Bablon Et Al.,2019

Download as pdf or txt
Download as pdf or txt
You are on page 1of 55

Accepted Manuscript

Interactions between volcanism and geodynamics in the southern


termination of the Ecuadorian arc

Mathilde Bablon, Xavier Quidelleur, Pablo Samaniego, Jean-Luc


Le Pennec, Laurence Audin, Hervé Jomard, Stéphane Baize,
Céline Liorzou, Silvana Hidalgo, Alexandra Alvarado

PII: S0040-1951(18)30423-2
DOI: https://doi.org/10.1016/j.tecto.2018.12.010
Reference: TECTO 128001
To appear in: Tectonophysics
Received date: 9 May 2018
Revised date: 10 December 2018
Accepted date: 11 December 2018

Please cite this article as: Mathilde Bablon, Xavier Quidelleur, Pablo Samaniego, Jean-
Luc Le Pennec, Laurence Audin, Hervé Jomard, Stéphane Baize, Céline Liorzou, Silvana
Hidalgo, Alexandra Alvarado , Interactions between volcanism and geodynamics in the
southern termination of the Ecuadorian arc. Tecto (2018), https://doi.org/10.1016/
j.tecto.2018.12.010

This is a PDF file of an unedited manuscript that has been accepted for publication. As
a service to our customers we are providing this early version of the manuscript. The
manuscript will undergo copyediting, typesetting, and review of the resulting proof before
it is published in its final form. Please note that during the production process errors may
be discovered which could affect the content, and all legal disclaimers that apply to the
journal pertain.
ACCEPTED MANUSCRIPT

Interactions between volcanism and geodynamics in the southern termination of the


Ecuadorian arc

Mathilde Bablon a*, Xavier Quidelleur a, Pablo Samaniego b, Jean-Luc Le Pennec b,c, Laurence
Audin d, Hervé Jomard e, Stéphane Baize e, Céline Liorzou f, Silvana Hidalgo c, Alexandra
Alvarado c

PT
a
GEOPS, Univ. Paris-Sud, CNRS, Université Paris-Saclay, 91405 Orsay, France

RI
b
Laboratoire Magmas et Volcans, Université Clermont Auvergne, CNRS, IRD, OPGC, F-

SC
63000 Clermont-Ferrand, France
c
NU
Instituto Geofísico, Escuela Politécnica Nacional, Ladrón de Guevara E11-253, Ap. 2759,

Quito, Ecuador
MA

d
Institut des Sciences de la Terre, Université de Grenoble - IRD - CNRS, 1381 Rue de la

piscine, 38400 Saint Martin d’Hères, France


D

e
Institut de Radioprotection et de Sûreté Nucléaire, 31 Avenue de la Division Leclerc, 92260
E

Fontenay-aux-Roses, France
PT

f
Université de Bretagne Occidentale, Domaines Océaniques IUEM, 29280 Plouzané, France
CE

Corresponding author: mathilde.bablon@u-psud.fr (M. Bablon).


AC

1
ACCEPTED MANUSCRIPT

ABSTRACT

This study focuses on the construction and evolution through time of volcanic edifices

located in the southern part of the Ecuadorian arc, with the objective to contribute to a better

understanding of the interactions between magmatism, slab geometry and the activity of

tectonic faults. Our new groundmass K-Ar ages obtained for a dozen volcanoes from the

PT
southern Quaternary arc are rather young, without ages older than 800 ka, and highlight an

increasing volcanic activity between 300 and 100 ka. These new temporal constraints suggest

RI
that a southward migration of the Ecuadorian arc occurred during the last 600 ka. We interpret

SC
this evolution as the result of major and recent geodynamic reorganization in Ecuador related
NU
to the activation of lithospheric faults and to the flexure of the slab at depth, following the

inland prolongation of the Grijalva fracture zone. Both phenomena could have been induced
MA

by the oblique subduction of the Nazca plate towards the convex margin of the South

American continent, and the coupling of these two plates along the Wadati-Benioff zone,
D

progressively intensified by the Carnegie ridge subduction. Southward migration of the slab
E

flexure at depth could have changed pressure and temperature conditions, favoring the partial
PT

melting of the mantle wedge and magma genesis, thereby inducing a southward migration of
CE

the Ecuadorian arc volcanoes. These results contribute therefore to a better knowledge of the

current arc dynamics in the Northern Andes, and provide insights into the use of
AC

geochronological data applied to volcanic rocks for studies of past tectonic activity in

Ecuador.

Keywords:

Ecuador
K-Ar dating
Pallatanga fault
Volcanic arc migration
Slab flexure

2
ACCEPTED MANUSCRIPT

Geodynamics

1. Introduction

The geodynamic setting of continental Ecuador is characterized by a considerable

volcanic activity since the Pleistocene, with more than 80 volcanoes active during the

Quaternary, including 25 active during the Holocene (e.g., Barberi et al., 1988; Hall and

PT
Beate, 1991; Hall et al., 2008; Bernard and Andrade, 2011). Eight of these volcanic edifices

are considered still active, namely with their latest activity during the so-called historical time

RI
(i.e. younger than 500 years BP). Four of them have been active during the last few years:

SC
Reventador, Cotopaxi, Tungurahua and Sangay volcanoes (www.igepn.edu.ec). The
NU
Ecuadorian volcanic arc originates from the subduction of the oceanic Nazca plate beneath the

South American continent (Fig. 1a). The volcanic front position is intimately linked to the
MA

geometry at depth of the subducting plate (e.g., Tatsumi, 1986; Guillier et al., 2001; Syracuse

and Abers, 2006), which in turn is controlled by the structures of the subducted slab and the
D

convexity of the Ecuadorian margin. Among these structures, we can mention the Carnegie
E

ridge, which is the trace of the Galápagos hotspot activity on the Nazca plate, as well as the
PT

Grijalva fracture zone, an escarpment separating two oceanic crust segments of different ages,
CE

older than 30 Ma south and younger than 20 Ma north of the fracture zone (Lonsdale, 2005).

As recently pointed out by Yepes et al. (2016), these features induce a flexure of the slab
AC

under the Quaternary volcanic arc. Consequently, the slab dips at an angle of ~22° north of the

Carnegie ridge, while the dip is more pronounced below the southern termination of the arc

(30-35°), then significantly decrease south of the Grijalva fracture zone, with an angle of 12°

(Yepes et al., 2016). In addition, this part of the Andean chain has a large number of active

crustal faults, notably along the Chingual-Cosanga-Pallatanga-Puná fault system (e.g., Baize

et al., 2015; Alvarado et al., 2016; Fig. 1a), mainly related to the oblique subduction of the

3
ACCEPTED MANUSCRIPT

Nazca plate. Consequently, combined seismic and volcanic hazards in Ecuador are a high

threat for both local populations and infrastructure. The Ecuadorian volcanic arc is, therefore,

particularly suitable for the investigation of interactions between geodynamics, tectonics and

volcanism. Despite a thorough investigation of kinematics along crustal faults (e.g., Winter et

al., 1993; Lavenu et al., 1995; Ego et al., 1996; Tibaldi et al., 2007; Alvarado et al., 2014;

Baize et al., 2015; Alvarado et al., 2016; Champenois et al., 2017), few Quaternary data with

PT
accurate timing control are available for fault activity prior to the Holocene. On the other

RI
hand, focused geochronological data are available for several volcanic edifices, such as

SC
Cayambe (Samaniego et al., 2005), Chacana (Opdyke, 2006), Atacazo-Ninahuilca (Hidalgo,

2006; Hidalgo et al., 2008), Pichincha (Robin et al., 2010), Imbabura (Le Pennec et al., 2011),
NU
Chimborazo (Samaniego et al., 2012), Chachimbiro (Bernard et al., 2014), Antisana (Hall et

al., 2017) and Tungurahua volcanoes (e.g., Hall et al., 1999; Le Pennec et al., 2013; Bablon et
MA

al., 2018a). Although some studies also focused on the development of the Ecuadorian arc at a

more regional scale (e.g., Hall and Wood, 1985; Barberi et al., 1988; Hall and Beate, 1991;
D

Opdyke et al., 2006), the overall chronology of the arc development is still poorly
E

documented.
PT

The present study combines new geochronological data of volcanic activity and
CE

mapping of the main faults in the southern termination of the Ecuadorian arc (Monzier et al.,

1999a), where numerous volcanic edifices and active faults appear to have been concomitant
AC

or influencing each other. This work aims to examine the potential influence of the

geodynamics on the arc development, and to contribute to a better understanding of the large

increase of volcanic activity in Ecuador during the Quaternary.

2. Geological and geodynamical context of the Ecuadorian arc

2.1. Orogeny and tectonics of the Ecuadorian Andes

4
ACCEPTED MANUSCRIPT

The subduction of the former oceanic Farallón plate beneath the continental South

American lithosphere began during the Late Triassic - Early Jurassic period (James, 1971;

Aspden et al., 1987). Between ca. 75 and 55 Ma, oceanic terranes were accreted to the

Ecuadorian and Colombian margins in several tectonic episodes (e.g., Spikings et al., 2001;

Kerr et al., 2002; Hughes and Pilatasig, 2002; Jaillard et al., 2008; Jaillard et al., 2009). The

current Andean range originates from the subduction of the Nazca plate, formed after the

PT
breakup of the older Farallón plate between 23 and 27 Ma, related to the activity of the

RI
Galápagos spreading center (Hey, 1977; Lonsdale and Klitgord, 1978; Pennington, 1981;

SC
Meschede and Barckhausen, 2001; Sallarès and Charvis, 2003; Lonsdale, 2005).

The width of the Quaternary volcanic arc ranges between ~60 and 150 km in front of
NU
the Carnegie Ridge (Figure 1a), and the volcanoes are distributed north of 2°S in Ecuador.

They mostly lie to the north of the inland prolongation of the Grijalva fracture zone, located
MA

roughly below Chimborazo and Carihuairazo-Puñalica volcanoes (Fig. 1b). However, some

edifices from the southern termination of the arc, such as Sangay, Altar, Igualata and
D

Tungurahua volcanoes, as well as Puyo, Calpi and Licto cones, are located south of the inland
E

prolongation of the Grijalva fracture zone (Yepes et al., 2016; Ancellin et al., 2017; Narvaez
PT

et al., 2018). Ecuadorian volcanoes are distributed in the Western and Eastern Cordilleras, the
CE

Interandean Valley and the back-arc area (e.g., Hall and Beate, 1991; Hall et al., 2008). The

volcanic front of the Western Cordillera is mainly composed of andesitic to dacitic compound
AC

stratovolcanoes, while the intermediate arc, including the Interandean Valley and the Eastern

Cordillera is mostly made up of andesitic stratovolcanoes. The calc-alkaline magmatic series

of these volcanoes contrasts with the alkaline products emitted from back-arc edifices (e.g.,

Hall et al., 2008; Fig. 1b). The basement of the Western Cordillera and the western part of the

Interandean Valley is composed of accreted oceanic terranes (e.g., Jaillard et al., 2009;

Fig. 1b), whereas the Eastern Cordillera basement is made up of Paleozoic to Mesozoic

5
ACCEPTED MANUSCRIPT

plutonic and metamorphic rocks with a mostly continental affinity. Lastly, the back-arc is

composed of late Cretaceous sedimentary deposits, which overlie Paleozoic-Lower

Cretaceous metamorphic rocks and the continental Precambrian craton (Coltorti and Ollier,

2000; Barragán et al., 2005; Hall et al., 2008). Moreover, the fast (~6 cm.a-1; Trenkamp et al.,

2002; Kendrick et al., 2003; Nocquet et al., 2014) and oblique (~N81°E; Kendrick et al.,

2003; Bird, 2003; DeMets et al., 2010; Nocquet et al., 2014) subduction of the Nazca plate,

PT
combined with the subduction of the Carnegie Ridge (Pennington et al., 1981; Gutscher et al.,

RI
1999; Witt et al., 2006; Fig. 1a), produce a displacement of the northwestern part of the South

SC
America continent (e.g., Egbue and Kellogg, 2010; Nocquet et al., 2014). This domain,

previously called the “Andean Block” (Pennington et al., 1981; Ego et al., 1996), and then the
NU
“North Andean Block” (Kellogg and Vera, 1995; Witt et al., 2006; Alvarado et al., 2014), has

been recently introduced as “North Andean Sliver” (Nocquet et al., 2014; Alvarado et al.,
MA

2016; Fig. 1a) in order to properly depict the internal crustal deformation of the block inferred

from geodesy. Paleoseismological, geodetic and field studies propose that the deformation
D

generated by this displacement is accommodated along a mature dextral strike-slip fault zone,
E

called the Chingual-Cosanga-Pallatanga-Puná (CCPP) fault system (Baize et al., 2015;


PT

Alvarado et al., 2016; Fig. 1a and b), with a current geodetic bulk velocity of ~8 mm.a-1
CE

(Nocquet et al., 2014). This fault system extends from the Gulf of Guayaquil to the Caribbean

Sea, and roughly corresponds to the previous Dolores-Guayaquil Megashear (e.g., Lavenu et
AC

al., 1995). In Ecuador, it crosses the Western Cordillera between ~2°S and ~1.7°S (Fig. 1a),

the Interandean Valley around ~1.5°S, through Igualata volcano (Fig. 1b), then goes along the

