Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                

Bablon Et Al 2020 - Glass Shard K-Ar Dating of The Chalupas Caldera Major Eruption

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Quaternary Geochronology 57 (2020) 101053

Contents lists available at ScienceDirect

Quaternary Geochronology
journal homepage: http://www.elsevier.com/locate/quageo

Research paper

Glass shard K-Ar dating of the Chalupas caldera major eruption: Main
Pleistocene stratigraphic marker of the Ecuadorian volcanic arc
Mathilde Bablon a, b, Xavier Quidelleur a, *, Giuseppe Siani a, Pablo Samaniego c, d,
Jean-Luc Le Pennec c, d, Julius Nouet a, C�eline Liorzou e, Santiago Santamaría a, Silvana Hidalgo d
a
Universit�e Paris-Saclay, CNRS, GEOPS, Orsay, 91405, France
b
Universit�e C^
ote d’Azur, CNRS, IRD, Observatoire de la C^ ote d’Azur, G�eoazur, Sophia Antipolis, 06560, Valbonne, France
c
Laboratoire Magmas et Volcans, Universit�e Clermont Auvergne, CNRS, IRD, OPGC, F-63000, Clermont-Ferrand, France
d
Instituto Geofísico, Escuela Polit�ecnica Nacional, Ladr�
on de Guevara E11-253, Ap. 2759, Quito, Ecuador
e
Universit�e de Bretagne Occidentale, Laboratoire G�eosciences Oc�ean, IUEM, 29280, Plouzan�e, France

A R T I C L E I N F O A B S T R A C T

Keywords: New K-Ar ages obtained on juvenile pumice glass shards indicate that the Chalupas ignimbrite, one of the main
K-Ar dating Pleistocene tephra markers of the Ecuadorian arc, was emplaced at 216 � 5 ka. Morphology and major and trace
Glass shards element contents of the glass shards are similar to those of ash layers from deep-sea cores and allow correlation
Ecuador
between continental deposits and marine tephra layers. Based on biostratigraphy and δ18O data, the age models
Chalupas
Ignimbrite eruption
of these cores support our K-Ar age. Fine ashes from the Chalupas eruption column have been found about 1000
km away from the source caldera, and the estimated deposit volume of both ignimbrite and co-ignimbrite de­
posits ranges from 200 to 265 km3. This suggests that the Chalupas event could have been strong enough to reach
the stratosphere and inject large amount of SO2 in both hemispheres, possibly impacting global temperatures. In
addition, the age of the Chalupas ignimbrite obtained here could provide a new radiometric constraint for the age
of isotope stage 7 recorded in orbitally-tuned δ18O deep-sea cores. This study highlights the relevance of K-Ar
dating applied to small glass shards from massive ignimbrite deposits, and the potential that it represents to
improve risk assessments in volcanic zones where dating crystals is not possible. Finally, detailed teph­
rochronology of deep-sea cores and correlation between marine ash-layers and continental volcanic deposits
constitute a strong tool to investigate the eruptive history of an active volcanic arc whose proximal products have
been eroded or deeply buried by younger deposit sequences.

1. Introduction magmatic processes, recurrence and impact of past major eruptions, and
hence better define the volcanic hazards and its frequency-magnitude
Major explosive caldera-forming eruptions are colossal volcanic relationships. As the time period between tephra emission and deposit
phenomenon that can eject large volumes of tephra and volatiles into the is mostly on the order of hours or days (Robock, 2000; Mills, 2000; Rose
atmosphere and spread tephra over wide areas. Such events can impact and Durant, 2009), dating tephra fallouts is a robust way to determine
the local population, infrastructures and natural environment, namely the age of the corresponding eruption. The 40Ar/39Ar single grain dating
the climate system at regional and even global scales (e.g. McCormick method applied on K-rich minerals, such as sanidine, biotite and even
et al., 1995; Robock, 2000; Cole-Dai, 2010). Tephrochronology is a plagioclase crystals, is commonly used to date tephra older than ~50 ka,
powerful tool for dating and correlating deposit sequences scattered i.e. for which 14C age determination is not possible. For instance, this
over large areas. The method consists in comparing tephra features, such technique has been used to accurately date the Young Toba Tuff and
as their age or position in the stratigraphic column, their chemical and investigate the relationship between this major eruptive event and
mineralogical composition, as well as glass shard morphology, in order global cooling (e.g., Williams et al., 2009; Storey et al., 2012; Mark et al.,
to match the deposits associated with a single volcanic event (e.g., Lowe, 2014), or to date the Yellowstone Tuffs and study the eruptive history of
2011). Consequently, tephrochronology allows us to investigate the caldera (e.g., Ellis et al., 2012; Matthews et al., 2015). In offshore

* Corresponding author.
E-mail address: xavier.quidelleur@u-psud.fr (X. Quidelleur).

https://doi.org/10.1016/j.quageo.2020.101053

Available online 6 February 2020


1871-1014/© 2020 Elsevier B.V. All rights reserved.
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

40
Central America this method was used to study characteristics, output Ar/39Ar technique is not always suitable for glass shard dating because
rates and recurrence of explosive eruptions (Schindlbeck et al., 2016a; of 39Ar recoil occurring during irradiation for non-crystalline phases and
b). Also, applications to North Atlantic marine cores have helped to small grains with high surface area to volume ratio (e.g., McDougall and
refine the age of ice-rafted debris recorded in offshore Iceland (Lacasse Harrison, 1999; Morgan et al., 2009), the potassium-argon (K-Ar)
and van den Bogaard, 2002). technique may be successfully employed (e.g., Milne, 1973; Pinti et al.,
As crystals may not be systematically present in pumice samples, 2001). In this paper, we focus on the Chalupas ignimbrite, one of the
glass may be the only juvenile material that can be dated. While the main Pleistocene stratigraphic markers in Ecuador, and we present new

Fig. 1. Location of the samples presented in this study. A) Map of Ecuador and the Gal� apagos Islands. Grey circles represent Quaternary volcanoes of Ecuador and
Colombia. Stars represent the location of ODP and Ninkovich and Shackleton (1975) drilling sites presented in the text. B) Zoom on the central part of the volcanic
arc, showing volcanoes and calderas mentioned in the text. The Chalupas caldera is represented in red, and the location of the Chalupas ignimbrite sample (16EQ01)
with a big white and red star. Small white and red stars indicate the location of Chalupas ignimbrite samples whose major and trace element content are provided in
Appendix A. Dashed lines are basal contours of volcanoes, while continuous lines are used for calderas.

2
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

K-Ar ages carried out on pumice glass shards and biotite crystals. We (Hammersley, 2003). Although this age is in agreement with strati­
also compare the morphology and geochemistry of glass shards of graphic relationships that have been described between the Chalupas
pumices with those of distal tephra fallouts from Ocean Drilling Project’s ignimbrite and deposits from Cotopaxi (Hall and Mothes, 2008a), Huisla
(ODP) Pacific deep-sea cores in order to provide support for our K-Ar age (Ordo�n
~ ez, 2012; Espín, 2014), Sagoatoa (Bablon et al., 2019a), Quilotoa
determination and investigate the volume and magnitude of this major (Hall and Mothes, 2008b), Carihuairazo (Ordo �n
~ ez, 2012), and Chim­
eruption. borazo (Bernard et al., 2008; Samaniego et al., 2012) volcanoes
(Fig. 1B), it is not supported by any description of data, the standards
2. Quaternary stratigraphic markers of the Ecuadorian arc used, age spectra, nor the inverse isochron.

