You are on page 1of 13

Journal of South American Earth Sciences 71 (2016) 235e247

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


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

Deglaciation chronology in the Me rida Andes from cosmogenic 10


Be
dating, (Gavidia valley, Venezuela)
Isandra Angel a, c, *, Franck A. Audemard M. b, Julien Carcaillet a, Eduardo Carrillo c,
Christian Beck d, Laurence Audin e
a
ISTerre, Universit
e de Grenoble 1, UMR CNRS 5275, F-38041, Grenoble, France
b
Fundacio n Venezolana de Investigaciones Sismolo gicas, FUNVISIS, El Llanito, Caracas, 1070, Venezuela
c
Instituto de Ciencias de la Tierra, Universidad Central de Venezuela, Apdo. 3805, Caracas, 1010-A, Venezuela
d
ISTerre, Universit
e de Savoie, UMR CNRS 5275, F-73376, Le Bourget-du-Lac, France
e
ISTerre, IRD, Universit
e de Grenoble 1, F-38041, France

a r t i c l e i n f o a b s t r a c t

Article history: In the Merida Andes, a detailed deglaciation history reconstruction is difficult to achieve due to scattered
Received 6 January 2016 deglaciation chronologies available. This paper contributes with 24 exposure ages of glacial landforms
Received in revised form sampled in the Gavidia valley. Exposure ages were obtained based on terrestrial cosmogenic nuclide 10Be
16 July 2016
dating. Results indicate deglaciation mainly occurred between ~21 ka and 16.5 ka and the complete
Accepted 8 August 2016
deglaciation occurred at ~16.0 ka. The glacier retreated in two different phases. The oldest one occurred
Available online 9 August 2016
since the LGM until middle OtD or the local climate event El Caballo Stadial. The youngest phase occurred
at ages younger than ~16.5 ka until complete deglaciation. A combination of topographic features and
Keywords:
Terrestrial cosmogenic nuclides dating
changes in the paleoclimate conditions at the end of the El Caballo Stadial seems leaded the fastest
TCN former glacier extinction. The topographic feature which seems contributed to the fastest glacier
Glacial landforms extinction was the low valley bottom slopes. In addition, exposure ages of the Gavidia valley were in-
Merida Andes tegrated with deglaciation chronologies from the central Me rida Andes to compare deglaciation his-
Venezuela tories. Asynchronous deglaciation histories were observed. Local paleotemperatures and
Pleistocene paleoprecipitations contrasts, different valleys aspects, insolation and catchments steepness could
Late Glacial explain different deglaciation histories.
LGM
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction chronologies data (e.g. Schubert, 1972, 1974, 1980, 1980a, 1992 and
1998). First deglaciation chronologies data are based on radio-
In the Merida Andes glacier landforms and glaciological studies carbon chronology (Schubert, 1970; Salgado-Labouriau and Schu-
were developed since the end of the XIXth century (Sievers, 1885; bert, 1977; Schubert and Rinaldi, 1987; Rull, 1998; Mahaney et al.,
Goering, 1962). Jahn (1912, 1925 and 1931) made first planimetric 2001; Dirszowsky et al., 2005; Stansell et al., 2005; Mahaney et al.,
measures for the existent glaciers. Glaciers surface have reduced 2007; Carrillo et al., 2008), Thermoluminescence (TL; Schubert and
from around 200 km2 (Schubert and Clapperton, 1990) to 2.9 km2 Vaz, 1987; Bezada, 1989), Optically Stimulated Luminescence (OSL,
(Schubert, 1992) from the Last Glacial Maximum (LGM) to the Mahaney et al., 2000) and more recently, Terrestrial Cosmogenic
middle of XX century. This glaciers surface was reduced from Nuclide (TCN) dating (Wesnousky et al., 2012; Angel et al., 2013;
0.33 km2 (Carrillo and Ye pez, 2008) to 0.017 km2 (Braun and Carcaillet et al., 2013; Guzma n et al., 2013).
Bezada, 2013) between 2008 and 2011. Limited deglaciation chronologies in the Me rida Andes make
Glacial landforms studies in the Me rida Andes derived into difficult to reconstruct a more detailed deglaciation history of the
glacial geological interpretations with limited deglaciation North of the Tropical Andes. The reconstruction of a more detailed
deglaciation history in this area, improves the Tropical glaciations
and paleoclimate knowledge. This knowledge allows to better
perform climatic projections and therefore, to minimize the impact
* Corresponding author. ISTerre, 1381 Rue de la Piscine, 38400, Saint Martin
d'Heres, France. of climate change on society.
E-mail address: iangel_ceballos@yahoo.com (I. Angel). In this article, we propose a new deglaciation chronology based

http://dx.doi.org/10.1016/j.jsames.2016.08.001
0895-9811/© 2016 Elsevier Ltd. All rights reserved.
236 I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247

on 24 new TCN ages from the Gavidia valley. We aim to precise (i) Quaternary uplift ranges between ~0.7 and 5 mm/a (Audemard and
the timing of the last deglaciation in the Gavidia valley, (ii) pa- Audemard, 2002; Wesnousky et al., 2012; Guzma n et al., 2013).
rameters driven the deglaciation history in this valley and, (iii) a Mainly glacial landforms studied in the MA since the XIXth
comparison with other deglaciation histories in the Merida Andes century are moraines, glacier cirques, glacier valleys, glacier lakes
mainly based on 10Be cosmonuclide dating. and paraglacial sediment deposits (e.g. Sievers, 1885; Jahn, 1912,
1925; Schubert, 1970, 1974; Schubert and Valastro, 1980; Schubert
and Rinaldi, 1987; Bezada, 1989; Schubert and Clapperton, 1990).
2. Regional setting In the MA tectonic control on glacial landforms has been recorded
and attributed to the Bocono  dextral fault (BF) (Audemard, 2003).
2.1. Geologic, tectonic and geomorphic settings The BF is NEeSW oriented and ~500 km long (Fig. 1). Quaternary
slip rates of the BF range between 3 and 14 mm/a (Audemard, 2003;
The Me rida Andes (MA) is located in the southwest of Venezuela
Schubert, 1980a; Soulas, 1985; Soulas et al., 1986; Audemard et al.,
(Fig. 1), it is ~N45 oriented and it extends over 400 km. The MA 1999; Wesnousky et al., 2012).
highest elevation is the Pico Bolivar at 4978 m a.s.l. The MA
orogenesis is mainly connected to the geodynamic interaction of
the Panama  Arc, Caribbean and South American plates (Taboada 2.1.1. Geomorphic description of the Gavidia valley
et al., 2000; Audemard and Audemard, 2002; Bermudez, 2009; The Gavidia valley is located in the Sierra Nevada in the central
Monod et al., 2010). These plates interactions provide that part of the MA (Fig. 1). It is a postglacial u-shaped valley which is

rida Andes. The dots indicate locations cited in the study (see text for details) B) Aerial photography (Google earth, 2014) of the
Fig. 1. A) Digital shaded-relief map of the central Me
Gavidia valley with main geomorphological landforms and valley contours.
I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247 237

deformed by two different inferred faults (Fig. 1B). Its headwalls are surfaces (roches moutonne es) are observed in the valley bottom. In
located at Pico Alto Santo Cristo (4230 m a.s.l.). the valley bottom, moraines are not developed as in other nearby
Quartz gneiss from the Iglesias Complex is the main bedrock glacier valleys (e.g Mucubají valley) and polished surfaces are
lithology (Hackley et al., 2005). In bedrock, foliation and glacier present in the valley side-walls (Fig. 2).
striations are easily recognizable (Fig. 2A). Striated and polished The Gavidia valley is separated in two main areas:

Fig. 2. Examples of the glacial landforms in the Gavidia valley. A) Striated bedrock in the axis of the Gavidia Valley, (GA-1203 location). B) Sub-vertical polished surface in the
northernmost stretch of the Gavidia valley (close to GA-1201). Glacial striations directions are indicated by white arrows C) Difference in orientation between glacier striation (white
arrows) and foliation (white lines) on a sub horizontal glacial surface. D) Sample position in the Gavidia valley (Google earth, 2014). White stars are polished surface, Grey stars are
roches moutonne es and only striated rock, and white dots are moraine boulders.
238 I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247