Eastern Cordillera north of 1°S. The onset of activity of the CCPP fault is still debated but

seems to have occurred during the Pliocene - Early Pleistocene (Lavenu et al., 1995). Indeed,

offshore sedimentary records indicate that the opening phase and high sedimentation rate of

the Gulf of Guayaquil occurred during the Early Pleistocene (Deniaud et al., 2001; Witt et al.,

6
ACCEPTED MANUSCRIPT

2006), while the Miocene fore-arc marine basins were uplifted during the Pliocene and filled

up with continental deposits during the Pleistocene (Spikings et al., 2001; Alvarado et al.,

2016). Local sections of this fault were studied onshore (Winter et al., 1993; Lavenu et al.,

1995; Dumont et al., 2005; Tibaldi et al., 2007; Alvarado et al., 2014; Baize et al., 2015;

Champenois et al., 2017), and 14C dating revealed that the fault system has been continuously

active during the Holocene, with an average slip rate of 2.5 to 5 mm.a-1 for the Pallatanga fault

PT
segment (Winter et al., 1993; Baize et al., 2015). Finally, the collision age of the Carnegie

RI
Ridge (Pennington et al., 1981; Gutscher et al., 1999; Witt et al., 2006; Fig. 1a) to the

SC
Ecuadorian margin is still debated and ranges between 1 and 15 Ma (e.g., Lonsdale and

Klitgord, 1978; Gutscher et al., 1999; Spikings et al., 2001). However, many evidences
NU
suggest that the collision occurred during the last 5 Ma, such as paleogeographic

reconstructions (Collot et al., 2009), sedimentation rates of the Gulf of Guayaquil (Deniaud et
MA

al., 2001; Witt et al., 2006), the uplift of the margin and basins (e.g. Gutscher et al., 1999;

Graindorge et al., 2004; Pedoja et al., 2006), as well as the volcanism distribution and
D

geochemistry (e.g., Hall and Wood, 1985; Barberi et al., 1988; Samaniego et al., 2010).
E
PT

2.2. Geological context of the southern termination of the Ecuadorian arc

The southern termination of the Ecuadorian arc is composed of a dozen volcanic


CE

edifices, active during the Plio-Quaternary (Hall and Wood, 1985; Barberi et al., 1988; Hall
AC

and Beate, 1991; Hall et al., 2008). This group of volcanoes is separated from the central and

northern part of the arc by a 50-100 km-long gap without active volcanism. A review of the

previously published data for these edifices is presented below, from the Western to the

Eastern Cordillera.

Sagoatoa volcano (4169 m above sea level (a.s.l.); Lat. 01°09’S; Long. 78°04’W) is

located north of our study area, on the western edge of the Interandean Valley. This volcano is

part of a large complex, with Pilisurco volcano (4508 m; Fig. 1b) located to the west. The

7
ACCEPTED MANUSCRIPT

erosion having affected its flanks appears important, with smooth surfaces incised by deep

valleys. Two K-Ar ages have been published, at 1.73 ± 0.35 and 1.40 ± 0.29 Ma (Lavenu et

al., 1992). Nevertheless, these ages should be considered with caution as they were carried out

on plagioclases and whole-rock, respectively, which, hence, may include phenocrysts

containing inherited radiogenic argon and/or weathered areas, possibly leading to erroneously

old ages (e.g., Harford et al., 2002; Samper et al., 2008).

PT
South of Sagoatoa, the activity of Carihuairazo volcano (5018 m a.s.l.; Lat. 01°24’S;

RI
Long. 78°45’W) began at least ~225 ka (Samaniego et al., 2012). In the Ambato basin,

SC
northeast of the volcano (Fig. 2a), four sequences of sector collapse and block-and-ash

deposits are interbedded with plinian fallout deposits from Huisla volcano. The debris
NU
avalanche deposits (DAD) of Carihuairazo located at the base of the sequence are older than

45 ka; the second sector collapse occurred before ~40 ka; and the more recent tephra fallout
MA

and debris avalanche deposits, which cover the sequence, are younger than ~40 ka based on
14
uncalibrated C age determinations (Clapperton, 1990; Ordóñez, 2012). Moreover, these
D

deposit sequences cover the deposits of the Chalupas ignimbrite, a major Ecuadorian
E
PT

stratigraphic marker dated at ~215 ka (Beate et al., 2006; Bablon et al., 2018b). Volcanic

activity of Carihuairazo ended with the growth of several domes on its eastern flank (Ordóñez,
CE

2012).

Puñalica (3988 m a.s.l.; Lat. 01°24’S; Long. 78°41’W; Fig. 2a), a morphologically
AC

fresh-looking edifice, is located on the northeast flank of Carihuairazo volcano, and may

correspond to its last volcanic activity (Clapperton, 1990). Puñalica is made up of basic

volcanic products that cover the older moraines of the Last Glacial Maximum (LGM), but do

not crop out on the youngest ones, which suggests that some parts of the volcano were

constructed between 14 and 18 ka (Clapperton, 1990).

8
ACCEPTED MANUSCRIPT

Further south, the activity of Chimborazo volcano (6268 m a.s.l.; Lat. 01°28’S; Long.

78°49’W) began at ~120 ka, and is considered still active (Clapperton, 1990; Barba et al.,

2008; Bernard et al., 2008; Samaniego et al., 2012). The volcano is composed of three

successive edifices, and experienced a major sector collapse at the end of Chimborazo I

construction, at ~60-65 ka (Samaniego et al., 2012). The debris avalanche deposit spread out

into the Riobamba basin (Fig. 2d), incorporating older deposits of the Chalupas ignimbrite.

PT
The Guano lava flow, in the northern edge of the basin (Fig. 2d), corresponds to volcanic

RI
products erupted after the sector collapse, as it locally overlies the avalanche deposit. The

SC
proximal outcrops of this debris avalanche deposit are covered by LGM moraines, suggesting

that both the Guano lava flow and the sector collapse are older than 33 ka (Samaniego et al.,
NU
2012). Moreover, ash flow deposits from Chimborazo II dated at ~43 ka also cover the sector

collapse deposits, implying that the latter is older than 43 ka (Samaniego et al., 2012).
MA

Located on the western edge of the Interandean Valley, northwest of Tungurahua

volcano (Fig. 1b), Huisla volcano (3763 m a.s.l.; Lat. 01°24’S; Long. 78°34’W) is an
D

andesitic edifice, previously called Cerro Llimpi (Stübel, 1897). On the basis of isopach maps
E
PT

and geochemistry, at least three tephra fallout deposits, exposed in the Ambato basin,

northwest of the volcano (Fig. 2), have been assigned to Huisla (Ordóñez, 2012). The two
CE

younger plinian eruptions occurred in the last 39 ka (Ordóñez, 2012). This edifice experienced
AC

a sector collapse of its southeastern flank (Espín et al., 2018), whose resulting debris

avalanche deposits cover the Chalupas ignimbrite (Bustillos, 2008; Espín, 2015). However, no

geochronological data from lava flows are available for this volcano, nor for the surrounding

volcanic edifices, such as Mulmul (3878 m a.s.l.; Lat. 01°26’S; Long. 78°33’W) and Igualata

(4430 m a.s.l.; Lat. 01°30’S; Long. 78°38’W) volcanoes.

The volcanic arc in the Interandean Valley ends to the south with the Calpi and Licto

scoria cones, in the southern edge of the Riobamba basin (Fig. 2d). Only geochemical data are

9
ACCEPTED MANUSCRIPT

available for these edifices (Ancellin et al., 2017); their periods of activity are still unknown

but stratigraphic evidences point out to an age older than the debris avalanche deposit of

Chimborazo volcano (Bernard et al., 2008; Clapperton, 1990).

In the Eastern Cordillera, only Tungurahua volcano (Fig. 1b) was studied in detail.

(e.g., Hall et al., 1999; Le Pennec et al., 2006; Bablon et al., 2018a). Its activity began at

~300 ka, and the volcano was erupting between 1999 and 2016. Made up of three successive

PT
edifices, it experienced two major sector collapses, at ~35 and ~3 ka (Hall et al., 1999; Le

RI
Pennec et al., 2013). In the Río Chambo valley, southwest of the volcano, the oldest avalanche

SC
deposit covers another avalanche deposit originated from Altar volcano (Bustillos, 2008).

However, the period of activity of the latter remains undocumented.


NU
The construction of Sangay volcano, the southernmost active volcano of the arc, began

at ~500-400 ka (Monzier et al., 1999b). This mainly andesitic stratovolcano is made up of


MA

three edifices and experienced two undated major flank collapses (Valverde, 2004).

Finally, the Puyo cones, made up of alkali basalts, were emplaced in the back-arc
D

around 200 ka (Hoffer et al., 2008).


E
PT

In this study, we present twenty-four new K-Ar ages on groundmass, and their corresponding

major and trace element contents, performed on lava flows from eight different volcanoes
CE

(Huisla, Mulmul, Igualata, Carihuairazo, Chimborazo, Sagoatoa, as well as Licto and Calpi
AC

cones; Fig. 1b), to investigate the temporal evolution of the southern termination of the

Ecuadorian arc.

3. Materials and methods

3.1. Rock sampling for K-Ar dating and geochemical analyses

10
ACCEPTED MANUSCRIPT

Twenty-six lava flows were sampled from several edifices (Fig. 2), mostly located

between Chimborazo and Tungurahua volcanoes (Fig. 1b). Whenever possible, we sampled

outcrops located in the deep incised valleys and in the summit area of the edifices, in order to

determine their whole period of activity.

Three lava flows were sampled from Sagoatoa volcano. Sample 16EQ03 is from a

distal lava flow located in the Río Cutuchi valley, northwest of Píllaro town (Fig. 2a). This

PT
lava flow is covered by the Chalupas ignimbrite, while samples 16EQ40 and 16EQ41 are

RI
located near the summit of the volcano.

SC
We sampled proximal (16EQ43) and distal (16EQ28) areas of the Guano lava flow

(Fig. 2d), which originated from Chimborazo volcano (Samaniego et al., 2012), and followed
NU
the southern flank of Igualata, north of the Riobamba basin. As this lava flow is cut and

slightly shifted by the Pallatanga fault (Baize et al., 2016), its age might bring valuable
MA

insights regarding the period of activity and the velocity of the fault displacement.

Two lava flows were sampled in order to constrain the timing of the construction of
D

Carihuairazo volcano. Sample CAR-14 (Fig. 2a) is from Tzunantza dome (also spelled Cerro
E
PT

Sunantza; Ordóñez, 2012), located at the foot of its northern flank, while RIO-14 is from the

young Puñalica edifice.


CE

The morphology of Igualata (Fig. 2d) is rather similar to that of Sagoatoa volcano,

except for its large E-W trending summit depression. This depression is similar to a pull-apart
AC

structure controlled by the NE-SW dextral Pallatanga fault. Samples 16EQ22 and 16EQ24 are

from lava flows interbedded with breccia layers, in the upper part of the northern flank, north

of this structure. Sample 16EQ23 is from a lava flow located at the summit of the volcano.

Because of agricultural fields and abundant vegetation present in the southern flank, we only

sampled two rocks, in the Patalú valley, which drains the southwest flank of Igualata volcano

down to the Riobamba basin (Fig. 2d). Sample 16EQ30 is from a juvenile block from a

11
ACCEPTED MANUSCRIPT

~10 m-thick block-and-ash flow deposit, which covers an older lava flow (sample 16EQ29),

located at the bottom of the valley.

Two basal lava flows were also sampled in the Río Chambo valley, at the foothill of

Mulmul (l6EQ14) and Igualata (16EQ27) volcanoes (Fig. 2). They are mapped as deposits

from the Mio-Pliocene Pisayambo formation (Litherland et al., 1993; Hughes and Pilatasig,

2002), and may represent an old and extinct volcanic activity. The outcrop of 16EQ27, more

PT
than 150 m-wide, presents tilted blocks, several dykes and numerous vesiculated enclaves

RI
with a phaneritic texture made up of plagioclase and amphibole phenocrysts (see Appendix A

SC
for photographs of thin sections).

Five lava flows were sampled in basal areas or in deep incised valleys on northern
NU
(16EQ04, RIO-111) and western (RIO-107, 16EQ05, 16EQ07) flanks of Huisla volcano

(Fig. 2b). Unfortunately, no lava from the terminal activity crops out at the summit of this
MA

edifice. We sampled four lava flows from Mulmul volcano (Fig. 2b), which partly grew in the

collapse amphitheater of Huisla volcano. Sample 16EQ08 is from a monogenetic breccia


D

deposit containing juvenile blocks included in an ash-rich matrix, which was probably related
E
PT

to a dome collapse, while samples 16EQ09, 16EQ10 and 16EQ11 are from lava flows from

the northwestern flank.


CE

Finally, three cones located in the Riobamba basin were sampled (Fig. 2). Samples

16EQ34, 16EQ35 and 17EQ114 are from slightly vesicular lava flows from Calpi cones,
AC

located along the Pallatanga fault, north-west of Riobamba city. The two former correspond to

proximal lava flows, while 17EQ114 is a distal lava flow covered by the debris avalanche

deposits of Chimborazo volcano. Sample 16EQ47, located in the southernmost part of our

study area, is from a lava flow from Licto cone (Cerro Tulabug), which is covered by varved

lacustrine deposits associated with the distal parts of the Chimborazo DAD.

12
ACCEPTED MANUSCRIPT

The freshest samples, selected after a careful examination of thin sections (Appendix A), were

dated by the potassium-argon (K-Ar) dating method applied to the groundmass, while the

whole-rock major and trace element contents were measured for all samples.

