Sie sind auf Seite 1von 23

Quaternary Science Reviews 53 (2012) 1e23

Contents lists available at SciVerse ScienceDirect

Quaternary Science Reviews


journal homepage: www.elsevier.com/locate/quascirev

Invited review

A review of Glacial and Holocene paleoclimate records from southernmost


Patagonia (49e55 S)
Rolf Kilian a, *, Frank Lamy b
a
b

Geologie, FBVI, Universitt Trier, Behringstr. 16, 54826 Trier, Germany


Alfred Wegener Institut fr Polar- und Meeresforschung, Am Alten Hafen 26, 27568 Bremerhaven, Germany

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 10 January 2012
Received in revised form
11 July 2012
Accepted 18 July 2012
Available online

Southern South America is the only landmass intersecting the southern westerly wind belt (SWW) that
inuences the large-scale oceanography and controls for example the outgassing of CO2 in the Southern
Ocean. Therefore, paleo-reconstructions from southernmost Patagonia are of global interest and an
increasing number of paleoclimate records have been published during the last decades. We provide an
overview on the different records mostly covering the Holocene but partly extending into the Late Glacial
based on a large variety of archives and proxies. We particularly discuss possible reasons for regionally
diverging palaeoclimatic interpretations and summarize potential climate forcing mechanisms. The
Deglacial and Holocene temperature evolution of the region including the adjacent Pacic Ocean indicates Antarctic pattern and timing consistent with glacier re-advances during the Antarctic Cold
Reversal. Some records indicate a signicant accumulation control on the glacier uctuations related to
changes in SWW strength and/or position. Reconstructions of Holocene changes in the SWW behaviour
provide partly inconsistent and controversially discussed pattern. While records from the hyperhumid
side point to a stronger or southward displaced SWW core during the Early Holocene thermal maximum,
records from the lee-side of the Andes show either no long term trend or the opposite, suggesting
enhanced westerlies during the late Holocene Neoglacial. Likewise, centennial-scale global or hemispheric cold intervals, such as the Little Ice Age, have been interpreted in terms of enhanced and reduced
SWW strength. Some SWW variations can be linked to changes in the El Nio-Southern Oscillation
(ENSO) consistent with instrumental climate data-sets and might be ultimately forced by solar variability. Resolving these inconsistencies in southernmost Patagonian SWW records is a prerequisite for
improving hemispheric comparisons and links to atmospheric CO2 changes.
2012 Elsevier Ltd. All rights reserved.

Keywords:
South America
Paleoclimate
Palynology
Glaciology
Speleothem
Southern hemispheric westerlies
SST
Holocene
Glacial
Quaternary

1. Introduction
The southernmost tip of South America is of particular interest
for paleoclimate reconstruction, since it is the only land-mass
intersecting the core of the southern westerly wind belt (SWW)
at latitudes between 49 and 53 S (Fig. 1). On a hemispheric scale,
SWW changes substantially contribute to the forcing of the deep
and vigorous Antarctic Circumpolar Current (ACC). Wind-induced
upwelling within the ACC in the Southern Ocean raises large
amounts of deep water to the oceans surface in this circumpolar
belt affecting the global thermohaline circulation (e.g. Marshall and
Speer, 2012) and atmospheric CO2 contents (e.g. Toggweiler et al.,
2006). Therefore, the SWW exerts a strong control on global
climate and oceanography.
* Corresponding author.
E-mail address: kilian@uni-trier.de (R. Kilian).
0277-3791/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.quascirev.2012.07.017

The rst paleoclimate records from Patagonia were based on


glacier advance reconstructions (Caldenius, 1932; Mercer, 1965)
and on pollen records from soil and peat records (Auer, 1933, 1958,
1960, 1974; Heusser, 1971). Multi-proxy reconstructions based on
lake and fjord sediment cores as well as stalagmites initiated only
during the past 10 years and constitute now a growing number of
partially high resolution records. The number of paleoclimaterelated publications particularly increased since ca 2005 (to more
than 25 publications/year; Fig. 2). Despite this large number of
publications and the global implications of the Patagonian paleoclimate, reconstructions of Glacial and Holocene SSW changes are
partly inconsistent and discussed controversially. Contrasting
inferences have been derived from proxy records in southern
Patagonia for example on multi-millennial time-scales during the
Holocene (e.g. Mayr et al., 2007a,b; Lamy et al., 2010; Moreno et al.,
2010; Waldmann et al., 2010; Fletcher and Moreno, 2011) but also
regarding shorter term climate variations on centennial time-

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Fig. 1. A) Southernmost South America with average annual precipitation New et al., 2002) and the modern annual mean sea surface temperature distribution of the surrounding
oceans (data from the NOAA-CIRES Climate Diagnostics Center http://www.cdc.noaa.gov/index.html). The Antarctic Circumpolar Current (ACC) and Cape Horn Current (CH) are
indicated. B) December to February zonal wind distribution over the Southern Hemisphere based on NCEP/NCAR reanalysis data (Kalnay et al., 1996) and C) correlation between
850 hPa zonal wind and precipitation (Garreaud et al., in press).

scales, e.g. during the globally known Little Ice Age (LIA; e.g. Moy
et al., 2008; Schimpf et al., 2011).
Besides SWW reconstructions which are primarily based on
precipitation proxies and their correlation to wind changes, paleotemperature reconstructions in southernmost Patagonia may help
to understand atmospheric and ocean temperature behaviour in
the Southern Hemisphere. These are for example essential for
evaluating coupled atmospheric-ocean circulation changes as well
as timing and dynamics of interhemispheric climate changes
including the bipolar seesaw (e.g. Stocker and Johnsen, 2003; Lamy
et al., 2007; Barker et al., 2009). Furthermore, the southern Patagonian Ice eld (PIF) is of special interest, because it constitutes the

Fig. 2. Number of published paleoclimate studies in southernmost Patagonia over the


past decades.

largest continental ice-sheet outside the polar regions (Casassa


et al., 2000) which is highly sensitive to temperature as well as
precipitation changes (Warren and Sudgen, 1993) and contributes
signicantly to global sea level changes (Rignot et al., 2003; Glasser
et al., 2011).
As with the paleoclimate data, also modelling studies on the
position and strength of the SWW that have been performed in
particular for the Last Glacial Maximum (LGM) are not yet
conclusive (Rojas et al., 2009). Holocene changes of the SWW have
also been addressed by for example mid-Holocene and preindustrial simulations (Wagner et al., 2007; Rojas and Moreno, 2011) and
modelling studies that focus on the impact of solar variability
during the past 3000 years (Varma et al., 2011).
Instrumental climate and weather data collected over the past
ca 50 years show that SSW extends more than 3000 km
northesouth with signicant latitudinal variations on seasonal to
decadal times-scales (Fig. 1A and B). Due to the few weather station
data in the Southern Hemisphere, it is likely that the SWW variability is spatially more complex than presently known (Garreaud,
2007; Garreaud et al., in press). Regarding paleo-SWW reconstructions, it is therefore important to note that past changes of the
SWW cannot be estimated from a record of a single site. Further
complicating is the west-east distribution of sites across the Patagonian Andes, where the SSW forces in one of the most pronounced
climate divide on earth (Fig. 1C). This strong precipitation gradient
produces different, highly sensitive ecosystems at the hyperhumid
western and evaporation controlled arid eastern side of the Andes.
Paleoclimate archives from such contrasting environments often
require different and not directly comparable proxies. Therefore,
local climate characteristics are very important in southern Patagonia as well as proxy calibration and proxy monitoring which has
only started within the past few years (e.g. Schimpf et al., 2011). A
further complication of paleoclimate reconstructions from Patagonia is the fact that some proxy records may have been affected by

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Glacial to Holocene coastline changes related to a regionally


distinct isostatic, eustatic, and neotectonic development of the
region. These factors result for example in a regionally different
timing of marine transgression into proglacial lakes and lowland
around the southern tip of South America (e.g. Anderson and
Archer, 1999; Kilian et al., 2007a).
This paper provides an overview on Glacial and Holocene
palaeoclimatic reconstructions based on a large variety of marine
and terrestrial archives as well as diverse proxies. We focus on
a compilation of records from both sides of the Andean climate
divide and discuss possible reasons for partly contrasting palaeoclimatic interpretations primarily regarding past changes of the
SWW. We conclude with hemispheric and global implications as
well as potential climate forcing mechanisms.
2. Characteristics of the present day climate and vegetation
The present-day latitudinal distribution of the SWW is monitored by abnormal high precipitation along the west coast of South
America and in the Andes between latitudes ca 30 S and >55 S
(Fig. 1A). During the last century, the annual variations of the SWW
strength and related precipitation have been recorded at some faroff located regional weather stations (Schneider et al., 2003;
Garreaud, 2007; Rasmussen et al., 2007; Carrasco et al., 2008;
Aravena and Luckman, 2009), but only few of these stations are
situated near to the Andean climate divide.
World-wide weather station data and numerical weather
prediction model output data since 1948 have been used to
construct an interpolated global climate data-set (NCEP/NCAR
reanalysis data: Kalnay et al., 1996; Fig. 1B and C). In summer, the
NCEP/NCAR data show the highest wind velocities across the South
American continent between 49 and 53 S. This marks the core of
the SWW (Fig. 1B) characterised also by a precipitation maximum
at the west coast and the Andes during summer (Schneider et al.,
2003). During winter, the SWW is broader and slightly displaced
northward, while the core section is relatively weaker. Interpolated
weather station data indicate a clear positive correlation between
precipitation and SSW strength at the continental margin and in
the Andes (r 0.4 to 0.8; Fig. 1C; Garreaud et al., in press). To the
east of the Andes this correlation soon becomes weaker (r 0.4 to
0.2) and negatively at the eastern steppe region towards the
Atlantic coast. In the lee of the Andes and in the Patagonian steppe
the dry winds produce high evaporation and negative water
balance, in particular during summer (Endlicher, 1991). Therefore
many lagoons dry out in summer and regional run-off is very
restricted in glacier-free eastern areas between 50 and 55  C.
The NCEP/NCAR data also show the relationship between
changes of Intertropical Convergence Zone (ITCZ), El Nino Southern
Oscillation (ENSO) and the SWW. During El Nio events, which
typically occur between December and March, the ITCZ is generally
weaker and/or southward displaced while the core section of SWW
is reduced and westerly winds are enhanced north of ca 45 S
(Schneider and Gies, 2004; Garreaud, 2007; Boucher et al., 2011).
The overall air temperature of southernmost South America is
predominately controlled by the SWW which in turn largely
reects SSTs within the Antarctic Circumpolar Current (ACC).
Occasionally, an atmospheric blocking of the SWW over Patagonia
(<20% of weather situations; Schneider et al., 2003; Garreaud et al.,
in press) enables moisture-rich easterly winds from the South
Atlantic to migrate over the Patagonian Steppe until the eastern
range of the Andes. Furthermore, northward advection of cold
Antarctic air masses can reach the southern tip of the continent and
sometimes could move further north reaching the subtropics/
topics. In general such weather situations are more likely during
winter. In the north eastern sector of Patagonia (north of 50 S) the

climate becomes more continental characterized by an increased


seasonality in temperatures of up to 20  C.
The Cape Horn Current branches from the northern ACC at
around 45 S southward and transports relatively warmer water
masses along the Pacic coast toward the Drake Passage (Fig. 1A;
Chaigneau and Pizarro, 2005). At 52 S, the open marine coastal
annual SSTs (ca 8.5  C) are ca 2  C warmer than air temperatures
measured at the coastal weather station Evangelistas (Fig. 4). This
station shows an average annual temperature of 6.5  C similar to
other weather stations across the continent at this latitude (e.g.
Punta Arenas; Schneider et al., 2003; Garreaud et al., in press),
indicating that the air temperatures are not signicantly increased
by the abnormal warm Cape Horn Current. However, the relatively
warm coastal water enters the fjord system as bottom water below
the supercial freshwater which is comparatively cold during
summer due to snow and glacier melting. This produces a strong
thermohaline layering in the fjords (Kilian et al., 2007a).
Distinct vegetational pattern occur in southernmost South
America denoting the extremely different moisture conditions over
the region. In the Andes and the western fjords, the extremely high
precipitation (>4000 mm annual precipitation) and comparatively
low evaporation causes stagnant water in most soils, in particular
during summer. These conditions limit the plant development
which is characterised by the occurrence of species-poor plant
communities (e.g. Pisano, 1977; Moore, 1979, 1983; Kleinebecker
et al., 2007). Climatic and non-climatic factors controlling the
composition and distribution of plant communities within terrestrial ecosystems are poorly investigated. Steep slopes of fjords and
regions with tectonic fracture zones, both with very good run-off
conditions as well as sediment bars and moraines with coarse
clastic sediment cover, and efcient drainage conditions enable
locally better conditions for forest growth. These observations also
indicate that the relative quantity of trees versus peat bog species
and/or hygrophytes could reect besides climatic factors (precipitation and temperature) also changes of regional coast lines which
are controlled by eustasy, isostasy and tectonics. Dendroecological
studies indicate a signicant ecesis after the glacier retreat (Koch
and Kilian, 2005), highlighting the role of local conditions (e.g.
nutrient decit) and/or ecological features (e.g. dispersion
syndromes) in the plant colonization and succession dynamics.
Furthermore, disturbance events such as earthquake-related mass
ows (Waldmann et al., 2011) or volcanic ash deposition can
produce important vegetational perturbances (Kilian et al., 2006).
To the east of the climate divide (leeward side of the Andes),
a strong reduction in soil moisture is induced by a substantial
decrease in annual precipitation (<500 mm/yr) and higher summer
temperatures (>12  C) in combination with strong winds which
cause enhanced evaporation (Fig. 3). The result is an open and
treeless landscape known as the Patagonian steppe characterised
by summer dry stress for most plants (e.g. Endlicher and Santana,
1988; Endlicher, 1991).
Between these dissimilar environments, the hyper-humid in the
west and semi-desert in the east, a Sub Antarctic deciduous forest
grows in a transitional area which is characterized by annual
precipitation of 500e1000 mm/yr (Heusser, 1995). This forest is
partly strongly affected by mechanical wind-stress (Armesto et al.,
1992), in particular when more open parkland was developed. The
moderate climate conditions enable the development of more
species-rich vegetal communities (Heusser, 1995). Charcoal records
indicate that res shaped this vegetation sequence during the
Holocene (Heusser, 1995; Huber et al., 2004; Whitlock et al., 2007;
Markgraf and Huber, 2010). However it remains highly disputed
how far the res are of natural or anthropogenic origin. A clear
anthropogenic impact was produced by the activity of farmers
during the last 130 years.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Fig. 3. WeE Transect across the Andes at 52 S with annual precipitation, evaporation,
and annual wind velocities as well as the correlation between wind velocities and
precipitation (compiled from data of Schneider et al., 2003; Aravena and Luckman,
2009; Garreaud et al., in press).

3. Regional distribution of archives and used proxies


The locations of investigated sites and possible future archives
depend on the natural distribution of glaciers, ice elds, fjords,
lakes, peat land and forest which is predominately controlled by the
morphology and the strong precipitation gradient in southernmost
Patagonia (Figs. 3 and 4). An additional more practical factor is the
accessibility of sites by roads which is only possible on the eastern
side of the Andes and in the steppe region. The hyperhumid
western area with its steep fjord and mountain belt as well as the
extended westernmost island zone is only accessible by ship.

Terrestrial explorations in this totally unpopulated area are much


more difcult and require an extended logistic support. These
geographic differences explain, why 85% of published paleoclimate
studies are from regions to the east of the Andean climate divide.
Major sites discussed in the text are listed in Table 1 and their
location is shown in Fig. 4.
Mapping of Glacial and Holocene glacier extent also primarily
concerns the eastside of the Andes: the Cordillera Darwin
(Kuylenstierna et al., 1996), Central Strait of Magellan (e.g.
Clapperton et al., 1995; McCulloch et al., 2005; Kaplan et al., 2008;
Sugden et al., 2009), the glacial lake area within the Torres del Paine
area (e.g. Moreno et al., 2009b) as well as the Lago Argentino and
Lago Viedma areas (e.g. Wenzens, 2005; Ackert et al., 2008; Kaplan
et al., 2011; Strelin et al., 2012). Only few investigations document
glacier uctuations from the central and western range of the
Andes. These include the Isla Santa Ines (Aravena, 2007) and Gran
Campo Nevado area (Koch and Kilian, 2005; Schneider et al., 2007;
Mller and Schneider, 2010). In the western area, many moraines
are subaquatic and therefore have only been locally explored by
multibeam and echo sounding (Kilian et al., 2007b; Breuer et al., in
press-a).
The Northern and Southern Patagonian Ice Fields (PIF) may
provide suitable sites for ice drilling. However, only few areas of
these ice elds are elevated enough to prevent summer melting of
the ice/snow surface related rn formation which complicates the
palaeoclimatic analyses. Frequent and extended eolian redistribution of snow further hinders palaeoclimatic exploration of icecores. Even if there are some areas appropriate areas with low ice
ow rates, the extremely high snow accumulation across the ice

Fig. 4. Regional map of southernmost Patagonia with site locations and numbering (used in the text and Table 1) and regional geography.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Table 1
Locations of paleoclimate archives and their stratigraphic range as well as regional weather stations.
Core

Location

Marine sediment records


MD07-3124
Concepcin F.
MD07-3128
Pacic coast
MD07-3132
Whiteside Channel
Vo1
Estero Vogel
Es1
Seno Skyring
Sk1
Seno Skyring
BA-1
Bahia Arevalo
LO-1
Bahia Lobo
LU-1
Quenca Luca
Sg-1
Seno Glaciar
PALM-2
Isla Parlamento
TM1
Cabo Tamar
JPC
Marinelli fjord
Lake sediment cores
ART-1
Arthuro
CH-1
Chandler
TML-1
Tamar
LMP-I
Muy Profundo
HB-KL I-III
Hambre
P8
Guanaco
Various cores
Potrok Aike
CAK 99 (various)
Cardiel
LF06 (various)
Fagnano
Peat records
P14
Puerto Eden
P13
Cerro Frias
P12
Lago Argentino
P11
Meseta La Torre 1-2
P10
Torres del Paine
P9
Nandu Vega
P7
Rio Rubens
P6
Punta Arenas
P5
Estancia Esmeralda
P4
Puerto del Hambre
P3
Passo Garibaldi
P2
Haberton
P1
Isla de los Estados
GC1
Gran Campo Nevado
GC2
Gran Campo Nevado
Sky1
Skyring
Pbr2
Puerto del Hambre

Lat. South

Long. West

Core length [m]

Water depth [m]

Stratigr range [ka]

References

50 310
52 400
53 440
52 490
52 340
52 370
52 420
52 450
52 460
52 470
52 470
52 540
54 250

74 580
75 34
70 190
72 330
72 160
71 420
73 240
73 160
73 190
73 230
73 390
73 470
69 300

22.2
33.0
21.7
4.6
4.1
4.8
9.9
8.2
1.6
0.6
8.9
4.8
13.5

564
1030
301
25
78
62
5
83
320
520
44
31
30e200

0e12.0
0e60.0
0e15.0
0e13.3
0e>5.0
0e18.0
0e>6.0
0e0.8
0e1.2
0e0.6
0e16.0
0e15.0
0e16.0

Kissel et al. (2007)


Caniupn et al. (2011)
Kissel et al. (2007)
Kilian et al. (2007b)
Baeza (2005)
Kilian et al. (2007b)
Breuer et al. (in press-b)
Breuer et al. (in press-b)
Breuer et al. (in press-b)
Breuer et al. (in press-b)
Lamy et al. (2010)
Lamy et al. (2010)
Boyd et al. (2008)

53 300
52 490
52 540
52 440
53 360
50 520
51 580
48 550
54 350

72 560
72 540
73 480
73 130
70 57
72 520
70 220
71 150
68 000

0.42
6.5
7.6
2.0
12.2
6.0
<122
15
8.0

42
16
20
205
17
16
76
76
200

0e2.1
0e12.5
0e15.0
0e4.2
0e17.3
0e16.0
0e52.0
0e15.0
0e12.0

Breuer et al. (in press-a)


Kilian et al. (2006)
Lamy et al. (2010)
Breuer et al. (in press-a)
Hermanns and Biester (2011)
Moreno et al. (2010)
Recasens et al. (2011)
Gilli et al. (2005)
Waldmann et al. (2011)

49 080
50 260
50 350
50 310
50 590
50 560
52 040
53 090
53 300
53 360
54 53
54 540
54 500
52 470
52 480
52 310
53 380

74 250
72 430
72 550
72 030
72 400
72 460
71 310
70 570
70 350
70 550
66 100
67 10
64 400
72 570
72 560
72 080
70 580

2.5

0e10.0
0e9.0
0e13.5
0e8.5
0e13.0
0e12.5
0e17.0
0e16.8
0e17.0
0e17.3
0e13.0
0e17.0
0e15.0
0e2.5
0e14.3
0e6.5
0e15.0

Ashworth et al. (1991)


Mancini et al. (2005)
Wille and Schbitz (2009)
Schbitz (1991)
Heusser (1995)
Villa-Martnez and Moreno (2007)
Huber et al. (2004)
Heusser (1995)
McCulloch and Davis (2001)
McCulloch and Davis (2001)
Huber et al. (2004)
Markgraf (1993)
Ponce et al. (2011)
Biester et al. (2002)
Lamy et al. (2010)
Biester et al. (2003)
Biester et al. (2003)

3.0
8.6
7.0
7.5
3.0
8.5
2.8
10.2
3.5
1.7
2.7
2.8
6.5

Cave and Stalagmites

Location

Lat. South

Long. West

Core length [m]

Altitude [m a.s.l.]