In Ecuador, at least three large caldera systems originate from past 4. Materials and methods
Quaternary volcanic activities. In the northern part of the current
Ecuadorian arc, the ~10 km-wide Chalpat� an caldera is located within 4.1. Sampling and preparation of samples for K-Ar dating
the Interandean Valley, near the border between Ecuador and Colombia
(Fig. 1B). This poorly documented caldera could have collapsed after a The Chalupas ignimbrite deposit was sampled towards the base of a
powerful eruption that occurred during the Quaternary (Barberi et al., ~30 m-high quarry outcrop, in the vicinity of Ambato city, 55 km
1988). Further south, Chalupas and Chacana (Fig. 1B) are the widest southwest of the caldera (sample 16EQ01; Fig. 1B).
caldera systems in Ecuador. The Chacana volcanic system is located in The pumice blocks were crushed manually in an agate mortar, sieved
the Eastern Cordillera, about 30 km from Quito, the capital city of and ultrasonically rinsed with deionized water. A magnetic separator
Ecuador. It experienced two large plinian eruptions at ~160–190 ka was used to remove the andesite xenoliths, and grains were separated
(Bigazzi et al., 1992; Hall and Mothes, 1997, 2008c), whose deposits, the using heavy liquids within a narrow density range (see below). Such
Pifo fall layers, constitute one of the main Pleistocene stratigraphic procedure allows the preservation of the massive structure of the glass
markers of the northern part of the Ecuadorian arc (e.g., Robin et al., shards and minerals.
2009; Alvarado et al., 2014; Bablon et al., 2019b). About 50 km south of The morphology of the glass shards, studied using a Scanning Elec­
Chacana, Chalupas caldera results from a major Pleistocene ignimbrite tron Microscope (SEM), shows a clear difference between the dense glass
eruption. In this study, we focus on the latter as only an unpublished shards, massive and rather smooth (Fig. 2A and B), and which were used
40
Ar/39Ar date is available for the Chalupas ignimbrite (Hammersley, for dating, and the light glass shards, significantly stringier and more
2003; Beate et al., 2006; Mothes and Hall, 2008). vesicular (Fig. 2C and D). Note that glass shards (80–125 μm; ρ ¼
Because of its large-scale dispersion, as well as its rather unique 2.32–2.43) selected from the Chalupas ignimbrite represent less than 5
rhyolite composition and easily recognizable facies, the Chalupas vol% of the initial sample. In addition, biotite crystals (>125 μm; ρ ¼
ignimbrite deposits constitute an important stratigraphic marker of the 2.85–3.28) were also separated for dating.
southern and central parts of the Ecuadorian volcanic arc in the Late Age measurements were performed using the unspiked K-Ar
Pleistocene record. Thereby, it is important to precisely date this event, Cassignol-Gillot technique (Gillot et al., 2006) in the GEOPS laboratory
in order to provide a geochronological constraint to the construction of at Orsay (Paris-Sud University, France). This technique is suitable for
Ecuadorian volcanoes, as well as to the Late Pleistocene deformation of dating groundmass of young lava flows, as well as K-feldspars from
the volcanic and volcano-clastic deposits located within the Interandean explosive deposits, and has been successfully applied for studying
Valley (e.g., Alvarado et al., 2014). In addition, dating events associated Quaternary subduction volcanic products from many places, such as
to such major eruptions allow us to investigate their recurrence through Ecuador (Bablon et al., 2018, 2019a), Argentina or Lesser Antilles (e.g.,
time and improve hazard assessment in this populated area of the Andes, Samper et al., 2009; Germa et al., 2010, 2011; Ricci et al., 2015).
and may allow correlation with climatic records to infer any possible Comparison with the age of the Matuyama-Brunhes paleomagnetic field
global effect of these eruptions. reversal has shown that it can also be used to date glass shards from
volcanic tephra (Pinti et al., 2001). However, this technique requires
3. The Chalupas ignimbrite fresh glass shards, as both the amorphous nature of glass and the large
surface to volume ratio of the grains favor a high mobility of K and/or Ar
The Chalupas volcanic complex (Lat. 00� 470 S; Long. 78� 180 W) is (Cerling et al., 1985). It seems that the K-Ar method is most successful in
located in the Eastern Cordillera of Ecuador, about 50 km from Quito dating volcanic glass, since a significant recoil effect, induced by the
(Fig. 1B), to the west-southwest of Cotopaxi volcano. This volcanic sample irradiation, biases 40Ar/39Ar ages of glass shards toward erro­
system is characterized by a conspicuous 16 k-wide and 18 km-long neous older values (McDougall and Harrison, 1999; Morgan et al.,
caldera (Mothes and Hall, 2008; Bernard and Andrade, 2011) that was 2009). The full description of analytical procedures, standards and un­
formed during a large ignimbrite eruption of rhyolitic composition. certainty calculation is given in Bablon et al. (2018). Ages have been
Andesite to dacite lava flows crop out southwest of the caldera and are calculated using decay constants of Steiger and Ja €ger (1977). The 40Ar
the main remnants of the pre-caldera volcanic system (Barberi et al., signal calibration was performed using an age of 24.21 Ma for HD-B1
1988; Beate et al., 2006). Afterwards, Quilindan ~ a andesitic stratovol­ biotite standard (Hess and Lippolt, 1994). Age results are given in
cano grew in the center of the caldera (Co �rdova, 2018; Co�rdova et al., Table 1 at the 1-σ confidence level. Each argon measurement has been
2018, Fig. 1B). Channelized by relief and volcanoes already present carried out on ~0.9 and ~0.5 g of glass shards and biotite crystals,
during the eruption, the Chalupas ignimbrite was deposited in the respectively.
Interandine Valley south of Quito. Within a radius of about 50 km, the
proximal deposits of the Chalupas ignimbrite are a few tens of meters 4.2. Tephra layers in deep-sea cores
thick (e.g., Beate et al., 2006; Hall and Mothes, 2008a, 2008b; Ordo �n
~ ez,
2012; this study). They consist of a non-welded juvenile pyroclastic For this study, we selected two deep-sea cores from Leg 202 ODP Site
material made up of poorly sorted pumiceous blocks in an ash-rich 1239, located near the Ecuadorian margin, and ODP Site 1240, located
matrix. Pumices contain rare biotite and plagioclase crystals, as well closer to the Gala�pagos Islands (Fig. 1A; Shipboard Scientific Party,
as accessory apatite and Fe-Ti oxides. Onshore tephra fallout deposits 2003). These cores present many marine ash layers visible to the naked
associated to this event spread out to the west and crop out as far as the eye throughout the sedimentary sequence, 24 for the proximal ODP Site
coast, ~300 km from the caldera (Jackson et al., 2019). 1239 and 8 for the distal ODP Site 1240, covering the last ~15 and ~2.3
Beate et al. (2006) reported an unpublished 40Ar/39Ar age of 211 � Ma, respectively (Shipboard Scientific Party, 2003). To investigate the
14 ka for the Chalupas ignimbrite, obtained on plagioclase crystals correlation between these marine deposits and the proximal deposits of

3
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

Fig. 2. Scanning Electron Microscopy (SEM) images of glass shards. A-B) Dense, blocky glass shards of the dated fraction of continental Chalupas ignimbrite. C-D)
Light glass grain fragments of Chalupas ignimbrite, with a lower density due to a significant increase of vesicle content. E-F) Glass shards from the proximal ODP Site
1239, deposited at ~225 ka. G-H) Glass shards from the distal ODP Site 1240, deposited at ~225 ka.

the Chalupas eruption, we focused on ash layers younger than 300 ka. 1240, the ash layer is located between 25.69 and 25.71 mcd, and be­
One layer is significantly thicker than the others, and reaches ~16 tween 25.69 and 25.72 mcd, for holes B and C, respectively (Shipboard
cm-thick in Site 1239, and ~3 cm-thick in Site 1240. Within cores from Scientific Party, 2003). It is composed of millimetric glass shards with
Site 1239, the ash layer is located between 6.65 and 6.83 mcd (meters some calcareous nannofossils, foraminifera and diatoms (mainly
composite depth), between 6.87 and 7.07 mcd, and between 6.72 and Gephyrocapsa spp., Globorotalia menardii and Neogloboquadrina dutertrei;
6.88 mcd, for holes A, B and C, respectively. Within cores from Site Shipboard Scientific Party, 2003), the proportion of which significantly

4
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

Fig. 3. A) K2O vs SiO2 diagram with composition fields of all whole-rock data obtained for the Ecuadorian arc (purple, pink, yellow and green for the Western
Cordillera, Interandean Valley, Eastern Cordillera and back-arc, respectively) and the Gal� apagos Islands (brown stars; Georoc database). HK, MK: high-K and
medium-K calc-alkaline series. B) Detail on composition of rhyolite pumices studied in this work. Squares: Whole-rock measurements (Table 2; Appendix A; Barberi
et al., 1988; Bryant et al., 2006; Ancellin et al., 2017; C�
ordova, 2018). Circles: measurements of single glass shards (Table 3; Jackson et al., 2019). Triangle: ~220 ka
“Layer L00 from deep-sea cores (Ninkovich and Shackleton, 1975; see below). Blue field represents compositions that we associate to a single event of Chalupas
ignimbrite. C) MgO/Fe2O3 ratio as a function of CaO/K2O ratio. Legend as in Fig. 3B. Error bars are included in the size of the symbols. D) Incompatible elements
normalized to primitive mantle spider diagram (Sun and McDonough, 1989) of continental sample 16EQ01 (Chalupas ignimbrite; Table 2) and ODP marine tephra
(Table 4). Purple and hatched grey domains represent the range of all values obtained for the Chalupas ignimbrite (Appendix A) and the Quaternary Ecuadorian arc
(Georoc database).

