8404–8409 (2014).
54. Demirci, S. et al. Mitochondrial DNA diversity of modern, ancient and wild
sheep (Ovis gmelinii anatolica) from Turkey: new insights on the evolutionary
history of sheep. PLoS ONE 8, e81952 (2013).
55. Meadows, J. R. S., Hiendleder, S. & Kijas, J. W. Haplogroup relationships
between domestic and wild sheep resolved using a mitogenome panel.
Heredity 106, 700–706 (2011).
56. Cai, D. et al. Early history of Chinese domestic sheep indicated by ancient
DNA analysis of Bronze Age individuals. J. Archaeol. Sci. 38, 896–902 (2011).
57. Meadows, J. R. S., Cemal, I., Karaca, O., Gootwine, E. & Kijas, J. W. Five
ovine mitochondrial lineages identified from sheep breeds of the Near East.
Genetics 175, 1371–1379 (2007).
58. Bruford, M. W. & Townsend, S. J., in Documenting Domestication: New
Genetic and Archaeological Paradigms (eds Zeder, M. A., Bradley, D. G. &
Emshwiller, E. A.) pp 306–316 (Univ. of California Press, 2006).
59. Arbuckle, B. S. & Atici, L. Initial diversity in sheep and goat management in
Neolithic south-western. Asia. Levant. 45, 219–235 (2013).
60. Hesse, B. Slaughter patterns and domestication: the beginnings of pastoralism
in western Iran. Man 17, 403–417 (1982).
61. Fijn, N., Living with Herds: Human–Animal Coexistence in Mongolia
(Cambridge Univ. Press, 2011).
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
ARTICLES
NATURE HUMAN BEHAVIOUR
62. Frachetti, M. D., Smith, C. E., Traub, C. M. & Williams, T. Nomadic ecology
shaped the highland geography of Asia’s Silk Roads. Nature 543, 193 (2017).
63. Jerardino, A., Fort, J., Isern, N. & Rondelli, B. Cultural diffusion was the main
driving mechanism of the Neolithic transition in Southern Africa. PLoS ONE
9, e113672 (2014).
64. van Klinken, G. J. Bone collagen quality indicators for palaeodietary and
radiocarbon measurements. J. Archaeol. Sci. 26, 687–695 (1999).
65. Reimer, P. J. et al. The IntCal20 Northern Hemisphere radiocarbon age
calibration curve (0–55 cal kBP). Radiocarbon 62, 725–757 (2020).
66. Ramsey, C. B. Bayesian analysis of radiocarbon dates. Radiocarbon 51,
337–360 (2009).
67. Wintle, A. G. & Prószyńska, H. TL dating of loess in Germany and Poland.
PACT 9, 547–554 (1983).
68. Fedorowicz, S. et al. Loess–paleosol sequence at Korshiv (Ukraine):
chronology based on complementary and parallel dating (TL, OSL), and
litho-pedosedimentary analyses. Quat. Int. 296, 117–130 (2013).
69. Frechen, M. Systematic thermoluminescence dating of two loess profiles from
the Middle Rhine Area (FRG). Quat. Sci. Rev. 11, 93–101 (1992).
70. Behrensmeyer, A. K. Taphonomic and ecologic information from bone
weathering. Paleobiology 4, 150–162 (1978).
71. van Doorn, N. L., Hollund, H. & Collins, M. J. A novel and non-destructive
approach for ZooMS analysis: ammonium bicarbonate buffer extraction.
Archaeol. Anthropol. Sci. 3, 281 (2011).
72. Welker, F. et al. Palaeoproteomic evidence identifies archaic hominins
associated with the Châtelperronian at the Grotte du Renne. Proc. Natl Acad.
Sci. U. S. A. 113, 11162–11167 (2016).
73. Rendu, W. Hunting behavior and Neanderthal adaptability in the Late
Pleistocene site of Pech-de-l’Azé I. J. Archaeol. Sci. 37, 1798–1810 (2010).