Stratigr range [ka]

References

MA Cave (MA1)
GC Cave (GC 1)

Bahia Arevalo
Seno Glaciar

52 417
52 481

73 230 3
73 163

0.3
0.3

20
35

0e4.5
2.4e8.0

Schimpf et al. (2011)

Automatic Weather Station Location

Lat. South Long. West Annual range of precip [m/yr] Altitude [m a.s.l.] Remarks

Evangelistas
Felix
Skyring
Passo
Bahamondes
Arevalo
Punta Arenas

52 240
52 570
52.570
52 450
52 480
52 410
53 080

Pacic coast
Str. of Magellan
Seno Skyring
Northern GCN
Ba. Bahamondes
Bahia Arevalo
Punta Arenas

75 060
74 040
71 510
73 010
72 560
73 160
70 530

1.5e5.5
3.2e7.0
0.6e1.1
7.2e9.8
3.3e7.1
4.8
0.2e0.8

elds (reaching >10,000 mm water equivalent per year (Rivera,


2004), strongly limits the possible length of ice records. So-far
only at the 4100 m high Mt. San Valentin within the Northern PIF
up to 122 m long ice cores have been drilled successfully. This site is
unaffected by summer melting and the record includes at least the
last 150 years up to 85 m core depth. Still undated older ice of the
lowermost core represents an earlier Holocene snow accumulation
(Vimeux et al., 2008, 2009).
In SE Patagonia, lake sediment records have been obtained from
numerous sites including Lago Fagnano (e.g. Waldmann et al., 2010;
Moy et al., 2011), Lago Hambre (e.g. Hermanns and Biester, 2011;
Breuer et al., in press-a), Lago Guanaco (Moreno et al., 2010), Laguna
Potrok Aike (e.g. Mayr et al., 2007a; Haberzettl et al., 2009) and
Lago Cardiel (e.g. Stine and Stine, 1990; Gilli et al., 2005; Fig. 4). As
close to the Andes all lakes haven been covered by glaciers,

30
15
8
380
26
90
30

Since
Since
Since
Since
Since
Since
Since

AD
AD
AD
AD
AD
AD
AD

References
1990
1910
2001
2000
1999
2007
1888

Aravena and Luckman (2009)


Schneider et al. (2003)
Schneider et al. (2003)
Schneider et al. (2003)
Aravena and Luckman (2009)

paleoclimate records only start after the ice retreat between 17 and
14 ka BP (e.g. Lamy et al., 2010). Exceptions are records from the
non-glaciated Laguna Potrok Aike that reach back to ca 50 ka BP
(e.g. Recasens et al., 2011; Lis-Pronovost et al., 2012). On the
western side of the Andes at around 52 300 S Late Glacial to Holocene records have been published from Lake Tamar, Lake Arthuro,
Lake Muy Muy and Lake Chandler (Fig. 4; e.g. Kilian et al., 2007b;
Lamy et al., 2010; Breuer et al., in press-b). Lake sediment studies
are mostly based on multiple palaeoclimatic proxies, typically
including pollen, and geochemistry and isotopy of organic material
as well as granulometry and geochemistry of siliciclastic sediment
components. XRF core scanner records of some lake sediment cores
reach up to yearly-resolution (e.g. Lake Fagnano and Tamar).
Late Glacial and Holocene peat cores have been obtained
from >17 sites at both sides of the Andes (Sites in Fig. 4 and Table 1;

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

e.g. Isla de los Estados, Haberton, Garibaldi, Puerto del Hambre,


Punta Arenas, Rio Rubens; Gran Campo Nevado). The records often
start during the Late Glacial with a minerogenic peat composition
which partly evolved to exclusively ombrogenic composition in the
Holocene (only aerial deposition above ground water level;
McCulloch and Davis, 2001; Biester et al., 2003). They have been
investigated in particular with respect to pollen and microscopic
charcoal abundances (e.g. Pendall et al., 2001 and Chapter 6.2.). The
complex regional distribution of vegetation communities and
resulting modern pollen assemblages is not yet fully understood in
southern Patagonia (see 2; Heusser, 1995). Therefore, palaeoclimatic inferences from past pollen records are often complicated
by local conditions at different sites (e.g. soil-types, drainage, slope,
facing) and plant ecology (e.g. dispersal, disturbance history)
(McCulloch and Davis, 2001; Markgraf et al., 2003; Huber et al.,
2004). In addition, climate implications based on peat decomposition rates, anthropogenic Hg and Pb accumulation (Biester et al.,
2002, 2003), sea spray-related halogenide and heavy metal
contents (Biester et al., 2006) as well as dust deposition (Sapkota
et al., 2007) have been investigated in peat cores.
27 tree ring sites have been studied in southern Patagonia
(Fig. 4; Villalba et al., 2009), including a relatively large number of
sites in the forested area around the Beagle Channel (e.g. Villalba
et al., 2003). From these and further records located close Punta
Arenas, in Torres del Paine area, and at Lago Argentino, regional
paleotemperature reconstructions covering the past several
hundred years have been recently compiled by Neukom et al.
(2010a). From the hyperhumid western side, only few tree-ring
records of the cypress Pilgerodendron Uviferum have been presented from the Gran Campo area (Koch and Kilian, 2002), Isla
Navarino (Aravena et al., 2002) and near to the Pio XI Glacier to the
west of the southern PIF (Rigozo et al., 2007). However, it remains
still unclear, how far temperature or frequent stagnant water
water stress controls the tree ring growth in this hyperhumid
area.
Several open marine and fjord sediment cores have been drilled
between 1989 and 1995 by RV Polar Duke and the Italian RV O.G.S.
Explora. Anderson and Archer (1999) investigated sediment cores
from Bahia Inutil in Tierra del Fuego and the White Side Channel
(Fig. 4) which give implications for marine transgression, but only
limited paleoclimate information. To our knowledge, only few cores
from the RV O.G.S. Explora cruises have been dated and paleoclimatologically investigated in more detail (e.g. Brambati, 2000;
Marinoni et al., 2008). In the marine records the Uk37-index of
alkenones formed by cocolithophorides provides calibrated
temperatures (e.g. Prahl and Wakeham, 1987; Prahl et al., 1988)
which typically reect annual SSTs but may contain rather a spring
or summer SST signal at higher latitudes as alkenone producing
algae do not grow year-round (e.g. Prahl et al., 2010).
More recently, during 15 cruises of RV Gran Campo II since 2002,
sediment basins of the western fjord zone between latitudes 49 to
54 S and former proglacial lakes such Seno Skyring and Seno
Otway have been explored systematically by echo sounding and
multibeam (e.g. Breuer et al., in press-b) and numerous sediment
cores have been recovered in the western fjord region (Fig. 4).
These cores have been investigated applying diverse terrigenous
(e.g. accumulation of terrestrial organic carbon and siliciclastic
material) and marine proxies (e.g. aquatic organic carbon, biogenic
carbonate, and opal) and provide records of continental paleoclimate and marine paleoproductivity changes.
During the IMAGES XV-MD159 Pachiderme cruise of RV Marion
Dufresne in 2007, several long Calypso sediment cores have been
taken from the fjords and Pacic continental slope between 49 and
53 S (Kissel et al., 2007; unpublished cruise report). Among those
the open ocean core MD07-3128 provides the rst regional SST and

IRD-related ice extent reconstruction back to 60 ka BP (Caniupn


et al., 2011).
4. Age control
Age control largely depends on the palaeoclimatic archive and
we here only shortly review the major methods applied to Patagonian records. The stratigraphic background of most archives is
based on 14C ages which have been typically calibrated to calendar
ages in publications of the last decade. However, only some records
consider the Southern Hemisphere calibration curve of McCormac
et al. (2004) which gives 50e80 years younger Holocene ages. In
particular lake sediments often include leaves and macro plant
remnants for reliable radiocarbon dating. Moy et al. (2011)
demonstrate that 14C ages of bulk organic material may be 4
to >6 ka older than coevally deposited macro plant remnants and
pollen concentrates. In sediment core ES1 located in Seno Skyring,
14
C of bulk organic material is around 8 ka older than a well dated
tephra layer (Baeza, 2005). Problems could also occur, if dated plant
remnants have been deposited by terrestrial mass ows which
often include older plant material from eroded peaty soils. In peat
cores, root activity has often modied the structure and layering of
organic material which alter the depth-relationship of 14C ages of
macro remnants.
14
C-based stratigraphies of marine and in particular fjord sediment cores have to deal with variable marine reservoir ages in the
range of 200e800 years. Shallow and freshwater-rich coastal sites
exhibit low reservoir effects of 200e300 years (Kilian et al., 2007a).
For the Puerto Natales region, a mollusc shell collected AD 1939
(pre-bomb) gave a reservoir age of 530 years (Ingram and Southon,
1996), but no exact information have been published for its origin.
Deeper sediment core sites (>300 m water depth) include signicantly older Pacic water which may cause higher reservoir ages
from 550 to 800 years (Caniupn et al., 2011). Since the degree of
wind-induced mixing of supercial freshwater with older fjord
bottom water may have changed during the Holocene, we suppose
that there are also still unknown long-term changes in the reservoir
effect.
Only some piston sediment cores include dating of the uppermost sediment section by the 210Pb method (e.g. Appleby and
Oldeld, 1992) in order to prove a sub-recent sediment surface
which was also applied successfully for dating of the supercial
part of peat cores (Biester et al., 2002).
Tephrochronology represents an important stratigraphic tool for
the Magellan region, since Late Glacial and Holocene tephra layers
are relatively well dated here. Best known are eruptions from the
Reclus volcano at w15.0 ka BP, Mt. Burney at w9.0 ka BP, Hudson at
w7.2 ka BP, Mt. Burney at w4.2 ka BP, Aguilera at w3.0 ka BP and
Mt. Burney at w2.0 ka BP (Stern, 1990, 1992, 2008; McCulloch and
Davis, 2001; Kilian et al., 2003; McCulloch et al., 2005; Moy et al.,
2008). Volcanic glasses of the different volcanoes can be clearly
distinguished due to their characteristic K and Ti contents (Stern
and Kilian, 1996; Stern, 2008). Due to anoxic conditions in the
sediment, most lakes have well preserved even mm-thick tephra
fall deposits (Kilian et al., 2003). In many peat cores thin tephra
layers are often smeared by the root activity which makes their
identication difcult. Some sediment records exhibit disseminated tephra or thin tephra layers after stronger erosion events,
which could be reworked tephra. Tephra particles originating from
the 4.2 and 2.0 ka BP Mt. Burney eruptions have also been found
within stalagmite layers which grew within an open cave. These
have been used as independent control for the detritus correction
of Th/U ages (Schimpf et al., 2011).
Several Patagonian moraines have been dated during the last
decade by using cosmogenic nuclide ages which improves the

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

chronologies of glacier uctuations, in particular for the Late


Glacial. However, most recent comparisons between 14C and
cosmogenic nuclide constrains for moraine ages suggest an
underestimation of the radionuclide ages by 1e2 ka related to
higher Late Glacial production rates (Putnam et al., 2010; Kaplan
et al., 2011).
Finally, a rst study applying high resolution paleomagnetic
dating methods have been applied to Holocene and Glacial sediment sequences at Laguna Potrok Aike (Lis-Pronovost et al., 2012).
This study suggests a large potential of paleomagnetic dating
applicable in both terrestrial and marine archives.
5. Glacial records
5.1. Full glacial (60e20 ka BP)
Paleoclimate records reaching back to the last full glacial period
in southern Patagonia are scarce because of the coverage of vast
areas by the Patagonian ice-sheet. One example is Laguna Potrok
Aike where up to 122 m long lake sediment cores reach back to ca
50 ka BP (Recasens et al., 2011). A preliminary pollen record from
this site suggests signicantly colder glacial temperatures and
probably reduced humidity. More detailed paleoenvironmental
studies based on the glacial records from Laguna Potrok Aike are
currently in progress.
A second example includes marine sediments from the Pacic
continental margin. The only published records from this area are
based on sediment core MD07-3128 recovered at 53 S off the Pacic
entrance of the Strait of Magellan (Caniupn et al., 2011). This welldated core reaches back to ca 60 ka BP (Fig. 5) and provides detailed
information on surface water changes and ice-rafted debris (IRD)
deposition. Alkenone-derived sea surface temperatures (SST) were
up to 8  C lower during the full glacial compared to the early
Holocene and reveal substantial millennial-scale uctuations that
are partly similar to those observed in SST changes at ODP Site 1233
at the Pacic margin further north (41 S) and temperature reconstructions from Antarctic ice cores (Fig. 5E). The strong glacial
cooling at site MD07-3128 implies a substantial northward expansion of polar water masses and the Southern Ocean fronts with the
sub-Antarctic front probably located close to the site during the full
glacial. An interesting feature of the SST record is a long-term
warming trend of 2  C from ca 50 to 25 ka BP and a successive
3  C cooling trend culminating at ca 19 ka BP when the coldest SST
were recorded. These trends have been related to variable supply of
cold melt water from the adjacent Patagonian ice-sheet (Caniupn
et al., 2011). A more direct proxy for changes in the extent of the
Patagonian ice-sheet comes from the IRD (Fig. 5B). This record
shows pronounced IRD pulses between ca 30 and 18 ka BP that
partly overlap with glacial advances reconstructed on the eastern
side of the ice-sheet in southern Patagonia (Kaplan et al., 2008;
Sugden et al., 2009) and further north at the Pacic margin of the
northern ice-sheet (Kaiser and Lamy, 2010). Patagonian ice-sheet
advances have been mechanistically linked to dust maxima recorded in Antarctic ice-cores (Caniupn et al., 2011; Kaiser and Lamy,
2010; Sugden et al., 2009; Fig. 5A) since the ne-grained glacial
detritus from the eastern Patagonian outwash plains was detected
as the major dust source area for Antarctica. Interestingly, terrestrial
reconstructions indicate a maximum ice-extent in southern Patagonia at around 25 ka BP and a slightly retreating ice extent already
before the global LGM (Fig. 5; Kaplan et al., 2008). This regional
retreat was explained by reduced precipitation during the coldest
glacial interval due to northward displaced westerlies. A weaker
southern SWW margin would have enabled colder supercial melt
water to migrate westward. This could explain the observed cooling
trend from 25 to 19 Ka BP at the MD07-3128. Such an accumulation-

driven early glacier retreat is consistent with ice-sheet modelling


results that suggest reduced precipitation in southern Patagonia
during the Last Glacial Maximum (Hulton et al., 2002).
5.2. Deglaciation (20e10 ka BP)
Along the central Strait of Magellan rapid glacier retreat started
after the last major set of moraines were built prior to 17.5 ka BP
(Moraine stage D: McCulloch and Bentley, 1998; McCulloch et al.,
2005; Kaplan et al., 2008; Sugden et al., 2009) leaving the Puerto
del Hambre site around 80 km southward ice-free by 17.3 ka BP
(McCulloch and Davis, 2001). Ice retreat in the Seno Skyring region
occurred largely in phase (Fig. 6). However, there are indications of
a slightly earlier response in this area before 18 ka BP, most likely
because the Seno Skyring glacier catchment was more sensitive to
a small initial elevation increase of the equilibrium-line altitude
(Kilian et al., 2007b). Deglacial warming in the Southeast Pacic
started likewise earlier (between 18 and 19 ka BP) both at site
MD07-3128 directly offshore the Strait of Magellan (Caniupn et al.,
2011) and further north at ODP Site 1233 (Lamy et al., 2007). In both

Fig. 5. Glacier extent in southern Patagonia compared to SST and Antarctic temperature changes during the last Glacial. A: Glacial stages in the Magellan Strait region
(Sugden et al., 2009). B: Non-sea-salt calcium from EPICA Dome C (EDC) (a proxy for
dust content changes in Antarctic ice cores (Fischer et al., 2007). C: Percentage of
>150 mm carbonate-free sediment fraction as a proxy for IRD in core MD07-3128 off
the Pacic entrance of the Magellan Strait (Caniupn et al., 2011); D: Timing of glacial
ice-sheet variations in the Magellan Strait region based on 10Be ages (ice retreat
towards the lower left; Kaplan et al., 2008). E: Alkenone SST record of core MD07-3128
(Caniupn et al., 2011); F: Alkenone SST record of ODP Site 1233 at the Pacic continental slope at 41 S (Kaiser and Lamy, 2010). G: Antarctic surface temperature changes
(deviation from mean of the last millenium; Jouzel et al., 2007) plotted on the new
Lemieux-Dundon time-scale (Lemieux-Dudon et al., 2010). Grey bars indicate
synchronous positive temperature excursions.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

SST records, the initial warming lasted until ca 15 ka BP and was


followed by plateau and a second warming step starting at ca 12.5 ka
BP leading into the pronounced early Holocene warming period
(Fig. 6). The SST plateau largely coincides with the Antarctic Cold
Reversal (ACR) known from ice-cores. Along the Strait of Magellan
pollen assemblages indicate more humid conditions after 16.9 ka PB
until at least 14.6 ka BP which may have caused a restricted readvance forming moraine stage E at the Northern Isla Dawson
(McCulloch and Davis, 2001). This limited and partly accumulationdriven re-advance partly overlaps with the ACR. The more than
100 km ice retreat from moraine stage E towards the Cordillera

Fig. 6. Deglaciation history from 22 to 8 ka BP: A: d18O record from NGRIP (North
Greenland Ice Core Project Members, 2004); B: Ice retreat phases at Lago Argentino
(Kaplan et al., 2011), C: Glacier retreat stages C to E and suggested retreat lengths along
the Central Strait of Magellan (McCulloch et al., 2005) as well as retreat constrains (red
squares) by Boyd et al. (2008) with a small early Holocene re-advance (CD-RE). D:
Retreat lengths from the Seno Skyring (Kilian et al., 2007b). E: Alkenone SST record of
core MD07-3128 (Caniupn et al., 2011); F: Antarctic surface temperature changes
(deviation from mean of the last millenium; Jouzel et al., 2007) plotted on the new
Lemieux-Dundon time-scale (Lemieux-Dudon et al., 2010). G: Global sea level estimates compiled by Siddall et al. (2003). The marine transgression at the western
entrance of the Strait of Magellan (MT West, Kilian et al., 2007b) and its eastern
Atlantic entrance (MT East, McCulloch and Morello, 2009) are indicated.

