Subido por Héctor Orellana

Orellana et al., 2023. paper publicado

Anuncio
Quaternary Science Reviews 313 (2023) 108174
Contents lists available at ScienceDirect
Quaternary Science Reviews
journal homepage: www.elsevier.com/locate/quascirev
Spatial analysis of paleoclimate variations based on proxy records in
the south-central Andes (18 - 35 S) from 32 to 4 ka
ctor Orellana a, b, d, *, Claudio Latorre b, c, d, **, Juan-Luis García a, d, Fabrice Lambert a, e
He
lica de Chile, Chile
Instituto de Geografía, Pontificia Universidad Cato
Instituto de Ecología & Biodiversidad (IEB), casilla 653, Santiago, Chile
c
lica de Chile, Chile
Departamento de Ecología, Pontificia Universidad Cato
d
lica de Chile, Chile
Centro UC Desierto de Atacama, Pontificia Universidad Cato
e
Center for Climate and Resilience Research, Chile
a
b
a r t i c l e i n f o
a b s t r a c t
Article history:
Received 8 August 2022
Received in revised form
22 May 2023
Accepted 4 June 2023
Available online xxx
The long-term climate dynamics of the central Andes are part of an ongoing international research effort
to reconstruct past climatic variations and sensitivity to different regional and global drivers during the
last 50,000 years. The large number of diverse records, however, makes it difficult to compare results
without an integrated spatial analysis that considers the nature of the record and whether they are
integrating environmental conditions across a large basin (i.e., a lake record) or at a very local scale (such
as a rodent midden). We compiled 92 records from the southern sector of the central Andes (SCA, 18 35 S). Recalibrated records were further compared by converting the original author's interpretation into
a scale of relative moisture anomalies (compared to the present) that ranges from 2 (very dry) to very
wet (þ2). Moisture anomaly maps were generated for intervals at 4, 6, 9.5, 14, 17, 21 and 32 ka BP (103
calibrated 14C years before present) using records within a 5% age uncertainty. Our compilations show a
surprising degree of agreement in the extent and magnitude of past climate changes during late Pleistocene, but less spatial agreement during the Holocene. The TRACE21 transient climate model shows
similar results, with better agreement during the Pleistocene compared to the Holocene. Our analyses
not only reveal discrepancies between proxy record interpretations at sites from the same region but
show which regions in the SCA require more study.
© 2023 Elsevier Ltd. All rights reserved.
Handling Editor: Dr Giovanni Zanchetta
1. Introduction
The long-term climate dynamics of the South-Central Andes (or
SCA, 18-35 S) is still poorly understood (Garreaud et al., 2010; Rech
et al., 2010, 2019; Norton and Schlunegger., 2011; Rohrmann et al.,
2016). This is due to at least two causes: 1) the lack of continuous
time series that span the last Glacial- Interglacial transition and 2)
the few records that do span this interval continuously are often at
odds in terms of the extent and timing of dry versus wet phases
(often even within the same basin, e.g., Baker et al., 2001; Fornari
et al., 2001; Placzek et al., 2006, 2013).
lica de
* Corresponding author. Instituto de Geografía, Pontificia Universidad Cato
Chile, Chile.
** Corresponding author. Departamento de Ecología, Pontificia Universidad
lica de Chile, Santiago, Chile.
Cato
E-mail addresses: [email protected] (H. Orellana), [email protected]
(C. Latorre).
https://doi.org/10.1016/j.quascirev.2023.108174
0277-3791/© 2023 Elsevier Ltd. All rights reserved.
Due to their latitudinal position spanning both tropical and
extratropical rainfall belts, the SCA contain a large number of
paleorecords used to reconstruct past environmental and climatic
variability during the late Quaternary. For example, sedimentary
and geomorphological records have provided evidence for lake
highstands and periods of extended drought (Sylvestre et al., 1999;
Grosjean et al., 1997a,b; Baker et al., 2001a; Grosjean et al., 2001;
Moreno et al., 2007) and glacial advances (Zech et al., 2006; 2007;
Smith et al., 2009; Aguilar, 2010; Ward et al., 2015). In contrast,
paleoecological records from the Atacama Desert and semiarid
central Chile have been used to infer changes in the elevation of
vegetation belts as species descend or ascend the Andes driven by
past wet and dry phases, respectively (Maldonado et al., 2005;
2005, 2006; Latorre et al., 2003; 2003, 2006; De Porras et al., 2017;
Díaz et al., 2019). Multiproxy studies have also been used to infer
paleoenvironments at archaeological sites (Gil et al., 2005;
Yacobaccio and Morales, 2005; Navarro et al., 2010; Tchilinguirian
et al., 2014; Grana et al., 2016). Many of these paleoclimate
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
2009). As with SC, precipitation in CC occurs during the winter
months, however, mid-latitude storm fronts circulating entrenched
in the Southern Westerly Winds (SWW) become increasingly
frequent as the South Pacific Subtropical High (SPSH) weakens and
migrates northwest (May to September) (Garreaud et al., 2003)
(Fig. 2).
To the east of the Andes, the NWA and Cuyo subregions (22 and
35 S) span along the eastern flank of the Andes. The NWA subregion lies above 2200 up to c. 6900 m.a.s.l., between 28 and 36 S.
Average annual rainfall is 350 mm/year that drops to less than
250 mm/year below 2200 m.a.s.l. in the central plains of Argentina.
The rainfall in this subregion results from tropical and subtropical
moisture advected southwards from the Amazon and Chaco basins
by the low-level northeast jet (NELLJ, 10 -20 S) that intensifies
during the summer months associated with the formation of the
Chaco Low and brings the moisture that feeds convective storms as
far south as 35 S (Garreaud et al., 2003). In contrast, winter rains
from the Pacific are restricted to the high Andes but are crucial in
maintaining snow cover throughout the year (Masiokas et al.,
2012).
reconstructions are not without some controversy. For instance, the
extent and timing of wet versus dry conditions during the Last
Glacial Maximum (LGM) (Thompson et al., 1998; Baker et al., 2001a;
Baker et al., 2001b; Fornari et al., 2001; Placzek et al., 2006) and the
Middle Holocene (Northgrippian) (Betancourt et al., 2000; Grosjean
et al., 2001; Quade et al., 2008; Grosjean et al., 2003).
There are many issues with estimating past hydroclimate from
records, spanning from potential geochronological problems
(unquantified or unknown radiocarbon reservoir effects, e.g., Geyh
et al., 1999), proxy interpretation, and integration of local versus
regional signals (Grosjean et al., 2003). Few past studies have
contrasted the different proxy records from the SCA for specific
timeframes. Such an approach is needed to produce a more complete picture of past environmental conditions for any given period.
Moreover, we still lack a thorough understanding of the geographic
extent of specific paleoclimate events (e.g., the LGM or during the
Last Termination). In this paper we reviewed 92 published paleoclimate records from the SCA with the goal of providing maps for
specific time windows that span the last 32 ka BP (see Supplementary Materials). These records are diverse in terms of types (i.e.,
biological or geological), stratigraphic continuity (from continuous
lake cores to discrete glacial deposits to rodent middens) and
dating methods (radiocarbon, luminescence, cosmogenic-exposure
dating, etc.). Instead of judging the quality or reviewing the interpretations of the published records, we compiled all of them into
paleoclimate maps for specific time-windows that could then easily
be contrasted with paleoclimate model outputs (we use the TRACE
21k transient model, Liu et al., 2009). We extracted a quantitative
moisture index from each of the 92 records, which we then discuss
in terms of the magnitude and direction of past climate change in
the SCA. In this way, the different records can be contrasted with
each other within the same time window, thus revealing any spatial
gaps due to a lack of data and/or conflicting interpretations.
3. Methods
We reviewed 92 paleoclimatic records from the SCA published
since 2020 (see supplemantary information). Our review adds to
and expands on a similar recent review (Tchilinguirian et al., 2014)
that provided a regional synthesis of various proxy records dated to
the Middle Holocene (the Northgrippian) for NWA and SCA. In our
case, we increased the time scale to include all known terrestrial
records since 32 ka BP restricted to those occurring from the subtropical latitudes to tropical latitudes (18 e35 S) and excluded
other sites that are associated with the influence of the lowland
east tropical in the Amazon basin (such as those from the Cordillera
Oriental from Bolivia).
The 92 records reviewed (numbered from 1 to 92, each site and
corresponding number are in Supplementary Materials represent a
diverse array of proxy records including alluvial stratigraphies,
moraines, ice cores, saltpan and lake cores, plant macrofossils,
paleowetland sediments, rodent middens, shoreline (paleolake)
extent, paleodune sequences, pollen, paleosols and diatom records
(Fig. 3). To enable comparisons across records, we converted the
original author's paleoclimate interpretations into an ordinal
moisture anomaly scale that spans from 2 (much drier than
present) to þ2 (much wetter than present) with 0 corresponding to
modern climate (or similar condition). For example, if an author
states that their records are very wet relative to the present it is þ2,
if it was only wetter than present then þ1, the same if the record
shows more arid or arid (2; 1) and 0 if it is similar to present
conditions. We also incorporated a measure that semi-quantifies
into two overall categories the implicit spatial extent represented
by each record in terms of their climate signal: a) the local scale or
point-source signal (e.g., <25 km2 for a single rodent midden) and
b) a basin-wide signal (>25 km2, e.g., a lake core from a large
watershed, such as the Tagua Tagua basin, 35 S). This is graphically
represented by a buffer of either 25 or 50 km, depending on
whether the record represents a local or regional climatic signal.
Once classified, records were then collated into a series of
paleoclimate maps for seven different time windows at 4.0, 6.0, 9.5,
14, 17, 21 and 32 ka BP. These time frames were chosen because they
represent available RCM simulations that have been developed by
the Paleoclimate Modelling Intercomparison Project (PMIP)
(Rehfeld and Brown, 2021). Time windows were further compiled
into the geographic sub-regions described above. Each timewindow is a map with a snapshot of relative moisture conditions
at a given moment including records that were dated to within a 5%
2. Physiography and present climate
The SCA as a region lies between 18 and 35 S, which encompasses a total of eight distinct subregions within Bolivia (1 subregion), Chile (5 subregions) and Argentina (2 subregions). These are
the Bolivian Altiplano, the Northern Atacama Desert, the Central
Atacama Desert, the Southern Atacama Desert, Semiarid Chile,
Central Chile, and Northwest Argentina and the Cuyo Region
(Fig. 1). The physiography and present-day climate for each of the
eight subregions are further detailed below.
The Bolivian Altiplano (abbreviated BA), with an average
elevation of 3800 m a.s.l., is located between the fold and thrust
belt of the eastern Cordillera and the active volcanic arc of the
western Cordillera (16 e23 S) (Rigsby et al., 2005; Jordan et al.,
2014). For this study, we consider the southern Altiplano corresponding to the Salar de Uyuni (17 e23 S) basin where most SCA
paleoclimatic records have been described. This subregion is
characterized by a tropical influence and elevations that reach
5500 m a.s.l. (Tapia et al., 2019).
In Chile, the region between 17 and 27 S known as the Atacama
Desert is divided into three subregions: Northern Atacama (NA, 17
to 21 S), Central Atacama (CA, 21 to 24 S) and Southern Atacama
(SA, 24 e27 S). At elevations higher than 3000 m a.s.l. these regions, together with the BA and Northwest Argentina (NWA) are
characterized by dry and cold winters (June to August) and rainy
summers (December to March) with overall rainfall decreasing
from north to south. Farther south, the subregions between 27 32 S and 32 - 35 S are termed Semiarid Chile (SC) and Central Chile
(CC), respectively. Precipitation in the SC is highly dependent on the
Andean orography, and it varies from 25 mm/year along the coast
to approximately 300 mm/year in the Andes (Falvey and Garreaud,
2
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 1. Map of the south-central Andes showing the geographic extent of regions discussed. Coastal Range, Central Depression, Altiplano-Puna, Northern Atacama (NA), Central
Atacama (CA), Southern Atacama (SA), Semiarid Chile (SC), Central Chile (CC), Northwest Argentina (NWA), and Cuyo subregions (CR).
et al., 2013). We could not recalibrate cosmogenic ages as most
authors did not include the necessary data in their original publications. Recalibrated 14C ages are expressed as ka BP whereas other
ages (e.g., UeTh, OSL, etc.) are in ka only. All maps were drawn
using ArcGIS 10.5 and the SRTM topographic base. For each site
compiled into our database we discuss the evidence used by the
authors in assigning our moisture anomaly category. In many cases,
also included assigning rainfall seasonality (summer versus winter)
suggested in publications (see Supplementary Materials).
Comparisons to our time-windows were made against paleoclimate simulations that were provided by the TRaCE 21k transient
simulation (Liu et al., 2009). This simulation uses a fully coupled
model (National Center for Atmospheric Research Community
Climate System Model version 3) that is forced by time-varying
orbital parameters, greenhouse gases, time-varying ice-sheet
extent and topography, and freshwater forcing to the oceans from
the retreating ice sheets (Liu et al., 2009).
range of that age (see Supplementary Materials). It is important to
note that our maps represent a “consensus” view as they incorporate any record that falls within the defined time interval. If the
same record has a date within the range of one of the temporal
windows with a different climatic signal (wet or dry) then the
signal closest to the target time interval was chosen. For example, if
the same record shows “wet” at 21.2 ka BP and “dry” at 20.5 ka BP
for the interval 21 ± 1.2 ka, the “accepted” (e.g., mapped) record is
the one closest in age to the 21 ka BP date, in this case 21.2 ka BP.
The qualitative nature of this analysis is based on the interpretation
that each author makes of their records (we did not reinterpret
records), that in some cases may not be very accurate of actual
relative moisture conditions, but is nevertheless a good element to
spatially compare records and to see differences with paleoclimatic
models.
Records that use radiocarbon chronologies were recalibrated
with Calib7.0.4 (Reimer et al., 2013) and the SHCAL13 curve (Hogg
3
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 2. Average annual precipitation over South America (Worldclim2 data, Fick and Hijmans, 2017). A) Distribution of average summer precipitation (DJF). B) Distribution of average
winter precipitation (JJA). The SCA are also shown for reference (red rectangle). Major wind directions associated with extra-tropical rainfall (orange arrows) and average directions
of tropical rainfall winds (black arrows) are also shown. Other major circulation features shown are the SPSH (South Pacific Subtropical High) and the NELLJ (North Easterly LowLevel Jet).
del Chaco I [60] (25 S) also show the presence of taxa that today
occur 700 m higher in elevation and their presence has been
attributed to an increase in winter precipitation before 32 ka BP,
although the ages are reported as >33.8 ka BP (Maldonado et al.,
2005) and are thus infinite dates and were not considered here in
our review. Increased dune formation at Llano Agua de los Burros
[65] (27 S) in the SA between 41 and 28 ka BP suggests an increase
in the source of fluvial sediments, as a result of increase rainfall in
the Andes (Nash et al., 2018).
Records from SC and CC show positive moisture anomalies at 32
ka BP. Exposure dates (10Be) on lateral moraine boulders indicative
~ a Rosa [78], a currently
of extensive glaciation at Cordon Don
unglaciated area with elevations over 4500 m. a.s.l, yielded ages of
32 ± 3 ka (Zech et al., 2007). The 10Be age was calculated using the
Kubik and Ivo-Ochs (2004) calibration data and the scaling system
of Desilets and Zreda (2003) (5.25 10Be atoms a1 g 1 SiO2 for
neutron spallation and geomagnetic corrections). Further south,
pollen evidence from a lake core obtained at the Tagua Tagua basin
[89] shows an increase in percentages of the forest taxa Nothofagus
olbiqua-type which is indicative of a much wetter environment
s
compared to the mediterranean environment today (Valero-Garce
et al., 2005).