Table 1
K-Ar data obtained in this study. Column headings indicate sample name, coordinates (Universal Transverse Mercator coordinate system, WGS84 spheroid, Zone 17),
dated material, potassium concentration in percent, radiogenic argon content in percent and 1011 atoms per gram, ages and weighted mean age in ka (1-σ uncertainty).
40 40
Sample Longitude (m) Latitude (m) Phase K (%) Ar* (%) Ar* x 1011 (at/g) Age � 1σ (ka) Mean age (ka)

Chalupas ignimbrite
16EQ01 769765 9867875 Glass shards 3.54 � 0.04 5.16 � 0.10 7.856 � 0.155 213 �5 216 ± 5
5.18 � 0.10 8.009 � 0.158 217 �5
5.09 � 0.10 8.063 � 0.152 218 �5
Biotite crystals 6.66 � 0.07 12.26 � 0.10 19.740 � 0.161 284 �5 286 ± 5
7.14 � 0.10 20.091 � 0.282 289 �6

decreases in the central part of the layer. For each ash layer from both (Site 1240) cores, respectively. Then, glass shards were individually
Sites 1239 and 1240, we collected three to six 2 cm-thick samples every selected by hand under a magnifier and mounted on epoxy resin beads
5 cm depth of the core section. Sample preparation for electron micro­ before polishing.
probe analyses consisted in removing clay by washing and manually
sieving each sample at 63 or 40 μm for proximal (Site 1239) and distal

5
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

4.3. Geochemical analyses 5.2. Geochemical analyses

Whole-rock major and trace element content of the continental Results of whole-rock major and trace element contents of conti­
Chalupas ignimbrite sample and trace element content of marine tephra nental Chalupas ignimbrite are provided in Table 2. Analyses show that
samples were measured by ICP-AES (Inductively Coupled Plasma - the chemical composition of the Chalupas ignimbrite is highly homo­
Atomic Emission Spectrometry) at the Laboratoire G�eosciences Oc�ean of geneous at the deposit scale (Table 2; Appendix A).
the Universit�e de Bretagne Occidentale (Brest, France), following the Major element content of the dated continental sample of the Cha­
procedure detailed in Cotten et al. (1995). Relative uncertainties are lupas ignimbrite and marine tephra layers, determined using electron
�2% for major elements and �5% for trace elements. microprobe on single glass shards, are given in Table 3, and trace
Major element measurements of single glass shards were performed element content measured by ICP-AES are provided in Table 4.
on CAMECA SX-100 and SX-FIVE electron microprobes at the University Overall, marine tephra from proximal and distal ODP cores (light and
Pierre and Marie Curie (CAMPARIS, France). Relative uncertainties are dark blue respectively; Fig. 3A and B), and the continental Chalupas
<0.6% for SiO2; ~1% for Al2O3, <3% for CaO and MgO, <4% for Na2O, ignimbrite (grey circles and squares; Fig. 3A and B), display a rather
<5% for Fe2O3, <6% for K2O, 10% for TiO2 and ~30% for MnO. homogeneous rhyolitic composition, with SiO2 and K2O contents higher
than 74.3% and 3.3%, respectively. They belong to the high-K calc-
5. Results alkaline magmatic suite (orange field; Fig. 4A), which characterizes
products from the Eastern Cordillera in Ecuador. Whole-rock measure­
5.1. K-Ar age of Chalupas ignimbrite ments of minor and trace element contents show that the Chalupas
ignimbrite and the proximal and distal marine tephra layers are homo­
K-Ar measurements have been performed on dense glass shards geneous and enriched in most incompatible elements (e.g., Rb, Ba, Th,
(Fig. 2A and B) and biotite crystals from continental Chalupas ignim­ K) and depleted in Nb, Sr and Ti (Fig. 3D), typical of subduction-related
brite. Dated dense glass shards have a fresher aspect and a more massive volcanic products.
structure than lighter and vesiculated shards (Fig. 2C and D) that have
been removed using heavy liquids. In fact, as vesicles contain gases, they 6. Discussion
are prone to a high contamination in atmospheric argon, and the content
of radiogenic argon is thus more difficult to determine precisely. 6.1. Use of K-Ar dating method for determining the age of quaternary
The potassium content of glass shard samples is 3.54 � 0.04%, and pumice deposits
reaches 6.66 � 0.07% for biotite crystals (Table 1). The content of
radiogenic argon, which derives from the radioactive decay of 40K, Glass shards from the Chalupas ignimbrite have a significantly
ranges between 5.1 and 12.3 � 0.10%. From these values combined with younger age than biotite crystals (216 � 5 and 286 � 5 ka, respectively;
the 40K decay constants (Steiger and Ja
€ger, 1977), we obtain ages of 216 Table 1). Glass shards are fragments of the juvenile fraction that solid­
� 5 and 286 � 5 ka for the glass shards and biotite crystals from the ified during the eruption, whereas the biotite phenocrysts could crys­
Chalupas ignimbrite, respectively (Table 1). tallize earlier in the magmatic reservoir. The older age of biotite crystals
could therefore be explained by the presence of inherited radiogenic
argon and/or excess argon trapped in the differentiated magma, where
partial pressure of argon prior eruption is high and not equilibrated with
atmosphere (e.g., Morgan et al., 2009; Hora et al., 2010). Alternatively,
the potassium leaching favored by the lamellar structure of biotite
Table 2 crystals may also result in older ages. However, we note that the K
Whole-rock major and trace element composition of our Chalupas ignimbrite
content of Chalupas biotite phenocrysts (6.7%; Table 1) is similar to the
sample (16EQ01), with mean and standard deviation of analyses performed on
mean values of 6.9% obtained for biotite crystals contained in rhyolite
20 different outcrops of the Chalupas ignimbrite across Ecuador (Appendix A;
Barberi et al., 1988; Bryant et al., 2006; Ancellin et al., 2017; C�
ordova, 2018),
products from worldwide arc volcanoes (1562 measurements; Georoc
measured by ICP-AES. All analyses of major elements were brought down to a database) and 6.8% obtained for dacite and rhyolite Ecuadorian pum­
total of 100% on a water-free basis. ices (61 measurements from Monzier et al., 1999; Samaniego et al.,
2005; Garrison et al., 2011; Chiaradia et al., 2014; Bernard et al., 2014;
Whole-rock Chalupas ignimbrite
B�eguelin et al., 2015), which suggests that our biotite crystals are rather
Major elements (wt.%) Trace elements (ppm) fresh. As both phenomena (i.e., inherited and/or excess argon and
16EQ01 Other outcrops 16EQ01 Other outcrops leaching) may induce older ages for biotite crystals, we propose that the
SiO2 74.72 74.64 � 0.49 Sc 2.06 2.06 � 0.23 age of the event associated with the eruption of the Chalupas ignimbrite
TiO2 0.24 0.25 � 0.02 V 13.46 13.32 � 3.76 is 216 � 5 ka, as obtained from glass shards. The very small amount of
Al2O3 13.77 13.89 � 0.24 Cr ≪ 3.65 � 2.02 plagioclase phenocrysts present in the sample prevented us from using
Fe2O3 1.46 1.42 � 0.18 Co 0.32 3.00 � 2.18
this mineral phase for dating. However, our age obtained on glass shards
CaO 1.22 1.25 � 0.08 Ni 2.07 � 1.12
refines and agrees with the unpublished 40Ar/39Ar age of 211 � 14 ka

MgO 0.34 0.33 � 0.03 Rb 191.42 172.76 � 5.58
MnO 0.05 0.05 � 0.00 Sr 201.14 207.94 � 6.50 obtained from plagioclase crystals (Hammersley, 2003 in Beate et al.,
K2O 4.29 4.38 � 0.23 Y 12.41 13.22 � 1.92 2006).
Na2O 3.87 3.74 � 0.29 Zr 153.25 162.35 � 13.49
P2O5 0.04 0.04 � 0.01 Nb 10.12 11.29 � 0.36
L.O.I. 3.47 Ba 1099.24 1059.00 � 49.29
6.2. Age of marine tephra and correlation with the Chalupas eruption
La 30.9 31.05 � 1.11
Ce 55.08 56.49 � 2.52 The occurrence of widespread volcanic ash layers in deep-sea sedi­
Nd 19.88 19.50 � 4.32 ments is related to the transport in the atmosphere to the ocean of glass
Sm 3.37 3.58 � 0.19
shards, which rapidly descend the water column and reach the seafloor
Eu 0.42 0.60 � 0.04
Gd 2.82 2.67 � 0.29 in a few days (e.g., Wiesner et al., 1995; Carey, 1997; Manville and
Dy 2.07 2.07 � 0.14 Wilson, 2004; Wetzel, 2009), i.e. instantaneously with respect to
Er 1.09 1.26 � 0.07 geological time-scales.
Yb 1.24 1.25 � 0.08 The stratigraphic age of the analyzed rhyolitic tephra occurring
Th 21.7 21.95 � 1.33
within both ODP 1239 and 1240 Sites can be inferred using the oxygen

