ArticlePDF Available

Impact of climate change on the transition of Neanderthals to modern humans in Europe

Authors:
  • Emil Racovita Institute of Speleology
  • University of South Florida / Babes-Bolyai University
  • Formerly - at International Atomic Energy Agency (IAEA)

Abstract and Figures

Significance A causality between millennial-scale climate cycles and the replacement of Neanderthals by modern humans in Europe has tentatively been suggested. However, that replacement was diachronous and occurred over several such cycles. A poorly constrained continental paleoclimate framework has hindered identification of any inherent causality. Speleothems from the Carpathians reveal that, between 44,000 and 40,000 years ago, a sequence of stadials with severely cold and arid conditions caused successive regional Neanderthal depopulation intervals across Europe and facilitated staggered repopulation by modern humans. Repetitive depopulation–repopulation cycles may have facilitated multiple genetic turnover in Europe between 44,000 and 34,000 years ago.
Content may be subject to copyright.
Impact of climate change on the transition of
Neanderthals to modern humans in Europe
Michael Staubwasser
a,1
, Virgil Dr ˘
agus
¸in
b
, Bogdan P. Onac
c,d
, Sergey Assonov
a,e
, Vasile Ersek
f
, Dirk L. Hoffmann
g
,
and Daniel Veres
d
a
Institute of Geologie and Mineralogy, University of Cologne, 50674 Cologne, Germany;
b
Emil Racovit
¸˘
a Institute of Speleology, Romanian Academy, 010986
Bucharest, Romania;
c
School of Geosciences, University of South Florida, Tampa, FL 33620;
d
Emil Racovit
¸˘
a Institute of Speleology, Romanian Academy,
400006 Cluj-Napoca, Romania;
e
Terrestrial Environment Laboratory, Environmental Laboratories, Department of Nuclear Applications, International Atomic
Energy Agency, 1400 Vienna, Austria;
f
Department of Geography and Environmental Sciences, Northumbria University, Newcastle upon Tyne NE1 8ST,
United Kingdom; and
g
Department of Human Evolution, Max Planck Institute for Evolutionary Anthropology, 04103 Leipzig, Germany
Edited by Richard G. Klein, Stanford University, Stanford, CA, and approved July 30, 2018 (received for review May 19, 2018)
Two speleothem stable isotope records from East-Central Europe
demonstrate that Greenland Stadial 12 (GS12) and GS10at 44.3
43.3 and 40.840.2 kawere prominent intervals of cold and arid
conditions. GS12, GS11, and GS10 are coeval with a regional pat-
tern of culturally (near-)sterile layers within Europes diachronous
archeologic transition from Neanderthals to modern human Auri-
gnacian. Sterile layers coeval with GS12 precede the Aurignacian
throughout the middle and upper Danube region. In some records
from the northern Iberian Peninsula, such layers are coeval with
GS11 and separate the Châtelperronian from the Aurignacian.
Sterile layers preceding the Aurignacian in the remaining Châtel-
perronian domain are coeval with GS10 and the previously
reported 40.0- to 40.8-ka cal BP [calendar years before present
(1950)] time range of Neanderthalsdisappearance from most of
Europe. This suggests that ecologic stress during stadial expansion
of steppe landscape caused a diachronous pattern of depopulation
of Neanderthals, which facilitated repopulation by modern humans
who appear to have been better adapted to this environment. Con-
secutive depopulationrepopulation cycles during severe stadials of
the middle pleniglacial may principally explain the repeated replace-
ment of Europes population and its genetic composition.
Central Europe
|
speleothems
|
millennial-scale climate cycles
|
stable isotopes
|
MiddleUpper Paleolithic transition
The replacement of Neanderthals by modern humans is
recorded across Europe in a diachronous and culturally complex
succession of distinct stone tool assemblages from the Middle
Upper Paleolithic transition (MUPT) roughly between 48 and 36 ka
cal BP [calendar years before present (1950)] (1, 2). The succession
often includes a regionally distinct transitionalassemblage of local
origin, or an intrusive Initial Upper Paleolithic assemblage between
the Neanderthal Mousterian and modern human Aurignacian. The
oldest anatomically modern human remains from Europefound in
East-Central Europe and radiocarbon dated to 40.638.6 ka cal BP
(68% probability) toward the time of Neanderthalsdisappearance
from most of continental Europecarry genetic evidence for species
interbreeding four to six generations earlier (36). This individual,
however, represents a population that did not contribute to the
genome of modern humans present in glacial Europe after the
MUPT (7), and the archeologic record provides no site with in-
dication of local coexistence. Within a few millennia after the
MUPT, at least two other genetically distinct modern human pop-
ulations came to subsequently dominate Middle Pleniglacial Europe.
During the entire interval, northern hemispheric climate went
through several millennial-scale DansgaardOeschger (DO) cold
cycles (8, 9). A causality between climate change, the archeologic
succession, and modern humansgenetic makeup has been tenta-
tively suggested but not demonstrated (1, 2, 7). Below, we present
the climatic history of continental Europe during the MUPT and
derive the impact of climate change on MUPT demography, which
may have led to the apparent repetitive genome turnover reported
for Europes human population during the middle pleniglacial.
The MUPT spans five DO cycles approximately between
Greenland Interstadial 12 (GI12) and Greenland Stadial 8 (GS8)
(9) for which climate change over continental Europe is poorly
constrained. The paleoclimatic and environmental context is
known with sufficient resolution and age-control only along the
continents western and southern fringe. In the Aegean and
Black Sea region, records of sea surface temperature (10),
coastal ice-rafted detritus (IRDc) (11), pollen assemblages (12
15), and stable isotopes in speleothems (16) suggest a DO-type
response without the clear prominence of ice-rafting intervals
(Heinrich stadials) seen in the Atlantic domain (17). Forest was
generally more abundant in Europe during interstadials, while
steppe landscape advanced during stadials (12). A taiga and
tundra shrub/forest landscape covered the eastern European
plains, with some loess deposition east of the Carpathians (12,
18, 19). The middle and lower Danube Plain was a steppe
landscape with continuous loess deposition (20). A temperate
open forest in the mountains of the southern Balkan passed into
a xerophytic steppe toward the Aegean Sea (12, 13, 15). Boreal
forest with birch and pine trees was present at 50° N in Western
Europe (Eifel maar lakes) but began to degrade after GI12,
44.5 ka cal BP (21). Two sparsely dated records from the
Western Carpathians (Safarka, Jablunka) suggest a dense taiga
forest landscape (14). For the upper Danube Plain, pollen
(Füramoos) and loess/paleosol profiles (Willendorf II, Nussloch)
Significance
A causality between millennial-scale climate cycles and the
replacement of Neanderthals by modern humans in Europe has
tentatively been suggested. However, that replacement was
diachronous and occurred over several such cycles. A poorly
constrained continental paleoclimate framework has hindered
identification of any inherent causality. Speleothems from the
Carpathians reveal that, between 44,000 and 40,000 years ago,
a sequence of stadials with severely cold and arid conditions
caused successive regional Neanderthal depopulation intervals
across Europe and facilitated staggered repopulation by mod-
ern humans. Repetitive depopulationrepopulation cycles may
have facilitated multiple genetic turnover in Europe between
44,000 and 34,000 years ago.
Author contributions: M.S., V.D., and B.P.O. designed research; M.S. and V.D. performed
research; V.D., B.P.O., S.A., and D.L.H. analyzed data; and M.S., V.D., B.P.O., V.E., and D.V.
wrote the paper.
The authors declare no conflict of interest.
This article is a PNAS Direct Submission.
This open access article is distributed under Creative Commons Attribution-NonCommercial-
NoDeriv atives L icense 4.0 (CC BY-N C-ND).
1
To whom correspondence should be addressed. Email: m.staubwasser@uni-koeln.de.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.
1073/pnas.1808647115/-/DCSupplemental.
www.pnas.org/cgi/doi/10.1073/pnas.1808647115 PNAS Latest Articles
|
1of6
EARTH, ATMOSPHERIC,
AND PLANETARY SCIENCES
ANTHROPOLOGY
suggest a cold steppe environment with few conifers during in-
terstadials, and a tundra landscape with cryosoil formation dur-
ing stadials (12, 22, 23). During GIs, a temperate forest-steppe
prevailed west (marine records) and south of the Alps (Monticchio,
Castiglione, and Lagaccione) (12). Dust deposition in the Eifel
maar lakes and speleothem carbon isotopes from Villars Cave
suggest increasing aridity in Western Europe across the MUPT
(24, 25).
The MUPT Speleothem Paleoclimate Record of East-Central
Europe
Here, we present two speleothem records from East-Central
Europe (Romania). Stalagmite POM1 is from Ascuns˘
a Cave
(AC), South Carpathians, 50 km north of the Danube Valley at
1,050-m altitude (SI Appendix, Fig. S1). Stalagmite 1152 is from
T˘
aus
¸oare Cave (TC), East Carpathians, at 950-m altitude, 2.5°
north of the AC site. We dated the speleothems by U-Th and
measured stable carbon and oxygen isotopes continuously at
decadal resolution (SI Appendix, Fig. S2 and Table S1).
The AC speleothem δ
13
C record reproduces the DO pacing
of the Greenland ice core record during the MUPT (Fig. 1). The
TC δ
18
O record reproduces some aspects of the Greenland record.
The climatic meaning of the AC δ
13
C data may be constrained
within context of other records from the region, particularly the
Black Sea (Fig. 1). The coherency of the AC δ
13
C record with
southern Black Sea IRDc (11) implies that speleothem δ
13
C
responded to changes in the length of the sea ice and winter frost
season. Chemically, speleothem δ
13
C depends on the proportion of
low δ
13
C aqueous CO
2
derived from microbial respiration of soil
organic matter above the cave (26). Extended frost and shorter
plant growth seasons reduces the supply of fresh soil organic matter
(SOM) and increases the proportion of CO
2
from old SOM in drip
water, which has up to 8higher δ
13
C than fresh SOM under
comparable conditions (27). Other potential controls are insuffi-
cient to explain the 6amplitude of DO cycles in the speleothem.
For example, the effect of a seasonally variable CO
2
degassing rate
from drip water may add a kinetic fractionation effect on the δ
13
C
of precipitating speleothem calcite (28). However, recent moni-
toring showed that interannual variability in δ
13
C of precipitating
calciteinACislessthan1.5(29). Similarity between the AC δ
13
C
and δ
18
O records is limited (SI Appendix,Fig.S3)another
qualitative argument against major kinetic isotope fractionation.
Also, a compositional change in SOM δ
13
C can be ruled out as
variability in Central European paleosol profiles does not exceed
1during the MUPT (30). Finally, variable moisture availability
may have influenced soil formation above the cave and speleothem
δ
13
C. The Danube loess record suggests such an interval of en-
hanced moisture availability after GI8 (20). The AC-speleothem
δ
18
O record does indeed suggest some change to that effect be-
tween GS8 and GI7, but the overall pacing across the MUPT re-
sembles that of southeastern Mediterranean records and shows
little coherency with DO cycles (see below). In general, soil for-
mation intervals recorded by low values in AC-δ
13
C are coeval with
the range of ages obtained for paleosols in the upper Danube
Willendorf II profile (Fig. 2). This suggests that the entire Danube
region was inside the same climate zone and responded coherently
to temperature change during DO cycles of the MUPT.
Little variance is observed in the East Carpathian TC δ
13
C
record (SI Appendix, Fig. S3). Here, pyrite weathering and en-
hanced limestone dissolution in the host rock obscures any po-
tential influence on δ
13
C from SOM (SI Appendix). That record
is not considered in this study.
In general, speleothem δ
18
O reflects a combination of calcite
precipitation temperaturethe annual average temperature in-
side the caveand factors controlling δ
18
O in drip water. Within
a few days, water and dissolved CO
2
equilibrate isotopically in
the aquifer above the cave, and drip water composition reflects
regional hydroclimatology, that is, moisture source, rain-out
history, local rain-out temperature, and the annual distribution
of rainfall (26, 31). The AC δ
18
O signals overall amplitude is
4. Such a large change is beyond reasonable temperature
variability (28) and suggests a dominant influence of changes in
hydrology on AC δ
18
O. The dissimilarity of AC δ
18
O to the co-
herent δ
13
C and Greenland temperature records (SI Appendix,
Fig. S3) renders a temperature control of the AC δ
18
O record
unlikely. Hydroclimatic dominance on speleothem δ
18
O is typical
for the entire Eastern Mediterranean domains Holocene record
including at AC, which reflects significant variability in rainfall
seasonality and moisture source proportions (31). The similar
pacing between AC δ
18
O and southeastern Mediterranean δ
18
O
record (SI Appendix, Fig. S4) may indicate that the underlying
synoptic-scale mechanism was also active during MIS3.