On the eastern flank of the Andes, records from the NWA also
show evidence for increased moisture, as suggested by La Poma
[53]. This record suggests important mass movements during a wet
phase that occurred between 40 and 25 ka BP. These mass movements dammed valleys forming paleolakes with finely laminated
sediments and suggest greater inter and intra-annual fluctuations
in precipitation compared to the present (Trauth and Strecker,
1999; 2000, 2003).
4. Results
For each of the seven time-windows mapped based on a total of
92 paleoclimatic records (including those that were recalibrated)
we explain the general trends regarding moisture anomalies.
4.1. 32 ka BP
A total of ten records dated between 33.9 and 31.4 ka BP were
included (Fig. 4). There are only 3 negative records from BAativeand
seven with positive moisture anomalies during this interval along
the SCA. Correlations of radiocarbon dated alluvial facies in sediment cores obtained in the Desaguadero valley (~18 S), which includes cores from Caquingora [4] (numbers in italics here refer to
the same keyed number in Fig. 3, this is the proxy record number)
and Huana Khaua [15]) are indicative of drier (or intermittently
wet) conditions from 38 to 26 ka BP (Rigsby et al., 2005). This dry
phase was preceded and succeeded by lacustrine facies and paleowetlands indicative of wetter climates before and after this interval in the Desaguadero valley (Rigsby et al., 2005). A different
result is reported from a core obtained from the Uyuni salt flat [26]
in Baker et al. (2001), where lacustrine muds suggest a wet event
from 33.4 to 31.8 ka BP. In contrast, a core from the same Uyuni salt
flat [25] suggests arid conditions due to the presence of salt crusts
dated to the interval spanning 35e22 ka BP (Fornari et al., 2001).
Several records from the Atacama Desert (CA and SA) provide
evidence for positive moisture anomalies at 32 ka BP. These include
a U-series dated salt core obtained from Salar de Atacama [42] and
AMS 14C-dated rodent middens from Quebrada Juncal [61]. Large
chevron halite crystals formed during a saline perennial lake, and
show wetter conditions from 54.3 to 15.3 ka BP (Bobst et al., 2001).
Extralocal taxa (Phacelia pinnatifida, Jarava frigida) found at higher
elevations today occur in midden macrofloras at Quebrada Juncal
[61], which can be attributed to an increase in local rainfall (Díaz
et al., 2012). Pollen from rodent middens collected at Quebrada
4.2. 21 ka BP
For this map, we compiled all published records from the SCA
dated to the interval 22.2e20.2 ka BP spanning the Last Global
4
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 3. Location and proxy type of the 92 records compiled in this study. See Supplementary Materials for a list of site references.
during much wetter conditions in relation to the present, such as
the Tauca (18.1e14.1 ka), Inca Huasi (ca. 46 ka) or Ouki (120e98 ka)
cycles (Placzek et al., 2006, 2013). In contrast, three other records
from elevations above 3000 m asl (SCA) have been interpreted as
evidence for major increases in precipitation during the LGM. Low
d18O values at Nevado Sajama [5] are interpreted as evidence for
increased snow accumulation and wetter conditions between ~21
and 18 ka from a core drilled into the summit (Thompson et al.,
1998) (Fig. 5). Very wet conditions have also been interpreted
from much higher values of natural g-ray radiation in mud
(lacustrine muds and planktonic diatoms) present in a salt core in
the Uyuni basin [26] (Baker et al., 2001). This contrasts with evidence from a 121 m salt core also drilled in the Salar de Uyuni basin
[25] that shows no evidence for increased lake levels during the
LGM, such as mud layers with diatoms, and instead shows salt crust
development from ~37.9 to ~19.2 ka BP (Fornari et al., 2001; for
further discussion of these ages, see also Placzek et al., 2013). A
wetter than today climate was also interpreted from the Huana
Khau record [15] denoting increased accumulation of voluminous
Glacial Maximum (LGM- Mix et al., 2001; Hughes et al., 2013)
(Fig. 5). A total of 16 records distributed across the Bolivian Altiplano, the Central Atacama, Central Chile and the Cuyo Region
provide information on past climate during the LGM. These include
fluctuations of paleolake levels based on the dating of preserved
shorelines in Uyuni Basin [21] (Placzek et al., 2006, 2013), salt cores
from Salar de Uyuni [26] (Baker et al., 2001; Fornari et al., 2001),
alluvial deposits in the Desaguadero Floodplain [15] (Rigsby et al.,
2005) and ice cores from Nevado Sajama [5] (Thompson et al.,
1998).
The resulting proxy-based climate reconstruction during the
LGM (Fig. 5) reveals a complex scenario. For example, shoreline
data from the Salar de Uyuni [23] (BA) suggest only a slight increase
in paleolake levels from 24 to 20.5 ka BP (ca. 5 m above the salt flat
surface, the Sajsi event), (Placzek et al., 2006, 2011, 2013). The
presence of a shallow paleolake during the Sajsi phase (paleolake
cycles - 24 to 20.5 ka BP) has been interpreted as forming due to
reduced evaporation in a much colder and relatively dry climate in
relation to other paleolake cycles in the Uyuni basin, which formed
5
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 4. Records compiled for the 32 ka BP interval. See Supplementary Materialsfor list of site numbers and references.
extralocal shrub Baccharis tola, which would indicate a slight increase in mean annual precipitation at 3000 m a.s.l (~10e20 mm,
Latorre et al., 2006). Rodent middens further south at Tilocalar [47]
reveal the presence of a very cold, drier than present and sparsely
vegetated ecosystem (Betancourt et al., 2000; Latorre et al., 2003).
A lake core from Laguna Miscanti [44] points to not only much drier
conditions during the LGM but also to much colder, as indicated by
glaciofluvial outwash derived from the adjacent eastern cordillera
in the Desaguadero floodplain (BA) that was dated using bulk
radiocarbon dating on muds (Rigsby et al., 2005.).
Further south, most records from higher elevations
(>2800e4500 m a.s.l) in CA are indicative of negative or no moisture anomalies during the LGM. A rodent midden collected at Río
n, 2985 m a.s.l.) [32] indicates the presence of the
Salado (El Sifo
6
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 5. Records compiled for the 21 ka BP (LGM) interval. See Supplementary Materials for list of site numbers and references.
moraines using 10Be and 36Cl yielded 20.8 ± 1.5 to 18.2 ± 2.8 ka in
the Chajnantor Plateau [39] (4585 m a.s.l.) (Ward et al., 2015), are
synchronous with the exposure ages for debris flow boulders (36Cl)
dated to 20.2 ka [40] in the same sector, but 2000 m a.s.l below the
Chajnantor Plateau (Cesta and Ward, 2016). In both cases, expanded
glaciers and enhanced alluvial activity have been interpreted as a
wetter than present climate during the LGM.
the complete absence of pollen from this portion of the record
(Grosjean et al., 2001) which is also absent from the Lake Titicaca
pollen LGM record (Bolivia) (Paduano et al., 2003). In contrast,
“chevron” halite (instead of the usual efflorescent halite crust)
preserved in the Salar de Atacama [42] salt core has been interpreted as evidence for wetter conditions (Bobst et al., 2001). On the
other hand, cosmogenic surface-exposure dating of boulders on
7
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
versus a mud phase by Baker et al. (2001) in the same Uyuni basin
[25].
Paleoecological evidence from the NA suggests very wet conditions. Abundant fossil leaf litter deposits of the willow-like
Escallonia agustifolia tree were present along much of the Lluta
valley [11] up to 15 m higher up on the slopes. As these slopes today
are hyperarid, their presence implies a much higher groundwater
table than today between 38 and 15.4 ka BP. Groundwater tables
plummeted after 15.4 ka BP, most likely caused by intensified
incision due to increased stream power and runoff from the Río
Lluta headwaters in the Andes Cordillera (Veit et al., 2016). In situ
rhizoliths and plant macrofossils occur along fluvial terraces at
Sipuca [30] in the Pampa del Tamarugal basin (Nester et al., 2007;
Gayo et al., 2012), which today lacks vegetation (absolute desert).
These plant macrofossils and rhizoliths were dated between 17.8
and 16.9 ka BP, indicative of increased discharge at relatively low
elevations (1250 m a.s.l, Central Depression).
Other records from the Pampa del Tamarugal, such as Quebrada
Maní [29], also suggest wetter conditions, associated with extensive
paleowetlands that developed between 17.6 and 15 ka BP
(Workman et al., 2020) due to elevated regional groundwater tables. Positive moisture anomalies have been established in the CA
n in the Río Salado basin
from rodent middens collected at El Sifo
[32] with rainfall increases of at least 60 mm/yr in relation to the
present day (a 75% increase) (Latorre et al., 2006). Further south,
along the coastal range at Sierra el Buitre [56], a rodent midden
dated to 17.3 ka BP shows the presence of extralocal Andean species
(Adesmia atacamensis; Cryptantha diffusa) indicative of longdistance dispersal events that were likely facilitated by increased
runoff across hyperarid sectors in the Atacama Desert (Díaz et al.,
2012).
Formation of crystalline halite in a saline perennial lake have
also been interpreted as wet conditions between 53.4 and 15.3 ka
BP in the Salar de Atacama basin [42] (Bobst et al., 2001). In
contrast, pollen and sediment limnogeochemical analyses from
Laguna Miscanti [44] suggest the disappearance of this freshwater
lake under drier conditions than today between 22 and 14 ka BP
(Grosjean et al., 2001).
In the SA, much wetter conditions have also been inferred from
rodent midden pollen in the higher elevations (3450 m), which
show abundant steppe grasses at Quebrada del Chaco III [57]. Today,
these steppe grasses occur at slightly higher elevations and dominate the landscape at 4000 m a.s.l (Maldonado et al., 2005; Latorre
et al., 2005).
Further south in the SC and CC, a terminal moraine in the
headwaters of the Río Turbio valley [75] (>4000 m a.s.l) was dated
to 17e12 ka BP and is indicative of increased moisture availability
attributed to an increase in summer precipitation (Riquelme et al.,
2011). Elevated moisture availability between 19.5 and 17 ka BP was
also inferred from an increase in forest taxa and associated decrease
in Amaranthaceae pollen in the Tagua Tagua record from central
s et al., 2005). In NWA, the presence of a
Chile [89] (Valero-Garce
saline lake from 21 to 13.8 ka BP in the Guayatayoc Salt Flat [43], is
interpreted as a climate much wetter than today.
Further south, rodent midden records from the SA have been
interpreted as evidence for a wetter LGM. These include rodent
midden records from Cerro Tres Tetas [50] (1900 m a.s.l), which
reveal the presence of considerably more diverse desert floras than
those present today (Díaz et al., 2012). Pollen (especially the presence of Asteraceae- Chaethantera type) and plant macrofossils from
LGM middens collected at Quebrada Chaco I [60] (2600 m a.s.l. - a
hyperarid, plantless environment today) show the presence of
numerous Andean taxa (>3000 m a.s.l) that included perennials
and annuals as well as elevated species richness (Latorre et al.,
2005; Maldonado et al., 2005).
Even further south at 32 S, a wet period likely ended by ~22 ka
BP, when paleosols were buried by eolian sand dunes in coastal CC
indicative of a return to arid or similar to present conditions (García
et al., 2019). At 34 S, a medial moraine indicative of an ice advance
dated to 20.4 ka was dated in the Cachapoal valley [87] (1505 m
a.s.l) suggesting cold and wet conditions in Central Chile (Charrier
et al., 2019). Pollen and geolimnogeological evidence from a sediment lake core from Laguna Tagua Tagua (Central Chile) [89] are
suggestive of very wet conditions between 40 and 21 ka (Fig. 5) as
indicated by increases in arboreal pollen and aquatic vegetation,
among other limnological proxies. This period was followed by
more arid conditions than present (~21e19.5 ka BP) as inferred
from the increase of Amaranthaceae pollen percentages along with
s et al., 2005).
evidence for a decreased lake level (Valero-Garce
Increased aridity has been described for CR, where sand dune
deposition occurred at ~22 ka BP near San Rafael [88] at ~34 S
(Tripaldi et al., 2010).
4.3. 17 ka BP
There are many records available for this time interval (n ¼ 23,
Fig. 6). These are from the BA, NA, CA, SA, SC and CC and the CR.
Records from BA all indicate considerably wetter climates
throughout the region. These records are diverse, such as the
Sajama ice core [5] which shows both low aerosol concentrations
and very negative d18Oice values that resulted from increased snow
accumulation and much wetter conditions (Thompson et al., 1998).
A major glacial advance occurred at Tunupa volcano [24] was also
determined from 3He exposure dating on pyroxene phenocrysts
from boulders resting on moraines (Blard et al., 2009). Three
sediment cores (with alluvial/fluvial stratigraphy) from the Desaguadero floodplain in the Bolivian Altiplano (Manquiri [1], Zambrana [2] and Huana Khaua [15]) suggest abundant discharge from
large tributaries in the Desaguadero river and the formation of large
lakes in the central and southern Desaguadero river floodplain, all
of which have been interpreted as the prevalence of a large positive
moisture anomaly (Rigsby et al., 2005). These records are also
consistent with evidence for planktic diatoms (Cyclotella striata)
and charophyte algae preserved in shoreline features (Pakollo
Jahuira [18], Pisalaque[20] and Churacari Bajo [23]) that suggest a
shoreline 27 m higher than the current Salar de Uyuni surface
(Sylvestre et al., 1999). Exposed sections at Vinto [7] and Toledo [12]
revealed geosols (a paleosol stratigrapically interbedded between
two paleolake cycles), indicative of landscape stability which caps
the lake transgressions to elevations below 3700 m. a.s.l from 95 to
17 ka (Placzek et al., 2006, 2013). Evidence for a prominent paleolake known as the Tauca phase (18.1e14.1 ka BP), is found
throughout the Uyuni basin, reaching a maximum extent of
60,000 km2 and a water depth of 140 m over what today is a dry salt
flat between 16.4 and 14.1 ka BP) (Placzek et al., 2006, 2013). Very
wet conditions are also recorded in a salt core at Uyuni [26], which
reveals alternating phases of muds (paleolake) and salt (gypsum
and halite) for the last 50 ka (Baker et al., 2001). In contrast, Fornari
et al. (2001) evidence a salt phase which suggests arid condition
4.4. 14 ka BP
A total of 33 records with dates between 14.3 and 13.7 ka BP
occur throughout the SCA (Fig. 7). 45% of these come from the BA;
15% is from the NA, 15% from the CA and 9% from the SA. 9% records
are from the SC and 3% is from CC. The remainder are from NWA
(3%).