6
M. Bablon et al.
Table 3
Major element contents of single glass shards measured by electron microprobe. All analyses were recalculated to a total of 100%. We report only analysis with oxide sum above 93%. Std: standard deviation.
Leg ODP 202 - Site 1239 - Hole C - Section 1H5 - 43-45 cm

wt.% Mean Std


SiO2 78.96 77.27 77.47 76.27 76.61 75.42 76.50 76.28 76.00 76.66 77.08 76.59 75.44 76.11 77.92 76.31 76.68 0.90
TiO2 0.27 0.29 0.09 0.19 0.33 0.38 0.21 0.10 0.12 0.20 0.07 0.12 0.22 0.19 0.18 0.09 0.19 0.09
Al2O3 13.49 14.31 14.44 14.08 14.28 13.76 13.26 13.80 14.00 14.22 14.17 13.81 14.20 13.73 13.67 13.88 13.94 0.32
Fe2O3 0.81 1.13 1.10 1.06 0.91 0.98 1.10 0.60 0.68 0.63 0.53 0.90 1.08 1.22 1.27 0.71 0.92 0.24
CaO 1.22 1.00 1.14 1.15 1.00 0.91 1.09 0.81 0.82 0.84 0.69 0.86 1.22 1.19 1.15 0.71 0.99 0.19
MgO 0.20 0.22 0.19 0.16 0.26 0.23 0.23 0.10 0.14 0.13 0.14 0.06 0.29 0.25 0.27 0.11 0.19 0.07
MnO 0.01 0.01 0.10 0.03 0.03 0.08 0.02 0.03 0.04 0.08 0.21 0.13 0.07 0.01 0.03 0.03 0.05 0.06
K2O 3.30 3.87 3.70 4.58 3.86 5.09 4.54 4.98 4.88 4.21 3.97 4.28 4.05 3.86 3.69 5.27 4.26 0.57
Na2O 1.77 1.90 1.78 2.48 2.72 3.16 3.06 3.29 3.32 3.03 3.14 3.25 3.42 3.46 1.81 2.95 2.78 0.63
Leg ODP 202 - Site 1240 - Hole C - Section 3H2 - 128-130 cm

wt.% Mean Std


SiO2 76.88 76.02 77.75 75.96 75.85 76.71 75.87 75.80 75.00 77.36 76.17 75.53 76.38 76.25 0.76
TiO2 0.14 0.23 0.17 0.11 0.25 0.17 0.20 0.22 0.21 0.13 0.11 0.20 0.23 0.18 0.05
Al2O3 13.94 13.53 12.77 13.30 14.03 13.83 14.24 14.15 14.48 13.85 14.08 13.67 13.87 13.83 0.44
7

Fe2O3 0.94 1.31 0.96 1.16 1.25 1.00 1.08 1.06 1.25 0.91 1.01 1.07 1.04 1.08 0.13
CaO 0.95 1.05 1.02 0.95 1.15 0.82 1.02 1.13 0.81 1.14 1.00 1.24 0.98 1.02 0.13
MgO 0.19 0.24 0.18 0.19 0.24 0.18 0.23 0.18 0.26 0.26 0.20 0.24 0.17 0.21 0.03
MnO 0.13 0.07 0.02 0.22 0.05 0.03 0.07 0.03 0.06 0.03 0.08 0.03 0.11 0.05 0.08
K2O 3.71 4.63 4.53 4.73 4.12 4.79 4.60 4.20 4.77 3.54 4.20 4.52 4.76 4.39 0.41
Na2O 3.12 2.91 2.59 3.38 3.17 2.49 2.70 3.23 3.16 2.83 3.32 3.51 2.45 2.99 0.35
Chalupas ignimbrite (16EQ01)

wt.% Mean Std


SiO2 75.45 75.50 75.90 74.37 75.44 75.21 74.51 75.94 75.29 75.10 76.10 74.37 75.27 0.59
TiO2 0.26 0.16 0.21 0.37 0.24 0.23 0.24 0.19 0.35 0.25 0.21 0.25 0.25 0.06
Al2O3 14.32 13.72 14.11 13.83 14.65 14.31 14.89 13.93 14.53 13.76 14.36 14.12 14.21 0.37
Fe2O3 1.18 0.54 0.75 1.06 0.39 0.99 0.63 0.97 0.94 0.93 0.54 1.41 0.86 0.30
CaO 1.03 1.03 1.04 1.14 0.96 1.03 0.98 1.10 1.06 1.04 1.05 1.15 1.05 0.06
MgO 0.17 0.16 0.22 0.24 0.08 0.22 0.13 0.17 0.22 0.22 0.18 0.22 0.19 0.05
MnO 0.08 0.02 0.05 0.08 0.07 0.12 0.02 0.02 0.10 0.05 0.05 0.05 0.06 0.03

Quaternary Geochronology 57 (2020) 101053


K2O 4.39 4.58 4.19 4.36 4.58 4.51 4.49 4.24 4.43 4.36 4.37 4.51 4.42 0.12
Na2O 3.11 4.30 3.53 4.55 3.59 3.38 4.10 3.44 3.09 4.28 3.13 3.92 3.70 0.51
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

Table 4 Rippert et al., 2017).


Trace element composition of the bottom and the top parts of marine tephra The rhyolitic ash layer lies between 6.72 and 6.88 mcd in ODP Site
layers from ODP 1239 and 1240 Sites, measured by ICP-AES. 1239C. This stratigraphic position corresponds to a cooling event that
Whole-rock ODP tephra layers occurred during the interglacial marine isotope stage 7d (sensu Rails­
Site 1239 - Hole B - Section 1H5 Site 1240 - Hole B - Section 3H5
back et al., 2015, Fig. 4), as pointed out by both alkenone temperature
estimate and increase of benthic δ18O values (Rinco �n-Martínez et al.,
Trace 93–95 cm 98–100 cm 5–7 cm (top 13–15 cm
2010; red curve, Fig. 4). Based on the ODP Site 1239 age model
elements (top of the (bottom of the of the layer) (bottom of the
(ppm) layer) layer) layer) (Rinco�n-Martínez et al., 2010), this tephra is dated at ~ 225 ka. Simi­
larly, in the most distal ODP Site 1240, the rhyolitic ash layer lies be­
Sc 2.40 2.23 2.47 2.65
V 16.09 16.78 31.21 22.21
tween 25.69 and 25.72 mcd. It is in an equivalent stratigraphic position
Cr 4.95 10.36 13.79 10.04 than in ODP Site 1239, as indicated by the δ18O record performed on
Co 1.44 2.30 1.61 1.75 planktonic foraminifera (Rippert et al., 2017, Fig. 4), and the isotopic
Ni 6.01 6.49 17.13 10.42 stage 7d cooling event inferred by sea-surface temperature re­
Rb 146.04 126.99 88.04 152.24
constructions (Pena et al., 2008).
Sr 196.15 192.95 298.73 226.68
Y 11.45 10.86 11.98 11.76 Bowles et al. (1973), and Ninkovich and Shackleton (1975), identi­
Zr 138.58 127.32 120.63 141.74 fied a similar rhyolitic ash layer, the “Layer L00 , off the coast of Colombia,
Nb 9.95 8.95 7.14 9.30 Ecuador and northern Peru. Ninkovich and Shackleton (1975) suggested
Ba 989.84 876.87 927.54 993.81 that the volcanic source is located in southern Colombia or Ecuador,
La 21.04 20.63 24.28 24.02
Ce 46.58 43.54 41.78 45.16
based on the thickness of the deposits and the chemical composition of
Nd 14.71 15.56 16.30 17.01 glass shards. The δ18O records on the benthic foraminifera Uvigerina
Sm 3.10 2.71 4.19 2.89 proboscidea within V19-28 and V19-29 cores (Fig. 1A) show that the
Eu 0.61 0.60 0.46 0.60 volcanic event associated with the “Layer L00 occurred during the first
Gd 2.06 1.97 2.35 2.43
cooling episode of the interglacial marine stage 7 (Ninkovich and
Dy 1.59 1.67 1.62 1.73
Er 1.28 1.22 1.68 1.09 Shackleton, 1975).
Yb 1.08 1.01 1.18 1.16 Marine tephra and the Chalupas ignimbrite present homogeneous
Th 18.08 17.88 15.78 18.31 contents for all major elements, with similar MgO/Fe2O3 and CaO/K2O
ratios (Fig. 3A, B and C). Marine tephra are slightly more enriched in
SiO2 and present a larger K2O range than continental ignimbrite deposits
isotope record (δ18O) from these deep-sea cores previously published by
(Fig. 3B). The whole-rock sample of the Chalupas ignimbrite is
Rinco�n-Martínez et al. (2010), Dyez et al. (2016), and Rippert et al.
composed of a mixture of glass and crystals, whereas single shard
(2017), respectively. The age model of δ18O records was based on the
measurements of marine tephra were performed on pristine glass, which
global δ18O stack of Lisiecki and Raymo (2005), through δ18O mea­
concentrates incompatible elements and is slightly more differentiated.
surements on the benthic foraminifera Cibicidoides wuellestorfii and on
Another hypothesis is that marine tephra might originate from a major
the planktonic foraminifera Globorotaloides hexagonus for ODP Sites
explosion that occurred in early eruptive phase of the event, while
1239 and 1240, respectively (Fig. 4; Rinco �n-Martínez et al., 2010;
continental ignimbrite deposits might have been emplaced latter, when