The TC speleothem δ
18
O record has an amplitude of 1
and a different pacing compared with AC δ
18
O(SI Appendix, Fig.
S3). This suggests that the East Carpathians were in a different
hydroclimatic regime during the MUPT. What defines the East
Carpathian regime may be investigated by exploring the partial
coherency between TC δ
18
O and the other paleoclimate record.
After 60 ka, the δ
18
O record shows low-amplitude variability
with a few pronounced minima of 0.5that are coeval with
some of the Greenland stadials (Fig. 1). These include GS3-H2,
GS5-H3, and GS13-H5. Other minima just after 34 ka and at
A
B
C
D
E
Fig. 1. (A) Greenland: North Greenland Ice Core Project (NGRIP) tempera-
ture (8). (B) South Carpathians: δ
13
C of stalagmite POM1 from AC. (C)
Southern Black Sea: coastal IRD abundance in core M72-5 (11). (D)East
Carpathians: δ
18
O of stalagmite 1152, TC. (E) Northern Black Sea: TEX
86
summer sea surface temperatures (33). The gray bars indicate Heinrich sta-
dials. A map with locations of records is available in SI Appendix, Fig. S1.
2of6
|
www.pnas.org/cgi/doi/10.1073/pnas.1808647115 Staubwasser et al.
41 ka correspond to GS7 and GS10 (Fig. 1). The latter lasted
from 40.66 ±0.47 ka
U-Th
,P=95%)directly datedto an in-
terpolated end date of 39.70 ±0.43 ka (32). Although there
may be a few centuries of overlap with GS9-H4, the measured
age interval of 40.739.8 ka
U-Th
,±0.47 to ±0.3 uncertainty,
matches GS10 (40.8040.16 ka
GICC05
) better than GS9-H4 (from
39.90 to 38.22 ka
GICC05
) (9). A poorly resolved relative δ
18
O
maximum at 39.6 ±0.4 ka
U-Th
and a subsequent broad relative
δ
18
O minimum in the TC record appear to be coeval with GI9
and GS9-H4, which apparently had a lesser impact on Central
European climate. The TC δ
18
O record thus responded to stadial
cooling in the Atlantic but with a regionally controlled ampli-
tude. Similarity with the northern Black Sea paleotemperature
archive suggests this response is temperature driven (Fig. 1). For
two reasons, TC δ
18
O likely records the local summer temper-
ature signal. First, the northern Black Sea record shows that
summer sea surface temperature between 40 and 20 ka cal BP
dropped by 2 °C during stadials from a MIS3 average of 4C
(33). Assuming published global empirical temperature rela-
tionships for cave calciteΔδ
18
O
c
/ΔT=0.18 (28)and for
rainwater in midlatitudeswhere Δδ
18
O
w
/ΔT=0.58 (34)the
0.40.6δ
18
O amplitude in the TC speleothem would corre-
spond to a comparable 1.01.5 °C temperature change. Second,
palynologic, geomorphologic, and geochemical studies support
this interpretation as follows. During the last glacial, the low-
lands north of the East Carpathians were situated inside the
tundra biome, where field evidence of ice wedges indicates
permafrost conditions (3537). Deep winter frost or even dis-
continuous permafrost is a likely scenario for the >1,000-m al-
titude of the ground above TC in the East Carpathians. At 2040 ka,
the average ground temperature of the northern Black Sea
drainage basincomprising the Danube, Dniestr, Dniepr, and
Don Lowlandswas 4 °C (38). Common adiabatic gradients of
610 °C per km of altitude imply that annual average ground
temperatures above TC were lower than 2 °C. This suggests an
alpine near-permafrost environment. Permafrost conditions may
have been continuous during extreme stadials, thus preventing
groundwater recharge to the cave and interrupting speleothem
growth. During GS12the coldest of all stadials in the Green-
land recordthe TC speleothem shows a hiatus coeval with the
most pronounced δ
13
C maximum in the AC speleothem further
south. In the Willendorf II loess profile, this interval correlates
with tundra gley horizon C9 with evidence of permafrost (22)
(Fig. 2). C9 is constrained in time by paleosols C8-3 above and
D1 underneath with dating ranges similar to GI11 and GI12,
respectively. The Willendorf II profile is located at the same
latitude as the TC speleothem, but only at 230-m altitude.
Continuous permafrost conditions at 950-m altitude in the East
Carpathians were thus likely during GS12. Seasonally permeable
frost or discontinuous permafrost at other times would restrict
speleothem growth to the summer season. Unlike the hydro-
logically dominated AC δ
18
O record, TC δ
18
O-only records re-
gional summer temperature change.
Paleoclimatic Context of the Middle Pleniglacial in Europe
A reduction of forest and expansion of steppe biomes occurred
during all stadials (1215, 21), but climate records between the
Atlantic and the Black Sea show a spatially heterogeneous response
to DO cycles (Fig. 2). GS13-H5 is apparent in the TC speleothem
δ
18
O record but unlike in the Atlantic domain, its amplitude is
small compared with subsequent stadials. Southern Black Sea
IRDc data (Fig. 1) indicate less sea ice than during subsequent
stadials despite apparent colder annual sea surface temperature
(Fig. 2). This suggests a different seasonality with less severe win-
ters. Thus, extreme cold was unlikely in East-Central Europe. The
extreme aridity apparent in the Aegean region (13) may not have
been as severe in the Balkan and Black Sea region (15, 16). GS12,
from 43.444.3 ka
GICC05
or 43.344.0 ka
U-Th
(AC), is a very
A
B
C
D
E
F
G
H
Fig. 2. (A) Greenland: NGRIP temperature (8). (B) Southern Black Sea: TEX
86
sea temperature core M72-5 (10). (C) South Carpathians: AC stalagmite
POM1 δ
13
C. (D) Upper Danube: Willendorf loess/paleosol profile, paleosol
ages, age probability density functions (age-pdf), and stratigraphically con-
strained loess and gley horizons C7-2, C7-3, and C9 (22). (E) East Carpathians:
TC stalagmite 1152 δ
18
O. (F) Eifel maar lakes, ELSA dust stack. (G) Western
Massive Central: Villars Cave stalagmites δ
13
C(25).(H) Iberian Atlantic margin:
TEX
86
sea surface temperature core MD95-2042 (17). Numbers indicate
Greenland stadials at the end of the respective DO cycle.A map with locations
of records is available in SI Appendix,Fig.S1.
Staubwasser et al. PNAS Latest Articles
|
3of6
EARTH, ATMOSPHERIC,
AND PLANETARY SCIENCES
ANTHROPOLOGY
prominent stadial in Greenland and Central Europe (Fig. 2), where
permafrost changed the upper Danube lowland cold steppe into a
tundra (see above). Malacological data from Willendorf II and the
Eifel maar dust record suggests dryer conditions also in Western
and Central Europe (22, 24) (Fig. 2). A cooling step likely occurred
over Western Europe, supported by evidence of deep frost in the
loess/paleosol record of northern France at 4344 ka cal BP (39).
However, this stadial is less prominent in the Atlantic record (17).
GS11 is present in all continental records, but not as pronounced as
GS12. A common cooling trend superimposed over DO cycles 129
is apparent in all records (Fig. 2). GS10from 40.140.9 ka
GICC05
and 39.740.7 ka
U-Th
(TC)is a very prominent stadial over
continental Europe, coeval with another cooling step in Western
Europe (Fig. 2), but also a reduction in moisture and reduced
speleothem growth (25). The first occurrence of loess in the
Willendorf II profile (C7-2) around that time (22) and another
strong dust peak in the Eifel maar record suggest stronger aridity
than during GS12. Although present in all continental records, this
stadial is unremarkable in the Atlantic and in the Black Sea (Fig.
2). However, annual sea surface temperature and winter coastal
ice abundance in the Black Sea show conflicting results for GS10,
which could again indicate a seasonality change. GS9-H4 is a long-
lasting stadial that was less prominent in the Central European
record but coeval with significant cooling and aridity in Western
Europe (Fig. 2). Over the Atlantic and the Black Sea, GS9-H4 is a
much more significant stadial. Subsequent GI8 is as warm as GI12
in Central Europe and the Black Sea region, with soil formation in
the Danube region and East Carpathians (19, 20, 22), but colder in
Western Europe and the Atlantic. Dust deposition in the Eifel
maar records was also high during GI8, unlike GI13. The causes to
this regionally heterogeneous response to DO climate cycles re-
main uncertain, but a variable influence of the Siberian high has
been suggested (20).
Following the MUPT, GS8 from 36.6 to 35.5 ka
GICC05
and GI7
from 35.5 to 34.7 ka
GICC05
are difficult to separate from each
other in the speleothem records. The Black Sea record shows
severe cooling during GS8 (Fig. 2). The AC-δ
18
O record suggests
a significant fluctuation of the regional hydrologic condition
during GS8 and GI7 in agreement with extended soil formation
in the Danube Valley and the East Carpathians (19, 22). Higher
moisture availability may have masked the temperature signal in
AC-δ
13
C. After GI7, a long-term cooling trend is superimposed
on DO variability throughout Europe (Fig. 2), in many places
coincident with a long drying trend (20, 25).
The Relationship Between Demography and Climate Change
During GS12 and GS10
GS12 and GS10 stand out as the most significant stadials in
Central and Western Europe during the MUPT with severe cold,
aridity, retreat of woodland, and expansion of steppe biomes.
The consequence for human populations present is reflected in
regional population decline (40) and in hunted game species,
which in Western Europe changed from bovine dominated to
reindeer dominated (41). Climatic and environmental con-
straints may not per se prove causality between the archeologic
succession, the rapid replacement of humansgenome, and cli-
mate change, respectively. However, they do allow testing such a
scenario for chronological feasibility: The biome changes during
GS12 and 10 likely forced the population in open woodland
habitats throughout Europe to adapt their subsistence strategy
or habitat track their preferred biome to survive. Where biomes
changed significantly, adaptation may not have been possible.
During GS12 and 10, the permafrost boundary encroached the
upper Danube region and East Carpathians. Both speleothems
presented here suggest a rapid onset of stadial cooling and aridity
in continental Europe within a few decades (Fig. 2). Depopulation
would have been the consequence of failure to adapt or migrate
away in time.
A number of chronologically well-constrained archeologic
records allow for testing directly the feasibility of a regional
depopulation scenario during severe stadials. Culturally (nearly)
sterile layers have been reporteddirectly
14
C dated or con-
strained by
14
C-dated cultural layers above and belowin some
regions and mark an archeological hiatus of several centuries
duration. While such layers in some places represent short-term
sedimentologic events without apparent discontinuity of habita-
tion (42), at many sites their time span is multicentennial. These
layers mostly relate chronologically to the time of GS12 and
GS10, and suggest regionally widespread depopulation for many
centuries between two subsequent cultures. (Fig. 3) (SI Appendix,
Table S2). During GS12, this is the case for sites from the upper
and middle Danube region. At Geissenklösterle, near-sterile
(geologic) layer 17 separates Mousterian and Aurignacian as-
semblages (43) and is directly dated to 44.442.7 and 43.5
42.0 ka cal BP by two samples, with a Bayesian model age of
42.942.1 ka cal BP (44). At Sesselfelsgrotte, sterile layer F is
directly constrained by Mousterian layer G1, 45.844.3 ka cal BP,
and the late Mousterian layer E3, 42.240.5 ka cal BP, both
recalibrated (4, 45). At Willendorf II, sterile layer C9 and sub-
sequent layer C8-3 containing Aurignacian artifacts are con-
strained between two dated paleosols D1 and C8-2 (Fig. 2) (22).
Mousterian Neanderthal presence in the middle Danube region
has been redated to until 45 ka cal BP, just before GS12 (46).
Two sites with Aurignacian finds, at Keilberg-Kirche (47), 43.6
41.7 ka cal BP recalibrated (4), and at Pes-kCave, 43.841.2 ka
cal BP (48), immediately postdate GS12. These fit the scenario
of a cultural hiatus but lack local constraint by older assemblages.