- The low part (~3250e~3600 m a.s.l) is NW-SE oriented and it is 16.7 ± 1.4 ka and from Sierra del Norte at 15.2 ± 0.9 ka could be
4 km long with a valley bottom slope ~ 5 . It exhibits perpen- related to this period (Wesnousky et al., 2012).
dicular hanging valleys (Fig. 1B) and roches moutonne es (Fig. 2). Stansell et al. (2005) concluded that in Sierra del Norte, glaciers
A ~40 m high step across the valley defines the upper limit of had significantly retreated by 15.70 ka BP. Then, several minor
this part, close to sample GA-1207 (Fig. 2). glacier advances and retreats between 14.85 and 13.83 ka BP also
- The high part (~3600e~4200 m a.s.l) is NE-SW orientated and it happened. However, these authors do not specify how long glaciers
is ~5 km long. In this section, valley slopes range between 3 and did retreat. A moraine inset into the former laterofrontal moraines
7. In this upper area two different glacier cirques have influence of La Culata (Sierra del Norte) was interpreted as a small Late Glacial
(Fig. 1B) and it is filled by gelifraction deposits. In addition, small readvance at weighted average exposure age of 14.1 ± 1.0 ka
moraines are present with ridges <10 m high. A ~100 m high (Wesnousky et al., 2012).
step is across the glacier valley close to sample GA-1212 (Fig. 2B).
2.3. Present-day climatic conditions
2.2. Paleoclimatic setting
Climate in the northern tropics is mainly controlled by the
Quaternary climate studies in the MA are based on analysis of Intertropical Convergence Zone (ITCZ) fluctuation. The ITCZ fluc-
lacustrine, fluvial and glacial deposits and paleosols (e.g. Schubert, tuation is highly dependent on the seasonal cycle of the solar
1974; Schubert and Valastro, 1980; Bradley et al., 1985; Salgado- declination (Benn et al., 2005).
Labouriau et al., 1988; Weingarten et al., 1991; Salgado-Labouriau Moisture is predominantly derived from evaporation over the
et al., 1992; Rull, 1998; Mahaney et al., 2000; Dirszowsky et al., tropical Atlantic and evapotranspiration from the Orinoco River
2005; Rull et al., 2005; Stansell et al., 2005; Mahaney et al., 2007; Basin which are advected toward the Andes by easterly trade winds
Carrillo et al., 2008; Rull et al., 2010; Stansell et al., 2010). Lake (Pulwarty et al., 1998). Modern climate data from the MA also
sediments, fluvial terraces and moraines are the main landforms demonstrate the pervasive influence of equatorial Pacific sea sur-
studied (e. g. Salgado-Laboriau and Schubert, 1976; Stansell et al., face temperatures (SSTs) (Polissar et al., 2013).
2007; Wesnousky et al., 2012; Carcaillet et al., 2013). The study of Precipitation is highly seasonal with a maximum during the
lake sediments provides a continuous record during a specific boreal summer and minimum during winter (Pulwarty et al., 1998).
period (e.g. Carrillo, 2006; Stansell et al., 2007). In contrast, fluvial However, more recently, Poveda et al. (2006) established also
terraces and moraines provide a discontinuous record of the pa- maximum precipitations during fall and spring boreal seasons.
leoclimatic setting and the deglaciation history (e.g. Kirkbride and Precipitation patterns are also affected by orographic controls and
Brazier, 1998). local mountain circulation systems (Pulwarty et al., 1998; Poveda
In the MA the Quaternary paleoclimatic setting is closely related et al., 2006).
to glaciations periods. Glaciations periods have been mainly Concerning average temperature in the MA, it varies little
established based on studies of two moraine complexes which have seasonally while daily temperature variations may be as much as
been recognized and mapped between 2600 and 2800 m a.s.l. and 20  C (Schubert and Clapperton, 1990). These temperature varia-
2900e3500 m a.s.l. These moraines complexes were observed in tions are determined by insolation, solar radiation and cloudiness
Paramo de La Culata, P aramo de Piedras Blancas and Sierra de Santo factors (Monasterio and Reyes, 1980).
Domingo. These two complexes were first pointed out by Royo y
Go mez (1959) and later assigned to the Early and Late Me rida 3. Methods and materials
Glaciation (Schubert, 1970, 1974). Early Me rida Glaciation is poorly
constrained and established between ~60 and ~90 ka (Mahaney 3.1. Morphostratigraphic relation of glacial landforms and
et al., 2000, 2001, 2010, 2011; Dirszowsky et al., 2005). Then, implications on reconstruction of the deglaciation histories
interstadial conditions were established between ~60 and ~25 ka
with temperatures ~5  C warmer than the Last Glacial Maximum Chronological information about the deglaciation history of a
(LGM). This period is locally named El Pedregal Interstadial paleo or a former glacier can be deduced from morphostratigraphic
(Schubert and Valastro, 1980; Schubert and Vivas, 1993; analyses of glacial landforms. Spatial distribution of glacial land-
Dirszowsky et al., 2005; Rull, 2005; Mahaney et al., 2000; forms in a landscape provides qualitative chronological information
Guzma n et al., 2013). (e.g. Hughes et al., 2005). For example, in a valley, moraines at low
Late Me rida Glaciation (Late Wisconsin) (related to moraines elevations were deposited by the glacier before the upper ones.
between 2600 and 2800 m a.s.l.) is better constrained and range When the glacier retreats, glacial landforms such as moraines or
between 25 and 13 ka (Schubert and Clapperton, 1990). This period roches moutonne es at lower elevations are first exposed, then,
includes the LGM which is established in the MA between 22.75 glacier landforms at highest elevation are lately exposed.
and 19.96 Cal ka BP (Schubert and Rinaldi, 1987). In the Mucubají Ice marginal moraines (a lateral, a frontal or a latero-frontal) are
valley the LGM is related to an exposure age of a frontal moraine at created during a period of the glacier equilibrium with climate
18.14 ± 2.11 ka (weighted average, Carcaillet et al., 2013). followed by a period of positive mass balances (e.g. Benn and Evans,
In the MA during the LGM, the equilibrium line altitudes (ELAs) 2010). During a period of positive mass balances, more mass
were 850e1420 m a.s.l. lower than the present (Stansell et al., accumulation occurs in a glacier, so the glacier grows, it advances
2007). Modern ELAs are estimated between 4880 and 4470 m and it creates an ice marginal moraine. Therefore, an ice marginal
a.s.l. So, local LGM temperatures were at least 8.8 ± 2  C cooler than moraine denotes a glacier advance. The lowest ice marginal
today (Stansell et al., 2007). This assumption is based on ELAs moraine limits the maximum glacier advance in a valley. The frontal
variation and a combined energy and mass-balance equation to or a latero-frontal moraine related to the maximum glacier advance
account for an ELA lowering (Stansell et al., 2007). is denoted as a terminal moraine.
During the Late Glacial (LG), El Caballo Stadial, a cold period In the deglaciation histories and paleoclimate reconstruction is
dated at 16.5 ± 0.3 ka BP was identified based on pollen content of important to know that studies of glacial landforms represent
fluvioglacial sediments from Mesa del Caballo section (Rull, 1998). discontinues or partial record of the deglaciation history (e.g.
Temperatures were around 7  C lower than today (Rull, 1998). Local Kirkbride and Brazier, 1998). In addition, all glacier advances
Last Glacial Maximum (LLGM) established for Sierra Nevada at occurred in a valley could be poorly represented by moraines
I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247 239

distribution in a landscape. This is because of different processes been reported between 0.3 and 0.5 m.Myr1 (Smith et al., 2005)
contributing to remove ancient moraines: a) a more extended and and 0.45 m.Myr1 (Kelly et al., 2013). Comparison of extreme
younger glacier advance, b) proglacial erosion, c) denudation and, exposure ages with 0 and 4.5 m.Myr1 gives no significant differ-
d) gravitational processes (landslides) (Kirkbride and Brazier, 1998). ences at the “post-LGM scale”. For this reason and because of the
Because of this, integration with proglacial sediments studies, as for clear presence of the striations and polished surfaces on sampled
example from a proglacial lake or terraces, complements de- bedrock, erosion rate of 0 m.Myr1 was considered for calculations.
glaciations or glaciations reconstruction studies. Based on valley side-wall samples GA-1208L and GA-1216L, glacier
thickness could be more than 50e100 m. This glacier thickness
10
3.2. Be nuclide dating implications on deglaciation studies should be enough to avoid TCN production during the last climatic
cycle (i.e. during the presence of paleo or former glacier).
The interaction between cosmic rays and chemical targets of the
Earth environment produces Beryllium-10 (10Be). It is a cosmogenic 4. Results
nuclide isotope (Half-life 1.36 ± 0.07 Ma) (Gosse and Phillips, 2001).
This cosmogenic nuclide isotope is named in-situ 10Be when is Twenty-four rock samples were collected in the Gavidia valley,
10
formed in the first meters of the lithosphere exposed to cosmic rays Be concentrations range between 2.25 ± 0.08 (GA-1201) and
(Terrestrial Cosmogenic Nuclide, TCN). It thus constitutes a suitable 5.09 ± 0.16 (GA-1214)*105 atoms per gram of quartz (at.g1)
tool for dating exposure age of rock surfaces. Rock surfaces of glacial (Table 1). 10Be concentrations once computed as exposure ages,
landforms are exposed to the cosmic rays after deglaciation or ice results range between 13.98 ± 1.19 and 27.22 ± 1.32 ka (Table 1).
retreat so; exposure ages or deglaciation chronologies denote time Analytical uncertainties of exposure ages are in general lower than
since glacial landform is ice free. It is important to note that 10% with only GA-1205 around 14% (Table 1). Table 1 summarizes
deglaciation chronologies are assumed as minimum values. This is exposure ages and input data used in the Cronus online calculator
because of potential post-deglaciation processes that could erode 2.2 (Balco et al., 2008).
glacial landforms surfaces (Nishiizumi et al., 1989; Briner and Exposure ages could be divided by ages younger and older than
Swanson, 1998; Siame et al., 2000; Gosse and Phillips, 2001; ~20 ka (Fig. 3). Exposure ages of polished surfaces from the valley
Dunai, 2010; Balco, 2011). side-walls are older than those from the valley bottom (Fig. 3). Two
In contrast to moraines, where eroded boulders may result in exceptions are evidenced, GA-1214L (16.59 ± 0.64 ka) and GA-1215L
dispersed ages; bedrock surfaces are largely insensitive to post- (16.06 ± 0.63 ka) which are significant different and young. These
glacial disturbance. Therefore, bedrock surfaces provide a more polished surfaces could be affected by post deglaciation processes
accurate age control. However, when bedrock surfaces were not (covered by sediments or focused limited erosion). For that, GA-
enough eroded by glaciers (less than 2e3 m), inherited or initial in- 1214L and GA-1215L were considered as outliers and will not be
situ 10Be is present in glacial landforms after deglaciation. This considered in the discussion.
inherited in-situ 10Be implies overestimated exposure ages (Gosse GA-1218L (27.22 ± 1.32 ka) is the oldest exposure age. However,
et al., 1995; Guido et al., 2007; Balco, 2011). it is located at low elevation compared to GA-1216L and GA-1217L
(Table 1), it thus suggests a significant content of 10Be inherited.
3.3. Samples collection This sample was thus considered as outlier and will not be
considered in the discussion as well.
Samples were mainly collected in the Gavidia valley in glacial
landforms from the valley bottom and the valley side-wall (Fig. 2). 5. Discussion
Glacial landforms mainly sampled were striated and/or polished
rocks (including roches moutonne es). Striated and/or polished 5.1. Deglaciation history in the Gavidia valley and its potential
rocks sampled were at least 50 cm high from the valley bottom in driven topographical features
order to minimize potential cover by superficial deposits.
TCN 10Be chemical targets were prepared at the cosmogenic Exposure ages or deglaciation chronologies denote time since
laboratory at ISTerre, France, following procedures adapted from glacial landforms are ice free. Exposure ages of glacial landforms
Brown et al. (1991) and Merchel and Herpers (1999). Measure- from the valley bottom decrease with increasing elevation (Fig. 2).
ments were carried out at the French National AMS facility However, exposure ages from high elevation (>~3500 m a.s.l.
(Accelerator Mass Spectrometry) at ASTER in Aix-en-Provence. including the Gavidia valley glacier cirque and glacier cirques sur-
Ages were calculated using the online Cronus calculator (Balco roundings) are not significantly different (Fig. 3). The exposure ages
et al., 2008). results indicate deglaciation happened during the Late Me rida
A recent 10Be production rate (i.e., production rates by neutron Glaciation.
spallation appropriate for sea-level, high-latitude sites SLHL) was Bedrock surfaces are largely unaffected by postglacial processes
determined in the tropical Andes; 3.78 ± 0.09 at g1.yr1 when compared to moraine boulder surfaces. Therefore, bedrock surfaces
erosion is 0 cm/yr (Kelly et al., 2013). In this study we used Lago dating provides more accurate deglaciation ages (Gosse et al., 1995;
Argentino calibration in Cronus online calculator with a 10Be pro- Guido et al., 2007; Balco, 2011). However, deglaciations ages from
duction rate of 3.81 ± 0.13 at g1.yr1 (Kaplan et al., 2011). This bedrock are not related to a glacier advances as deglaciations ages
production rate was used because there are no significant differ- of moraines are.
ences with Kelly et al. (2013) value and calibration was available in The glacial landforms sampled in the Gavidia valley at lowest
Cronus online calculator. Ages were computed using the scaling elevation were a polished rock from the valley side-wall (GA-1201)
scheme of time dependent model from Lal (1991) modified by and roches moutonne es from the valley bottom (GA-1203, GA-1205
Stone (2000). This scaling scheme was selected because it considers and GA-1206). Exposure ages of these glacial landforms prevent to
the geomagnetic field variation (Balco et al., 2008). This is impor- infer time of the maximum glacier advance. However, instead of
tant in the present study because the magnetic modulation has a determining the maximum glacier advance, the maximum exten-
particularly critical effect in the vicinity of the magnetic Equator sion of the former glacier could be established. The maximum
where the samples were collected. glacier extension happened between ~21 ka and ~16.5 ka. In
Long-term erosion rates of boulders in the tropical Andes have contrast, the complete valley deglaciation happened at ~16.0 ka
240
Table 1
Exposure ages or TCN results from the Gavidia valley.

