The repeated replacement model

Transcription

The repeated replacement model
Quaternary International 247 (2012) 38e49
Contents lists available at ScienceDirect
Quaternary International
journal homepage: www.elsevier.com/locate/quaint
The repeated replacement model e Rapid climate change and population
dynamics in Late Pleistocene Europe
Marcel Bradtmöller a, Andreas Pastoors a, Bernhard Weninger b, Gerd-Christian Weniger a, *
a
b
Neanderthal Museum, Talstrasse 300, 40822 Mettmann, Germany
University of Cologne, Institute for Prehistoric Archaeology, Weyertal 125, 50923 Cologne, Germany
a r t i c l e i n f o
a b s t r a c t
Article history:
Available online 19 October 2010
The disappearance of Neanderthals from the Palaeolithic record in Europe remains an enigma, even after
more than 150 years of research. This paper identifies Rapid Climate Change during the Glacial period as
a major factor that influences a variety of cultural, economic and demographic processes during the
European Palaeolithic. In particular, and in agreement with many previous authors, climatic deterioration
is put forward to explain multiple population breakdown during the European Palaeolithic, as well as to
explain corresponding major cultural changes. Taking the archaeological record of the Iberian Peninsula
as a case study, the Repeated Replacement Model (RRM) is proposed to explain population turnover in
Europe during the most extreme climatic phases of the Glacial, the occurrence of North Atlantic Heinrich
Events (HE). The strong aridity of the Mediterranean during HEs appears to have limited settlement
refugia to such an extreme extent that communication networks and cultural traditions broke down and
were subsequently reorganized under different socio-cultural conditions. The transition from the Middle
Palaeolithic to the Aurignacian during HE 4 is one of these cultural turnover periods, which saw the final
(macro-scale) extinction of Neanderthals and their widespread replacement by Anatomically Modern
Humans. More specifically, and recognizable by comparisons with other climatically extreme Glacial
periods (i.e. HE 3, and HE 2), the model excludes the survival of geographically wider (supra-regional)
human networks, but it does allow for (micro-scale) survival of scattered groups. From this model, some
kind of admixture between Neanderthals and incoming groups of modern humans would indeed have
been possible on a small scale. If this climatic scenario turns out to be correct, the most spectacular thing
about Neanderthal disappearance might actually lie in the seemingly unspectacular nature of the
processes involved.
Ó 2010 Elsevier Ltd and INQUA.
1. Introduction
Looking back on more than 150 years of research on Neanderthals means to look back on a number of highly controversial
debates, especially within the last two decades. A large number of
hypotheses and speculations concerning the transition from
Neanderthals to anatomically modern humans as well as the
second phenomena of technological transition from Middle to
Upper Palaeolithic techno-complexes have been put forward by
expert members of the various scientific disciplines. Despite this
plethora of opinions, the high variability of Late Pleistocene climatic
change in Europe is now widely acknowledged. This new palaeoclimatological understanding provides an important stimulus for
further archaeological discussions. Researchers must now seriously
allow for the possibility that the observed climatic and
* Corresponding author.
E-mail address: [email protected] (G.-C. Weniger).
1040-6182/$ e see front matter Ó 2010 Elsevier Ltd and INQUA.
doi:10.1016/j.quaint.2010.10.015
environmental changes were also influential in terms of cultural
variability. In support of climate forcing as major factor for cultural
change, see (e.g.) van Andel and Davies (2003), d’Errico and
Sanchez-Goñi (2003), Mellars (2006), Shea (2008), Finlayson and
Carrión (2007), and Sepulchre et al. (2007). There are alternative
proposals, which place less emphasis on the climatic background
for cultural change in the Palaeolithic (e.g. Tzedakis et al., 2007;
Roebroeks, 2008; Banks et al. 2008a).
Talking about Neanderthals means to talk about humans living
at the very periphery of the Pleistocene human world. These
humans represent pioneering populations living on Mediterranean
shores, just as at high Northern latitudes. In consequence e and
keeping in mind the increasing number of high-resolution climatic
records that demonstrate the extreme rapidity and amplitude of
past climatic changes e it must now be recognized that the Late
Pleistocene in Europe was a period during which human populations repeatedly experienced significant environmental
switches, which occurred rapidly on decadel time scales and in
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
many different geo-climatic regions (Clement and Peterson, 2008;
van Andel et al., 2003). During the Late Pleistocene, considerable
parts of Europe must have represented risky environments for
humans. Assuming this climatic and environmental variability had
major implications for Neanderthals, as well as for modern
humans. This paper combines environmental data with chronological data from archaeological sites in Iberia to develop a model
for cultural change in Late Pleistocene Europe.
2. Models for climate change
A brief look at Greenland climate curves between 50 ka and 20
ka calBP displays the high climatic variability in Europe during the
Glacial period (Fig. 1: Upper). The Dansgaard/Oeschger cycles of
sudden climatic amelioration (Greenland Interstadials: GIS) and
subsequent coolings (Greenland Stadials: GS) are well recorded in
North Atlantic air temperatures (v18O: Grootes et al., 1993) as well
as in the GISP2 glaciochemical record (e.g. nss [Kþ]: Mayewski et al.,
1997). Previous research on the European and Near Eastern Palaeolithic has placed strong emphasis on chronological applications of
the stable oxygen isotope (d18O) records from Greenland ice-cores.
The Greenland d18O-record has indeed been helpful, e.g. in the
initial construction of an extended Glacial 14C-age calibration curve
(Jöris and Weninger, 1998), but its direct archaeological applications are to some large extent restricted to the North Atlantic.
Indeed, for archaeological applications the Greenland d18O-record
has certain limitations. Perhaps most notably, in the d18O-record
there is no evidence for the existence of the Litte Ice Age (ca.
1600e1929 AD) but which shows up strongly in the GISP2 glaciochemical record (Mayewski et al., 2004). Although presently not
used by the archaeological community, the GISP2 chemical ion
records (Ca2þ, Mg2þ, Naþ, Cl, Kþ, Cl) appear to have an equally
wide applicability, and also some complementary use in archaeological applications, along with the d18O-record. Significantly, the
39
GISP2 glaciochemical record demonstrates the high variability of
Northern Hemispheric atmospheric circulation patterns during the
last 60 ky. Further, and also of considerable interest for archaeological research, is the recent discovery (Rohling et al., 2002) that
the severe coolings observed in eastern Mediterranean sea-surface
temperatures appear to be related to a quasi-periodic expansion
and intensification of the Siberian High. These severe coolings (here
termed ‘Rapid Climate Change’, RCC) can be seen to run systematically through the Last Glacial into the Holocene and up to
modern times. The occurrence of RCC periods are presently bestidentified in the GISP2 [Kþ] record, but also show up in other highresolution records (e.g. d13C, SufularCave, NW-Anatolia: Fleitmann
et al., 2009). In the present paper, in search of the expected
extreme impact of Heinrich Events (HE), the GISP2-[Kþ] record is
therefore included. To achieve better time-scale compatibility with
U/Th-Hulu-based calibrated 14C-ages cf. below, we have transferred
the Greenland GISP2-[Kþ]- record to the U/Th-Hulu time-scale.
Together (Fig. 1), these records demonstrate a long sequence of
changes in the strength of the North Atlantic Ocean circulation
system, as well as in Northern Hemispheric atmospheric circulation
patterns during the Glacial. Altogether, at least 12 major climate
cycles are recorded in Greenland ice-cores between 50 ka and 20 ka
calBP, including the Late Glacial Maximum (LGM). However, similar
cycles (e.g. Fig. 1, Lower) show up in marine cores taken from the
Alborean Sea, (e.g. Moreno et al., 2005; Jiménez-Espejo et al., 2007;
Fletcher and Sánchez Goñi, 2008) as well as the Iberian margin
(Roucoux et al., 2005; Naughton et al., 2007, 2009; Salgueiro et al.,
2010). These records provide important information as to corresponding changes of neighbouring terrestrial environments. It is
a major task for palaeoclimatologists and environmental
researchers to accurately synchronise the records from the
different geographic realms (ice-core, atmosphere, marine sourced)
and to transfer their information contents to the much less-well
documented terrestrial regimes. Today, in archaeological research,
it must be considered that such rapid and intensive climatic
changes had an important impact on geographic frontiers.
3. Human expansion during the Glacial
Fig. 1. Selected Climatic Records 50 kae20 ka calBP Upper: Greenland GISP2 stable
oxygen isotope (Grootes et al., 1993) as proxy for North Atlantic air temperature and
Greenland non-sea salt (nss) [Kþ] record (Mayewski et al., 1997) as proxy for the
strength of the Siberian High (Rohling et al., 2002), with GISP2-ages tuned to U/ThHulu ages (Wang et al., 2001) according to the Greenland GISP2Hulu time-scale
Weninger and Jöris (2008). Greenland stadial/interstadial numbers are according to
Johnsen et al., (1992). Lower: Sea-Surface Temperature (SST) in the W-Mediterranean
as derived from marine core MD952043 (Cacho et al., 2001) and Ice Surging on the
Iberian Atlantic Margin (Martrat et al., 2007).