Records from the Bolivian Altiplano mostly show very positive
moisture anomalies (þ2). For instance, planktonic diatoms and
other aquatic plants in sediment cores obtained from the
8
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 6. Records compiled for 17 ka BP interval. See Supplementary Materials for list of site numbers and references.
associated with the final phase of “Tauca Paleolake”, which is dated
from 18.1 to 14.1 ka BP (Placzek et al., 2006). Moreover, glaciers
were advancing in the Nevado Sajama, such as Huaji Jihuata [6]at
14 ± 0.5 ka BP. These ice advances occurred between 16.9 and 11.3
ka BP (Smith et al., 2009). In contrast, a brief positive d18Oice
excursion occurs in the Nevado Sajama ice core [5] from 14.3 to 14.0
ka BP indicative of very arid conditions, together with increased
Desaguadero Floodplain at Tejopa [3] and Huana Khau [15] show
evidence for much wetter conditions compared to today (Rigsby
et al., 2005). A similar condition occurs at Salar de Uyuni based
on evidence from a salt core and elevated shorelines. Shorelines
~ a [13], Salar de
occurring in the northern Salar Coipasa [16], Pazn
Uyuni [21], northern Salar de Uyuni [19], southern Salar de Uyuni
[27], and in the southern sector of Salar de Coipasa [22] are all
9
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 7. Records compiled for the 14 ka BP interval. See Supplementary Materials for list of site numbers and references.
was originally interpreted from the presence of alluvial sands along
the margins of the Coipasa and Uyuni salars deposited under
generally dry conditions (Sylvestre et al., 1999). These sands contained freshwater mollusk shells reworked from preexisting Tauca
lake deposits (Sylvestre et al., 1999). A few other records (Alianza
[24] and Huacuyo [17]) in the Uyuni basin, however, are suggestive
dust in the ice core (Thompson et al., 1998). This possibly correlates
~ a event” dated between 14.1 and 13.8 ka BP
with the arid “Tican
(Sylvestre et al., 1999; Baker et al., 2001; Placzek et al., 2006)
~ a [17] and in a salt core from
apparent in the Uyuni basin at Tican
the same basin [26]. This event occurred between the paleolake
Tauca (18.1e14.1 ka BP) and Coipasa (c. 13 to 11 ka BP) phases and
10
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
short dry event occurred at c. 9.5 ka BP [10] (Moreno et al., 2007).
Pollen from Andean Forest taxa at Laguna Seca [8] suggest a climate
cooler and wetter than today until about 9 ka BP (Baied and
Wheeler, 1993). An in-situ tree trunk found in alluvial sediments
at Pampa del Tamarugal has been dated directly to 9.4 ka BP cal. BP,
indicative of wet conditions at the time it grew as the hyperarid
environment today has no vegetation (Nester et al., 2007).
In contrast, onset of aridity is inferred in the CA. First, efflorescent halite crust began to accumulate in the Salar de Atacama salt
core [42] at 9.5 ka BP (Bobst et al., 2001). Aridity also occurred along
the Río San Salvador [34] where channel downcutting (which occurs during droughts) is evident between 9.8 and 6.4 ka BP (Tully
et al., 2019). In the SA, paleowetlands dried out around 9.7 ka BP,
with sediments deposited afterwards forming under low groundwater tables typical of a phreatic playa at Quebrada Guanaqueros
[49], El Salto [51] and Punta Negra [52] (Quade et al., 2008).
Further evidence for negative moisture anomalies in the CA
comes from the disappearance of steppe grasses and the predominance of prepuna taxa (more arid) on open slopes and small
mountain ranges from almost all published rodent midden records,
including those at Cerros de Aiquina [33], Lomas de Tilocalar [48],
Cordon de Tuina [35] and El Hotel- Paso Barros Arana [37] (Latorre
et al., 2003; 2003). The pollen record from these same rodent
middens at Cerros de Aiquina supports these conclusions (De
Porras et al., 2017). In contrast, plant macrofossils from El Sifon
rodent middens [32] have been identified as Andean steppe
grasses, which do not exist at present at this site, implying the
persistence of slightly wetter conditions (Latorre et al., 2006).
Positive moisture anomalies are also inferred from geochemical
and pollen evidence at Laguna Miscanti [44] (Grosjean et al., 2001).
SA midden pollen records from Quebrada La Zorras [53] show
pollen assemblages typical of more transitional communities,
characterized by Amaranthaceae, Brassicaceae and Adesmia, all
indicative of wetter conditions than present, but drier than assemblages from the Late Pleistocene (De Porras et al., 2017). Two SA
records indicate arid conditions. A stratigraphic sequence of
groundwater (paleowetland) deposits from Aguada Alto Varas [55]
shows massive breccia layers interpreted as alluvial fan channels
and the absence of organic layers at this time (which are interpreted as times of spring inactivity) due to overall arid conditions
ez et al., 2016).
(Sa
SC records all show arid conditions. Exposure 10Be ages on
boulders show that buried glacier ice at Valle Encierro [69] had
disappeared by 9.3 ka BP (Zech et al., 2006). Glacier retreat and
deposition of alluvial debris cones in the Río Turbio valley [73] are
also interpreted as negative moisture anomalies (Riquelme et al.,
2011). Local dominance of herbaceous and coastal shrubs in pollen assemblages between 9.5 and 9.0 ka BP at Santa Julia [79] also
indicate drier conditions (Maldonado et al., 2005). This is further
supported by geomorphological and alluvial stratigraphy at Quebrada Santa Julia [79] (Ortega et al., 2012), where the presence of
clayey muds as well as scant and fine-grained alluvial deposits
provide further evidence for dry conditions, concomitant with
sporadic precipitation events that were able to transport very fine
sands. In contrast, intense snow accumulation at high elevations
accompanied by alluvial fan deposition has been interpreted as an
increase of precipitation associated with the southern Westerlies,
such as those present at Juntas Turbio y Seco [73] (Veit, 1996).
Positive moisture anomalies are also inferred from the presence of
swamp forest indicators Myrtaceae and Escallonia at the coastal
n, 2006). A lake
Palo Colorado site [81] (Maldonado and Villagra
core from Laguna El Cepo [76], shows the start of an arid phase at
9.5 ka BP, inferred from lack of a coarse fraction and lower C/N
values suggesting arid climates (Tiner et al., 2018).
CC records all show negative moisture anomalies. The lack of
of wetter conditions, as evidenced by the presence of lacustrine
sediment at 14.1 ka BP (Sylvestre et al., 1999) before the collapse of
the Tauca paleolake cycle.
Records from the NA suggest wetter conditions. Sedimentary
facies including diatom oozes suggest very wet conditions in a core
from Lago Chungara [10] (Moreno et al., 2007). Steppe grasses and
other extralocal taxa (Cumulopuntia boliviana, Oreocereus leucotrichus and Solanum sp.) occurred at lower elevations than today at
14.4e11.6 ka BP as revealed by rodent middens from Quebrada La
Higuera [14] (Mujica et al., 2015). Similarly, alluvial deposits and
wetland outcrops with plant macrofossils occur in the hyperarid
Pampa del Tamarugal basin in places such as Lomas de la Sal [31]
and Chipana [28] between 17.6 and 14.1 ka BP (Gayo et al., 2012;
Pfeiffer et al., 2018). In contrast, more recent stratigraphic evidence
from Quebrada Mani [29] also in the Pampa del Tamarugal basin,
shows evidence for an erosive unconformity with deposition of
eolian sands from to >13 - <15 ka BP (Workman et al., 2020).
Abundant grass pollen together with very low Amaranthaceae
percentages correlates with a major lake transgression at Laguna
Miscanti [44] (Grosjean et al., 2001) in CA. Further evidence for wet
conditions is inferred from extensive wetlands that formed during
this interval along the western Andean piedmont at Tilomonte [45]
and Salar de Punta Negra [52] (Rech et al., 2002; Quade et al., 2008).
In contrast, the Salar de Atacama salt core in [42] shows the formation of subaerial halite between 16.5 and 11.4 ka BP (Bobst et al.,
2001). A rodent midden from El Sifon [32] in the Río Salado basin
directly dated to 14.2 ka BP shows that mean annual rainfall
declined to modern conditions, as this midden lacks plant macrofossils from the Andean steppe or other wet indicator taxa (Latorre
et al., 2006).
Intense aridity with lowered groundwater tables has been
inferred from alluvial sediment sections in Alto Varas [55] in the SA
(S
aez et al., 2016). The only exception is a brief wet interlude at c.
ez et al., 2016).
14.5 ka BP evidenced by a black peat layer (Sa
Records show wetter climates in SC and CC. Exposure ages on
boulders resting on lateral moraines LM and the terminal moraine
M-II B yielded an age of 14 ± 0.6 ka BP and indicate a late glacial
advance in the Encierro valley [69] (Zech et al., 2006). Aguilar
(2010) however, proposes that these glacial sediments were
reworked (Encierro valley [67)], and the moraines were altered
(moraine settlement) around 14 ka BP. Further south in the headwaters of the Turbio valley [75], a glacial advance has been dated to
17e12 ka BP based on glacial lacustrine succession (Riquelme et al.,
2011). A large decrease in Amaranthaceae pollen and concomitant
increase in lake levels is also interpreted as positive moisture
s et al.,
anomalies in the Tagua Tagua basin [89] (Valero-Garce
2005). In NWA, similar wet conditions occurred at Guayatayoc
[43], where evidence for a aline lake occurred from 21 to 13.8 ka BP.
4.5. 9.5 ka BP
A total of 31 records distributed across the SCA (Fig. 8) comprise
this interval, including records from BA, NA, CA, SA, SC, CC, NWA
and CR. Overall, the records available show highly variable moisture
anomalies even within the same subregions.
For example, two records from BA show evidence for wetter
conditions. These include a paleowetland record from Huana Khaua
[15], which has been interpreted as wetter than present conditions
between 10 and 7.9 ka BP (Rigsby et al., 2005). Decreased soluble
anions (SO4e and Cl-) and dust remained low in the Nevado Sajama
ice core [5] from 14 until 9 ka BP, indicative of wetter climates and
higher paleolake levels across the BA (Thompson et al., 1998).
Concomittantly with the BA, NA records also show positive
moisture anomalies. A low Ca/Ti occurs in the Lago Chungara (lake
core) c. 11 to 7.5 ka BP and is linked to wetter conditions, although a
11
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 8. Records compiled for the 9.5 ka BP interval. See Supplementary Materials for list of site numbers and references.
taxa almost disappeared and were replaced by grasses along with
s et al., 2005).
lowered lake levels (Valero-Garce
NWA records all indicate positive moisture anomalies. For
example, planktonic diatoms have been described from the Hornillos II [41] archaeological site (Yacobaccio and Morales, 2005) and
pollen at Laguna Aculeo [86] has been attributed to more arid
conditions and very low plant productivity (Jenny et al., 2002),
coupled with low lake levels inferred from the presence of evaporites indicative of increased arid conditions (Jenny et al., 2002). A
similar situation occurred at Tagua Tagua [89] where regional forest
12
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
broad floodplains with organic soils dominated by grasses imply an
elevated and stable water table at Pasto Chicos [46] (Tchilinguirian
et al., 2014). A wetter climate is also inferred from sedimentary
facies changes and diatom records at Antofagasta de la Sierra [62]
(Grana et al., 2016). In contrast, CR records show more arid than
current conditions during the Early Holocene. Pollen evidence from
a cave sediment record at the rockshelter Gruta el Indio [90] suggests a lower than present precipitation regime and higher than
present temperatures between 12 and 8.5 ka BP (Gil et al., 2005).
occurs in the Salar de Atacama salt core [42] suggesting a positive
moisture anomaly (Bobst et al., 2001).
SA records show negative moisture anomalies. These include
rodent middens from the higher elevation site of Quebrada el Chaco
III [57] which have low taxa richness and low levels of steppesubnival and puna taxa (Maldonado et al., 2005; Latorre et al.,
2005). A lake core from Laguna Negro Francisco [63] shows abundant evaporites, indicating very arid climate (Grosjean et al.,
1997a,b). A rodent midden record from Quebrada Juncal [61],
however, shows no moisture anomaly (climate similar to modern)
based on species richness and composition that are similar to today
around 6.1 ka (Díaz et al., 2012).
Three SC records all have been interpreted as more arid than
today. Strong, infrequent storms and subsequent debris flows in an
overall arid climate have been cited to explain the formation of
down-cut alluvial fans along the Río Turbio valley [67] (Riquelme
et al., 2011). A peat layer at Pabellon [71] dated to 5.9e5.5 ka BP
has been interpreted as the product of very low sedimentation and
strongly reduced alluvial activity, interpreted as arid conditions
(Veit, 1996). A similar situation occurred at Quebrada el Transito
[66], where the presence of wetland deposits and the soil formation
on the alluvial fans have been interpreted as low sedimentation
rates associated with arid conditions between 7 and 5.5 ka BP
et al., 2017). A predominance of open vegetation and her(Cabre
~
baceous taxa documented in the Nague
I pollen record [80] denotes
n,
increased aridity for the region also (Maldonado and Villagra
2002). “Pollen-starved” sediments with only scant presence of
semiarid vegetation (only shrub taxa) have been interpreted as
evidence for extremely arid climates in relation to today between
8.7 to ~5.7 ka BP at the Palo Colorado [81] swamp forest
n, 2006). Just 26 km north of Palo Colorado
(Maldonado and Villagra
at Quebrada Santa Julia [79], geomorphological and alluvial sedimentary evidence suggest episodic and torrential rainfall events
under an arid climate setting between 8.6 and 5.7 ka BP (Ortega
et al., 2012). Coarser and poorly sorted grain-size distributions
and higher C/N ratios occur in Laguna el Cepo [76], which implies
arid conditions from 9.5 to 5.5 ka BP (Tiner et al., 2018).
Paleosol, pollen and lake CC records also show negative moisture anomalies. A sparse herbaceous flora was present under
warmer and drier conditions at Quintero II [83] (Villa-Martinez and
Villagr
an, 1997). Formation of an A-horizon (soil) near Valparaiso
[84] has been interpreted as a period of aridity, where the fog is
reduced in these erosive periods. The soils reflect the lack of deep
water infiltration, because they are only characterized by the
accumulation of organic material (A-horizons) (Veit, 1996).
Elevated percentages of Amaranthaceae and low arboreal pollen
also occur in a lake core from Laguna Aculeo [86] (Jenny et al.,
2002). Extensive wetland surface fluctuations with relatively high
and variable content of Typha (herbaceous), almost no arboreal
pollen and low lake levels all occurred in the Tagua Tagua record
s et al., 2005).
[89], indicative of very arid conditions (Valero-Garce
Across the Andes, NWA records such as Pastos Chicos [46] show
an increase of shrub/steppe pollen taxa and the absence of peat
sediments until downcutting suggests a decrease in vegetation
cover and sediment supply, interpreted as drier climatic conditions
between 6 and 4.2 ka BP (Tchilinguirian et al., 2014). Aridity may
have been even more intense at Antofagasta de la Sierra where
diatom assemblages at Laguna Colorado [62] exhibit a low planktonic to benthic species ratios, indicating a shallow lake from 7.2 to
3.5 ka BP. River flow at Antofagasta de la Sierra was also ephemeral
rather than perennial c. 6.7e4.8 ka BP, as inferred from diatom
proxies, due to increased aridity (Grana et al., 2016).
Further south, CR records also show contrasting evidence. The
Salinas 2 record [92] indicates wetter conditions as suggested by
elevated percentages of Ephedra and the presence of lacustrine clay
4.6. 6.0 ka BP
The Middle Holocene (Northgrippian) interval has a total of 30
records in the SCA which were compiled into our database (Fig. 9).
A large proportion of these records agree with each other as to
widespread negative moisture anomalies present throughout the
SCA. Some important exceptions in the CA and CR, however, are
also present. Three BA records show opposite trends. A36Cl exposure date on a ground moraine plateau on the eastern flank of
Nevado Sajama [6] indicates that ice must have retreated after 6 ka
BP, but the presence of glacial deposits in advanced positions indicates that it may not have been as dry and warm as current
conditions (Smith et al., 2009). In contrast, the Sajama ice core
shows enhanced dust deposition associated with an elevated
snowline and decreased net accumulation [5] (Thompson et al.,
1998). A sequence of alluvial sediments at Huana Khau [15] and a
lack of lacustrine sediments in the río Desaguadero are interpreted
as drier than today from 7.9 to 4.5 ka BP (Rigsby et al., 2005).