Samples Latitude Longitude Elevation Site information Sample Shielding Scaled production 10Be/9Be blank 10
Be Ages(d, e)

thickness factor(a) rates corrected(b) concentraction(c)

Spallations Muons Value Uncertainty Value External Internal


uncertainty uncertainty
 
N W m cm atoms/g/yr 10-13 103 atoms/gQtz/yr ka

GA-1201 8.694 70.940 3198 Polished rock (valley-side 3 0.488 8.88 0.480 2.04 225.00 8.31 21.679 1.085 0.899
wall)
GA-1203 8.674 70.919 3374 Roche moutonne e 1 0.942 19.10 0.510 1.90 294.39 22.92 13.977 1.188 1.178
GA-1205 8.667 70.912 3428 Roche moutonne e 1.5 0.927 19.24 0.515 2.48 334.48 43.33 15.640 2.096 2.207
e

I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247


GA-1206 8.663 70.909 3494 Roche moutonne 2 0.920 19.66 0.523 4.22 375.30 19.66 17.025 1.062 0.983
GA-1207 8.662 70.906 3568 Roche moutonne e 2 0.989 21.93 0.533 4.17 410.76 18.19 16.752 0.933 0.817
GA-1208 8.654 70.909 3635 Polished rock (valley-side 6 0.543 12.15 0.535 2.23 306.78 16.78 21.760 1.402 1.344
L wall)
GA-1209 8.653 70.908 3592 Boulder in a lateral moraine 1.5 0.934 21.05 0.539 3.24 364.73 19.10 15.589 0.972 0.892
GA-1211 8.648 70.910 3654 Roche moutonne e 3 0.973 22.33 0.543 3.83 388.66 14.74 15.664 0.794 0.650
GA-1212 8.646 70.911 3737 Roche moutonne e 3.5 0.829 19.73 0.553 2.64 374.10 18.62 16.870 1.015 0.926
GA-1213 8.641 70.916 3810 Roche moutonne e 2 0.955 23.85 0.569 4.21 444.47 15.66 16.658 0.812 0.647
GA-1214 8.648 70.915 3884 Polished rock (valley-side 5 0.999 25.21 0.571 4.80 467.38 16.45 16.587 0.808 0.644
L wall)
GA-1215 8.648 70.916 3870 Polished rock (valley-side 5 0.985 24.59 0.567 4.21 440.599 15.77 16.059 0.789 0.632
L wall)
GA-1216 8.647 70.915 3840 Polished rock (valley-side 2.3 0.523 13.22 0.572 1.88 363.55 24.36 23.337 1.757 1.790
L wall)
GA-1217 8.647 70.915 3820 Polished rock (valley-side 3.5 0.520 12.89 0.565 2.70 359.80 21.77 23.633 1.643 1.640
L wall)
GA-1218 8.644 70.914 3805 Polished rock (valley-side 2.5 0.524 13.00 0.566 4.44 426.25 14.71 27.221 1.316 1.101
L wall)
GA- 8.646 70.926 3909 Roche moutonne e 2 0.958 25.08 0.583 4.48 438.48 13.71 15.740 0.723 0.540
1301FE
GA- 8.643 70.930 3964 e
Roche moutonne 3 0.922 24.59 0.588 4.46 459.75 14.62 16.717 0.774 0.587
1302FE
GA- 8.641 70.930 3929 e
Roche moutonne 4.5 0.922 23.88 0.579 3.75 37.93 20.96 14.122 0.928 0.865
1303FE
GA-1301 8.647 70.924 3930 Roche moutonnee 4 0.995 25.88 0.580 4.23 461.34 14.75 16.032 0.744 0.563
GA-1302 8.649 70.923 3964 Roche moutonnee 1 0.973 26.38 0.597 2.23 469.50 16.66 16.029 0.784 0.625
GA-1303 8.655 70.927 4150 Roche moutonnee 5 0.992 28.36 0.611 6.13 507.51 15.91 16.071 0.739 0.555
GA-1304 8.626 70.933 4208 Striated rock 5 0.984 28.88 0.619 5.30 499.87 15.51 15.549 0.713 0.531
GA-1305 8.625 70.933 4197 Striated rock 2 0.985 29.49 0.628 6.26 509.09 15.89 15.699 0.713 0.533
GA-1306 8.631 70.924 3945 Roche moutonnee 2.5 0.982 26.06 0.587 4.72 454.03 17.19 15.811 0.795 0.652
a
The topographic scaling factor has been calculated following the method of Dunne et al. (1999).
b
AMS analyses have been carried out at the French AMS facility ASTER.
c
Beryllium 10 concentrations were calibrated against NIST Standard Reference Material 4325 using its certified 10Be/9Be ratio of 2.79 1011 and a 10Be half life of 1.387 ± 0.012 106 yr (Korschinek et al., 2009; Chmeleff et al.,
2010). Results have been corrected from the chemical blank (10Be/9Beblank ¼ 2.625 ± 0.32  1015). Propagated uncertainties include counting statistics, a conservative estimate of 1% for instrumental variability, the uncertainty
of the standard deviation and chemical blank.
d
Ages have been computed with the Cronus Calculator (Balco et al., 2008) using the time-dependent production rate of Lal (1991) modified by Stone (2000).
e
Internal uncertainties consider the analytical uncertainties, while the external uncertainties include 3.4% uncertainty in the production rate (Kaplan et al., 2011) and 8% uncertainty in the 10Be decay constant. In the Results
section, results are presented with the external uncertainties. No correction for snow or other coverage have been taken into account, the sample thickness correction has been calculated with a 2.7 density factor.
I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247 241

10
Fig. 3. Exposure ages of the Gavidia valley determined by the in-situ Be dating (Google earth, 2014). * Identifies samples collected on the valley side-walls (see text for details).