As shown in Fig. 2, the northern limit of human expansion in
Europe prior to the Gravettian (i.e. before 30 ka calBP) is located
around 53 e55 N (van Andel et al., 2003; Roebroeks, 2008). In
regions further to the east, humans might have reached at that time
even higher latitudes (Pavlov et al., 2004). This expectation, that
human frontiers in Europe during the Late Pleistocene were
continually on the move, is supported by several investigations
(Weniger and Orschiedt, 2000; Bocquet-Appel and Demars, 2000;
Bocquet-Appel et al., 2005; Gamble, 1993; Housley et al., 1997;
Gamble et al., 2004). In the present paper, to achieve more
detailed insight into these pulsations, selected radiocarbon and
(adapted) environmental data from the Stage Three Project (van
Andel and Davies, 2003), which is aimed at simulating the
human frontier for selected extreme climatic scenarios (Barron and
Pollard, 2003), were used.
As a main simulation component, the modeling studies
provided by the Stage Three Project are based on the isoline for the
geographic distribution of annual/daily average snow cover. Davies
and Gollop (2003, 143) proposed, prior to 30 ka calBP, a preference
of humans to occupy regions with snow cover below 60 days per
year and an average tolerance up to 180 days a year. However, as
shown in Fig. 2, such minimum-maximum simulations document
strong geographic shifts of the isolines and, in consequence, it is
expected that in the transitions from Interstadials to Stadials the
human frontiers must have displayed some pronounced oscillating
effect. The distribution of snow cover isolines provides evidence
40
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
Fig. 2. Upper: Simulation of Snow Cover (in days) during Stadial conditions (Stage 3 Database, Barron et al., 2002). For Stadial conditions we use here the simulation for the LGM,
since the Stage 3 Projekt simulation of OIS 3 cold events consistently turned out too warm (Davies and Gollop, 2003, 135). Sea level reconstructions are taken from Lea et al., (2002).
Coastline shapes are taken from GEBCO_08 Grid, version 20091120,http://www.webco.net. Lower: Simulation of Snow Cover (in days) during Interstadial conditions (Stage 3
Database, Barron et al., 2002). Sea level reconstructions are taken from Lea et al., (2002). Coastline shapes are taken from the GEBCO_08 Grid, version 20091120,http://www.webco.
net.
that a number of regions in Europe, and most notably the Northern
Mediterranean, Southwest France and the circum Black Sea regions,
may all have been used as possible refuge areas for human populations. Similar configurations have been described for Western
Europe by Boquet-Appel and Demars (2000).
According to such modeling approaches, hunter gather populations in Europe outside of these refuge areas would have had
two possibilities to react to extreme climatic changes: bear up or
take flight. Gamble (1993) developed a three-phase model to
describe corresponding reaction patterns of hunter-gatherers
towards climate forcing. His model starts with a compaction
(“downturn”) phase followed by a refuge phase and a subsequent
expansion (“upturn”) phase. This model was further refined by
separating the upturn phase into a pioneering and residential phase
(Housley et al., 1997). It was then used to explain population variability during cold events around the LGM and GS 2 in Europe
(Gamble et al., 2004, 247). Similar to Gamble (1993) and Housley
et al. (1997), the authors believe that climate forcing indeed had
tremendous impact both on the organization of social networks
and on population density. Recently, Hulbin and Roebroeks (2009)
argued that the periodical abandonment of some areas in the
northern environments of Neanderthals supports a model of local
extinction rather than a habitat tracking model. From this point of
view, both models can be applied in parallel to the differentiation of
population events. Both provide useful descriptions of population
dynamics during the entire European Late Pleistocene. To further
understand the relation between population dynamics and climatic
variations, on this broader temporal and geographic level, a four
steps cascade model is proposed, based on the adaptive cycle
representation of societal changes (Fig. 3). Hence, although strongly
based on previous considerations, the cascade model goes a step
further to allow for extreme climatic variability, and especially for
the increasingly high temporal resolution of underlying climate
records, as are now firmly established for the Late Pleistocene in
Europe (Fig. 4).
The adaptive cycle model was generated from observations in
ecosystems (Holling, 2001; Holling and Gunderson, 2002) and
successfully adapted to human social-ecological systems (Abel
et al., 2006; Folke, 2006). It represents an ambitious attempt at
presenting a framework aimed at describing and comparing the
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
Fig. 3. Adaptive Cycle Model (after Holling, 2001; Walker et al., 2006).
internal dynamics of social-ecological systems (Holling, 2001). The
adaptive cycle model not only sheds light on situations of crises but
also on the subsequent phases of reorganization. The cascade
model integrates the well known process of systems changes
summed up in the adaptive cycle model with climate change as
a vital parameter, during all phases of change including an escalation of human reaction pattern. The basic assumption is that, over
time, the structures and functions of systems may change both as
a result of internal dynamics, but also due to external influences,
and most notably due to climatic instability. These changes (societal
and climatic) result in four characteristic phases (r-phase, K-phase,
U-phase and a-phase) which are defined by Walker et al. (2006, 2).
The r-phase is a moment of growth with high availability of
resources, the establishment of internal structures and connections, and of high resilience. During the K-phase (conservation
41
phase) the system growth slows down, and internal structures are
determined. The system is now typically less flexible and more
vulnerable to external disturbances. In the model of Walker et al.
(2006), the U-phase describes the crises, in which parts of the
system breakdown and numerous functional structures and
networks are subsequently reordered. The a-phase of the cycle
describes the reorganization of the system. In this phase, new
diversity is established, innovative strategies of resource exploitation are developed, technology and tool production are adapted,
and networks are reorganized, all of which creates new structures
and new contents for exchange between subsystems. The different
phases of the adaptive cycle may run on different time scales:
whereas the phases r and K may last for some long time periods,
crises and reorganization represent extremely rapid processes
(Redman and Kinzig, 2003).
Some features of the modes of reorganization described by
Walker et al. (2006) are especially important when adapted as
background for the cascade model. “During release and reorganization, the system is most vulnerable to change, because it is in
these phases that the effects of the linkages between the system of
interest and systems at other scales become more pronounced. [.]
Some resources in the system are depleted following the release
phase, whereas others are recombined, reused, and rebuilt.
Resources can also be acquired. The dynamics of these resources
and their interrelations determine how the system reorganizes.”
(Walker et al., 2006: 3-4). They summarized their analysis by
stating that multiple modes of reorganization are possible, and that
any new modes will be most apparent during the phases that
follow from the crises.
The proposed cascade model (Fig. 4) represents the hierarchic
succession of different modes of system reorganization (x-axis),
each of which is described in terms of climatic variability (y-axis).
External agents in form of release impacts lead to vital disturbances
of the socio-ecological system. These are used as markers to
Fig. 4. Archaeological Cascade Model.
42
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
describe the beginning of the release phase (U). The model focuses
on climatic instability, and hence the restricted tolerance of the
system towards climatic stress is used as the basic release factor. As
reaction of the system towards increasingly extreme climates, four
modes of societal reorganization (a) are defined: resistance, retreat,
micro-extinction, and population breakdown. The system may end
in macro-extinction, but also in subsequent immigration, or else
the system may recover due to actions of the scattered remainder of
the population. The primary mode of reorganization as reaction to
climate change is termed ‘resistance’. In this case, the subsistence
patterns can be described as a cultural adjustment within the same
territory. Increasing climatic instability leads to retreat into the
following reorganization mode. This corresponds to the observation that human groups may occasionally only abandon certain
parts of their settlement area. Due to accelerated climatic forcing,
a first reaction would be micro-extinction of peripheral groups. This
mode of reaction has been described by various researchers (e.g.
Stiner and Kuhn, 2006; Hulbin and Roebroeks, 2009). Finally, the
breakdown of the meta-population will lead to macro-extinction,
corresponding to the complete collapse of the cultural system.
Basically, minor disturbances can be understand as components of
the normal system development, considering the plausible
assumption that “periods of gradual change and periods of rapid
transition coexist and complement one another.” (Folke, 2006: 258)
In this version of the model, the optimal ability to react on climatic
instability is resistance; the worst case reaction is population
breakdown. Finally, in each mode of reorganization the adaptive
cycle is open to change in both directions (down and upwards) and
is therefore able to react towards increasing as well as decreasing
climatic instability, except for the irreversible breakdown of population (macro-extinction). Limitations of dating resolution
preclude strict exclusion of the possibility of the survival of a few
scattered groups in the definition of population breakdown.
Changes in social structure, settlement pattern/mobility or technology are indeed highly probable as reaction towards climatic
change e.g. during Greenland stadial/interstadial transitions. One
main problem in applying the research model outlined above is the
limited archaeological visibility of these changes. Only a very few
items in the fragmentary and coarse grained Late Pleistocene
archaeological archives will be sufficiently sensitive to the climatic
changes to allow the necessary model testing. The majority of
palaeolithic archives have only very low potential for detection of
cultural changes. Nevertheless, there is the possibility that the
different modes of reorganization may be identifiable as variations of
techno-complexes. Under this assumption, which remains to be
tested, a complete change in techno-complex would be indicative for
a complete population breakdown (e.g. Blackwell and Buck, 2003).