NA records such as Laguna Seca [8] shows a shift from clayey silt
to peat together with vegetation changes (i.e., Brassicaceae, Azorella) between 7 and 5 ka BP indicative of a lowering of the water
table and negative moisture anomalies (Baied and Wheeler, 1993).
Limnogeochemical evidence from Lago Chungara (lake core) [10]
shows that this period was possibly the most arid during the Holocene, indicating that the lake went through an “aridity crisis”
from ~8.6e6.4 ka BP which then recovered to modern levels by 6 ka
BP (Moreno et al., 2007). Despite the large uncertainties of the
Chungara age-depth model (Giralt et al., 2007) this arid period is
further constrained by AMS 14C dates on terrestrial charcoal from
nearby Hakenasa rock shelter, which shows that humans abandoned this site between ~7.9e6.9 ka BP, presumably in response to
increased aridity but then reoccupied it afterwards (Moreno et al.,
2009; Osorio et al., 2011).
CA records disagree over whether the interval was wetter or
drier than today. Lack of aquatic pollen and the presence of gypsum
layers in the Laguna Miscanti core [44] are interpreted as strong
negative moisture anomalies (Grosjean et al., 2001). Increased
aridity is also suggested by the stratigraphy and incision (downcutting) of in-stream wetlands deposits, along a 4.3-km reach of the
Río San Salvador between 6.4 and 3.5 ka BP (Tully et al., 2019).
Middle Holocene fine-sediment (in-stream wetland) deposits at
Quebrada de Puripica [38] have been interpreted as either having
formed by side-canyon damming during very arid conditions
(Grosjean et al., 1997a,b) or more recently, as evidence for elevated
groundwater tables during a wetter climate (Rech et al., 2003).
Slightly wetter conditions than present have also been inferred
from rodent midden records at Lomas de Tilocalar [47] and Pampa
Vizcachilla [36], which show evidence for the presence of sparse
puna shrubs, C4 annual plants (indicative of summer rainfall) and
other extralocal perennial taxa (Betancourt et al., 2000; Latorre
et al., 2003; 2003). The presence of steppe grasses at Pampa Vizcachilla [36] points to a two-fold rainfall increase (Latorre et al.,
2003). Further south in Quebrada La Zorra [53], rodent midden
pollen shows puna taxa (Adesmia) at 6.1 ka, which could indicate a
brief wetter episode (De Porras et al., 2017). Patchy primary halite
13
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 9. Records compiled for the 6 ka BP interval. See Supplementary Materials for list of site numbers and references.
4.7. 4.0 ka BP
layers in a core from a peat mound (Markgraf, 1983). Large increases in Poaceae, Larrea and Prosopis along with other Monte
Desert taxa occur in Agua la Cueva [82] all indicative of a strong
positive moisture anomaly (Navarro et al., 2010). In contrast, only
slightly further south a pollen record from Gruta el Indio [90]
suggests very arid conditions (Gil et al., 2005).
A total of 29 different records across the SCA were compiled for
this time interval. Interpretations are diverse, with 16 records
interpreted as positive moisture anomalies, seven as negative and
six showing modern climates (Fig. 10).
14
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 10. Records compiled for the 4 ka BP interval. See Supplementary Materialsfor list of site numbers and references.
the presence of increased moisture in the Desaguadero valley
(Rigsby et al., 2005). Just across the border, NA records show an
increase of aquatic and Andean arboreal pollen at Laguna Seca [8]
interpreted as evidence for a positive moisture anomaly (Baied and
Wheeler, 1993). Geolimnological proxies at Lago Chungara [10]
indicate environmental conditions similar to today (Moreno et al.,
2007).
BA records include enhanced dust concentration associated
with an elevated snowline and high concentration soluble aerosol
(SO2
4 ; CI ) at Nevado Sajama [5] suggestive of arid conditions
(Thompson et al., 1998). In contrast, exposure ages on moraines
show a glacial advance at Huaqui Jihuata [6] in the western
Cordillera between 4.7 and 3.3 ka BP (Smith et al., 2009). Lacustrine
strata at Huana Khaua [15] and other localities have been argued for
15
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
core from a peat mound at Salinas 2 [92] shows clay layers that
formed during a lake transgression coupled with a large decrease in
Ephedra and increases in Poaceae pollen, all indicative of increased
precipitation (wet condition) (Markgraf, 1983). Organic rich levels
formed by wetlands dominated by Cyperaceae, Juncaceae and
Poaceae are found along the distal facies of the Arroyo Bayo alluvial
fan at Agua Buena [85] are associated with a highly productive
environments and increased available moisture (Navarro et al.,
2010).
Most CA records indicate positive moisture anomalies. At Quebrada Puripica [38] the presence of perched, exposed fine-grained
sediments throughout this valley have been interpreted as periods of elevated groundwater tables and increased aquifer
recharge (Rech et al., 2003). Patchy primary halite occurs in Salar de
Atacama [42] is also indicative of wetter conditions (Bobst et al.,
2001). Slight increases of summer flowering grasses in rodent
middens have been interpreted as evidence for wetter conditions at
Lomas de Tilocalar [47] (Latorre et al., 2003). Elevated plant species
richness and the presence of extralocal Andean species at Sierra del
Buitre [58] have also been interpreted as evidence for wetter conditions in the coastal cordillera (Díaz et al., 2012). Pollen and geolimnological evidence from Laguna Miscanti [44] however, shows
climate similar to today (Grosjean et al., 2001). Incision (downcutting) of in-stream wetlands deposits suggest arid conditions
between 6.4 and 3.5 ka BP at Río San Salvador [35] (Tully et al.,
2019). Midden pollen records of prepuna and puna taxa, characterized for Amaranthaceae, Brassicaceae and Malvaceae at Quebrada Zorras [52], are indicative of drier conditions than present
(De Porras et al., 2017). In contrast, SA records show wet conditions
as indicated by a black peat layer in the alluvial section at Quebrada
ez et al., 2016). Geolos Sapos [55] and dated to 3.9 ka BP (Sa
limnological evidence from further south at Laguna del Negro
Franscisco [63] has been interpreted as indicating arid conditions
(Grosjean et al., 1997a,b).
SC records show increased snow cover and alluvial deposition at
several sites, including Los Loros [64], Las Rojas [74] and Junta Río
Toro [72], all of which are interpreted as evidence for strong positive moisture anomalies (Veit, 1996). Basal peats, which have been
14
C-dated to 4.4e3.9 ka BP in the Valeriana valley [68] have also
been interpreted as evidence for increased moisture (Grosjean
et al., 1998). This positive moisture anomaly is also present in the
Río Turbio Valley [75], where a second paraglacial episode occurred
after 7.8 ka BP, but most likely later than 5.5 cal ka BP, as dated by
wood present in debris cones. This episode is characterized by
increased fluvial transport capacity and reflects a change towards
wetter conditions (Riquelme et al., 2011). Swamp forests appear
and arboreal taxa such as Luma chequen and Escallonia revoluta
~
show that climate was similar to the present at Nague
I [80]
(Maldonado and Villagran, 2002) although nearby Palo Colorado
[81] has been interpreted as wetter conditions, based on the ren,
appearance of swamp forest taxa (Maldonado and Villagra
2006). At Laguna el Cepo [70], however, a change in mean grain
size and sorting parameters occurred between 4.1 and 2.2 ka BP,
suggest increased frequency of storms and relatively dry conditions
as compared to present (Tiner et al., 2018).
In contrast to most SC paleorecords, CC records mostly indicate
negative moisture anomalies (i.e., 1). The predominance of halophytes at Quintero II [83] before the modern swamp forest vegetation appeared, suggests arid conditions (Villa-Martinez and
n, 1997). An increased arboreal component and lower perVillagra
centages of Amaranthaceae have been interpreted as evidence for
increased moisture at Laguna Aculeo [86] but these percentages
were still below those seen in more recent sediments (Jenny et al.,
2002). Similar pollen assemblages have also been documented at
Tagua Tagua [89] and share the same interpretation: climate
became increasingly wetter, but not as wet as that inferred from
s et al., 2005).
more recent assemblages (Valero-Garce
NWA records show negative anomalies at 26 S with positive
anomalies further south. Sedimentary facies and diatom assemblages from a fluvial deposit at Antofagasta de la Sierra [60] are
interpreted as very arid conditions (Grana et al., 2016). Further
south, however, hygrophytic communities expanded (Asteraceae
and Cyperaceae), probably linked to fluvial cutoffs and back swamp
environments at La Estancada Creek [80] (Rojo et al., 2012). A short
4.8. Comparison of paleorecord time windows with the TRACE-21
model simulation
The comparison of our time windows with TRACE-21 model
simulations are described below (Fig. 11). There are no simulations
available with this model for 32 ka BP. The simulation shows a wet
LGM (21 ka BP), with higher rainfall than at present for the entire
SCA. While positive precipitation anomalies occur in the BA, this
contrasts with paleorecords from Laguna Miscanti in Chile and
Salar de Uyuni in Bolivia which indicate arid conditions. Positive
anomalies persist at 17 ka and are in agreement with most records,
all of which indicate very wet conditions across the entire SCA. At
14 ka the model suggests a positive precipitation anomaly in the BA,
~ a and Salar de
which disagrees with data from the Sajama, Tican
Uyuni paleorecords, all of which suggest arid conditions.
At 9.5 ka, modelled precipitation anomalies tend to be more
positive than the moisture anomalies observed in paleorecords
found south of the CA, where a clear tendency to more arid conditions is seen. The model results agree more with BA paleorecords,
where a trend of positive moisture anomalies is apparent. At 6 ka,
the modeling suggests similar to present conditions for the entire
SCA, which disagrees with the paleorecords as these show more
arid conditions in CC and NWA. In contrast, records from the Atacama and Altiplano vary between arid and wet. At 4 ka the model
shows very homogeneous precipitation anomalies throughout in
the SCA, versus the large spatial differences seen in the proxies.
5. Discussion
By compiling records into maps for any given time interval, we
provide an explicitly spatial view of past climate change across the
SCA during the late Quaternary. Although such comparisons can
reveal conflicting records at local spatial scales (i.e., within a same
basin or watershed), they can also aid in defining multi-scale responses to drivers of the climate system, such as local summer
insolation during the LGM or ENSO-like regime shifts during the
Holocene. We discuss these records and the implications of these
maps (temporal windows) for each of our selected intervals below.
5.1. Spatial extent and climate change drivers in the SCA between
32 and 4 ka BP
Our compilation of records at 32 ka BP shows wet conditions at
least as far north as 22 in the SCA. An increase in winter precipitation, either due to a northward shift of the austral winter frontal
systems or to an increased frequency of cutoff lows from the Pacific,
which today reach 23 S (Vuille and Ammann, 1997) is one possible
explanation. BA proxies suggest dry anomalies in the Desaguadero
Valley, which have been attributed to decreased summer insolation
and a weakened SASM (Rigsby et al., 2005). Placzek et al. (2013)
suggest that the wet and dry phases over the Bolivian Altiplano
may not be directly related to astronomical forcing, at least over the
last 130 ka but may be related to cold stadials in the north Atlantic
during the last 100 ka. Shoreline features around the Uyuni basin
are lacking for this interval (46e24 ka ) (Placzek et al., 2006, 2013),
16
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Fig. 11. TRACE-21 model simulations for the South-Central Andes for A) 21.5e20.5 ka BP, B) 17.5e16.5 ka BP, C) 14.5e13.5 ka BP, D) 10e9 ka BP, E) 6.5e5.5 ka BP, and F) 4.5e3.5 ka BP.
except for the “Minchin” phase (Minchin, 1882; Steinmann et al.,
1904; Servant and Fontes, 1978; Fornari et al., 2001) dated to
73e30 ka . Using UeTh ages on 14C-dated “Minchin” deposits,
Placzek et al. (2006) show that this phase is probably an artifact of
radiocarbon dating of much older carbonate deposits. Positive
moisture anomalies at 32 ka BP are present in records from the CA
and southwards, as suggested by the Salar de Atacama salt core and
Quebrada del Chaco midden pollen record. It is very likely that
these are due to northward shifts of the austral winter frontal
systems, which today are rare but can occur in latitudes as far north
as 23 - 25 S (Vuille and Ammann, 1997). This is further supported
by the fact that positive moisture anomalies are present in all records south of these latitudes and on both sides of the Andes.
Our analyses show major differences in moisture anomalies
throughout the SCA at 21 ka BP during the global LGM. For the high
Andes (>3000 m a.s.l) between 17 and 23 S, most records suggest
either very wet or wet conditions, which have partly been attributed to maximum summer insolation in the southern hemisphere
(~20 ka) (Berger and Loutre., 1991; Baker et al., 2001; Rigsby et al.,
2005). Greater warming caused by increased insolation would favor
maximum deep convection over Central South America (e.g., the
Chaco Low), intensifying the South American monsoon (Zhou and
Lau, 1998; Placzek et al., 2006, 2013). However, the occurrence of
major paleolake phases in the BA coupled with hydrologic budget
models, were not always consistent with the maximum local
summer insolation, at least over the last 130 ka BP (Placzek et al.,
2013).
Another proposed mechanism for explaining the increase in
precipitation in the high Andes (>3000 m a.s.l) at 21 ka BP is
derived from the modern synoptic circulation regime of Andean
climate (Garreaud et al., 2003) and simulated for the LGM (Vizy and
Cook, 2007). These studies point out that precipitation variability in
the SCA is not directly controlled by moisture changes in the
Amazon Basin but rather by favorable easterly advection of moisture. This could explain why previous LGM climate models show
more aridity in relation to today in the Amazon while it was wetter
in the high Andes (>3000 m a.s.l) (Van der Hammen and
Hooghiemstra, 2000; Vizy and Cook, 2007). Increased precipitation over the Altiplano (high Andes) occurs when the absence of
convection over the Amazon in the austral spring strengthens the
zonal geopotential height gradient resulting in anomalous easterly
flow between the Amazon and Altiplano. Instead of the low-level
flow paralleling the Andes (e.g., northwesterly), the flow impinges more directly onto the eastern slopes of the Andes during
the LGM, where it rises and transports moisture from the Amazon
up into the mountains, enhancing convection over the Andes
(Fuenzalida and Rutllant. 1987; Chaffaut et al., 1998; Garreaud,
1999; Vizy and Cook, 2007). A lake core in northern Amazonia
suggests that during glacial times there was no positive correlation
between precipitation in the Altiplano and western Amazonia
(Cordeiro et al., 2011). Finally, PMIP3 LGM simulations suggest a
reduction Amazon Basin summer rainfall (Bermanm et al., 2016)
linked to colder tropical Atlantic Sea surface temperatures and a
reduction in ocean to continent advection (Arraut and Satyamurty,
2009).
Our LGM map also shows contrasting reconstructions for the CA
with both wet and dry anomalies in the high Andes (17 e27 S,
3000e4500 m a.sl.). Such contrasting local differences could be due
to the same climate mechanisms proposed for the entire Altiplano
(17 -27 S) i.e., that summer precipitation extended south of 22 S
along the Pacific side of the Andes, which even today occurs to 27 S
(Houston, 2006). The spatial scale representation of each proxy
(from local to basin-wide) could also be important in explaining
these discrepancies. Rodent middens are local records of climatic
variability at (sub) decadal timescales (Grosjean et al., 2003)
although the duration of a depositional episode is generally
17
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
airflow across the tropics, causing a reduction of moisture availability in the tropical lowlands to the east (Garreaud et al., 2003).