Fig. 4. δ18O (‰) record of the planktonic


Globorotaloides hexagonus (green curve;
modified from Rippert et al., 2017) and
benthic Cibicidoides wuellerstorfi (red curve,
modified from Rinc� on-Martínez et al., 2010)
foraminifera as a function of depth, for the
ODP Sites 1240 and 1239 (Hole C), respec­
tively. Depth of the ash layer presented in
this study are represented by circles (symbols
as in Fig. 3), and its thickness is indicated in
parentheses. White and yellow fields repre­
sent marine isotope stages (MIS) defined by
Lisiecki and Raymo (2005). Substages of MIS
7 from Railsback et al. (2015).

8
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

the magma reservoir was emptying, emitting slightly less differentiated (areas of ~1.6 � 106 km2, 5.4 � 105 km2 and 1.7 � 105 km2, respec­
and more homogeneous magma. In addition, because the marine tephra tively; Fig. 5A) in order to derive a plot of ash thickness decay rate (i.e.
layer off the Ecuadorian coast is significantly thicker than in the distal thickness against square root of isopach area). Because we have only
cores (Ninkovich and Shackleton, 1975; Shipboard Scientific Party, three contour lines, we calculated the minimal ash volume using the
2003), and based on its trace element signature that is typical of arc single isopach approach of Legros (2000), which yields bulk
magmas (light and dark blue fields; Fig. 3D), it cannot originate from the co-ignimbrite ash volumes in the range of about 60–140 km3, with an
Gala�pagos Islands (brown stars; Fig. 3A). average close to 100 km3. A more realistic approximation that assumes
Considering the good match between the major element composition an exponential thickness decay rate from the source and a fit to our data
of marine tephra, and continental pumice samples combined to our new (using two or three isopachs on a single exponential segment, Pyle,
age determination, we can consider that the “Layer L” (Ninkovich and 1989) yields volumes in the range of about 120–155 km3, values that are
Shackleton, 1975), the ~6.8 m-deep ash layer in ODP Site 1239 (Fig. 4), considered reliable based on our data set.
the ~25.7 m-deep ash layer in ODP Site 1240, and the continental The bulk volume of the continental ignimbrite around the caldera
Chalupas ignimbrite deposit (sample 16EQ01) are related to the same and in the Interandean Valley (beige field, total area of ~3150 km2;
~216 ka Chalupas eruption. Fig. 5B) was estimated on a gross scale through delineating the envelope
It should be emphasized that the K-Ar age that we obtain for the of known outcrops (red triangles; Beate et al., 2006; Bernard et al., 2008;
Chalupas ignimbrite (216 � 5 ka; Table 1) is slightly younger than that Bustillos, 2008; Hall and Mothes, 2008a, 2008b; Ordo �n
~ ez, 2012;
inferred from age models of deep-sea cores (~225 ka), although these Co�rdova, 2018; Jackson et al., 2019; this study; Fig. 5B), extrapolating
ages are in agreement within 2-sigma uncertainty. Age models in ODP the distribution of the ignimbrite in nearby valleys, and assuming from
Sites 1239 and 1340 have been reconstructed using the global orbitally- literature data and our field observations an average thickness of about
tuned δ18O benthic stack of Lisiecki and Raymo (2005), based on 57 25–35 m. Results point to a bulk ignimbrite volume in the range of
worldwide cores, and with an average uncertainty of ~4 ka for the last 80–110 km3 (95 � 15 km3), slightly above previous literature estimates
million years (Lisiecki and Raymo, 2005). It is worth mentioning that of 80–200 km3 (Beate et al., 2006; Hall and Mothes, 2008a).
benthic δ18O records are related to sea-level changes, and that the δ18O The total bulk deposit volume of both ignimbrite and co-ignimbrite
stack of Lisiecki and Raymo (2005) is essentially based on this signal. thus amounts to about 200–265 km3 (~230 � 30 km3). This result
However, Skinner and Shackleton (2005) state that during the main ranks the 216 � 5 ka Chalupas eruption at VEI 7, consistent with the
climatic transitions, local changes in deep-water hydrography could associated ~16 � 18 km caldera size (Lipman, 1999; Geshi et al., 2014).
account for a lag deep-eastern Pacific temperatures of ~4 ka when As a comparison, the volume erupted at Chalupas caldera is about two
compared to the North Atlantic ones. Such further uncertainties should times larger than volumes released by historic eruptions of Tambora,
therefore be taken into account for age model reconstructions in this Samalas, Krakatoa and Santorini (e.g., Lipman, 1999; Self et al., 2004;
area. Hence, our new K-Ar age combined to δ18O records offers a great Lavigne et al., 2013; Kara �tson et al., 2018). Therefore, the Chalupas
potential to consider the Chalupas tephra as a new chronostratigraphic event may be the largest Quaternary eruption from the Northern Andean
marker for the marine isotopic stage 7 in the East Equatorial Pacific Volcanic Zone and it can be envisioned that such an eruption could have
Ocean. caused a short-term climate forcing, with a possible global cooling.
Measurements of sulphide contents contained in fluid inclusions of the
6.3. K-Ar dating of marine glass shards glass shards are needed to quantify the amount of volatiles injected into
the atmosphere, and to better infer the climatic impact of this
In order to compare the age of both Chalupas continental ignimbrite super-eruption.
and marine co-ignimbrite, we tentatively performed K-Ar dating on glass
shards from the ODP 1239C core (section 1H5, intervals 43–45 and 7. Conclusions
38–40 cm), following the same sample preparation as for 16EQ01 (see
Methods). Although the K content of both marine deposits is about 3.5%, The present study demonstrates the reliability of the K-Ar dating
in agreement with measurements performed on continental deposits method performed on glass shards from a massive continental ignim­
(16EQ01; Table 1), the K-Ar ages we obtained for both marine deposits brite deposit. This method is a reliable tool to date Quaternary pyro­
are meaningless, with values older than 700 ka for both of them. There is clastic deposits. The samples require a careful selection of only very
therefore probably a high radiogenic 40Ar content in excess. It should be fresh material, and removal of all altered and/or vesiculated glass
emphasized that deep-sea tephra are distal deposits made of thin par­ shards. We provide a new K-Ar age of 216 � 5 ka for the Chalupas
ticles that cooled rapidly during eruption. On the other hand, pumices ignimbrite.
within the massive (>50 m) proximal ignimbrite deposit probably Based on glass shard morphology, whole-rock and glass geochemical
cooled much slower, thereby allowing equilibrium with atmosphere. data, as well as isopach map of deposits, we have correlated the Cha­
Consequently, any initial excess 40Ar would have been removed from lupas ignimbrite to marine ash-layers from several cores offshore
16EQ01 ignimbrite glass, while it was trapped within the glass of distal Ecuador. It shows that the Chalupas eruption could have injected be­
tephra. Further analyses are required to investigate this important issue tween 200 and 265 km3 of pyroclastic material into the atmosphere, and
regarding the wider applicability of our dating technique. that ashes have reached areas located more than 900 km away from the
source caldera.
6.4. Spatial distribution and volume of Chalupas eruption Our K-Ar age obtained for Chalupas ignimbrite also provides reliable
independent timing control for the δ18O records of isotope stage 7,
There is no clear evidence of Plinian fall deposits associated with the which is tuned by orbitally derived times-scales.
216 � 5 ka Chalupas ignimbrite, but the offshore ash fall layer preserved Finally, we hope that our study will reinforce the interest of jointly
in deep-sea cores (Ninkovich and Shackleton, 1975; Shipboard Scientific studying continental volcanic deposits and marine tephra layers to
Party, 2003) and correlated in this study to that eruption (Figs. 3 and 4) better characterize the major Pleistocene eruptions of the Ecuadorian
seemingly corresponds to a co-ignimbrite ash deposit that shows a arc. Such an approach also appears to be the only way to identify
consistent thinning pattern from a source in sub-central Ecuador at catastrophic eruptions that occurred during early stages of the current
Chalupas caldera. Although offshore thickness could have been biased Ecuadorian arc construction, and could help to better define the recur­
by ocean currents, slumping or bioturbation (Manville and Wilson, rence of such major events.
2004; Wetzel, 2009; Lowe, 2011), the ash layer thickness data allowed
us designing a tentative isopach map with 1, 7 and 15 cm contour lines