Darwin was probably very fast, since marine sedimentation in the


Marinelli Fjord started at around 14 ka BP (Boyd et al., 2008).
A second warming step occurred during the Northern Hemisphere Younger Dryas period. The possible occurrence of a Younger
Dryas (YD) cold period in southern South America has been intensively debated over the past decade as pollen records in the Chilean
Lake District and on Chilo Island have been interpreted in terms of
a YD cooling (Moreno et al., 1999, 2001) whereas pollen records
from the Taitao peninsula and the Chonos archipelago, revealed no
such cooling during the YD (Bennett et al., 2000). Further south at
Lago Argentino, Ackert et al. (2008) reconstructed advancing
glaciers during the YD that culminated around 11 ka BP. This glacier
advance on the lee-side of the Andes was attributed an increased
precipitation from easterly sources and does not imply colder
conditions. However, a recent recalculation of the cosmogenic
nuclide dates from Lago Argentino suggests that this Lago Argentino
glacier advance took place earlier and most of the advance falls into
the ACR (Fig. 6; Kaplan et al., 2011). This would be consistent with
ACR glacier advances in the region of the Strait of Magellan Strait
(Sugden et al., 2005) and in the Torres del Paine area (Moreno et al.,
2009b). Taken together, the Deglacial records conrm the Antarctic
timing of paleoclimate changes in southern Patagonia.
Pollen records indicate the development of a treeless and open
landscape (tundra-like) during the deglaciation, dominated by
eurythermal herbs and shrubs like Acaena, Empetrum, Gunnera and
Poaceae, followed by the expansion of woody elements (e.g.
Nothofagus) in response to an increase in temperature and moisture (e.g. Markgraf, 1993; Heusser, 1995, 2003; Huber et al., 2004;
Villa-Martnez and Moreno, 2007; Mancini, 2009; Moreno et al.,
2009a, 2010; Markgraf and Huber, 2010; Ponce et al., 2011).
Interesting is the dissimilar timing and structure in the forest
expansion between northern and southern sites located along the
east of the Andes between latitudes 49 to 55 S. We observe that in
northern sites (Cerro Frias, Brazo Sur, Vega andu, Lago Guanaco,
Rio Rubens and Punta Arenas; Fig. 7), the Nothofagus expansion
occurs between >16 and 11 ka BP and gradually, whereas in more
southern locations (Puerto del Hambre, Estancia Esmeralda,
Puerto Haberton, Paso Garibaldi, Isla de los Estados) its transitions
occurs much later (between 11.0 and 9 ka BP) and more abrupt.
The humidity-sensitive and primarily thermophilous nature of
Nothofagus suggests: (i) Favourable (unfavourable) moisture and/
or temperature conditions during the deglaciation and earliest
Holocene in northern (southern) sites, or (ii) non-climatic features
(underlying mechanisms) controlling the forest expansion
dynamics. Based on genetical evidences, Premoli et al. (2010)
suggest that Nothofagus trees survived full glacial conditions
within ecological niches on Tierra del Fuego at latitudes 54 S.
Similar conditions must be occurred around the latitude 52 S
where Nothofagus pollen present in glacial sediments of the Potrok
Aike record suggest small sheltered patches of Nothofagus
surviving at the foot of the Andes (Wille et al., 2007; Recasens
et al., 2011). Therefore northern and southern locations have the
same Nothofagus dispersal potential, and explain their dissimilitude in these terms is not likely. Nevertheless at some sites, the
deglacial Nothofagus expansion was probably delayed hundreds or
thousands of years with respect to the deglacial climate change
(Heusser, 1995; Pisano, 1977) due to long term soil evolution and
consolidation within metastable glacial detritus (Heusser, 1995;
Pisano, 1977; Ballantyne, 2002). Therefore if climate was the main
responsible mechanism in the dynamic of the Nothofagus expansion, it may indicate a northerly position of the SWW core until the
earliest Holocene (Lamy et al., 2010).
During the Late Glacial glacier retreat, a stepwise marine
transgression occurred into the proglacial lakes, forming an
extended Patagonian fjord system. Kilian et al. (2007a) determined

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Fig. 7. Nothofagus pollen records from sites within a precipitation range of


400e1200 mm/year at the eastern side of the Andes (Isla de los Estados: Ponce et al.,
2011; Haberton and Passo Garibaldi: Pendall et al., 2001; Puerto del Hambre I: Heusser,
1995; Puerto del Hambre II: McCulloch and Davis, 2001; Punta Arenas, Rio Rubens:
Huber et al., 2004; Lago Guanaco in the Torres del Paine Park: Moreno et al., 2010). The
relative amount of Nothofagus, the dominant tree species, indicates the vicinity to the
transition between Magellan forest and open parkland and/or steppe vegetation.

a marine transgression to the western Strait of Magellan at around


14.3 ka BP, related to melt water pulse 1A (Fig. 6) and nearly coeval
to rst marine inuences in the Marinelli fjord (Boyd et al., 2008).
The Atlantic transgression to the Strait of Magellan occurred later at
around 9 ka BP when the further global sea level rise passed the
shallower eastern sills (McCulloch and Morello, 2009). The marine
transgression cross-cut the Andes and reached the Seno Otway at
around 14 ka BP and Seno Skyring as well as Seno Ultimo Esperanza
at around 10.3 ka BP (Kilian et al., 2007a; Stern et al., 2011).
6. Holocene
6.1. Paleovegetation
The climate transition zone to the east of the Andes
with annual precipitation of 500e1000 mm/yr was probably

characterised by rather open Nothofagus pumilio-dominated


subantarctic deciduous forest during most of the Holocene (e.g.
Heusser, 1995; McCulloch and Davis, 2001). Its local evolution
and eastward expansion towards the steppe may have been
controlled by available moisture, in particular during summer
(e.g. Pisano, 1977). In this area the available moisture depends on
the complex interplay between precipitation and amount of
evaporation which is controlled by wind velocities, humidity,
temperature as well as on site-specic soil-types and drainage
conditions of underlying substrate. In particular, the climate
control on the degree of evaporation at different sites have been
poorly quantied and therefore remains a major problem for
palaeoclimatic interpretations based on pollen assemblages in
this region.
Eight Nothofagus pollen records from this humidity-sensitive
transition zone to the east of the Andes between latitudes 49 to
55 S are summarised in Fig. 7. The records of the southernmost
locations Isla de los Estados (54 S, Ponce et al., 2011) and Haberton
(54 S; Pendall et al., 2001) show a relatively late and gradual forest
expansion between 11.0 and 5.0 ka BP. High values of Ericaceae
(>80%) and low values of Nothofagus (>20%) on Isla del Estado
suggest an open and perturbed vegetation, in particular between
10.3 and 8.3 ka BP, which has been related to dry and warm climatic
conditions (Unkel et al., 2010).
The Puerto del Hambre and Passo Giribaldi records (e.g. Heusser,
1995; Heusser et al., 2000; McCulloch and Davis, 2001) show a full
evolved forest (with >75% Nothofagus pollen) rst after 8.5 ka BP
when regional temperatures dropped by around 1.5  C (Chapter
6.2.) and the SWW strength probably decreased (Chapter 6.3.; Lamy
et al., 2010). Previously, between w12.0 and 9.0 ka BP, the Puerto
del Hambre and Estancia Esmeralda records show increased pollen
degradation as well as humied peat deposits indicating very low
effective moisture (McCulloch and Davis, 2001). This could indicate
less precipitation and/or higher evaporation rates possibly related
to stronger westerlies combined with higher temperatures (Lamy
et al., 2010, see Chapter 6.3.).
The Hambre record shows a signicant reduction of tree pollen
during the last 5.5 ka BP (Figs. 7 and 13F; see also Chapter 6.6. on
possible human impact), whereas the pollen records from Isla de
los Estados (Ponce et al., 2011), Haberton, Passo Giribaldi (Pendall
et al., 2001), and Punta Arenas (Heusser, 1995) show evolved
forest vegetation after 5.5 ka BP and persisting throughout the Late
Holocene. The Estancia Esmeralda site on the northern Isla Dawson
shows only restricted Nothofagus pollen (<40%) reecting the
signicantly lower precipitation at this site situated further east
along the regional precipitation gradient (Figs. 1 and 2).
Records from the more northern sites at Punta Arenas (Heusser,
1995) and Rio Rubens (Huber et al., 2004; Markgraf and Huber,
2010) as well as Lago Guanaco (Moreno et al., 2010) and Laguna
Vedga andu (Villa-Martnez and Moreno, 2007) in the Torres del
Paine area and locations close to Brazo Sur of Lago Argentino
(Mancini, 2009; Wille and Schbitz, 2009) show a more gradual
evolution to a denser forest starting between 16 and 14 ka BP and
lasting throughout the Holocene. This has been interpreted also as
an eastward migration of the forestesteppe transition. The relatively high Nothofagus pollen (up to 50%) at the Punta Arenas site
between 16 and 12 ka BP are extraordinary and difcult to explain.
In this case the Late Glacial section of the peat/sediment core could
have been contaminated by Holocene material during drilling with
a Hiller corer.
The above discussed pollen records does not exhibit a synchronous behaviour during phases of regionally known climate
perturbations, like the 8.5 to 7.5 ka BP period, the beginning of the
Neoglacial at ca 5.5 ka BP, the 3.5 to 2.5 ka BP period, the Medieval
Climate Anomaly (MCA), and the LIA, which have been found in

10

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Fig. 8. Pollen records from the hyperhumid western side of the Andes: (A) GC1 site near the Gran Campo Nevado (Fesq-Martin et al., 2004; Lamy et al., 2010) and (B) TM1 site on
Tamar Island (Lamy et al., 2010).

many other records (e.g. Figs. 10 and 12). This individual behaviour
may be related to local microclimates and individual sensitivities of
the ecosystems as well as locally distinct available moisture
depending on wind exposition and local soil types.
Pollen records from more eastern sites in the Patagonian steppe
have been published from Laguna Potrok Aike (e.g. Mayr et al.,
2007a, 2009) and from Lago Cardiel (Markgraf et al., 2003). These
pollen assemblages are partly controlled by a long distance transport of Nothofagus from Andean sites and/or by local moisture
conditions which are slightly inversely correlated to westerly wind
strength due to regular precipitation events derived from South
Atlantic sources (Fig. 10D). Therefore their interpretation on the
local paleovegetation and their climate relationship is not
straightforward. The SWW-relationship of the Laguna Potrok Aike

tree pollen record related to long distance pollen transport from


Andean sources is discussed in Chapter 6.3.
Only two well-dated pollen records have been presented from
the hyperhumid side of the southern Andes up to now. They include
the Gran Campo Nevado area (Fesq-Martin et al., 2004) and Lake
Tamar (Lamy et al., 2010) along the western Strait of Magellan (sites
GC2 and TML in Fig. 4). Since the climate sensitivity of these
hyperhumid ecosystems is poorly investigated, the pollen variations must be evaluated with care. In general the precipitationdependent height of stagnant water in the peaty soils as well as
local drainage conditions control to what extent peat bogs or,
alternatively, forest evolved.
In the GC2-record of the Gran Campo area the amount of
humidity sensitive hygrophyte pollen was interpreted within the

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

11

Fig. 9. Holocene paleotemperatures (and precipitation): A: Combined alkenone SST


record from core GeoB 3313-1 (Lamy et al., 2002) and ODP Site 1233 (Lamy et al., 2007)
at the Pacic continental slope at 41 S (the published GeoB 3313-1 SST data have been
corrected by 0.7  C in order to match the overlapping period with the ODP 1233
record). B: Tree ring-based temperature reconstruction for southernmost South
America (Neukom et al., 2010a); C: Alkenone SST record of core MD07-3128 (Caniupn
et al., 2011) D: Diatom-based SST-record from core PS2102-2 in the South Atlantic at
53 S (Bianchi and Gersonde, 2004); E: Terrestrical organic carbon accumulation in core
TML-1 from Lake Tamar as a record for paleo-precipitation (Lamy et al., 2010). F:
Antarctic surface temperature changes (deviation from mean of the last millenium;
Jouzel et al., 2007) plotted on the new Lemieux-Dundon time-scale (Lemieux-Dudon
et al., 2010).

context of other multiproxy records to reect humidity and


related SWW-strength during the Holocene (Fig. 8A; Lamy et al.,
2010). Cyperacae and Astelia which characterize plants of the
peat and wetland are more frequent in the Early Holocene
between 11.2 and 8.5 ka BP and decrease towards the Neoglacial.
Nothofagus growth which is hampered by high stagnant water
levels became more frequent during the Middle and Late Holocene. Nothofagus pollen show a strong perturbation at the GC-2
site associated with a large peak of the pioneer plant Gunnera
from 4.1 to 2.9 ka BP. This was probably produced by local forest
decay and millennial scale regional perturbation of plant
communities due to long-term sulphur release and additional
acidication of soils after deposition of a sulfur-bearing tephra
layer from the 4.15 ka Mt. Burney eruption (Kilian et al., 2006).
During the last millennium, a higher amount of peat land versus
tree pollen could indicate increased moisture and/or may reect
regionally up to the 1.5  C lower temperatures after 0.7 ka BP
(Neukom et al., 2010a,b).

Fig. 10. Precipitation and SWW-related records: A: Lake level reconstruction of Lago
Cardiel (Stine and Stine, 1990; Ariztegui et al., 2010); B: Magnetic susceptibility record
of core CAR 99-9P from Lago Cardiel as an indicator for westerly wind related drift
deposition in the lake (Gilli et al., 2005). C: Pollen-based reconstruction of precipitation near Brazo Sur (Lago Argentino; Tonella et al., 2009). D: Andean forest taxa in
a sediment core of Laguna Potrok Aike as wind indicator (Mayr et al., 2007a, 2009); E:
Pollen record of Laguna Guanaco in the Paine National Park as proxy for humidity
changes related to the SWW (Moreno et al., 2010). F and G: Clay/silt ratio and illite
content of core SK1in the eastern Seno Skyring as proxy for wind-induced long
distance eastward sediment transport (Lamy et al., 2010). H: Terrestrial organic carbon
accumulation in Tamar Lake sediment core TML-1 as a paleo-precipitation proxy (Lamy
et al., 2010). I: C/N ratios in sediment core PC-18 of Lago Fagnano as indicator for
precipitation-dependent terrestrial plant supply (Moy et al., 2008) and J: Fe content
record of core LF06-16 from Lago Fagnano as proxy for precipitation-related siliciclastic
sediment-input (Waldmann et al., 2010). Thick line is 80-point running average.

The Lake Tamar (Site TML in Fig. 4) pollen record shows


a distinctive forest succession (Fig. 8B; Lamy et al., 2010), ranging
from a Nothofagus-dominated forest (w13.2e9.3 ka BP), to
NothofaguseDrimys forest (w9.3e5.2 ka BP), and leading to the

12

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Fig. 11. Compilation of selected paleoclimate records and Holocene glacier advances in
the southernmost Andes: A: Paleoprecipitation based on tree pollen from Lake
Guanaco in Paine National Park 50 S (Moreno et al., 2010), B: Alkenone SST record of
ODP Site 1233 at the Pacic continental slope at 41 S (Kaiser et al., 2005; Lamy et al.,
2007); C: Alkenone SST record of core MD07-3128 (Caniupn et al., 2011), D) Terrestrial
Corg accumulation in core TML1 from Lake Tamar as a precipitation proxy (Lamy et al.,
2010), E: d18O values of stalagmite MA1 in the Southern Andes at 53 S as indicator for
precipitation (Schimpf et al., 2011), F: Glacier advance phases in the southernmost
Andes from 1: Wenzens and Wenzens (1998), 2: Aniya (1995), 3: Mercer (1982), 4:
Masiokas et al. (2009a) and from the Antarctic Peninsula and adjacent islands (Hall,
2009). GI: Magnetic susceptibility as indicator for glacial clay in core JPC67 core
from the Marinelli fjord next to Cordillera Darwin (Boyd et al., 2008) and H: Ice rafted
debris (IRD) in the South Atlantic at 53 S (Hodell et al., 2001).

establishment of mixed evergreen forest (w5.2 ka BP to present).


The establishment of NothofaguseDrimys forest and nally the
mixed evergreen forest was interpreted as an important precipitation decrease towards the late Holocene. High relative percentage
of Misodendrum, a hemiparasite that infests several Nothofagus
species (Markgraf et al., 1981), during the early Holocene
(w11.3e9.3 ka BP) has been interpreted as Nothofagus forest
growing under limited ecological conditions (e.g. swamp soils). This
palaeoclimatic interpretation is consistent with high values of C/N
and accumulation rates of terrestrial organic carbon in Lake Tamar
triggered by enhanced precipitation during this period (Lamy et al.,
2010). Nevertheless high values in the total pollen concentration
and inux parameters during the early Holocene suggest denser
plant cover than during the late Holocene.
6.2. Paleotemperature
Tree-ring widths (e.g. Villalba et al., 2003; Neukom et al., 2010a,
b), hydrogen isotopes of mosses (Pendall et al., 2001), C and O

Fig. 12. High resolution records interpreted in terms of SWW behaviour, ENSO
changes, sun activity, position, strength of the ITCZ, and/or Antarctic temperatures: A:
10
Be-based sun activity reconstruction (Steinhilber et al., 2009); B: Ti content record of
ODP Site 1002 from the Cariaco basin, indicating position and strength of the ITCZ
(Haug et al., 2001); C: Grey scale variations in a sediment core from Laguna Pallcacocha
in the Ecuadorian Andes as a proxy for ENSO events (Moy et al., 2002); D: Colour-scale
variations as indicator for ENSO-related ood-derived lithic grains in sediment core
106KL from the Peruvian coast (Rein et al., 2005); E: Fe content changes in core in
LF06-16 from Lago Fagnano as indicator for SWW-related precipitation (Waldmann
et al., 2010). F and G: d18O values and Y content in the stalagmite MA1 in the
Southern Andes at 53 S as indicator for drip rates and SWW strength (Schimpf et al.,
2011). H: Antarctic surface temperature changes (deviation from mean of the last
millennium; Jouzel et al., 2007).

isotopes of stalagmites (Mhlinghaus et al., 2008, 2009), and the


UK37 index of alkenones (e.g. Caniupn et al., 2011) in marine
sediments were used to calculate calibrated Holocene temperatures. While most of these records show relatively low temperatures for the LIA after 0.7 ka BP, there are no clear trends or
uctuations for the whole Holocene except for the open ocean
record of Caniupn et al. (2011) which reveals a well-dened early
Holocene SST maximum. In the following we discuss the different
paleotemperature reconstructions and possible reasons for
inconsistencies.
Pendall et al. (2001) presented a Late Glacial and Holocene
temperature reconstruction which is based on a hydrogen isotope
record of mosses from a peat bog at Tierra del Fuego. This record
does not show clear Holocene uctuations within the given error
range and the effect of likely changes in the rain water isotopy
remains unclear. The d13C record from bulk organic material in lake
sediments of the Lago Fagnano was interpreted to reect partly
temperature-controlled
aquatic
bioproductivity
indicating