Another proposed mechanism is the change in aridity over the
Amazon (Vizy and Cook, 2007). Several authors (Betancourt et al.,
2000; Garreaud et al., 2003; Placzek et al., 2006, 2013; Gayo
et al., 2012) have proposed that the Tauca phase could have been
~ a-like conditions in the equatorial
due by the prevalence of La Nin
Pacific between 18 and 13 ka (Palmer and Pearson, 2003). ENSO is
strong driver of modern interannual variability today over the Altiplano and northern Chile (Vuille and Keimig, 2004) and wet years
~ a years
in this area are often concurrent with strong La Nin
(Houston, 2006) indicative of a strong link between moisture levels
in the SCA and tropical Pacific SST gradients. Several researchers
have proposed high-latitude drivers in phase with events in the
North Atlantic (for example, Heinrich Events), and propose that
these exert a first-order control over Altiplano hydroclimate over
the last 130 ka (Placzek et al., 2013; Martin et al., 2018) although not
all paleolake events are coeval with Heinrich events (Placzek et al.,
2013).
As previously mentioned, almost all Atacama Desert records
(except Laguna Miscanti) show wet anomalies coeval with the
Tauca phase during CAPE I. Increased summer rainfall may have
reached as far south as 30 S, in the Turbio and Encierro valleys
(Aguilar, 2010; Riquelme et al., 2011). This would suggest that CAPE
I was a probably the most extensive period of increased summer
precipitation examined here, even at subtropical latitudes.
Further south in CC, the pollen record from the Tagua Tagua lake
core also suggests an increase in winter precipitation between 19.5
s et al.,
and 17 ka BP and attributed to the local LGM (Valero-Garce
2005). Indeed, deglaciation began in southern Chile at 17.8 ka BP
(Palacios et al., 2020). This would imply that temperatures as well
as rainfall increased throughout the region after the global LGM.
Although most BA records suggest very wet conditions attributed to the Tauca phase, there is some evidence for increased
~ a event dated from 14.1 to 13.8
aridity at 14 ka BP, such as the Tican
ka BP (Sylvestre et al., 1999). This event, however, was considered
restricted to the Uyuni and Coipasa basins (Sylvestre et al., 1999;
Placzek et al., 2006). Records form the Atacama Desert have shown
that this arid event may have been considerably more widespread
and could have impacted the height of the water table in places
such as Salar de Punta Negra, from 13.8 to 12.7 ka BP (Quade et al.,
2008) and Alto Varas from 14.5 to 12.2 ka BP (S
aez et al., 2016). This
arid event likely terminated CAPE I, and prolonged conditions like
~ o phenomenon (which today weaken
those of a persistent El Nin
the flow of easterly moisture across the Andes) have been invoked
to explain its onset (Kienast et al., 2006; Makou et al., 2010).
Most records in the Atacama (18 -27 S) and NWA (Guayatayoc)
suggest positive moisture anomalies at the end of the CAPE I at 14
ka BP (Grosjean et al., 2001; Moreno et al., 2007; Gayo et al., 2012;
Steinmann et al., 1904; Mujica et al., 2015; Pfeiffer et al., 2018).
A14C-dated rodent midden record from Río Salado (22 S) shows the
presence of arid plant communities with rainfall amounts similar to
today at 14.2 ka BP (Latorre et al., 2006). In contrast, midden pollen
from the Quebrada del Chaco record in the SA shows a mixed
regime of both summer and winter rainfall increases from 17 to 14
ka BP (Maldonado et al., 2005). This positive anomaly also occurs in
Semiarid Chile, where glacial advances occurred from 15 to 14 ka BP
in the headwaters of the Encierro and Turbio valleys and have been
attributed to increased summer rainfall (Zech et al., 2006; Aguilar,
2010; Riquelme et al., 2011). A colder and wetter climate is suggested by an increase in forest pollen taxa, associated with a
stronger influence of the westerlies at Laguna Tagua Tagua in
Central Chile, which shows colder and wetter climates between ~14
s et al., 2005).
and 11.5 ka BP (Valero-Garce
unknown (Betancourt and Saavedra, 2002). While limited spatially,
these records can be collated across a large area (several basins-see
lez-Pinilla et al., 2021) and are easily 14C-dated. This is very
Gonza
different from the lake records, which integrate climate from
decadal to sub-millennial timescales across an entire basin and
therefore show low-frequency variability (Grosjean et al., 2001,
2003). And there are other known limitations associated with lake
records in the Andes, such as unknown and variable reservoir effects in the Chilean altiplano (Geyh et al., 1999; Grosjean et al.,
2001). Clearly, more LGM records are needed from the Central
Atacama Desert to resolve these issues using different geochronological techniques such as OSL and UeTh dating.
SA records all show wetter LGM climates than today. Increased
runoff and higher water tables resulting from precipitation increases have been proposed to explain the presence of more diverse
plant communities in the hyperarid core of Atacama Desert (Díaz
et al., 2012). Just as for the 32 ka BP interval, an increase in
winter rainfall associated with a northward expansion of the
westerlies likely occurred at least as far north as 25 S, as suggested
by Maldonado et al. (2005). Indeed, increased presence of Nothofagus dombeyi-type pollen at Tagua Tagua during the LGM has
attributed to a northward shift in the westerlies (e.g., Heusser,
1990). Cosmogenic ages on glacial moraines in the Cachapoal valley provide further evidence for wet conditions, as glaciers are
mostly sensitive to precipitation at these latitudes (34-35 S, Zech
et al., 2008; Herrera, 2016; Charrier et al., 2019).
Overall, our maps show that records with positive moisture
anomalies occur throughout the SCA at 17 ka BP, perhaps the
wettest interval over the last 32 ka BP discussed here. Positive
moisture anomalies are so widespread throughout the SCA at this
time that it has been termed the CAPE (Central Andean Pluvial
Event) (Latorre et al., 2006, Quade et al., 2008; Placzek et al., 2009;
Gayo et al., 2012; Workman et al., 2020; Pfeiffer et al., 2018). The
term “CAPE” was originally coined for the Atacama Desert by
Latorre et al. (2006) as the “Central Atacama Pluvial Event” based
on rodent middens that showed two distinctive pluvials at
17.5e16.2 ka BP and 11.7e9.5 ka BP interrupted by an arid event at
14.2 ka B P (see Fig. 7). Quade et al. (2008) and Placzek et al. (2009)
expanded this term to “Central Andean Pluvial Event” based on
paleowetland evidence from the Salar de Punta Negra basin (3070,
m a.s.l, 24.5 S), and established two phases of increased available
moisture at 15.9e13.8 ka BP and 12.7e9.7 ka BP that were found
throughout the SCA between 18 and 27 S. Based on evidence for
positive moisture anomalies from the northern Atacama, Gayo et al.
(2012) and Pfeiffer et al. (2018) revised the CAPE chronology, which
were divided into CAPE I (17.5e14.2 ka BP) and CAPE II (13.8e9.7 ka
BP) separated by an arid period at 14.2e13.8 ka BP. More recent
evidence based on mean annual rainfall anomalies derived from
rodent body size changes show that in the CA, CAPE I and II
occurred from 15.9 e 14.8 and 13.0e8.6 ka BP, respectively
lez-Pinilla et al., 2021).
(Gonza
During CAPE I, paleolake Tauca spread to cover 60,000 km2 of
the southern Bolivian Altiplano (Sylvestre et al., 1999; Baker et al.,
2001; Rigsby et al., 2005; Blard et al., 2009; Placzek et al., 2006,
2013) between 18.1 and 14.1 ka, and shorelines formed ~140 m
above the current surface of the Uyuni basin (Placzek et al., 2006,
2013). As previously mentioned, several explanations have been
proposed for the formation of this and other paleolakes over the
last 130 ka, including summer insolation and links to North Atlantic
stadials (Baker et al., 2001; Rowe et al., 2002; Fritz et al., 2004;
Rigsby et al., 2005). Changes in global temperature have also been
proposed based on climate models (Blodgett et al., 1997; Garreaud
et al., 2003) that show the impacts of warming contrasts between
the northern and southern hemispheres during glacial periods.
Such contrasts generate moisture anomalies that increase easterly
18
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
increased moisture during the Middle Holocene (Rech et al., 2002,
2003). Further evidence for a brief interlude of increased moisture
in the Central Atacama is present in the Salar de Atacama salt core
(Bobst et al., 2001).
A pollen record from Pastos Chicos and the absence of organic
pedogenic horizons in Antofagasta de la Sierra rivers also imply
negative moisture anomalies in NWA at 6 ka BP (Tchilinguirian
et al., 2014; Grana et al., 2016). Pulses of increased rainfall also
occurred in these records (Grana et al., 2016; Morales, 2011). MidHolocene NWA aridity has been attributed to weakening of the
South American low-level jet (SALLJ) since when strengthened, it
would lead to an intensification of the South Atlantic convergence
zone (SACZ) and favor the penetration of cold fronts with an
improved convection area ahead in the exit region from SALLJ
(bringing tropical humid air masses from the Amazon to southern
Brazil and northern Argentina). The SALLJ mostly occurs during the
warm season south of 20 S, whereas to the north it occurs
throughout the year (Marengo et al., 2004; Vera et al., 2006).
SC and CC also show arid climates in our 6 ka BP maps, except for
a diatomite layer in the Tagua Tagua core (Fig. 9), which indicates
s et al., 2005). As with the 9.5
present-day conditions (Valero-Garce
ka BP interval, a southward displacement of the westerly storm
tracks and the strengthening of the SPH associated with strong La
~ a-like conditions are commonly cited explanations (Maldonado
Nin
n, 2002, 2006; Carre
et al.,
et al., 2005; Maldonado and Villagra
2012; Tiner et al., 2018). In contrast, CR records suggest wet conditions (Markgraf, 1983). These conditions could be associated with
increased advection of moisture from the Atlantic (Hebe, 1997; Gil
et al., 2005), whereas south of 33 S frontal systems decreased in
frequency (Gil et al., 2005).
Finally, our 4 ka BP map shows increases in moisture throughout
the SCA. For instance, ice advanced from 4.7 to 3.3 ka BP at Nevado
Sajama (Smith et al., 2009), even though the Sajama ice core shows
increased aridity (Thompson et al., 1998). Differential responses
across west versus east facing slopes may partly explain this (Smith
et al., 2009). Positive moisture anomalies at 4 ka BP have been
linked to increased ENSO activity throughout the SCA, which was
more pronounced compared to the Middle Holocene (Sandweiss
et al., 2001; Mix et al., 2001; Donders et al., 2008). Even so, ENSO
variability decreased between 4.5 and 3.5 ka BP compared to its
variability after 3.5 ka BP (Sandweiss et al., 2001; Mix et al., 2001;
Donders et al., 2008). Another possible explanation for increased
moisture compared to the 6 ka BP has been linked to increased solar
irradiance, which in turn would drive increased activity of the
SASM at least on millennial timescales (Novello et al., 2016; De
Porras et al., 2017).
Positive moisture anomalies also prevailed in SA and CC records
n,
at 4 ka BP (Veit, 1996; Jenny et al., 2002; Maldonado and Villagra
2006; Riquelme et al., 2011) associated with a weakening of the
n, 2002, 2006; Maldonado et al.,
SPH (Maldonado and Villagra
2005). CMIP5/PMIP3 simulations show that during cold periods
(such as during the European Little Ice Age) the SPH contracts, with
expansion occurring during warmer periods (Flores-Aqueveque
et al., 2020). This implies that global temperatures drove a weakening of the SPH and increased available moisture across the region
at the onset of the Late Holocene (cold period) and its expansion
with increased aridity during the Middle Holocene (warm period).
Some paleorecords show arid conditions, however, and these have
~ a-like conditions from ~5 to 4 ka BP (Carre
been linked to La Nin
et al., 2012). Paleotemperatures estimated from the marine
bivalve Mesodesma donacium d18O values show colder SSTs than
present, associated with increased coastal upwelling and therefore
an intensified SPH and strengthened Humboldt Current system
BA records at 9.5 ka BP begin to show shifts towards increased
aridity, which has been attributed to a summer insolation minimum at 10 ka BP (Rigsby et al., 2005) although this was not the
most arid phase in the Bolivian Altiplano (which occurs during the
Middle Holocene). NA records also indicate arid conditions, evidenced by decreases in the lake levels in the Chilean altiplano ca.
9.5 to 9 ka BP (Moreno et al., 2007; Baied and Wheeler, 1993). CA
and SA records show widespread aridity, with the disappearance of
Andean steppe taxa and the prevalence of prepuna species in rodent midden records at Cerros de Aiquina, Lomas de Tilocalar,
n de Tuina, El Hotel- Paso Barros Arana and Quebrada del
Cordo
Chaco III (Latorre et al., 2003; 2003; Maldonado et al., 2005). This
was mostly coeval with a decrease in groundwater levels along the
río San Salvador, Salar de Punta Negra and in Alto Varas, around 9.8
ka BP (Quade et al., 2008; S
aez et al., 2016; Tully et al., 2019), which
has been attributed to the end of CAPE II (Quade et al., 2008; Gayo
ez et al., 2016; Tully et al., 2019).
et al., 2012; Sa
The heterogeneity in CA proxies in relation to moisture anomalies and the end of CAPE events could reflect the temporal scale
preserved in each record, which can vary from centuries to
millennia. As previously discussed, lacustrine proxies (lake cores)
integrate climate from decades to centuries and therefore show
low-frequency variability (Grosjean et al., 2001, 2003). Groundwater tables also respond more slowly to onset of drought (c. 500
yrs, Grosjean et al., 2003; Pigati et al., 2014). In contrast, rodent
middens likely represent much shorter timescales, although the
length of a durational episode needs to be assessed on a midden-tomidden basis.
Our 9.5 ka BP map shows that SC and CC climates were clearly
drier than today further south (Jenny et al., 2002; Maldonado and
n, 2002; Valero-Garce
s et al., 2005; Zech et al., 2006;
Villagra
Maldonado et al., 2005; Ortega et al., 2012; Tiner et al., 2018)
(Fig. 8). Explanations for this increase in aridity involve a southward shift of winter storm tracks entrained within the southern
westerlies and a strengthening of the SPH (Jenny et al., 2002;
n, 2002, 2006; Valero-Garce
s et al., 2005;
Maldonado and Villagra
Maldonado et al., 2005; Tiner et al., 2018). The onset of strong La
~ a-type conditions between 13 and 8.6 ka BP could have also
Nin
increased regional aridity (Vargas et al., 2006; Ortega et al., 2012).
Arid conditions also prevailed east of the Andes in CR, although
aridity intensified here during the Middle Holocene and has been
n, 1993;
attributed to the strengthening of the SPH (Villagra
Yacobaccio and Morales, 2005). In contrast, NWA records all suggest wet conditions associated with increased summer rainfall
(Yacobaccio and Morales, 2005; Tchilinguirian et al., 2014; Grana
pez
et al., 2016), possibly linked to a CAPE-type phenomenon (Lo
and Galli, 2015). On the other hand, these wetter climates may
have been caused by ENSO-like conditions and sea surface temperature variability across the tropical Atlantic (Trauth et al., 2000,
2003).
Our 6 ka BP map reveals that two-thirds of all records show
strong negative moisture anomalies across the SCA (Baied and
Wheeler, 1993; Grosjean et al., 1997a,b; Thompson et al., 1998;
Grosjean et al., 2001; Rigsby et al., 2005; Placzek et al., 2009; Tully
et al., 2019). Widespread aridity during the Middle Holocene has
~ a-like conditions (Koutavas
been attributed to a prevalence of La Nin
et al., 2012; Sandweiss et al.,
et al., 2002; Donders et al., 2008; Carre
2020). Nevertheless, a third of all the records show positive moisture anomalies during this interval. For example, increased moisture occurred from 6.5 to 5 ka BP at Laguna Miscanti, which have
been attributed to centennial-scale moisture “pulses” (Grosjean
et al., 2003). Paleowetlands formed in the Puripica, Tulan and Río
Salado drainages at the same time, and are cited as evidence for
19
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
et al., 2003) whereas deeper lakes (Chungar
a or Miscanti) will
not reflect very high-frequency changes or specific events, and
therefore mostly record long-term climatic trends (Vuille and
Baumgartner, 1993; Grosjean et al., 2003).