74. Stutz, A. J. Polarizing microscopy identification of chemical diagenesis in
archaeological cementum. J. Archaeol. Sci. 29, 1327–1347 (2002).
75. Geusa, G. et al. in Osteodental Biology of the People of Portus Romae
(Necropolis of Isola Sacra, 2nd-3rd Cent. AD) (eds Bondioli, L. &
Macchiarelli, R.) (Roma, 1999).
76. Lieberman, D. E., Deacon, T. W. & Meadow, R. H. Computer image
enhancement and analysis of cementum increments as applied to teeth of
Gazella gazella. J. Archaeol. Sci. 17, 519–533 (1990).
77. Gorgé, O. et al. Analysis of ancient DNA in microbial ecology. Methods Mol.
Biol. 1399, 289–315 (2016).
78. Rohland, N., Harney, E., Mallick, S., Nordenfelt, S. & Reich, D. Partial
uracil-DNA-glycosylase treatment for screening of ancient DNA. Philos.
Trans. R. Soc. Lond. B Biol. Sci. 370, 20130624 (2015).
79. Kircher, M., Sawyer, S. & Meyer, M. Double indexing overcomes inaccuracies
in multiplex sequencing on the Illumina platform. Nucleic Acids Res. 40,
e3 (2012).
80. Ginolhac, A., Rasmussen, M., Gilbert, M. T. P., Willerslev, E. & Orlando, L.
mapDamage: testing for damage patterns in ancient DNA sequences.
Bioinformatics 27, 2153–2155 (2011).
81. Li, H. & Durbin, R. Fast and accurate long-read alignment with Burrows–
Wheeler transform. Bioinformatics 26, 589–595 (2010).
82. Purcell, S. et al. PLINK: a tool set for whole-genome association and
population-based linkage analyses. Am. J. Hum. Genet. 81, 559–575 (2007).
83. Li, X. et al. Whole-genome resequencing of wild and domestic sheep
identifies genes associated with morphological and agronomic traits.
Nat. Commun. 11, 2815 (2020).
84. Li, H. et al. The sequence alignment/map format and SAMtools.
Bioinformatics 25, 2078–2079 (2009).
85. DePristo, M. A. et al. A framework for variation discovery and genotyping
using next-generation DNA sequencing data. Nat. Genet. 43, 491 (2011).
86. Mallick, S. et al. The Simons Genome Diversity Project: 300 genomes from
142 diverse populations. Nature 538, 201–206 (2016).
87. Kumar, S., Stecher, G., Li, M., Knyaz, C. & Tamura, K. MEGA X: molecular
evolutionary genetics analysis across computing platforms. Mol. Biol. Evol. 35,
1547–1549 (2018).
88. Stecher, G., Tamura, K. & Kumar, S. Molecular evolutionary genetics analysis
(MEGA) for macOS. Mol. Biol. Evol. 37, 1237–1239 (2020).
89. Tamura, K. & Nei, M. Estimation of the number of nucleotide substitutions
in the control region of mitochondrial DNA in humans and chimpanzees.
Mol. Biol. Evol. 10, 512–526 (1993).
90. Murphy, D. J. People, Plants and Genes: The Story of Crops and Humanity
(Oxford Univ. Press, 2007).
91. Lv, F.-H. et al. Mitogenomic meta-analysis identifies two phases of
migration in the history of eastern Eurasian sheep. Mol. Biol. Evol. 32,
2515–2533 (2015).
Acknowledgements
The authors thank D. Paul and S. Palstra for performing radiocarbon dating of tooth
enamel, and E. Rannamäe for assistance with manuscript preparation. Cementum
analyses were funded through the CemeNTAA project, via the French National Agency
for Research (ANR-14-CE31-0011). Geological investigations were supported by the
National Science Center, Poland (grant no. 2018/29/B/ST10/00906). Sampling for
ZooMS, DNA and radiocarbon analysis (Golden Valley Laboratory) and lithic analysis
of Obishir V were supported by RSF project no. 19-78-10053, ‘The emergence of
food-producing economies in the high mountains of interior Central Asia’. Ancient
DNA analyses were conducted with the support of the palaeogenomic platform from the
UMR5199 PACEA Universite de Bordeaux and the European Research Council under
the European Union’s Horizon 2020 research and innovation programme under grant
agreement no. 804884-DAIRYCULTURES. The funders had no role in study design, data
collection and analysis, decision to publish or preparation of the manuscript.