All of the above suggest widespread Mousterian Neanderthal
Fig. 3. The temporal pattern of Greenland stadials (red), prominent events
in the Central European speleothem record (blue), and culturally (near-)
sterile layers in archeologic records of Western and Central Europe (black).
For stadials and speleothems, the bar shows the duration. For archeologic
layers, the bar shows the68% age interval (ka, cal BP) defined by the youngest
date of the preceding layer and the oldest date of the succeeding layer. For
details on archeologic radiocarbon chronologies, see SI Appendix,TableS2.A
map with the locations of records is available in SI Appendix,Fig.S1.
4of6
|
www.pnas.org/cgi/doi/10.1073/pnas.1808647115 Staubwasser et al.
depopulation of the upper and middle Danube Valley and
except for the tributary Altmühl Valley (49)repopulation by
Aurignacian after GS12.
Circumstantial evidence points to a similar impact of GS12 in
northern Italy, where two sites, Riparo Mochi on the Ligurian
Mediterranean coastlayer Hand at Grotta di Fumane in the
Venetian Pre-Alpslayer A7contain sterile layers probably
coeval with GS12. However, reliable dating is only available for
the base of the subsequent archeologic layer (50, 51) and age
uncertainty is large (Fig. 3). A muted response to GS12 may be
inferred for the site of Grotte Mandrin, Rhone Valley, where
sterile layers separate strata C, B3, and B2 with post-Neronian-II
artifacts (50). Layers B3B1 of this latest regional Mousterian
expression were dated in the range between 42.7 and 45.0 ka BP
(6). These layers may indicate environmental change but do not
appear to reflect longer breaks in the sites occupation. No such
layers associated with GS12 are reported from archeologic sites
further west. This suggests that the environmental impact of
GS12 may have been less severe in Western Europe.
Sterile or near-sterile layers, or cultural hiatuses in the
archeologic succession also occur around the time of GS10 (Fig.
3). This time is coeval with Neanderthalsdisappearance from
Western Europe (6). Sites include the Grotte-du-Rennethe
low-density artifact layer VIII of intermittent Châtelperronian
occupation (52); Les Cottéssterile unit 5 (53); and Saint Césaire
layer Ejo inf (6). At Willendorf II, dated paleosols with lithic
artifacts in C8 and C7 broadly constrain the sterile gley and loess
horizons (C7-3 and C7-2) at that time (22) (Fig. 2). Subsequent
archeologic layers are generally Aurignacian and document
modern human expansion or reoccupation during GI9 and GI8
but under more arid conditions in Europe (Fig. 2).
A few archeologic profiles from around the Pyreneesat
Labeko Koba, Abric Romani, and possibly at Isturitz, but not at
LArbreda (5456)contain sterile layers coeval with GS11 (Fig.
3). These represent a regional cultural hiatus between the
Châtelperronian and the (Proto-)Aurignacian. However, apart
from a long-term cooling and drying trend over Western Europe,
an unambiguously dated climatic cold event cannot be detected
in the present speleothem records for that time (Fig. 2).
A general case for likely adaptation of modern humans in
response to climate change during the MUPT has already been
made (48, 57, 58). Increasing cold and aridity around the onset of
GS12 (Fig. 2) appears to mark transitions in the archeologic
sequence attributable to Neanderthals that suggest some adap-
tation as well. GS12 is coeval with the transition from Mouste-
rian to archaic Uluzzian, and then to evolved Uluzzian in
northern Italy (59), and from Mousterian to Châtelperronian in
Western Europe (53). However, late Neanderthals may have had
a less diverse diet than modern humans (60). In open grasslands,
Neanderthalsexclusive diet was meat from terrestrial animals,
whereas modern human Aurignacian also exploited plant and
aquatic foods (60, 61). The frequently observed cultural hiatuses,
however, do not suggest a direct competitive displacement of
Neanderthals, but rather a higher vulnerability to rapid envi-
ronmental change and ecologic stress in the open landscape
during cold and arid GS12GS10. While Neanderthals did not
survive GS12 in most of the Danube steppe and tundra, modern
humans may have been more capable to adapt and habitat track
the expanding steppe in Central Europe. GS10 likely caused a
repeat of that process in Western Europe.
A depopulation scenario was suggested for GS13-H5 as a
trigger for the first intrusion of modern humans into Europe,
represented by the Bohunician stone tool assemblage in Moravia
(Central Europe) and further east (13, 62). Because a potential
cultural hiatus is difficult to prove in the archeologic record of
Moravia at given accuracy and precision of applied radiometric
dating methods (63), this scenario may not currently be tested.
There is also not as strong a climatic evidence in support.
Cooling likely was less severe in East-Central Europe compared
with subsequent stadials (see above). Severe aridity was apparent
between the northern Aegean coastal region and the Levant (13,
64) but not over the Balkan (15), the Adriatic and the Black Sea
regions (16, 65), or Central-East Europe.
The Moravian record suggests contemporaneity of modern
humansthe Bohunicianand Neanderthalsthe Szeletian
at close geographic proximity during GI12 and perhaps later but
with some chronologic uncertainty (63, 66). During GI11-10, the
situation is comparable for the Aurignacian and Mousterian on
the upper Danubeat Keilberg-Kirche and Sesselfelsgrotte (see
above)and for Aurignacian and Szeletian on the middle
Danubeat Pes-kCave and Szeleta Cave (42, 67) (SI Appen-
dix, Fig. S1). Here, however, the interbreeding between species
only six generations before the lifetime of the Oase Cave modern
human specimen in that time range (5) principally confirms true
contemporaneity along the northern fringe of the steppe land-
scape in the Danube Valley. Nonetheless, in the subsequent 6
millennia after GS10, the modern human genome in Europe was
replaced twice with a different genetic lineage harboring a much
older Neanderthal ancestry (7). One modern European genetic
branch found inside Goyet Cave (Belgium), was attributed to the
late Aurignacian and dated to 35 ka cal BPcoeval with GI7.
This interval of modern human repopulation may have followed
the extended cold interval from GS10 to GS9-H4, over 4,000 y
long and only interrupted briefly by the 250-y-long GI9 (Fig. 2).
A different genetic branch then dominated Europe after 34 ka cal
BPcoeval to the end of the severe and 1,000-y-long GS7 (Fig.
2). This lineage was attributed to the Gravettian (7), who repo-
pulated Europe during yet another time of low population
density (68). The discussion above lays out a general scenario of
depopulationrepopulation cycles associated with steppe land-
scape expansion following extreme or long stadials. The com-
parable timing of stadials and population changes seen in the
archeologic and genetic record suggests that millennial-scale
climate cycles may have been the pacesetter for Europes de-
mographic history during the Middle Pleniglacial.
ACKNOWLEDGMENTS. This research was supported by Deutsche For-
schungsgemeinschaft Funding (SFB 806, TP B2) (to M.S.). V.D. acknowledges
support by theEuropean Social Fund, Sectoral Operational Programme Human
Resources Development, Contract POSDRU 6/1.5/S/3—“Doctoral Studies:
Through Science Towards Society,PCE-2016-0179 Grant CARPATHEMS, and
IFA-CEA C4-08 (FREem). Part of the isotopic analysis were funded by the Uni-
versity of South Florida via an internal grant (to B.P.O.).
1. Hoffecker JF (2009) Out of Africa: Modern human origins special feature:
The spread of modern humans in Europe. Proc Natl Acad Sci USA 106:16040
16045.
2. Hublin J-J (2015) The modern human colonization of western Eurasia: When and
where? Quat Sci Rev 118:194210.
3. Trinkaus E, et al. (2003) An early modern human from the Pes
¸tera cu Oase, Romania.
Proc Natl Acad Sci USA 100:1123111236.
4. Reimer PJ, et al. (2013) Intcal13 and Marine13 radiocarbon age calibration curves 0
50,000 years cal Bp. Radiocarbon 55:18691887.
5. Fu Q, et al. (2015) An early modern human from Romania with a recent Neanderthal
ancestor. Nature 524:216219.
6. Higham T, et al. (2014) The timing and spatiotemporal patterning of Neanderthal
disappearance. Nature 512:306309.
7. Fu Q, et al. (2016) The genetic history of ice age Europe. Nature 534:200205.
8. Kindler P, et al. (2014) Temperature reconstruction from 10 to 120 kyr b2k from the
NGRIP ice core. Clim Past 10:887902.
9. Rasmussen SO, et al. (2014) A stratigraphic framework for abrupt climatic changes
during the last glacial period based on three synchronized Greenland ice-core records:
Refining and extending the INTIMATE event stratigraphy. Quat Sci Rev 106:1428.
10. Wegwerth A, et al. (2015) Black Sea temperature response to glacial millennial-scale
climate variability: Black Sea temperature and DO cycles. Geophys Res Lett 42:
81478154.
11. Nowaczyk NR, et al. (2012) Dynamics of the Laschamp geomagnetic excursion from
Black Sea sediments. Earth Planet Sci Lett 351-352:5469.
12. Fletcher WJ, et al. (2010) Millennial-scale variability during the last glacial in vege-
tation records from Europe. Quat Sci Rev 29:28392864.
Staubwasser et al. PNAS Latest Articles
|
5of6
EARTH, ATMOSPHERIC,
AND PLANETARY SCIENCES
ANTHROPOLOGY
13. Müller UC, et al. (2011) The role of climate in the spread of modern humans into
Europe. Quat Sci Rev 30:273279.
14. Feurdean A, et al. (2014) Climate variability and associated vegetation response
throughout Central and Eastern Europe (CEE) between 60 and 8 ka. Quat Sci Rev 106:
206224.
15. Panagiotopoulos K, et al. (2014) Climate variability over the last 92 ka in SW Balkans
from analysis of sediments from Lake Prespa. Clim Past 10:643660.
16. Fleitmann D, et al. (2009) Timing and climatic impact of Greenland interstadials re-
corded in stalagmites from northern Turkey. Geophys Res Lett 36:L19707.
17. Darfeuil S, et al. (2017) Sea surface temperature reconstructions over the last 70kyr
off Portugal: Biomarker data and regional modeling. Paleoceanography 31:4065.
18. Bokhorst MP, et al. (2011) Atmospheric circulation patterns in central and eastern
Europe during the Weichselian Pleniglacial inferred from loess grain-size records.
Quat Int 234:6274.
19. Haesaerts P, et al. (2010) Charcoal and wood remains for radiocarbon dating upper
Pleistocene loess sequences in eastern Europe and central Siberia. Palaeogeogr
Palaeoclimatol Palaeoecol 291:106127.
20. Obreht I, et al. (2017) Shift of large-scale atmospheric systems over Europe during late
MIS 3 and implications for modern human dispersal. Sci Rep 7:5848.
21. Sirocko F, et al. (2016) The ELSA-vegetation-stack: Reconstruction of landscape evo-
lution zones (LEZ) from laminated Eifel maar sediments of the last 60,000 years.
Global Planet Change 142:108135.
22. Nigst PR, et al. (2014) Early modern human settlement of Europe north of the Alps
occurred 43,500 years ago in a cold steppe-type environment. Proc Natl Acad Sci USA
111:1439414399.
23. Moine O, et al. (2017) The impact of last glacial climate variability in west-European
loess revealed by radiocarbon dating of fossil earthworm granules. Proc Natl Acad Sci
USA 114:62096214.
24. Sirocko F, et al. (2013) Multi-proxy dating of Holocene maar lakes and Pleistocene dry
maar sediments in the Eifel, Germany. Quat Sci Rev 62:5676.
25. Genty D, et al. (2010) Isotopic characterization of rapid climatic events during OIS3
and OIS4 in Villars Cave stalagmites (SW-France) and correlation with Atlantic and
Mediterranean pollen records. Quat Sci Rev 29:27992820.
26. Dreybrodt W, Scholz D (2011) Climatic dependence of stable carbon and oxygen
isotope signals recorded in speleothems: From soil water to speleothem calcite.
Geochim Cosmochim Acta 75:734752.
27. Wang G, Jia Y, Li W (2015) Effects of environmental and biotic factors on carbon
isotopic fractionation during decomposition of soil organic matter. Sci Rep 5:11043.
28. Tremaine DM, Froelich PN, Wang Y (2011) Speleothem calcite farmed in situ: Modern
calibration of δ
18
O and δ
13
C paleoclimate proxies in a continuously-monitored natural
cave system. Geochim Cosmochim Acta 75:49294950.