Furthermore, completely opposite interpretations by different
authors often arise for the same record or locality. The fine-grained
diatomites, black mats and other wetland deposits at Quebrada
Puripica are a good example (Grosjean et al., 1997a,b; Rech et al.,
2003). These authors propose different theoretical frameworks to
interpret the same record. Grosjean et al. (1997) suggest that an
increase in the size and frequency of discharge events results in
increased erosion of the active channel during wet periods with
fine-grained sediments deposited during arid phases. This is contrary to a base level model used by Rech et al. (2003), which interprets such periods of downcutting as arid phases. Such base level
models have been applied successfully to interpret paleowetlands
across the Atacama Desert (Rech et al., 2002; Rech et al., 2003;
ez et al., 2016; Pfeiffer et al., 2018; Tully et al.,
Quade et al., 2008; Sa
2019; Workman et al., 2020).
Dating uncertainties remain an important issue throughout the
SCA. Chronologies of glacial advances and retreat are based mostly
on cosmogenic dating techniques in which sample ages can vary
significantly depending on the rate of cosmogenic production
considered. For example, Zech et al. (2006) considers a production
rate of 5.25 10Be Atoms a1 g (Desilet and Zedra, 2003) for the
Encierro valley glacial features, whereas Aguilar (2010) uses a rate
of 4.5 10Be atoms a1g (based on Blaco et al., 2008). Unknown and/
or variable reservoir effects are widely known issues for radiocarbon age-depth models in the altiplano lakes (Geyh et al., 1999;
Moreno et al., 2007). Many of these issues have been resolved by
UeTh dating of carbonates, which for the altiplano often show that
samples are much older than the corresponding radiocarbon ages
(Placzek et al., 2006). Indeed, most researchers recommend that
radiocarbon ages on carbonates (e.g., stromatolites, bioherm, tufas,
marls, etc.) in the SCA should be considered maximum ages (Geyh
et al., 1999; Grosjean et al., 2001; Pfeiffer et al., 2018).
Despite these many limitations, our maps reveal overall how
different regions across the SCA responded to different drivers
across the last glacial-interglacial cycle (Table 1). For example, records show increased moisture throughout the southern sector of
the SCA (18 -34 S), while the high Andes (17 -27 S) remained dry
et al., 2012). This agrees with explanations for the presence
(Carre
of negative moisture anomalies in other records, such as Laguna el
Cepo (Tiner et al., 2018).
Positive moisture anomalies also occur throughout the CR in
Argentina (Markgraf, 1983; Navarro et al., 2010; Rojo et al., 2012),
although these have been linked to increased advection of moisture
from the Atlantic. Such advection today is strongest in September
and remains until January (Hebe, 1997), whereas tropically sourced
moisture (Amazon and Chaco basins) is predominant in summer.
These factors can enhance each other and lead to higher rainfall in
the summer (Hebe, 1997). Increased frequency of frontal systems of
Pacific origin (winter rainfall) have also been cited to explain
increased Late Holocene moisture as these would be linked to a
n., 1993).
weakening of the SPSH (Capitanelli, 1967; Villagra
5.2. Proxy sensitivity and spatial integration
Our maps show important differences between regional and
local records for any given time interval (Table 1). Such differences
can arise due to multiple factors, such as spatial representation,
proxy sensitivity, chronological uncertainties, and record temporal
resolution (e.g., Grosjean et al., 2003). For example, records that
represent entire watersheds (59), such as salt or lake cores
including the Salar de Atacama, Uyuni-Coipasa basin, Lago
and Laguna Miscanti are often (but not always) at odds
Chungara
with records that are more spatially restricted, such as rodent
middens (34) (Sylvestre et al., 1999; Bobst et al., 2001; Fornari et al.,
2001; Grosjean et al., 2001; Placzek et al., 2006; Moreno et al.,
2007) (Latorre et al., 2003; 2003, 2006; Maldonado et al., 2005;
Díaz et al., 2012; Mujica et al., 2015). As previously discussed, this
difference in spatial representation is also accompanied by temporal differences as well, depending on the resolution of the record
which can vary from decades to centuries. Paleosols and records of
groundwater table variations (i.e., paleowetlands) have inherent
lags associated with system response times and therefore respond
very slowly to climatic variations (Veit, 1996; Grosjean et al., 2003;
Pigati et al., 2014). Such lags have been tentatively quantified for
groundwater systems in the Atacama, and these systems typically
respond on the order of 500 years (Rech et al., 2016). Otherwise, the
same type of proxy may respond differently, as in the case of large
salt flats likely respond immediately to precipitation (Grosjean
Table 1
Interregional comparison of paleorecords across the South-Central Andes for all time intervals.
20
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
from Central Atacama to Central Chile (35 S), unlike the NWA.
Window 6 ka BP shows an extended or very arid period on both
sides of the Andes. Finally, at 4 ka, we see humid conditions in the
Bolivian Altiplano, Central Atacama, Southern Atacama and Cuyo
Region, while Northern Atacama and Central Chile, the conditions
become similar to the current ones, except in the Argentine
Northwest which has arid conditions.
at 32 ka. Another example is at 4 ka BP, where NA and CC records
indicate conditions similar to the present, while the other records
show wet conditions along the SCA.
5.3. TRACE-21 model simulation
The TRACE-21 model is forced by orbital and insolation changes
as well as freshwater discharges into the North Atlantic, which
directly affects Atlantic Meridional Overturning Circulation (AMOC)
(Liu et al., 2009; Karger et al., 2021). This discharge can be associated with Heinrich events, such as Heinrich 1, which occurred just
before the 17 ka window (17.5 ka). This time window was a wet
period especially in the Altiplano and the elevated areas of the
Atacama (>3000 m.a.s.l) (Fig. 11), mostly because ice discharge in
the North Atlantic is linked to the intensification of the South
American Monsoon and its southward displacement (Placzek et al.,
lez-Pinilla et al.,
2013; Stríkis et al., 2018; Martin et al., 2018; Gonza
2021). This mechanism generates greater transport of moisture
from the tropical Atlantic and a southward shift of the Intertropical
Convergence Zone (Broccoli et al., 2006), which int turn leads to
greater advection of moisture to the Andes (Garreaud et al., 2003).
This coincides with almost all the records in that sector (17 ka-time
window). On the other hand, since the model is strongly linked to
the forcings that affect the northern hemisphere (AMOC) and these
are strongly represented in the Late Pleistocene, models such as
TRACE-21 or PIMP do a poor job of adequately simulating Holocene
lez-Pinilla et al., 2021). Adding to this
records (also see Gonza
problem is the fact that the models are run at a global scale and
downscaling becomes more complex.
Author contribution statement
All authors have made substantial contributions to this
submission.
This manuscript was based on the creation of both a database
and a state-of-the-art review of the records in the south-central
ctor OrelAndes, which were carried out by the lead author (He
s). Claudio Latorre and Juan Luis García collaborated in
lana Corte
the interpretation of the proxies and discussion of the manuscript,
respectively. Fabrice Lambert carried out the paleoclimatic
modeling. Claudio Latorre secured funding.
Declaration of competing interest
The authors declare that they have no known competing
financial interests or personal relationships that could have
appeared to influence the work reported in this paper.
Data availability
The data that has been used is confidential.
6. Conclusions
Acknowledgments
The meta-analysis presented here of all published records from
the South-Central Andes suggest how proxies used to reconstruct
past climate change have different sensitivities to climatic variability and/or different geographic representativity (record is a
local or regional representation of paleoclimatic conditions). Dating
issues clearly can account for some of the discrepancies seen in our
maps, especially between proxies from very similar areas, most
notably in the Salar de Uyuni basin. These issues include problems
variable reservoir effects in 14C-dated records or evolving parameter estimation with cosmogenic dating techniques. By aggregating
different records with specific time intervals, our analyses may in
fact compound these issues at the local scale (i.e. within a basin) but
can also provide further insights into the spatial signals across a
broader area, not only revealing overall regional climate responses
but also how these shifts over time in intensity and direction (e.g.,
during the Middle and Late Holocene a drier Central Chile almost
always correlates with wetter climates directly across the Andes in
the Cuyo Region but not during the Late Pleistocene (in the 32e14
ka BP maps), when it was much wetter everywhere.
Our maps based on specific temporal windows show that at 32
ka BP climate was more wetter than today in most of the SCA (22 35 S), except in the Desaguadero Valley and the Uyuni basin in the
Bolivian Altiplano. In contrast, the 21 ka BP map shows wetter
conditions in most of the Bolivian Altiplano, but other records from
Northern Chile show drier climates than today. All LGM records
from the Central Atacama show positive anomalies on the western
side of the Andes, as opposed to the Cuyo Region in the eastern
Andes. The 17 ka BP map shows very wet conditions throughout the
SCA, as does the 14 ka BP map, although this presents some
important exceptions to this trend, which are inferred at the sites of
Quebrada Maní, Río Salado and in the Domeyko mountain range in
Southern Atacama. At 9.5 ka BP, the Bolivian Altiplano and Northern
Atacama show wet anomalies, which change more arid southward
We thank ANID FONDECYT grants 1191568, 1201786 (Chile
grants) for funding as well as ANID PIA grant FB210006 to the
Institute of Ecology and Biodiversity. HO received an ANID Doctoral
Fellowship used to complete this study. Additional funding was
provided by ANID FSEQ210021 and Millennium Nucleus UPWELL
(ANID NCN19_153) to CL.
Appendix A. Supplementary data
Supplementary data to this article can be found online at
https://doi.org/10.1016/j.quascirev.2023.108174.
References
n y transporte de material en la vertiente Occidental de los
Aguilar, G., 2010. Erosio
Andes semi
aridos del Norte de Chile (27 -32 S): desde un enfoque a gran escala
n cuaternaria de un Sistema fluvial. In:
temporal y espacial, hasta la evolucio
lica del Norte, 1Thesis to qualifity for the PhD in Geology at Universidad Cato
317.
Arraut, J., Satyamurty, P., 2009. Precipitation and water vapor transport in the
Southern Hemisphere with emphasis on the South American region. J. Appl.
Meteorol. Climatol. 48 (9), 1902e1912.
Baied, C., Wheeler, J., 1993. Evolution of high Andean Puna ecosystems: environment, climate, and culture change over the last 12000 years in the Central
Andes. Mt. Res. Dev. 13 (2), 145e156.
Baker, P., Rigsby, C., Seltzer, G., Fritz, S., Lowenstein, T.K., Bacher, N., Veliz, C., 2001.
Tropical climate changes at millennial and orbital timescales on the Bolivian
Altiplano. Nature 409.
Berger, A., Loutre, M.F., 1991. Insolation values for the climate of the last 10 million
years. Quat. Sci. Rev. 10, 297e317.
Bermanm, A., Silvestri, G., Tonello, M., 2016. Differences between last glacial
maximum and present-day temperature and precipitation in southern south
America. Quaternary Sci. Rev. 150, 221e233.
Betancourt, J., Latorre, C., Rech, J., Quade, J., Rylander, K., 2000. A 22,000-year record
of monsoonal precipitation fro northern Chile0 s Atacama Desert. Science 289,
1542e1546.
todo
Betancourt, J., Saavedra, B., 2002. Paleomadrigueras de roedores, un nuevo me
rica. Revista Chilena
para el estudio del Cuaternario en zonas aridas de Sudame
21
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
Desert (northern Chile). Quat. Res. 52, 143e153.
Gil, A., Zarate, M., Neme, G., 2005. Mid Holocene paleoenvironments and the
archaeological paleorecord of southern Mendoza, Argentina. Quat. Int. 132,
81e94.
s, B., Pueyo, J.J., GonGiralt, S., Moreno, A., Bao, R., Saez, A., Prego, R., Valero-Garce
zalez-Samperiz, P., Taberner, C., 2007. Statistical approach to disentangle environmental forcings in a lacustrine record: the Lago Chungara case (Chilean
Altiplano). J. Paleolimnol. 40 (1), 195e215.
lez-Pinilla, F.J., Latorre, C., Rojas, M., Houston, J., Rocuant, M.I., Maldonado, A.,
Gonza
Santoro, C.M., Quade, J., Betancourt, J.L., 2021. High-and low-latitude forcings
drive Atacama Desert rainfall variations over the past 16,000 years. Sci. Adv. 7
(38), eabg1333. https://doi.org/10.1126/sciadv.abg1333.
Grana, L., Tchilinguirian, P., Olivera, D., Laprida, C., Maidana, N., 2016. Síntesis
paleomabiental en Antofagasta de la Sierra: heterogeniedad ambiental y ocu~ os cal AP. Intersecc. Antropol. 4,
paciones humanas en los últimos 7200 an
19e32.
~ ez, L., Cartajena, I., Messerli, B., 1997a. Mid-Holocene climate and
Grosjean, M., Nun
culture change in the Atacama Desert, northern Chile. Quat. Res. 48, 239e246.
Grosjean, M., Valero-Garces, Geyh, A., Messerli, B., Schotterer Schreier, H., Kelts, K.,
1997b. Mid and late Holocene limnogeology of Laguna del Negro Francisco,
northern Chile, and its palaeoclimatic implications. Holocene 7 (2), 151e159.
Grosjean, M., Geyh, M.A., Messerli, B., Schreier, H., Veit, H., 1998. A late Holocene
(<2600 BP) glacial advance in the south-central Andes (29 S), northern Chile.
Holocene 8 (4), 473e479.
Grosjean, M., Van Leeuwen, J.F.N., Van der Knaap, W.O., Geeyh, M., Ammann, B.,
~ ez, L., Valero-Garce
s, B., 2001. A 2200O 14C year BP
Tanner, W., Messerli, B., Nun
sediment and pollen record of climate change from Laguna Miscanti (23 S),
northern Chile. Global Planet. Change 28, 35e51.
~ ez, L., 2003. From proxy data to paleoGrosjean, M., Cartajena, I., Geyh, M., Nun
climate interpretation: the mid-Holocene paradox of the Atacama Desert,
northern Chile. Palaeogeography, Paleaeoclimatology, Palaeoecology 194,
247e258.
gimen de la precipitacio
n en Argentina
Hebe, M., 1997. Factores que determinan el re
Subtropical. In: PhD Thesis in Atmospheric Science. Universidad de Buenos
Aires.
n de las altitudes de las líneas de equilibrio en glaciares
Herrera, M., 2016. Estimacio
~ a para el último ciclo glacial-interglacial en los Andes de Santiago,
de montan
Chile Central. Tesis de Posgrado, Universidad de Chile, Departamento de Geología, p. 171.
Heusser, C., 1990. Ice age vegetation and climate of subtropical Chile. Palaeogeography, Palaeoclimatology, Palaeoecology 80, 107e127.
Hogg, A., Hua, Q., Blackwell, P., Niu, M., Buck, C., Guilderson, T., Heaton, T., Palmer, J.,
Reimer, P., Reimer, R., Turney, C., Zimmerman, S., 2013. SCHAL13 Southern
hemisphere calibration, 0-50.000 years Cal BP. Radiocarbon 55 (4), 1889e1903.