Author contributions
W.T.T.T. and S.Sh. designed the research, collected data, conducted analysis and wrote
the manuscript. M.P., C.P., A.A., W.R., C.J., T.H., and C.W. collected data, conducted
analysis and helped to write the manuscript. M.T.K., G.B., S.Sc., G.H., R.Sp., R.St., J.M.,
A.S., S.F., L.O., K.D. and A.K. collected data, conducted analysis and assisted in data
interpretation. All authors reviewed the manuscript.
Competing interests
The authors declare no competing interests.
Additional information
Supplementary information The online version contains supplementary material
available at https://doi.org/10.1038/s41562-021-01083-y.
Correspondence and requests for materials should be addressed to W.T.T.T. or S.S.
Peer review information Nature Human Behaviour thanks Suzanne Birch, Laurent
Frantz and Eve Rannamae for their contribution to the peer review of this work.
Reprints and permissions information is available at www.nature.com/reprints.
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
© The Author(s), under exclusive licence to Springer Nature Limited 2021
NATuRE HuMAN BEHAviOuR | www.nature.com/nathumbehav
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Last updated by author(s): Sep 11, 2020
Reporting Summary
Nature Research wishes to improve the reproducibility of the work that we publish. This form provides structure for consistency and transparency
in reporting. For further information on Nature Research policies, see our Editorial Policies and the Editorial Policy Checklist.
nature research | reporting summary
Corresponding author(s): William Taylor
Statistics
For all statistical analyses, confirm that the following items are present in the figure legend, table legend, main text, or Methods section.
n/a Confirmed
The exact sample size (n) for each experimental group/condition, given as a discrete number and unit of measurement
A statement on whether measurements were taken from distinct samples or whether the same sample was measured repeatedly
The statistical test(s) used AND whether they are one- or two-sided
Only common tests should be described solely by name; describe more complex techniques in the Methods section.
A description of all covariates tested
A description of any assumptions or corrections, such as tests of normality and adjustment for multiple comparisons
A full description of the statistical parameters including central tendency (e.g. means) or other basic estimates (e.g. regression coefficient)
AND variation (e.g. standard deviation) or associated estimates of uncertainty (e.g. confidence intervals)
For null hypothesis testing, the test statistic (e.g. F, t, r) with confidence intervals, effect sizes, degrees of freedom and P value noted
Give P values as exact values whenever suitable.
For Bayesian analysis, information on the choice of priors and Markov chain Monte Carlo settings
For hierarchical and complex designs, identification of the appropriate level for tests and full reporting of outcomes
Estimates of effect sizes (e.g. Cohen's d, Pearson's r), indicating how they were calculated
Our web collection on statistics for biologists contains articles on many of the points above.
Software and code
Policy information about availability of computer code
Data collection
No software used
Data analysis
Radiocarbon dates
Bayesian analysis of radiocarbon dates was performed in OxCal v. 4.4 (https://c14.arch.ox.ac.uk/oxcal.html)
ZooMS
Mass spectrometry identifications were performed using the open-source software MMass (http://www.mmass.org/)
Cementum analysis
Measurements of the increments were conducted using ImageJ (https://imagej.nih.gov/ij).
PCR amplification
Sequences were visually inspected and manually curated, assembled and aligned using Geneious software suite v9.1.8 (https://
www.geneious.com).