9
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

Fig. 5. A) Tentative isopach map (1, 7 and 15 cm) of the 216 � 5 ka Chalupas offshore co-ignimbrite ash fall deposit. Location of the cores shown as orange, yellow
and green circles and thickness data are from Ninkovich and Shackleton (1975)and Shipboard Scientific Party (2003). B) Reconstructed distribution of the Chalupas
ignimbrite (beige field). Location of selected exposure sites shown as red triangles and ignimbrite thickness values are from Beate et al. (2006); Bernard et al. (2008);
Bustillos (2008); Hall and Mothes (2008a; 2008b); Ord� on~ ez (2012); C�
ordova (2018); Jackson et al. (2019); and this study. The Chalupas caldera is shown with a solid
star, and its rim is depicted with a dotted line.

Declaration of competing interest International Ocean Discovery Program (IODP) for providing us with the
core samples, and Guillaume Leduc for details regarding δ18O data ob­
The authors declare that there is no conflict of interest. tained at ODP 1239 Site. This work is part of the Laboratoire Mixte In­
ternational “S�eismes et Volcans dans les Andes du Nord” program, an
Funding Ecuadorian-French cooperation program between the IG-EPN and the
French Institut de Recherche pour le D�eveloppement (IRD). This is
Field work and analyses (preparations of samples, K and Ar mea­ LGMT contribution number 155 and Laboratory of Excellence ClerVolc
surements for K-Ar ages, ICP-AES and electron microprobe analyses for contribution number 399.
major and trace elements) were supported by INSU CNRS TelluS Aleas
([INSU 2015-ALEAS], [INSU 2016-ALEAS]) and LMI IRD ([2012-16 LMI Appendix A. Supplementary data
SVAN-IRD]) programs.
Supplementary data to this article can be found online at https://doi.
org/10.1016/j.quageo.2020.101053.
Acknowledgments

The authors are grateful to Roelant van der Lelij and Patricia Mothes References
for their detailed reviews, constructive comments and suggestions,
Alvarado, A., Audin, L., Nocquet, J.M., Lagreulet, S., Segovia, M., Font, Y., Lamarque, G.,
which contributed to improve this manuscript. The authors also wish to Yepes, H., Mothes, P., Rolandone, F., Jarrín, P., Quidelleur, X., 2014. Active
thank the Instituto Geofísico of Quito (IG-EPN) for their support during tectonics in Quito, Ecuador, assessed by geomorphological studies, GPS data, and
field work. We are grateful to Ama€elle Landais and Rita Traversi for their crustal seismicity. Tectonics 33, 67–83. https://doi.org/10.1002/2012TC003224.
Ancellin, M.-A., Samaniego, P., Vlast�elic, I., Nauret, F., Gannoun, A., Hidalgo, S., 2017.
help and advice concerning ice core data, as well as the Chalupas quarry Across-arc versus along-arc Sr-Nd-Pb isotope variations in the Ecuadorian volcanic
manager who gave us permission to sample. Special thanks to the arc. G-cubed 18, 1163–1188. https://doi.org/10.1002/2016GC006679.