29. Dr˘
agus
¸in V, et al. (2017) Transfer of environmental signals from the surface to the
underground at Ascuns˘
a Cave, Romania. Hydrol Earth Syst Sci 21:53575373.
30. Hatté C, et al. (2013) Excursions to C4 vegetation recorded in the upper Pleistocene
loess of surduk (Northern Serbia): An organic isotope geochemistry study. Clim Past 9:
10011014.
31. Dr˘
agus
¸in V, et al. (2014) Constraining Holocene hydrological changes in the Carpathian
Balkan region using speleothem
18
O and pollen-based temperature reconstructions.
Clim Past 10:13631380.
32. Scholz D, Hoffmann DL (2011) StalAgeAn algorithm designed for construction of
speleothem age models. Quat Geochronol 6:369382.
33. Ménot G, Bard E (2012) A precise search for drastic temperature shifts of the past
40,000 years in southeastern Europe: Abrupt SST event in southeastern Europe.
Paleoceanography, 27:PA2210.
34. Rozanski K, Araguas-Araguas L, Gonfiantini R (1993) Isotopic patterns in modern
global precipitation. Climate Change in Continental Isotope Records, Geophysical
Monograph, eds Swart PK, et al. (American Geophysical Union, Washington, DC), pp
136.
35. Huntley B, et al. (2003) European vegetation during marine oxygen isotope Stage-3.
Quat Res 59:195212.
36. van Huissteden K, Vandenberghe J, Pollard D (2003) Palaeotemperature reconstruc-
tions of the European permafrost zone during marine oxygen isotope stage 3 com-
pared with climate model results. J Quat Sci 18:453464.
37. Vandenberghe J, et al. (2014) The last permafrost maximum (LPM) map of the
northern hemisphere: Permafrost extent and mean annual air temperatures, 25
17 ka BP. Boreas 43:652666.
38. Sanchi L, Menot G, Bard E (2014) Insights into continental temperatures in the
northwestern Black Sea area during the last glacial period using branched tetraether
lipids. Quat Sci Rev 84:98108.
39. Antoine P, et al. (2016) Upper Pleistocene loess-palaeosol records from northern
France in the european context: Environmental background and dating of the middle
palaeolithic. Quat Int 411:424.
40. Morin E (2008) Evidence for declines in human population densities during the early
Upper Paleolithic in western Europe. Proc Natl Acad Sci USA 105:4853.
41. Discamps E, Jaubert J, Bachellerie F (2011) Human choices and environmental con-
straints: Deciphering the variability of large game procurement: From mousterian to
Aurignacian times (MIS 5-3) in southwestern France. Quat Sci Rev 30:27552775.
42. Vandevelde S, Brochier JÉ, Petit C, Slimak L (2017) Establishment of occupation
chronicles in Grotte Mandrin using sooted concretions: Rethinking the middle to
upper paleolithic transition. J Hum Evol 112:7078.
43. Conard NJ, Bolus M (2003) Radiocarbon dating the appearance of modern humans
and timing of cultural innovations in Europe: New results and new challenges. J Hum
Evol 44:331371.
44. Higham T, et al. (2012) Τesting models for the beginnings of the Aurignacian and the
advent of figurative art and music: The radiocarbon chronology of Geißenklösterle.
J Hum Evol 62:664676.
45. Böhner U (2008) Sesselfelsgrotte IVDie Schicht E3 der Sesselfelsgrotte und die
Funde am Abri Schulerloch (Franz Steiner Verlag Stuttgart, Stuttgart, Germany).
46. Devièse T, et al. (2017) Direct dating of Neanderthal remains from the site of Vindija
cave and implications for the middle to upper paleolithic transition. Proc Natl Acad Sci
USA 114:1060610611.
47. Uthmeyer T (1996) Ein bemerkenswert frühes Inventar des Aurignacien von der
Freilandfundstelle Keilberg-Kirchebei Regensburg. Archaeologisches Korrespondenzblatt
26:233248.
48. Davies W, et al. (2015) Evaluating the transitional mosaic: Frameworks of change
from Neanderthals to Homo sapiens in eastern Europe. Quat Sci Rev 118:211242.
49. Richter J (2016) Leave at the height of the party: A critical review of the middle pa-
leolithic in western central Europe from its beginnings to its rapid decline. Quat Int
411:107128.
50. Douka K, Grimaldi S, Boschian G, del Lucchese A, Higham TF (2012) A new chro-
nostratigraphic framework for the upper palaeolithic of Riparo Mochi (Italy). J Hum
Evol 62:286299.
51. Higham T, et al. (2009) Problems with radiocarbon dating the middle to upper pa-
laeolithic transition in Italy. Quat Sci Rev 28:12571267.
52. Hublin J-J, et al. (2012) Radiocarbon dates from the Grotte du Renne and Saint-
Césaire support a Neandertal origin for the Châtelperronian. Proc Natl Acad Sci USA
109:1874318748.
53. Talamo S, et al. (2012) A radiocarbon chronology for the complete middle to upper
Palaeolithic transitional sequence of Les Cottés (France). J Archaeol Sci 39:175183.
54. Barshay-Szmidt C, et al. (2018) Radiocarbon dating the Aurignacian sequence at
Isturitz (France): Implications for the timing and development of the proto-
aurignacian and early Aurignacian in western Europe. J Archaeol Sci Rep 17:809838.
55. Wood RE, et al. (2014) The chronology of the earliest upper palaeolithic in northern
Iberia: New insights from LArbreda, Labeko Koba and La Viña. J Hum Evol 69:91109.
56. Camps M, Higham T (2012) Chronology of the Middle to Upper Palaeolithic transition
at Abric Romaní, Catalunya. J Hum Evol 62:89103.
57. Banks WE, dErrico F, Zilhão J (2013) Human-climate interaction during the early
upper paleolithic: Testing the hypothesis of an adaptive shift between the proto-
Aurignacian and the early Aurignacian. J Hum Evol 64:3955.
58. Riel-Salvatore J, Popescu G, Barton CM (2008) Standing at the gates of Europe: Hu-
man behavior and biogeography in the southern Carpathians during the late Pleis-
tocene. J Anthropol Archaeol 27:399417.
59. Douka K, et al. (2014) On the chronology of the Uluzzian. J Hum Evol 68:113.
60. Richards MP, Trinkaus E (2009) Out of Africa: Modern human origins special feature:
Isotopic evidence for the diets of European Neanderthals and early modern humans.
Proc Natl Acad Sci USA 106:1603416039.
61. El Zaatari S, Grine FE, Ungar PS, Hublin J-J (2016) Neandertal versus modern human
dietary responses to climatic fluctuations. PLoS One 11:e0153277.
62. Hoffecker JF (2011) The early upper Paleolithic of eastern Europe reconsidered. Evol
Anthropol 20:2439.
63. Skrdla P (2017) Middle to Upper Paleolithic transition in Moravia: New sites, new
dates, new ideas. Quat Int 450:116125.
64. Torfstein A, Goldstein SL, Stein M, Enzel Y (2013) Impacts of abrupt climate changes in
the levant from last glacial dead sea levels. Quat Sci Rev 69:17.
65. Tzedakis PC, et al. (2004) Ecological thresholds and patterns of millennial-scale cli-
mate variability: The response of vegetation in Greece during the last glacial period.
Geology 32:109112.
66. Nigst PR (2014) First modern human occupation of Europe: The middle Danube region
asacasestudy.Living in the Landscape: Essays in Honour of Graeme Barker, McDonald
Institute Monographs, eds Boyle K, Rabett RJ, Hunt CO (McDonald Institute for
Archaeological Research, University of Cambridge, Cambridge, UK), pp 3547.
67. Hauck TC, et al. (2016) Neanderthals or early modern humans? A revised C-14 chro-
nology and geoarcheological study of the Szeletian sequence in Szeleta Cave
(Kom. Borsod-Abauj-Zemplen) in Hungary. Archaeologisches Korrespondenzblatt 46:
271290.
68. Bicho N, Cascalheira J, Gonçalves C (2017) Early Upper Paleolithic colonization across
Europe: Time and mode of the Gravettian diffusion. PLoS One 12:e0178506.
6of6
|
www.pnas.org/cgi/doi/10.1073/pnas.1808647115 Staubwasser et al.
www.pnas.org/cgi/doi/10.1073/pnas. 115
1
Supplementary Information for
The impact of climate change on the transition of Neanderthals to modern
humans in Europe
Michael Staubwasser, Virgil Drăgușin, Bogdan P. Onac, Sergey Assonov, Vasile Ersek,
Dirk L. Hoffmann, Daniel Veres
Michael Staubwasser
Email:
m.staubwasser@uni-koeln.de
This PDF file includes:
Supplementary text
Figs. S1 to S4
Tables S1 to S2
References for SI reference citations
Other supplementary materials for this manuscript include the following:
Datasets S1
1808 476
2
Supplementary Information Text
Materials and Methods.
Figure S1 shows the location of paleoclimate and environmental records used for
comparison, in addition to archaeologic sites discussed in the text. The two speleothems
presented in this study are from two sites in the East and South Carpathians. Tăușoare Cave (TC)
is situated in the Rodnei Mountains of the East Carpathians, northern Romania (47º 26’ N, 24º 31’
E), at an altitude of 950 m. The cave has a surveyed length of over 8.5 km and is 329 m deep and
hosts an abundance of gypsum crusts and flowers along with the presence of various mixed cation
sulfates (1, 2). Sulfate derives from weathering of pyrite present in the calcite and bituminous
sediments. The process provides significant acidity for calcite dissolution in this cave. Thus, the
isotopic composition of the host-rock calcite rather than soil CO2 dominates the δ13C record of TC
speleothems. Stalagmite 1152 was retrieved from the “Dining Room” at about 625 m from and
about 190 m below the entrance. It is 27 cm long and is made of dark-brown dense calcite (Fig.
S2). The growth axis of stalagmite T-1152 is stable for the first 23 cm and changes direction for
the last 4 cm. Ascunsă Cave (AC) is situated in the Mehedinți Mountains, South Carpathians,
southwestern Romania, at an altitude of 1050 m (45º 00’ N, 22º 36’ E). The cave is 400 m long
and over 200 m deep, formed mostly on the contact between Barremian-Aptian limestone and
Turonian-Senonian wildflysch (melange) (3). Stalagmite POM1 was retrieved from the “White
Chamber”, at about 80 m from the entrance. It is ~30 cm long and composed of alternations of
dense dark calcite and lighter, less dense calcite (Fig. S2). The growth axis presents several slight
direction changes.
For U-Th dating, samples were drilled parallel to growth layers. U and Th were analyzed
by MC-ICP-MS (Neptune+, Thermo) after adding a 229Th – 236U double spike and subsequent
ion-exchange chromatographic separation of U and Th from the sample matrix (4) (Tab. S1). A
detrital Th-correction was applied assuming the average crustal 232Th/238U activity ratio of 1,250
± 0.625. Detrital 230Th correction amounts to within 50 – 350 years for the AC-speleothem, and
10 – 260 years for the TC-speleothem. The final chronology was modeled using the STALAGE
software package (5). Results can be found in Dataset S1.
For stable C and O isotope ratios samples were drilled continuously at 250 µm (TC) and
at 300 µm (AC) resolution and analyzed by gas-IRMS (Thermo 253, and Thermo Delta+) at the
University of Cologne and (TC) and the University of South Florida (AC). Results can be found
in Dataset S1. The commonly applied ‘Hendy-Test’ to exclude kinetic domination of isotope
fractionation was not performed. Both speleothems show relatively low growth rates and poor
resolution of growth layering at the 250 - 300 µm sampling resolution level. The criterion of the
Hendy-Test is to look for covariance of δ13C and δ18O within the same growth layer, which may
hint at potential kinetically dominated isotope fractionation as a result e.g. of rapid degassing or
rapid calcite precipitation. Because the two speleothems generally show little covariance between
δ13C and δ18O overall, the criterion to meet the Hendy-Test is essentially fulfilled. In addition,
cave monitoring water isotope data (δD, δ18O) suggests that oxygen isotopes fractionate at
equilibrium between drip water and calcite, and carbon isotopes may at the worst contain a
kinetic fractionation of 1.0 - 1.5 ‰ (6).