3.2. K-Ar dating method and protocol

The K-Ar dating method was applied to the groundmass, using the unspiked

PT
Cassignol-Gillot technique (Cassignol and Gillot, 1982). This technique was developed for

Quaternary volcanoes, whose lavas contain low radiogenic argon (40Ar*; Gillot et al., 2006). It

RI
has been shown to be particularly suitable for dating young calk-alkaline lavas, such as in

SC
Ecuador for Tungurahua volcano (Bablon et al., 2018a), as well as in Argentina or in the

Lesser Antilles (Samper et al., 2009; Germa et al., 2010; Germa et al., 2011; Ricci et al.,
NU
2015). The description of sample preparation and analytical procedures, standards used and
MA

uncertainty calculations are given in Bablon et al. (2018a). Samples were crushed and sieved

to 63-80, 80-125 or 125-250 µm, based on the size of the phenocrysts, which must be
D

removed, and on the groundmass proportion. The technique relies on the detection of the very
E

40
small difference between the isotopic Ar/36Ar ratio of the groundmass extracted from the
PT

sample and the atmospheric ratio. The result corresponds to the quantity of 40Ar* produced by

the radioactive decay of 40K since the eruption. Therefore, together with the potassium content
CE

of the groundmass and the 40K decay constants (Steiger and Jäger, 1977), we can calculate the
AC

age of the lava sample. Both potassium and argon measurements were carried out at the

GEOPS laboratory in Orsay (Paris-Sud University, France). All measurements were

performed at least twice in order to check their reproducibility within uncertainty, except for

samples 16EQ28, 16EQ34 and 16EQ35, for which the argon content was measured five times,

in order to improve the precision calculated for the mean age. As the five ages obtained for

each aliquot are consistent within uncertainty, we can calculate the mean age by averaging

each analysis, weighted by the inverse of its variance. The final age uncertainty of these three

13
ACCEPTED MANUSCRIPT

samples is obtained as the reciprocal square root of the five reciprocal variances sum (Taylor,

1997). Ages reported throughout this study are given at the 1-σ confidence level.

3.3. Geochemical analyses

The analytical procedure for measurement of major and trace element content is

detailed in Cotten et al. (1995). Agate-crushed powders of the twenty-three whole-rock

PT
samples were analyzed by ICP-AES (Inductively Coupled Plasma - Atomic Emission

Spectrometry), at the LGO (Laboratoire Géosciences Océan) of the Université de Bretagne

RI
Occidentale (Brest, France). Relative uncertainties are lower than 2% and 5%, for major and

SC
trace elements, respectively. Major element concentrations were recalculated to a total of

100% on a water free-basis. Major and trace element concentrations are given in Table 2.
NU
MA

4. Results

4.1. K-Ar groundmass dating


D

Twenty-five K-Ar ages obtained in this study are presented in Table 1 and Fig. 2. All
E

samples have porphyritic textures with a variable amount of plagioclase, ortho- and
PT

clinopyroxene, Fe-Ti oxides, olivine, and amphibole phenocrysts (Appendix A). Overall, the
CE

groundmass represents about up to 50% of the rocks, and generally contains microlites of

plagioclase, pyroxene and Fe-Ti oxides into a glassy matrix. The K content of the separated
AC

groundmass ranges between 1.3 and 3.1 wt.%, and radiogenic argon content ranges between

0.04% and 24.88%. The mean density of the dated fractions is about 2.66 g.cm-3. The ages

obtained for each edifice are detailed below, from the Western to the Eastern Cordillera.

Both samples from Sagoatoa volcano have similar ages: 826 ± 12 ka for the summit

(16EQ40), and 799 ± 12 ka for the eastern distal lava flow (16EQ03; Fig. 2a). These ages

14
ACCEPTED MANUSCRIPT

probably represent the terminal activity of the volcano, and, although we do not have any lava

flow from its initial stages, Sagoatoa volcano is the oldest edifice of our study.

Further south, the sample from Tzunantza dome (CAR-14; Fig. 2a), to the north of

Carihuairazo volcano, is dated at 512 ± 9 ka. Given that Carihuairazo lavas have been dated

at ~225 ka by Samaniego et al. (2012), this age may correspond to an older, pre-Carihuairazo

activity. The volcanic activity around Carihuairazo area may have ended with the formation of

PT
Puñalica edifice, at 18 ± 3 ka (RIO-18).

RI
Guano lava flow unit samples, related to the post-collapse activity of Chimborazo,

SC
show two significantly different ages, which indicate that the proximal and distal areas do not

belong to a single lava flow. Indeed, the proximal sample yields 30 ± 3 ka (16EQ43), while
NU
the distal sample suggests the existence of a Holocene lava flow emission from Chimborazo

volcano, with an age of 4 ± 2 ka (16EQ28; Fig. 2c).


MA

In the Interandean Valley, Igualata volcano seems to be older than the neighboring

volcanoes, based on the high erosion of its flanks and its morphology. However, the oldest
D

sample is dated at 376 ± 10 ka (16EQ23; Fig. 2d), although we could not sample the base of
E
PT

this edifice, due to the lack of outcrops. On the southwestern flank, the lava flow located in

the valley incised along the Pallatanga fault is dated at 237 ± 9 ka, while the pyroclastic flow
CE

deposit sequence, dated at 107 ± 11 ka (Fig. 2d), is the most recent outcrop dated from this
AC

volcano. Eastward, the two lava flows sampled as basement lavas yielded ages of 358 ± 6 and

371 ± 7 ka, for sample 16EQ27 located south of Igualata volcano, and for sample 16EQ14

located on the eastern foot of Mulmul volcano, respectively. Both ages are too young to be

associated with the old volcanic basement (Litherland et al., 1993; Hughes and Pilatasig,

2002), and they are in the same range as those obtained for Igualata volcano, with a similar

geochemical composition (see below). Thus, these lava flows should be related to the base of

this edifice.

15
ACCEPTED MANUSCRIPT

Huisla volcano appears older than Igualata and coeval with Tzunantza dome, north of

Carihuairazo, with five ages ranging between 612 ± 10 ka (16EQ07) and 492 ± 9 ka (RIO-

111; Fig. 2b). The ages obtained for Mulmul volcano are consistent with the fact that it partly

grew within the sector collapse amphitheater of Huisla volcano. Indeed, ages are younger than

those of Huisla, and range between 174 ± 3 ka (16EQ08) and 145 ± 4 ka (16EQ11; Fig. 2b).

Finally, the timing of construction of Licto and Calpi cones, within the Riobamba

PT
basin, is significantly different. In fact, the lava flow of Licto cone is dated at 183 ± 9 ka

RI
(16EQ47), while the proximal lava flows of Calpi monogenic cones display ages of 9 ± 3 and

SC
8 ± 5 ka (16EQ34 and 16EQ35, respectively; Fig. 2c). The distal lava flow of Calpi

(17EQ114), covered by the debris avalanche deposits of Chimborazo, was emitted at


NU
62 ± 4 ka. Consequently, Calpi cones could have been constructed in at least two stages. The

oldest stage occurred before the major sector collapse event of Chimborazo volcano
MA

(Clapperton, 1990; Bernard et al., 2008; Samaniego et al., 2012), and the youngest stage

corresponds to the construction of the cones during the Holocene. Detailed field investigations
D

on this part are needed in order to better support this hypothesis.


E
PT

4.2. Whole rock geochemical analyses


CE

Geochemical analyses of the twenty-seven samples show that they belong to the

medium-K calc-alkaline series in the K2O vs. SiO2 diagram (Peccerillo and Taylor, 1976;
AC

Fig. 3a). Sample RIO-111, from Huisla volcano, plots in the limit between the medium and

high-K calc-alkaline series. Silica contents range between 52.8 and 65.9 wt.%, and K2O

content between 1.5 and 2.6 wt.%. Hence, lavas from this study are mainly basaltic andesites

and andesites, with some dacitic rocks, such as the youngest lava flow of Huisla (RIO-111),

the lava flow at the summit of Sagoatoa (16EQ40), and the pyroclastic flow deposit of

Igualata volcano (16EQ30; Fig. 3a). The Licto cone lava flow is a basaltic andesite, as well

16
ACCEPTED MANUSCRIPT

as those from Puñalica (RIO-18) and from the southern Calpi cone (16EQ35), while other

samples from Calpi cones (16EQ34 and 17EQ114) are andesitic lavas with silica content of

~61%. All samples from Huisla volcano plot in the field of both in-situ lava flows and blocks

of the collapse deposits (Bustillos, 2008; Fig. 3a). Finally, we note that most samples have

lower potassium content than Tungurahua lava flows (Fig. 3a), located in the Eastern

Cordillera. Indeed, the content of most of the incompatible elements, including K, increases

PT
with the distance from the trench, interpreted as a lower degree of mantle partial melting

RI
(Barragán et al., 1998; Hidalgo et al., 2012; Ancellin et al., 2017).

SC
Spider diagrams of trace elements normalized to the primitive mantle, and diagrams

of Rare Earth Elements (REE) normalized to chondrites (Sun and McDonough, 1989; Fig. 3b
NU
and 3c), show enrichment of Large-Ion Lithophile Elements (LILE; Rb, Ba, and K) and Light

REE (LREE; La, Ce, Nd), as well as a depletion of High-Field Strength Elements (HFSE;
MA

Nb, Ti and Y). These patterns are typical of arc magmas. The content of incompatible

elements is similar to our samples, except for Y and the heavy REE (HREE; Dy, Er, Yb),
D

which are enriched in samples of Mulmul and Sagoatoa volcanoes (red and pink lines
E

respectively; Fig. 3b and 3c). Finally, the content of major, trace and incompatible elements
PT

is similar for both Guano lava flows.


CE

Crustal processes such as fractional crystallization govern the composition of lava

flows, yielding a negative correlation between compatible elements (e.g., MgO, Fe2O3, CaO,
AC

Cr, Ni; Fig. 4a and b) and SiO2, due to the crystallization of olivine, pyroxene, Fe-Ti oxide

and plagioclase phenocrysts in the magma chamber, and a positive correlation for Na2O

(Fig. 4c) and for incompatible elements (e.g., Rb and Ba; Fig. 4d). These trends are in

agreement with previous analyses carried out on other volcanoes from the southern part of

the arc (grey dots in Fig. 4), including lavas from Tungurahua, Chimborazo, Huisla, and

Sangay. However, there is no single and unequivocal interpretation about the origin of the

17
ACCEPTED MANUSCRIPT

temporal variation of the geochemistry, which may be related to changes in the deep mantle

source or crustal processes. Our data, despite age differences, reflect the same macro-process

related to the subduction context, and do not result from local, superficial or minor

phenomena. For that reason, we prefer not to use these major and trace element analyses for

interpretative purposes about volcanism-geodynamics relationships.

PT
5. Discussion

RI
5.1. Comparison with previously published data

SC
Our new ages (Table 1) are in the range of those previously reported at Carihuairazo,

Chimborazo and Tungurahua volcanoes (Barba et al., 2008; Samaniego et al., 2012; Le
NU
Pennec et al., 2013; Bablon et al., 2018a). However, both ages from Sagoatoa volcano are
MA

younger than published K-Ar whole-rock and plagioclase ages (1.73 ± 0.35 and 1.40 ± 0.29

Ma respectively; Lavenu et al., 1992), obtained for the same lava flow as 16EQ03

(799 ± 12 ka), and for a lava flow located on the eastern flank of the volcano, respectively.
E D

This large difference of about 1 Ma can be explained by the presence of inherited radiogenic
PT

argon in the phenocrysts, which are necessarily included in the whole-rock analysis. Indeed, if

they were incorporated late in the magma chamber, they did not have time to be reset before
CE

the eruption, hence producing a bias towards older values for the whole-rock ages. Whole-
AC

rock samples may also contain concealed weathered areas, where the potassium may have

been leached, and, hence, this will also induce too old ages.

The age obtained for Tzunantza dome (512 ± 9 ka), north of Carihuairazo volcano, is

significantly older than lava flows of its southern flank, dated at ~225 ka (Samaniego et al.,

2012). We thus propose that this new age may correspond to an older, pre-Carihuairazo

volcanic activity. Moreover, the lava sample of Puñalica edifice, dated at 18 ± 3 ka, confirms

18
ACCEPTED MANUSCRIPT

that it was constructed at the end of the LGM, before the emplacement of the younger

moraines during the Younger Dryas (10-12 ka; Clapperton, 1990).