(Fig. 3). This is because of not significant differences of exposure Inside a valley, a topographical feature such as valley bottom
ages at elevations > ~3500 m a.s.l. until the glacier cirque is slope could play a role in the deglaciation history. In a flat area
observed. larger glaciers grow whereas non plateau topography will restrict
Glacier cirques nearby the Gavidia valley at elevations >3850 m glaciation to smaller glaciers (Sugden and John, 1976). In addition,
a.s.l. have similar aspect and average elevation (related to samples in flat areas, deglaciation implies bigger ice free valley extensions.
GA-1302 FE and GA-1303 FE, Fig. 3, Table 1). Exposure ages of this The fastest deglaciation in the Gavidia valley could be also related to
area have not significant differences compared to the Gavidia valley the low slopes of the valley bottom (~5e7 ). In a flatter area any
high part results (related to samples GA-1304, GA-1305 and GA- climatic change providing deglaciation would significantly impact
1306, Fig. 3, Table 1). Therefore, deglaciation of glacier cirques larger areas, as probably occurred in the Gavidia valley.
nearby the Gavidia valley at elevations >3850 m a.s.l. could have
occurred during the same time at ~16.0 ka. Factors driving the 5.2. Global, regional and local paleoclimate data related to the
deglaciation in the Gavidia valley could have influenced in the same Gavidia deglaciation history
way in this area, because of similar orientation and average
elevation of these nearby glacier cirques. In this section, a comparison between the Gavidia valley expo-
Regarding the ice retreat, deglaciation happened in two periods. sure ages and global (i.e. d18O values in the GISP2; Stuiver et al.,
The first one happened from ~21 ka to ~16.5 ka and yields an 1995), regional (i.e. Sea Surface Temperatures -SST- deduced from
average glacier retreat rate of ~0.26 km/ka. The second one Mg/Ca in sediment of the Cariaco Basin; Lea et al., 2003) and local
happened after ~16.5 ka and yields an average retreat rate of ~4.70 paleoclimate proxies records is made. Local paleoclimate proxies
km/ka. A significant extension of the Gavidia valley (~5.5 km) was records used were obtained from studies of fluvioglacial terraces in
ice free in a period shorter than ~ 0.5 ky. the Pa ramo de La Culata, Mucubají or Miranda (Salgado-Laboriau
Side-wall valley samples (GA-1208L and GA-1216L) indicate a and Schubert, 1976; Salgado-Labouriau et al., 1977, 1988; Maha-
glacier maximum thickness ranging between 50 m and 100 m. ney et al., 2008). Also local paleoclimate records used were ob-
These values are similar to thickness estimations in the Mucubají tained from studies of glaciofluvial or glaciolacustrine sediments
valley post LGM (Carcaillet et al., 2013). Based on thickness esti- from Mesa del Caballo or El Pedregal fan complex (Schubert and
mations from samples GA-1208L, GA-1216L, GA-1209 and GA-1212, Rinaldi, 1987; Rull, 1998, 2005; Dirszowsky et al., 2005; Mahaney
vertical ice thinning rates can be estimated. Thinning rates range et al., 2010). Finally, local paleoclimate records used for compari-
between 7 m/ka to 18 m/ka (Fig. 2). However, these preliminary son were obtained from studies of paleosols (Mahaney et al., 2007)
results need to be recalculated with more data. Thinning rates and lake sediments from the Sierra Nevada, Pa ramos Piedras
compare to horizontal ice retreat rates is less significant. Blancas and Agua Blanca (Rull et al., 2005; Stansell et al., 2005;
242 I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247

Polissar et al., 2006; Rull et al., 2010; Stansell et al., 2014). paleoclimate records are needed.
The maximum extension of the Gavidia former glacier (at ~21
ka) is related to the cold temperatures in the Nord Hemisphere 5.3. Deglaciation histories and chronologies in the M
erida Andes
based on more negative values of d18O (Stuiver et al., 1995). It is also mainly based on the 10Be nuclide dating and their potential driving
related to low SST in the North of South America (Fig. 6) (Lea et al., factors
2003). At local scale, palynological analysis of Mesa del Caballo
section confirms cold temperatures during the LGM between 22.75 5.3.1. Deglaciation histories and chronologies in the M
erida Andes
ka and 19.96 ka (Schubert and Rinaldi, 1987). Local LGM tempera- mainly based on the 10Be nuclide dating
ture was at least 8.8 ± 2  C cooler than today and this is the lowest Thirty-seven exposure ages have been recently published
one recorded in the MA (Stansell et al., 2007). (Table 2). We recalculate these exposure ages with the same 10Be
A glacier grows or advances because of positive mass balances production rate of 3.81 ± 0.13 at g1 yr1 (Kaplan et al., 2011) and
and it is directly related to climate (Bennett and Glasser, 2009). the same scaling scheme (Lal, 1991; Stone, 2000). The recalculation
Coldest temperatures that prevailed in the central MA during the of original published exposure ages with the production rate of
LGM forced glaciers growth (Stansell et al., 2007). This assumption Kaplan et al. (2011) provide updated values. For the Mucubají val-
is because of paleoenvironmental records indicate drier climate ley, updated values are ~15% older than original ones (Table 2).
conditions during the LGM (Bradbury et al., 1981; Bradley et al., Concerning La Culata moraine, updated values are ~10% older than
1985; Weingarten et al., 1991; Salgado-Labouriau et al., 1992). the original dataset of Wesnousky et al. (2012).
Coldest temperatures in the central MA seem to also allow the Updated published and unpublished (this study) exposure ages
maximum former glacier growth in the Gavidia valley at ~21 ka. of glacial landforms from the MA range between 10.95 ± 0.97 ka
These cold temperatures probably prevailed until younger ages and and 23.63 ± 1.64 ka (Figs. 4 and 5, Tables 1 and 2). Deglaciation of
it contributed to the lowest ice retreat rates between ~21 ka and these glacial landforms happened during the Late Pleistocene,
~16.5 ka (~0.26 km/ka). However, In the Gavidia valley or nearby specifically, during the Late Me rida Glaciation.
valleys, is not possible to corroborate cold temperatures between The youngest and oldest deglaciation ages are obtained from the
21 and 16.5 ka because there are no available paleoclimate records. Sierra Nevada glacial landforms (10.52 ± 0.48 ka and 23.63 ± 1.64
The Oldest Dryas (OtD) was established in the North Hemi- ka) (modified exposure ages from Wesnousky et al., 2012; Carcaillet
sphere as a cold period ranging between ~17.50 and ~14.60 cal ka BP et al., 2013 and this study). The glacial landform with the youngest
(Blunier et al., 1998). In the North Hemisphere other short cold deglaciation age is located in the Mucubají valley at 4067 m a.s.l.
events are recognized. These short cold events are named the whereas the oldest one is located in the Gavidia valley at 3805 m
Heinrich events. Heinrich event 1 (H1) is established at ~16.8 ka a.s.l.
(Hemming, 2004) inside the OtD period. The OtD may have affected Because some time lag could happen between a glacier advance
northern South America between 17.50 and 17.00 cal ka BP based and deglaciation, we have made the hypothesis that the age of the
on the SST record (Lea et al., 2003). In the MA the OtD and the H1 glacier advance is older than the age of abandonment of the ter-
could be related to a local cold period identified at 16.5 ± 0.3 ka minal or lateral-terminal moraine. However, considering the sen-
named El Caballo Stadial (Rull, 1998). During El Caballo Stadial sitive response of tropical glaciers to climate changes (Kaser and
temperature was 7  C lower than the present value (Fig. 6). This Osmaston, 2002), minimum time lag between glacier advance
cold period was identified as the last cold event recorded by the and landform deglaciation are expected.
Gavidia former glacier (Fig. 6). Here, former glacier drastically re- In the exposure ages discussions is important to keep in mind
treats even before the end of the OtD. This drastic retreat could be that a moraine boulder dated on the crest top of a terminal moraine
actually related to the end of the El Caballo Stadial or the H1. represents a single exposure history of each individual boulder
Paleo-temperature or paleo-precipitation or both variables (Ivy-Ochs et al., 2007). The integrated population of ages, however,
seem have drastically changed in the Gavidia valley at ~16.5e16.0 indicates the average age of the moraine deglaciation (Putkonen
ka close to the end of the El Caballo Stadial related to the H1 and the and Swanson, 2003).
OtD. Here, it is important to note that in the MA, careful attention is In the Gavidia valley, the last glacier maximum extension
necessary to have to relate a short paleoclimate event and exposure happened between ~21 ka and ~16.5 ka. Meanwhile, completely
ages. For example, exposure ages of Los Zerpa terminal moraine valley deglaciation happened at around 16.0 ka. To the NE of the
were originally related to the Younger Dryas (YD) (Carcaillet et al., Gavidia valley, in the Mucubají valley, a recent geomorphological
2013). Whereas a weighted updated exposure age for Los Zerpa is interpretation indicates that the maximum glacier advance is
17.57 ± 0.52 ka (Table 2). In this case, it appears more related to the related to a damaged frontal moraine at elevations lower than
OtD. These different interpretations are due to the 10Be production 3589 m a.s.l. (Angel, 2016). Lowest frontal moraines dated based on
rate lack in the MA and even, until recently, in the Tropical Andes in-situ 10Be dating are located at 3589 m a.s.l. and 3620 m a.s.l.
(high elevations low latitudes) (e.g. Wesnousky et al., 2012; (Carcaillet et al., 2013). Exposure ages of these moraines are
Carcaillet et al., 2013). However, recently 10Be production rates 20.66 ± 1.79 ka and 19.09 ± 0.92 ka (MU09-01 and MU09-02
have been determined in the Tropical Andes (e.g. Blard et al., 2013; Table 2) (modified ages from Carcaillet et al., 2013). Carcaillet
Kelly et al., 2013). The 10Be production rate used in this study (from et al. (2013) dated moraines agree with a previous outwash fan
Kaplan et al., 2011) is not significantly different from that of Kelly radiocarbon dated located at 3400 m a.s.l. This outwash fan is
et al. (2013). So, a more recent interpretation on exposure ages in located down valley of MU09-02 and has a basal age of 19.08 ± 0.82
14
the MA will probably not significantly change. C kyr BP (Schubert and Rinaldi, 1987). These chronologies of
Paleo-temperature and paleo-precipitation changes in the glacial landforms suggest the maximum glacier advance seems not
Gavidia valley provided negative mass balances. These negative to be much older than around 20 ka. It occurred maybe during the
mass balances produced the fastest former glacier deglaciation. The same local LGM (defined between 22.75 and 19.96 Cal ka BP,
deglaciation in other areas of the central MA at ~16.5 ka, specifically Schubert and Rinaldi, 1987).
in La Culata, Sierra del Norte, has been related to paleo- The deglaciation history of the Mucubají valley at ages between
precipitation changes (Wesnousky et al., 2012). However, to make the LGM-Late Glacial (LG) involves multiple less significant glacier
a more accurate interpretation about the paleo-parameters con- readvances (in the range 18e12 ka). This assumption is based on
trolling negative mass balances in the Gavidia valley, more local different moraines located along the valley between 3572 m a.s.l.
I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247 243