Regarding the association of the TC-stalagmite’s pronounced low δ18O interval between
~ 40.7 and 39.7 ka with GS-10 (between 40.8 and 40.2 ka in the GICC05 chronology) (5), as
opposed to an association with GS-9 – H4 (between 39.9 and 38.2 ka in the GICC05 chronology),
it is worthwhile to assess potential chronologic inaccuracy due to the correction for detrital Th.
The onset of the low δ18O has been dated directly to 40.66 ± 0.47 ka (40.92 ± 0.47 ka without
detrital Th correction) (Tab. S1). An activity ratio for 230Th/232Th of 101 suggests a small detrital
contribution – amounting to an age correction of ~ 260 years. However, detrital 230Th just
3
amounts to ~0.5% of the total measured 230Th activity based on the average crustal Th/U activity
ratio (1.20) for a detrital contamination. To bring the resulting 40.7 ka onset age in line with the
39.9 ka BP onset age of GS-9 – H4 would require a low detrital Th/U ratio of less than 0.4, well
outside the lower boundary for crustal Th/U of 0.625 and unusual for the given setting.
The Ascunsă Cave δ18O record
Stable O-isotopes combine information on temperature and a number of hydrologic
aspects that may be constrained by data comparison in regional and temporal context. Throughout
the Holocene, the amplitude of δ18O variability in the western Mediterranean, the Alps, and
northern Romania – including TC – was significantly lower (~ 0.5 ‰) than in the eastern
Mediterranean and southern Romania – including AC (7). An even larger amplitude difference
between the AC18O and the TC-δ18O existed during MIS-3. (Fig. S3). This cannot be explained
by a different temperature regime alone (see main article) but must reflect a significant
hydroclimatic contribution to the AC speleothem δ18O records. The pacing of δ18O is very
different between AC and TC, and between AC and Greenland temperatures (8, 9). Thus, it is
unlikely, that the two speleothems were inside the same hydrologic regime but record latitudinal
differences in rain-out history from the same source. However, there is a similar structure
between AC-δ18O record and eastern Mediterranean marine and speleothem records (Fig. S4)
(10). This pattern-similarity between the AC-δ18O record and the eastern Mediterranean is
observed only when the AC-stalagmite's δ18O record is plotted with an inverted scale relative to
the speleothem record from the Levant (Fig. S4). Numeric climate simulations of the last glacial
(11) demonstrate the underlying causality for this apparent inverse correlation between the East
Carpathians and the Eastern Mediterranean. All simulated modes of atmospheric circulation
patterns that contributed to the average glacial precipitation over European and the Mediterranean
show seasonal precipitation anomalies of opposite sign between the eastern Mediterranean and
east-central Europe. This synoptic-scale relationship is corroborated by a similar relationship
between the speleothems of AC and Karaca Cave (12), eastern Turkey.
Interstadial GI-7 stands out in the AC-δ18O record with an anomalous isotope event (Fig.
S3), a minimum more than 1 ‰ lower than all other features in the record. It is also a prominent
feature in the speleothem and marine records from the eastern Mediterranean (Fig. S4). This
could suggest that supply and influence of Mediterranean moisture in the region of the South
Carpathians was highly variable between interstadials. Although the loess record of the Danube
Valley appears to confirm this observation within given chronologic uncertainty (13), there is,
however, no apparent explanation available as to the cause of this moist interval during GI-7.
4
Fig. S1. Map of the southern and central Europe with rivers (blue: the Danube), coastlines for sea
level 40 m and 60 m lower than at present, and sites discussed in the main text. Geologic records
(dots): (1) Tăușoare Cave; (2) Ascunsă Cave; (3) core MD04-2790; (4) core M72/5-25-GC1; (5)
Willendorf-II paleosol/loess profile; (6) Eifel Maar Lakes; (7) Villars Cave; (8) core MD95-2042.
Archeologic sites: (a1) Sesselfelsgrotte; (a2) Keilberg-Kirche; (a3) Geissenklösterle; (a4) Riparo
Mochi; (a5) Willendorf-II; (a6) Grotta di Fumane; (a7) Grotte Mandrin; (a8) Pes-kὅ Cave; (a9)
Szeleta Cave; (a10) Vindija Cave; (b1) Labeko Koba; (b2) Abric Romani; (b3) Isturitz; (c1)
Grotte du Renne; (c2) Les Cottés; (c3) Saint Césaire; (d1) Oase Cave; (d2) Goyet Cave. The map
was generated with the GeoMapApp software, http://www.geomapapp.org (14).
5
Fig. S2. U-Th age models and photographs of a) stalagmite 1152, Tăușoare Cave, and b)
stalagmite POM1, Ascunsă Cave.
6
Fig. S3. A: Greenland NGRIP ice core temperatures on the GICC05 time scale with numbered
interstadials (8). B: δ13C of stalagmite 1152 from Tăușoare Cave. C: δ13C of stalagmite POM1
from Ascunsă Cave. D: δ18O of stalagmite 1152 from Tăușoare Cave. E: δ18O of stalagmite
POM1 from Ascunsă Cave.
7
Fig. S4. A: Greenland NGRIP ice core temperatures on the GICC05 time scale with numbered
interstadials (8). B: Eastern Mediterranean and Levantine records (10): δ18O data of planktonic
foraminifera (G. ruber) from eastern Mediterranean marine sediment core MD9501, and δ18O data
of a stalagmite from Soreq cave (Israel). C: δ18O data of stalagmite POM1 from Ascunsă Cave.
Note that the scale for the two stalagmites is inverted with respect to each other.
8
Table S1. U-Th data for stalagmites POM1, Ascunsă Cave, and 1152, Tăușoare Cave. Corrected values include detrital Th correction.
Sample Distance
(mm)
238U
(ng/g) ±
232Th
(ng/g) ±
(230Th/232Th)
activity
ratio
±
(230Th/238U)
activity
ratio
±
(234U/238U)
activity
ratio
± uncorrected
age (ka) ± co rrected
age (ka) ±
(
234
U/
238
U)
initial
activity
ratio
± Comments
1152 top 0,50 1019,5 6,6 3,442 0,034 6,31 0,15 0,0070 0,0002 1,4296 0,0024 0,53 0,01 0,47 0,04 1,4305 0,0024
1152 / XIII 16,25 963,1 34,4 3,054 0,467 97,29 1,28 0 ,1009 0,0038 1 ,4464 0,0033 7,87 0,30 7,81 0,30 1,4567 0,0034
1153 / XII 37,38 1304,6 50,9 2,486 0,125 235,37 2,60 0,1468 0,0013 1 ,5133 0,0035 11,08 0,10 11,04 0,11 1,5298 0,0036
1152 / XI 38,88 1231,0 40,9 1,438 0,054 404,15 5 ,16 0,1545 0,0016 1,5394 0,0039 1 1,48 0,13 11 ,46 0,13 1,5573 0,0040
1152 / X 78,50 2632,7 73,8 1,672 0,062 1069,73 15,75 0,2223 0,0019 1,7095 0 ,0036 15,07 0,14 15,06 0,14 1,7405 0,0037
1152 / IX 81,50 952,4 28,6 2,950 0,101 306,65 2 ,91 0,3108 0,0027 1,8252 0,0044 2 0,10 0,19 20 ,06 0,20 1,8740 0,0045
1152 / VIII 97,63 1854,4 70,8 3,301 0,259 578,76 7,41 0,3371 0,0028 1 ,7996 0,0060 22,31 0,22 22,28 0,22 1,8520 0,0062
1152 / VII 114,88 1674,3 61,1 1 ,603 0,552 1061,09 7,72 0,3323 0 ,0045 1,7126 0,0029 2 3,21 0,35 23 ,20 0,35 1,7610 0,0031
1152 / VI 135,88 942,6 33,0 1,656 0,279 583,06 4,91 0,3352 0,0034 1 ,6309 0,0050 24,75 0,29 24,72 0,29 1,6769 0,0052
1152 XXII 149,63 1086 ,5 53,1 1,051 0,054 1161,09 11,40 0,3674 0 ,0025 1,6158 0,0034 27 ,71 0,22 27,69 0,22 1,6661 0,0036
1152 / V 155,00 1397,5 39,1 1,318 0,040 1151,27 11,89 0,3552 0,0024 1,5133 0,0037 28,77 0 ,24 28,75 0,24 1,5568 0,0039
1152 /IV 174,63 931,3 30,0 2,383 0,085 4 72,40 4,68 0 ,3955 0,0031 1,5418 0,0034 31,80 0,30 31,75 0 ,30 1,5931 0,0037
1152 XVII 179,88 802,2 16 ,3 1,246 0,047 802,42 22,17 0,4078 0,0022 1,5125 0 ,0031 33,68 0,22 33,65 0,22 1,5639 0,0033
1152 XVI 187,88 612,2 13,9 3,442 0,122 243 ,85 6,17 0,4486 0 ,0023 1,5034 0,0033 37,91 0,25 37,80 0,25 1,5610 0,0035
1152 / III 193,63 517,7 21,8 7,404 0,315 101,37 1,00 0,4744 0,0041 1,4903 0 ,0049 40,92 0,45 40,66 0,47 1,5521 0,0054
1152 XXI 196,25 506,4 12,8 3,987 0,106 188 ,71 1,28 0,4862 0 ,0029 1,5296 0,0031 40,81 0,31 40,67 0,31 1,5953 0,0034
1152 XV 201,25 566,7 14,1 3,243 0,093 269 ,37 3,24 0,5044 0 ,0031 1,4390 0,0033 46,02 0,37 45,91 0,37 1,5006 0,0036
1152 XIV 207,25 639,8 12,8 2,372 0,052 433 ,22 3,25 0,5255 0 ,0023 1,4452 0,0030 48,12 0,29 48,05 0,29 1,5105 0,0033
1152 / II 215 ,00 570,5 16,6 3,212 0,100 292,59 2,68 0,5389 0 ,0039 1,4293 0,0040 50 ,36 0,48 50,25 0,49 1,4955 0,0044
1152 / I 234,88 1182 ,4 38 ,7 6,035 0,229 360,50 2,83 0,6021 0,0035 1,4059 0,0034 59,21 0,49 59,11 0,49 1,4803 0,0038
1152/ base 264,00 622,8 3,5 5,433 0,033 225,32 0,50 0 ,6432 0 ,0023 1,3694 0,0027 66,98 0,37 66,81 0,38 1,4472 0,0031
POM 1 / top 32 ,5 0,2 0 ,576 0,005 22,93 0,51 0,1331 0,0031 1,2610 0,0040 12,14 0,30 11,74 0,36 1,2710 0,0041 outside measured profile
POM 1 / III 29,67 38,1 0,2 1,923 0,017 20,95 0,18 0,3460 0,0028 1 ,2746 0,0032 34,18 0,34 33,07 0,61 1,3055 0,0040
POM1/X 94,33 45,2 0,3 0,164 0,005 292,85 3,83 0,3472 0,0040 1,1799 0,0044 37,69 0,54 37,61 0 ,54 1,2003 0,0048
POM 1 / V 119,67 84,8 0 ,4 0,620 0,006 140,69 1,06 0,3366 0,0021 1,1306 0,0027 38,32 0,31 38,14 0,32 1,1457 0,0030
POM1/VIII
130,33
70,1
0,4
0,183
0,006
422,27
5,43
0,3602
0,0027
1,1942
0,0036
38,79
0,38
38,73
0,38
1,2168
0,0039
POM1/VII 153,33 37,3 0,2 0,313 0,003 160,76 1 ,64 0,4406 0,0038 1,3985 0,0047 40,53 0,45 40,37 0,46 1,4477 0,0051
POM1/XI 175,33 27,7 0,1 0,079 0,002 473,33 5,60 0,4390 0,0042 1,3776 0,0048 41,13 0,50 41,07 0,50 1,4244 0,0052
POM 1 / II 226,67 27,7 0,2 0,386 0,004 104,84 1,52 0,4777 0,0063 1 ,4225 0,0040 43,73 0,71 43,46 0,72 1,4794 0,0045
POM 1 / I 228,33 20,9 0 ,1 0,396 0,004 80,09 1,01 0,4955 0,0056 1,4684 0,0050 4 3,91 0,63 43 ,56 0,64 1,5324 0,0056
POM 1 XXII 267,00 24,3 0,1 0 ,239 0,003 1 49,77 1,29 0 ,4825 0,0035 1,3771 0,0031 46,11 0,43 45,92 0 ,43 1,4304 0,0034
POM 1 XXIII 289,00 26,4 0,1 0 ,236 0,006 1 62,37 1,71 0 ,4752 0,0042 1,3742 0,0036 45,38 0,51 45,20 0 ,51 1,4262 0,0040
POM 1 /base 33,2 0 ,2 29,626 0 ,162 2,47 0 ,02 0,7209 0,0054 1,3700 0,0036 78,29 0,88 59,06 7,83 1,5708 0,0819 discarded from age model
9
Table S2. Published 14C ages of archaeologic sequences included in Figure 3 and discussed in the text.
site
map
signature
(Fig. S1)
stratigraphic
position / layer
chronologic
position in
sequence
sample material conventional
14C age, ka
ka, cal BP (
p
= 68%,
INTCAL13), original
calibration (c) /
Bayesian model (m);
recalibrated (r)
preceding
layer, min
age ka, cal
BP (p =
68%)
lower layer
boundary
ka, cal BP
(p = 68%)
upper layer
boundary ka,
cal BP (p =
68%)
succeding
layer, max
age ka, cal
BP (p =
68%)
Reference
original authors'
artifact identification
Grotte du
Rennes at Arcy-
sur-Cure
c1
VII
oldest
EVA
-
95
bone
34.81 ± 0.21
39.5
-
40.3
(m)
15
Protoaurignacian
VIII
model
transition
VIII/VII
Bayesian model
age 40.7 - 41.4 (m)
Chatelperronian, low
artifact density
model
transition
IX/VIII
Bayesian model
age 41.2 - 41.6 (m)
IX
youngest
EVA
-
29
bone
35.50 ± 0.22
41.0
-
42.1 (m)
Chatelperronian
summary VIII
41.0
41.6
40.7
40.3
Saint Cesaire c2
ejo sup
n.a.
n.a.