Further to the south, the Riobamba basin is filled by 10-12 km3 of debris avalanche

deposits from the huge sector collapse of Chimborazo volcano (Bernard et al., 2008;

Samaniego et al., 2012). The avalanche deposit related to this event is locally covered by the

distal Guano lava flows unit, the LGM moraines (18-25 ka), and the Río Blanco ash flow

PT
sequence (42-43 ka; Samaniego et al., 2012). The age of 30 ± 3 ka, obtained here for a

RI
proximal lava flow from the Guano unit, yields a maximum age for the LGM (Clapperton,

SC
1990). Moreover, it implies that the sector collapse occurred before 30 ka, which is roughly in

agreement with the stratigraphic data and the ages of Samaniego et al. (2012). However, the
NU
age obtained for the distal Guano lava flow (4 ± 2 ka), whose levees are well preserved

(Fig. 5a), is significantly younger than the previously published 40Ar/39Ar groundmass age of
MA

60 ± 11 ka (sample RIO-5; Samaniego et al., 2012). Unfortunately, this age is not supported

by any 40Ar/39Ar age spectra, nor inverse isochron. Furthermore, it was the only age from that
D

study that was obtained using stepwise laser heating, hence without temperature control (M.
E
PT

Fornari, pers. com.). The lack of detail information prevents us from explaining such large
40
differences, but we note that the groundmass fraction used for Ar/39Ar dating was selected
CE

by hand, without using heavy liquids to obtain a narrow density range (2.53-2.60), as was the
39
AC

case here for K-Ar dating. Similarly, undetected Ar recoil, occurring during irradiation,
40
could also have significantly biased the Ar/39Ar results. The elevation profile of the valley

across the distal Guano lava flow (Fig. 5c) shows that a ~120 m-thick sequence of the

Chimborazo avalanche deposits, which also crop out on the southern flank of Igualata volcano

(Bernard et al., 2008), was eroded by the San Andrés river before the emission of this lava

flow. The fact that the base of the flow is at the same elevation as the present day bottom of

the valley (Fig. 5c) strongly suggests that the river had no time to further incise the valley, and

19
ACCEPTED MANUSCRIPT

argues in favor of the very young age of 4 ± 2 ka obtained from our measurements (16EQ28;

Table 1). Finally, this age implies that Guano lava flows were emitted as two different units.

An older proximal unit was erupted after the sector collapse of Chimborazo and before the

LGM period, whereas a younger unit includes the distal lava flow. This suggests that a pulse

of lava emission occurred during the Holocene, probably associated with the explosive

activity reported by Barba et al. (2008).

PT
It was previously described that the Chimborazo collapse deposits cover scoria fallouts

RI
from Calpi cones located southwest of the Riobamba basin (Bernard et al., 2008). New field

SC
observations indicate that a distal lava flow can be also related to the activity of Calpi cones.

The distal lava flow from Calpi cones is unambiguously covered by the debris avalanche and
NU
tephra fallout deposits from Chimborazo volcano, while Calpi proximal lavas do not seem

covered by any volcanic product. Consequently, Calpi cones could have been constructed in at
MA

least two stages. The remnants of the oldest stage, represented by the distal lava flow that lies

under the collapse deposits, were emitted at 62 ± 4 ka (17EQ114), while we have dated the
D

youngest stage at 9 ± 3 and 8 ± 5 ka (16EQ34 and 16EQ35; Table 1), which is responsible of
E
PT

the formation of the cones and proximal lava flows. Directly overlying the lava flow of the

oldest construction stage, the debris avalanche deposits emplaced during the Chimborazo
CE

major sector collapse have therefore a well constrain maximum age of 62 ± 4 ka. This new
AC

age confirms the previous hypothesis that the sector collapse event occurred ~65-60 ka, based

on stratigraphy and 40Ar/36Ar ages of Chimborazo lava flows (Samaniego et al., 2012).

Although no lava flow from Huisla volcano was previously dated, three plinian fallout

deposits, exposed in the Ambato basin (Fig. 2a), have been attributed to the late activity of

this volcano (Ordóñez, 2012). They contain grey and white pumice clasts, crystals of

plagioclase, amphibole, pyroxene and biotite, as well as some metamorphic clasts with a

petrography close to the basement of the Eastern Cordillera, and andesite clasts similar to lava

20
ACCEPTED MANUSCRIPT

flows of the Huisla-Mulmul complex (Ordóñez, 2012) and incorporated during magma ascent.
14
According to uncalibrated C age determinations, the two younger tephra fallout deposits

were erupted during the last 40 ka, while the first plinian eruption is older (Ordóñez, 2012).

However, our ages indicate that Huisla volcano was constructed between 612 ± 10 ka and

492 ± 9 ka, and collapsed before the onset of Mulmul construction, dated here at ~180 ka

(Fig. 5b; Table 1). As there is no evidence of activity for Huisla volcano after its sector

PT
collapse, the proposed age of the tephra fallout deposits, in particular for the two younger

RI
plinian eruptions, is therefore much too young to be associated with Huisla volcano. On the

SC
other hand, Mulmul volcano is younger than 180 ka, with major element compositions of its

lavas similar to those of Huisla (Fig. 3a; Fig. 4). Therefore, we suggest that plinian fall
NU
deposits of the Ambato basin may originate from a late activity of Mulmul volcano. This

hypothesis remains consistent with the isopach maps (Ordóñez, 2012), although no proximal
MA

tephra fallout deposit seems to crop out in the vicinity of Mulmul. Moreover, based on

stratigraphy, Bustillos (2008) and Espín (2015) showed that Huisla sector collapse is younger
D

than the Chalupas ignimbrite, dated at ~215 ka (Beate et al., 2006; Bablon et al., 2018b).
E
PT

Since this destabilization occurred before the construction of Mulmul volcano, which is partly

constructed inside the amphitheater, we therefore propose that the sector collapse of Huisla
CE

volcano occurred between 215 and 180 ka.


AC

5.2. Edifice history of the southern arc volcanism

Based on our new data and previous studies of Chimborazo, Tungurahua and Sangay

volcanoes (Monzier et al., 1999b; Samaniego et al., 2012; Bablon et al., 2018a), a

reconstruction of the eruptive history of the E-W section from the southern Ecuadorian arc can

be proposed as follows.

21
ACCEPTED MANUSCRIPT

The development of the southern part of the arc began with the construction of the

andesitic Huisla volcano in the Interandean Valley, around 600 ka. Before, the southernmost

edifice of the Middle Pleistocene arc was probably Sagoatoa volcano, which was constructed

before 800 ka (Fig. 6a and b). Huisla volcano activity ended ~500 ka ago, when andesitic

domes were emplaced at the current location of Carihuairazo volcano, in the Western

Cordillera. This activity was probably followed by the onset of Igualata and Sangay

PT
construction, older than 400 ka (Monzier et al., 1999b). At 300 ka, the end of construction of

RI
the main edifice of Igualata volcano (Fig. 6c) was coeval with the onset of Tungurahua

SC
activity, in the Eastern Cordillera (Fig. 6d; Bablon et al., 2018a). Then, the volcanic activity

seems to increase between 300 and 100 ka (Fig. 6d and e), with the construction of
NU
Carihuairazo (Samaniego et al., 2012), and Tungurahua (Bablon et al., 2018a) volcanoes, then

those of Licto cone, Mulmul volcano and Puyo cones in the back-arc (Hoffer et al., 2008), as
MA

well as the onset of Chimborazo activity ~120 ka (Samaniego et al., 2012). The construction

of Mulmul volcano followed the Huisla sector collapse (Fig. 6c) and the large ignimbrite
D

eruption of Chalupas caldera that occurred at ~215 ka, 75 km to the north (Beate et al., 2006;
E
PT

Bablon et al., 2018b). The construction stages of Altar volcano, located south of Tungurahua

volcano (Fig. 1b), are poorly documented, but it started before ~35 ka since its debris
CE

avalanche deposit related to its western flank collapse crops out under the Tungurahua DAD
AC

(Bustillos, 2008) dated at that age (Le Pennec et al., 2013). At least seven major sector

collapses occurred within the last 100 ka. They are evidenced by the debris avalanche deposits

from Chimborazo in the Riobamba basin (Clapperton, 1990; Bernard et al., 2008; Samaniego

et al., 2012), from Carihuairazo in the Ambato basin, and from Altar and Tungurahua in

Chambo and Pastaza valleys (Hall et al., 1999; Bernard et al., 2008; Bustillos, 2008; Bablon et

al., 2018a; Fig. 2). In addition, Mulmul volcano reactivated and produced at least three plinian

eruptions, whose deposits are found in the Ambato basin, interbedded with the avalanche

22
ACCEPTED MANUSCRIPT

deposits of Carihuairazo volcano (Fig. 6f; Ordóñez, 2012). Finally, the volcanic activity of

Tungurahua continued during the Holocene, as well as the construction of Puñalica, Sangay

volcano (Monzier et al., 1999b), the youngest construction stage of Calpi cones (Fig. 6f), and

the emission of the ~15 km-long distal Guano lava flow from Chimborazo volcano.

5.3. Relationship between volcanism and the deep geometry of the slab

PT
Overall, the development of volcanoes from the southern Ecuadorian arc seems to have

migrated southward. Such migration can also be inferred at larger scale. Indeed, the onset of

RI
volcanism in the northern part of the Ecuadorian arc is older than for volcanoes from the

SC
southern part (Fig. 7). Although many edifices remain to be studied in detail, such as Altar

and Sangay volcanoes, we point out that northern edifices began their activity earlier or close
NU
to 1 Ma (Cayambe, Pichincha, Chacana, Pan de Azúcar and Atacazo, for instance; Samaniego
MA

et al., 2005; Opdyke et al., 2006; Hidalgo et al., 2008; Hoffer, 2008; Robin et al., 2010), while

volcanic activity in the southern arc started later and is younger than 0.6 Ma.
D

In general, the volcanic front corresponds to the line formed by the volcanoes closest
E

to the subduction zone and generally parallel to the trench axis. These edifices are typically
PT

located about 105-110 km above the dipping plate (e.g., Tatsumi, 1986; England et al., 2004;
CE

Syracuse and Abers, 2006), where the temperature and pressure conditions allow melting of

the mantle wedge and magma genesis. In Ecuador, the depth of the slab below the volcanic
AC

front varies between 85 and 100 km for the northern part of the arc, and between 100 and

130 km for southern termination (Fig. 8). Moreover, the distribution of volcanoes along the

Ecuadorian arc is not uniform. At least four volcanic clusters stand out, separated by along-arc

segments without Quaternary volcanic activity (Fig. 7). For these segments, the spacing

between volcanic centers is greater than the volcano-spacing average of ~25 km (Fig. 7). To

the north, the Colombian arc consists of a single alignment of volcanoes (Hall and Wood,

23
ACCEPTED MANUSCRIPT

1985), which extends into northern Ecuador, forming a first volcanic group of rather small-

size edifices, less than 10 km wide (Fig. 7). Further south, separated from the first cluster by a

20-25 km interval without Quaternary volcanic activity, the second group extends from

Chachimbiro to Iliniza volcanoes (Fig. 7) and includes the greatest number of volcanoes and

back-arc volcanism. The width of the arc increases significantly between 0.5°N and 1°S and

reaches 140 km wide, in front of the subducting Carnegie ridge (Fig. 1a). Southward, where

PT
the volcanic activity is younger, the arc becomes narrower again. The third group of volcanoes

RI
is mainly represented by Pilisurco-Sagoatoa complex, Chinibano and Quilotoa volcanoes

SC
(Fig. 1b and 7). They are separated from the previous group by about 35 km, where only

Quilotoa volcano, the westernmost edifice, is constructed. Finally, south of a 10 km wide


NU
interval without Quaternary edifices, the fourth group consists of the southern termination of

the Ecuadorian arc, and ends with a single edifice, Sangay volcano. The northeastern Japan
MA

arc presents similar volcanic clusters, interpreted as a result of thermal anomalies within the

mantle wedge, which increase melting rate and magma production (Tamura et al., 2002).
D

Moreover, the geographic orientation of the Ecuadorian Quaternary arc changes significantly
E
PT

from the northern to the younger southern part, although it seems to have been rather straight

during the Pliocene (yellow areas west of Chimborazo volcano; Fig. 7). More specifically, the
CE

line formed by the volcanic front in Colombia has a NE trending orientation, which changes to
AC

a NNE trend in northern and central Ecuador, then to a NW trend in the southern part of the

arc, south of 1°S. A recent seismic study focused on the geometry of the slab below Ecuador

has shown that the subduction of the Grijalva fracture zone between Farallón and Nazca plates

(Fig. 1a), as well as the change of convergence obliquity resulting from the convex shape of

the continental margin, produce a flexure of the slab at depth (Yepes et al., 2016; Fig. 8). We

observe that the young volcanoes from the southern termination of the arc investigated here

are located just above this flexure. By contrast, volcanoes with an older activity are located

24
ACCEPTED MANUSCRIPT

above an area where the dip of the slab is shallower (Yepes et al., 2016; Fig. 8), whereas only

two Quaternary edifices (Sangay and Altar volcanoes) are present south of the Grijalva

fracture zone prolongation. We infer that the regional volcanism and the development of the

arc are related to the slab geometry at depth. Consequently, the curvature of the volcanic front

could have been initiated by the change in the slope of the slab induced by the flexure area,

while the recent southward migration of volcanic activity highlighted here may have been

PT
partly triggered by a displacement of both the flexure and the Grijalva fracture zone at depth

RI
below the arc in the last 600 ka. Indeed, the fast and oblique convergence of the Nazca plate

SC
may induce a translation of the Grijalva fracture zone, and then, a displacement of the slab

flexure southward through time. The resulting changes of pressure and temperature conditions
NU
may favor mantle partial melting, and hence magma genesis. Such relationships between the

slab geometry and the volcanism migration have been previously described in the Central
MA

Andes (e.g., Ramos and Folguera, 2005; 2009). Assuming that the average velocity and

motion direction of the Nazca plate during the last million years is similar to the current GPS
D

data (~6 cm.a-1 and ~N81°E relative to South America; Trenkamp et al., 2002; Kendrick et al.,
E
PT

2003; Bird, 2003; DeMets et al., 2010; Nocquet et al., 2014), that the angle between the

Grijalva fracture zone and the motion direction of the Nazca plate is 10 ± 1°, that
CE

the angle between the Grijalva fracture zone and the margin is 45 ± 5° (Fig. 8),
AC

and using the same approach as used by Hampel (2002) to estimate the displacement of the

Nazca ridge in Peru, we propose that the southward translation of the Grijalva fracture zone

along the trench occurred at = 15 ± 1 km.Ma-1 . Such value is in

agreement with the paleogeographic reconstructions of the Nazca plate based on marine

magnetic anomalies and the inferred past location of the Grijalva fracture zone (Hey, 1977;

Meschede and Barckhausen, 2001; Collot et al., 2009). However, this velocity seems too low

25
ACCEPTED MANUSCRIPT

to account alone for the southward volcanism migration since 600 ka, and a slab deformation

without a significant displacement of the Grijalva fracture zone may have occurred to explain

the distribution of volcanoes along the southern Ecuadorian arc.