10
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

Bablon, M., Quidelleur, X., Samaniego, P., Le Pennec, J.-L., Lahitte, P., Liorzou, C., Geshi, N., Rush, J., Acocella, V., 2014. Evaluating volumes for magma chambers and
Bustillos, J.E., Hidalgo, S., 2018. Eruptive chronology of Tungurahua volcano magma withdrawn for caldera collapse. Earth Planet Sci. Lett. 396, 107–115.
(Ecuador) revisited based on new K-Ar ages and geomorphological reconstructions. https://doi.org/10.1016/j.epsl.2014.03.059.
J. Volcanol. Geoth. Res. 357, 378–398. https://doi.org/10.1016/j. Gillot, P.-Y., Hildenbrand, A., Lef�evre, J.-C., Albore-Livadie, C., 2006. The K/Ar dating
jvolgeores.2018.05.007. method : principle, analytical techniques, and application to Holocene volcanic
Bablon, M., Quidelleur, X., Samaniego, P., Le Pennec, J.-L., Audin, L., Jomard, H., eruptions in Southern Italy. Acta Vulcanol. 18, 55–66.
Baize, S., Liorzou, C., Hidalgo, S., Alvarado, A., 2019a. Interactions between Hall, M.L., Mothes, P.A., 1997. Origin and age of the upper Cangahua formation,
volcanism and geodynamics in the southern termination of the Ecuadorian arc. Tumbaco valley (Ecuador). Suelos volc� anicos endurecidos, Mem. III Symp. Intern.
Tectonophysics 751. https://doi.org/10.1016/j.tecto.2018.12.010, 54–14. ORSTOM (EPN, Quito, dec. 1996) 19–28.
Bablon, M., Quidelleur, X., Samaniego, P., Le Pennec, J.-L., Santamaria, S., Liorzou, C., Hall, M., Mothes, P., 2008a. The rhyolitic–andesitic eruptive history of Cotopaxi volcano,
Hidalgo, S., Eschbach, B., 2019b. Volcanic history reconstruction in northern Ecuador. Bull. Volcanol. 70, 675–702. https://doi.org/10.1007/s00445-007-0161-2.
Ecuador: insights for eruptive and erosion rates on the whole Ecuadorian arc. Bull. Hall, M., Mothes, P., 2008b. Quilotoa volcano - Ecuador: an overview of young dacitic
Volcanol. (submitted for publication). volcanism in a lake-filled caldera. J. Volcanol. Geoth. Res. 176, 44–55. https://doi.
Barberi, F., Coltelli, M., Ferrara, G., Innocenti, F., Navarro, J.M., Santacroce, R., 1988. org/10.1016/j.jvolgeores.2008.01.025.
Plio-quaternary volcanism in Ecuador. Geochem. Mag. 125, 1–14. Hall, M., Mothes, P., 2008c. The Chacana caldera complex in Ecuador. IOP Conf. Ser.
Beate, B., Hammersley, L., DePaolo, D., Deino, A.I., 2006. La Edad de la Ignimbrita de Earth Environ. Sci. 3 https://doi.org/10.1088/1755-1307/3/1012004.
Chalupas, prov. de Cotopaxi, Ecuador, y su importancia como marcador Hammersley, L., 2003. The Chalupas Caldera. Ph.D. thesis. University of California,
afico. 6th Jornadas en Ciencias de la Tierra (EPN, Quito), pp. 68–71. Abstract
estratigr� Berkeley, p. 201.
book. Hess, J.C., Lippolt, H.J., 1994. Compilation of K-Ar measurements on HD-B1 standard
B�
eguelin, P., Chiaradia, M., Beate, B., Spikings, R., 2015. The Yanaurcu volcano (Western biotite. 1994 status report, Phanerozoic time Scale. Bull. Liais. Inform. IUGS
Cordillera, Ecuador): a field, petrographic, geochemical, isotopic and Subcomm. Geochronol. 122.
geochronological study. Lithos 218–219, 37–53. https://doi.org/10.1016/j. Hora, J.M., Singer, B.S., Jicha, B.R., Beard, B.L., Johnson, C.M., de Silva, S.,
lithos.2015.01.014. Salisbury, M., 2010. Volcanic biotite-sanidine 40Ar/39Ar age discordances reflect Ar
Bernard, B., Andrade, D., 2011. Volcanes Cuaternarios del Ecuador Continental. Map 1, partitioning and pre-eruption closure in biotite. Geology 38, 923–926.
500000. Jackson, L.J., Horton, B.K., Beate, B.O., Bright, J., Breecker, D.O., 2019. Testing stable
Bernard, B., van Wyk de Vries, B., Barba, D., Leyrit, H., Robin, C., Alcaraz, S., isotope paleoaltimetry with Quaternary volcanic glasses from the Ecuadorian Andes.
Samaniego, P., 2008. The Chimborazo sector collapse and debris avalanche: deposit Geology 47, 411–414. https://doi.org/10.1130/G45861.1.
characteristics as evidence of emplacement mechanisms. J. Volcanol. Geoth. Res. Kar�atson, D., Gertisser, R., Telbisz, T., Vereb, V., Quidelleur, X., Druitt, T., Nomikou, P.,
176, 36–43. https://doi.org/10.1016/j.jvolgeores.2008.03.012. K�osik, S., 2018. Towards reconstruction of the lost Late Bronze Age intra-caldera
Bernard, J., Kelfoun, K., Le Pennec, J.-L., Vallejo Vargas, S., 2014. Pyroclastic flow island of Santorini, Greece. Sci. Rep. 8 https://doi.org/10.1038/s41598-018-25301-
erosion and bulking processes: comparing field-based vs. modeling results at 2.
Tungurahua volcano, Ecuador. Bull. Volcanol. 76 https://doi.org/10.1007/s00445- Lacasse, C., van den Bogaard, P., 2002. Enhanced airborne dispersal of silicic tephras
014-0858-y. during the onset of Northern Hemisphere glaciations, from 6 to 0 Ma records of
Bigazzi, G., Coltelli, M., Hadler, N.J.C., Araya, A.M.O., Oddone, M., Salazar, E., 1992. explosive volcanism and climate change in the subpolar North Atlantic. Geology 30,
Obsidian-bearing lava flows and pre-Columbian artifacts from the Ecuadorian Andes: 623–626.
first new multidisciplinary data. J. S. Am. Earth Sci. 6, 21–32. https://doi.org/ Lavigne, F., Degeai, J.-P., Komorowski, J.-C., Guillet, S., Robert, V., Lahitte, P.,
10.1016/0895-9811(92)90014-. Oppenheimer, C., Stoffel, M., Vidal, C.M., Surono, Pratomo I., Wassmer, P.,
Bowles, F.A., Jack, R.N., Carmichael, I.S.E., 1973. Investigation of deep-sea volcanic ash Hajdas, I., Hadmoko, D.S., de Belizal, E., 2013. Source of the great A.D. 1257
layers from equatorial Pacific cores. Geol. Soc. Am. Bull. 84, 2371–2388. mystery eruption unveiled, Samalas volcano, Rinjani Volcanic Complex, Indonesia.
Bryant, J.A., Yogodzinski, G.M., Hall, M.L., Lewicki, J.L., Bailey, D.G., 2006. Proc. Natl. Acad. Sci. Unit. States Am. 110, 16742–16747. https://doi.org/10.1073/
Geochemical constraints on the origin of volcanic rocks from the andean northern pnas.1307520110.
volcanic zone, Ecuador. J. Petrol. 47, 1147–1175. https://doi.org/10.1093/ Legros, F., 2000. Minimum volume of a tephra fallout deposits estimated from a single
petrology/egl006. isopach. J. Volcanol. Geoth. Res. 96 (1–2), 25–32. https://doi.org/10.1016/S0377-
Bustillos, J.E., 2008. Las avalanchas de escombros en el sector del volcan Tungurahua. 0273(99)00135-3.
Facultad de ingenieria en geologia y petroleos. Escuela Politecnica Nacional, Quito. Lipman, P., 1999. Caldera. In: Sigurdsson, Haraldur (Ed.), Encyclopedia of Volcanoes,
Carey, S., 1997. Influence of convective sedimentation on the formation of widespread ISBN 0-12-643140-X.
tephra fall layers in the deep sea. Geology 25 (9), 839–842. https://doi.org/ Lisiecki, L.E., Raymo, M.E., 2005. A Pliocene-Pleistocene stack of 57 globally distributed
10.1130/0091-7613(1997)025<0839:IOCSOT>2.3.CO. benthic δ18O records. Paleoceanography 20. https://doi.org/10.1029/
Cerling, T.E., Brown, F.H., Bowman, J.R., 1985. Low-temperature alteration of volcanic 2004PA001071.
glass: hydration, Na, K, 18O and Ar mobility. Chem. Geol. Isot. Geosci. 52, 281–293. Lowe, D.J., 2011. Tephrochronology and its application: a review. Quat. Geochronol. 6,
Chiaradia, M., Barnes, J.D., Cadet-Voisin, S., 2014. Chlorine stable isotope variations 107–153. https://doi.org/10.1016/j.quageo.2010.08.003.
across the Quaternary volcanic arc of Ecuador. Earth Planet Sci. Lett. 396, 22–33. Manville, V., Wilson, C.J.N., 2004. Vertical density currents: a review of their potential
https://doi.org/10.1016/j.epsl.2014.03.062. role in the deposition and interpretation of deep-sea ash layers. J. Geol. Soc. 161,
Cole-Dai, J., 2010. Volcanoes and climate. Wiley Interdisciplinary Rev.: Clim. Change 1, 47–958.
824–839. https://doi.org/10.1002/wcc.76. Mark, D.F., Petraglia, M., Smith, V.C., Morgan, L.E., Barfod, D.N., Ellis, B.S., Pearce, N.J.,
C�
ordova, M., 2018. Identificaci� on y caracterizaci�
on de los últimos productos eruptivos de Pal, J.N., Korisettar, R., 2014. A high-precision 40Ar/39Ar age for the Young Toba
la fase resurgente de la caldera de Chalupas, vol. 133. Escuela Polit�ecnica Nacional, Tuff and dating of ultra-distal tephra: forcing of Quaternary climate and implications
Quito. for hominin occupation of India. Quat. Geochronol. 21, 90–103. https://doi.org/
C�
ordova, M., Mothes, P., Hall, M., Telenchana, E., 2018. Identification and 10.1016/j.quageo.2012.12.004.
characterization of the youngest eruptive products of the Chalupas Caldera, Ecuador: Matthews, N.E., Vazquez, J.A., Calvert, A.T., 2015. Age of the Lava Creek supereruption
an update on the caldera. Cities on Volcanoes 10. Abstract ID: 438. and magma chamber assembly at Yellowstone based on 40Ar/39Ar and U-Pb dating of
Cotten, J., Le Dez, A., Bau, M., Caroff, M., Maury, R.C., Dulski, P., Fourcade, S., Bohn, M., sanidine and zircon crystals. G-cubed 16, 2508–2528. https://doi.org/10.1002/
Brousse, R., 1995. Origin of anomalous rare-earth element and yttrium enrichments 2015GC005881.
in subaerially exposed basalts: evidence from French Polynesia. Chem. Geol. 119, McCormick, P., Thomason, L.W., Trepte, C.R., 1995. Atmospheric effects of the Mt
115–138. Pinatubo. Nature 373, 399–404.
Dyez, K.A., Ravelo, A.C., Mix, A.C., 2016. Evaluating drivers of Pleistocene eastern McDougall, I., Harrison, T.M., 1999. Geochronology and Thermochronology by the
40
tropical Pacific sea surface temperature. Paleoceanography 31, 1054–1069. https:// Ar/39Ar Method, second ed. Oxford University Press, United States.
doi.org/10.1002/2015PA002873. Mills, M.J., 2000. Volcanic aerosol and global atmospheric effects. In:
Ellis, B.S., Mark, D.F., Pritchard, C.J., Wolff, J.A., 2012. Temporal dissection of the Sigurdsson, Haraldur (Ed.), Encyclopedia of Volcanoes, ISBN 0-12-643140-X.
Huckleberry Ridge Tuff using the 40Ar/39Ar dating technique. Quat. Geochronol. 9, Milne, J.D.G., 1973. Mount Curl Tephra, a 230000-year-old implications for Quaternary
34–41. https://doi.org/10.1016/j.quageo.2012.01.006. chronology marker bed in New Zealand, and its. N. Z. J. Geol. Geophys. 16,
Espín, P.A., 2014. Caracterizaci� on geol�
ogica y litol�
ogica de los dep�ositos lah�
aricos de 519–532. https://doi.org/10.1080/00288306.1973.10431375.
Mera, provincia de Pastaza. Ph.D. tesis. Facultad de Geología y Petr� oleos, EPN, Monzier, M., Robin, C., Samaniego, P., Hall, M.L., Cotten, J., Mothes, P., Arnaud, N.,
Quito, p. 213. 1999. Sangay volcano, Ecuador: structural development, present activity and
Garrison, J.M., Davidson, J.P., Hall, M., Mothes, P., 2011. Geochemistry and petrology of petrology. J. Volcanol. Geoth. Res. 90, 49–79.
the most recent deposits from Cotopaxi volcano, northern volcanic zone, Ecuador. Morgan, L.E., Renne, P.R., Taylor, R.E., WoldeGabriel, G., 2009. Archaeological age
J. Petrol. 52, 1641–1678. https://doi.org/10.1093/petrology/egr023. constraints from extrusion ages of obsidian: examples from the Middle Awash,
Germa, A., Quidelleur, X., Gillot, P.Y., Tchilinguirian, P., 2010. Volcanic evolution of the Ethiopia. Quat. Geochronol. 4, 193–203. https://doi.org/10.1016/j.
back-arc Pleistocene Payun Matru volcanic field (Argentina). J. S. Am. Earth Sci. 29, quageo.2009.01.001.
717–730. https://doi.org/10.1016/j.jsames.2010.01.002. Mothes, P., Hall, M.L., 2008. Rhyolitic calderas and centers clustered within the active
Germa, A., Quidelleur, X., Lahitte, P., Labanieh, S., Chauvel, C., 2011. The K-Ar andesitic belt of Ecuador’s Eastern Cordillera. IOP Conf. Ser. Earth Environ. Sci. 3
Cassignol–Gillot technique applied to western Martinique lavas: a record of Lesser https://doi.org/10.1088/1755-1307/3/1/012007.
Antilles arc activity from 2 Ma to Mount Pel�ee volcanism. Quat. Geochronol. 6, Ninkovich, D., Shackleton, N.J., 1975. Distribution, stratigraphic position and age of ash
341–355. https://doi.org/10.1016/j.quageo.2011.02.001. layer “L”, in the Panama Basin region. Earth Planet Sci. Lett. 27, 20–34.