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

13

chronologies from the hyperhumid southern Andes do not show


a temperature-relationship when compared with 100 year weather
station data from Evangelistas (Koch and Kilian, 2002; Rigozo et al.,
2007).
Based on the distinct temperature dependence of the kinetically
controlled C and O isotope fractionation in two stalagmites for the
southernmost Andes at latitude 52 S (Ma1 and MA2 records in
Mhlinghaus et al., 2008, 2009), paleotemperatures have been
calculated for the past 5 ka BP. However, this reconstruction does
not consider temperature dependent changes in the isotopic
composition of rain water (Schimpf et al., 2011), in particular,
during periods of signicantly lower temperature, like during the
LIA (1.5  C; see tree-ring temperatures in Fig. 9).
An alkenone-temperature record was obtained from site MD073128 at the continental margin off the Pacic entrance of the
Magellan Strait (Caniupn et al., 2011). This record provides high
time resolution for the Last Glacial between 55 and 13 ka BP (Fig. 5;
Chapter 5), whereas the Holocene record has low time resolution.
However, the record shows an early Holocene temperature
maximum between 11.5 and 8.5 ka BP (Fig. 9) similar to open
marine records off Chile further north (e.g. ODP Site 1233, 41 S;
Kaiser et al., 2005). Holocene fjord SST records from southern
Patagonia have not yet been published. Fjord alkenone SST records
from the northern Patagonian fjords (Sepulveda et al., 2009) and
on-going studies in the south (Magaly Caniupan and Claudia Aracena, personal communication) suggest SST cooling between 2.5
and 3.5 ka BP and the LIA period and hint to the large potential of
this method for regional paleotemperature reconstructions. The
2.5e3.5 ka cooling has also previously been suggested based on
peat records in southern Patagonia (Van Geel et al., 2000;
Chambers et al., 2007) and might be a global event (Wanner et al.,
2011).
6.3. Paleoprecipitation, water mass balance and possible
relationship to SWW strength

Fig. 13. Relationship between human occupation in the Magellan region and Tierra del
Fuego and paleoclimate records: A: Antarctic surface temperature changes (deviation
from mean of the last millennium; Jouzel et al., 2007) plotted on the new LemieuxDundon time-scale (Lemieux-Dudon et al., 2010). B: Terrestrial organic carbon accumulation in Tamar Lake sediment core TML-1 as paleo-precipitation proxy for the
Andean area (Lamy et al., 2010). C and D: Y content and C/N ratios in a sediment core
from Lake Hambre as a proxy for terrestrial erosion around the lake (Hermanns and
Biester, 2011). E: Human occupation phases at Tierra del Fuego (Morello et al., 2012).
F: Charcoal as well as Graminae and Nothofagus pollen Puerto del Hambre (Heusser,
1995). G: Black arrows indicate human occupation phases (for details and references
see text) and global sea level curve after Siddall et al. (2003) and present day water
depths at Primera and Segund Angostura of the Strait of Maggellan.

a Holocene Hypsithermal between 7.0 and 5.0 ka BP (Moy et al.,


2011). However, this record does not cover the Early Holocene.
For southernmost Patagonia tree-ring based temperatures have
been reconstructed for the last 1.2 ka BP (Villalba et al., 2003;
Neukom et al., 2010a,b) indicating a signicant temperature
decrease of about 1.5  C between 0.7 and 0.5 ka BP (Fig. 9).
However, due to the limited length of most tree ring chronologies,
only the last 0.6 ka BP have smaller error bars and are more reliable
(Neukom et al., 2010a,b). Though average annual temperatures (at
the same altitude) do not change signicantly across the Andean
divide, tree-ring based temperatures from the eastern slopes of the
Andes, in particular north of 52 S, may respond to the higher
summer temperatures (higher seasonality). Furthermore, tree-ring

A major focus of paleoclimate research in southern Patagonia is


the reconstruction of changes in the intensity and latitudinal
position of the SWW (e.g. Gilli et al., 2005; Villa-Martnez and
Moreno, 2007; Mayr et al., 2007a, b, 2009; Moy et al., 2008, 2011;
Moreno et al., 2009a, 2010; Kastner et al., 2010; Lamy et al., 2010;
Waldmann et al., 2010; Fletcher and Moreno, 2011; Fletcher and
Moreno, 2012). The SWW crosses the entire southern tip of South
America (Fig. 1B; e.g. Garreaud et al., in press) and therefore pure
wind proxies should give similar results from different sites.
However, such proxies (e.g. wind-related transport of pollen or
specic clay-types in supercial waters) are rare and their relation
to wind strength is not always straightforward. Therefore, most
SWW reconstructions are based on the relationship between
precipitation and SWW strength. As shown in Fig. 1C, the western
side of the Andes the western fjord system exhibit a good correlation of r 0.5e0.8 between precipitation and westerly wind
strength while the eastern slopes of the Andes (e.g. Punta Arenas
and Torres del Paine area) exhibit a weaker correlation of
r 0.2e0.4. In this area the wind-related effect on evaporation for
the local water balance becomes important, in particular during
summer. Further to the east, in the steppe region, westerly wind
strength is slightly negatively correlated with precipitation due to
increased easterly moisture from the Atlantic, in particular when
the SWW strength is weak. This applies for example to Lago Cardiel
and Laguna Potrok Aike.
Already 20 years ago (Stine and Stine, 1990), lake level changes
of Lago Cardiel (Site in Fig. 4) have been used to depict the regional
water balance during the Late Glacial and Holocene (Fig. 10A).
These reconstructions are consistent with more recent studies (e.g.

14

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Ariztegui et al., 2010; Cusminsky et al., 2011) on precipitation/


evaporation ratios deduced from lacustrine microfossil assemblages. However, the lake catchment receives precipitation originating from the westerlies during summer months, but also
signicant moisture form Atlantic sources in particular during
winter. The lake level increased signicantly from 80 m
(compared to present day) after 11.5 ka BP and reached
a maximum of 60 m at around 10 ka BP (Fig. 10B). After 8 ka BP,
only minor changes of 20 m are recorded. Gilli et al. (2005)
suggested that the magnetic susceptibility of late and middle
Holocene sediment records in this lake reect westerly wind
strength. Between 8 and 5 ka BP, this record indicates higher
westerly wind intensities which afterwards decrease until the
present (Fig. 10B).
Tonella et al. (2009) calibrated a pollen record from a site close
to Lago Argentino (Site location see in Fig. 4) with respect to
precipitation. For this propose, they investigated surface pollen in
a W-E transect and compared them with the present day precipitation gradient. Their record indicates a successive precipitation
increase from 12.0 to 3.5 ka BP (from ca 300e600 mm/year) and
a subsequent general decrease superimposed by a secondary
maximum around the Medieval Climate Anomaly and
a signicant decrease during the LIA (Fig. 10C). However, the calibration has a large error range of about 150 mm/yr and the impact
of evaporation on the available moisture in the soils has not been
considered clearly.
Mayr et al. (2007a, 2009) show the relative contribution of
Andean Forest Taxa (AFT) during the past 12 ka BP based on
a sediment record from Laguna Potrok Aike (Site location in Fig. 4)
(Fig. 10D), situated at 52 S around 80 km east of the Andean forest
belt. They interpreted the AFT content as predominately winddependent eastward pollen transport from forested Andean sites.
The highly resolved record shows pronounced decadal to centennial variability, but no signicant long term trend throughout the
Holocene. A marked increase of AFT associated with a low lake level
is registered between 8.5 and 7.5 ka BP (Fig. 10D). This period was
interpreted as the strongest Holocene SWW phase characterised by
low precipitation and pronounced evaporation. For the same
period, pollen records from Lago Guanaco (Moreno et al., 2010,
Fig. 10E) and Vega andu (Villa-Martnez and Moreno, 2007) (both
Torres del Paine region) show the lowest tree pollen content during
the past 12 ka BP. In contrast to the results of Mayr et al. (2007a,
2009), the period from 8.5 to 7.5 ka BP was interpreted as the
weakest SWW phase of the Holocene. The systematic increase of
tree pollen in both Torres del Paine records has been interpreted as
increasing westerly wind strength during the Holocene, culminating during the LIA (Villa-Martnez and Moreno, 2007; Moy et al.,
2008; Moreno et al., 2010). This interpretation is based on an
assumed positive (Fig. 1C) of tree pollen, precipitation and SWW
strength. More importantly, the impact of wind velocities and
temperature on evaporation and related available moisture
remains unclear. Lago Guanaco is relatively alkalic (Moy et al.,
2008), suggesting that evaporation is an important factor in this
area which dries out soils and plants, especially during the windy
summer periods (see discussion in Lamy et al., 2010). An alternative
interpretation of the Torres del Paine pollen records could be that
the available moisture in the soils of the hilly catchment was
strongly reduced during windier and warmer periods (e.g. between
ca 7.5 and 8.5 ka BP interval) or more generally during the early and
middle Holocene. Such an alternative interpretation would be
consistent with previous pollen interpretation in this region
(Heusser, 1995) and records west of the climate divide (Lamy et al.,
2010; see below) (Fig. 10E).
Further to the south at Tierra del Fuego, sediment records from
Lago Fagnano (Fig. 4) have been used for westerly wind

reconstructions (Waldmann et al., 2010; Moy et al., 2011). Though


less pronounced than the western fjord region (see below), this
area is characterised by a clear positive correlation of precipitation
and wind strength (Fig. 1C). Waldmann et al. (2010) used the iron
content record of a sediment core from the lake as a proxy for
precipitation-dependent siliciclastic sediment input (Fig. 10J). This
ultra-high resolution record shows pronounced decadal and some
centennial changes, but may also depict single extreme storm and/
or precipitation periods. There is however no clear long term trend
in this record. Moy et al. (2011) use C/N ratios from another sediment core retrieved from Lago Fagnano to characterize the possibly
rain-dependent supply of terrestrial organic matter. In contrast to
the above described iron record, the continuously increasing C/N
ratios since 8 ka BP are interpreted to reect a continuous increase
in precipitation and SSW strength throughout the Holocene,
culminating in the LIA. As such this record would be compatible
with the tree pollen interpretations based on the above discussed
Torres del Paine records (Villa-Martnez and Moreno, 2007;
Moreno et al., 2010). However, the millennial-scale changes of the
Vega andu pollen record are also not visible in the C/N record
from Lago Fagnano. An increased accumulation of terrestrial
organic matter in the lake sediment could alternatively also be
explained by increasing thicknesses of peaty soils throughout the
Holocene which would enable more terrestrial organic carbon
supply during the Late Holocene, even during similar or less
intensive precipitation and runoff. Pollen records from Puerto del
Hambre (McCulloch and Davis, 2001) within a similar climatic
setting at the lee-side of the Andes which have been recently
interpreted to reect regional precipitation and SWW strength
(McGlone et al., 2010) do not clearly show a correlation with the C/
N record of the nearby Lake Hambre (Fig. 13B; Hermanns and
Biester, 2011).
Lamy et al. (2010) present a Holocene multiproxy compilation
from sites within the hyperhumid Andes where the present-day
correlation between precipitation and SWW strength is highest
(Fig. 1C). An increase in precipitation and SWW strength was
deduced from humidity sensitive pollen (Fig. 8), precipitationdependent terrestrial organic carbon accumulation in lake and
fjord sediments, salinity dependent decrease in biogenic carbonate
accumulation and the >80 km long-distance eastward transport of
illite-rich Andean clay from the Patagonian Batholith towards the
eastern sector of the former proglacial Lake Seno Skyring (two of
the proxy records are shown in Fig. 10FeH). These records indicate
a stronger SWW from 12.5 to 8.5 ka BP, a transitional weaker SWW
phase from 8.5 to 5.5 ka BP and a subsequent interval (Neoglacial)
with a more variable but in general lower SSW strength. These
results contrast especially with the above described pollen interpretation form the Torres del Paine area (Villa-Martnez and
Moreno, 2007; Moreno et al., 2010) and the C/N record from Lago
Fagnano given by Moy et al. (2011).
Taken together, the above presented records (most of them
shown in Fig. 10) of Holocene westerly wind strength changes do
presently not provide a consistent picture concerning the behaviour of the SWW during the Holocene. While pollen interpretations
from the Torres del Paine region and the C/N record of Lago Fagnano
suggest weak westerlies in the Early Holocene and stronger westerlies during the Neoglacial (culminating during the LIA), records
from the hyperhumid western side of the Andes in general show
the reverse pattern. The only possibility to explain these contrasting trends would be subtle latitudinal shifts of the westerlies
causing contrasting wind intensities over short distances (ca
150 km between Torres del Paine and the Gran Campo area).
Considering the modern extension of the SWW core, such smallscale shifts for explaining contrasting precipitation trends in
southern Patagonia appear rather unlikely.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Other records such as the AFT from Potrok Aike and the Fe
intensities from Lago Fagnano indicate high short term variability, but only subtle long term trends. Independent of the
individual SWW related interpretation of precipitation records,
the general trends allow to distinguish characteristic periods.
These periods include the latest Glacial to early Holocene from
12.5 to 8.5 ka BP and a period from 8.5 to 7.5 ka BP which is
possibly related to global climate reorganization related to the
8.2 ka BP event which was observed at many sites of the
Northern Hemisphere and also at lower latitudes between 8.4
and 7.9 ka BP (e.g. Rohling and Plike, 2005). From 7.5 to 5.5 ka
BP most records reveal a transitional phase leading into a distinct
late Holocene interval. Pronounced perturbations are also
obvious in the period from 3.5 to 2.5 ka BP and during the LIA
after 0.65 ka BP (Fig. 12).
6.4. Glacier uctuations
On a global scale, Holocene glacier uctuations are often
ascribed to temperature and related ablation changes (Anderson
et al., 2010), especially during the past few centuries (Oerlemans,
2005). In temperate very humid mountain ranges such as the
New Zealand alps and the Norwegian mountains, winter snow
accumulation and thus precipitation related to atmospheric circulation changes may play a major role (e.g. Rother and Shulmeister,
2006; Nesje et al., 2008; Anderson et al., 2010). Due to the
extremely high precipitation, glacier advances may thus also
largely be accumulation driven in southern Patagonia as suggested
already by Warren and Sudgen (1993) who also speculated that
ablation may play a major role on the drier eastern side of the
Andes. In the following, we summarise the available information on
Holocene glacier advances and discuss their forcing in the context
of the paleoclimate background data (i.e. paleotemperature and
paleoprecipitation reconstructions).
In Fig. 11F, we provide a compilation of Holocene glacier
advances in the southernmost Andes (e.g. Mercer, 1982; Aniya,
1995; McCulloch et al., 2000; Glasser et al., 2004; Wenzens,
2005; Masiokas et al., 2009a) and the Antarctic Peninsula (Hall,
2009) located at the southern margin of the SWW. Only few Early
Holocene advances (mostly around ca 8 ka BP) have so-far reconstructed from southern Patagonia, for example from the Lago
Argentino and Viedma area (Wenzens and Wenzens, 1998;
Wenzens, 1999, 2005). These studies are consistent with reconstructed glacier advances at the NPIF (Douglass et al., 2005) the
Antarctic Peninsula (Hall, 2009). However, the early Holocene
records are primarily based on moraine dating and subsequent
more extended late Holocene glacier advances might have overprinted or eroded earlier moraines. The only early Holocene glacier
advance record not based on moraine dating, is a glacier clay
deposition record from a fjord core north of Cordillera Darwin
(Fig. 11G) which reveals a minor advance at ca 8 ka BP (Boyd et al.,
2008).
The next dated glacier advances (from ca 4.9e5.4 ka BP) at the
beginning of the Neoglacial are recorded at several sites in the
southern Andes (Mercer, 1982; Aniya, 1995; Glasser et al., 2004;
Wenzens, 2005) and coincide with glacier advances on a global
scale (Porter, 2000; Magny and Haas, 2004) and increased ice
rafted detritus recorded in the South Atlantic (Hodell et al.,
2001). A subsequent minor advance between ca 3.9 and 4.3 ka
BP occurred at several locations in the southern Andes, but
probably not at Cordillera Darwin (Kuylenstierna et al., 1996).
However, an IRD peak in the South Atlantic also indicates
increased Antarctic drift ice during this period (Fig. 11H; Hodell
et al., 2001). At many sites at the eastern (Fig. 11F) and western
side of the Andes, glacier advances were recorded for the interval

15

between ca 2.0e2.7 ka BP (e.g. Mercer, 1982; Glasser et al., 2004).


This advance period appears also in the glacial clay deposition
record of the Cordillera Darwin (Boyd et al., 2008, Fig. 11G). At
some sites at the western side of the Andes further advances
occurred during the period from 0.9 to 1.2 ka BP which roughly
coincides with the global MCA (Masiokas et al., 2009a). At other
eastern sites, these moraines may have been overrun by more
extended advances during the global LIA. The LIA advance period
in southern Patagonia ranged from 0.6 to 0.05 ka BP. Most
moraines are well preserved at nearly all glacier systems. During
this period, many glaciers reached their maximum extension at
around 0.55 to 0.4 ka BP and afterwards experienced a successive
retreat until the last century (Koch and Kilian, 2005, Kilian et al.,
2007a; Masiokas et al., 2009a,b).
The few available Holocene temperature reconstructions (see
6.2.) suggest that the early Holocene from 11.5 to 8.5 ka BP was
characterised by 1.5e2  C higher temperatures compared to the
Late Holocene (e.g. Fig. 11B and C). Thus early Holocene advances
require signicantly higher precipitation than during the Middle
and Late Holocene. Such accumulation-driven advances, would
only be conceivable under very humid conditions and might
therefore support the western Andean paleoprecipitation reconstructions by Lamy et al. (2010) (Fig. 11D) compared with the
contrasting interpretations from the Torres del Paine region
(Fig. 11A; Moreno et al., 2010). Also the Neoglacial advances
appear to coincide with short-term comparatively warm and
humid late Holocene intervals as recorded in stalagmites (Schimpf
et al., 2011) superimposed on a long-term cooling (Fig. 11BeD).
Tree-ring data suggest substantially, ca 1e1.5  C colder temperatures around the LIA which might have forced southern Patagonian
LIA glacier advances in spite of probably more arid conditions
(Fig. 11E) (Schimpf et al., 2011).
6.5. High resolution records and event stratigraphy
High resolution records (1e10 yr resolution) have been obtained
from Laguna Potrok Aike (e.g. Habelzettl et al., 2007, 2009), Laguna
Vizcachas (Fey et al., 2009), Lake Tamar (Lamy et al., 2010), Lake
Hambre (Fig. 13; Hermanns and Biester, 2011), Lake Fagnano
(Waldmann et al., 2011) and from staglamites of the Marcelo Arevalo (MA) cave (e.g. MA1 stalagmite record; Schimpf et al., 2011).
These records may be able to depict single storm events (e.g. Lake
Tamar, Lamy et al., 2010) or can show mass ows related to earth
quakes which appeared at Lago Fagnano throughout the Holocene
within recurrence times of 0.2e1.0 ka (Waldmann et al., 2011).
Although all the above named records are situated at the southern
margin of the present-day core of the SWW (Figs. 1 and 4) at latitudes 52 30S to 54 S, correlations on a decadal to centennial scale
are often difcult. However, the precipitation-dependent iron
concentrations in the LF06-16 sediment core from Lago Fagnano
(Fig. 13E; Waldmann et al., 2010) and the drip rate/precipitationdependent d18O values as well as Y contents of the MA1 stalagmite (MA cave near Gran Campo Nevado) exhibit surprisingly
synchronous pattern on a centennial and decadal scale (Fig. 12FeG)
although all three proxies have largely unknown sensitivities and
threshold levels. A major difference between both records concerns
the phase between 3.5 and 2.5 ka BP for which a relatively week
SWW is indicated in the stalagmite record. At Lago Fagnano situated east of the climate divide a generally higher precipitation is
recorded during this period which can be explained by a higher
amount of easterly-derived precipitation due to a weaker SWW.
However, during the last 5 ka both records show seven pronounced
precipitation and SWW maxima with durations of 0.1e0.25 ka as
well as SWW minima (1e7a and b in Fig. 12E and F). The pattern
of both SWW-related records also show considerable correlations