Table 2
10 rida Andes (Wesnousky et al., 2012; Carcaillet et al., 2013).
Be dating inventory of the Me

Samples Latitude Longitude Elevation Sample information Original ages (ka) Modified ages (ka)(a) Reference

Value External uncertainty Internal uncertainty


 
N W m Sampled landforms Location ka ka

VEN1 8.7601 71.0516 3391 Boulder moraine La Culata 15.9 ± 0.8 17.06 1.00 0.90 Wesnousky et al., 2012
VEN3 8.7654 71.0477 3457 Boulder moraine La Culata 15.4 ± 0.7 16.61 0.89 0.76 Wesnousky et al., 2012
VEN4 8.7656 71.0476 3458 Boulder moraine La Culata 14.2 ± 0.4 15.38 0.65 0.43 Wesnousky et al., 2012
VEN5 8.7663 71.0471 3467 Boulder moraine La Culata 14.4 ± 0.6 15.63 0.82 0.76 Wesnousky et al., 2012
VEN6 8.7689 71.0457 3508 Boulder moraine La Culata 16.2 ± 0.6 17.39 0.88 0.73 Wesnousky et al., 2012
VEN7 8.7697 71.0467 3472 Boulder moraine La Culata 14.3 ± 0.7 15.52 0.91 0.81 Wesnousky et al., 2012
VEN8 8.7697 71.0466 3472 Boulder moraine La Culata 13.1 ± 0.6 14.27 0.82 0.72 Wesnousky et al., 2012
VEN9 8.7705 71.0466 3477 Boulder moraine La Culata 13.2 ± 0.5 14.38 0.75 0.62 Wesnousky et al., 2012
VEN11 8.7721 71.0464 3501 Boulder moraine La Culata 15.4 ± 0.7 16.61 0.93 0.82 Wesnousky et al., 2012
VEN12 8.7726 71.0463 3500 Boulder moraine La Culata 14.7 ± 0.7 15.96 0.94 0.85 Wesnousky et al., 2012
VEN13 8.7729 71.0393 3657 Boulder moraine La Culata 14.2 ± 0.5 15.38 0.76 0.61 Wesnousky et al., 2012
VEN14 8.7729 71.0395 3653 Boulder moraine La Culata 16.3 ± 1.2 17.52 1.42 1.43 Wesnousky et al., 2012
VEN15 8.7727 71.0400 3651 Boulder moraine La Culata 14.7 ± 0.7 15.89 0.96 0.87 Wesnousky et al., 2012
VEN19 8.8141 70.8006 3255 Boulder moraine La Victoria 18.6 ± 3.7 19.75 4.00 4.38 Wesnousky et al., 2012
VEN20 8.8142 70.8006 3258 Boulder moraine La Victoria 16.9 ± 1.0 18.07 1.17 1.10 Wesnousky et al., 2012
VEN21 8.8142 70.8010 3260 Boulder moraine La Victoria 15.3 ± 0.6 16.49 0.86 0.73 Wesnousky et al., 2012
VEN23 8.8139 70.7993 3243 Boulder moraine La Victoria 15.1 ± 0.5 16.37 0.76 0.58 Wesnousky et al., 2012
VEN25 8.8121 70.7881 3115 Boulder moraine Los Zerpa 17.7 ± 1.4 18.92 1.60 1.63 Wesnousky et al., 2012
VEN26 8.8120 70.7873 3104 Boulder moraine Los Zerpa 15.0 ± 1.6 16.20 1.74 1.80 Wesnousky et al., 2012
VEN27 8.8117 70.7875 3105 Boulder moraine Los Zerpa 17.8 ± 0.5 18.92 0.85 0.63 Wesnousky et al., 2012
VEN28 8.8118 70.7873 3106 Boulder moraine Los Zerpa 16.9 ± 0.9 18.20 1.10 1.01 Wesnousky et al., 2012
LZ09-01 8.8117 70.7884 3127 Boulder moraine Los Zerpa 13.84 ± 1.74 15.78 1.56 1.60 Carcaillet et al., 2013
LZ09-02 8.8117 70.7874 3113 Boulder moraine Los Zerpa 12.48 ± 1.39 14.33 1.14 1.17 Carcaillet et al., 2013
Mu09- 8.8009 70.8279 3620 Boulder frontal moraine Mucubají 16.78 ± 1.54 19.09 0.92 0.74 Carcaillet et al., 2013
01
Mu09- 8.7954 70.8343 3589 Boulder frontal moraine Mucubají 18.14 ± 2.11 20.66 1.79 1.85 Carcaillet et al., 2013
02
Mu09- 8.7951 70.8267 3572 Boulder frontal moraine Mucubají 15.66 ± 1.67 17.79 1.31 1.28 Carcaillet et al., 2013
03
Mu09- 8.7874 70.8233 3607 Boulder frontal moraine Mucubají 13.27 ± 1.20 15.19 0.69 0.51 Carcaillet et al., 2013
04
Mu09- 8.7850 70.8229 3615 Boulder frontal moraine Mucubají 13.32 ± 1.48 15.24 1.22 1.20 Carcaillet et al., 2013
05
Mu09- 8.7852 70.8224 3620 Boulder frontal moraine Mucubají 15.96 ± 1.80 18.14 1.49 1.50 Carcaillet et al., 2013
06
Mu09- 8.7790 70.8197 3697 Striated bedrock Mucubají 13.84 ± 1.32 15.78 0.88 0.76 Carcaillet et al., 2013
07
Mu09- 8.7785 70.8189 3727 Striated bedrock Mucubají 14.20 ± 1.34 16.17 0.86 0.73 Carcaillet et al., 2013
08
Mu09- 8.7667 70.8129 4067 Striated bedrock Mucubají 9.08 ± 0.82 10.52 0.48 0.36 Carcaillet et al., 2013
10
Mu09- 8.7633 70.8119 4213 Boulder moraine Mucubají 10.63 ± 0.96 12.30 0.56 0.41 Carcaillet et al., 2013
11
Mu09- 8.7659 70.8121 4091 Striated bedrock Mucubají 9.48 ± 1.11 10.95 0.97 0.97 Carcaillet et al., 2013
12
Mu09- 8.7689 70.8164 3982 Moraine boulder Mucubají 9.73 ± 0.88 11.23 0.52 0.40 Carcaillet et al., 2013
13
Mu09- 8.7719 70.8152 3862 Moraine boulder Mucubají 9.93 ± 1.16 11.46 1.01 1.01 Carcaillet et al., 2013
14
Mu09- 8.7758 70.8161 3804 Striated bedrock Mucubají 12.86 ± 1.57 14.75 1.38 1.40 Carcaillet et al., 2013
15
a
Modified ages correspond to recalculation using Lago Argentino production rate (Kaplan et al., 2011) and the scaling scheme model of Lal (1991) modified by Stone (2000).

and 4200 m a.s.l. (Table 2). Multiple readvances involved the former In the Mucubají valley the glacier retreated in two distinct pe-
glacier equilibrium with climate conditions characterized by cold riods (Carcaillet et al., 2013). The oldest one occurred between the
temperatures or high precipitations which provided more glacier LGM and the Late Glacial (~15 ka) with ~25 m/ka as ice retreat rate
mass accumulation. (Carcaillet et al., 2013). The second one occurred between 15 ka and
TCN exposure ages from the Mucubají valley provide a degla- the Early Holocene (~11 ka) with ~310 m/ka as ice retreat rate. In
ciation history until the Pleistocene/Holocene transition (at ~ 11 ka; addition, Late Pleistocene/Early Holocene climate change seems to
Table 2). The Holocene activity of the Mucubají glacier seems to prevent glacier development and produced a faster glacier retreat
have occurred at least until 6.25 cal kyr BP (5.46 ± 0.40 14C kyr BP ka (Carcaillet et al., 2013).
from aquatic moss, Stansell et al., 2005). This glacier activity was At the NE of the Mucubají valley, at La Victoria terminal moraine
inferred from the Laguna de Mucubají sediment descriptions, (located at ~ 3200 m a.s.l.), deglaciation began at 17.16 ± 0.71 ka
characterized by low clastic content and low magnetic suscepti- (weighted average) (Table 2). This deglaciation age is similar to Los
bility (Stansell et al., 2005; Carrillo, 2006; Carrillo et al., 2008). A Zerpa terminal moraine which is located beside La Victoria at
more recent glacier activity seems to have occurred between around 3100 m a.s.l. (17.57 ± 0.52 ka weighted average, modified
(0.64 ± 0.05 14C kyr BP) 0.61 Cal. kyr BP and 0.13 Cal. kyr BP (Stansell from Wesnousky et al., 2012; Carcaillet et al., 2013).
et al., 2005). In Sierra del Norte at La Culata moraine (located
244 I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247

rida Andes determined by the in-situ 10Be. All the exposure ages were recalculated with the Lago Argentino 10Be production rate (Kaplan
Fig. 4. Published exposure ages in the Me
et al., 2011) from previous works: ages with * correspond to Wesnousky et al. (2012) and ** correspond to Carcaillet et al. (2013).

moraine was interpreted as a small LG readvance related to the


4200
weighted average exposure age of 14.1 ± 1.0 ka (Wesnousky et al.,
2012).