16
Aurignacian 0
ejop sup / ejo inf
model
transition
Bayesian model
age 39.7 - 41.1 (m) sterile (ejo inf)
ejop sup
youngest
OxA
-
21699
bone
36.00 ± 0.70
40.3
-
41.3 (m)
Chatelperronian
summary ejo inf
41.3
41.1
39.7
n.a.
Les Cottes c3
4
oldest
S
-
EVA 9713
bone
35.15 ± 0.28
39.3
-
40.0
17
Protoaurignacian
5/4
upper
boundary
Bayesian model
age 39.5 - 40.3 (m)
sterile (layer 5)
6/5
lower
boundary
Bayesian model
age 40.8 - 41.6 (m)
6
youngest
S
-
EVA 13666
bone
36.23 ± 0.21
41.3
-
41.7 (m)
Chatelperronian
summary 5
41.3
41.6
39.5
40.0
Europe -
Bayesian age
model 40.0 - 40.8 40.8 40.0 16
Neanderthal
disappearance
Labeko Koba b1
VII
oldest
OxA
-
21766
bone
36.85 ± 0.80
41.0
-
41.6
(m)
18
Protoaurignacian
VIII / VII
upper
boundary
Bayesian model
age 41.4 - 42.0 (m)
sterile
VIII
n.a.
IX
-
upper
youngest
OxA
-
21972
bone
36.55 ± 0.75
41.6
-
42.1 (m)
IX
-
upper
oldest
OxA
-
23199
bone
38.40 ± 0.90
41.7
-
42.2 (m)
IX
-
lower / IX
-
upper
lower
boundary 41.9 - 42.4 (m)
IX
-
lower
youngest
OxA
-
22560
bone
37.40 ± 0.80
42.2
-
42.6 (m)
Chatelperronian
summary IX
-
upper 42.2 42.2 41.6 41.6
10
Table S2 continued. Published 14C ages of archaeologic sequences included in Figure 3 and discussed in the text.
site
map
signature
(Fig. S1)
stratigraphic
position / layer
chronologic
position in
sequence
sample materia l conventional
14C age, ka
ka, cal BP (
p
= 68%,
INTCAL13), original
calibration (c) /
Bayesian model (m);
recalibrated (r)
preceding
layer, min
age ka, cal
BP (p =
68%)
lower layer
boundary
ka, cal BP (p
= 68%)
upper layer
boundary
ka, cal BP (p
= 68%)
succeding
layer, max
age ka, cal
BP (p =
68%)
Reference
original authors' artifact
identification
Abric Romani b2
A
oldest
OxA
-
X
-
2095
-
46
shell
36.09 ± 0.23
41.3
-
41.7
(m)
19
Aurignacian 0
AR3/AR6 av. U-series
Bayesian model
age
41.6
-
42.6 (U
-
Th
age) sterile
B
youngest
OxA
-
12025
shell
39.06 ± 0.35
42.8
-
43.5 (m)
Mousterian
summary
AR3/AR6 42.8 42.8 41.7 41.7
Sesselfelsgrotte
a1
E3
GrN
-
7153
charcoal
37.1 ± 1.1
40.7
-
42.3
(r)
20
MMO
G1
GrN
-
6848
charcoal
41.84 ± 1.10
44.2
-
46.2 (r)
Mousterian
G1
GrN
-
20302
bone
39.95 ± 0.92
42.9
-
44.4 (r)
G1
GrN
-
20303
bone
41.37 ± 1.06
43.6
-
45.7 (r)
G1 GrN-21528
combined bone
from GrN-20302 &
20303
41.39 ± 0.58 44.3 - 45.4 (r)
summary F
44.3
n.a
n.a
42.3
Keilberg a2 layer 2, base
KN
-
4690
charcoal
37.50 ± 1.45
40.5
-
42.9 (r)
21 Aurignacian
KN
-
4691
charcoal
37.50 ± 1.25
40.7
-
42.8 (r)
KN
-
4692
charcoal
38.6 ± 1.20
41.7
-
43.6 (r)
Geissenklösterle
a3
16
oldest
OxA
-
21722
bone
41.8
-
42.7 (m)
22
Early Aurignacian
17/16
upper boundary
42.4
-
43.7 (m)
17
OxA
-
21658
bone
42.4
-
43.3 (m)
sterile
17
OxA
-
21657
bone
42.4
-
43.4
(m)
18/17
lower boundary
42.5
-
44.8 (m)
18
youngest
OxA
-
21720
bone
42.5
-
44.5 (m)
Mousterian
summary 17
42.5
43.4
42.4
42.7
Riparo Mochi a4
G
oldest
Rome
-
2
charcoal
37.4 ± 1.3
41.4
-
42.3 (m)
23
Protoaurignacian
transition H/G range
Bayesian model
age 41.6 - 42.8 (m) sterile
transition I/H range
Bayesian model
age 41.8 - 44.0 (m) Mousterian
summary H
n.a.
44.0
41.6
42.2
WIllendorf 5/a5
C8.2
average of 9 OxA
& GrA dates charcoal 41.7 - 42-9 (c)
24
Early Aurignacian
C9
n.a.
n.a.
D1
average of 9 OxA
& GrA dates charcoal 43.9 - 46.6 (c) IUP
summary C9
43.9
42.9
11
References
1. Onac BP, Drăguşin V, Papiu F, Theodorescu C-T (2019) Rodna Mountains: Izvorul Tăusoarelor Cave
(Pestera de la Izvorul Tăusoarelor). Cave and Karst Systems of Romania, eds Ponta GML, Onac BP
(Springer International, Cham), pp 47–56.
2. Onac BP, White WB, Viehmann I (2001) Leonite [K2Mg(SO4)2·4H2O], konyaite [Na2Mg(SO4)2·5H2O]
and syngenite [K2Ca(SO4)2·H2O] fromTausoare Cave, Rodnei Mts, Romania. Min Mag 65(1):103-109.
3. Drăguşin V, Vlaicu M, Isverceanu E (2019) Mehedinti Mountains: The Cave from Mohilii Creek
(Ascunsă Cave). Cave and Karst Systems of Romania, eds Ponta GML, Onac BP (Springer
International, Cham), pp 171–173.
4. Hoffmann D, et al. (2007) Procedures for accurate U and Th isotope measurements by high precision
MC-ICPMS. Int J Mass Spec 264(2-3): 97-109.
5. Scholz D, Hoffmann DL (2011) StalAge – An algorithm designed for construction of speleothem age
models. Quat Geochronol 6(3–4):369–382.
6. Dragusin V, et al. (2017) Transfer of environmental signals from the surface to the underground at
Ascunsă Cave, Romania. Hydrol Earth Syst Sci 21(10):5357–5373.
7. Drăguşin V, et al. (2014) Constraining Holocene hydrological changes in the Carpathian–Balkan region
using speleothem δ18O and pollen-based temperature reconstructions. Clim Past 10(4):1363–1380.
8. Rasmussen SO, et al. (2014) A stratigraphic framework for abrupt climatic changes during the Last
Glacial period based on three synchronized Greenland ice-core records: refining and extending the
INTIMATE event stratigraphy. Quat Sci Rev 106:14–28.
9. Kindler P, et al. (2014) Temperature reconstruction from 10 to 120 kyr b2k from the NGRIP ice core.
Clim Past 10(2):887–902.
10. Almogi-Labin A, et al. (2009) Climatic variability during the last ~90ka of the southern and northern
Levantine Basin as evident from marine records and speleothems. Quat Sci Rev 28:2882-2896.
11. Hofer D, et al. (2012) Simulated winter circulation types in the North Atlantic and European region for
preindustrial and glacial conditions: Glacial Circulation Types. Geophys Res Lett 39(15): DOI
10.1029/2012GL052296.
12. Rowe PJ et al. (2012) Speleothem isotopic evidence of winter rainfall variability in northeast Turkey
between 77 and 6 ka. Quat Sci Rev 45:60-72.
13. Obreht I, et al. (2017) Shift of large-scale atmospheric systems over Europe during late MIS 3 and
implications for Modern Human dispersal. Sci Rep 7:5848.
14. Ryan, W.B.F, et al (2009) Global Multi-Resolution Topography synthesis, Geochem. Geophys.
Geosyst., 10, Q03014, doi:10.1029/2008GC002332.
15. Hublin J-J, et al. (2012) Radiocarbon dates from the Grotte du Renne and Saint-Césaire support a
Neanderthal origin for the Châtelperronian. Proc Natl Acad Sci U S A 109(46):18743–18748.
16. Higham T, et al. (2014) The timing and spatiotemporal patterning of Neanderthal disappearance.
Nature 512(7514):306–309.
17. Talamo S, et al. (2012) A radiocarbon chronology for the complete Middle to Upper Palaeolithic
transitional sequence of Les Cottés (France). J Archaeol Sci 39(1):175–183.
18. Wood RE, et al. (2014) The chronology of the earliest Upper Palaeolithic in northern Iberia: New
insights from L’Arbreda, Labeko Koba and La Vina. J Hum Evol 69:91–109.
19. Camps M, Higham T (2012) Chronology of the Middle to Upper Palaeolithic transition at Abric
Romani, Catalunya. J Hum Evol 62(1):89–103.
20. Böhner U (2008) Sesselfelsgrotte IV - Die Schicht E3 der Sesselfelsgrotte und die Funde am Abri
Schulerloch (Franz Steiner Verlag Stuttgart, Germany).
21. Uthmeyer T (1996) Ein bemerkenswert frühes Inventar des Aurignacien von der Freilandfundstelle
“Keilberg-Kirche” bei Regensburg. Archaeologisches Korrespondenzblatt 26:233–248.
22. Higham T, et al. (2012) Τesting models for the beginnings of the Aurignacian and the advent of
figurative art and music: The radiocarbon chronology of Geißenklösterle. J Hum Evol 62(6):664–676.
23. Douka K, et al. (2012) A new chronostratigraphic framework for the Upper Palaeolithic of Riparo
Mochi (Italy). J Hum Evol 62(2):286–299.
24. Nigst PR, et al. (2014) Early modern human settlement of Europe north of the Alps occurred 43,500
years ago in a cold steppe-type environment. Proc Natl Acad Sci 111(40):14394–14399.

Supplementary resource (1)

... More recent versions have also attempted to take other factors (e.g. random events or climate fluctuations) into account, in varying degrees (Finlayson, 2004;van Andel and Davies, 2004;Finlayson, 2005;Finlayson and Carrión, 2007;Finlayson, 2008;Jennings et al., 2011;Müller et al., 2011;Sørensen, 2011;Stewart and Stringer, 2012;El Zaatari et al., 2016;Timmermann et al., 2022;Kolodny and Feldman, 2017;Shultz et al., 2019;Melchionna et al., 2018;Staubwasser et al., 2018;Timmermann, 2020Bicho and Carvalho, 2022;Vahdati et al., 2022;Zeller et al., 2023;Ruan et al., 2023;Margari et al., 2023). ...