Houston, J., 2006. Variability of precipitation in the Atacama Desert: its causes and
hydrological impact. Int. J. Climatol. 26, 2181e2198.
Hughes, P., Gibbard, P., Ehlers, J., 2013. Timing of glaciation during the last glacial
cycle: evaluating the concept of global Last Glacial Maximun (LGM). Earth Sci.
Rev. 125, 171e198.
Jenny, B., Valero-Garces, B., Villa-Martinez, R., Urrutia, R., Geyh, M., Veit, H., 2002.
Early to mid-Holocene aridity in central Chile and the southern westerlies: the
Laguna Aculeo record (34 S). Quat. Res. 58 (2), 160e170.
Jordan, T., Kirk-Lawlor, N., Blanco, N., Rech, J., Cosentino, N., 2014. Landscape
modification in response to repeated onset of hyperarid paleoclimate state
since 14 Ma, Atacama Desert, Chile. GSA Bulletin 126 (7e8), 1016e1046.
Karger, D.N., Nobis, M.P., Normand, S., Graham, C.H., Zimmermann, N.E., 2021.
CHELSA-TraCE21k v1. 0. Downscaled transient temperature and precipitation
data since the last glacial maximum. Clim. Past Discuss 1e27.
Kienast, M., Kienast, S.S., Calvert, S.E., Eglinton, T.I., Mollenhauer, G., Francois, R.,
Miz, A.C., 2006. Eastern Pacific cooling and Atlantic overturning circulation
during the last deglaciation. Nature 443, 846e849.
~ o-like pattern in
Koutavas, A., Lynch-Stieglitz, J., Marchitto Jr., Sachs, J.P., 2002. El Nin
ice age tropical Pacific sea surface temperatures. Science 297 (5579), 226e230.
Latorre, C., Betancourt, J., Rylander, K.A., Quade, J., Matthei, O., 2003. A vegetation
history from the arid prepuna of northern Chile (22-23 S) over the last 13500
years. Palaeogeogr. Palaeoclimatol. Palaeoecol. 194, 223e246.
Latorre, C., Betancourt, J., Rech, J., Quade, J., Holmgren, C., Placzek, C., Maldonado, A.,
Vuille, M., Rylander, K., 2005. Late quaternary history of the Atacama Desert. In:
Smith, M., Hesse, P. (Eds.), 23 S: Archaeology and Environmental History of the
Southern Deserts. National Museum of Australia Press, pp. 73e90.
Kubik, P., Ivy-Ochs, S., 2004. A re-evaluation of the 0e10 ka 10Be production rate for
€fels (Austria) landslide. Nuclear Inexposure dating obtained from the Ko
struments and Methods in Physics Research B 223e224, 618e622.
Latorre, C., Betancourt, J., Arroyo, M.T.K., 2006. Late Quaternary vegetation and
climate history of perennial river canyon in the Río Salado basin (22 S) of
Northern Chile. Quat. Res. 65, 450e466.
Liu, Z., Otto-Bliesner, He, F., Brady, E.C., Tomas, R., Clark, P.U., Carlson, A.E., LynchStieglitz, J., Curry, W., Brook, E., Erickson, D., Jacob, R., Kutzbach, J., Cheng, J.,
2009. Transient simulation of last deglaciation with a new mechanism for
bolling-allerod warming. Science 325, 310.
pez, R., Galli., 2015. Hydrological change during the Pleistocene-Holocene tranLo
sition associated with the Last Glacial Maximum-Altithermal in the eastern
border of northern Puna. Andean Geology, 42(1), 1e19.
Makou, M.C., Eglinton, T.I., Oppo, D.W., Hughen, K.A., 2010. Postglacial changes in el
~ o and La Nin
~ a bahavior. Geology 38, 43e46.
Nin
de Historia Natural 75, 527e546.
Blaco, G., Stone, J., Lifton, N., Dunai, T., 2008. A complete and easily accessible means
of calculationg surface exposure ages or erosion rates from 10Be and 26Al
measurements. Quat. Geochronol. 3 (3), 174e195.
, J., Farley, K.A., Fornari, M., Jimenez, N., Ramirez, V., 2009. Late local
Blard, P.H., Lave
glacial maximum in the central Altiplano triggered by cold and locally-wet
conditions during the paleolake Tauca episode (17-15 ka, Heinrich 1). Quat.
Sci. Rev. 28, 1e14.
Blodgett, T., Lenters, J., Isacks, B., 1997. Constraints on the origin of paleolake expansions in the Central Andes. Earth Interact. 1, 1e29.
Bobst, a., Lowenstein, T., Jordan, T., Godfrey, L., Lung Ju, T., Luo, S., 2001. A 106 ka
paleoclimate paleorecord from drill core of the Salar de Atacama, northern
Chile. Palaeogeogr. Palaeoclimatol. Palaeoecol. 173, 21e42.
Broccoli, A., Dahl, K., Stouffer, R., 2006. Response of the ITCZ to Northern Hemisphere cooling. Geophysical Research Letters 33, L01702.
, A., Aguilar, G., Riquelme, R., 2017. Holocene evolution and geochronology of
Cabre
semiarid fluvial system in the western slope of the Central Andes: AMS 14C data
nsito River Valley, Northern Chile. Quat. Int. 438 (A), 20e32.
in El Tra
ficos, No.
Capitanelli, R., 1967. Climatología de Mendoza. Boletín de Estudios Geogra
54-57, 1-441.
, M., Azzoug, M., Bentaleb, I., Chase, B., Fontugne, M., Jackson, D., Ledru, M.-P.,
Carre
Maldonado, A., Sachs, J., Schauer, A., 2012. Mid-holocene mean climate in the
South-eastern pacific and its influence on south America. Quat. Int. 253, 55e66.
Cesta, J., Ward, D., 2016. Timing and nature of alluvial fan development along the
Chajnantor Plateau, northern Chile. Geomorphology 273, 412e427.
Chaffaut, I., Coudrian-Ribstein, A., Michelot, J.L., Pouyaud, B., 1998. Precipitation ad
altitude du Nord-Chili, oirgine des sources de vapeaur et donnes isotopique, vol.
27. Bulletion Institute Frances Etudies Andines, pp. 367e384.
Charrier, R., Iturriazaga, L., Carretier, S., Regard, V., 2019. Geomorphologic and
glacial evolution of the cachapoal and southern maipo catchments in the andean principal cordillera, Central Chile (34 -35 S). Andean Geol. 46 (2),
240e278.
Cordeiro, R.C., Turcq, B., Sifeddine, A., Lacerda, L.D., Silva, E.V., Gueiros, B., Potty, Y.P.,
dua, E.O., Patchinelam, S.R., 2011. Biogeochemical indicators of
Santelli, R.E., Pa
environmental changes from 50 ka to 19 ka in a humid region of the Brazilian
Amazon. Palaeogeogr. Palaeoclimatol. Palaeoecol. 299 (3e4), 426e436.
De Porras, M.E., Maldonado, A., Pol-Holz, De, Latorre, C., Betancourt, J.L., 2017. Late
Quaternary environmental dynamics in the Atacama Desert reconstructed from
rodent midden pollen paleorecords. J. Quat. Sci. 32 (6), 665e684.
Desilets, D., Zreda, M., 2003. Spatial and temporal distribution of secondary cosmicray nucleon intensities and applications to in situcosmogenic dating. Earth
Planet. Sci. Lett. 206, 21e42.
Díaz, F.P., Latorre, C., Carrasco-Puga, G., Wood, J.R., Wilmshurst, J.M., Soto, D.C.,
rrez, R.A., 2019. Multiscale climate change impacts on plant
Cole, T.L., Gutie
diversity in the Atacama Desert. Global Change Biol. 25, 1733e1745.
Díaz, F.P., Latorre, C., Maldonado, A., Quade, J., Betancourt, J., 2012. Rodent middens
reveal episodic, long-distance plant colonizations across the hyperarid Atacama
Desert over the last 34000 years. J. Biogeogr. 39, 510e525.
Donders, T.H., Wagner, F., Visscher, H., 2008. Integration of proxy data and model
scenarios for the mid-Holocene onset of modern ENSO activity. Quat. Sci. Rev.
27, 571e579.
Falvey, M., Garreaud, R., 2009. Regional cooling in warming world: recent temperature trends in the southeast Pacific and along the west coast of subtropical
South America (1979-2006). J. Geophys. Res. 114, 1e16.
Fick, S.E., Hijmans, R.J., 2017. WorldClim 2: new 1-km spatial resolution climate
surface for global land areas. Int. J. Climatol. 00, 1e14.
Flores-Aqueveque, V., Rojas, M., Aguirre, C., Arias, P., Gonzalez, C., 2020. South Pacific Subtropical High from the Late Holocene to the end of the 21st century:
insights from climate proxies and general circulation models. Clim. Past 16,
79e99.
raud, G., 2001. Dating paleolakes in the central altiplano of
Fornari, M., Risaher, F., Fe
boliviano. Palaeogeogr. Palaeoclimatol. Palaeoecol. 172, 269e282.
Fritz, S.C., Baker, P., Lowenstein, T.K., Seltzer, G.O., Rigsby, C.A., Dwyer, G.S.,
Tapia, P.M., Arnold, K.K., Ku, T.-L., Luo, S., 2004. Hydrologic variation during the
last 170,000 years in the southern hemisphere tropics of South America. Quat.
Res. 61, 95e104.
Fuenzalida, H., Rutllant, J., 1987. Origen del vapor de agua que precipita sobre el
Altiplano de Chile, paper presented at II Congreso InterAmericano de Meteorología. Am. Meteorol. Soc., Buenos Aires.
n, M., Luthgens, C., 2019. Multi-millennial scale
García, J.L, Andrade, B., Caldero
climate variability during MIS 3 and MIS 2 inferred from luminescence dating of
coastal sand dunes and buried paleosol sequences in central Chile, 32 S. Journal
of Quaternary Science 34 (3), 203e214.
Garreaud, R., 1999. Multi-scale analysis of the summertime precipitation over the
central Andes. Mon. Weather Rev. 127, 901e921.
Garreaud, R., Vuille, M., Clement, A., 2003. The climate of the Altiplano: observed
current conditions and mechanisms of past changes. Palaeogeogr. Palaeoclimatol. Palaeoecol. 3054, 1e18.
Garreaud, R., Molina, A., Farias, M., 2010. Andean uplift, ocean cooling and Atacama
hyperaridity: a climate modeling perspective. Earth Planet Sci. Lett. 292, 39e50.
Gayo, E.M., Latorre, C., Jordan, T., Nester, P., Estay, S., Ojeda, K., Santoro, C., 2012. Late
Quaternary hydrological and ecological changes in the hyperarid core of the
northern Atacama Desert (~21 S). Earth Sci. Rev. 113, 120e140.
~ ez, L., Schotterer, U., 1999. Radiocarbon reservoir effect
Geyh, M., Grosjean, M., Nun
and timing of the late-glacial/early Holocene humid phase in the Atacama
22
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
n, C., 2002. Paleoenvironmental changes in the semiarid
Maldonado, A., Villagra
coast of Chile (~32 S) during the last 6200 cal, years inferred from a swampforest pollen record. Quat. Res. 58, 130e138.
n, C., 2006. Climate variability over the last 9900 cal yr BP
Maldonado, A., Villagra
from swamp forest pollen record along the semiarid coast of Chile. Quat. Res.
66, 246e258.
n, C., 2005. Pollen analyses from a
Maldonado, A., Betancourt, J., Latorre, C., Villagra
50000 -yr rodent midden series in the southern Atacama Desert (25 30'S).
J. Quat. Sci. 20 (5), 493e507.
Marengo, J., Rodriguez, W., Celeste, A., Nicolini, M., 2004. Climatology of the low
level jet east of andes as derived from the NCEP-NCAR reanalyses: characteristics and temporal variability. J. Climatol. 17 (12), 2261e2280.
Markgraf, V., 1983. Late and postglacial vegetational and paleoclimatic changes in
subantarctic, temperate, and arid environments in Argetina. Palynology 7,
43e70.
, J., Condom, T., Pre
maillon, M., Jomelli, V., Brunstein, D.,
Martin, L., Blard, P.-H., Lave
Lupker, M., Charreau, J., Maiotti, V., Tibarí, B., , ASTER Team., Davy, E., 2018. Lake
Tauca highstand (Heinrich stadial 1) driven by a southward shift of the Bolivian
high. Sci. Adv. 4 (8), eaar2514.
Masiokas, M., Villalba, R., Luckman, B., Le Quesne, C., Aravena, J.C., 2012. Snowpack
variation in the Central Andes of Argentina and Chile, 1951-2005: large-scale
atmospheric influences and implications for water resources in the region.
J. Clim. 19, 6334e6352.
Minchin, J.B., 1882. Notes of journey though part of the Andes table-land of Bolivia
in 1882. Proc. Royal Geograph. Soc. Month Record. Geograph. New Month Ser. 4,
671e676.
Mix, A., Bard, E., Schneider, R., 2001. Environmental processes of the ice age: land,
oceans, glaciers (EPILOG). Quat. Sci. Rev. 20, 627e657.
Morales, M., 2011. Arqueología Ambiental del Holoceno temprano y medio en la
Puna Seca Argentina. Modelos paleoambientales multi-escalas y sus implicancias para la arqueología de cazadores-recolectores. In: BAR South American
Archaeology Series 15. Archaeopres, Oxford.
Moreno, A., Giralt, S., Valero-Garces, B., Saez, A., Bao, R., Prego, R., Pueyo, J.J., Gonriz, P., Taberner, C., 2007. A 14 kyr record of the tropical Andes: the
zalez-Sampe
sequence (18 S, northern Chilean Altiplano. Quat. Int. 161, 4e21.
Lago Chungara
Moreno, A., Santoro, C., Latorre, C., 2009. Climate change and human occupation ini
the northernmost Chilean Altiplano over the last ca. 11500 cal. a BP. J. Quat. Sci.
24 (4), 373e382.
Mujica, M.I., Latorre, C., Maldonado, A., Gonzalez-Silvestre, L., Pinto, R., De PolHolz, R., Santoro, C., 2015. Late Quaternary climate change, relict populations
and present-day refugia in the northern Atacama Desert: a case study fromo
Quebrada La Higuera (18 S). J. Biogeogr. 42, 76e88.
Nash, D.J., Bateman, M.D., Bullard, J.E., Latorre, C., 2018. Late Quaternary coastal
evolution and aeolian sedimentation in the tectonically-active southern Atacama Desert, Chile. Palaeogeogr. Palaeoclimatol. Palaeoecol. 490, 546e562.
Navarro, D., Mehl, A., Zarate, m., Paez, M.M., 2010. Mid-late Holocene environments
of Agua Buena locality (34 50'S: 69 56'W), mendoza, Argentina. Earth Environ.
Sci. 9, 1e8.
Nester, P., Gayo, E., Latorre, C., Jordan, T., Blanco, N., 2007. Perennial stream
discharge in the hyperarid Atacama Desert of northern Chile during the latest.
Pleistocene 104 (50), 19724e19729.
Norton, K., Schlunegger, F., 2011. Migrating deformation in the Central Andes from
enhanced orographic rainfall. Nat. Commun. 2. Article number: 584.
Novello, V., Vuille, M., Cruz, F., Strikis, N., De Paula, M., Edwards, R., Cheng, H.,
Karmann, I., Jaqueto, P., Trindade, R., 2016. Centennial-scale solar forcing of the
South American Monsoon System recorded in stalagmites. Sci. Rep. 6, 24762.
Ortega, C., Vargas, G., Rutllant, J., Jackson, D., Mendez, C., 2012. Major hydrological
regime change along the semiarid western coast of South America during the
early Holocene. Quat. Res. 78, 513e527.