11
M. Bablon et al. Quaternary Geochronology 57 (2020) 101053

Ord�on~ ez, J., 2012. Dep�


ositos volc�
anicos del Pleistoceno Tardío en la cuenca de. Ambato: Samper, A., Quidelleur, X., Komorowski, J.-C., Lahitte, P., Boudon, G., 2009. Effusive
caracterizaci� on, distribuci�
on y origen. Ph.D. tesis. Facultad de Geología y Petr� oleos history of the Grande D�ecouverte volcanic complex, southern Basse-Terre
(Quito, EPN), p. 213. (Guadeloupe, French west Indies) from new K-Ar Cassignol–Gillot ages. J. Volcanol.
Pena, L.D., Cacho, I., Ferretti, P., Hall, M.A., 2008. El Ni~ no–Southern Oscillation-like Geoth. Res. 187, 117–130. https://doi.org/10.1016/j.jvolgeores.2009.08.016.
variability during glacial terminations and interlatitudinal teleconnections. Schindlbeck, J.C., Kutterolf, S., Freundt, A., Alvarado, G.E., Wang, K.-L., Straub, S.M.,
Paleoceanography 23, PA3101. https://doi.org/10.1029/2008PA001620. Hemming, S.R., Frische, M., Woodhead, J.D., 2016a. Late Cenozoic
Pinti, D.L., Quidelleur, X., Chiesa, S., Ravazzi, C., Gillot, P.-Y., 2001. K-Ar dating of an tephrostratigraphy offshore the southern Central American Volcanic Arc: 1. Tephra
early Middle Pleistocene distal tephra in the interglacial varved succession of ages and provenance. G-cubed 17, 4641–4668. https://doi.org/10.1002/
Pianico-Sellere (Southern Alps, Italy). Earth Planet Sci. Lett. 188, 1–7. 2016GC006503.
Pyle, D.M., 1989. The thickness, volume and grainsize of tephra fall deposits. Bull. Schindlbeck, J.C., Kutterolf, S., Freundt, A., Straub, S.M., Vannucchi, P., Alvarado, G.E.,
Volcanol. 51 (1), 1–15. 2016b. Late Cenozoic tephrostratigraphy offshore the southern Central American
Railsback, L.B., Gibbard, P.L., Head, M.J., Voarintsoa, N.R.G., Toucanne, S., 2015. An Volcanic Arc: 2. Implications for magma production rates and subduction erosion. G-
optimized scheme of lettered marine isotope substages for the last 1.0 million years, cubed 17, 4585–4604. https://doi.org/10.1002/2016GC006504.
and the climatostratigraphic nature of isotope stages and substages. Quat. Sci. Rev. Self, S., Gertisser, R., Thordarson, T., Rampino, M.R., Wolff, J.A., 2004. Magma volume,
111, 94–106. https://doi.org/10.1016/j.quascirev.2015.01.012. volatile emissions, and stratospheric aerosols from the 1815 eruption of Tambora.
Ricci, J., Quidelleur, X., Lahitte, P., 2015. Volcanic evolution of central Basse-Terre Geophys. Res. Lett. 31, 1–4. https://doi.org/10.1029/2004GL020925.
Island revisited on the basis of new geochronology and geomorphology data. Bull. Shipboard Scientific Party, 2003. Sites 1238, 1239 and 1240. In: Mix, A.C.,
Volcanol. 77 https://doi.org/10.1007/s00445-015-0970-7. Tiedemann, R., Blum, P., et al. (Eds.), Proceedings of Ocean Drill Program, Initial
Rinc�on-Martínez, D., Lamy, F., Contreras, S., Leduc, G., Bard, E., Saukel, C., Blanz, T., Report, p. 202.
Mackensen, A., Tiedemann, R., 2010. More humid interglacials in Ecuador during Skinner, L.C., Shackleton, N.J., 2005. An Atlantic lead over Pacific deep-water change
the past 500 kyr linked to latitudinal shifts of the equatorial front and the across Termination I: implications for the application of the marine isotope stage
Intertropical Convergence Zone in the eastern tropical Pacific. Paleoceanography 25, stratigraphy. Quat. Sci. Rev. 24, 571–580. https://doi.org/10.1016/j.
PA2210. https://doi.org/10.1029/2009PA001868. quascirev.2004.11.008.
Rippert, N., Max, L., Mackensen, A., Cacho, I., Povea, P., Tiedemann, R., 2017. Steiger, R.H., J€ager, E., 1977. Subcommission on geochronology: convention on the use
Alternating influence of northern versus southern-sourced water masses on the of decay constants in geo- and cosmochronology. Earth Planet Sci. Lett. 36, 359–362.
equatorial pacific subthermocline during the past 240 ka. Paleoceanography 32, Storey, M., Roberts, R.G., Saidin, M., 2012. Astronomically calibrated 40Ar/39Ar age for
1256–1274. https://doi.org/10.1002/2017PA003133. the Toba supereruption and global synchronization of late Quaternary records. Proc.
Robin, C., Eissen, J.-P., Samaniego, P., Martin, H., Hall, M., Cotten, J., 2009. Evolution of Natl. Acad. Sci. Unit. States Am. 109, 18684–18688. https://doi.org/10.1073/
the late Pleistocene Mojanda–Fuya Fuya volcanic complex (Ecuador), by progressive pnas.1208178109.
adakitic involvement in mantle magma sources. Bull. Volcanol. 71, 233–258. Sun, S.-s., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
https://doi.org/10.1007/s00445-008-0219-9. implications for mantle composition and processes. Geol. Soc. Lond. Spec. Publ. 42,
Robock, A., 2000. Volcanic eruptions and climate. Rev. Geophys. 38, 191–219. 313–345. https://doi.org/10.1144/GSL.SP.1989.042.01.19.
Rose, W.I., Durant, A.J., 2009. Fine ash content of explosive eruptions. J. Volcanol. Wetzel, A., 2009. The preservation potential of ash layers in the deep-sea: the example of
Geoth. Res. 186, 32–39. https://doi.org/10.1016/j.jvolgeores.2009.01.010. the 1991-Pinatubo ash in the South China Sea. Sedimentology 56, 1992–2009.
Samaniego, P., Martin, H., Monzier, M., Robin, C., Fornari, M., Eissen, J.-P., Cotten, J., Wiesner, M.G., Wang, Y., Zheng, L., 1995. Fallout of volcanic ash to the deep South China
2005. Temporal evolution of magmatism in the northern volcanic zone of the Andes: Sea induced by the 1991 eruption of Mount Pinatubo. Geology 23, 885–888.
the geology and petrology of Cayambe volcanic complex (Ecuador). J. Petrol. 46, Williams, M.A.J., Ambrose, S.H., van der Kaars, S., Ruehlemann, C., Chattopadhyaya, U.,
2225–2252. https://doi.org/10.1093/petrology/egi053. Pal, J., Chauhan, P.R., 2009. Environmental impact of the 73ka Toba super-eruption
Samaniego, P., Barba, D., Robin, C., Fornari, M., Bernard, B., 2012. Eruptive history of in south Asia. Palaeogeogr. Palaeoclimatol. Palaeoecol. 284, 295–314. https://doi.
Chimborazo volcano (Ecuador): a large, ice-capped and hazardous compound org/10.1016/j.palaeo.2009.10.009.
volcano in the Northern Andes. J. Volcanol. Geoth. Res. 221 (222), 33–51. https://
doi.org/10.1016/j.jvolgeores.2012.01.014.

12

You might also like