5.4. Relationship between volcanism and crustal fault activity

The intense volcanic activity since ~500 ka and the numerous sector collapse events

PT
that seem to occur in the last 100 ka in the southern termination of the arc may also be

interpreted in light of the tectonic activity occurring along the Chingual-Cosanga-Pallatanga-

RI
Puná (CCPP; Alvarado et al., 2016; Fig. 1a) fault system, and especially along the Pallatanga

SC
fault segment. Several edifices, such as Calpi cones, Igualata, Huisla, and Mulmul volcanoes,
NU
line up with the fault (Fig. 2b; Fig. 6). Their location and morphology suggest a close

relationship between the tectonic activity of the Pallatanga fault and the location and evolution
MA

of volcanoes within the Interandean Valley. The volcanic edifice construction that appears to

be most affected by the interaction with the tectonic activity is Igualata volcano. The
D

construction of its main structure occurred approximately between 380 and 350 ka (Table 1).
E

Then, after an apparent period of quiescence of ~100 ka, it experienced at least two periods of
PT

activity, at 237 ± 9 ka with the andesitic lava flow, and at 107 ± 11 ka with the three dacitic
CE

pyroclastic flows unit. Both units were emitted along the Pallatanga fault (Baize et al., 2016).

Moreover, their emplacement within the Patalú valley (Fig. 2c) incised by the recurring
AC

seismic activity of the Pallatanga segment strongly suggests that the fault had already an

impact on the morphology at ~250 ka, and therefore that it was already active at that time. In

the southern foothill of Igualata volcano, the distal Guano lava flow is sheared by the

Pallatanga fault (Fig. 4a). Taking into account its age of 4 ± 2 ka (Table 1) and a lateral offset

of at least 20 m (Baize et al., 2016), a displacement velocity of 3.3 to 10 mm.a-1 can be

calculated. This result is in agreement with the Holocene slip rate obtained for the Pallatanga

fault (Winter et al., 1993) and GPS data for the bulk CCPP fault velocity (Nocquet et al.,

26
ACCEPTED MANUSCRIPT

2014). Further south, Calpi cones are quite isolated from other volcanoes and sit along the

fault system, southwest of the Riobamba basin (Fig. 6). Characterized by a relatively primitive

composition (Monzier et al., 1999a; Ancellin et al., 2017; Fig. 3a), their eruption could have

been favored by fracture openings along the fault possibly following large earthquakes.

Similarly, Huisla and Mulmul volcanoes are constructed at the intersection between the

northeastern known tip of the Pallatanga fault and a N-S suture zone located along the Eastern

PT
Cordillera (Alvarado et al., 2016). These fracture zones may also have facilitated the rise of

RI
magma. Moreover, as evidenced at Mount St Helens (e.g., Endo et al., 1981; Tilling et al.,

SC
1990; Lagmay et al., 2000) and in Ecuador (Andrade et al, 2018), fault activity may also favor

sector collapses. It seems to have been the case for Huisla volcano, since the Pallatanga fault
NU
is roughly aligned with the orientation of its flank collapse amphitheater (Fig. 5b) that

occurred between 215 and 180 ka (Table 1). In addition, the morphology of Mulmul volcano
MA

shows that its southeastern side is more eroded than Huisla volcano, even though Mulmul is

much younger (Table 1). The dismantling of both edifices could have been promoted by a
D

recent activity along these segments of the Pallatanga fault (Fig. 5b), triggering local
E

collapses, thereby enhancing the effects of erosion and river incision. The gradual erosion of
PT

the banks of the deep Río Chambo valley could also have played a role in the dismantling of
CE

Mulmul’s eastern flank.

Finally, the morphological comparison of the two largest edifices studied here,
AC

Sagoatoa and Igualata volcanoes, highlights the impact of the tectonic activity on erosion of

volcanoes. Both volcanoes are located in the Interandean Valley. Sagoatoa volcano seems

rather unaffected by tectonics, although no survey focused on the western edge of the

Interandean Valley was performed, except south of Chimborazo (Baize et al., 2015). On the

contrary, Igualata is cross-cut by the Pallatanga fault (Baize et al., 2016; Fig. 6), forming a

sigmoidal graben-like structure in its central area. Such structure is characteristic of a locally

27
ACCEPTED MANUSCRIPT

extensive deformation, typical of interactions between strike-slip faults and volcanic edifices

(Mathieu and van Wyk de Vries, 2001; Lagmay et al., 2000; Andrade et al, 2018). The

different tectonic settings of Sagoatoa and Igualata volcanoes may thus influence their

different erosional processes. A simple way to compare the degree of erosion affecting each

volcano is to calculate the circularity index (Grosse et al., 2009; 2012) of a contour line

located mid-height of the volcano. This index is defined by , with A being the area

PT
contained within the contour line and P its perimeter. C tends towards 1 for a quasi-circular

RI
line, which would characterize non-eroded flanks of a young edifice, while it gets closer to 0

SC
for a highly eroded volcano. Circularity index of Sagoatoa volcano is ~0.34 for the contour

line at 3480 m (unaffected by the presence of Pilisurco to the west), while the circularity index
NU
for Igualata is ~0.19 at 3770 m. Igualata volcano is thus much more eroded than Sagoatoa,

despite the fact that it is only half the age. The presence of faults facilitated the erosion of
MA

volcanic edifices, more specifically by destroying the edifice during earthquakes, but also by

driving water circulation (e.g., Traineau et al., 1989). The striking morphologic difference
D

between Sagoatoa and Igualata volcanoes can thus be related to the impact of the shear of the
E
PT

Pallatanga fault on Igualata volcano, which favored erosion and led to the formation of the

wide valleys observed in the present-day landscape. However, they could have experienced
CE

different climatic conditions that would increase the apparent tectonic impact on erosion, as
AC

Igualata volcano has a summit elevation that currently reaches 4430 m and could have been

capped by a glacier during the Younger Dryas and/or Neoglacial periods (Clapperton, 1990;

Samaniego et al., 2012), contrary to the Sagoatoa summit, which culminates at 4169 m and

may not have experienced any glacial erosion since the LGM.

To summarize, our new ages suggest that the Pallatanga fault is active since at least

~350 ka, i.e. at the end of Igualata main edifice construction (Fig. 6c). It shows an ongoing

activity, as evidenced by Huisla sector collapse that occurred at 215-180 ka, by Chimborazo

28
ACCEPTED MANUSCRIPT

collapse deposits that are sheared and present fault related pressure ridges (Baize et al., 2016)

since at least ~62 ka (Fig. 6d), and by the distal Guano lava flow, displaced by the fault during

the last 6 ka (Fig. 6d).

5.5. Geodynamical evolution of the Ecuadorian range in the last million years

The new ages provided in this study (Table 1) highlight the southward migration of the

PT
volcanic activity of the Ecuadorian arc since 600 ka (Fig. 6 and 7). It correlates with (1) a

change of the position and orientation of the volcanic front (Fig. 7), (2) the geometry of the

RI
slab at depth and the presence of a flexure zone (Fig. 8), and (3) the activity of the Pallatanga

SC
fault segment of the CCPP fault system (Fig. 6 and 8). We interpret these interactions by

proposing the following geodynamic scenario.


NU
Since the beginning of the subduction of the Carnegie Ridge, which could have
MA

occurred between 15 and 1 Ma (e.g., Lonsdale and Klitgord, 1978; Spikings et al., 2001),

coupling between the Nazca plate and South American continent increased (Spikings et al.,
D

2001; Espurt et al., 2008; Nocquet et al., 2014). This strain accumulation intensifies at 3 to
E

6 Ma, contributing to the uplift of the Ecuadorian Andes and forearc basins (Spikings et al.,
PT

2001; Pedoja et al., 2006; Alvarado et al., 2016). The oblique convergence of the Nazca plate
CE

led to re-activation of N-S suture zones segments (Alvarado et al., 2016), progressively

reorganized and connected with the more recent Pallatanga and Chingual oblique fault
AC

systems. While the dip of the subducting plate is too flat in Peru and Southern Ecuador to

trigger partial melting of the mantle wedge (e.g., Gutscher et al., 1999), pressure and

temperature conditions favored the magmatic production in the northern part of Ecuador, and

the Quaternary volcanic arc developed in front of the Carnegie ridge (Fig. 7). The convex

geometry of the margin, the presence of the over thickened Carnegie ridge, together with the

stress increase related to its arrival in the trench area, lead to the formation of a major slab

29
ACCEPTED MANUSCRIPT

flexure at depth (Yepes et al., 2016), driven by the Grijalva fracture zone subduction (Fig. 8).

Additionally, the detailed seismic tomography of Ecuador also evidences a curvature of the

Mohorovičić discontinuity between 1 and 2°S (Araujo, 2013). Relative to the South American

continent margin, and due to the oblique convergence, the location of this slab flexure slightly

moved southward as the oceanic plate was subducted. The fast evolution of the slab depth, the

width of the flexure and its slope, may have initialized the migration of the volcanic arc for

PT
the last 600 ka. Moreover, as the arc magmatism occurred above the subducting plate from a

RI
rather constant depth, between 90 and 120 km in Ecuador (Fig. 8), the location and orientation

SC
of the volcanic front changed markedly, from ~27°NE to ~320°NW, following the new

orientation of the slab at depth. Finally, the activity of major faults and the presence of suture
NU
zones (Alvarado et al., 2016) favored magma ascent and participated in the important

development of the arc since 400-300 ka (Fig. 7). In parallel, magmatic dikes and sporadic
MA

intrusion of magma through crustal faults may have also contributed to the development of the

Chingual-Cosanga-Pallatanga-Puná (CCPP) fault system activity, which remains very active


D

today (Winter et al., 1993; Tibaldi et al., 2007; Alvarado et al., 2014; Baize et al., 2015;
E
PT

Alvarado et al., 2016; Champenois et al., 2017).

Along the American Pacific coast, several areas are characterized by the subduction of
CE

an aseismic ridge, such as the Cocos ridge in Central America, the Nazca ridge and the Inca

Plateau in Peru, and the Juan-Fernandez ridge in Chile (e.g., Espurt et al., 2008). However, the
AC

influence of these ridges on geodynamics is variable. Indeed, at these three locations the

geometry of the slab is rather flat due to a higher buoyancy of the oceanic crust (e.g., Pilger,

1984; Gutscher et al., 2000), while the Nazca plate presents a deep flexure below Ecuador.

This can be related to the age, and therefore density difference of the subducting oceanic crust

across the Grijalva fracture zone. In addition, the flat slab of Mexico does not prevent the

development of a Quaternary volcanic arc, contrary to Peru and Chile, where no recent

30
ACCEPTED MANUSCRIPT

volcanism is present (e.g., Manea et al., 2017). Gutscher et al. (2000) suggest that volcanism

can occur during the early stages of a flat subduction, and wane progressively with the lack of

mantle wedge and a prolonged cooling of the subducting lithosphere. Following this

hypothesis, the present geodynamic setting of Ecuador would be at an early stage of the slab

geometry evolution, preceding a long-term shallowing of the Nazca plate and a gradual

decrease of volcanism in the far future (Gutscher et al., 2000; Espurt et al., 2008; Ramos and

PT
Folguera, 2009).

RI
SC
6. Conclusions

The twenty-five new K-Ar ages performed on groundmass for young volcanic activity
NU
characterization presented in this study provide new temporal constraints on the development

of the Ecuadorian arc. Although the periods of activity of some volcanoes remain poorly
MA

documented at the regional scale, such as Sangay, Altar or Pilisurco, edifices from the

southern termination of the arc appear significantly younger than northern volcanoes
D

(Samaniego et al., 2005; Opdyke et al., 2006; Hidalgo et al., 2008; Hoffer, 2008; Robin et al.,
E

2010; Alvarado et al., 2014). An intensification of the volcanic activity seems to have
PT

occurred since 600 ka, with the construction of Huisla, Igualata, Sangay (Monzier et al.,
CE

1999b), Tungurahua (Hall et al., 1999; Le Pennec et al., 2006; Bablon et al., 2018a), then

Carihuairazo, Chimborazo (Samaniego et al., 2012) and Mulmul volcanoes, in the vicinity of
AC

the Pallatanga fault. In addition, the magmatic composition of most recent strombolian cones,

such as those of Calpi, Licto and Puñalica, contrasts with the rest of the arc by showing a

more primitive signature. We attribute the southward migration of volcanism to a recent

reorganization of the geodynamic context of Ecuador, linked to the oblique convergence of

the Nazca plate, and both Grijalva fracture zone and Carnegie Ridge subduction. The coupling

between the subducting plate and the South American continent, intensified with the increase

31
ACCEPTED MANUSCRIPT

of crustal material accumulating into the trench and the oceanic crust age difference across the

Grijalva fracture zone, triggered the formation of a slab flexure at depth (Yepes et al., 2016).