April 2020
Library construction and sequencing
Sequence authenticity was estimated on libraries subjected to either a partial or no USER treatment using MapDamage. We drew alleles at
random using PileupCaller (https://github.com/stschiff/sequenceTools), rendering the individuals from the dataset homozygous for each
locus. The two files were then merged using Plink v1.9. We then conducted principal component analysis using smartPCA implemented in
EIGENSOFT. We mapped our filtered reads to both the Ovis aries genome assembly (international Sheep Genome consortium build Oarv3.1)
and the Capra hircus genome assembly (ARS1) using BWA. We produced a multiple genome alignment of our newly reconstructed sequences
with Geneious software suite v9.1.8. Gaps were removed from the alignment and a Maximum Parsimony tree was built using MEGA version
X.Phylogenetic trees were visualized and manipulated in FIGTREE v1.4.1 (http://tree.bio.ed.ac.uk/software/figtree/).
1
Content courtesy of Springer Nature, terms of use apply. Rights reserved
For manuscripts utilizing custom algorithms or software that are central to the research but not yet described in published literature, software must be made available to editors and
reviewers. We strongly encourage code deposition in a community repository (e.g. GitHub). See the Nature Research guidelines for submitting code & software for further information.
Data
Policy information about availability of data
All manuscripts must include a data availability statement. This statement should provide the following information, where applicable:
- Accession codes, unique identifiers, or web links for publicly available datasets
- A list of figures that have associated raw data
- A description of any restrictions on data availability
Sequencing data can be found on the European Nucleotide Archive: [to be added upon upload]
nature research | reporting summary
Graphs
All graphs were produced and edited in R (https://cran.r-project.org/) and Adobe Photoshop
Field-specific reporting
Please select the one below that is the best fit for your research. If you are not sure, read the appropriate sections before making your selection.
Life sciences
Behavioural & social sciences
Ecological, evolutionary & environmental sciences
For a reference copy of the document with all sections, see nature.com/documents/nr-reporting-summary-flat.pdf
Behavioural & social sciences study design
All studies must disclose on these points even when the disclosure is negative.
Study description
This study is a mixed-methods, interdisciplinary analysis of archaeological materials from the site of Obishir V, in Kyrgyzstan. All
available archaeological materials (as of 2017) were selected for the analysis, of which the best preserved specimens (n=6) were
selected for genomic sequencing.
Research sample
All available skeletal remains from 2017 and early excavations were chosen for study. For ZooMS analysis, all materials from one
square of the excavation (R8) were selected for study.
Sampling strategy
All available materials from the archaeological assemblage were chosen for study.
Data collection
Researchers were blind to any hypothesis during data collection, which did not involve human participants.
Excavation
Excavations at Obishir V site carried out between 2015-2019 by a joint Russian-Kyrgyzstan archaeological expedition. In 2015–2016,
researchers working at the site excavated an area of 8 m2 adjoining the 1968–1969 excavation area (Supplementary Appendix A).
Each artifact (lithic and faunal remains) greater than 1 cm in size was individually piece-plotted with 1 mm accuracy using a Leica
Total Station TS02 plus in tandem with the Trimble software EDM Mobile, and documented along with contextual information (
orientation, dip, inclination, find category, stratigraphic level, etc.). Fragments smaller than this were collected in bulk for each 0.25
m2 area. Features (natural or anthropogenic) were recorded with the total station using outlines, surface points and breaklines.
Large surfaces, such as walls of old excavation pits, sections and archaeological horizons with numerous artefacts were also
documented using photogrammetry and 3D-scanning. Digital records were supplemented by field drawings and notes. All sediment
obtained during the excavation was wet-screened with mesh of a size of 2.0 mm, and finds were documented according to excavated
square, layer and depth. Stratigraphic units were defined during the excavation based on grain size composition, consistency, color,
presence of erosional surfaces and accumulations of limestone clasts.