16

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

with the total solar irradiance (Fig. 12A; Steinhilber et al., 2009;
Vieira et al., 2011), the latitudinal position and/or intensity of the
ITCZ (Fig. 12B; Haug et al., 2001) as well as the only two existing
continuous and highly resolved records of El Nio events represented by colour intensity variations in a sediment core from
Laguna Pallcacocha in Ecuador (Fig. 12C; Moy et al., 2002) and the
amount of lithic grains in a sediment core from the Peruvian coast
(Fig. 12D; Rein et al., 2005). Both cores show in common several
phases with an increased number of El Nio events during the
period between 3.5 and 2.5 ka BP as well as during several shorter
phases. In the Peruvian coastal record the MCA was identied as
a period of dominant La Nia situations while the Laguna Pallcacocha record shows increased El Nio events after around 1.2 ka BP.
However, the youngest age control point in the Pallcacocha record
is at 1.32 ka BP and low sedimentation rates during more frequent
La Nia could imply a later onset of this El Nio event period in this
record as shown in Fig. 12CeD.
6.6. Possible anthropogenic interferences in paleoclimate records
The possible human impact on ecosystems and paleoclimate
proxies is difcult to evaluate, because quantitative estimates of
the population and there regional distribution are not well known.
Nevertheless, we summarise in the following the available information on the inhabitation history and their possible relation to
paleoclimate. Thereafter, we discuss potential anthropogenic
perturbations in the paleoclimate records. The steppe region of
southernmost Patagonia was populated by terrestrial huntergatherer after ca 12.5 ka BP (e.g. Borrero, 1999; Legoupil, 2009;
Orquera et al., 2011) when most glacier systems of the Magellanes
region have been retreated nearly to present-day extension (e.g.
McCulloch and Bentley, 1998; Kilian et al., 2007a; Boyd et al.,
2008). These humans occupied also Tierra del Fuego soon after
12.0 ka BP when the lower global sea level allowed for a land
bridge in the sector of Primera and Segunda Angostura of the Strait
of Magellan (McCulloch and Morello, 2009; Morello et al., 2012,
Figs. 4 and 13G). This early Holocene period between 11.5 and
8.5 ka BP was probably very arid in the steppe region and partly at
the eastern slopes of the Andes, characterised by high charcoal
presence within regional pollen records (Fig. 13F; e.g. Heusser,
1995; McCulloch and Davis, 2001) (see also Chapters 6.2. and
6.3.). This arid climate (with strong wind-induced drying of plants)
abets the expansion of res, but a possibly stronger SWW
(following Lamy et al., 2010) (Fig. 13B) decreases the probability of
thunderstorms which could cause natural res. Heusser (1995)
suggested that at least some res of this period are related to
human activity. However, the causes for natural versus anthropogenic re-activity remain complex and therefore strongly
disputed, especially for open forest regions to the east of the Andes
(e.g. Heusser, 1994; Huber et al., 2004; Whitlock et al., 2007;
Markgraf and Huber, 2010), whereas charcoal is nearly absent in
records of the hyperhumid Patagonian rain forest (Fesq-Martin
et al., 2004).
The period from 8.5 to 5 ka BP was probably less arid to the east
of the Andes and pollen records from the Puerto del Hambre as well
as Estancia Esmeralda on the Northern Isla Dawson indicate a more
expanded and stable forest ecosystem with restricted re activity
(Figs. 7 and 13F; Heusser, 1995; McCulloch and Davis, 2001). Only
few remnants of inhabitants have been found on Tierra del Fuego
during this period. This may be related to its island location after
the formation of the Western Strait of Magellan at around 9 ka BP
caused by the global sea levels rise (Fig. 13G; McCulloch et al., 2005;
McCulloch and Morello, 2009). Pollen records from Puerto del
Hambre also indicate a stable forest system between 10.0 and
5.5 Ka BP (Fig. 13F). After around 6.5 ka BP, an increased number of

settling ages of maritime hunteregatherer were reported from


Andean fjords and the southern sector of the Strait of Magellan
Strait near Puerto el Hambre and Santa Ana (Prieto et al., 2009;
Torres, 2009; Orquera et al., 2011). This occupation period with
canoeing Indians may explain a renewed inhabitation of Tierra del
Fuego after 6.5 ka BP (Morello et al., 2012). However signicant
anthropogenic impacts on proxy records of this period are not
clearly documented.
A reduced and more open Nothofagus forest together with
more regular charcoal events occur during the last 5.5 ka BP in
the pollen record of Puerto del Hambre although palaeoclimatic
background data only show moderate climate variations during
this period (Fig. 13A and B) and other regional pollen records
from e.g. Punta Arenas, Haberton and Paso Giribaldi do not show
a signicant forest reduction (Fig. 7). Therefore it can be questioned whether these changes in the forest ecoton near to Puerto
del Hambre may reect anthropogenic perturbation. The Lake
Hambre sediment record indicates a pronounced increase of Late
Holocene erosion characterised by very high siliciclastic accumulation rates (high Y-content of the sediment; Fig. 13E) and very
high deposition rates of terrestrial plants (high C/N ratios;
Hermanns and Biester, 2011), in particular from 5.5 to 4.3 ka BP.
This coincides with higher charcoal contents in the Puerto del
Hambre record (Fig. 13F; Heusser, 1995) and frequent radiocarbon
ages of settling remnants in this area (Prieto et al., 2009; Torres,
2009; Orquera et al., 2011). This settling activity may have
provoked more frequent anthropogenic res with forest
destruction and related erosion.
7. Hemispheric/interhemispheric linkages and climate
forcing mechanisms
7.1. Insolation
Insolation changes are regarded as the major forcing factor
and pacemaker for glacial-interglacial changes in global climate
(e.g. Hays et al., 1976). In the traditional Milankovic theory,
primarily summer insolation at 65 N and its impact on Northern
Hemisphere glaciation has been involved for explaining global
ice-ages (Imbrie et al., 1992). For the Southern Hemisphere (SH)
where summer insolation changes are out-of-phase to the north,
changes in austral spring insolation have been recently proposed
to explain the warming during the Last Glacial termination (e.g.
Stott et al., 2007). Modelling studies suggest that these spring
insolation changes combined with variations in greenhouse gas
are sufcient to explain most of the SH deglacial warming and
the cooling during the course of the Holocene (Renssen et al.,
2005; Timmermann et al., 2009). Though small in amplitude,
changes in mean annual insolation may additionally play a role
in particular during the Holocene as suggested by climate
modelling studies (Liu et al., 2003). The effect on orbital-scale
insolation forcing on the SWW during the Holocene has also
been recently addressed in modelling studies (Wagner et al.,
2007; Varma et al., 2011). Wagner et al. (2007) suggest for
example enhanced austral summer and reduced austral winter
westerlies in the middle Holocene over southernmost Patagonia
induced by insolation-driven meridional changes in temperature
gradients.
During the Deglacial, the impact of insolation changes may be
partly obscured by the antiphased temperature excursions of the
ACR and Younger Dryas, probably related to the bipolar seesaw
effect (e.g. Stocker, 1998; Stenni et al., 2011; Chapter 7.3.). Holocene
SST records within the SWW and ACC belt off Chile, including the
southern Patagonian site MD07-3128 and the northern Patagonian
ODP Site 1233, are compared with orbital-scale insolation changes

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

in Fig. 14. These records show an Early Holocene thermal maximum


between 11.5 and 8.5 ka BP likewise known from Antarctica
(Masson et al., 2000) which may be related to the maximum in
average annual insolation at ca 10 ka BP (Fig. 14H). However, this
maximum represents only 0.4 W/m2 higher insolation and
modelling results suggest that annual insolation may account for
only 0.5  C early Holocene warming in the Southern Ocean (Liu
et al., 2003). Summer insolation reaches a pronounced maximum
during the Late Holocene, opposite to the general slight cooling in
the SST records whereas Antarctic temperature records show no
clear trends during the middle and late Holocene (Fig. 14). The
former considerations suggest that the link between proxy data and
insolation changes in the Southern Hemisphere is not
straightforward.

17

7.2. Sun activity


Total solar irradiance (TSI) changes have been recently reconstructed for the whole Holocene based on changes of the Earths
magnetic eld and the 14C (Vieira et al., 2011) as well as 10Be
behaviour (Steinhilber et al., 2009). The TSI variations are in the
range of 1 W/m2 which is higher than the above discussed
Holocene changes of the average annual insolation (0.4 W/m2).
Phases of low TSI have been related with Holocene North Atlantic
drift ice events (Bond et al., 2001) and also to global climate
perturbations (Wanner et al., 2008, 2011). In southern Patagonia
such probably sun-related temperature depressions occurred
especially during the period between from 3.5 to 2.5 ka BP, in
particular around 2.7 ka BP (Chambers et al., 2007; Van Geel et al.,
2000; Mhlinghaus et al., 2009). As discussed in Chapter 6.2., tree
ring-based temperature reconstructions (Neukom et al., 2010a, b)
and all SST records (Fig. 9) show a temperature depression of
1.0e1.5  C during the LIA period which follows in general the wellknown sun activity minima.
The evaluation of early and mid-Holocene inuences of the TSI
is more difcult. Fjord SST reconstruction have to be considered
with care during this time interval, since high melt water uxes
during spring and summer may have depressed the SSTs signicantly. Furthermore wind-induced changes in transport rates of
drift ice in the fjords after major calving events may have impacted
SSTs. Furthermore, most TSI changes appeared on a scale of some
decades to a century which is relatively short compared to the
time resolution of fjord sediment cores. Nevertheless, in Fig. 11 we
compare SWW-sensitive high resolution lake and stalagmite
records with the TSI values of the last 5 ka BP (Fig. 11). There is
a clear relationship between reconstructed SWW strength and the
TSI record. Since SWW changes are driven especially by latitudinal
SST gradients on the Pacic and vertical atmospheric temperature
gradients, we expect that the sun has a clear inuence on the
thermal structure of the Southern Hemisphere within a time
resolution of decades to centuries. This linkage is also conrmed
by sun related cyclicities (11, 22, 90 and 210 yrs) which have been
detected in tree ring records near the Pio XI glacier on the western
side of the southern Patagonian Ice Field (Rigozo et al., 2007) and
in a stalagmite record from the Andes at 53 S (Schimpf et al.,
2011).
7.3. Ocean currents

Fig. 14. Relationship between insolation and hemispheric temperatures: A: Summer


insolation at 50 N (Berger and Loutre, 1991). B: d18O record from the NGRIP ice-core as
indicator for the Northern Hemisphere temperature evolution (North Greenland Ice
Core Project Members, 2004); C: Alkenone SST record of ODP Site 1233 at the Pacic
continental slope at 41 S (Kaiser and Lamy, 2010); D: Alkenone SST record of core
MD07-2120 near New Zealand at 46 S (Pahnke and Sachs, 2006): E: Alkenone SST
record of core MD07-3128 (Caniupn et al., 2011); F: Diatom-based SST-record in
sediment core PS2102-2 from the South Atlantic at 53 S (Bianchi and Gersonde, 2004);
G: Summer insolation at 50 S (Berger and Loutre, 1991) and H: Average annual insolation at 50 S (Berger and Loutre, 1991).

The maritime climate of Southern Patagonia is strongly affected


by oceanographic changes especially in the Southern Pacic. The
directly off the Strait of Magellan located site MD07-3128 reveals
temperature changes that are largely following an Antarctic
timing as known from ice-core records (e.g. EPICA Community
Member, 2006) and other marine records offshore Chile (Lamy
et al., 2004; Kaiser et al., 2005) and New Zealand (Pahnke et al.,
2003). This includes millennial-scale temperature changes during
the Last Glacial as well as the timing and structure of the Deglacial
warming. Terrestrial records from southern Patagonia partly
conrm the Antarctic timing, e.g. through the existence of glacier
advances during the Antarctic Cold Reversal (e.g. Sugden et al.,
2005; Moreno et al., 2009b). In general, millennial-scale temperature changes in Antarctica over the Last Glacial may be consistently
explained by the bipolar seesaw concept that suggests an out-ofphase millennial-scale climate pattern between the Northern and
Southern Hemisphere during the Last Glacial (e.g. Stocker, 1998;
Stenni et al., 2011). Therefore, changes in ocean currents provide
a substantial forcing for southern Patagonia. Also for the early
Holocene warm period, large-scale global circulation changes
involving seesaw-like SST pattern have been proposed (e.g. Masson

18

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

et al., 2000) that contributed to the strengthening of the core SWW


in southern Patagonia (Lamy et al., 2010). During the late Holocene,
warming in the Eastern Tropical Pacic and cooling further south
possibly enhanced the mid- to low-latitude SST gradient resulting
in stronger winds at the northern margin of the SWW (Lamy et al.,
2010).
7.4. Atmospheric circulation
Atmospheric circulation changes within the SWW have been
involved as a key forcing mechanism both on glacialeinterglacial
time-scales as well as during the Glacial and Deglacial millennialscale climate events (e.g. Toggweiler et al., 2006; Lamy et al.,
2007; Toggweiler and Joellen Russell, 2008). Based on a simplied general circulation model, Toggweiler et al. (2006) suggested
that an equator-ward shifted SWW during the glacial allowed more
respired CO2 to accumulate in the deep ocean. During Glacial
terminations, the southward moving SWW reduced polar stratication and enhanced upwelling of deep water masses around
Antarctica that would then have released large amounts of the
stored CO2 to the atmosphere. Though the glacial northward shift of
the SWW is not clearly conrmed by more sophisticated model
studies (e.g. Rojas et al., 2009), it is consistent with paleoceanographic and palaeoclimatic reconstructions from the Southeast
Pacic (e.g. Heusser, 1989; Lamy et al., 1999, 2004, 2007; Moreno
et al., 2001). However, most studies of glacial westerlies were so
far based only on records from the northern margin of the SWW.
Based on opposing rainfall trends across the northern margin and
the core of the SWW belt during the Holocene, Lamy et al. (2010)
suggested that glacial westerlies may resemble more the present
winter situation with a northward expansion of westerly winds and
not necessarily a latitudinal shift of the wind belt as a whole.
Furthermore, the SWW plays a major role for coupling of both
hemispheres despite out-of-phase summer insolation signals. For
the last deglaciation, Denton et al. (2010) suggested that the
combined inuence of the oceanic bipolar seesaw and the southward displacement of the SWW allowed the high southern latitudes to warm as a result of melting and collapse of NH ice sheets
into the North Atlantic. These authors further proposed that the
ITCZ plays a major role in transferring the Northern Hemisphere
climate signals southward to the SWW.
Though it is generally assumed that the ITCZ was located further
south during Northern Hemisphere cold phases including the LGM,
a reconstruction of the Equatorial Front-ITCZ system off northwest
South America over glacial-interglacial cycles reveals a northward
ITCZ position (Rincn-Martnez et al., 2010). The authors suggest
that this northward position during the Last Glacial may have been
controlled by enhanced northward advection of cold water masses
with the Humboldt Current system and was restricted to the
eastern tropical Pacic. The same study likewise suggests
predominance of La Nia-like conditions during the Glacial,
whereas other authors propose that average conditions during Last
Glacial were more similar to modern El Nio years (e.g. Koutavas
et al., 2002). Based on the modern climatology, more La Nia-like
conditions during the Last Glacial would however imply enhanced
winds in the core SWW over Southern Patagonia and a reduction
north of ca 45 S (Schneider and Gies, 2004) which is not consistent
with the available paleodata (see 5.1.).
Modelling studies (e.g. Cane, 2005) suggest that the ITCZ was
located north of its present position and the frequency and strength
of El Nio events were strongly reduced during the early Holocene,
consistent with reconstructions based on paleo-data from the
eastern tropical Pacic and adjacent subtropical South America (e.g.
Moy et al., 2002). Assuming that reduced El Nio events are connected with more La Nia-like conditions this would imply

a stronger SWW over Southern Patagonia during the early Holocene (Schneider and Gies, 2004) consistent with enhanced rainfall
on the western hyperhumid side of the Andes at these latitudes
(Lamy et al., 2010). Since ca 8 ka BP, the number of El Nio events
recorded in Laguna Pallcacocha in Ecuador (Moy et al., 2002) largely
increased in particular during the past ca 3.5 ka. This middle and
late Holocene increase in El Nio strength and frequency has been
related to orbital-scale seasonal insolation changes close to the
Equator (Clement et al., 2000; Cane, 2005).
The frequency and amplitude of El Nio events also varied on
sub-orbital time-scales during the Holocene and were linked to
changes in the SWW strength and the TSI (see Chapter 7.2.). Seven
phases of increased and decreased strength of the SWW core have
been detected in high-resolution Patagonian stalagmite and lake
sediment records from latitudes 52 300 to 54 300 S. Phases of low
sun activity are related to low SWW strength and increased El
Nio situations and vice versa. Rein et al. (2005) suggested for
example a pronounced reduction of El Nio events during the
MCA consistent with higher rainfall and thus a stronger SWW as
recorded in both records (Fig. 11). A reduced precipitation and
SWW strength occurs especially between 2.5 and 3.5 ka BP and
from 0.7 to 0.1 ka BP. when increase El Nio events have been
recorded at the Peruvian coast (Rein et al., 2005) and in the
Laguna Pallcacocha record (Fig. 11), again consistent with the
modern El NioeSWW relationship in the Southern Patagonia
(Schneider and Gies, 2004).
Taken together, atmospheric circulation changes within the
SWW are the major forcing factor for paleoclimate changes in
Southern Patagonia. Past SWW changes are strongly linked to
changes in the tropical climate system especially latitudinal shifts
of the ITCZ and changes in the ENSO system. During the Holocene,
latitudinal shifts of the ITCZ and related SWW changes were
controlled by millennium-scale Northern Hemispheric insolation
changes as well as centennial scale changes of sun activity (TSI).
8. Conclusions
The southern tip of South America represents a key area for the
understanding of global paleoclimate and therefore more than 150
paleoclimate reconstructions have been published only during the
last 5 years. These records give new implications concerning
paleotemperature as well as changes in SSW strength and/or latitudinal shifts. However, when comparing these records many
inconsistencies or partly opposite interpretations appear which are
related to restricted proxy understanding and/or unavailable calibration. This concerns also the individual sensitivity and threshold
of each proxy with respect to climate change and an often nonlinear proxy-climate relationship. Some proxies are highly sensitive to even short term changes or occasional climate events
whereas others have high natural variability often unrelated to
climate or do not reach a threshold for registering paleoclimate
uctuations of a given amplitude.
The few existing full glacial records indicate that the Patagonian
Ice Sheet experienced signicant oscillations between 60 and 12 ka
BP which have been primarily controlled by paleotemperatures and
the intensity and latitudinal position of the SWW. The paleotemperatures of this period are clearly related to those reconstructed from Antarctic Ice cores. The Deglacial also follows
Antarctic pattern with an Antarctic Cold Reversal and no signicant
cooling during the Younger Dryas. During the Holocene most
glacier advances occurred after 5.5 ka BP and were probably
primarily driven by short term (a few hundreds of years) increased
SWW strength and related increased precipitation/accumulation.
Only the LIA advances likely controlled by reduced ablation and
thus temperature.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