Strains were accommodated at the surface by the activation of the major CCPP fault system.

Moreover, the new ages allow us to estimate that the Pallatanga fault was active since at least

350 ka, with a displacement velocity of about 3.3 to 10 mm.a-1 since the last 6 ka. Our

hypothesis relating volcanic arc development in direct connection with the upper plate and

PT
slab geodynamics in Ecuador seems to favor a future increase of the number of volcanoes to

RI
the south of the current active arc.

SC
Acknowledgments
NU
The authors warmly thank Victor Ramos and an anonymous reviewer for their

constructive comments and detailed reviews that greatly improved this manuscript. We thank
MA

members of Instituto Geofísico (Escuela Politécnica Nacional) of Quito for their support

during field trips, and the French Institut de Recherche pour le Développement (IRD), through
D

the Laboratoire Mixte International program “Séismes et Volcans dans les Andes du Nord”,
E
PT

which made this work possible. We also wish to thank Valérie Godard for her quality

manufacturing of the thin sections, Fanny Soler for her help during preparations and
CE

measurements of Sagoatoa, Calpi and Guano samples, as well as Santiago Santamaría for
AC

useful discussions, Pierre Lahitte for his advices and remarks concerning the impact of erosion

on the morphology, and Sébastien Lénard for his suggestions on the earlier draft of the

manuscript. This work is dedicated to the memory of our colleague Michel Monzier,

volcanologist at IRD, which was the first to describe the unusual characteristic of the southern

termination of the Ecuadorian arc. This publication was financially supported by INSU CNRS

TelluS Aleas (INSU 2015-ALEAS) and LMI IRD (2012-16 LMI-SVAN-IRD) programs. This

32
ACCEPTED MANUSCRIPT

is LGMT contribution number XXX and Laboratory of Excellence ClerVolc contribution

number XXX.

PT
RI
SC
NU
MA
E D
PT
CE
AC

33
ACCEPTED MANUSCRIPT

References

Alvarado, A., Audin, L., Nocquet, J.M., Lagreulet, S., Segovia, M., Font, Y., Lamarque, G., Yepes, H.,
Mothes, P., Rolandone, F., Jarrín, P., Quidelleur, X., 2014. Active tectonics in Quito, Ecuador,
assessed by geomorphological studies, GPS data, and crustal seismicity. Tectonics 33, 67–83.
doi:10.1002/2012TC003224
Alvarado, A., Audin, L., Nocquet, J.M., Jaillard, E., Mothes, P., Jarrín, P., Segovia, M., Rolandone, F.,
Cisneros, D., 2016. Partitioning of oblique convergence in the Northern Andes subduction zone:
Migration history and the present-day boundary of the North Andean Sliver in Ecuador. Tectonics
35, 1048–1065. doi:10.1002/2016TC004117
Ancellin, M.-A., Samaniego, P., Vlastélic, I., Nauret, F., Gannoun, A., Hidalgo, S., 2017. Across-arc

PT
versus along-arc Sr-Nd-Pb isotope variations in the Ecuadorian volcanic arc. Geochem. Geophys.
Geosystems 18, 1163–1188. doi:10.1002/2016GC006679
Andrade, S. D., van Wyk de Vries, B., Robin, C., 2018. Imbabura volcano (Ecuador): The influence of

RI
dipping-substrata on the structural development of composite volcanoes during strike-slip
faulting. J. Volcanol. Geotherm. Res., accepted.
Araujo, S., 2013. The Ecuadorian Moho. La Granja, Revista de Ciencias de la Vida, Universidad

SC
Politécnica Salesiana, Ecuador. 18 (2), 43–47.
Aspden, J.A., McCourt, W.J., Brook, M., 1987. Geometrical control of subduction-related magmatism:
the Mesozoic and Cenozoic plutonic history of Western Colombia. J. Geol. Soc. 144, 893–905.
NU
Bablon, M., Quidelleur, X., Samaniego, P., Le Pennec, J.-L., Lahitte, P., Liorzou, C., Bustillos, J.E.,
Hidalgo, S., 2018a. Eruptive chronology of Tungurahua volcano (Ecuador) revisited based on
new K Ar ages and geomorphological reconstructions. J. Volcanol. Geotherm. Res., 357, 378-
398. doi: 10.1016/j.jvolgeores.2018.05.007
MA

Bablon, M., Quidelleur, X., Samaniego, P., Le Pennec, S., Hidalgo, 2018b. New unspiked K-Ar ages
of Holocene lava flows and pumices from the Ecuadorian arc. Cities on Volcanoes 10, Naples,
Italy, pp. 637.
Baize, S., Audin, L., Winter, T., Alvarado, A., Pilatasig Moreno, L., Taipe, M., Reyes, P., Kauffmann,
P., Yepes, H., 2015. Paleoseismology and tectonic geomorphology of the Pallatanga fault
D

(Central Ecuador), a major structure of the South-American crust. Geomorphology 237, 14–28.
doi:10.1016/j.geomorph.2014.02.030
E

Baize, S., Audin, L., Alvarado, A., Champenois, J., 2016. Earthquake fault segmentation in the Central
PT

Andes, Ecuador. Extended Abstract, 7th International INQUA Meeting on Paleoseismology,


Active Tectonics and Archaeoseismology, Crestone, Colorado, USA.
Barberi, F., Coltelli, M., Ferrara, G., Innocenti, F., Navarro, J.M., Santacroce, R., 1988. Plio-
Quaternary Volcanism in Ecuador. Geochem. Mag. 125, 1–14.
CE

Barragán, R., Geist, D., Hall, M., Larson, P., Kurz, M., 1998. Subduction controls on the compositions
of lavas from the Ecuadorian Andes. Earth Planet. Sci. Lett. 154, 153–166.
doi:10.1016/S0012821X(97)00141-6
AC

Barragán, R., Baby, P., Duncan, R., 2005. Cretaceous alkaline intra-plate magmatism in the
Ecuadorian Oriente Basin: Geochemical, geochronological and tectonic evidence. Earth Planet.
Sci. Lett. 236, 670–690. doi:10.1016/j.epsl.2005.03.016
Barba, D., Robin, C., Samaniego, P., Eissen, J.-P., 2008. Holocene recurrent explosive activity at
Chimborazo volcano (Ecuador). J. Volcanol. Geotherm. Res. 176, 27–35.
doi:10.1016/j.jvolgeores.2008.05.004
Beate, B., Hammersley, L., DePaolo, D., Deino, A.I., 2006. La Edad de la Ignimbrita de Chalupas,
Prov. De Cotopaxi, Ecuador, y su importancia como marcador estratigráfico. 6th Jornadas en
Ciencias de la Tierra, EPN, Quito, pp. 68–71.
Béguelin, P., Chiaradia, M., Beate, B., Spikings, R., 2015. The Yanaurcu volcano (Western Cordillera,
Ecuador): A field, petrographic, geochemical, isotopic and geochronological study. Lithos 218–
219, 37–53. doi:10.1016/j.lithos.2015.01.014
Bernard, B., van Wyk de Vries, B., Barba, D., Leyrit, H., Robin, C., Alcaraz, S., Samaniego, P., 2008.
The Chimborazo sector collapse and debris avalanche: Deposit characteristics as evidence of

34
ACCEPTED MANUSCRIPT

emplacement mechanisms. J. Volcanol. Geotherm. Res. 176, 36–43.


doi:10.1016/j.jvolgeores.2008.03.012
Bernard, B., Andrade, D., 2011. Volcanes Cuaternarios del Ecuador Continental. Map 1:500000.
Bernard, B., Hidalgo, S., Robin, C., Beate, B., Quijozaca, J., 2014. The 3640–3510 BC rhyodacite
eruption of Chachimbiro compound volcano, Ecuador: a violent directed blast produced by a
satellite dome. Bull. Volcanol. 76. doi:10.1007/s00445-014-0849-z
Bird, P., 2003. An updated digital model of plate boundaries. Geochem. Geophys. Geosyst. 4(3),
doi:10.1029/2001GC000252
Bryant, J.A., Yogodzinski, G.M., Hall, M.L., Lewicki, J.L., Bailey, D.G., 2006. Geochemical
Constraints on the Origin of Volcanic Rocks from the Andean Northern Volcanic Zone, Ecuador.
J. Petrol. 47, 1147–1175. doi:10.1093/petrology/egl006
Bustillos, J.E., 2008. Las avalanchas de escombros en el sector del volcán Tungurahua. Facultad de

PT
ingeniería en geología y petróleos, Escuela Politécnica Nacional, Quito.
Cassignol, C., Gillot, P.-Y., 1982. Range and effectiveness of unspiked potassium-argon dating:
experimental groundwork and application. Odin GS Ed Numer. Dating Stratigr. John Wiley &

RI
Sons, New York, 159–179.
Castillo, P.R., Janney, P.E., Solidum, R.U., 1999. Petrology and geochemistry of Camiguin Island,

SC
southern Philippines: insights to the source of adakites and other lavas in a complex arc setting.
Contrib. Mineral. Petrol. 134, 33–51.
Champenois, J., Baize, S., Vallée, M., Jomard, H., Alvarado, A., Espin, P., Ekstrom, G., Audin, L.,
2017. Evidences of surface rupture associated with a low magnitude (Mw5.0) shallow earthquake
NU
in the Ecuadorian Andes: Andean earthquake surface rupture. J. Geophys. Res. Solid Earth.
doi:10.1002/2017JB013928
Clapperton, C.M., 1990. Glacial and volcanic geomorphology of the Chimborazo-Carihuairazo Massif,
MA

Ecuadorian Andes. Trans. R. Soc. Edinburgh Earth Sci. 81, 91–116.


Collot, J.-Y., Michaud, F., Alvarado, A., Marcaillou, B., Sosson, M., Ratzov, G., Migeon, S.,
Calahorrano, A., Pazmino, A., 2009. Visión general de la morfología submarina del margen
convergente de Ecuador-Sur de Colombia: implicaciones sobre la transferencia de masa y la edad
D

de la subducción de la Cordillera de Carnegie. Comisión Nacional del Derecho del Mar (CNDM).
Primera Edición. 47–74.
E

Coltorti, M., Ollier, C.D., 2000. Geomorphic and tectonic evolution of the Ecuadorian Andes.
Geomorphology 32, 1–19.
PT

Cotten, J., Le Dez, A., Bau, M., Caroff, M., Maury, R.C., Dulski, P., Fourcade, S., Bohn, M., Brousse,
R., 1995. Origin of anomalous rare-earth element and yttrium enrichments in subaerially exposed
basalts: Evidence from French Polynesia. Chem. Geol. 119, 115–138.
CE

DeMets, C., Gordon, R.G., Argus, D.F., 2010. Geologically current plate motions. Geophys. J. Int.
181, 1–80. doi:10.1111/j.1365-246X.2009.04491.x
Dumont, J.F., Santana, E., Vilema, W., Pedoja, K., Ordóñez, M., Cruz, M., Jiménez, N., Zambrano, I.,
AC

2005. Morphological and microtectonic analysis of Quaternary deformation from Puná and Santa
Clara Islands, Gulf of Guayaquil, Ecuador (South America). Tectonophysics 399, 331–350.
doi:10.1016/j.tecto.2004.12.029
Egbue, O., Kellogg, J., 2010. Pleistocene to Present North Andean “escape.” Tectonophysics 489, 248–
257. doi:10.1016/j.tecto.2010.04.021
Ego, F., Sébrier, M., Lavenu, A., Yepes, H., Egues, A., 1996. Quaternary state of stress in the Northern
Andes and the restraining bend model for the Ecuadorian Andes. Tectonophysics 259, 101–116.
Endo, E.T., Malone, S.D., Noson, L.L., Weaver, C.S., 1981. Locations, magnitudes and statistics of the
March 20-May 18 earthquake sequence, in The 1980 eruptions of Mount St. Helens, Washington,
U. S. Geol. Surv. Profess. Paper 1250, 93–107.
England, P., Engdahl, R., Thatcher, W., 2004. Systematic variation in the depths of slabs beneath arc
volcanoes. Geophys. J. Int. 156, 377–408. doi:10.1111/j.1365-246X.2003.02132.x
Espín, P.A., 2014. Caracterización geológica y litológica de los depósitos laháricos de Mera, provincia
de Pastaza. Facultad de Geología y Petróleos, Escuela Politécnica Nacional, Quito.