As mentioned above, the climatic records for the period
between 50 ka and 25 ka calBP (Fig. 1) show approximately 12
major climate switches. To be consequent in model application,
prior to the testing phase, it must apriori be assumed that each of
the 12 stadials would have forced human population to reduce their
settlement area. Further, by apriori application of the cascade
model, each of the observed 12 periods of rapid climate change
would have triggered some kind of cultural change, even if different
each time, by the hunter-gatherer populations. However,
comparing the number (N ¼ 4) of generally recognized major
cultural changes (i.e. Middle Palaeolithic, Transitional Industries,
Aurignacian, Gravettian, Solutrean) with the number (N ¼ 12) of
stadial/interstadial switches in the time frame under study, it
becomes readily apparent that these numbers do not correspond at
all well. This disagreement applies even allowing for variations in
the established archaeological definitions of cultural variability that
is largely based on variations in observed tool technologies. Clearly,
therefore, even prior to its first application, the cascade model of
cultural reaction towards climate change needs some major
adjustments. Notably, why are the number of major cultural
changes so strikingly different from the number of stadial-interstadial switches?
Four possible explanations are proposed:
The temporal resolution of the archaeological record, that is
largely based on radiocarbon dating, is too low to detect (rapid)
climate impact on cultural systems
The archaeological definitions of cultural techno-complexes
are largely insensitive to underlying rapid societal changes.
The climatic impact of an average stadial-interstadial switch
was not strong enough to trigger the complete reaction chain,
as envisaged in the cascade model.
The resilience of human groups within their cultural system
was high enough, in many cases, to buffer the climatic impact.
Allowing for such considerations e and as an alternative to the
clearly possible alternative that the cascade model does not properly capture the social realities of Palaeolithic hunter-gatherers e it
is expected that cultural change in Late Pleistocene Europe needs
a very robust climatic signal indeed before it becomes visible in the
coarse grained archaeological record. The initial test is whether (at
least) the major cultural switches in the Palaeolithic sequence show
any indication of corresponding to (at least) the strongest climatic
switches that occurred in the period under study. Clearly, the
largest cultural switches are given for complete breakdown of
hunter-gatherer populations; the largest climatic switches are
given for Greenland stadials which show major ice-rafting events in
the North Atlantic (Heinrich Events). In the following methodological case study we test this association of human population
breakdown with Heinrich Events using the archaeological 14Crecord of the Iberian Peninsula.
4. Iberian case study
The Iberian Peninsula has a reasonably well-studied archaeological record and numerous sets of high-resolution multi-proxy
palaeoclimatological data that display good synchronicity between
terrestrial and marine climatic sequences, and which are consistent
in showing abrupt climate changes (Moreno et al. in press). The
sequence of palaeolithic techno-complexes (Fig. 5) ranges from
the Mousterian, through the Aurignacian and Gravettian, up to the
Solutrean. Altogether, some 200 different archaeological sites can
be associated with the time-period under study (i.e. 50e20 ka
calBP, according to available radiocarbon dates). The one disadvantage of the Iberian setting is the weak evidence for the Châtelperronian, which should represent the so-called transitional
industries in Southwestern Europe. The existence of the Châtelperronian is highly controversial (Maíllo-Fernández, 2007).
For the purposes of this study, with compiled published 14C-data
from Palaeolithic sites of Iberia, the aim is to establish a radiometric
chronology on the absolute time-scale, which can be directly
compared with available high-resolution palaeoclimate records.
For this purpose, the data were calibrated with CalPal-software,
using the CalPal-2007-Hulu 14C-age calibration curve. For all
practical purposes, this 14C-age calibration produces results that are
identical to the recently established and now internationally recommended INTCAL09-calibration (Reimer et al., 2009). In consequence, the resulting archaeological chronology can be compared
with the palaeoclimatic sequence. Cultural changes from one
techno-complex to the next are located during Greenland stadials
GS3, GS5 and GS9. These specific stadials coincide roughly with
Heinrich Events (HE) with numbers HE2, HE3 and HE4. The most
clearly detectable (major) cultural changes appear linked only to
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
43
Fig. 5. 14C sequence Iberia and climatic record. Upper: Summed calibrated probability of 14C-ages for cultural sequences on the Iberian Peninsula and France. From top to bottom:
Solutrean (N ¼ 77, Spain and Portugal), Gravettian (N ¼ 70, Portugal and Spain), Early Upper Palaeolithic (EUP, N ¼ 81, Spain and Portugal), Middle Palaeolithic (MP, N ¼ 101, Spain
and Portugal). Age calibration using CalPal-2007-Hulu curve (cf. Fig. 7). HE1-HE4 e Heinrich Events are shown shaded with ages according to Fletcher and Sanchez Goni (2008),
transferred to Greenland GISP2Hulu time-scale (Weninger and Jöris, 2008). Filtered and supplemented 14C-data assembled from the INQUA-database (Vermeersch, 2009). Note the
unreliability of MP and EUP 14C-data (see text) and the approximate synchronicity of major changes in culturally-defined 14C-data assemblages both with (certain) North Atlantic ice
surges as well as with periods that show significant expansion and strengthening of the Siberian High, as well as with Heinrich Events HE1-HE3.Lower: Greenland non-sea salt (nss)
[Kþ] record (Mayewski et al., 1997) as proxy for the strength of the Siberian High (Rohling et al., 2002) with GISP2-age model tuned to the HuluCave U/Th-ages (Wang et al., 2001) by
Weninger and Jöris (2008). Greenland stadial/interstadial numbering according to Johnsen et al., (1992) in comparison to Ice Surging on the Iberian Atlantic Margin based on C37:4
alkenone measurements from core MD01-2444 (Martrat et al., 2007).
those Greenland stadials that are associated with Heinrich Events
(Fig. 5). However, even a brief glance at the available radiocarbon
data from the Iberian Peninsula, when classified according to the
major techno-complexes, demonstrates that the different cultural
entities show a significant overlap in time. In previous research this
temporal overlap of the 14C-data has often been used as an argument in support of a corresponding overlapping of underlying
archaeological culture, and often also used to demonstrate contact
between human populations. In addition, the overlap of 14C-data
has been used as evidence for the assumed long coexistence of the
Mousterian and Aurignacian (e.g. Carrión et al., 2008; Finlayson and
Carrión, 2007; Zilhão, 2006; Banks et al. 2008a). However, the
’historical’ probability of such assumed cultural patterns has been
seriously contested, since this interpretation does not correctly
allow for many problematic aspects of the applied radiocarbon
dating method e.g. large-scale sample contamination and stratigraphic reworking (Weninger and Jöris, 2008). Not unexpectedly,
therefore, in the transition from the Mousterian to the Aurignacian,
and to some less extent from the Aurignacian to the Gravettian, the
available (unfiltered) radiocarbon data from the Iberian Peninsula
display some clearly unacceptable temporal overlap of these
techno-complexes. The given overlap of 14C-data does not correspond at all well to the known succession of the cultural technocomplexes based on the stratigraphic observations made at
a multitude of different sites. For example, based solely on radiocarbon ages, we may (wrongly) assume an overlapping of the
Mousterian and Aurignacian by more than 10000 years, as shown
in Fig. 5.
An inherently more reliable chronological background to the
questions at stakeethe reaction of human populations to rapid
climate change during the Late Pleistocene period e can be
provided by detailed crosschecking of given radiocarbon data
against the stratigraphical evidence. The results, as provided e.g. by
Jöris et al. (2003) and Zilhão (2006), are in clear contradiction to the
widely assumed temporal interstratification of techno-complexes.
Significantly, there is e up to now e not one single archaeological
site known from anywhere on the Iberian Peninsula, at which the
postulated interstratification between Mousterian and Aurignacian, between Aurignacian and Gravettian, or Gravettian and
Solutrean levels has survived this critical approach. Indeed, this
applies elsewhere in Europe. Combining the critically filtered
radiocarbon record with the equally critically filtered stratigraphic
record, the basic conclusion is that the previously postulated
chronological overlap of Upper Palaeolithic techno-complexes has
been falsified at all sites. It remains to be established, by new
excavations and with enhanced radiocarbon dating resolution,
whether these present observations can be confirmed in all regions
of Europe.
In the study region, the transition from the final Mousterian to
a disputed Châtelperronian and the Aurignacian now allows for
more detailed discussion. While Northern Iberia displays an
archaeological record which is similar to that of Western and
Central Europe, recent investigations have argued for the delayed
existence of Mousterian industries in Southern Iberia, up to at least
25 ka calBP (Delson and Harvati, 2006; Zilhão and Pettitt, 2006).