5.3.2. Potential factors driving different deglaciation histories and


4000
chronologies in the M erida Andes
Differences in climate, local environmental conditions or topo-
graphic features of glacier catchments and former glacier features
3800 (e.g. steepness, aspect, cloudiness, speed, insolation, hypsometry,
precipitation, temperatures, etc.) could explain different former
Elevation (m)

glaciers deglaciation histories in an area (Paterson, 1994 in Bennett


3600 and Glasser, 2009; Barr and Lovell, 2014). Different deglaciation
histories or chronologies in the northern Sierra de Santo Domingo
during the LGM (including Mucubají, Victoria and Zerpa glaciers)
were mainly driven by the coldest temperatures (Stansell et al.,
PLEISTOCENE

3400
2007). In addition, the gradient in ELA analysis suggests also
HOLOCENE

strong influence of local conditions, such as precipitation, cloud


cover and valley aspect (Stansell et al., 2007).
3200 Favorable climate conditions during the LGM contributed to the
maximum Gavidia former glacier extension at around 21 ka and the
glacier maximum advance in the Mucubají valley. In the Mucubají
3000
valley favorable intermittent climate conditions led to positive
0 5 10 15 20 25 30 mass balances between the LGM and LG (~18e12 ka). These posi-
Ages (ka) tive mass balances provided multiple former glacier readvances at
Fig. 5. Altitudinal distribution of the exposure ages in the Merida Andes obtained by Mucubají in contrast to the Gavidia former glacier which
the in-situ 10Be dating (This study, Wesnousky et al., 2012; Carcaillet et al., 2013). Blue completely disappeared at ~16.0 ka.
and red diamonds are dating of the western side of the study area (Gavidia and La Different deglaciation histories in the MA could have been also
Culata). Orange, yellow and grey dots are dating of the eastern part (Mucubají, La driven by different topographic features of the valleys. Gavidia
Victoria and Los Zerpa). (For interpretation of the references to colour in this figure
valley is separated from the Chama valley by a narrow deep gorge
legend, the reader is referred to the web version of this article.)
whereas the Culata-Chama and Mucubají-Santo Domingo conflu-
ences are wider (Fig. 4). These different topographical features
at ~ 3400e3500 m a.s.l.), deglaciation began at ~16.5 ka based on probably confer different atmospheric circulation which controlled
recalculated exposure ages from Wesnousky et al. (2012). In addi- humidity, air temperature and rainfall.
tion, authors infer a readvance based on a minor moraine that inset In addition, the Mucubají valley bottom profile is characterized
into the former laterofrontal moraine of La Culata. This minor by a lower flat area (3550e3700 m a.s.l. with slopes ~5 ) and an
I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247 245

A YD OD OtD LGM whereas accumulation area in the Mucubají valley is NW-SE


oriented.
4200
6. Conclusions

4000 The first deglaciation chronology of the Gavidia valley is based


on TCN dating (10Be in-situ) and a detailed analysis of glacial
landforms. Deglaciation happened in two separated periods. The
first one occurred between the LGM (21 ka) and the Oldest Dryas
3800
Elevation (m)

Stadial (OtD at ~16.5 ka) - El Caballo Stadial. During this period, cold
climate conditions seem to be maintained and led a low ice retreat
rate (~0.26 km/ka). In contrast, the second shorter period occurred
3600 after the El Caballo Stadial (at ~16.5 ka) with the fastest ice retreat
rate (~4.7 km/ka) followed by glacier full extinction. The fastest
glacier extinction at ~16.0 ka was produced by changes in paleo-
3400 climate conditions at the end of the El Caballo Stadial connected to
the low valley bottom slopes (around 5-7 ).
In addition, different deglaciation histories were evidenced in
3200 the Me rida Andes. Local paleotemperatures and paleo-
precipitations contrasts, different valleys aspects and steepness
could account for different deglaciation histories.
0 5 10 15 20 25
B -33

-35 Acknowledgements

-37
δ18O0/00

This research was founded by INSU-CNRS-IRD (France). FUN-


VISIS (Project Geodin amica Integral de los Andes de Me rida,
-39
GIAME-FONACIT 2012002202), CIGIR (Venezuela) and FONACIT
-41 (Venezuela). 10Be analyses were performed at the French Acceler-
ator Mass Spectrometry Facility, ASTER (INSU-CNRS). We thank
-43
28 Maurice Arnold, Georges Aumaitre and Karim Keddadouche for
C assistance during AMS measurements. We thank Eduardo Barreto
27
for his valuable help during fieldwork, Francis Coeur for rock
SST (°C)

26 crushing and Alisse Geigher for paleoglaciological discussions.


25 Finally, we thank the staff of both Santo Domingo Hotel (Santo
Domingo) and Mucoposada Michicaba (Gavidia). Thanks to anon-
24
ymous reviewers for their comments, which significantly improve
D 25 this contribution.
2 24
Estimated ΔT (°C)

0 References

-2 Angel, I., Carrillo, E., Carcaillet, J., Audemard, F.A., Beck, C., 2013. Geocronología con
 topo cosmoge
el iso nicos 10Be, aplicacio n para el estudio de la din
amica glaciar
-4 cuaternaria en la regio  n central de los Andes de Me rida. GEOS 44, 73e82.
Angel, I., 2016. Late Pleistocene Deglaciation Histories in the Central Me rida Andes
-6 (Venezuela). PhD. Sc Thesis. Universite  de Grenoble Alpes-Universidad Central
de Venezuela, Francia, Venezuela, 234 pp.
Audemard, F.A., 2003. Geomorphic and geologic evidence of ongoing uplift and
0 5 10 15 20 25 deformation in the Me rida Andes, Venezuela. Quat. Int. 101e102, 43e65.
Ages (ka)
Audemard, F.E., Audemard, F.A., 2002. Structure of the Me rida Andes, Venezuela:
relations with the South America-Caribbean geodynamic interaction. Tectono-
Fig. 6. A) Altitudinal distribution of the Gavidia exposure ages versus global (B),
physics 345, 299e327.
regional (C) and local (D) paleoclimatic proxy records. Dots are samples collected on Audemard, F.A., Pantosti, D., Machette, M., Costa, C., Okumura, K., Cowan, H.,
the valley bottom while diamonds are samples of the valley side-walls. B) GISP2 d18O Diederix, H., Ferrer, C., Sawop Participants, 1999. Trench investigation along the
values (‰) (Stuiver et al., 1995). C) Sea Surface Temperature (SST) deduced from Mg/Ca Me rida section of the Bocono  fault (central Venezuelan Andes), Venezuela. In:
sediment record of the Cariaco Basin (Lea et al., 2003). D) Noticeable climatic events in Pavlides, S., Pantosti, D., Peizhen, Z. (Eds.), Earthquakes, Paleoseismology and
the Me rida Andes deduced from palynological investigations (Based on Salgado- Active Tectonics. Selected Papers to 29th General Assembly of the Association of
Laboriau and Schubert, 1976; Salgado-Labouriau et al., 1977; Schubert and Rinaldi, Seismology and Physics of the Earth's Interior (IASPEI), Thessaloniki, Greece,
1987; Salgado-Laboriau et al., 1988; Rull, 1998; Dirszowsky et al., 2005). Cold periods August 1997. Tectonophysics, 308, pp. 1e21.
based on GISP2 curves are in grey bands. YD: Younger Dryas, OD: Older Dryas and OtD: Balco, G., 2011. Contributions and unrealized potential contributions of cosmogenic
Oldest Dryas. nuclide exposure dating to glacier chronology, 1990e2010. Quat. Sci. Rev. 30,
3e27.
Balco, G., Stone, J.O., Lifton, N.A., Dunai, T.J., 2008. A complete and easily accessible
means of calculation surface exposure ages or erosion rates from 10Be and 26Al
upper steep area (3700e4600 m a.s.l. with slopes ~20 ). Whereas measurements. Quat. Geochronol. 3, 174e195.
Barr, I.D., Lovell, H., 2014. A review of topographic controls on moraine distribution.
the Gavidia valley bottom profile is not steep, it is uniform and
Geomorphology 226, 44e64.
characterized by slopes ~ 5-7. A lower average slope in the Gavidia Bermúdez, M., 2009. Cenozoic Exhumation Patterns across the Venezuelan Andes:
valley contributed to the fastest glacier withdrawal. Insights from Fission-track Thermochronology. Ph.D. Thesis. Universite  Joseph
Insolation can be another strong playing factor driving different Fourier, France, 305 pp.
Benn, D.I., Owen, L.A., Osmaston, H.A., Seltzer, G.O., Porter, S.C., Mark, B., 2005.
deglaciation histories because of different accumulation areas as- Reconstruction of equilibrium-line altitudes for tropical and sub-tropical gla-
pects. Accumulation area in the Gavidia valley is NE-SW oriented ciers. Quat. Int. 138e139, 8e21.
246 I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247