Osorio, D., Jackson, D., Ugalde, P., Latorre, C., De Pol-Holz, R., Santoro, C., 2011.
Hakenasa cave and its revelance for the peopling of Andean altiplano. Antioquity 85, 1194e1208.
Paduano, G., Bush, M., Baker, P., Fritz, S., Seltzer, G., 2003. A vegetation and fire
history of Lake Titicaca since the last glacial maximum. Palaeogeogr. Palaeoclimatol. Palaeoecol. 194, 259e279.
-Reygonosa, J., Andre
s, N., Angel, I.,
Palacios, D., Stoke, C., Phillips, F., Clague, J., Alcala
Blard, P-Henri., Briner, J., Hall, B., Dahms, D., Hein, A., Jomelli, V., Mark, B.,
z-Selem, L.,
Martini, M., Moreno, P., Riedel, j., Sagredo, E., Stansell, N., Vsque
Ward, D., 2020. The deglaciation of the Americas during the Last Glacial
Termination. Earth-science Reviews 203, 1e58.
Palmer, M.R., Pearson, P.N., 2003. A 23,000-year record of surface water pH and
PCO2 in the western Ecuatorial Pacific Ocean. Science 300, 480e482.
Pfeiffer, M., Latorre, C., Santoro, C., Gayo, E.M., Rojas, R., Carrevedo, M.L.,
McRostie, V., Finstad, K., Heimsath, A., Jungers, M., De Pol-Holz, R.,
Amundson, R., 2018. Chronology, stratigraphy and hydrological modelling of
extensive wetlands and paleolakes in the hyperarid core of the Atacama Desert
during the late quaternary. Quat. Sci. Rev. 197, 224e245.
Pigati, J., Rech, J., Quade, J., Bright, J., 2014. Desert wetland in the geologic record.
Earth Sci. Rev. 132, 67e81.
Placzek, C., Quade, J., Patchett, P.J., 2006. Geochronology and stratigraphy of late
Pleistocene lake cycles on the southern Bolivian Altiplano: implications for
causes of tropical climate change. GSA Bulletin 118 (5e6), 515e532.
Placzek, C., Quade, J., Betancourt, J.L., Patchett, J., Rech, J., Latorre, C., Matmon, A.,
Holmgren, C., English, N., 2009. Climate in the dry central Andes over geologic,
millennial, and internannual timescales. Ann. Mo. Bot. Gard. 96, 386e397.
Placzek, C., Quade, J., Patchett, P.J., 2011. Isotopic tracers of paleohydrological change
in large lakes of the Bolivian Altiplano. Quat. Res. 75, 231e244.
Placzek, C., Quade, J., Patchett, P.J., 2013. A 130 ka reconstructing of rainfall on the
Bolivian Altiplano. Earth Planet Sci. Lett. 363, 97e108.
Quade, J., Rech, J., Betancourt, J., Latorre, C., Quade, B., Rylander, K., Fisher, T., 2008.
Paleowetlands and regional climate change in the central Atacama Desert,
northern Chile. Quat. Res. 69, 343e360.
Rech, J., Quade, J., Betancourt, J., 2002. Late quaternary paleohydrology of the
Atacama Desert (lat 22 -24 S), Chile. GSA Bulletin 114 (3), 334e348.
Rech, J., Pigati, J., Quade, J., Betancourt, J., 2003. Re-evaluation of mid Holocene
deposits at Quebrada Puripica, northern Chile. Palaeogeogr. Palaeoclimatol.
Palaeoecol. 194, 207e222.
Rech, J., Currie, B.S., Schullenberger, E.D., Dunagan, S.P., Jorda, T., Blanco, N.,
Tomlinson, A.J., Rowe, H.D., Houston, J., 2010. Evidence for the development of
the Andean rain shadow from a Neogene isotopic record in the Atacama Desert,
Chile. Earth Planet Sci. Lett. 292 (3e4), 371e382.
Rech, J., Tully, C., Latorre, C., 2016. Stream Response to Climate Change, Atacama
Desert,
Chile.
https://serc.carleton.edu/vignettes/collection/36630.html.
(Accessed 2 August 2020).
Rech, J., Currie, B.s., Jordan, T., Riquelme, R., Lehmann, S.B., Kirk-Lawlor, N., Li, S.,
Gooley, J., 2019. Massive middle miocene gypsic paleosols in the atacma Desert
and the formation of the central andean rain-shadow. Earth Planet Sci. Lett. 56,
184e194.
Rehfeld, K., Brown, J., 2021. Interannual-to-interdecadal variability in PMIP simulations at the local to global scale. PAGES Magazine 29 (2), 98e99.
Reimer, P., Bard, E., Bayliss, A., Beck, J., Blackwell, P., Ramsey, C., Buck., C., Cheng, Hai.,
Edwards, R., Friedrich, M., Grootes, P., Guilderson, T., Haflidason, H., Hajdas, I.,
Christine, H., Heaton, T., Hogg, A., Hughen, K., Kaiser, K., Staff, R., 2013. IntCal13
and MARINE13 radiocarbon age calibration curves 0-50,000 years cal BP.
Radiocarbon 55, 1869e1887.
Rigsby, C., Bradbury, J.P., Baker, P., Rollins, S., Warren, M., 2005. Late Quaternary
palaeolakes, rivers, and wetlands on the Bolivian Altiplano and their palaeoclimatic implications. J. Quat. Sci. 20 (7e8), 671e691.
Riquelme, R., Rojas, C., Aguilar, G., Flores, P., 2011. Late Pleistocene-early Holocene
paraglacial and fluvial sediment history in the Turbio valley, semiarid Chilean
Andes. Quat. Res. 75, 166e175.
Rohrmann, A., Sachse, D., Mulch, A., Pingel, H., Tofelde, S., Alonso, R., Strecker, M.,
2016. Miocene orographic uplift forces rapid hydrological change in the
southern central Andes. Nat. Sci. Rep. 6, 35678.
Rojo, L.D., Mehl, A., Paez, M., Zarate, M., 2012. Mid-to late Holocene pollen and
alluvial record of the arid Andean piedmont between 33 and 34 S, Mendoza,
Argentina: inferences about floodplain evolution. J. Arid Environ. 77, 110e122.
Rowe, H.D., Dunbar, R.B., Mucciarone, D.A., Seltzer, G.O., Baker, P., Fritz, S., 2002.
Insolation,moisture balance and climate change on the South American altiplano since the Lastg Glacial Maximun. Climatic Change 52, 175e199.
ez, A., Godfrey, L., Herrera, C., Chong, G., Pueyo, J., 2016. Timing of wet episodes in
Sa
Atacama Desert over the last 15 ka. The groundwater discharge deposits (GWD)
from Domeyko Range at 25 S. Quat. Sci. Rev. 154, 82e93.
Sandweiss, D., Maasch, K., Burger, R., Richardson, J., Rollins, H., Clement, A., 2001.
~ o frequencies: climate records and cultural conVariation in Holocene El Nin
sequences in ancient Peru. Geology 29 (7), 603e606.
Sandweiss, D., Andrus, C.F.T., Kelley, A., Maasch, K., Reitz, E., Roscoe, P., 2020.
Archaeological climate proxies and the complexities of reconstructing Holocene
~ o in coastal Peru. Proc. Natl. Acad. Sci. U. S. A. 117 (15), 8271e8279.
El Nin
Servant, M., Fontes, J.C., 1978. Les lac quaternaries des hauts plateaux des Andes
boliviennes: premieres interpretations paleoclimatique. Evolution recentes des
hauts plateau en Bolivie. Cahiers ORSTROM . Serie Geologie 10 (1), 9e23.
Smith, C., Lowell, T., Caffee, M., 2009. Lateglacial and Holocene cosmogenic surface
exposure age glacial chronology and geomorphological evidence for the presence of cold-based glaciers at Nevado Sajama, Bolivia. J. Quat. Sci. 24 (4),
360e372.
Steinmann, D., Hok, H., Bistram, A., 1904. Zur Geologie des suedestland Bolivien.
Zentralblatt fur Mineralogie 5, 1e4.
Stríkis, N., Cruz, F., Barreto, E., Naughton, F., Vuille, M., Cheng, H., Voelker, A.,
Zhang, H., Karmann, I., Edwards, R., Auler, A., Santos, R., Sales, H., 2018. South
American monsoon response to iceberg discharge in the North Atlantic. Proc.
Natl. Acad. Sci. U.S.A. 115, 3788e3793.
Sylvestre, F., Servant, M., Servant-Vildary, S., Causse, C., Fournier, M., Ybert, J.P., 1999.
Lake-Level chronology on the southern Bolivian altiplano (18 -23 S) during
late-glacial time and the early Holocene. Quat. Res. 51, 54e66.
Tapia, J., Murray, J., Omachea, M., Tirado, N., Nordstrom, D.K., 2019. Origin, distribution, and geochemistry of arsenic in the Altiplano-Puna plateau of Argentina,
Bolivia, Chile, and Perú. Science of the Total Environment 678, 309e325.
Tchilinguirian, P., Morales, M.R., Oxman, B., Lupo, L.C., Olivera, D.E., Yacobaccio, H.D.,
2014. Early to middle Holocene transition in the Pastos Chicos record dry puna
of Argentina. Quat. Int. 330, 171e182.
Thompson, L.G., Davis, M.E., Mosley-Thompson, E., Sowers, T.A., Henderson, K.A.,
Zagorodnov, V.S., Lin, P.N., Mikhalenko, V.N., Campen, R.K., Bolzan, J.F., ColeDai, J., Francou, B., 1998. A 25000-year tropical climate history from Bolivian Ice
Cores. Science 282, 1858e1864.
Tiner, R., Negrini, R., Antinao, J., Mcdonald, E., Maldonado, A., 2018. Geophysical and
geochemical constratints on the age and paleoclimate implications of Holocene
lacustrine cores from the Andes of Central Chile. J. Quat. Sci. 33 (2), 150e165.
Trauth, M., Strecker, M.R., 1999. Formation of landslide dammed lakes during a wet
period between 40,000-25,000 yr B.P in northwestern Argentina.
23
H. Orellana, C. Latorre, J.-L. García et al.
Quaternary Science Reviews 313 (2023) 108174
ciclo glacial-postglacial en Sudamerica, vol. 22. Bulletin del Institut Francais d’
etudes Andines, pp. 243e258.
n, C., 1997. Historia de la vegetacio
n de bosques panVilla-Martinez, R., Villagra
tanosos de la costa de Chile central durante el Holoceno medio y tardío. Rev.
Chil. Hist. Nat. 70, 391e401.
Vizy, E.K., Cook, K.H., 2007. Relation between Amazon and high andes precipitation.
J. Geophys. Res. Atmos. 112, 1e14. D07107.
Vuille, M., Ammann, C.M., 1997. Regional snowfall patterns in the high, arid Andes.
Climatic Change 36, 413e423.
Vuille, M., Baumgartner, M., 1993. Hydrologic investigations in the north Chilean
Altiplano using Landsat-MSS and TM-data. Geocarto Int. 8, 35e45.
Vuille, M., Keimig, F., 2004. Interannual variability of summertime convective
cloudiness and precipitation in the central andes derived from ISCCP-B3 Data.
J. Clim. 7, 3334e3348.
Ward, D., Cesta, J.M., Galewsky, J., Sagredo, E., 2015. Late Pleistocene glaciatios of the
arid subtropical Andes and new results from the Chajnantor Plateau, northern
Chile. Quat. Sci. Rev. 128, 98e116.
, E., Calogero, Santoro., Ugalde, P., De Pol-Holz, R.,
Workman, T., Jason, Rech., Gayo
Capriles, J., Latorre, C., 2020. Landscape evolution and the environmental
context of human occupation of the southern pampa del tamarugal, Atacama
Desert, Chile. Quaternary Science Reviews 243, 106502.
Yacobaccio, H., Morales, M., 2005. Mid-Holocene environment and human occupation of the Puna (Susque, Argentina). Quat. Int. 132, 5e14.
Zech, R., Kull, C., Veit, H., 2006. Late Quaternary history in the Encierro Valley,
northern Chile (29 S), deduced from 10Be surface exposure dating. Palaeogeogr.
Palaeoclimatol. Palaeoecol. 234, 277e286.
Zech, R., Kull, C., Kubik, P.W., Veit, H., 2007. Exposure dating of late glacial and pre~ a Rosa, northern/cCentral Chile ~31 S).
LGM moraines in the Cordon de Don
Clim. Past 3, 1e14.
Zech, R., May, J., Kull, C., Ilgner, J., Kubik, P., Veit, H., 2008. Timing of the late quaternary glaciation in the andes from 15 to 40 S. J. Quat. Sci. 23 (6e7), 635e647.
Zhou, J., Lau, K.M., 1998. Does a monsoon climatic exist over South America.
J. Climatol. 11, 1020e1040.
Palaeogeograpy, Paleoclimatology, Paleoecology 153, 277e287.
Trauth, M., Alonso, R., Haselton, K., Hermanns, R., Strecker, M., 2000. Climate change
and mass movements in the NW Argentine Andes. Earth Planet Sci. Lett. 179,
243e256.
Trauth, M., Bookhagen, B., Marwan, N., Strecker, M., 2003. Multiple landslide cluster
record Quaternary climate changes in the northwestern Argentine Andes.
Paleogeography, Paleoclimatology, Paleoecology 194, 109e121.
licas y fluviales del
Tripaldi, A., Zarate, M., Brook, G. (Eds.), 2010. Sucesiones eo
Pleistoceno tardío-Holoceno de la planicie Sanrafaelina: paleoambientes y
paleoclima. In: Reference to chapter in an edited book. Sociedad Argentina de
Antropología, pp. 95e121.
Tully, C., Rech, J., Workman, T., Santoro, C., Capriles, J.M., Gayo, E., Latorre, C., 2019.
In-stream wetland deposits, megadrought, and cultural change in the northern
Atacama Desert, Chile. Quat. Res. 91, 63e80.
s, B., Jenny, B., Rondanelli, M., Delgado-Huertas, A., Burns, S., Veit, H.,
Valero-Garce
Moreno, A., 2005. Palaeohydrology of Laguna de Tagua Tagua (34 30'S) and
moisture fluctuations in central Chile for the last 46000 yr. J. Quat. Sci. 20 (7e8),
625e641.
Van der Hammern, T., Hooghiemstra, H., 2000. Neogene and Quaternary history of
vegetation, climate, and plant diversity in Amazonia. Quat. Sci. Rev. 19 (8),
725e742.
Vargas, G., Rutllant, J., Ortlieb, L., 2006. ENSO tropical-extratropical climate teleconnections and mechanisms for Holocene debris flows along the hyperarid
coast of western South America (17 -24 S). Earth Planet Sci. Lett. 249, 467e483.
Veit, H., 1996. Southern westerlies during the Holocene deduced from geomorphological and pedological studies in the norte chico, northern Chile (27-33 S).
Palaeogeogr. Palaeoclimatol. Palaeoecol. 123, 107e119.
Veit, H., Jan-Hendrik, M., Madella, A., Delunel, R., Schluneger, F., Szidat, S.,
Capriles, J., 2016. Palaeo-geoecological significance of Pleistocene trees in the
Lluta valley, Atacama Desert. J. Quat. Sci. 31 (3), 203e213.
Vera, C., Baez, J., Douglas, M., Emamanuel, C.B., 2006. The South American low-level
jet experiment. Bull. Am. Meteorol. Soc. 87 (1), 63e78.
n, C., 1993. Una interpretacio
n clima
tica del registro palinolo
gico del ultimo
Villagra
24
Descargar