35
ACCEPTED MANUSCRIPT

Espín, P.A., Mothes, P.A., Hall, M.L., Valverde, V., Keen, H., 2018. The “Mera” lahar deposit in the
upper Amazon basin: Transformation of a late Pleistocene collapse at Huisla volcano, central
Ecuador. J. Volcanol. Geotherm. Res., accepted.
Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna, C., Brusset, S., 2008. Flat
subduction dynamics and deformation of the South American plate: Insights from analog
modeling. Tectonics 27, TC3011. doi:10.1029/2007TC002175
Germa, A., Quidelleur, X., Gillot, P.Y., Tchilinguirian, P., 2010. Volcanic evolution of the back-arc
Pleistocene Payun Matru volcanic field (Argentina). J. South Am. Earth Sci. 29, 717–730.
doi:10.1016/j.jsames.2010.01.002
Germa, A., Quidelleur, X., Lahitte, P., Labanieh, S., Chauvel, C., 2011. The K–Ar Cassignol–Gillot
technique applied to western Martinique lavas: A record of Lesser Antilles arc activity from 2Ma
to Mount Pelée volcanism. Quat. Geochronol. 6, 341–355. doi:10.1016/j.quageo.2011.02.001

PT
Gillot, P.-Y., Hildenbrand, A., Lefèvre, J.-C., Albore-Livadie, C., 2006. The K/Ar dating method :
principle, analytical techniques, and application to Holocene volcanic eruptions in Southern Italy.
Acta Vulcanol. 18, 55–66.

RI
Graindorge, D., Calahorrano, A., Charvis, P., Collot, J.-Y., Bethoux, N., 2004. Deep structures of the
Ecuador convergent margin and the Carnegie Ridge, possible consequence on great earthquakes

SC
recurrence interval. Geophysical Research Letters 31. doi: 10.1029/2003GL018803
Grosse, P., van Wyk de Vries, B., Petrivonic, I.A., Euillades, P.A., Alvarado, G.E., 2009.
Morphometry and evolution of arc volcanoes. Geology 37, 651–654. doi:10.1130/G25734A.1
Grosse, P., van Wyk de Vries, B., Euillades, P.A., Kervyn, M., Petrinovic, I.A., 2012. Systematic
NU
morphometric characterization of volcanic edifices using digital elevation models.
Geomorphology 136, 114–131. doi:10.1016/j.geomorph.2011.06.001
Guillier, B., Chatelain, J.-L., Jaillard, E., Yepes, H., Poupinet, G., Fels, J.-F., 2001. Seismological
MA

evidence on the geometry of the orogenic system in central-northern Ecuador (South America).
Geophys. Res. Lett. 28, 3749–3752.
Gutscher, M.-A., Malavieille, J., Lallemand, S., Collot, J.-Y., 1999. Tectonic segmentation of the
North Andean margin: impact of the Carnegie Ridge collision. Earth Planet. Sci. Lett. 168, 255–
270.
D

Gutscher, M.-A., Maury, R., Eissen, J.-P., Bourdon, E., 2000. Can slab melting be caused by flat
subduction? Geology 28, 535–538.
E

Hall, M.L., Wood, C.A., 1985. Volcano-tectonic segmentation of the northern Andes. Geology 13,
PT

203–207.
Hall, M.L., Beate, B., 1991. El Volcanism o Plio-Cuaternario en los Andes del Ecuador. El Paisaje
Volcánico de la Sierra Ecuatoriana, Corp. Edit. Nac., Quito, pp. 5–18.
CE

Hall, M.L., Robin, C., Beate, B., Mothes, P., Monzier, M., 1999. Tungurahua Volcano, Ecuador:
structure, eruptive history and hazards. J. Volcanol. Geotherm. Res. 91, 1–21.
Hall, M.L., Samaniego, P., Le Pennec, J.L., Johnson, J.B., 2008. Ecuadorian Andes volcanism: A
review of Late Pliocene to present activity. J. Volcanol. Geotherm. Res. 176, 1–6.
AC

doi:10.1016/j.jvolgeores.2008.06.012
Hall, M.L., Mothes, P.A., 2008. Quilotoa volcano - Ecuador: An overview of young dacitic volcanism
in a lake-filled caldera. J. Volcanol. Geotherm. Res. 176, 44–55.
doi:10.1016/j.jvolgeores.2008.01.025
Hall, M.L., Mothes, P.A., Samaniego, P., Militzer, A., Beate, B., Ramón, P., Robin, C., 2017. Antisana
volcano: A representative andesitic volcano of the eastern cordillera of Ecuador: Petrography,
chemistry, tephra and glacial stratigraphy. J. South Am. Earth Sci. 73, 50–64.
doi:10.1016/j.jsames.2016.11.005
Hampel, A., 2002. The migration history of the Nazca Ridge along the Peruvian active margin: a re-
evaluation. Earth Planet. Sci. Lett. 203, 665–679.
Harford, C.L., Pringle, M.S., Sparks, R.S.J., Young, S.R., 2002. The volcanic evolution of Montserrat
using 40Ar/39Ar geochronology. Geol. Soc. Lond. Mem. 21, 93–113.
doi:10.1144/GSL.MEM.2002.021.01.05

36
ACCEPTED MANUSCRIPT

Hayes, G.P., Wald, D.J., Johnson, R.L., 2012. Slab1.0: A three-dimensional model of global
subduction zone geometries. J. Geophys. Res. Solid Earth 117, B01302. doi
:10.1029/2011JB008524
Hey, R., 1977. Tectonic evolution of the Cocos-Nazca spreading center. Geol. Soc. Am. Bull. 88,
1404–1420.
Hidalgo, S., 2006. Les interactions entre magmas calco-alcalins “classiques” et adakitiques : exemple
du complexe volcanique Atacazo-Ninahuilca (Equateur). Université Blaise Pascal, Clermont-
Ferrand II.
Hidalgo, S., Monzier, M., Almeida, E., Chazot, G., Eissen, J.-P., van der Plicht, J., Hall, M.L., 2008.
Late Pleistocene and Holocene activity of the Atacazo–Ninahuilca Volcanic Complex (Ecuador).
J. Volcanol. Geotherm. Res. 176, 16–26. doi:10.1016/j.jvolgeores.2008.05.017
Hidalgo, S., Gerbe, M.C., Martin, H., Samaniego, P., Bourdon, E., 2012. Role of crustal and slab

PT
components in the Northern Volcanic Zone of the Andes (Ecuador) constrained by Sr–Nd–O
isotopes. Lithos 132–133, 180–192. doi:10.1016/j.lithos.2011.11.019
Hoffer, G., 2008. Fusion partielle d’un manteau métasomatisé par un liquide adakitique: approches

RI
géochimique et expérimentale de la genèse et de l’évolution des magmas de l’arrière-arc
équatorien. Université Blaise Pascal, Clermont-Ferrand II.

SC
Hoffer, G., Eissen, J.-P., Beate, B., Bourdon, E., Fornari, M., Cotten, J., 2008. Geochemical and
petrological constraints on rear-arc magma genesis processes in Ecuador: The Puyo cones and
Mera lavas volcanic formations. J. Volcanol. Geotherm. Res. 176, 107–118.
doi:10.1016/j.jvolgeores.2008.05.023
NU
Hughes, R.A., Pilatasig, L.F., 2002. Cretaceous and Tertiary terrane accretion in the Cordillera
Occidental of the Andes of Ecuador. Tectonophysics 345, 29–48.
Jaillard, E., Bengtson, P., Ordoñez, M., Vaca, W., Dhondt, A., Suárez, J., Toro, J., 2008. Sedimentary
MA

record of terminal Cretaceous accretions in Ecuador: The Yunguilla Group in the Cuenca area. J.
South Am. Earth Sci. 25, 133–144. doi:10.1016/j.jsames.2007.08.002
Jaillard, E., Lapierre, H., Ordonez, M., Alava, J.T., Amortegui, A., Vanmelle, J., 2009. Accreted
oceanic terranes in Ecuador: southern edge of the Caribbean Plate? Geol. Soc. Lond. Spec. Publ.
D

328, 469– 485. doi:10.1144/SP328.19


James, D.E., 1971. Plate tectonic model for the evolution of the Central Andes. Geological Society of
E

America bulletin 82, 3325–3346.


Kellogg, J. N., Vega, V., 1995. Tectonic development of Panama, Costa Rica, and the Colombian
PT

Andes: Constraints from Global Positioning System geodetic studies and gravity. In Mann, P.,
ed., Geologic and Tectonic Development of the Caribbean Plate Boundary in Southern Central
America, Boulder, Colorado, Geological Society of America Special Paper 295.
CE

Kendrick, E., Bevis, M., Smalley, ., Brooks, B., Vargas, .B., Laur a, E., Fortes, L.P.S., 2003. The
Nazca–South America Euler vector and its rate of change. J. South Am. Earth Sci. 16, 125–131.
doi:10.1016/S0895-9811(03)00028-2
AC

Kerr, A.C., Aspden, J.A., Tarney, J., Pilatasig, L.F., 2002. The nature and provenance of accreted
oceanic terranes in western Ecuador: geochemical and tectonic constraints. J. Geol. Soc. 159,
577– 594. doi:10.1144/0016-764901-151
Lagmay, A.M.F., van Wyk de Vries, B., Kerle, N., Pyle, D.M., 2000. Volcano instability induced by
strike-slip faulting. Bull. Volcanol. 62, 331–346. doi:10.1007/s004450000103
Lavenu, A., Noblet, C., Bonhomme, M.G., Egüez, A., Dugas, F., Vivier, G., 1992. New K-Ar age dates
of Neogene and Quaternary volcanic rocks from the Ecuadorian Andes: Implications for the
relationship between sedimentation, volcanism, and tectonics. J. South Am. Earth Sci. 5, 309–
320.
Lavenu, A., Winter, T., Dávila, F., 1995. A Pliocene–Quaternary compressional basin in the
Interandean Depression, Central Ecuador. Geophys. J. Int. 121, 279–300.
doi:10.1111/j.1365246X.1995.tb03527.x
Le Pennec, J.-L., Hall, M.L., Robin, C., Bartomioli, E., 2006. Tungurahua volcano, Late Holocene
Activity. Fourth Conference Cities on Volcanoes, IAVCEI, Quito (Ecuador).

37
ACCEPTED MANUSCRIPT

Le Pennec, J.L., Ruiz, A.G., Eissen, J.P., Hall, M.L., Fornari, M., 2011. Identifying potentially active
volcanoes in the Andes: Radiometric evidence for late Pleistocene-early Holocene eruptions at
Volcán Imbabura, Ecuador. J. Volcanol. Geotherm. Res. 206, 121–135.
doi:10.1016/j.jvolgeores.2011.06.002
Le Pennec, J.-L., De Saulieu, G., Samaniego, P., Jaya, D., Gailler, L., 2013. A Devastating Plinian
Eruption at Tungurahua Volcano Reveals Formative Occupation at 1100 cal BC in Central
Ecuador. Radiocarbon 55, 1199–1214.
Litherland, M., Aspden, J.A., Egüez, A., 1993. Mapa geológico de la República del Ecuador
1:1000000. British Geological Survey y CODIGEM
Lonsdale, P., Klitgord, K.D., 1978. Structure and tectonic history of the eastern Panama Basin. Geol.
Soc. Am. Bull. 89, 981–999.
Lonsdale, P., 2005. Creation of the Cocos and Nazca plates by fission of the Farallon plate.

PT
Tectonophysics 404, 237–264. doi:10.1016/j.tecto.2005.05.011
Manea, V.C., Manea, M., Ferrari, L., Orozco-Esquivel, T., Valenzuela, R.W., Husker, A.,
Kostoglodov, V., 2017. A review of the geodynamic evolution of flat slab subduction in Mexico,

RI
Peru, and Chile. Tectonophysics 695, 27–52. doi:10.1016/j.tecto.2016.11.037
Mathieu, L., van Wyk de Vries, B., 2011. The impact of strike-slip, transtensional and transpressional

SC
fault zones on volcanoes. Part 1: Scaled experiments. J. Struct. Geol. 33, 907–917.
doi:10.1016/j.jsg.2011.03.002
Meschede, M., Barckhausen, U., 2001. The relationship of the Cocos and Carnegie ridges: age
constraints from paleogeographic reconstructions. Int. J. Earth Sci. 90, 386–392.
NU
doi:10.1007/s005310000155
Monzier, M., Robin, C., Hall, M.L., Cotten, J., Samaniego, P., 1999a. Geochemistry and tectonics at
the Southern termination of the Northern Volcanic Zone (Riobamba volcanoes, Ecuador):
MA

Preliminary results. Fouth ISAG, Goettingen (Germany), 516-518.