SSW air masses circulate within few weeks around the globe
and have a clear oceanic control. Therefore Holocene temperatures
are best represented by oceanic SST records which show similar
overall pattern at various sites within the northern ACC and SWW
belt. These are characterised by an Early Holocene thermal
maximum (11.5e8.5 ka BP) and relatively constant 1.5  C lower
temperatures during the last 8 ka BP. Therefore regional terrestrial
temperature anomalies in southernmost Patagonia may be very
weak. Southern Hemisphere insolation has only a minor or no
inuence on this pattern, but periods of low sun activity between
3.5 and 2.5 ka BP and especially the LIA are characterised by
a signicant drop in temperature up to 1.5  C in the fjord SST
records and in terrestrial paleotemperature reconstructions.
Different reconstructions of the SWW strength and related
precipitation show an inconsistent or partly opposite behaviour.
Some records indicate a weak SWW core in the early Holocene and
a strengthening towards the Late Holocene, whereas others show
the reverse pattern or do not reveal a clear trend. These inconsistencies may partly be related to unclear relations between precipitation and SWW strength. Furthermore, restricted knowledge of
the seasonally variable wind- and temperature-related effect of
evaporation on the available moisture in soils and/or the general
water balance might have let to conicting results. Resolving these
inconsistencies in southernmost Patagonian SWW records is
a prerequisite for improving hemispheric comparisons with e.g.
New Zealand records (e.g. Fletcher and Moreno, 2012) and links to
atmospheric CO2 changes (Hodgson and Sime, 2010).
For the last 5 ka BP high resolution records show clear linkages
on a decadal and centennial scale between SWW strength (at the
southern margin of its present-day core at latitudes 52 300 to
54 300 S) and the behaviour of ENSO, the ITCZ and sun activity. Low
sun activity is related to weak and/or northward displaced SWW
and to more frequent El Nio events as well as a southward displaced and/or weaker ITCZ.
A denser grid of sediment core records is required to better
cover the whole precipitation gradient across the Andean climate
divide in southernmost Patagonia. In this respect systematic transects across the climate gradient with comparable archives and
proxies would be desirable. These studies should be accompanied
by more extensive long-term climate and ocean monitoring,
particularly in the western fjord region.
Acknowledgements
We thank Helge Arz, Rene Garreaud, Jerme Kaiser, Christopher
Moy, Michael Mayr, Nicolas Waldmann, Luis Vieira, and Rodrigo
Villa-Martnez for sharing data and discussions. Jean Pierre Francois
provided very valuable comments on pollen issues. Oscar Baeza and
Francisco Rios gave important technical support. We further
acknowledge the constructive reviews by Robert McCulloch and an
anonymous reviewer that helped to improve the manuscript. The
German Research Foundation is thanked for research grants Ki 456/
8 to Ki 456/12, LA 1273/3-2, LA1273/5-1, and LA1273/7-1.
References
Ackert Jr., R.P., Becker, R.A., Singer, B.S., Kurz, M.D., Caffee, M.W., Mickelson, D.M.,
2008. Patagonian glacier response during the Late GlacialeHolocene transition.
Science 321 (5887), 392e395.
Anderson, D.M., Archer, R.B., 1999. Preliminary evidence of early deglaciation in
southern Chile. Palaeogeography, Palaeoclimatology, Palaeoecology 146 (1e4),
295e301.
Anderson, B., MacKintosh, A., Stumm, D., George, L., Kerr, T., Winter-Billington, A.,
Fitzsimons, S., 2010. Climate sensitivity of a high-precipitation glacier in New
Zealand. Journal of Glaciology 56, 114e128.
Aniya, M., 1995. Holocene glacial chronology in Patagonia -Tyndall and Upsala
glaciers. Arctic and Alpine Research 27 (4), 311e322.

19

Appleby, P.G., Oldeld, F., 1992. Application of lead-210 to sedimentation studies.


Uranium series disequilibrium: applications to environmental problems. In:
Ivanovich, M., Harmon, R.S. (Eds.), Uranium Series Disequilibrium: Applications
to Environmental Problems. Clarendon Press, Oxford, pp. 731e778.
Aravena, J.-C., 2007. Reconstructing climate variability using tree rings and glacier
uctuations in southernmost Chilean Andes. PhD thesis, University of Western
Ontario, London, Ontario, Canada.
Aravena, J.C., Luckman, B.H., 2009. Spatio-temporal rainfall patterns in southern
south America. International Journal of Climatology 29, 2106e2120.
Aravena, J.C., Lara, A., Wolodarsky-Franke, A., Villalba, R., Cuq, E., 2002. Tree-ring
groth patterns and temeparture reconstruction from Nathofagus pumilio
(Fagaceae) en el lmete arbores superior de la Patagonia austral de Chile. Revista
Chilena de Historia Natural 75, 361e375.
Ariztegui, D., Gilli, A., Anselmetti, F.S., Goni, R.A., Belardi, J.B., Espinosa, S., 2010.
Lake-level changes in central Patagonia (Argentina): crossing environmental
thresholds for Lateglacial and Holocene human occupation. Journal of Quaternary Science 25 (7), 1092e1099.
Armesto, J.J., Casassa, I., Dollenz, O., 1992. Age structure and dynamics of Patagonian
beech forests in Torres del Paine National Park, Chile. Vegetatio 98, 13e22.
Ashworth, A.C., Markgraf, V., Villagran, C.,1991. Late Quaternary climatic history of the
Chilean Channels based on fossil pollen and beetle analysis, with an analysis of
the modern vegetation and pollen rain. Journal of Quaternary Science 6, 279e291.
Auer, V., 1933. Verschiebungen der Wald-und Steppengebiete Feuerlands in postglazialer Zeit. Acta Geographica 5, 1e313.
Auer, V., 1958. The Pleistocene of Fuego-Patagonia. Part II. The history of the ora
and vegetation. Annales Academiae Scientiarum Fennicae III. Geologica-Geographica 50, 1e239.
Auer, V., 1960. The Quaternary history of Fuego-Patagonia. Proceedings of the Royal
Society of London. Series B, Biological Sciences. A Discussion on the Biology of
the Southern Cold Temperate Zone 152 (949), 507e516.
Auer, V., 1974. The isorhythmicity subsequent to Fuego-Patagonian and Fennoscandian ocean level transgressions and regressions of the latest glaciation.
Annales Academiae Scientiarum Fennicae III. Geologica-Geographica 115, 1e88.
Baeza, O., 2005. Lake and fjord sediments as Late Glacial to Holocene environmental
and climate archives of the southernmost Andes at 53 S, Chile. Ph.D. thesis,
Albert-Ludwigs-Universitt Freiburg, Germany.
Ballantyne, C.K., 2002. Paraglacial geomorphology. Quaternary Science Reviews 21,
1935e2017.
Barker, S., Diz, P., Vautravers, M.J., Pike, J., Knorr, G., Hall, I.R., Broecker, W.S., 2009.
Interhemispheric Atlantic seesaw response during the last deglaciation. Nature
457 (7233), 1097e1102.
Bennett, K.D., Haberle, S.G., Lumley, S.H., 2000. The Last GlacialeHolocene transition in southern Chile. Science 290, 325e328.
Berger, A., Loutre, M.F., 1991. Insolation values for the climate of the last 10 million
years. Quaternary Science Reviews 10, 297e317.
Bianchi, C., Gersonde, R., 2004. Climate evolution at the last deglaciation: the role
of the southern ocean. Earth and Planetary Science Letters 228 (3e4),
407e424.
Biester, H., Kilian, R., Franzen, C., Woda, C., Mangini, A., Schler, H.F., 2002.
Elevated mercury accumulation in a peat bog of the Magellanic Moorlands,
Chile (53 S) e an anthropogenic signal from the Southern Hemisphere. Earth
and Planetary Science Letters 201 (3e4), 609e620.
Biester, H., Martinez-Cortizas, A., Birkenstock, S., Kilian, R., 2003. Effect of peat
decomposition and mass loss on historic mercury records in peat bogs from
Patagonia. Environmental Science Technology 37 (1), 32e39.
Biester, H., Selimovc, D., Hemmerich, S., Petri, M., 2006. Halogens in pore water of
peat bogs e the role of peat decomposition and dissolved organic matter.
Biogeosciences 3, 53e64.
Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M.N., Showers, W., Hoffmann, S.,
Rusty Lotti-Bond, R., Hajdas, I., Bonani, G., 2001. Persistent solar inuence on
north Atlantic climate during the Holocene. Science 294 (5549), 2130e2136.
Borrero, L.A., 1999. Human dispersal and climatic conditions during Late Pleistocene
times in Fuego- Patagonia. Quaternary International 53/54, 93e99.
Boucher, ., Guiot, J., Chapron, E., 2011. A millennial multi-proxy reconstruction of
summer PDSI for Southern South America. Climate of the Past 7, 957e974.
http://dx.doi.org/10.5194/cp-7-957-2011.
Boyd, B., Anderson, J., Wellner, J., Fernndez, R., 2008. The sedimentary record of
glacial retreat, Marinelli Fjord, Patagonia: regional correlations and climate ties.
Marine Geology 255 (3e4), 165e178.
Brambati, 2000. Paleoclimatic and Paleoenvironmental Records in Sediments from
the Southern Ocean (Strait of Magellan and Ross Sea), Terra Antarctica Reports
4, pp. 1e41.
Breuer, S., Kilian, R., Baeza, O., Lamy, F., Arz H. Holocene denudation rates from the
superhumid southernmost Patagonian Andes (53 S) deduced from lake sediment budgets. Geomorphology, in press-a.
Breuer, S., Kilian, R., Weinrebe, W. Schrner, D., Behrmann, J. Glacier and tectonic
control on fjord morphology and sediment deposition in the Magellan region
(53 S). Marine Geology, in press-b.
Caldenius, C., 1932. Las glaciaciones cuaternarias de la Patagonia y Tierra del Fuego.
Geograska Annaler 14, 1e164.
Cane, M.A., 2005. The evolution of El Nio, past and future. Earth and Planetary
Science Letters 230 (3e4), 227e240.
Caniupn, M., Lamy, F., Lange, C.B., Kaiser, J., Arz, H.W., Kilian, R., Baeza-Urrea, O.,
Aracena, C., Hebbeln, D., Kissel, C., Laj, C., Mollenhauer, G., Tiedemann, R., 2011.
Millennial-scale sea surface temperature and Patagonian Ice Sheet changes off

20

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

southernmost Chile (53 S) over the past w60 kyr. Paleoceanography 26


(PA3221), 1e10.
Carrasco, J.F., Osorio, R., Casassa, G., 2008. Secular trend of the equilibrium-line
altitude on the western side of the southern Andes, derived from radiosonde
and surface observations. Journal of Glaciology 54 (186), 538e550.
Casassa, G., Rivera, A., Aniya, M., Naruse, R., 2000. Caractersticas glaciolgicas del
Campo de Hielo Patagnico Sur. Anales del Instituto de la Patagonia, Serie
Ciencias Naturales 28, 5e22.
Chaigneau, A., Pizarro, O., 2005. Surface circulation and fronts of the South Pacic
Ocean, east of 120 W. Geophysical Research Letters 32 (L08605).
Chambers, F., Mauquoy, D., Brain, S., Blaauw, M., Daniell, J., 2007. Globally
synchronous climate change 2800 years ago: proxy data from peat in South
America. Earth and Planetary Science Letters 253 (3e4), 439e444.
Clapperton, C.M., Sugden, D.E., Kauffman, D., McCulloch, R.D., 1995. The last glaciation in central Magellan Strait, southernmost Chile. Quaternary Research 44,
133e148.
Clement, A.C., Seager, R., Cane, M.A., 2000. Suppression of El Nio during the
Mid-Holocene by changes in the Earths orbit. Paleoceanography 15 (6),
731e737.
Cusminsky, G., Schwalb, A., Prez, A.P., Pineda, D., Viehberg, F., Whattley, R.,
Markgraf, V., Gilli, A., Aritztegui, D., Anselmetti, F.S., 2011. Late Quarternary
environmental changes in Patagonia as inferred from lacustrine fossil and
extant ostracods. Biological Journal of the Linnean Society 103, 397e408.
Denton, G.H., Anderson, R.F., Toggweiler, J.R., Edwards, R.L., Schaefer, J.M.,
Putnam, A.E., 2010. The Last Glacial termination. Science 328 (5986),
1652e1656.
Douglass, D.C., Singer, B.S., Kaplan, M.R., Ackert, R.P., Mickelson, D.M., Caffee, M.W.,
2005. Evidence of early Holocene glacial advances in southern South America
from cosmogenic surface-exposure dating. Geology 33, 237e240.
Endlicher, W., 1991. Patagonien e klima- und agrarkologische Probleme an der
Magellanstrae. Geographische Rundschau 43, 143e151.
Endlicher, W., Santana, A., 1988. El clima del sur de la Patagonia y sus aspectos
ecolgicos. Un siglo de Mediciones Climatolgicas en Punta Arenas. Anales
Instituto Patagonia, Serie Ciencias Naturales 18, 57e86.
EPICA community members, 2006. One-to-one coupling of glacial climate variability in Greenland and Antarctica. Nature 444, 195e198.
Fesq-Martin, M., Friedmann, A., Peters, M., Behrmann, J., Kilian, R., 2004. Late-glacial
and Holocene vegetation history of the Magellanic rain forest in southwestern
Patagonia, Chile. Vegetation History and Archaeobotany 13 (4), 249e255.
Fey, M., Korr, C., Maidana, N.L., Carrevedo, M.L., Corbello, H., Dietrich, S.,
Haberzettl, T., Kuhn, G., Lcke, A., Mayr, C., Ohlendorf, C., Paez, M.A.,
Quintana, F.A., Schbitz, F., Zolitschka, B., 2009. Paleoenvironmental changes
during the last 1600 years inferred from the sediment record of a cirque lake in
Southern Patagonia (Laguna Viscachas, Argentina. Paleogeography, Paleoclimatology, Paleoecology 281 (3e4), 363e375.
Fischer, H., Fundel, F., Huybrechts, P., Rybak, O., Oerter, H., Twarloh, B., Wegener, A.,
Ruth, U., Udisti, R., Becagli, S., Castellano, E., Fattori, I., Severi, M., Wolff, E.,
Littot, G., Rthlisberger, R., Mulvaney, R., Hutterli, M., Kaufmann, P., Federer, U.,
Lambert, F., Bigler, M., Stauffer, B., Hansson, M., Jonsell, U., de Angelis, M.,
Gabrielli, P., Boutron, C., Siggaard-Andersen, M.-L., Steffensen, J.P., Barbante, C.,
Gaspari, V., Wagenbach, D., 2007. Reconstruction of millennial changes in dust
emission, transport and regional sea ice coverage using the deep EPICA ice cores
from the Atlantic and Indian Ocean sector of Antarctica. Earth and Planetary
Science Letters 260, 340e354.
Fletcher, M.S., Moreno, P.I., 2011. Zonally symmetric changes in the strength and
position of the Southern Westerlies drove atmospheric CO2 variations over the
past 14 k.y. Geology 39 (5), 419e422.
Fletcher, M.S., Moreno, P.I., 2012. Have the Southern Westerlies changed in a zonally
symmetric manner over the last 14,000 years? A hemisphere-wide take on
a controversial problem. Quaternary International 253, 32e46.
Garreaud, R., 2007. Precipitation and circulation covariability in the extratropics.
Journal of Climate 20 (18), 4789e4797.
Garreaud, R., Lopez, P., Minvielle, M., Rojas, M. Large scale control on the Patagonia
Climate. Journal of Climate, in press.
Gilli, A., Ariztegui, D., Anselmetti, F., Mckenzie, J., Markgraf, V., Hajdas, I.,
Mcculloch, R., 2005. Mid-Holocene strengthening of the Southern Westerlies in
South America e Sedimentological evidences from Lago Cardiel, Argentina
(49 S). Global and Planetary Change 49 (1e2), 75e93.
Glasser, N.F., Harrison, S., Winchester, V., Aniya, M., 2004. Late Pleistocene and
Holocene palaeoclimate and glacier uctuations in Patagonia. Global and
Planetary Change 43 (1e2), 79e101.
Glasser, N.F., Harrison, S., Jansson, K.N., Anderson, K., Cowley, A., 2011. Global sealevel contribution from the Patagonia Iceelds since the Little Ice Age
maximum. Nature Geoscience 4 (5), 303e307.
Habelzettl, T., Corbella, H., Fey, M., Janssen, S., Lcke, A., Mayr, C., Ohlendorf, C.,
Schbitz, F., Schleser, G.H., Wille, M., Wulf, S., Zolitschka, B., 2007. Late glacial
and Holocene wetedry cycles in southern Patagonia: chronology, sedimentology and geochemistry of a lacustrine record from Laguna Potrok Aike,
Argentina. The Holocene 17 (3), 297e310.
Haberzettl, T., Anselmetti, F.S., Bowen, S.W., Fey, M., Mayr, C., Zolitschka, B.,
Ariztegui, D., Mauz, B., Ohlendorf, C., Kastner, S., Lcke, A., Schbitz, F., Wille, M.,
2009. Late Pleistocene dust deposition in the Patagonian steppe e extending
and rening the paleoenvironmental and tephrochronological record from
Laguna Potrok Aike back to 55 ka. Quaternary Science Reviews 28 (25e26),
2927e2939.