Other authors argue for a shorter coexistence of Neanderthals and
AMH up to w 30 ka calBP (e.g. Vega Toscana, 2005; Maroto et al.,
2005; Vaquero, 2006). In a recent paper, Zilhão et al. (2010) has
assembled documents for sites in Southern Iberia that may be of
use in addressing the question of continuity or discontinuity
between Neanderthals and AMH. Zilhão et al. (2010) argue that
traces of Aurignacian settlement can only be seen in a later phase of
the Aurignacian, dating after 36 ka calBP, and that Middle Palaeolithic settlement came to an end prior to this age limit. Following
44
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
this argument, a possible coexistence of Middle Palaeolithic
complexes and Aurignacian in Iberia appears restricted to a very
small time frame (in the order of max 1000 years). To date, the
presence of a late Aurignacian in Southern Iberia is based on a small
number of sites, that display only limited excavated surface areas
and have supplied only scarce and scattered finds. However, the
radiocarbon record in support of these hypotheses is still now
highly limited. There are only four sites at stake, and they display
highly heterogeneous radiocarbon data, in addition to one site
which has a TL-date (Zilhao et al., 2010; Fig. 12).
The situation on the Iberian Peninsula is not unusual. The
tremendous obstacle of how to obtain reliable 14C-ages from
organic samples older than 30 ka has been stressed by numerous
authors (e.g. Maroto et al., 2005; Weninger and Jöris, 2008), but the
dating problems are still unfortunately often neglected in the
archaeological community. As shown by recent 14C-radiometric
research, the different samples (e.g. charcoal and bone) can react in
many different ways, such that dating results may spread over
several thousand years, even under quasi-ideal technical conditions. New laboratory methods for chemical pre-treatment (e.g.
ultrafiltration and enhanced Acid-Base-Acid methods) have
demonstrated that further reductions in sample contamination are
feasible (e.g. Higham et al., 2006; Brock et al., 2007; Hüls et al.,
2007). Notwithstanding these recent advances in chemical processing, at the present state-of-research there is still no guarantee
that measured 14C-ages beyond the critical limit of w30 ka 14C-BP
have some reliable chronological meaning.
The already difficult situation is further complicated by the
existence of strong secular variations of Glacial atmospheric 14Clevels which must be accounted for in the construction of 14C-based
archaeological (calendric time-scale) chronologies. Previously, due
to unresolved differences between data sets (van der Plicht, 2000;
Bronk Ramsey et al., 2006) the INTCAL04-group (Reimer et al.,
2004) refrained from recommendation of 14C-age calibration
beyond 16 ka 14C-BP. More recently, the INTCAL09-group (Reimer
et al., 2009) has ratified a Glacial calibration curve, which is
extended to include 14C-ages beyond this limit. As shown in Fig. 6,
the new INTCAL09-curve is practically identical to the CalPal-2007Hulu calibration (Weninger and Jöris, 2008). This is not unexpected,
since both curves are to some large extent based on the same (Hulu
U/Th-tuned) Cariaco basin marine data (Hughen et al., 2006).
Despite this agreement, there are remaining differences in curve
construction, including data selection and curve smoothness. Most
important, however, in the CalPal-2007-Hulu-based calibration
(Weninger and Jöris, 2008) the U/Th-tuning is applied to all
archaeologically relevant age models (i.e. not only Greenland icecores but comparisons with other climate proxies e.g. stalagmites)
To conclude, then, although further refinements to Glacial 14C-age
calibration are pending, there is now at least reasonable international agreement on which age models and data to use in Glacial
14
C-age calibration. For ages beyond 30 ka 14C-BP the main problem
is thus reduced to the known difficulties in how to obtain uncontaminated ages on samples (bone and charcoal) from the archaeological record. Going further back in time (beyond the limits of
radiocarbon dating), there are other dating inconsistencies to be
solved e.g. between Greenland ice-core age models (Anderson
et al., 2007) and the U/Th-data derived from stalagmite proxies
(e.g. Chiu et al., 2007; Fleitmann et al., 2009), However, these are
generally well-covered by given quantitative errors and such
differences are therefore not problematic, but rather provide
a challenge for further refinement of science-based archaeological
chronologies.
Backed by such considerations towards the recent history and
reliability of present time-scales, the overall working approach to
these problems is, first, to treat archaeological radiocarbon dates
Fig. 6. Comparison of CalPal-2007-Hulu and INTCAL09 Glacial 14C-Age Calibration.
Upper: Composite U/Th-based Glacial 14C-age calibration showing data (1s error
bars) of Fairbanks et al. (2005), Bard et al., 2004, Voelker et al., 2000, and Hughen et al.,
2006 used in the construction of CalPal-2007Hulu calibration curve (Weninger and Jöris,
2008) in the time-window 50.0e10.0 ka calBP/46.0e9.0 ka 14C-BP, in comparison to
INTCAL09-calibration curve (Reimer et al., 2009). Used for comparison only (i.,e. not
included in CalPal-2007Hulu construction) are the Arabian Speleothem data (van der
Plicht et al., 2004). Lower: Greenland GISP2 stable oxygen isotope (Grootes et al.,
1993) and Greenland non-sea salt [Kþ] record (Mayewski et al., 1997), with GISP2ages according to the Greenland GISP2Hulu time-scale (Weninger and Jöris (2008).
older than 30 ka generally as minimum age-estimates. This caution
also applies to the use of presently rather broadly defined (both in
stratigraphic and cultural terms) archaeological 14C-databases available for the European Palaeolithic (e.g. as provided with the CalPalprogramhttp://www.calpal.deor the Radiocarbon Palaeo-lithic Database Europe v.10 of Vermeersch (2009); cf. http://ees.kuleuven.be/
geography/projects/14c-palaeolithic/download/).
An intersite comparison of cultural sequences for any area as
wide as the Iberian Pensinsula will indeed have to address the
important question of differentiating between the (true) cultural
and the (potentially distorted) site interstratifications. Although
a temporal coexistence of Middle Palaeolithic and Upper Palaeolithic techno-complexes over 5000 years and even more is indicated by radiocarbon dating, for many regions of Europe, up to
present (and based on modern excavations) no reliable interstratifications of these techno-complexes have been recorded in a single
site, perhaps with one exception (Buran-Kaya, Krim) (Chabai et al.,
2004).
A second conclusion is that, for the time-being, the stratigraphic
method still remains by far the most reliable chronological tool for
archaeological research. Indeed, in strong contrast to the presently
highly error-prone natural scientific dates, the stratigraphic
method is still today capable of providing some clear indication of
the true temporal succession of cultural complexes. Importantly,
concerning the transition from Middle to Upper Palaeolithic in
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
Fig. 7. Dansgaard-Oeschger-Cycle and Adaptive Cycle Model.
Iberia by stratigraphic analysis, different types of stratigraphies can
be distinguished:
Sites with rich Middle Palaeolithic occupation sequence typically come to an end with settlement discontinuity into the Upper
Palaeolithic. At some, reoccupation - if at all - restarts later in the
Gravettian or Solutrean, or even Neolithic, although natural sedimentation processes continue.
Sites with a continuous settlement from the Middle Palaeolithic to the Upper Palaeolithic sequence. At some, a hiatus of
sedimentation or sterile layers are recorded between the final
Middle Palaeolithic and the early Upper Palaeolithic.
Sites with an onset of occupation in the Upper Palaeolithic
only; underlying sediments are recorded but are archaeological
sterile.
These patterns clearly express different settlement histories of
sites and regions within Iberia. In combination with detailed sedimentological analysis the distribution of this stratigraphical
patterning could open new venues for further research independent from radiocarbon chronology.
5. Discussion
Typically, the main cultural changes in Iberia are linked to
Greenland stadials that are combined with one of the Heinrich
Events. Therefore, the possibility that differences between standard
Greenland stadials and those associated with a Heinrich Event may
exist must be discussed. The standard scheme of a DO-Event
(Ganopolski and Rahmstorf, 2001) is rapid climatic amelioration
followed by an extended cooling phase and subsequent restart of
the cycle (Fig. 7 upper). Combined with the adaptive cycle, this
process is translated into the four steps model of rapid reorganization, growth, conservation, release and subsequent reorganization (Fig. 7 lower). For the Iberian Peninsula, several
reconstructions have suggested an especially strong environmental
impact of HE which is more severe than during standard G-Stadials.
For example, in the western part of the peninsula a severe decline of
45
arboreal plant cover during HE is recorded (Roucoux et al., 2005) as
well as decreasing temperatures and increasing aridity (Naughton
et al., 2009). In southern Iberia, several environmental reconstructions show strong signs of severe aridity both along the
southern coast as well as on the Mediterranean coastal sections
(Sepulchre et al., 2007, d’Errico et al., 2006; Jiménez-Espejo et al.,
2007; Fletcher and Sánchez Goñi, 2008; Vegas et al., 2009).
Reconstruction of rainfall patterns show that during HE 4 annual
precipitation was below 100 mm/y, in parallel to a significant
reduction in vegetation-cover to less than 25% of modern values
(Sepulchre et al., 2007, 286). These observations provide strong
evidence that the key to understanding the unusually strong
impact of HE on hunter-gatherer groups recognized as change in
lithic assemblages is to be sought in widespread desertification.
This is especially clear in the pollen spectra from several marine
cores e.g. from the Alborean Sea (Jiménez-Espejo et al., 2007;
Fletcher and Sánchez Goñi, 2008) and in the Fuentilljeo core
(Vegas et al., 2009). The North Atlantic Heinrich Events clearly
coincide with periods of strong aridity on the Iberian Peninsula.