Bennett, M., Glasser, N., 2009. Glacial Geology Ice Sheets and Landforms, second ed. 64, 265e280.
Wiley-Blackwell. 385 pp. Jahn, A., 1931. Los p aramos venezolanos. Bol. la Soc. Venez. Ciencias Nat. 1 (3),
Benn, D.I., Evans, D.J.A., 2010. Glaciers and Glaciation, second ed. Hodder Education, 93e132.
Abingdon, United Kingdom, p. 802. Kaplan, M.R., Strelin, J.A., Schaefer, J.M., Denton, G.H., Finkel, R.C., Schwartz, R.,
Bezada, M., 1989. Geología Glacial del Cuaternario de la regio  n de Santo Domingo Putnam, A.E., Vandergoes, M.J., Goehring, B.M., Travis, S.G., 2011. In-situ
-Pueblo Llano - Las Mesitas (Estados Me rida y Trujillo). PhD. Sc Thesis. Instituto cosmogenic 10Be production rate at Lago Argentino, Patagonia: implications for
Venezolano de Investigaciones Científicas, Venezuela, 245 pp. late-glacial climate chronology. Earth Planet. Sci. Lett. 309, 21e32.
Blard, P.H., Braucher, R., Lave , J., Bourle
s, D., 2013. Cosmogenic 10Be production rate Kaser, G., Osmaston, H., 2002. Tropical Glaciers. Cambridge University Press, Cam-
calibrated against 3He in the high Tropical Andes (3800e4900 m, 20e22 S). bridge, 207 pp.
Earth Planet. Sci. Lett. 382, 140e149. Kelly, M.A., Lowell, T.V., Applegate, P.J., Phillips, F.M., Schaefer, J.M., Smith, C.A.,
Bradbury, J.P., Leyden, B., Salgado-Labouriau, M., Lewis Jr., W.M., Schubert, C., Kim, H., Leonard, K.C., 2013. A locally calibrated, late glacial 10Be production rate
Binford, M.W., Frey, D.G., Whitehead, D.R., Weibezahn, F.H., 1981. Late quater- from low-latitude, high-altitude site in the Peruvian Andes. Quat. Geochronol.
nary environmental history of Lake Valencia, Venezuela. Science 214, 1e16 (in press), Corrected Proof.
1299e1305. Kirkbride, M.P., Brazier, V., 1998. A critical evaluation of the use of glacial chro-
Bradley, R.S., Yuretich, R., Salgado-Labouriau, M.L., Weingarten, B., 1985. Late Qua- nologies in climatic reconstruction, with reference to New Zealand. Quat. Proc.
ternary paleoenvironmental reconstruction using lake sediments from the 6, 55e64.
Venezuelan Andes: preliminary results. Z. für Gletscherkd. Glazialgeol. 21, Korschinek, G., Bergmaier, A., Faestermann, T., Gerstmann, U.C., Knie, K., Rugel, G.M.,
97e106. Poutivtsev, M., Remmert, A., 2009. A new value for the half-life of 10Be by
Braun, C., Bezada, M., 2013. The history and disappearance of glacier in Venezuela. heavy-ion elastic recoil detection and liquid scintillation counting. Nucl. Ins-
J. Lat. Am. Geogr. 12, 85e124. trum. Methods Phys. Res. B 268 (2), 187e191.
Blunier, T., Chappellaz, J., Schwander, J., Dallenbach, A., Stauffer, B., Stocker, T.F., Lal, D., 1991. Cosmic ray labeling of erosion surfaces: in situ nuclide production rates
Raynaud, D., Jouzel, J., Clausen, H.B., Hammer, C.U., Johnsen, S.J., 1998. Asyn- and erosion models. Earth Planet. Sci. Lett. 104, 429e439.
chrony of Antarctic and Greenland climate change during the last glacial period. Lea, D.W., Pak, D.K., Peterson, L.C., Hughen, K.A., 2003. Synchronicity of tropical and
Nature 394, 739e743. high-latitude Atlantic temperatures over the last glacial termination. Science
Briner, J.P., Swanson, T., 1998. Using inherited cosmogenic 36Cl to constrain glacial 301, 1361e1364.
erosion rates of the Cordilleran ice sheet. Geology 26, 3e6. Mahaney, W.C., Milner, M.W., Kalm, V., Dirszowsky, R.W., Hancock, R.G.V.,
Brown, E.T., Edmond, J.M., Raisbeck, G.M., Yiou, F., Kurz, M.D., Brook, E.J., 1991. Beukens, R.P., 2008. Evidence for a younger Dryas glacial advance in the Andes
Examination of surface exposure ages of Antartic moraines using in situ pro- of northwestern Venezuela. Geomorphology 96, 199e211.
duced 10Be and 26Al. Geochim. Cosmochim. Acta 55, 2269e2283. Mahaney, W.C., Volli, K., Menzies, J., Milner, M.W., 2010. Reconstruction of the early
Carcaillet, J., Angel, I., Carrillo, E., Audemard, F.A., Beck, C., 2013. Timing of the last Me rida, pre-LGM glaciation with comparison to late glacial maximum till,
deglaciation in the Sierra Nevada of the Me rida Andes, Venezuela. Quat. Res. 80 northwestern Venezuelan Andes. Sediment. Geol. 226, 29e41.
(3), 482e494. Mahaney, W.C., Russell, S.E., Milner, M.W., Kalm, V., Bezada, M., Hancock, R.G.V.,
Carrillo, E., Yepez, S., 2008. Evolucio  n de glaciares en Venezuela: glaciares de los Beukens, R.P., 2001. Paleopedology of middle Wisconsin/Weichselian paleosols
Picos Humboldt y Bonpland. Bol. Geol. 42, 97e108. in the Me rida Andes, Venezuela. Geoderma 104, 215e237.
Carrillo, E., 2006. L'Enregistrement se dimentaire de la sismicite  re
cente le long de la Mahaney, W.C., Dirszowsky, R.W., Milner, M.W., Harmsen, R., Finkelstein, S.A.,
frontiere sudoccidentale de la plaque caraïbe (Faille de Bocono ): Modalites et Kalm, V., Bezada, M., Hancock, R.G.V., 2007. Soil stratigraphy and plantesoil
chronologie. Contribution  a l'estimation de l’ale a sismique re gional. Ph.D. interactions on a late glacialeHolocene fluvial terrace sequence, Sierra Nevada
Thesis. Universite  de Savoie, France, p. 335. National Park, northern Venezuelan Andes. J. S. Am. Earth Sci. 23, 46e60.
Carrillo, E., Beck, C., Audemard, F.A., Moreno, M., Ollarves, R., 2008. Disentangling Mahaney, W.C., Milner, M.W., Voros, J., Kalm, V., Hütt, G., Bezada, M.,
late quaternary climatic and seismo-tectonic controls on lake Mucubají sedi- Hancock, R.G.V., Autreiter, S., 2000. Stratotype for the Me rida glaciation at
mentation (Me rida Andes, Venezuela). Palaeogeogr. Palaeoclimatol. Palaeoecol. Pueblo Llano in the northern Venezuela Andes. J. S. Am. Earth Sci. 13, 761e774.
259, 284e300. Merchel, S., Herpers, U., 1999. An update on radiochemical separation techniques
Chmeleff, J., von Blanckenburg, F., Kossert, K., Jakob, J., 2010. Determination of the for the determination of long-lived radionuclides via Accelerator Mass Spec-
10
Be half-life by multicollector ICP-MS and liquid scintillation counting. Nucl. trometry. Radiochim. Acta 84, 215e219.
Instrum. Methods Phys. Res. B 268 (2), 192e199. Monasterio, M., Reyes, S., 1980. Diversidad ambiental y variacio  n de la vegetacio n en
Dirszowsky, R.W., Mahaney, W.C., Hodder, K.R., Milner, M.W., Kalm, V., Bezada, M., los pa ramos de los Andes Venezolanos. In: Monasterio, M. (Ed.), Estudios
Beukens, R.P., 2005. Lithostratigraphy of the Me rida (Wisconsinan) glaciation Ecolo gicos en los Paramos Andinos. Editorial de la Universidad de Los Andes,
and Pedregal interstade, Me rida Andes, northwestern Venezuela. J. S. Am. Earth Me rida, pp. 47e91.
Sci. 19, 525e536. Monod, B., Dhont, D., Hervoue €t, Y., 2010. Orogenic float of the Venezuelan Andes.
Dunai, T., 2010. Cosmogenic Nuclides: Principles, Concepts and Applications in the Tectonophysics 490, 123e135.
Earth Surface Sciences. Cambridge University Press, 180 pp. Nishiizumi, K., Winterer, E.L., Kohl, J.R., Klein, J., Middleton, R., Lal, D., Arnold, J.R.,
Dunne, J., Elmore, D., Muzikar, P., 1999. Scaling factors for the rates of production of 1989. Cosmic ray production rates of 10Be and 26Al in quartz from glacially
cosmogenic nuclides for geometric shielding and attenuation at depth on polished rocks. J. Geophys. Res. 94, 17907e17915.
sloped surfaces. Geomorphology 27 (1e2), 3e11. Paterson, W.S.B., 1994. The Physics of Glaciers, third ed. Pergamon, Oxford.
Goering, 1962. Venezuela, El Ma s Bello País Tropical. Ediciones de la Universidad de Polissar, P.J., Abbott, M.B., Wolfe, A.P., Bezada, M., Rull, V., Bradley, R.S., 2006. Solar
los Andes, Me rida, 172 pp. modulation of little ice age climate in the tropical Andes. Proc. Natl. Acad. Sci.
Google earth V 7.1.2.2041, August 1, 2014. Gavidia valley, Me rida Andes. 8.665889’N, 103, 8937e8942.
-70.908993’W, Eye Alt 13.31 Km. SIO, NOAA, U.S. Navy, NGA, GEBCO. GeoEye. Polissar, P.J., Abbott, M.B., Wolfe, A.P., Vuille, M., Bezada, M., 2013. Synchronous
http://www.googleearth.com. interhemispheric Holocene climate trends in the tropical Andes. Proc. Natl.
Gosse, J.C., Phillips, F.M., 2001. Terrestrial in situ cosmogenic nuclides: theory and Acad. Sci. 110, 14551e14556.
application. Quat. Sci. Rev. 20, 1475e1560. Poveda, G., Waylen, P.R., Pulwarty, R.S., 2006. Annual and inter-annual variability of
Gosse, J.C., Klein, J., Evenson, E.B., Lawn, B., Middleton, R., 1995. 10Be dating of the the present climate in northern South America and southern Mesoamerica.
duration and retreat of the last Pinedale glacial sequence. Science 268, Palaeogeogr. Palaeoclimatol. Palaeoecol. 234, 3e27.
1329e1333. Pulwarty, R.S., Barry, R.G., Hurst, C.M., Sellinger, K., Mogollon, L.E., 1998. Precipita-
Guido, Z.S., Ward, D.J., Anderson, R.S., 2007. Pacing the post-last glacial maximum tion in the Venezuelan Andes in the Context of regional climate. Meteorol.
demise of the Animas valley glacier and the San Juan Mountain ice cap, Colo- Atmos. Phys. 67, 217e237.
rado. Geology 35, 739e742. Putkonen, J., Swanson, T., 2003. Accuracy of cosmogenic ages for moraines. Quat.
Guzma n, O., Vassallo, R., Audemard, F., Mugnier, J.-L., Oropeza, J., Ye pez, S., Res. 59, 255e261.
Carcaillet, J., Alvarado, M., Carrillo, E., 2013. 10Be dating of river terraces of Santo Royo y Go  mez, J., 1959. El Glaciarismo Pleistoceno en Venezuela. Bol. Inf. Asoc.
Domingo river, on Southeastern flank of the Me rida Andes, Venezuela: tectonic Venez. Geol. Minería Petro leo 2, 333e357.
and climatic implications. J. S. Am. Earth Sci. 48, 85e96. Rull, V., 1998. Palaeoecology of pleniglacial sediments from the Venezuelan Andes.
Hackley, P.C., Urbani, F., Karlsen, A.W., Garrity, C.P., 2005. Geologic Shaded Relief Palynological record of El Caballo stadial, sedimentation rates and glacier
Map of Venezuela. Open-file Report - U. S. Geological Survey. 2005. U. S. retreat. Rev. Palaeobot. Palynology 99, 95e114.
Geological Survey, Reston, VA, United States. Pages: 2 sheets. Rull, V., 2005. A Middle Wisconsin interstadial in the northern Andes. J. S. Am. Earth
Hemming, S.R., 2004. Heinrich events: massive late Pleistocene detritus layers of Sci. 19, 173e179.
the north Atlantic and their global climate imprint. Rev. Geophys. 42 http:// Rull, V., Stansell, N.D., Montoya, E., Bezada, M., Abbott, M.B., 2010. Palynological
dx.doi.org/10.1029/2003RG000128RG1005. signal of the younger Dryas in the tropical Venezuelan Andes. Quat. Sci. Rev. 29,
Hughes, P.D., Gibbard, P.L., Woodward, J.C., 2005. Quaternary glacial records in 3045e3056.
mountain regions: a formal stratigraphical approach. Episodes 28, 85e92. Rull, V., Abbott, M.B., Polissar, P.J., Wolfe, A.P., Bezada, M., Bradley, R.S., 2005. 15,000-
Ivy-Ochs, S., Kerschner, H., Schlüchter, C., 2007. Cosmogenic nuclides and the dating yr pollen record of vegetation change in the high altitude tropical Andes at
of late glacial and early Holocene glacier variations: the Alpine perspective. Laguna Verde Alta, Venezuela. Quat. Res. 64, 308e317.
Quat. Int. 164e165, 53e63. Salgado-Labouriau, M.L., Schubert, C., 1976. Palynology of Holocene peat bogs from
Jahn, A., 1912. La cordillera venezolana de los Andes.: Revista Te cnica del Ministerio central Venezuelan Andes. Palaeogeography, Paleoclimatology. Palaeoecology
de Obras Públicas 2, No. 21, pp. 451e488. 19, 147e156.
Jahn, A., 1925. Observaciones glaciolo  gicas en los Andes Venezolanos. Cult. Venez. Salgado-Labouriau, M.L., Schubert, C., 1977. Pollen analysis of a peat bog from
I. Angel et al. / Journal of South American Earth Sciences 71 (2016) 235e247 247