Hall, B.L., 2009. Holocene glacial history of Antarctica and the sub-Antarctic islands.
Quaternary Science Reviews 28, 2213e2230.
Haug, G.H., Hughen, K.A., Sigman, D.M., Peterson, L.C., Rhl, U., 2001. Southward
migration of the Intertropical Convergence Zone through the Holocene. Science
293, 1304e1308.
Hays, J.D., Imbrie, J., Shackleton, N.J., 1976. Variations in the Earths orbit: pacemaker
of the Ice Ages. Science 194 (4270), 1121e1132.
Hermanns, Y.-M., Biester, H., 2011. A Holocene record of mercury accumulation in
a pristine lake in Sothernmost South America (53 S) e climatic and environmental drivers. Biogeosciences Discussions 8 (4), 6555e6588.
Heusser, C.J., 1971. Pollen and Spores of Chile, Modern Types of the Pteridophyta,
Gymnospermae, and Angiospermae. The University of Arizona Press, USA,
176 pp.
Heusser, C.J., 1989. Late Quaternary vegetation and climate of southern Tierra del
Fuego. Quaternary Research 31 (3), 396e406.
Heusser, C.J., 1994. Paleoindians and re during the late Quaternary in southern
South America. Revista Chilena de Historia Natural 67 (4), 435e443.
Heusser, C.J., 1995. Three Late Quaternary pollen diagrams from southern Patagonia
and their paleoecological implications. Palaeogeography, Palaeoclimatology,
Palaeoecology 118 (1e2), 1e24.
Heusser, C.J., 2003. Ice Age Southern Andes e a Chronicle of Paleoecological Events.
In: Developments in Quaternary Science, vol. 3. Elsevier, Amsterdam.
Heusser, C., Heusser, E., Lowell, T., Moreira, A., Moreira, S., 2000. Deglacial palaeoclimate at Puerto del Hambre, subantarctic Patagonia, Chile. Journal of
Quaternary Science 15 (2), 101e114.
Hodell, D.A., Kanfoush, S.L., Shemesh, A., Crosta, X., Charles, C.D., Guilderson, T.P.,
2001. Abrupt cooling of Antarctic surface waters and sea ice expansion in the
South Atlantic sector of the Southern Ocean at 5000 cal yr B.P. Quaternary
Research 56, 191e198.
Hodgson, A., Sime, L.C., 2010. Southern westerlies and CO2. Nature Geoscience 3,
66e667.
Huber, U.M., Markgraf, V., Schbitz, F., 2004. Geographical and temporal trends in
lLate Quaternary pre histories of Fuego-Patagonia, South America. Quaternary
Science Reviews 23 (9e10), 1079e1097.
Hulton, N.R.J., Purves, R.S., Mcculloch, R.D., Sugden, D.E., Bentley, M.J., 2002. The
Last Glacial maximum and deglaciation in southern South America. Quaternary
Science Reviews 21 (1e3), 233e241.
Imbrie, J., Boyle, E.A., Clemens, S.C., Duffy, A., Howard, W.R., Kukla, G., Kutzbach, J.,
Martinson, D.G., McIntyre, A., Mix, A.C., Molno, B., Morley, J.J., Peterson, L.C.,
Pisias, N.G., Prell, W.L., Raymo, M.E., Shackleton, N.J., Toggweilerl, J.R., 1992. On
the structure and origin of major glaciation cycles: 1. Linear responses to
Milankovic forcing. Paleoceanography 7 (6), 701e738.
Ingram, B.L., Southon, J.R., 1996. Reservoir ages in eastern Pacic coastal and
estuarine waters. Radiocarbon 38 (3), 573e582.
Jouzel, J., Masson-Delmotte, V., Cattani, O., Dreyfus, G., Falourd, S.,
Hoffmann, G., Minster, B., Nouet, J., Barnola, J.M., Chappellaz, J., Fischer, H.,
Gallet, J.C., Johnsen, S., Leuenberger, M., Loulergue, L., Luethi, D., Oerter, H.,
Parrenin, F., Raisbeck, G., Raynaud, D., Schilt, A., Schwander, J., Selmo, E.,
Souchez, R., Spahni, R., Stauffer, B., Steffensen, J.P., Stenni, B., Stocker, T.F.,
Tison, J.L., Werner, M., Wolff, E.W., 2007. Orbital and millennial Antarctic
climate variability over the past 800,000 years. Science 317 (5839),
793e796.
Kaiser, J., Lamy, F., 2010. Links between Patagonian Ice Sheet uctuations and
Antarctic dust variability during the Last Glacial period (MIS 4-2). Quaternary
Science Reviews 29 (11e12), 1464e1471.
Kaiser, J., Lamy, F., Hebbeln, D., 2005. A 70-kyr sea surface temperature record off
southern Chile (Ocean Drilling Program Site 1233). Paleoceanography 20,
PA4009. http://dx.doi.org/10.1029/2005PA001146.
Kalnay, E., Kanamitsu, M., Kistler, R., Collins, W., Deaven, D., Gandin, L., Iredell, M.,
Saha, S., White, G., Woollen, J., Zhu, Y., Leetmaa, A., Reynolds, R., Chelliah, M.,
Ebisuzaki, W., Higgins, W., Janowiak, J., Mo, K.C., Ropelewski, C., Wang, J.,
Jenne, R., Joseph, D., 1996. The NCEP/NCAR 40-year reanalysis project. Bulletin
of the American Meteorological Society 77 (3), 437e470.
Kaplan, M.R., Fogwill, C.J., Sugden, D.E., Hulton, N.R.J., Kubik, P.W., Freeman, S.P.H.T.,
2008. Southern Patagonian glacial chronology for the Last Glacial period and
implications for Southern Ocean climate. Quaternary Science Reviews 27 (3e4),
284e294.
Kaplan, M.R., Strelin, J.A., Schaefer, J.M., Denton, G.H., Finkel, R.C., Schwartz, R.,
Putnam, A.E., Vandergoes, M.J., Goehring, B.M., Travis, S.G., 2011. In-situ cosmogenic 10Be production rate at Lago Argentino, Patagonia: implications for lateglacial climate chronology. Earth and Planetary Science Letters 309 (1e2), 21e32.
Kastner, S., Ohlendorf, C., Haberzettl, T., Lcke, A., Mayr, C., Maidana, N.I.,
Schbitz, F., Zolitschka, B., 2010. Southern hemispheric westerlies control
spatial distribution of modern sediments in Laguna Potrok Aike, Argentina.
Journal of Paleolimnology 44, 887e902.
Kilian, R., Hohner, M., Biester, H., Wallrabe-Adams, H., Stern, Ch, 2003. Holocene
peat and lake sediment tephra record from the southernmost Chilean Andes
(53-55 S). Revista Geolgica de Chile 30 (1), 47e64.
Kilian, R., Biester, H., Behrmann, J., Baeza, O., Fesq-Martin, M., Hohner, M.,
Schimpf, D., Friedmann, A., Mangini, A., 2006. Millennium-scale volcanic impact
on a superhumid and pristine ecosystem. Geology 34 (8), 609e612.
Kilian, R., Baeza, O., Steinke, T., Arvalo, M., Ros, C., Schneider, C., 2007a. Late
Pleistocene to Holocene marine transgression and thermohaline control on
sediment transport in the western Magellanes fjord system of Chile (53 S).
Quaternary International 161 (1), 90e107.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23


Kilian, R., Schneider, C., Koch, J., Fesq-Martin, M., Biester, H., Casassa, G., Arvalo, M.,
Wendt, G., Baeza, O., Behrmann, J., 2007b. Palaeoecological constraint on Late
Glacial and Holocene ice retreat in the Southern Andes (53 S). Global and
Planetary Change 59 (1e4), 49e66.
Kissel, C., Leau, H., The shipboard Scientic party, 2007. MD159-PACHIDERME
(IMAGES XV) Cruise Report. Ref OCE2007/0, 2007. In: Srie Les rapports de
campagnes la Mer, IPEV.
Kleinebecker, T., Hlzel, N., Vogel, A., 2007. Gradients of continentallty and moisture
in South Patagonian ombrotrophic peatland vegetation. Folia Geobotanica 42,
363e382.
Koch, J., Kilian, R., 2002. Dendroecological potencial of common tree species along
a transect across the southernmost Andes, Chile (53 ). Anales Instituto Patagonia, Serie Ciencias Naturales 30, 123e132.
Koch, J., Kilian, R., 2005. Little Ice Age glacier uctuations, Gran Campo Nevado,
southernmost Chile. The Holocene 15 (1), 20e28.
Koutavas, A., Lynch-Stieglitz, J., Marchitto, T.M., Sachs, J., 2002. El Nio-like pattern
in Ice Age tropical Pacic sea surface temperature. Science 297, 226e230.
Kuylenstierna, J.L., Rosqvist, G.C., Holmlund, P., 1996. Late-Holocene glacier variations in the Cordillera Darwin, Tierra del Fuego. Holocene 6 (3), 353e358.
Lamy, F., Hebbeln, D., Wefer, G., 1999. Highz resolution marine record of climate
change in mid-latitude Chile during the last 28,000 years based on terrigenous
sediment vparameters. Quaternary Research 51, 83e93.
Lamy, F., Rhlemann, C., Dierk Hebbeln, D., Wefer, G., 2002. High and low latitude
climate control on the position of the southern Peru-Chile Current during the
Holocene. Paleoceanography 17 (2). http://dx.doi.org/10.1029/2001PA000727.
Lamy, F., Kaiser, J., Arz, H., Ninnemann, U., Hebbeln, D., 2004. Antarctic timing of
surface water changes off Chile and Patagonian Ice Sheet response. Science 304
(5679), 1959e1962.
Lamy, F., Kaiser, J., Arz, H., Ninnemann, U., Hebbeln, D., Timm, O., Timmermann, A.,
Toggweiler, J., 2007. Modulation of the bipolar seesaw in the southeast pacic
during Termination 1. Earth and Planetary Science Letters 259 (3e4), 400e413.
Lamy, F., Kilian, R., Arz, H., Francois, J., Kaiser, J., Prange, M., Steinke, T., 2010.
Holocene changes in the position and intensity of the southern westerly wind
belt. Nature Geoscience 3 (10), 695e699.
Legoupil, D., 2009. Cerro Castillo rockshelter occupation: a residential site in
a panoramic viewing-point. Magallania 37 (1), 47e60.
Lemieux-Dudon, B., Blayo, E., Petit, J.-R., Waelbroeck, C., Svensson, A., Ritz, C.,
Barnola, J.-M., Narcisi, B.M., Parrenin, F., 2010. Consistent dating for Antarctic
and Greenland ice cores. Quaternary Science Reviews 29 (1e2), 8e20.
Lis-Pronovost, Agathe, St-Onge, G., Gogorza, C., Haberzettl, T., Preda, M., Kliem, P.,
Francus, P., Zolitschka, B., The PASADO Science Team, 2012. High-resolution
paleomagnetic secular variations and relative paleointensity since the Late
Pleistocene in southern South America. Quaternary Science Reviews. http://
dx.doi.org/10.1016/j.quascirev.2012.05.012.
Liu, Z., Brady, E., Lynch-Stieglitz, J., 2003. Global ocean response to orbital forcing in
the Holocene. Paleoceanography 18 (2), 1041. http://dx.doi.org/10.1029/
2002PA000819.
Magny, M., Haas, J.N., 2004. A major widespread climatic change around 5300 cal.
yr BP at the time of the Alpine Iceman. Journal of Quaternary Science 19 (5),
423e430.
Mancini, M.V., 2009. Holocene vegetation and climate changes from a peat pollen
record of the forest-steppe ecotone, Southwest of Patagonia (Argentina).
Quaternary Science Reviews 28 (15e16), 1490e1497.
Mancini, M., Paz, M., Prieto, A., Stutz, S., Tonello, M., Vilanova, I., 2005. MidHolocene climatic variability reconstruction from pollen records (32e52 S,
Argentina. Quaternary International 132 (1), 47e59.
Marinoni, L., Setti, M., Salvi, C., Lpez-Galindo, A., 2008. Clay minerals in late
Quaternary sediments from the south Chilean margin as indicators of provenance and palaeoclimate. Clay Minerals 43, 235e253.
Markgraf, V., 1993. Paleoenvironments and paleoclimates in Tierra del Fuego and
southernmost Patagonia, South America. Palaeogeography, Palaeoclimatology,
Palaeoecology 102 (1e2), 53e68.
Markgraf, V., Huber, U.M., 2010. Late and postglacial vegetation and re history in
Southern Patagonia and Tierra del Fuego. Palaeogeography, Palaeoclimatology,
Palaeoecology 297 (2), 351e366.
Markgraf, V., DAntoni, H., Ager, T., 1981. Modern pollen dispersal in Argentina.
Palynology 5, 235e254.
Markgraf, V., Bradbury, J.P., Schwalb, A., Burns, S.J., Stern, C., Ariztegui, D., Gilli, A.,
Anselmetti, F.S., Stine, S., Maidana, N., 2003. Holocene palaeoclimates of
southern Patagonia: limnological and environmental history of Lago Cardiel,
Argentina. The Holocene 13 (4), 581e591.
Marshall, J., Speer, K., 2012. Closure of the meridional overturning circulation
through Southern Ocean upwelling. Nature Geoscience. http://dx.doi.org/
10.1038/ngeo1391.
Masiokas, M.H., Rivera, A., Espizua, L.E., Villalba, R., Delgado, S., Aravena, J.C., 2009a.
Glacier uctuations in extratropical South America during the past 1000 years.
Palaeogeography, Palaeoclimatology, Palaeoecology 281 (3e4), 242e268.
Masiokas, M., Luckman, B., Villalba, R., Delgado, S., Skvarca, P., Ripalta, A., 2009b.
Little Ice Age uctuations of small glaciers in Monte Fitz Roy and Lago del
Desierto areas, south Patagonian Andes, Argentina. Palaeogeography, Palaeoclimatology, Palaeoecology 281, 351e362.
Masson, V., Vimeux, F., Jouzel, J., Morgan, V., Delmotte, M., Ciais, P., Hammer, C.,
Johnsen, S., Lipenkov, V.Y., Mosley-Thompson, E., Petit, J.-R., Steig, E.J.,
Stievenard, M., Vaikmae, R., 2000. Holocene climate variability in Antarctica
based on 11 ice-core isotopic records. Quaternary Research 54 (3), 348e358.

21

} cke, A., Ohlendorf, C.,


Mayr, C., Wille, M., Haberzettl, T., Fey, M., Janssen, S., Lu
Oliva, G., Schbitz, F., Schleser, G.H., Zolitschka, B., 2007a. Holocene variability of
the southern hemisphere westerlies in Argentinean Patagonia (52 S). Quaternary Science Reviews 26 (5e6), 579e584.
Mayr, C., Lcke, A., Stichler, W., Trimborn, P., Ercolano, B., Oliva, G., Ohlendorf, C.,
Soto, J., Fey, M., Haberzettl, T., 2007b. Precipitation origin and evaporation of
lakes in semi-arid Patagonia (Argentina) inferred from stable isotopes (d18O,
d2H). Journal of Hydrology 334, 53e63.
Mayr, C., Lcke, A., Maidana, N.I., Wille, M., Haberzettl, T., Corbella, H., Ohlendorf, C.,
Schbitz, F., Fey, M., Janssen, S., Zolitschka, B., 2009. Isotopic ngerprints on
lacustrine organic matter from Laguna Potrok Aike (southern Patagonia,
Argentina) reect environmental changes during the last 16,000 years. Journal
of Paleolimnology 42 (1), 81e102.
McCormac, F.G., Hogg, A.G., Blackwell, P.G., Buck, C.E., Higham, T.F.G., Reimer, P.J.,
2004. SHCal04 Southern Hemisphere calibration 0e11.0 cal Kyr BP. Radiocarbon
46 (3), 1087e1092.
McCulloch, R., Bentley, M.J., 1998. Late-glacial ice advances in the Strait of Magellan,
southern Chile. Quaternary Science Reviews 17, 775e787.
McCulloch, R., Davis, S., 2001. Late-glacial and Holocene palaeoenvironmental
change in the central Strait of Magellan, southern Patagonia. Palaeogeography,
Palaeoclimatology, Palaeoecology 173 (3e4), 143e173.
McCulloch, R., Morello, F., 2009. Evidencia glacial y paleoecolgica de ambientes
tardiglaciales y del Holoceno temprano. Implicaciones para el poblamiento
temprano de Tierra del Fuego. In: Salemme, M., Santiago, F., lvarez, M.,
Piana, E., Vzquez, M., Mansur, M.E. (Eds.), Arqueologa de Patagonia: una
mirada desde el ultimo conn. Editorial Utopas, Ushuaia, pp. 119e136.
McCulloch, R., Bentley, M., Purves, S., Hulton, N., Sudgen, R., Clapperton, Ch, 2000.
Climatic inferences from glacial and palaeoecological evidence at the Last
Glacial termination, southern South America. Journal of Quaternary Science 15
(4), 409e417.
McCulloch, R., Fogwill, C., Sugden, D., Bentley, M., Kubik, P., 2005. Chronology of the
last glaciation in central Strait of Magellan and Baha Intil, southernmost
South America. Geograska Annaler: Series A, Physical Geography 87 (2),
289e312.
McGlone, M.S., Turney, C.S.M., Wilmshurst, J.M., Renwick, J., Pahnke, K., 2010.
Divergent trends in land and ocean temperature in the Southern Ocean over the
past 18,000 years. Nature Geoscience 3 (9), 622e626.
Mercer, J.H., 1965. Glacier variations in southern Patagonia. Geogr. Rev. 55, 390e413.
Mercer, J.H., 1982. Holocene glacial variations in southern South America. Striae 18,
35e40.
Mller, M., Schneider, C., 2010. Calibration of glacier volumeearea relations from
surface extent uctuations and application to future glacier change. Journal of
Glaciology 56 (195), 33e40.
Moore, D.M., 1979. Southern oceanic wet-heathlands (including Magellanic Moorland). In: Specht, R.L. (Ed.), Heathlands and Related Shrublands, Ecosystems of
the World 9A. Elsevier, Amsterdam, pp. 489e497.
Moore, B.M., 1983. Flora of Tierra del Fuego. Anthony Nelson, Oswesty.
Morello, F., Borrero, L., Massone, M., Stern, C., Garca-Herbst, A., McCulloch, R.,
Arroyo-Kalin, M., Cals, E., Torres, J., Prieto, A., Martinez, I., Bahamonde, G.,
crdenas, P., 2012. Hunter-gatherers, biogeographic barriers and the development of human settlement in Tiera del Fuego. Antiquity 86, 71e87.
Moreno, P.I., Lowell, T.V., Jacobson Jr., G.L., Denton, G.H., 1999. Abrupt vegetation
and climate changes during the Last Glacial Maximum and last termination in
the Chilean Lake District: a case study from Canal de la Puntilla (41 S). Geograska Annaler: Series A 81, 285e311.
Moreno, P.I., Jacobson, G.L., Lowell, T.V., Denton, G.H., 2001. Interhemispheric
climate links revealed by a late-glacial cooling episode in southern Chile. Nature
409, 804e808. http://dx.doi.org/10.1038/35057252.
Moreno, P.I., Franois, J.P., Villa-Martnez, R.P., Moy, C.M., 2009a. Millennial-scale
variability in Southern Hemisphere westerly wind activity over the last 5000
years in SW Patagonia. Quaternary Science Reviews 28 (1e2), 25e38.
Moreno, P.I., Kaplan, M.R., Francois, J.P., Villa-Martinez, R., Moy, C.M., Stern, C.R.,
Kubik, P.W., 2009b. Renewed glacial activity during the Antarctic cold reversal
and persistence of cold conditions until 11.5 ka in Southwestern Patagonia.
Geology 37 (4), 375e378.
Moreno, P.I., Francois, J.P., Moy, C.M., Villa-Martnez, R., 2010. Covariability of the
Southern Westerlies and atmospheric CO2 during the Holocene. Geology 38 (8),
727e730.
Moy, C.M., Seltzer, G.O., Rodbell, D.T., Anderson, D.M., 2002. Variability of El Nino/
Southern oscillation activity at millennial timescales during the Holocene
epoch. Nature 420, 162e165.
Moy, C.M., Dunbar, R.B., Moreno, P., Francois, J.P., Villa-Martinez, R., Guilderson, T.P.,
Garreaud, R.D., 2008. Isotopic evidence for hydrologic change related to the
westerlies in SW Patagonia, Chile, during the last millennium. Quaternary
Science Reviews 27 (13e14), 1335e1349.
Moy, C.M., Dunbar, R.B., Guilderson, T.P., Waldmann, N., Mucciarone, D.A.,
Recasens, C., Ariztegui, D., Austin, J.A., Anselmetti, F.S., 2011. A geochemical and
sedimentary record of high southern latitude Holocene climate evolution from
Lago Fagnano, Tierra del Fuego. Earth and Planetary Science Letters 302 (1e2),
1e13.
Mhlinghaus, C., Scholz, D., Mangini, A., 2008. Temperature and precipitation
records from stalagmites grown under disequilibrium conditions: a rst
approach. PAGES News 16 (3), 19e20.
Mhlinghaus, C., Scholz, D., Mangini, A., 2009. Modelling fractionation of stable
isotopes in stalagmites. Geochimica et Cosmochimica Acta 73, 7275e7289.