Zooarchaeology and Zooarchaeology by Mass Spectrometry
For each specimen, we conducted taphonomic analysis and morphology-based species identifications using comparative faunal
collections held by the Russian Academy of Sciences-Siberia in their field station in Aidarkyen, Kyrgyzstan, and attempted refits for all
specimens to control for issues of fragmentation. We recorded taphonomic indicators such as rodent and carnivore damage, root
etching, and weathering (70), along with evidence for anthropogenic modification such as spiral fracturing, cut marks, and burning
via visual inspection (Supplementary Appendix C). In addition, we used collagen fingerprinting (or Zooarchaeology by Mass
Spectrometry, also known as ZooMS) to taxonomically identify a subset of this material, all derived from 2017 excavations. We
April 2020
Thermoluminescence dating
During the 2017 field season, two sediment samples were collected from the southern portion of the western excavation profile of
layer 5 at a depth of 295 and 335 cm below the surface. We measured deposit moisture in each sample, and after drying determined
dose rate (DR) using a MAZAR gamma spectrometer. Concentrations of 226Ra, 228Th, 40K in each sample were obtained from
twenty measurements lasting 2000 s each. We established equivalent dose (ED) on the 63-80 mm polymineral fraction, after 10%
HCl and 30% H2O2 washing and UV optical bleaching. The samples were irradiated with 20 Gy, 30 Gy, 40 Gy, 50 Gy and 100 Gy, doses
from 60Co gamma source. Before measurement, we heated the samples at 140 degrees for 3 hours. A sample pre-treated in this way
was used to determine the equivalent dose (ED) by the TL multiple-aliquot regenerative technique (67), according to the description
published by Fedorowicz et al. (68). Curve registration was performed on RA'94 (Mikrolab) thermoluminescence reader, coupled with
EMI 9789 QA photomultiplier. Finally, we calculated TL age according to Frechen (69).
2
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Cementum analysis
First, all analyzed teeth were 3D scanned using a MicroCT scanner, and washed in a 90% ethanol solution. A cast of the occlusal
surface was retained for later micro-wear analyses. Thin sections were prepared following recommendations by Rendu (73). Teeth
were embedded in epoxy within a vacuum chamber (at 0.2 bars pressure). They were then cut at a thickness of 250μm using an
Isomet5000 automatic saw, and ground to a thickness of 100μm with a PressiCube grinding machine. For each tooth, four
independent thin sections were prepared and analyzed independently. Observations were conducted under polarized light and
polarized/analyzed light both with and without the insertion of a lambda plate (74), using a Leica 2500P microscope equipped with a
Leica DC 120 camera. P
DNA Extraction
DNA was extracted from 6 teeth using silica-based purification in a dedicated laboratory for ancient DNA at PACEA (UMR 5199,
University of Bordeaux). We incubated between 100-250 mg of bone powder in a 1 mL extraction buffer ( 0.5 M EDTA, 0.25 M
Na2HPO4, 50 μg/mL proteinase K) for 48-72 h at 37°C. DNA was purified from the collected supernatant using a method adapted
from the QIAquick MinElute purification kit protocol (Qiagen).
nature research | reporting summary
analyzed all bone specimens from Square R8, Layers 2.2, 3, and 4, as well as all bone specimens from Square R8, Layer 1.3 using
ZooMS. To do so, we demineralized 10 mg of bone using 50mM ammonium bicarbonate buffer (Sigma-Aldrich) produced in UHQ
water and pH adjusted to 8.0 using Ammonium Hydroxide (1M), following the protocol outlined by (71). The extracted collagen was
digested into component peptides using trypsin (Pierce), and was purified using Pierce C18 tips (Thermo Scientific), eluting in 50%
acetonitrile (Sigma-Aldrich) and 0.1% TFA (Sigma-Aldrich). Samples were then spotted on Bruker AnchorChip with Bruker Peptide
Calibration Standard in calibration spots directly neighbouring the samples. Mass spectrometric analysis was conducted using a
Bruker Autoflex Speed LRF MALDI-TOF in the dedicated ZooMS Laboratory at the Max Planck Institute for the Science of Human
History in Jena, Germany.