... Models provide a range of possible scenarios, dependent on the model structure and the data used to constrain the model parameters. Scenarios range from a direct Modern Human intervention to climatic and random factors, or a mix of these (Timmermann and Friedrich, 2016;Kolodny and Feldman, 2017;Shultz et al., 2019;Melchionna et al., 2018;Staubwasser et al., 2018;Timmermann, 2020;Timmermann and Friedrich, 2016;Bicho and Carvalho, 2022;Vahdati et al., 2022). The only conclusion that we can draw from this motley group of disparate schemes is that it is possible to explain the empirical patternnamely the disappearance of the Neanderthalswith a multitude of factors and processes, both with and without the need of Modern Human interaction. ...
Article
Recent advances in the study of ancient DNA recovered from fossils and cave sediments have profoundly changed our views on the biological and cultural interactions between populations and lineages of fossil Homo in the Later Pleistocene of Eurasia. A spatiotemporally complex picture emerges, with multiple population admixture and replacement events. Focusing on the evidence from Western Eurasia, we consider here how the mapping out of between-species interactions based on fossil and material cultural evidence is being replaced by a broader approach. Traditional narratives about human migrations and the biological and/or cultural advantages of our own species over the Neanderthals are now giving way to the study of the biological and cultural dynamics of past human populations and the nature of their interactions in time and space.
... Rather than a rapid and straightforward replacement of Neanderthals by H. sapiens, the Middle to Upper Paleolithic transition (MUPT) was characterized by a mosaic of cultural and biological landscapes lasting several thousand years (6)(7)(8). Ancient DNA studies showed that our species and Neanderthals interbred (9) and, therefore, coexisted in some regions (10)(11)(12); nonetheless, in other areas of Europe, Neanderthals were quickly replaced by H. sapiens or even disappeared a few millennia before their arrival (13)(14)(15). The factors that triggered this spatiotemporal replacement pattern are unknown. ...
... On the basis of compared nucleotide diversity estimates, Neanderthal remains from central (R_13) and northern (R_12) Iberia dated to the MIS3 revealed more than sixfold lower diversity than Neanderthals recovered in Eastern Europe (38). Likewise, Neanderthal genetic bottlenecks or demographic vacuums have also been proposed for Northern Europe (R_16 and R_5) (14), the Carpathian region (R_2) (15), and Central and Northern Iberia (13,19). In all these regions where Neanderthals experienced a loss of genetic diversity or where archaeological data suggest a demographic hiatus before the arrival of H. sapiens, the ecosystems' productivity was low or unstable during MIS3. ...
Article
Full-text available
It has been proposed that climate change and the arrival of modern humans in Europe affected the disappearance of Neanderthals due to their impact on trophic resources; however, it has remained challenging to quantify the effect of these factors. By using Bayesian age models to derive the chronology of the European Middle to Upper Paleolithic transition, followed by a dynamic vegetation model that provides the Net Primary Productivity, and a macroecological model to compute herbivore abundance, we show that in continental regions where the ecosystem productivity was low or unstable, Neanderthals disappeared before or just after the arrival of Homo sapiens . In contrast, regions with high and stable productivity witnessed a prolonged coexistence between both species. The temporal overlap between Neanderthals and H. sapiens is significantly correlated with the carrying capacity of small- and medium-sized herbivores. These results suggest that herbivore abundance released the trophic pressure of the secondary consumers guild, which affected the coexistence likelihood between both human species.
... During the time investigated, a replacement of Neanderthal population by Aurignacian (Anatomically Modern Humans, AMH) took place in Europe [15,[40][41][42][43]. This so called Mid to Upper Paleolithic Transition (MUPT) was a process that took place roughly between 45 and 35 ka cal BP [44]. ...
... The reasons for the demise of Neanderthals at around 40,000 yr b2k are still under debate. Among others, factors such as abrupt climate deterioration [42] may have caused a population decline, although NH are assumed to have been the most cold-adapted of all hominins [45] (see [46] for further discussion). During cold phases, fire was only infrequently used, which may have been due to the fact that firewood became less available [47]. ...
Article
Full-text available
We present a record of pollen and spores of coprophilous fungi from a sediment core from Auel infilled maar, Eifel, Germany, covering the period from 42,000 to 36,000 yr b2k. We can show that vegetation cover was dominated by a boreal forest with components of steppe and cold-temperate wood taxa. The proportion of wood taxa was higher during interstadials, whereas steppe-vegetation became more prominent during stadials. During Heinrich stadial 4, temperate taxa are mostly absent. Spores of coprophilous fungi show that megaherbivores were continuously present, albeit in a larger number during stadials when steppe environment with abundant steppe herbs expanded. With the onset of Greenland stadial 9, forests became more open, allowing for steppe-environment to evolve. The shift in vegetation cover coincides with the highest values of herbivore biomass at the time that Neanderthal humans demised and Anatomically Modern Humans most probably arrived in Central and Western Europe. Megaherbivore biomass was a direct consequence of vegetation cover/availability of food resources and thus an indirect consequence of a changing climate. Herds of large herbivores following suitable (steppe) habitats may have been one cause of the migration of AMH into Europe, going along with their prey to productive hunting grounds.
... Competition with H. sapiens more generally is often listed as a cause of Neandertal extinction (e.g., Slimak et al. 2022, Timmermann 2020, along with competition with other Neandertal groups (Chang & Nowell 2020, Rios-Garaizar et al. 2022. Other researchers looked to external factors, such as climate change (Finlayson 2009, Staubwasser et al. 2018, Vahdati et al. 2022, Vernot et al. 2021; but see Columbu et al. 2020), including that driven by pole reversals in conjunction with grand solar minima (Cooper et al. 2021). Finally, many authors argue that a unique combination of factorsa perfect storm-led to Neandertal extinction, and if not for this unhappy accident, the fate of Neandertals (and presumably of humans) might have been different (Finlayson 2009, Shea 2008, Vahdati et al. 2022. ...
... In this study, we investigate the extent to which several fungi species are fractionating carbon isotopes during their development, as compared to the growth medium. For this experiment, we took four soil samples under grassland and forest cover in the Mehedinţi Mountains (SW Romania), where speleothem δ 13 C records were published (Drȃguşin et al., 2014;Staubwasser et al., 2018) and a cave monitoring program is ongoing (Drȃguşin et al., 2017(Drȃguşin et al., , 2020. Soil samples were taken from 10 cm and 30 cm depths, aliquots of 1 g from each sample were mixed with 9 mL of sterile water solution, then three-fold dilutions were obtained and 100 µL aliquots of each dilution were spread on Sabouraud dextrose agar (this medium contains glucose, meat peptone, casein peptone, and agar). ...
... Understanding the process of Homo sapiens' first dispersal across Europe and Asia remains a major objective in paleoanthropological science. The current prevailing model postulates that this event occurred during a succession of millennial-scale warm phases of the last glacial period, 45 to 40 thousand calendar years (ka) ago, immediately following an intense cold phase that caused a decline in resident Neanderthal populations (1)(2)(3). At odds with this view, however, are local proxy records from some of the earliest European Upper Paleolithic sites, suggesting a cold-phase dispersal (4)(5), as well as new claims of migrations occurring as early as 54 ka ago (6)(7) and even earlier (8). ...
Article
Full-text available
The dispersal of Homo sapiens in Siberia and Mongolia occurred by 45 to 40 thousand years (ka) ago; however, the climatic and environmental context of this event remains poorly understood. We reconstruct a detailed vegetation history for the Last Glacial period based on pollen spectra from Lake Baikal. While herb and shrub taxa including Artemisia and Alnus dominated throughout most of this period, coniferous forests rapidly expanded during Dansgaard-Oeschger (D-O) events 14 (55 ka ago) and 12 to 10 (48 to 41 ka ago), with the latter presenting the strongest signal for coniferous forest expansion and Picea trees, indicating remarkably humid conditions. These abrupt forestation events are consistent with obliquity maxima, so that we interpret last glacial vegetation changes in southern Siberia as being driven by obliquity change. Likewise, we posit that major climate amelioration and pronounced forestation precipitated H. sapiens dispersal into Baikal Siberia 45 ka ago, as chronicled by the appearance of the Initial Upper Paleolithic.
... Pollen data show a major environmental change marked by the expansion of Cichorioideaedominated grasslands comparable to findings from the Fimon PD and TdA records ( Supplementary Fig. S1). Although it's well known that pollen deposition in caves is complex and subjected to stochastic elements 89 , a suitable correlation with an "off-site" record corroborates the consistency of the Broion cave record that preserved the Fig. 5) may have played a role in triggering these dynamics in S-European ecosystems 4 , although its expression is less remarkable in the Mediterranean 90 . After the Uluzzian phase, the Proto-Aurignacian one developed in a context of extremely open and harsh conditions (Fig. 7) roughly corresponding to GS 10-9/HS4, which favoured hunting of alpine ibex and chamois 91 . ...
Article
Full-text available
Observation of high-resolution terrestrial palaeoecological series can decipher relationships between past climatic transitions, their effects on ecosystems and wildfire cyclicity. Here we present a new radiocarbon dated record from Lake Fimon (NE-Italy) covering the 60–27 ka interval. Palynological, charcoal fragments and sediment lithology analysis were carried out at centennial to sub-centennial resolutions. Identification of the best modern analogues for MIS 3 ecosystems further enabled to thoroughly reconstruct structural changes in the vegetation through time. This series also represents an “off-site” reference record for chronologically well-constrained Palaeolithic sites documenting Neanderthal and Homo sapiens occupations within the same region. Neanderthals lived in a mosaic of grasslands and woodlands, composed of a mixture of boreal and broad-leaved temperate trees analogous to those of the modern Central-Eastern Europe, the Southern Urals and central-southern Siberia. Dry and other grassland types expanded steadily from 44 to 43 ka and peaked between 42 and 39 ka, i.e., about the same time when Sapiens reached this region. This vegetation, which finds very few reliable modern analogues in the adopted Eurasian calibration set, led to the expansion of ecosystems able to sustain large herds of herbivores. During 39–27 ka, the landscape was covered by steppe, desert-steppe and open dry boreal forests similar to those of the modern Altai-Sayan region. Both Neanderthal and Sapiens lived in contexts of expanded fire-prone ecosystems modulated by the high-frequency climatic cycles of MIS 3.
... Competition with H. sapiens more generally is often listed as a cause of Neandertal extinction (e.g., Slimak et al. 2022, Timmermann 2020, along with competition with other Neandertal groups (Chang & Nowell 2020, Rios-Garaizar et al. 2022. Other researchers looked to external factors, such as climate change (Finlayson 2009, Staubwasser et al. 2018, Vahdati et al. 2022, Vernot et al. 2021but see Columbu et al. 2020), including that driven by pole reversals in conjunction with grand solar minima (Cooper et al. 2021). Finally, many authors argue that a unique combination of factorsa perfect storm-led to Neandertal extinction, and if not for this unhappy accident, the fate of Neandertals (and presumably of humans) might have been different (Finlayson 2009, Shea 2008, Vahdati et al. 2022. ...
Article
In this article, I first provide an overview of the Neandertals by recounting their initial discovery and subsequent interpretation by scientists and by discussing our current understanding of the temporal and geographic span of these hominins and their taxonomic affiliation. I then explore what progress we have made in our understanding of Neandertal lifeways and capabilities over the past decade in light of new technologies and changing perspectives. In the process, I consider whether these advances in knowledge qualify as so-called Black Swans, a term used in economics to describe events that are rare and unpredictable and have wide-ranging consequences, in this case for the field of paleoanthropology. Building on this discussion, I look at ongoing debates and focus on Neandertal extinction as a case study. By way of discussion and conclusion, I take a detailed look at why Neandertals continue to engender great interest, and indeed emotion, among scientists and the general public alike. 151
... Competition with H. sapiens more generally is often listed as a cause of Neandertal extinction (e.g., Slimak et al. 2022, Timmermann 2020, along with competition with other Neandertal groups (Chang & Nowell 2020, Rios-Garaizar et al. 2022. Other researchers looked to external factors, such as climate change (Finlayson 2009, Staubwasser et al. 2018, Vahdati et al. 2022, Vernot et al. 2021but see Columbu et al. 2020), including that driven by pole reversals in conjunction with grand solar minima (Cooper et al. 2021). Finally, many authors argue that a unique combination of factorsa perfect storm-led to Neandertal extinction, and if not for this unhappy accident, the fate of Neandertals (and presumably of humans) might have been different (Finlayson 2009, Shea 2008, Vahdati et al. 2022. ...