Laguna Victoria (Venezuelan Andes). Acta Científica Venez. 28, 428e432. 383e396.
Salgado-Labouriau, M.L., Schubert, M.L., Valastro, S.J., 1977. Paleoecologic analysis of Sievers, W., 1885. Über Schneeverha €ltnisse in der Cordillere Venezuelas. Jahres-
a late quaternary terrace from Mucubají, Venezuelan Andes. J. Biogeogr. 4, bericht der Geogr. Gesellschaft München 10, 54e57.
313e325. Smith, J.A., Finkel, R.C., Farber, D.L., Rodbell, D.T., Seltzer, G.O., 2005. Moraine
Salgado-Laboriau, M.L., Rull, V., Schubert, C., Valastro, J.R.S., 1988. The establishment preservation and boulder erosion in the tropical Andes: interpreting old surface
of vegetation after late Pleistocene deglaciation in the p aramo Miranda, Ven- exposure ages in glaciated valleys. J. Quat. Sci. 20 (7e8), 735e758.
ezuelan Andes. Rev. Palaeobot. Palynol. 55, 5e17. Soulas, J.-P., 1985. Neotecto nica del flanco occidental de los Andes de Venezuela
Salgado-Labouriau, M.L., Bradley, R.S., Yuretich, R., Weingarten, B., 1992. Paleoeco- entre 70 300 y 71000 W (Fallas de Bocono , Valera, Pin
~ ango y del Piedemonte). VI
logical analysis of the sediments of lake Mucubají, Venezuelan Andes. Congr. Geol. Venez. 4, 2690e2711. Caracas.
J. Biogeogr. 19, 317e327. Soulas, J.-P., Rojas, C., Schubert, C., 1986. Neotecto nica de las fallas de Bocono ,
Schubert, C., 1970. Glaciation of the Sierra de Santo Domingo, Venezuelan Andes. Valera, Tun ~ ame y Mene Grande. Excursio  n No. 4. VI Congr. Geol. Venez. 10,
Quaternaria 13, 225e246. 6961e6999. Caracas.
Schubert, C., 1972. Geomorphology and glacier retreat in the Pico Bolívar area, Si- Stansell, N.D., Abbott, M.B., Polissar, P.J., Wolfe, A.P., Bezada, M., Rull, V., 2005. Late
erra Nevada de Me rida, Venezuela. Z. für Gletscherkd. Glazialgeol. 8 (1e2), quaternary deglacial history of the Me rida Andes, Venezuela. J. Quat. Sci. 20
189e202. (7e8), 801e812.
Schubert, C., 1974. Late Pleistocene Merida glaciation, Venezuelan Andes. Boreas 3, Stansell, N.D., Polissar, P.J., Abbott, M.B., 2007. Last glacialmaximum equilibrium-
147e151. line altitude and paleo-temperature reconstructions for the Cordillera
Schubert, C., 1980. Contribucio n al inventario mundial de glaciares. Bol. la Soc. deMe rida, Venezuelan Andes. Quatern. Res. 67, 115e127.
Venez. Ciencias Nat. 34 (137), 267e279. Stansell, N.D., Abbott, M.B., Rull, V., Rodbell, D.T., Bezada, M., Montoya, E., 2010.
Schubert, C., 1980a. Morfología neotecto nica de una falla rumbo deslizante e Abrupt Younger Dryas cooling in the northern tropics recorded in lake sedi-
informe preliminar sobre la falla de Bocono , Andes meriden ~ os. Acta Científica ments from the Venezuelan Andes. Earth Planet. Sci. Lett. 293, 154e163.
Venez. 31, 98e111. Stansell, N.D., Polissar, P.J., Abbott, M.B., Bezada, M., Steinman, B.A., Braun, C., 2014.
Schubert, C., 1992. The glaciers of the Sierra Nevada de Me rida (Venezuela): a Proglacial lake sediment records reveal Holocene climate changes in the Ven-
photographic comparison of recent deglaciation. Erkunde 46, 58e64. ezuelan Andes. Quat. Sci. Rev. 89, 44e55.
Schubert, C., 1998. Glaciers of Venezuela. In: Williams, R.S., Ferrigno, J.G. (Eds.), Stone, J.O., 2000. Air pressure and cosmogenic isotope production. J. Geophys. Res.
Satellite Image Atlas of Glaciers of the World. U.S. Geological Survey Profes- 105, 23753e23759.
sional Paper 1386-I, Washington D.C. Stuiver, M., Grootes, P.M., Braziunas, T.F., 1995. The GISP2 d18O climate record of the
Schubert, C., Clapperton, C.M., 1990. Quaternary glaciations in the northern Andes Past 16,500 years and the role of the sun, ocean, and volcanoes. Quat. Res. 44
(Venezuela, Colombia and Ecuador). Quaternary Science Reviews 9 (2e3), (3), 341e354.
123e135. Sugden, D.E., John, B.S., 1976. Glaciers and Landscape: a Geomorphological
Schubert, C., Rinaldi, M., 1987. Nuevos datos sobre la cronología del estadio tardio Approach. Edward Arnold, London, p. 376.
de la Glaciacion Me rida, Andes Venezolanos. Acta Cient. Venez. 38, 135e136. Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H., Bijwaard, H., Olaya, J.,
Schubert, C., Valastro, S., 1980. Quaternary Esnujaque Formation, Venezuelan Rivera, C., 2000. Geodynamic of the northern Andes: subductions and intra-
Andes: preliminary alluvial chronology in a tropical mountain range. Z. Dtsch. continental deformation (Colombia). Tectonics 19 (5), 787e813.
Geol. Ges. 131, 927e947. Wesnousky, S.G., Aranguren, R., Rengifo, M., Owen, L.A., Caffee, M.W., Krishna
Schubert, C., Vaz, J.E., 1987. Edad termoluminiscente del complejo aluvial cua- Murari, M., Pe rez, O.J., 2012. Toward quantifying geomorphic rates of crustal
ternario de Timotes Andes Venezolanos. Acta Científica Venez. 38, 285e286. displacement, landscape development, and the age of glaciation in the Ven-
Schubert, C., Vivas, L., 1993. El Cuaternario de la Cordillera de Me rida, Andes Ven- ezuelan Andes. Geomorphology 141e142, 99e113.
ezolanos. Universidad de Los Andes/Fundacio  n POLAR, Me rida, Venezuela, 345 Weingarten, B., Salgado-Labouriau, M.L., Yuretich, R., Bradley, R., 1991. Late qua-
pp. ternary environmental history of the Venezuelan Andes. In: Yuretich, R. (Ed.),
Siame, L., Braucher, R.Y., Bourles, D., 2000. Les nucle
ides cosmoge niques in situ: de Late Quaternary Climatic Fluctuations of the Venezuelan Andes. University of
nouveaux outils en ge omorphologie quantitative. Bull. Soc. Geo l. Fr. 171 (4), Massachusetts, Amherst, MA, pp. 63e94.

You might also like