Similar observations are available from the Eastern Mediterranean
(Tzedakis et al., 2004), as well as from the Levant, where the water
levels of Lake Lisan display an extreme decline that also appears
linked to HE (Bartov et al., 2003). Combining these results, the
unusually strong environmental impact of stadial conditions associated with HE must have transformed the Mediterranean, at times,
from a refugial zone with milder climate into the opposite kind of
human landscape, namely a high-risk environment with semi-arid
climate. Such rapid climatic switches, associated with HE all around
the Mediterranean, must have had tremendous impact on settlement patterns for hunter-gatherers all over Europe, including
Turkey and the Near East. This applies especially in view of human
migration patterns between the different ecologies. On the Atlantic
Coast of Iberia and Southwestern France, ice-rafting suggests that
ice probably had a direct impact on the climate and the environment of the coastal areas and their adjacent hinterland.
In the course of G-stadials, the already reduced settlement area
(with associated local or regional micro extinctions) would have
been further minimized during HE, to an extent that environmental
stress finally led to macro-extinction on a supra-regional level and
thus, ultimately, a complete breakdown of the meta-population.
However, in contrast to similar climatic arguments put forward by
Sepulchre et al. (2007), aridity during HE 4 did not support, nor
does it explain the assumed (based on 14C-interpretation) longer
survival of Neanderthals in Southern Iberia, but would rather have
produced the opposite, namely a breakdown of the Neanderthal
population in that area. This conclusion is further supported by the
observation that immigrant populations of Aurignacian tradition
avoided southern Iberia for some long time, before they finally also
colonized the south e if at all.
The climatically-deterred expansion of Aurignacian settlement
itself came to an end during HE 3, with yet another population
breakdown and subsequent recolonization of the Iberian Peninsula
by Gravettian populations. This clear break between Aurignacian
and Gravettian assemblages is generally accepted for most of
Europe. As an exception, for the Swabian Alb the possibility of an insitu transition between both techno-complexes based on a small
sample of backed bladelet fragments in late Aurignacian levels and
a series of early radiocarbon dates for Gravettian levels has been
discussed (Conard and Moreau, 2004; Moreau, 2009). However,
this interpretation is rejected by Verpoorte (2005) and Svoboda
(2007) for various reasons. The main problem is the stratigraphical situation at key sites such as Geissenklösterle or Hohle
Fels which display strong cryoturbation activity. In the case of the
Gravettian at Geißenklösterle, refitting of stone tools provides
evidence that artefacts from one single occupation layer was now
46
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
Fig. 8. Repeated Replacement Model.
split up by gelifraction into six different sublevels (Moreau, 2009). It
may be further assumed that the documented vertical movement
and mixing of artifacts from different levels goes in parallel with
movement of corresponding sediments. Concerning the (supposedly) early radiocarbon dates from the Swabian Gravettian, it would
therefore be argued that similar processes cannot be excluded for
the sites in Northern Iberia, some of which show some unexpectedly early dates (Fig. 5). In conclusion, in the analysis of the 14C-data
distribution from the Iberian Peninsula, even seemingly major
temporal structures of the data must be interpreted as simply
representing noise.
A further turnover of human populations in Iberia and Europe
can be inferred for the period of HE 2 with the Solutrean (e.g.
Gamble et al., 2004; Banks et al. 2008a,b). However, this time the
release phase was not accompanied by complete breakdown of
population. Solutrean Groups in Western Europe apparently reorganized their social system, and no immigration is recognized. In
Western Europe, a new cultural tradition was established while in
Eastern Europe, the Gravettian tradition continued in its transferred form as Epigravettian (Banks et al., 2009). This was the first
time that in parts of Europe an HE was apparently buffered by the
cultural system and it was only Central Europe that was abandoned
by hunter-gatherer groups. For the first time, it appears, the cultural
legacy of European hunter-gatherers was obviously split up into
a Western and an Eastern World.
Although this proposed pattern of human population turnover
and cultural change is in accordance with available archaeological
14
C-data for the younger HE periods (Fig. 5), due to prevailing
inconsistencies of the radiocarbon method the corresponding
association of cultural switch and climate change during HE 4
although highly probable remains unclear. From a stratigraphical
point of view, HE 4 follows in sequences from Southern Italy and
the Balkan immediately on the Campagnian Imgnimbrite tephra
(CI) (Fedele et al., 2008; Hoffecker et al., 2008). At a few sites the CI
seals so-called transitional industries. Although redepositional
phenomena cannot be excluded this might indicate a cultural
change before HE 4. As an alternative scenario, it might be argued
that HE 5 resulted in the development of the so-called transitional
industries that came to an end during HE 4. For the Near East, Shea
(2008) proposes a similar association of climate forcing and population turnover during HE 5 and HE 4 and also comment on
chronological uncertainties. An additional methodological
complication may be the shortness of Heinrich Events: Roche et al.
(2004) estimate that HE 4 had a duration of 250 years only.
However, there is a question of how to define (and identify) HE 4 in
different archives. Most recently, Fletcher and Sánchez Goñi (2008)
have proposed data-oriented methods that are promising for
further (high-resolution) sub-division of HE based on marinearchived vegetation records from the Alborean Sea. To detect such
short termed events within the archaeological sequences is a real
challenge. However, the same is true for many other aspects of the
combined 14C-U/Th-radiometric and archaeological chronologies.
6. Conclusion
Rapid Climate Change during the Glacial period was a major
factor that influences a variety of cultural, economic and demographic processes during the European Palaeolithic. In particular,
and in agreement with many previous authors, climatic deterioration is put forward to explain multiple population breakdown
during the European Palaeolithic, as well as to explain corresponding major cultural changes. In the Repeated Replacement
Model (RRM) proposed here (Fig. 8), the most extreme climatic
phases of the Glacial are identified by the occurrence of North
Atlantic Heinrich Events (HE) and which are further interpreted as
representing the main climatic drivers for population turnover. The
strong aridity of the Mediterranean during HEs appears to have
limited settlement areas to such an extreme extent that communication networks and cultural traditions broke down and were
subsequently reorganized under different socio-cultural conditions. The transition from the Middle Palaeolithic to the Aurignacian during HE 4 is one of these cultural turnover periods, which
saw the final (macro-scale) extinction of Neanderthals and their
widespread replacement by Anatomically Modern Humans. More
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
specifically, and best recognizable by comparisons with other
climatically extreme Glacial periods (i.e. HE 3, and HE 2), the model
excludes the survival of geographically wider (supra-regional)
human networks, but it does allow for (micro-scale) survival of
scattered groups. Nevertheless, it can be inferred e less from the
data than from the modeling approach e that some kind of
admixture between Neanderthals and incoming groups of modern
humans would indeed have been possible, if only on a small scale. A
long-delayed survival of Neanderthals in Southern Iberia for many
thousands of years, as proposed by many previous researchers, can
be excluded on the base of erroneously interpreted radiocarbon
ages. Recent results of Neanderthal genome analysis stress the
possibility that a small amount of interbreeding between Neanderthals and modern humans may indeed have occurred, but
which appears to have taken place mainly in the Near East. Later
contacts in Europe are excluded to date (Green et al., 2010).
After Neanderthal macro-extinction during HE 4, the following
Aurignacian settlement system itself broke down during HE 3,
showing the same kind of macro-extinction as during HE 4
(Stringer et al., 2003). Europe and Iberia then saw the immigration
of a new population and the setting of again new cultural traditions
that are broadly classified as Gravettian. The morphology of
Gravettian populations shows a very specific and homogenous
pattern in Pleistocene Europe (Churchill et al., 2000), and this
supports the idea of immigration.
The Gravettian techno-complex is replaced after HE 2 in Iberia
by the Solutrean. This change is different from previous population
turnovers for two reasons:
First, the cultural continuum in Europe became more rugged.
Whereas in Iberia a clear succession is recognized, in Central and
Eastern Europe the situation is less clear. The end of the Gravettian
complex saw for the first time a separation of the European cultural
landscape into a western and an eastern area. In Eastern Europe, an
Epigravettian, although still poorly defined in terms of technology
and chronology, succeeds the Gravettian.
A second difference in comparison to previous cultural turnovers is the widely accepted evidence for a hiatus of human occupation in Central Europe after the Gravettian. The existence of such
a hiatus was proposed more than 20 years ago (Weniger, 1989;
Housley et al., 1997), although its intensity is still under discussion (Street and Terberger, 1999; Terberger and Street, 2002). This
clearly visible regional population breakdown occurs during the
Late Glacial Maximum. From the modeling perspective, this hiatus
has such high visibility in the archaeological record simply because
the LGM has such a long duration (many millennia). In comparison
to the LGM, the HEs represent much shorter periods of climate
deterioration. This readily explains why the HEs are inherently so
much more difficult to recognize in the archaeological record.