22

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23

Nesje, A., Bakke, J., Dahl, S.O., Lie, O., Matthews, J.A., 2008. Norwegian mountain
glaciers in the past, present and future. Global and Planetary Change 60, 10e27.
Neukom, R., Luterbacher, J., Villalba, R., Kttel, M., Frank, D., Jones, P.D., Grosjean, M.,
Wanner, H., Aravena, J.C., Black, D.E., Christie, D.A., DArrigo, R., Larra, A.,
Morales, M., Soliz-Gamboa, C., Srur, A., Urrutia, R., von Gunten, L., 2010a. Multiproxy summer and winter surface air temperature eld reconstructions for
southern South America covering the past centuries. Climate Dynamics 37
(1e2), 35e51.
Neukom, R., Luterbacher, J., Villalba, R., Kttel, M., Frank, D., Jones, P.D., Grosjean, M.,
Esper, J., Lopez, L., Wanner, H., 2010b. Multi-centennial summer and winter
precipitation variability in southern South America. Geophysical Research
Letters 37, L14708.
New, M., Lister, D., Hulme, M., Makin, I., 2002. A high-resolution data set of surface
climate over global land areas. Climate Research 21, 1e25.
North Greenland Ice Core Project Members, 2004. High resolution climate record of
the northern hemisphere reaching into the Last Interglacial period. Nature 431,
147e151.
Oerlemans, J., 2005. Extracting a climate signal from 169 glacier records. Science
308 (5722), 675e677.
Orquera, L.A., Legoupil, D., Piana, E.L., 2011. Littoral adaptation at the southern end
of South America. Quaternary International 239, 61e69.
Pahnke, K., Sachs, J.P., 2006. Sea surface temperatures of southern midlatitudes
0e160 kyr BP. Paleoceanography 21 (2).
Pahnke, K., Zahn, R., Eldereld, H., Schulz, M., 2003. 340,000-year centennial-scale
marine record of Southern Hemisphere climatic oscillation. Science 301 (5635),
948e952.
Pendall, E., Markgraf, V., White, J.W.C., Dreier, M., 2001. Multiproxy record of Late
PleistoceneeHolocene climate and vegetation changes from a peat bog in
Patagonia. Quaternary Research 55 (2), 168e178.
Pisano, E., 1977. Fitogeografa de Fuego-Patagonia Chilena. 1: Comunidades vegetales entre las latitudes 52 y 56 S. Anales del Instituto de la Patagonia 8,
121e150.
Ponce, J.F., Borromei, A.M., Rabassa, J., Martinez, O., 2011. Late Quaternary palaeoenvironmental change in western Staaten island (54.5 S, 64 W), Fuegian
Archipelago. Quaternary International 233 (2), 89e100.
Porter, S.C., 2000. Onset of Neoglaciation in the Southern Hemisphere. Journal of
Quaternary Science 15 (4), 395e408.
Prahl, F.G., Wakeham, S.G., 1987. Calibration of unsaturation patterns in longchain ketone compositions for paleotemperature assessment. Nature 330,
367e369.
Prahl, F.G., Muehhausen, L.A., Zahnle, D.L., 1988. Further evaluation of long-chain
alkenones as indicators of paleoceanographic conditions. Geochimica et Cosmochimica Acta 52, 2303e2310.
Prahl, F.G., Rontani, J.-F., Zabeti, N., Walinsky, S.E., Sparrow, M.A., 2010. Systematic
pattern in UK37-Temperature residuals for surface sediments from high latitude and other oceanographic settings. Geochimica et Cosmochimica Acta 74
(1), 131e143.
Premoli, A.C., Mathiasen, P., Kitzberger, T., 2010. Southernmost Nothofagus tress
enduring ice ages: genetic and ecological niche retrodiction reveal high latitude
(54 S) glacial refugia. Palaeogeography, Paleoclimatology, Paleoecology 298,
247e256.
Prieto, A., Cals, E., Morello, F., Torres, J., 2009. Sitios arqueologico Myren2, Tierra
del Fuego. Magallania 35 (2), 83e103.
Putnam, A.E., Denton, G.H., Schaefer, J.M., Barrell, D.J.A., Andersen, B.G., Finkel, R.C.,
Schwartz, R., Doughty, A.M., Kaplan, M.R., Schlchter, C., 2010. Glacier advance
in southern middle-latitudes during the Antarctic Cold Reversal. Nature Geoscience 3, 700e703.
Rasmussen, L.A., Conway, H., Raymond, C.F., 2007. Inuence of upper air conditions on the Patagonia iceelds. Global and Planetary Change 59 (1e4),
203e216.
Recasens, C., Ariztegui, D., Gebhardt, C., Gogorza, C., Haberzettl, T., Hahn, A., LiesPronovost, Kliem, Lcke, A., Maidana, N., Maryr, C., Ohlendorf, C., Schbitz, F., StOnge, G., Wille, M., Zolitschka, B., POSADA Science Team, 2011. New insights
into paleoenvironmental changes in Laguna Potrok Aike, southern Patagonia,
since the Late Pleistocene: the Posada multiproxy record. The Holocene. http://
dx.doi.org/10.1177/0959683611429833.
Rein, B., Lckge, A., Reinhardt, L., Sirocko, F., Wolf, A., Dullo, W.-C., 2005. El Nio
variability off Peru during the last 20,000 years. Paleoceanography 20, PA4003.
http://dx.doi.org/10.1029/2004PA001099.
Renssen, H., Goosse, H., Fichefet, T., Brovkin, V., Driesschaert, E., Wolk, F., 2005.
Simulating the Holocene climate evolution at northern high latitudes using
a coupled atmosphere-sea ice-ocean-vegetation model. Climate Dynamics 24
(1), 23e43.
Rignot, E., Rivera, A., Casassa, G., 2003. Contribution of the Patagonia iceelds of
south America to sea level rise. Science 302 (5644), 434e437.
Rigozo, N.R., Nordemann, D.J.R., da Silva, H.E., Souza Echer, M.P., Echer, E., 2007.
Solar and climate signal records in tree ring width from Chile (AD 1587e1994).
Planetary and Space Science 55 (1e2), 158e164.
Rincn-Martnez, D., Lamy, F., Contreras, S., Leduc, G., Bard, E., Saukel, C., Blanz, T.,
Mackensen, A., Tiedemann, R., 2010. More humid interglacials in Ecuador
during the past 500 kyr linked to latitudinal shifts of the equatorial front and
the Intertropical Convergence Zone in the eastern tropical Pacic. Paleoceanography 25, PA2210. http://dx.doi.org/10.1029/2009PA001868.
Rivera, A., 2004. Mass balance investigations at Glaciar Chico, Southern Patagonia
Iceeld, Chile. PhD thesis, Bristol University, UK. 303 pp.

Rohling, E.J., Plike, H., 2005. Centennial-scale climate cooling with a sudden cold
event around 8,200 years ago. Nature 434 (7036), 975e979.
Rojas, M., Moreno, P., 2011. Atmospheric circulation changes and neoglacial
conditions in the Southern Hemisphere mid-latitudes: insights from PMIP2
simulations at 6 kyr. Climatre Dynamics 37, 357e375.
Rojas, M., Moreno, P., Kageyama, M., Crucix, M., Hewitt, C., Abe-Ouchi, A.,
Ohgaito, R., Brady, E.C., Hope, P., 2009. The Southern Westerlies during the Last
Glacial Maximum in PMIP2 simulations. Climate Dynamics 32 (4), 525e548.
http://dx.doi.org/10.1007/s00382-008-0421-7.
Rother, H., Shulmeister, J., 2006. Synoptic climate change as a driver of late
Quaternary glaciations in the mid-latitudes of the Southern Hemisphere.
Climate of the Past 2, 11e19.
Sapkota, A., Cheburkin, A.K., Bonani, G., Shotyk, W., 2007. Six millenia of atmospheric dust deposition in southern south America (Isla Navarino, Chile). The
Holocene 17 (5), 561e572.
Schimpf, D., Kilian, R., Kronz, A., Simon, K., Sptl, C., Wrner, G., Deininger, M.,
Mangini, A., 2011. The signicance of chemical, isotopic, and detrital components in three coeval stalagmites from the superhumid southernmost Andes
(53 S) as high-resolution palaeo-climate proxies. Quaternary Science Reviews
30 (3e4), 443e459.
Schbitz, F., 1991. Holocene vegetation and climate in southern Santa Cruz,
Argentina. Bamberger Geographische Schriften 11, 235e244.
Schneider, C., Gies, D., 2004. Effects of El Nioesouthern oscillation on southernmost South America precipitation at 53 S revealed from NCEPeNCAR reanalyses and weather station data. International Journal of Climatology 24 (9),
1057e1076.
Schneider, C., Kilian, R., Santana, A., Butorovic, N., Casassa, G., 2003. Weather
observations across the southern Andes at 53 S. Physical Geography 24 (2),
97e119.
Schneider, C., Schnirch, M., Acna, C., Casassa, G., Kilian, R., 2007. Glacier inventory
of the Gran Campo Nevado Ice Cap in the Southern Andes and glacier changes
observed during the recent decades. Global and Planetary Change 59 (1e4),
87e100.
Seplveda, J., Pantoja, S., Hughen, K., Bertrand, S., Figueroa, D., Len, T., Drenzek, N.,
Lange, C., 2009. Late Holocene sea-surface temperature and precipitation
variability in northern Patagonia, Chile (Jacaf Fjord, 44 S). Quaternary Research.
http://dx.doi.org/10.1016/j.yqres.2009.06.010.
Siddall, M., Rohling, E.J., Almogi-Labin, A., Hemleben, Ch., Meischner, D., Schmelzer,
Smeed, D.A., 2003. Sea level uctuations during the Last Glacial cycle. Nature
423, 853e858.
Steinhilber, F., Beer, J., Frhlich, C., 2009. Total solar irradiance during the Holocene.
Geophysical Research Letters 36, L19704. http://dx.doi.org/10.1029/
2009GL040142.
Stenni, B., Burion, D., Frezzotti, M., Albani, S., Barbante, C., Bard, E., Barnola, J.M.,
Baroni, M., Baumgartner, M., Bonazza, M., Capron, E., Castellano, E.,
Chappellaz, J., Delmonte, B., Falourd, S., Genoni, L.-, Iacumin, P., Jouzel, J.,
Kipfstuhl, S., Landais, A., Lemieux-Dudon, B., Maggi, V., Masson-Delmonte, V.,
Mazzola, C., Minster, B., Montagnat, M., Mulvaney, R., Narcisci, B., Oerter, H.,
Parrenin, F., Petit, J.R., Ritz, C., Scarchilli, C., Schilt, A., Schpach, S., Schwander, J.,
Selmo, E., Sereri, M., Stocker, T.F., Udisti, R., 2011. Expression of the bipolar seesaw in Antarctic climate records durinmg the last deglaciation. Nature Geosience 4, 46e50.
Stern, C.R., 1990. Tephrochronology of southernmost Patagonia. National
Geographic Research 6 (1), 110e126.
Stern, C.R., 1992. Tefrocronologia de Magellanes: nuevos datos e implicaciones.
Anales del Instituto de la Patagonia 21, 129e141.
Stern, C.R., 2008. Holocene tephrochronology record of large explosive eruptions in
the southernmost Patagonia Andes. Bulletin of Volcanology 70 (4), 435e454.
Stern, C.R., Kilian, R., 1996. Role of the subducted slab, mantle wedge and continental crust in the generation of adakites from the Andean Austral Volcanic
Zone. Contributions to Mineralogy and Petrology 123 (3), 263e281.
Stern, C.R., Moreno, P.I., Villa-Martnez, R., Sagredo, E.A., Prieto, A., Labarca, R., 2011.
Evolution of ice-dammed proglacial lakes in ltima Esperanza, Chile: implications from the late-glacial R1 eruption of Recls volcano, Andean Austral
Volcanic Zone. Andean Geology 38 (1), 82e97.
Stine, S., Stine, M., 1990. A record from Lake Cardiel of climate change in southern
South America. Nature 345, 705e708.
Stocker, T.F., 1998. Climate change: the seesaw effect. Science 282 (5386), 61e62.
Stocker, T.F., Johnsen, S.J., 2003. A minimum thermodynamic model for the bipolar
seesaw.
Paleoceanography
18
(4),
1087.
http://dx.doi.org/10.1029/
2003PA000920.
Stott, L., Timmermann, A., Thunell, R., 2007. Southern Hemisphere and deep-sea
warming led deglacial atmospheric CO2 rise and tropical warming. Science
318 (5849), 435e438.
Strelin, J.A., Denton, G.H., Vandergoes, M.J., Ninnemann, U.S., Putnam, A.E., 2012.
Radiocarbon chronology of the late-glacial Puerto Bandera moraines, southern
Patagonian Iceeld, Argentina. Quaternary Science Reviews. http://dx.doi.org/
10.1016/j.quascirev. 2011.05.004.
Sugden, D., Bentley, M., Fogwill, M., Hulton, N., McCulloch, R., Purves, R., 2005. Lateglacial glacier events in southernmost South America: a blend of northern and
southern hemispheric climatic signals? Geograska Annaler: Series A, Physical
Geography 87 (2), 273e288.
Sugden, D.E., McCulloch, R.D., Bory, A.J.M., Hein, A.S., 2009. Inuence of Patagonian
glaciers on Antarctic dust deposition during the Last Glacial period. Nature
Geoscience 2 (4), 281e285.

R. Kilian, F. Lamy / Quaternary Science Reviews 53 (2012) 1e23


Timmermann, A., Timm, O., Stott, L., Menviel, L., 2009. The roles of CO2 and orbital
forcing in driving southern hemispheric temperature variations during the last
21,000 years. Journal of Climate 22 (7), 1626e1640.
Toggweiler, J.R., Joellen Russell, J., 2008. Ocean circulation in a warming climate.
Nature 451, 286e288.
Toggweiler, J.R., Russell, J.L., Carson, S.R., 2006. Midlatitude westerlies, atmospheric
CO2, and climate change during the ice ages. Paleoceanography 21, PA2005.
Tonella, M.S., Mancini, M.V., Sepp, H., 2009. Quantitative reconstruction of Holocene
precipitation changes in southern Patagonia. Quaternary Research 72, 410e420.
Torres, J., 2009. La pesca entre los cazadores recolectores terrestres de la Isla Grande
de Tierra del fuego, desde la prehistoria a tiempos etnogrcos. Magallania 37
(2), 109e138.
Unkel, I., Fernandez, M., Bjrk, S., Ljung, K., Wohlfarth, B., 2010. Records of environmental changes during the Holocene from Isla de los Estados (54.4 S),
southeastern Tierra del Fuego. Global and Planetary Change 74 (3e4), 99e113.
Van Geel, B., Heusser, C.J., Renssen, H., Schuurmans, C.J.E., 2000. Climatic change in
Chile at around 2700 BP and global evidence for solar forcing: a hypothesis. The
Holocene 10, 659e664.
Varma, V., Prange, M., Lamy, F., Merkel, U., Schulz, M., 2011. Solar-forced shifts of the
southern hemisphere westerlies during the Holocene. Climate of the Past 7 (2),
339e347. doi:10.5194/cp-7-339-2011.
Vieira, L.E.A., Solanki, S.K., Krivova, N.A., Usoskin, I., 2011. Evolution of the solar
irradiance during the Holocene. Astronomy and Astrophysics 531, A6.
Villa-Martnez, R., Moreno, P.I., 2007. Pollen evidence for variations in the southern
margin of the westerly winds in SW Patagonia over the last 12,600 years.
Quaternary Research 68 (3), 400e409.
Villalba, R., Lara, A., Boninsegna, J.A., Masiokas, M., Delgado, S., Aravena, J.C.,
Roig, F.A., Schmelter, A., Wolodarsky, A., Ripalta, A., 2003. Large-scale temperature changes across the southern Andes: 20th-century variations in the
context of the past 400 years. Climatic Change 59 (1e2), 177e232.
Villalba, R., Grosjean, M., Kiefer, T., 2009. Long-term multi-proxy climate reconstructions and dynamics in South America (LOTRED-SA): state of the art and
perspectives. Palaeogeography, Palaeoclimatology, Palaeoecology 281, 175e179.
Vimeux, F., De Angelis, M., Ginot, P., Magand, O., Casassa, G., Pouyaud, B., Falourd, S.,
Johnsen, S., 2008. A promising location in Patagonia for paleoclimate and
paleoenvironmental reconstructions revealed by a shallow rn core from
Monte San Valentn (Northern Patagonia Iceeld, Chile). Journal of Geophysical
Research 113, 1e20.
Vimeux, F., Ginot, P., Schwikowski, M., Vuille, M., Hoffmann, G., Thompson, L.G.,
Schotterer, U., 2009. Climate variability during the last 1000 years inferred from
Andean ice cores: a review of methodology and recent results. Palaeogeography, Palaeoclimatology, Palaeoecology 281 (3e4), 229e241.

23

Wagner, S., Widmann, M., Jones, J., Haberzettl, T., Lucke, A., Mayr, C., Ohlendorf, C.,
Schabitz, F., Zolitschka, B., 2007. Transient simulations, empirical reconstructions and forcing mechanisms for the Mid-Holocene hydrological climate in
southern Patagonia. Climate Dynamics 29 (4), 333e355.
Waldmann, N., Ariztegui, D., Anselmetti, F.S., Austin, J.A., Stern, C., Moy, C.M.,
Recasens, C., Dunbar, R., 2010. Holocene climatic uctuations and positioning of
the Southern Hemisphere westerlies in Tierra del Fuego (54 S), Patagonia.
Journal of Quaternary Science 25 (7), 1063e1075.
Waldmann, N., Anselmetti, F.S., Austin, J.A., Pirouz, M., Moy, C.M., Dunbar, R., 2011.
Holocene mass-wasting events in Lago Fagnano, Tierra del Fuego (54 S):
implications for paleoseismicity of the Magallanes-Fagnano transform fault.
Basin Research 23 (2), 171e190.
Wanner, H., Beer, J., Btikofer, J., Crowley, T.J., Cubasch, U., Flckiger, J., Goosse, H.,
Grosjean, M., Joos, F., Kaplan, J.O., Kttel, M., Mller, S., Prentice, I.C.,
Solomina, O., Stocker, T.F., Tarasov, P., Wagner, M., Widmann, M., 2008. Midto late Holocene climate change: an overview. Quaternary Sci. Rev. 27,
1791e1828.
Wanner, H., Solomina, O., Grosjean, M., Ritz, S., Jetel, M., 2011. Structure and
origin of Holocene cold events. Quaternary Science Reviews 30 (21e22),
3109e3123.
Warren, C., Sudgen, D.E., 1993. The Patagonian Iceelds: a Glaciological review.
Arctic and Alpine Research 25 (4), 316e331.
Wenzens, G., 1999. Fluctuations of outlet and valley glaciers in the southern Andes
(Argentina) during the past 13,000 years. Quaternary Research 51 (3), 238e247.
Wenzens, G., 2005. Glacier advances east of the Southern Andes between the Last
Glacial Maximum and 5,000 BP compared with lake terraces of the endorrheic
Lago Cardiel (49 S, Patagomgonia, Argentina). Zeitschrift fur Geomorphologie
N.F 49 (4), 433e454.
Wenzens, G., Wenzens, E., 1998. Late glacial and Holocene glacier advances in the
area Lago Viedma (Patagonia, Argentina). Zentralblatt Geologie Palontologie.
Teil I 1997 (3e6), 593e608.
Whitlock, C., Moreno, P.I., Bartlein, P., 2007. Climatic controls of Holocene re
patterns in southern South America. Quaternary Research 68 (1), 28e36.
Wille, M., Schbitz, F., 2009. Late-glacial and Holocene climate dynamics at the
steppe/forest ecotone in southernmost Patagonia, Argentina: the pollen record
from a fen near Brazo Sur, Lago Argentino. Vegetation History and Archaeobotany 18 (3), 225e234.
Wille, M., Maidana, N.I., Schbitz, F., Fey, M., Haberzettl, T., Janssen, S., Lcke, A.,
Mayr, C., Ohlendorf, C., Schleser, G.H., Zolitschka, B., 2007. Vegetation and
climate dynamics in South America: the microfossil record of Laguna Potrok
Aike, Santa Cruz, Argentina. Review of Paleobotany and Palynology 146,
234e246.

Das könnte Ihnen auch gefallen