PCR amplification
Purified DNA was amplified by qPCR using a LightCycler® 96 Instrument (Roche Life Science). Primer pairs were designed to amplify 2
regions of the mitochondrial DNA, a 130 bp region from cytB and 170bp from the hypervariable region (Supplementary Appendix F:
list of primers used and product sizes). We obtained each PCR product at least twice, conducting standard negative control to ensure
authenticity of the results. We used 1 μL of extract in 10 μL reaction with 1X FastStart Essential DNA Green Master and 1μM of each
primer. Products were sequenced by capillary electrophoresis at Genewiz (Leipzig, Germany).
Library construction and sequencing
We used DNA extracts to build double-stranded Illumina libraries following the protocol described in Gorge et al (77). We sequenced
whole genome shotgun libraries for ~ 500 000 reads on an Illumina NextSeq 500 at the Institut de Recherche Biomédicale des
Armées in Brétigny-sur-Orge, France, to assess DNA preservation. DNA extracts of five samples produced in Bordeaux were sent to
the dedicated clean room facilities of the Max Planck Institute for the Science of Human History. There, 20 ul of the extract were
used to build double stranded UDG-half libraries (78). The resulting genetic libraries were double indexed and further amplified (79)
Finally, all five samples were shotgun sequenced for ~20M reads on an Illumina HiSeq run with a 75-cycles single-end configuration.
Timing
Archaeological research was conducted in the summer of 2017, with ZooMS conducted in Fall 2017, and radiocarbon dating and DNA
sequencing from early 2018 through spring of 2020.
Data exclusions
No data were excluded from the analyses.
Non-participation
No human participants were used in the study
Randomization
No human participants were used in the study, and no data were allocated into groups
Reporting for specific materials, systems and methods
We require information from authors about some types of materials, experimental systems and methods used in many studies. Here, indicate whether each material,
system or method listed is relevant to your study. If you are not sure if a list item applies to your research, read the appropriate section before selecting a response.
Materials & experimental systems
Methods
n/a Involved in the study
n/a Involved in the study
Antibodies
ChIP-seq
Eukaryotic cell lines
Flow cytometry
Palaeontology and archaeology
MRI-based neuroimaging
Animals and other organisms
Human research participants
Dual use research of concern
April 2020
Clinical data
Palaeontology and Archaeology
Specimen provenance
Specimens all come from the archaeological site of Obishir V in Aidaryken, Kyrgyzstan, and research was conducted under Permit
#0040/02-05 issued by the The Field Committee of the Institute of History and Cultural Heritage of the National Academy of Sciences,
Kyrgyz Republic on 19 April, 2017. Specific provenance for each object is provided in the supplementary material.
3
Content courtesy of Springer Nature, terms of use apply. Rights reserved
All specimens are stored in the collections at the Institute for Archaeology and Ethnography, Siberian Branch of the Russian Academy
of Sciences, Novosibirsk, Russian Federation
Dating methods
Nine radiocarbon samples on identified and unidentified animal bone and tooth samples (n = 7) and charcoal (n = 2) were selected
from various depths across the stratigraphic unit of layers 2-4, and submitted for radiocarbon dating at the Oxford Radiocarbon
Accelerator Laboratory, the Center for Isotope Research at the University of Groningen, and the Golden Valley Laboratory at
Novosibirsk. A portion of tooth 20-1 was submitted to the University of Arizona Accelerator Mass Spectrometry Laboratory. Collagen
was obtained from 20-1 dentin using Acid-Base-Acid (ABA) pretreatment, gelatinization, 0.45 micron filtration, and ultrafiltration.
Quality control parameters: collagen yield (7.1%); carbon yield (42.9%); CN ratio (3.2); and d13C (19.7 +/- 0.1 per mil), indicated good
preservation (64). All dates were calibrated in OxCal using the INTCAL13 calibration curve (65). We performed Bayesian stratigraphic
phase modeling (using ordered phases and combining Layers 2-4) in the software OxCal with a uniform prior using all available
radiocarbon dates from Obishir V, and two thermoluminescence dates from Layer 5 (using OxCal’s C_date function), following
methods outlined by Ramsey (66). We also modeled the cultural occupation of Layer 2 separately, using only bone and tooth samples
using a uniform prior. The models provided good agreement, and repeating the analysis using a general outlier model did not
significantly alter estimated parameters. OxCal code used in the analysis, along with human tooth dating methods are outlined in
Supplementary Appendix B.