Article
In this article, I first provide an overview of the Neandertals by recounting their initial discovery and subsequent interpretation by scientists and by discussing our current understanding of the temporal and geographic span of these hominins and their taxonomic affiliation. I then explore what progress we have made in our understanding of Neandertal lifeways and capabilities over the past decade in light of new technologies and changing perspectives. In the process, I consider whether these advances in knowledge qualify as so-called Black Swans, a term used in economics to describe events that are rare and unpredictable and have wide-ranging consequences, in this case for the field of paleoanthropology. Building on this discussion, I look at ongoing debates and focus on Neandertal extinction as a case study. By way of discussion and conclusion, I take a detailed look at why Neandertals continue to engender great interest, and indeed emotion, among scientists and the general public alike. Expected final online publication date for the Annual Review of Anthropology, Volume 52 is October 2023. Please see http://www.annualreviews.org/page/journal/pubdates for revised estimates.
Chapter
This chapter focuses on the modalities of large fauna exploitation by human groups and their carnivorous competitors throughout the Paleolithic. It highlights the great diversity of hominin-animal interactions throughout human evolution, particularly in Africa and Eurasia. Carnivorous diet gradually increased within human populations, in parallel with the development of hunting capacities. There is evidence of regular animal meat and fat consumption by extinct hominins from 2 Ma onward, with the first occurrence prior to 3 Ma in Eastern Africa. The consumption of meat and fat may have had significant consequences on human evolution in terms of biology, culture and also in terms of energetic cost and benefit. Thus, the aim of this chapter is to present the main aspects, stages and degrees of complexity of hominin-animal relationships during the Paleolithic of the old world. Animals have undoubtedly always represented a major food, technological and symbolism resource for hunter-gatherer groups.
Article
Full-text available
We present here the results of a 4-year environmental monitoring program at Ascunsă Cave (southwestern Romania) designed to help us understand how climate information is transferred through the karst system and archived by speleothems. The air temperature inside the cave is around 7 °C, with slight differences between the upper and lower parts of the main passage. CO2 concentrations in cave air have a seasonal signal, with summer minima and winter maxima. These might indicate the existence of an organic matter reservoir deep within the epikarst that continues to decompose over the winter, and CO2 concentrations are possibly modulated by seasonal differences in cave ventilation. The maximum values of CO2 show a rise after the summer of 2014, from around 2000 to about 3500 ppm, following a rise in surface temperature. Using two newly designed types of water–air equilibrators, we were able to determine the concentration of CO2 dissolved in drip water by measuring its concentration in the equilibrator headspace and then using Henry's law to calculate its concentration in water. This method opens the possibility of continuous data logging using infrared technology, without the need for costly and less reliable chemical determinations. The local meteoric water line (δ2H = 7.7 δ18O + 10.1), constructed using monthly aggregated rainfall samples, is similar to the global one, revealing the Atlantic as the strongly dominant vapor source. The deuterium excess values, as high as 17 ‰, indicate that precipitation has an important evaporative component, possibly given by moisture recycling over the European continent. The variability of stable isotopes in drip water is similar at all points inside the cave, suggesting that the monitored drip sites are draining a homogenous reservoir. Drip rates, as well as stable isotopes, indicate that the transfer time of water from the surface is on the order of a few days.
Article
Full-text available
Soot marks, witnesses of past human activities, can sometimes be noticed in concretions (speleothem, travertine, carbonated crust, etc.) formed in cavities. We demonstrate here that these deposits, generally ignored in archaeological studies, turned out to be a perfectly suitable material for micro-chronological study of hominin activities in a site. At the Grotte Mandrin (Mediterranean France), thousands of clastic fragments from the rock walls were found in every archaeological level of the shelter. Calcareous crusts containing soot deposits are recorded on some of their surfaces. They appear in thin section as thin black laminae. Microscopic observation of these crusts revealed that they kept track of many occupations. We show that is possible to link them with the archaeological units identified during the excavation. Minimum Number of Occupations (MNO) can be built out of these sooted crusts. MNO are usually high and attest to the cumulative nature of each archaeological unit. They are witnesses of each occupation of hominin groups in each archaeological level of the cave. This study also shows that, in Grotte Mandrin, a very short time separates the first Middle/Upper Paleolithic transitional groups' occupations from those of the last Mousterians. The research perspectives on soot deposits are diversified and raise the possibility of studying multiple aspects of past human life, and in this case, to rethink the Middle/Upper Paleolithic transition, with an unmatched temporal resolution. Sooted concretion analysis provides high temporal resolution archaeology. There is a real possibility of extending this study with chronological implications to cavities of all ages and areas.
Article
Full-text available
Significance Radiocarbon dating of Neanderthal remains recovered from Vindija Cave (Croatia) initially revealed surprisingly recent results: 28,000–29,000 B.P. This implied the remains could represent a late-surviving, refugial Neanderthal population and suggested they could have been responsible for producing some of the early Upper Paleolithic artefacts more usually produced by anatomically modern humans. This article presents revised radiocarbon dates of the human bones from this site obtained using a more robust purification method targeting the amino acid hydroxyproline. The data show that all the Neanderthal remains are from a much earlier period (>40,000 cal B.P.). These revised dates change our interpretation of this important site and demonstrate that the Vindija Neanderthals probably did not overlap temporally with early modern humans.
Article
Full-text available
Understanding the past dynamics of large-scale atmospheric systems is crucial for our knowledge of the palaeoclimate conditions in Europe. Southeastern Europe currently lies at the border between Atlantic, Mediterranean, and continental climate zones. Past changes in the relative influence of associated atmospheric systems must have been recorded in the region’s palaeoarchives. By comparing high-resolution grain-size, environmental magnetic and geochemical data from two loess-palaeosol sequences in the Lower Danube Basin with other Eurasian palaeorecords, we reconstructed past climatic patterns over Southeastern Europe and the related interaction of the prevailing large-scale circulation modes over Europe, especially during late Marine Isotope Stage 3 (40,000–27,000 years ago). We demonstrate that during this time interval, the intensification of the Siberian High had a crucial influence on European climate causing the more continental conditions over major parts of Europe, and a southwards shift of the Westerlies. Such a climatic and environmental change, combined with the Campanian Ignimbrite/Y-5 volcanic eruption, may have driven the Anatomically Modern Human dispersal towards Central and Western Europe, pointing to a corridor over the Eastern European Plain as an important pathway in their dispersal.
Article
Full-text available
The characterization of Last Glacial millennial-Timescale warming phases, known as interstadials or Dansgaard-Oeschger events, requires precise chronologies for the study of paleoclimate records. On the European continent, such chronologies are only available for several Last Glacial pollen and rare speleothem archives principally located in the Mediterranean domain. Farther north, in continental lowlands, numerous high-resolution records of loess and paleosols sequences show a consistent environmental response to stadial-interstadial cycles. However, the limited precision and accuracy of luminescence dating methods commonly used in loess deposits preclude exact correlations of paleosol horizons with Greenland interstadials. To overcome this problem, a radiocarbon dating protocol has been developed to date earthworm calcite granules from the reference loess sequence of Nussloch (Germany). Its application yields a consistent radiocarbon chronology of all soil horizons formed between 47 and 20 ka and unambiguously shows the correlation of every Greenland interstadial identified in isotope records with specific soil horizons. Furthermore, eight additional minor soil horizons dated between 27.5 and 21 ka only correlate with minor decreases in Greenland dust records. This dating strategy reveals the high sensitivity of loess paleoenvironments to Northern Hemisphere climate changes. A connection between loess sedimentation rate, Fennoscandian ice sheet dynamics, and sea level changes is proposed. The chronological improvements enabled by the radiocarbon "earthworm clock" thus strongly enhance our understanding of loess records to a better perception of the impact of Last Glacial climate changes on European paleoenvironments.
Article
Full-text available
Significance Last Glacial millennial-timescale warming phases well-recorded in Greenland ice cores are relevant across the Northern Hemisphere. However, dating limitations in loess deposits inhibited characterizing their impact on the European Great Plain. Here, the radiocarbon dating of a large set of earthworm calcite granule samples from the Nussloch reference loess sequence (Rhine Valley, Germany) led to a straightforward chronological distinction of all soil horizons. Resulting correlations with Greenland interstadials between 50 and 20 ka also revealed more complex climate dynamics than interpreted from Greenland δ ¹⁸ O records. This study is a fundamental contribution to understanding links between mid- and high-latitude climate changes and their spatial and temporal impact on paleoenvironments and prehistoric population settlement in Europe.
Article
Full-text available
This study presents new models on the origin, speed and mode of the wave-of-advance leading to the definitive occupation of Europe’s outskirts by Anatomically Modern Humans, during the Gravettian, between c. 37 and 30 ka ago. These models provide the estimation for possible demic dispersal routes for AMH at a stable spread rate of c. 0.7 km/year, with the likely origin in Central Europe at the site of Geissenklosterle in Germany and reaching all areas of the European landscape. The results imply that: 1. The arrival of the Gravettian populations into the far eastern European plains and to southern Iberia found regions with very low human occupation or even devoid of hominins; 2. Human demography was likely lower than previous estimates for the Upper Paleolithic; 3. The likely early AMH paths across Europe followed the European central plains and the Mediterranean coast to reach to the ends of the Italian and Iberian peninsulas.
Chapter
Izvorul Tăușoarelor Cave, which held for many years the record for the deepest cave in Romania, is renowned for its steeply descending galleries, tectonic-controlled speleogenesis, and a rich assemblage of sulfate minerals (gypsum, mirabilite, arcanite, bassanite, epsomite, konyaite, leonite, and syngenite) that form a variety of speleothems. The enigmatic spherical concretions, known as “Tăușoare balls,” formed simultaneously with the limestone bedrock during its early diagenesis. The existence of hundreds of bones of cave bear (Ursus spelaeus) and brown bear (Ursus arctos) suggests the two species cohabited during the Quaternary in this part of the country. The cave represents the hibernation site for four bat species.
Chapter
Ascunsă Cave is situated in the Mehedinţi Mountains and is part of the Isverna cave system. The cave was discovered in the late 1970s by members of the Focul Viu Caving Club, who surveyed the first 13.5 m of it. Since 2008, the “Emil Racoviţă” Institute of Speleology and the Underwater and Cave Exploration Group explored and surveyed 691 m of passages totaling 185 m of vertical development. Studies based on stalagmites from this cave revealed the climate evolutions during the last glacial cycle.
Article
The site of Isturitz is clearly important to discussions of the emergence and development of the Aurignacian. It bears a long stratigraphic sequence of this period and has benefited from recent excavation and analysis. In this paper we present 18 new AMS radiocarbon dates (Normand excavation), covering the majority of the Aurignacian sequence at this site. Our dating was aimed at addressing two key questions of this period (1) what is the date of occupation of each of the Aurignacian variants (Protoaurignacian, PA, and Early Aurignacian, EA) at Isturitz and (2) how do the dates of PA and EA occupation at Isturitz compare to those of other nearby sites? To achieve this we dated well-provenienced, species-identified, humanly-modified faunal remains from layers of each Aurignacian variant at Isturitz, most including an ultrafiltration step. We built a Bayesian model from these to determine start/end dates of each layer/industry at this site. We also compiled a list of all recently-dated, ultrafiltered/ABOx-SC, more carefully sampled, results from PA and EA layers in France, Italy and Spain to see where Isturitz fit into these groups. Results indicate that at Isturitz, the PA started at 42.8–41.3 modelled BP (95% confidence interval) and the EA at least as far back as 41.6–39.7 modelled BP (95% confidence interval). These are among the earliest dates for both of these industries in western Europe and it is one of the only sites to have multiple old dates. Our results and comparisons confirm that the PA was the earlier of the two variants, but show that the EA and later PA overlap statistically within the western European region (between sites). Thus the possibility that the two variants existed partially contemporaneously must remain open as a working hypothesis. Whether the EA was initiated as a result of HS4 is equivocal with the new data.