Such features provide an enhanced understanding of another
pattern of cultural change. Immigration of new human groups from
Western Asian after the Gravettian as assumed for previous turnovers can be ruled out. As would follow from the Repeated
Replacement Model (RRM), the survival of some scattered human
groups would be expected, at least in certain places of Western
Europe. This survival would only have been possible if cultural
adaptation to the harsh environmental conditions in Europe since
the beginning of the Gravettian had taken place. However, it may
also be expected that during the LGM such abilities would not have
been sufficiently strong to cope with the worst climate conditions
in Central Europe between the Scandinavian and the Alpine ice
shields.
The reorganization of human groups after the Gravettian that
gave rise to the Solutrean in Iberia might also have been influenced
by a migration process. For decades (e.g. Pericot, 1942; Debénath,
1994), cultural input from the Aterian of North Africa up to an
47
immigration of humans has been discussed, albeit controversially.
Recently, Tiffagom (2005) provided new input to this discussion, by
proposing in addition to the well known existence of typological
similarities (stemmed points and bifacial retouch) technological
similarities (Levallois technique) as well between the Aterian and
the Solutrean in Southern Spain. This position has been contradicted by Alcaraz and Castaño (2007). To decide on such questions,
further research on both sides of the Mediterranean is crucial.
Acknowledgements
This paper was developed within the framework of the CRC 806
“Our Way to Europe” supported by the Deutsche Forschungsgemeinschaft (DFG). We thank two anonymous reviewers
for their helpful comments.
References
Abel, N., Cumming, D.H.M., Anderies, J.M., 2006. Collapse and reorganization in
social-ecological systems: questions, some ideas, and policy implications.
Ecology and Society 11 (1), 17. http://www.ecologyandsociety.org/vol11/iss1/
art17/.
Alcaraz, Castaño, M., 2007. El Ateriense del Norte de África y el Solutrense peninsular: contactos transgibraltareños en el Pleistoceno Superior? Munibe 58,
101e126.
Andersen, K.K., Bigler, M., Clausen, H.B., Dahl-Jensen, D., Johnsen, S.J.,
Rasmussen, S.O., Seierstad, I., Steffensen, J.P., Svensson, A., Vinther, B.M.,
Davies, S.M., Muscheler, R., Parrenin, F., Röthlisberger, R., 2007. A 60 000 year
Greenland stratigraphic ice core chronology. Clim. Past Discuss 3, 1235e1260.
Banks, W.E., d’Errico, F., Townsend Peterson, A., Kageyama, M., Sima., A., SánchezGoñi, M.-F., 2008a. Neanderthal extinction by competitive exclusion. PLoS ONE
3 (12), e3972. doi:10.1371/journal.pone.0003972.
Banks, W.E., d’Errico, F., Peterson, A.T., Vanhaeren, M., Kageyama, M., Sepulchre, P.,
Ramsteim, G., Jost, A., Lunt, D., 2008b. Human ecological niches and ranges
during the LGM in Europe derived from an application of eco-cultural niche
modeling. Journal of Archaeological Science 35, 481e491.
Bard, E., Menot-Combes, G., Rostek, F., 2004. Present status of radiocarbon calibration and comparison records based on Polynesian Corals and Iberian margin
sediments. Radiocarbon vol. 46 (3), 1189e1202.
Barron, E., Pollard, D., 2002. High-Resolution climate simulations of oxygen isotope
Stage 3 in Europe. Quaternary Research 58 (3), 296e309.
Bartov, Y., Goldstein, S.L., Stein, M., Enzel, Y., 2003. Catastrophic arid episodes in the
eastern Mediterranean linked with the North Atlantic Heinrich events. Geology
31, 439e442.
Blackwell, P.G., Buck, C.E., 2003. The Late Glacial human reoccupation of north-western
Europe: new approaches to space-time modeling. Antiquity 77, 232e240.
Bocquet-Appel, J.-P., Demars, P.-Y., Noiret, L., Dobrowsky, D., 2005. Estimates of
Upper Palaeolithic meta-population size in Europe from archaeological data.
Journal of Archaeological Science 32, 1656e1668.
Boquet-Appel, J.-P., Demars, P.-Y., 2000. Population kinetics in the Upper Palaeolithic in Western Europe. Journal of Archaeological Science 27, 551e570.
Brock, F., Bronk Ramsey, C., Higham, T., 2007. Quality assurance of ultrafiltered bone
dating. Radiocarbon 49 (2), 187e192.
Bronk Ramsey, C., Buck, C.E., Manning, S.W., Reimer, P., van der Plicht, H., 2006.
Developments in radiocarbon calibration for archaeology. Antiquity 80,
783e798.
Cacho, I., Grimalt, J.O., Canals, M., Sbaffi, L., Shackleton, N.J., Schoenfeld, J., Zahn, R.,
2001. Variability of the western Mediterranean Sea surface temperature during
the last 25,000 years and its connection with the Northern Hemisphere climatic
changes. Paleoceanography 16, 40e52.
Carrión, J.S., Finlayson, C., Fernández, S., Finlayson, G., Allué, E., López-Sáez, J.A.,
López-García, P., Gil-Romera, G., Bailey, G., González-Sampériz, P., 2008. A
coastal reservoir of biodiversity for Late Pleistocene human populations:
palaeoecological investigations in Gorham’s Cave (Gibraltar) in the context of
the Iberian Peninsula. Quaternary Science Reviews 27, 2118e2135.
Chabai, V., Monigal, K., Marks, A. (Eds.), 2004. The Middle Paleolithic and Early
Upper Paleolithic of Eastern Crimea. ERAUL, Liege.
Chiu, T.-C, Fairbanks, R.G, Cao, L., Mortlock, R.A, 2007. Analysis of the atmospheric
14
C record spanning the past 50,000 years derived from high-precision
230
Th/234U/238U and 231Pa/235U and 14C dates on fossil corals. Quaternary
Science Reviews 26, 18e36.
Churchill, S.E., Formicola, V., Holliday, T.W., Holt, B.M., Schuman, B.A., 2000. The
Upper Palaeolithic Population of Europe in an Evolutionary Perspective. In:
Roebroeks, W., Mussi, M., Svoboda, J., Fennema, K. (Eds.). Hunter of the golden
age, Leiden, pp. 31e57.
Clement, A.C., Peterson, L.C., 2008. Mechanisms of abrupt climate change of the last
glacial period. Reviews of Geophysics 46, 39.
Conard, N., Moreau, L., 2004. Current research on the Gravettian of the Swabian
Jura. Mitteilungen der Gesellschaft für Urgeschichte 13, 29e59.
48
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
d’Errico, F., Sanchez-Goñi, M.F., 2003. Neanderthal extinction and the millennia
scale climatic variability of OIS 3. Quaternary Science Reviews 22, 769e788.
d’Errico, F., Sánchez-Goñi, M.S., Vanhaeren, M., 2006. L’impact de la variabilité
climatique rapide des OIS3–2 sur le peuplement de l’Europe. In: Bard, E. (Ed.).
L’Homme face au Climat, Paris, 265e282.
Davies, W., Gollop, P., 2003. The Human Presence in Europe during the Last Glacial
Period II: Climate Tolerance and Climate Preferences of Mid- and Late Glacial
Hominids. In: van Andel, T.H., Davies, W. (Eds.). Neanderthals and modern
humans in the European landscape during the last glaciations, Cambridge, pp.
131e146.
Debénath, A., 1994. L‘Atérien du nord de l’Afrique et du Sahara. Sahara 6, 21e30.
Delson, E., Harvati, K., 2006. Return of the last Neanderthal. Nature 443, 76e763.
Fairbanks, R.G., Mortlock, R.A., Chiu, T.-C., Cao, L., Kaplan, A., Guilderson, T.P.,
Fairbanks, T.W., Bloom, A.L., Grootes, P.M., Nadeau, M.-J., 2005. Radiocarbon
calibration curve spanning 10,000 to 50,000 years BP based on paired
230
Th/234U/238U and 14C Dates on pristine Corals. Quaternary Science Reviews
25, 1781e1796.
Fedele, F.G., Giaccio, B., Hajdas, I., 2008. Timescales and cultural process at 40,000
BP in the light of the Campanian Ignimbrite eruption, Western Eurasia. Journal
Human Evolution 55, 834e857.
Finlayson, C., Carrión, J.S., 2007. Rapid ecological turnovers and its impact on
Neanderthal and other human populations. Trends in Ecology and Evolution
22 (4), 213e222.
Fleitmann, D., Cheng, H., Badertscher, S., Edwards, R.L., Mudelsee, M., Göktürk, O.M.,
Fankhauser, A., Pickering, R., Raible, C.C., Matter, A., Kramers, J., Tüysüz, O.,
2009. Timing and climatic impact of Greenland interstadials recorded in
stalagmites from northern Turkey. Geophysical Research Letters 36, L19707.
doi:10.1029/2009GL040050.
Fletcher, W.J., Sánchez Goñi, M.F., 2008. Orbital and sub-orbital climate impacts on
vegetation of the western Mediterranean basin over the last 48,000 yr.
Quaternary Research 70, 451e464.
Folke, C., 2006. Resilience: the emergence of a perspective for social-ecological
systems analyses. Global Environmental Change 16, 253e267.