Tick this box to confirm that the raw and calibrated dates are available in the paper or in Supplementary Information.
Ethics oversight
nature research | reporting summary
Specimen deposition
All research was conducted in concurrence with ethical guidelines of the Max Planck Institute for the Science of Human History, Jena,
Germany.
Note that full information on the approval of the study protocol must also be provided in the manuscript.
April 2020
4
Content courtesy of Springer Nature, terms of use apply. Rights reserved
Terms and Conditions
Springer Nature journal content, brought to you courtesy of Springer Nature Customer Service Center GmbH (“Springer Nature”).
Springer Nature supports a reasonable amount of sharing of research papers by authors, subscribers and authorised users (“Users”), for smallscale personal, non-commercial use provided that all copyright, trade and service marks and other proprietary notices are maintained. By
accessing, sharing, receiving or otherwise using the Springer Nature journal content you agree to these terms of use (“Terms”). For these
purposes, Springer Nature considers academic use (by researchers and students) to be non-commercial.
These Terms are supplementary and will apply in addition to any applicable website terms and conditions, a relevant site licence or a personal
subscription. These Terms will prevail over any conflict or ambiguity with regards to the relevant terms, a site licence or a personal subscription
(to the extent of the conflict or ambiguity only). For Creative Commons-licensed articles, the terms of the Creative Commons license used will
apply.
We collect and use personal data to provide access to the Springer Nature journal content. We may also use these personal data internally within
ResearchGate and Springer Nature and as agreed share it, in an anonymised way, for purposes of tracking, analysis and reporting. We will not
otherwise disclose your personal data outside the ResearchGate or the Springer Nature group of companies unless we have your permission as
detailed in the Privacy Policy.
While Users may use the Springer Nature journal content for small scale, personal non-commercial use, it is important to note that Users may
not:
1. use such content for the purpose of providing other users with access on a regular or large scale basis or as a means to circumvent access
control;
2. use such content where to do so would be considered a criminal or statutory offence in any jurisdiction, or gives rise to civil liability, or is
otherwise unlawful;
3. falsely or misleadingly imply or suggest endorsement, approval , sponsorship, or association unless explicitly agreed to by Springer Nature in
writing;
4. use bots or other automated methods to access the content or redirect messages
5. override any security feature or exclusionary protocol; or
6. share the content in order to create substitute for Springer Nature products or services or a systematic database of Springer Nature journal
content.
In line with the restriction against commercial use, Springer Nature does not permit the creation of a product or service that creates revenue,
royalties, rent or income from our content or its inclusion as part of a paid for service or for other commercial gain. Springer Nature journal
content cannot be used for inter-library loans and librarians may not upload Springer Nature journal content on a large scale into their, or any
other, institutional repository.
These terms of use are reviewed regularly and may be amended at any time. Springer Nature is not obligated to publish any information or
content on this website and may remove it or features or functionality at our sole discretion, at any time with or without notice. Springer Nature
may revoke this licence to you at any time and remove access to any copies of the Springer Nature journal content which have been saved.
To the fullest extent permitted by law, Springer Nature makes no warranties, representations or guarantees to Users, either express or implied
with respect to the Springer nature journal content and all parties disclaim and waive any implied warranties or warranties imposed by law,
including merchantability or fitness for any particular purpose.
Please note that these rights do not automatically extend to content, data or other material published by Springer Nature that may be licensed
from third parties.
If you would like to use or distribute our Springer Nature journal content to a wider audience or on a regular basis or in any other manner not
expressly permitted by these Terms, please contact Springer Nature at
onlineservice@springernature.com