Gamble, C., 1993. People on the move: interpretations of regional variation in
Palaeolithic Europe. In: Chapman, J.C., Dolukhanov, P. (Eds.), Cultural Transformations and Interactions in Eastern Europe. Aveburg, Aldershot, pp. 37e55.
Gamble, C.S., Davies, W., Pettitt, P.B., Richards, M., 2004. Climate change and
evolving human diversity in Europe during the last glacial. Philosophical
Transactions of the Royal Society B 359, 243e254.
Ganopolski, A., Rahmstorf, S., 2001. Rapid changes of glacial climate simulated in
a coupled climate model. Nature 409, 153e158.
Green, R.E., Stenzel, U., Kircher, M., Patterson, N., Li, H., Zhai, W., Hsi-Yang Fritz, M.,
Hansen, N.F., Durand, E.Y., Malaspinas, A.-S., Jensen, J.D., Marques-Bonet, T.,
Alkan, C., Prüfer, K., Meyer, M., Burbano, H.A., Good, J.M., Schultz, R., AximuPetri, A., Butthof, A., Höber, B., Höffner, B., Siegemund, M., Weihmann, A.,
Nusbaum, C., Lander, E.S., Russ, C., Novod, N., Affourtit, J., Egholm, M., Verna, C.,
Rudan, P., Brajkovic, D., Kucan, Z., Gusic, I., Doronichev, V.B., Golovanova, L.V.,
Lalueza-Fox, C., de la Rasilla, M., Fortea, J., Rosas, A., Schmitz, R.W.,
Johnson, P.L.F., Eichler, E.E., Falush, D., Birney, E., Mullikin, J.C., Slatkin, M.,
Nielsen, R., Kelso, J., Lachmann, M., Reich, D., Pääbo, S., 2010. A Draft sequence of
the Neandertal genome. Science 328, 710e722.
Grootes, P.M., Stuiver, M., White, J.W.C., Johnsen, S., Jouzel, J., 1993. Comparison of
oxygen isotope records from the GISP2 and GRIP Greenland ice core. Nature
366, 552e554.
Higham, T.F.G., Jacobi, R.M., Bronk Ramsey, C., 2006. AMS radiocarbon dating of
ancient bone using ultrafiltration. Radiocarbon 48 (2), 179e195.
Hoffecker, J.F., Anikovich, M.V., Sinitsyn, A.A., Holliday, V.T., Levkovskaya, G.M.,
Pospelova, G.A., Forman, S.L., Giaccio, B., 2008. From the Bay of Naples to the
river Don: the Campanian Ignimbrite eruption and the Middle-Upper Paleolithic transition in Eastern Europe. Journal Human Evolution 55, 858e870.
Holling, C., 2001. Understanding the Complexity of Economic, Ecological, and Social
Systems, Ecosystems, 390e405.
Holling, C., Gunderson, L.H., 2002. Panarchy: Understanding Transformations in
Human and Natural Systems, Washington, D.C.
Housley, R., Gamble, C., Street, M., Petitt, P., 1997. Radiocarbon evidence for the Late
glacial human recolonisation of Northern Europe. Proceedings of the Prehistoric
Society 63, 25e54.
Hulbin, J.J., Roebroeks, W., 2009. Ebb and flow or regional extinctions? On the
character of Neandertal occupation of northern environments. Comptes Rendus
Palevol 8, 503e509.
Hughen, K., Southon, J., Lehman, S., Bertrand, C., Turnbull, J., 2006. Marine-derived
14
C calibration and activity record for the past 50,000 years updated from the
CariacoBasin. Quaternary Science Reviews 25, 3216e3227.
Hüls, M.C., Grootes, P.M., Nadeau, M.-J., 2007. How clean is ultrafiltration cleaning of
bone collagen? Radiocarbon 49 (2), 193e200.
Jiménez-Espejo, F.J., Martínez-Ruiz, F., Finlayson, C., Paytane, A., Sakamoto, T.,
Ortega-Huertasa, M., Finlayson, G., Iijimaf, K., Gallego-Torres, D., Fa, D., 2007.
Climate forcing and Neanderthal extinction in Southern Iberia: insights from
a multiproxy marine record. Quaternary Science Reviews 26, 836e852.
Jöris, O., Weninger, B., 1998. Extension of the 14C calibration curve to ca. 40,000 cal BC
by Synchronizing Greenland 18O/16O ice core records and North Atlantic Foraminifera Profiles: a comparison with U/Th Coral data. Radiocarbon 40, 495e504.
Jöris, O., Alvarez, E., Weninger, B., 2003. Radiocarbon evidence of the Middle to
Upper Paleolithic transition in Southwestern Europe. Trabajos de Prehistoria 60,
15e38.
Maíllo-Fernández, J.-M., 2007. Le Châtelperronien en Espagne: mythes et réalités.
Une approche technologique, 2004. XXVI Congrès de la Société Prehistorique
Française, Avignon. 3, 94e103.
Maroto, J., Vaquero, M., Arriazabalaga, A., Baena, J., Carrión, E., Jorda, J.F.,
Martinón, M., Menéndez, M., Montes, R., Rosell, J., 2005. Problemática cronológica del final del Paleolítico Medio en el Norte Peninsular. Museo de Altamira
MONOGRAFÍAS 20, 101e114 (Santander).
Martrat, B., Grimalt, J.O., Shackleton, N.J., de Abreu, L., Hutterli, M.A., Stocker, T.F.,
2007. Four climate cycles of Recurring Deep and surface water Destabilizations
on the Iberian margin. Science 317, 502e507.
Mayewski, P.A., Meeker, L.D., Twickler, M.S., Whitlow, S., Yang, Q., Lyons, W.B.,
Prentice, M., 1997. Major features and forcing of high- latitude northern
hemisphere atmospheric circulation using a 110,000- year long glaciochemical
series. Journal of Geophysical Research 102, 26345e26366.
Mayewski, P.A., Rohling, E.E., Stager, J.C., Karlén, W., Maascha, K.A., Meeker, L.D.,
Meyerson, E.A., Gasse, F., van Kreveld, S., Holmgrend, K., Lee-Thorph, J.,
Rosqvist, G., Racki, F., Staubwasser, M., Schneider, R.R., Steig, E.J., 2004. Holocene
climate variability. Quaternary Research 62, 243e255.
Moreau, L., 2009. Das Siedlungsmuster im Achtal zur Zeit des älteren Gravettien.
Archäologisches Korrespondenzblatt 39 (1), 1e19.
Moreno, A., Cacho, I., Canals, M., Grimalt, J.O., Sánchez-Goñi, M.F., Shackelton, N.,
Sierro, F.J., 2005. Links between marine and atmospheric processes oscillating on
a millennial time-scale. A multi-proxy study of the last 50,000 yr from the Alboran
Sea (Western Mediterranean Sea). Quaternary Science Reviews 24, 1623e1636.
Moreno, A., González-Sampériz, P., Morellón, M., Valero-Garcés, B.L., Fletcher, W.J.,
in press, Northern Iberian abrupt climate change dynamics during the last
glacial cycle: A view from lacustrine sediments. Quaternary Science Reviews,
doi:10.1016/j.quascirev.2010.06.031.
Naughton, F., Sanchez Goñi, M.F., Desprat, S., Turon, J.-L., Duprat, J., Malaizé, B., Joli, C.,
Cortijo, E., Drago, T., Freitas, M.C., 2007. Present-day and past (last 25 000 years)
marine pollen signal off western Iberia. Marine Micropaleontology 62, 91e114.
Naughton, F., Sánchez Goñi, M.F., Kageyama, F., Bard, E., Duprat, J., Cortijo, E.,
Desprat, S., Malaizé, B., Joly, C., Rostek, F., Turon, J.-L., 2009. Wet to dry climatic
trend in north-western Iberia within Heinrich events. Earth and Planetary
Science Letters 284, 329e342.
Pavlov, P., Roebroeks, W., Svendsen, J.I., 2004. The Pleistocene colonization of
northeastern Europe: a report on recent research. Journal of Human Evolution
47, 3e17.
Pericot, L. 1942, La cueva del Parpalló. Gandia. Valencia.
Redman, C.L., Kinzig, A.P., 2003, Resilience of past landscapes: resilience theory,
society, and the longue durée. In: Conservation Ecology, Jg. 7. online http://
www.consecol.org/vol7/iss1/art14.
Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Bertrand, C.J.H.,
Blackwell, P.G., Buck, C.E., Burr, G.S., Cutler, K.B., Damon, P.E., Edwards, R.L.,
Fairbanks, R.G., Friedrich, M., Guilderson, T.P., Hogg, A.G., Hughen, K.A.,
Kromer, B., McCormac, F.G., Manning, S.W., Ramsey, C.B., Reimer, R.W.,
Remmele, S., Southon, J.R., Stuiver, M., Talamo, S., Taylor, F.W., van der Plicht, J.,
Weyhenmeyer, C.E., 2004. IntCal04 Terrestrial radiocarbon age calibration,
26-0 ka BP. Radiocarbon 46, 1029e1058.
Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Bronk
Ramsey, C., Buck, C.E., Burr, G.S., Edwards, R.L., Friedrich, M., Grootes, P.M.,
Guilderson, T.P., Hajdas, I., Heaton, T.J., Hogg, A.G., Hughen, K.A., Kaiser, K.F.,
Kromer, B., McCormac, F.G., Manning, S.W., Reimer, R.W., Richards, A.A.,
Southon, J.R., Talamo, S., Turney, C.S.M., van der Plicht, J., Weyhenmeyer, C.E.,
2009. IntCal09 and Marine09 radiocarbon age calibration curves, 0e50,000
years cal BP. Radiocarbon 51, 1111e1150.
Roche, D., Paillard, D., Cortijo, E., 2004. Constraints on the duration and freshwater
release of Heinrich event 4 through isotope modelling. Nature 432, 379e382.
Roebroeks, W., 2008. Time for the Middle to Upper Paleolithic transition in Europe.
Journal of Human Evolution 55, 918e926.
Rohling, E.J., Mayewski, P.A., Abu-Zied, R.H., Casford, J.S.L., Hayes, A., 2002. Holocene
atmosphere-ocean interactions: records from Greenland and the Aegean Sea.
Climate Dynamics 18, 587e593.
Roucoux, K.H., de Abreu, L., Shackleton, N.J., Tzedakis, P.C., 2005. The response of
NW Iberian vegetation to North Atlantic climate oscillations during the last 65
kyr. Quaternary Science Reviews 24, 1637e1653.
Salgueiro, E., Voelker, A.H.L., de Abreu, L., Abrantes, F., Meggers, H., Wefer, G., 2010.
Temperature and productivity changes off the western Iberian margin during
the last 150 ky. Quaternary Science Reviews 29, 680e695.
Sepulchre, P., Ramstein, G., Kageyama, M., Vanhaeren, M., Krinner, G., SánchezGoñi, M.-F., d’Errico, F., 2007. H4 abrupt event and late Neanderthal presence in
Iberia. Earth and Planetary Science Letters 258, 283e292.
Shea, J., 2008. Transitions or turnovers? Climatically-forced extinctions of Homo
sapiens and Neanderthals in the East Mediterranean Levant. Quaternary Science
Reviews 27, 2253e2270.
Stiner, M.C., Kuhn, S.L., 2006. Changes in the ‘Connectedness’ and Resilience of
Paleolithic Societies in Mediterranean Ecosystems. Human Ecology 34,
693e712.
Street, M., Terberger, T., 1999. The late Plenigacial and the human settlement of
Central Europe: new information from the Rhineland site of Wiesbaden-Igstadt.
Antiquity 73 (280), 259e272.
Stringer, C., Pälike, H., van Andel, T.H., Huntley, B., Valdes, P., Allen, J.R.M., 2003.
Climatic stress and the extinction of the Neanderthals. Cambridge. In: van
Andel, T.H., Davies, W. (Eds.), Neanderthals and Modern Humans in the European Landscape during the Last Glaciations, pp. 233e240.
M. Bradtmöller et al. / Quaternary International 247 (2012) 38e49
Svoboda, J.A., 2007. The Gravettian on the Middle Danube. Palaeo 19, 203e220.
Terberger, T., Street, M., 2002. Hiatus or continuity? New results for the question of
plenigacial settlement in Central Europe. Antiquity 76 (293), 691e708.
Tiffagom, M., 2005. El Solutrense de facies ibérica o la cuestión de los contactos
transmediterráneos (Europa, Africa) en el Último Máximo Glaciar. In: SanchidriánJ.L., MárquezA.M., FullolaJ.M. (Eds.), IV Simposio de Prehistoria Cueva de
Nerja. La Cuenca Mediterránea durante el Paleolítico Superior, 38.000e010.000
años, pp. 60e77.
Tzedakis, P.C., Frogley, M.R., Lawson, I.T., Preece, R.C., Cacho, I., de Abreu, L., 2004.
Ecological thresholds and patterns of millennial-scale climate variability: the
response of vegetation in Greece during the last glacial period. Geology 32,109e112.
Tzedakis, P.C., Hughen, K.A., Cacho, I., Harvati, K., 2007. Placing late Neanderthals in
a climatic context. Nature 449, 206e208.
van Andel, T.H., Davies, W., 2003. Neanderthals and Modern Humans in the European
Landscape during the Last Glaciation. Cambridge University Press, Cambridge.
van Andel, T.H., Davies, W., Weninger, B., 2003. The human presence in Europe
during the last glacial period I. Human migrations and the changing climate. In:
van Andel, T.H., Davies, W. (Eds.), Neanderthals and Modern Humans in the
European Landscape during the Last Glaciations. Cambridge University Press,
Cambridge, pp. 31e56.
van der Plicht, J. (Ed.), 2000. The 2000 Radiocarbon Varve/comparison Issue.
Radiocarbon, 42, pp. 313e482.
van der Plicht, J., Beck, J.W., Bard, E., Baillie, M.G.L., Blackwell, P.G., Buck, C.E.,
Friedrich, M., Guilderson, T.P., Hughen, K.A., Kromer, B., McCormac, F.G., Bronk
Ramsey, C., Reimer, P.J., Reimer, R.W., Remmele, S., Richards, D.A., Southon, J.R.,
Stuiver, M., Weyhenmeyer, C.E., 2004. NotCal04. Comparison/calibration 14C
records 26-50 cal kyr BP. Radiocarbon 46, 1225e1238.
Vaquero, M., 2006. El tránsito Paleolítico Medio/Superior en la Península Ibérica y la
Frontera del Ebro. Comentario a Zilhao. Pyrenae 37, 107e129.
Vega Toscana, L.G., 2005. El final del Paleolitico Medio y el inicio del Paleolitico
Superior: más allá de los datos Cantábricos. In: Montes R., Lasheras J.A. (Eds.),
Neandertales cantábricos, estado de la cuestión. Santander: Museo de Altamira,
pp. 541e556.
Vegas, J., Ruiz-Zapata, B., Ortiz, J.E., Gala, L., Torres, T., Gracía-Cortés, A., GilGarcía, M.J., Pérez-Gonzàlez, A., Gallardo-Millán, J.L., 2009. Identification of arid
49
phases during the last 50 cal ka BP from the Fuentillejo maar-lacustrine record
(Campo de Calatrava Volcanic Field, Spain). Journal of Quaternary Science.
doi:10.1002/jqs.
Vermeersch, P.M., 2009. Radiocarbon Palaeolithic database Europe v10, European
late Pleistocene Isotopic stages 2 & 3: their ecology and cultural adaptations.
Available under. http://ees.kuleuven.be/geography/projects/14c-palaeolithic/.
Verpoorte, A., 2005. The first modern humans in Europe? A closer look at the dating
evidence from the Swabian Jura (Germany). Antiquity 79, 269e279.
Voelker, A.H.L., Grootes, P.M., Nadeau, M.-J., Sarnthein, M., 2000. 14C levels in the
IcelandSea from 25-53 kyr and their Link to the Earth’s Magnetic Field intensity.
Radiocarbon 42, 437e452.
Walker, B.H., Gunderson, L., Kinzig, A.P., Folke, C., Carpenter, S., Schultz, L., 2006. A
handful of heuristics and some propositions for understanding resilience in
social-ecological systems. In: Ecology and Society, 11. http://www.
ecologyandsociety.org/vol11/iss1/art13 online.
Wang, Y.J., Cheng, H., Edwards, R.L., An, Z.S., Wu, J.Y., Shen, C.C., Dorale, J.A., 2001. A
high-resolution Absolute-Dated late Pleistocene Monsoon record from Hulu
Cave, China. Science 294, 2345e2348. http://www.sciencemag.org/cgi/content/
full/294/5550/2345/DC1. Supplemental Data from.
Weniger, G.-C., 1989. Germany at 18.000 BP. In: Soffer, O., Gamble, C. (Eds.), The
World at 18.000 BP. High Latitudes, vol. 1, pp. 171e192 (London).
Weniger, G.-C., Orschiedt, J., 2000. New answers to old questions? In: Orschiedt, J.,
Weniger, G.-C. (Eds.), Central and Eastern Europe from 50.000 e 30.000 BP.
Wissenschaftliche Schriften 2. Neanderthal Museum, Mettmann, pp. 8e12.
Weninger, B., Jöris, O., 2008. A 14C-age calibration curve for the last 60 ka: the
Greenland-Hulu U/Th timescale and its impact on understanding the Middle to
Upper Paleolithic transition in Western Eurasia. Journal of Human Evolution 55,
772e781.
Zilhão, J., 2006. Chronostratigraphy of the Middle-to-Upper Palaeolithic transition
in the Iberian peninsula. Pyrenae 1, 7e84.
Zilhão, J., Pettitt, P., 2006. On the New dates for Gorham’s Cave and the late survival
of Iberian Neanderthals. Before Farming 3, 95e122.
Zilhão, J., Davis, S.J.M., Duarte, C., Soares, A.M.M., Steier, P., Wild, E., 2010. Pego do
Diabo (Loures, Portugal): dating the emergence of Anatomical Modernity in
Westernmost Eurasia. PLoS ONE 5 (1), e8880. doi:10.1371/journal.pone.0008880.

Similar documents