Monzier, M., Robin, C., Samaniego, P., Hall, M.L., Cotten, J., Mothes, P., Arnaud, N., 1999b. Sanguay
volcano, Ecuador: structural development, present activity and petrology. J. Volcanol. Geotherm.
Res. 90, 49–79.
D

Narvaez, D.F., Rose-Koga, E.F., Samaniego, P., Koga, K.T., Hidalgo, S., 2018. Constraining magma
sources using primitive olivine-hosted melt inclusions from Puñalica and Sangay volcanoes
E

(Ecuador). Contributions to Mineralogy and Petrology 173:10, doi:10.1007/s00410-018-1508-8


Nocquet, J.-M., Villegas-Lanza, J.C., Chlieh, M., Mothes, P.A., Rolandone, F., Jarrin, P., Cisneros, D.,
PT

Alvarado, A., Audin, L., Bondoux, F., Martin, X., Font, Y., Régnier, M., Vallée, M., Tran, T.,
Beauval, C., Maguiña Mendoza, J.M., Martinez, W., Tavera, H., Yepes, H., 2014. Motion of
continental slivers and creeping subduction in the northern Andes. Nat. Geosci. 7, 287–291.
CE

doi:10.1038/ngeo2099
Opdyke, N.D., Hall, M., Mejia, V., Huang, K., Foster, D.A., 2006. Time-averaged field at the equator:
Results from Ecuador. Geochem. Geophys. Geosystems 7. doi:10.1029/2005GC001221
AC

Ordóñez, J., 2012. Depósitos volcánicos del Pleistoceno Tardío en la cuenca de. Ambato:
caracterización, distribución y origen. Escuela Politécnica Nacional, Quito.
Peccerillo, A., Taylor, S.R., 1976. Geochemistry of Eocene calc-alkaline volcanic rocks from
the kastamonu area, northern Turkey. Contrib. Mineral. Petrol. 58, 63–81.
Pedoja, K., Dumont, J.F., Lamothe, M., Ortlieb, L., Collot, J.-Y., Ghaleb, B., Auclair, M., Alvarez, V.,
Labrousse, B., 2006. Plio-Quaternary uplift of the Manta Peninsula and La Plata Island and the
subduction of the Carnegie Ridge, central coast of Ecuador. J. South Am. Earth Sci. 22, 1–21.
doi:10.1016/j.jsames.2006.08.003
Pennington, W.D., 1981. Subduction of the Eastern Panama Basin and seismotectonics of northwestern
South America. J. Geophys. Res. 86, 10753–10770.
Pilger, R.H., 1984. Cenozoic plate kinematics, subduction and magmatism: South American Andes. J.
Geol. Soc. 141, 793–802.
Ramos, V., Folguera, A., 2005. Tectonic evolution of the Andes of Neuquén: constraints derived from
the magmatic arc and foreland deformation, in: Veiga, G., Spalletti, L., Howell, J., Schwarz, E.

38
ACCEPTED MANUSCRIPT

(Eds.), The Neuquén Basin: A case study in sequence stratigraphy and basin dynamics. Geol. Soc.
Lond. Spec. Publ. 252, 15–35.
Ramos, V.A., Folguera, A., 2009. Andean flat-slab subduction through time. Geol. Soc. Lond. Spec.
Publ. 327, 31–54. doi: 10.1144/SP327.3
Ricci, J., Quidelleur, X., Lahitte, P., 2015. Volcanic evolution of central Basse-Terre Island revisited
on the basis of new geochronology and geomorphology data. Bull. Volcanol. 77.
doi:10.1007/s00445015-0970-7
Robin, C., Samaniego, P., Le Pennec, J.-L., Fornari, M., Mothes, P., van der Plicht, J., 2010. New
radiometric and petrological constraints on the evolution of the Pichincha volcanic complex
(Ecuador). Bull. Volcanol. 72, 1109–1129. doi:10.1007/s00445-010-0389-0
Sallarès, V., Charvis, P., 2003. Crustal thickness constraints on the geodynamic evolution of the
Galapagos Volcanic Province. Earth Planet. Sci. Lett. 214, 545–559. doi:10.1016/S0012-

PT
821X(03)00373-X
Samaniego, P., Monzier, M., Robin, C., Hall, M.L., 1998. Late Holocene eruptive activity at Nevado
Cayambe Volcano, Ecuador. Bull. Volcanol. 59, 451–459.

RI
Samaniego, P., Martin, H., Monzier, M., Robin, C., Fornari, M., Eissen, J.-P., Cotten, J., 2005.
Temporal Evolution of Magmatism in the Northern Volcanic Zone of the Andes: The Geology

SC
and Petrology of Cayambe Volcanic Complex (Ecuador). J. Petrol. 46, 2225–2252.
doi:10.1093/petrology/egi053
Samaniego, P., Robin, C., Chazot, G., Bourdon, E., Cotten, J., 2010. Evolving metasomatic agent in
the Northern Andean subduction zone, deduced from magma composition of the long-lived
NU
Pichincha volcanic complex (Ecuador). Contrib. Mineral. Petrol. 160, 239–260.
doi:10.1007/s00410-009-0475-5
Samaniego, P., Le Pennec, J.-L., Robin, C., Hidalgo, S., 2011. Petrological analysis of the pre-eruptive
MA

magmatic process prior to the 2006 explosive eruptions at Tungurahua volcano (Ecuador). J.
Volcanol. Geotherm. Res. 199, 69–84. doi:10.1016/j.jvolgeores.2010.10.010
Samaniego, P., Barba, D., Robin, C., Fornari, M., Bernard, B., 2012. Eruptive history of Chimborazo
volcano (Ecuador): A large, ice-capped and hazardous compound volcano in the Northern Andes.
J. Volcanol. Geotherm. Res. 221–222, 33–51. doi:10.1016/j.jvolgeores.2012.01.014
D

Samper, A., Quidelleur, X., Lahitte, P., Mollex, D., 2007. Timing of effusive volcanism and collapse
E

events within an oceanic arc island: Basse-Terre, Guadeloupe archipelago (Lesser Antilles Arc).
Earth Planet. Sci. Lett. 258, 175–191. doi:10.1016/j.epsl.2007.03.030
PT

Samper, A., Quidelleur, X., Boudon, G., Le Friant, A., Komorowski, J.C., 2008. Radiometric dating of
three large volume flank collapses in the Lesser Antilles Arc. J. Volcanol. Geotherm. Res. 176,
485–492. doi: 10.1016/j.jvolgeores.2008.04.018
CE

Samper, A., Quidelleur, X., Komorowski, J.-C., Lahitte, P., Boudon, G., 2009. Effusive history of the
Grande Découverte Volcanic Complex, southern Basse-Terre (Guadeloupe, French West Indies)
from new K–Ar Cassignol–Gillot ages. J. Volcanol. Geotherm. Res. 187, 117–130.
doi:10.1016/j.jvolgeores.2009.08.016
AC

Schiano, P., Monzier, M., Eissen, J.-P., Martin, H., Koga, K.T., 2010. Simple mixing as the major
control of the evolution of volcanic suites in the Ecuadorian Andes. Contrib. Mineral. Petrol. 160,
297– 312. doi:10.1007/s00410-009-0478-2
Smith, W.H.F., Sandwell, D.T., 1997. Global Sea Floor Topography from Satellite Altimetry and Ship
Depth Soundings. Science 277, 1956–1962.
Spikings, R.A., Winkler, W., Seward, D., Handler, R., 2001. Along-strike variations in the thermal and
tectonic response of the continental Ecuadorian Andes to the collision with heterogeneous oceanic
crust. Earth Planet. Sci. Lett. 186, 57–73.
Steiger, R.H., Jäger, E., 1977. Subcommission on geochronology: convention on the use of decay
constants in geo- and cosmochronology. Earth Planet. Sci. Lett. 36, 359–362.
Stübel, A., 2004 (reissue of 1897). Las montañas volcánicas del Ecuador: retratadas y descritas
geológica-topográficamente. Banco Central del Ecuador, Quito. ISBN: 9978-43-567-0

39
ACCEPTED MANUSCRIPT

Sun, S.-s., McDonough, W.F., 1989. Chemical and isotopic systematics of oceanic basalts:
implications for mantle composition and processes. Geol. Soc. Lond. Spec. Publ. 42, 313–345.
doi:10.1144/GSL.SP.1989.042.01.19
Syracuse, E.M., Abers, G.A., 2006. Global compilation of variations in slab depth beneath arc
volcanoes and implications. Geochem. Geophys. Geosystems 7. doi:10.1029/2005GC001045
Tatsumi, Y., 1986. Formation of the volcanic front in subduction zones. Geophys. Res. Lett. 13, 717–
720.
Tamura, Y., Tatsumi, Y., Zhao, D., Kido, Y., Shukuno, H., 2002. Hot fingers in the mantle wedge: new
insights into magma genesis in subduction zones. Earth Planet. Sci. Lett. 197, 105–116.
Taylor, J.R., 1997. Introduction to Error Analysis, the Study of Uncertainties in Physical
Measurements, 2nd Edition, University Science Books.
Tibaldi, A., Rovida, A., Corazzato, C., 2007. Late Quaternary kinematics, slip-rate and segmentation of

PT
a major Cordillera-parallel transcurrent fault: The Cayambe-Afiladores-Sibundoy system, NW
South America. J. Struct. Geol. 29, 664–680. doi:10.1016/j.jsg.2006.11.008
Tilling, R.I., Topinka, L., Swanson, D.A., 1990. Eruptions of Mount St. Helens: Past, Present, and

RI
Future. The Climactic Eruption of May 18, 1980. U.S. Geological Survey Special Interest
Publication.

SC
Traineau, H., Westercamp, D., Benderitter, Y., 1989. Case study of a volcanic geothermal system,
Mount Pelée, Martinique. J. Volcanol. Geotherm. Res. 38, 49–66.
Trenkamp, R., Kellogg, J.N., Freymueller, J.T., Mora, H.P., 2002. Wide plate margin deformation,
southern Central America and northwestern South America, CASA GPS observations. J. South
NU
Am. Earth Sci. 15, 157–171.
Valverde, V., 2014. Las avalanchas de escombros provenientes del volcán Sangay: caracterización
petrográfica-geoquímica. Escuela Politécnica Nacional, Quito.
MA

Yepes, H., Audin, L., Alvarado, A., Beauval, C., Aguilar, J., Font, Y., Cotton, F., 2016. A new view
for the geodynamics of Ecuador: Implication in seismogenic source definition and seismic hazard
assessment. Tectonics 35, 1249–1279. doi:10.1002/2015TC003941
Winter, T., Avouac, J.P., Lavenu, A., 1993. Late Quaternary kinematics of the Pallatanga strike-slip
fault (Central Ecuador) from topographic measurements of displaced morphological features.
D

Geophys. J. Int. 115, 905–920.


Witt, C., Bourgois, J., Michaud, F., Ordoñez, M., Jiménez, N., Sosson, M., 2006. Development of the
E

Gulf of Guayaquil (Ecuador) during the Quaternary as an effect of the North Andean block
PT

tectonic escape. Tectonics 25, TC3017. doi:10.1029/2004TC001723


CE
AC

40
ACCEPTED MANUSCRIPT

Table 1 (Bablon et al.)

40
Longitude Latitude K Ar* 40
Ar* x 1011 Age ± 1σ Mean age
Sample Location
(m) (m) (%) (%) (at/g) (ka) (ka)
Sagoatoa volcano
16EQ40 Lava flow, summit 759889 9872238 2.512 22.10 21.661 825 ± 12 826 ± 12
19.66 21.721 828 ± 12
16EQ03 Distal lava flow, west of 770638 9869872 1.639 24.88 13.710 801 ± 12 799 ± 12
Píllaro city 22.84 13.651 797 ± 12
Carihuairazo volcano
CAR-14 Tzunantza dome, N flank 753880 9851817 1.287 8.46 6.9239 515 ± 9 512 ± 9

PT
10.58 6.8446 509 ± 9
RIO-18 Distal lava flow from Cerro 761709 9848098 1.409 0.59 0.2953 20 ± 3 18 ± 3
Puñalica 0.52 0.2357 16 ± 3

RI
Chimborazo volcano
16EQ43 Proximal deposit of Guano 752451 9833515 2.132 0.90 0.6553 29 ± 3 30 ± 3
lava flow 1.02 0.6993 31 ± 3

SC
16EQ28 Distal deposit of Guano lava 759527 9834465 2.029 0.17 0.1871 9±5 4 ± 2*
flow, north of Riobamba city 0.10 0.0955 5±5
0.09 0.0965 5±5
0.06 0.0598 3±5
NU
0.04 0.0409 2±5
Igualata volcano
16EQ23 Lava flow, summit 762559 9834762 1.783 4.44 6.9751 374 ± 10 376 ± 10
3.99 7.0389 378 ± 11
MA

16EQ14 Distal lava flow, under 776294 9837640 1.967 9.78 7.5893 369 ± 6 371 ± 7
Mulmul 5.62 7.6959 375 ± 9
16EQ27 Distal SE lava flow 770383 9818984 1.455 12.87 5.4347 358 ± 6 358 ± 6
11.17 5.4406 358 ± 6
16EQ24 Lava flow, E flank 765091 9834997 1.320 7.94 4.6095 334 ± 6 337 ± 7
D

6.97 4.6994 341 ± 7


16EQ29 Lava flow, SW flank, Patulù 757733 9827827 2.130 2.98 5.2630 237 ± 9 237 ± 9
E

valley 2.99 5.2781 237 ± 9


16EQ30 Pyroclastic flow deposit, SW 757703 9827809 3.063 1.04 3.4471 108 ± 10 107 ± 11
PT

flank 1.02 3.4291 107 ± 11


Huisla-Mulmul volcanic complex
16EQ07 Lava flow, W flank 769568 9845822 1.532 10.28 9.9303 621 ± 11 612 ± 10
CE

15.30 9.7077 607 ± 9


16EQ04 Lava flow, N flank 771468 9849602 1.838 14.14 11.317 590 ± 9 587 ± 9
14.15 11.205 584 ± 9
RIO-107 Lava flow, NW flank 768948 9846937 1.413 6.29 7.9374 538 ± 11 547 ± 11
AC

6.61 8.2139 557 ± 12


16EQ05 Lava flow, W flank 769181 9845523 1.328 2.91 7.1965 519 ± 19 526 ± 20
2.88 7.3924 533 ± 20
RIO-111 Lava flow, N flank 770536 9848278 2.579 8.97 13.221 491 ± 9 492 ± 9
9.23 13.278 493 ± 9
16EQ08 Lava flow, N flank 773790 9843648 1.933 9.42 3.4885 173 ± 3 174 ± 3
8.14 3.5413 175 ± 3
16EQ09 Lava flow, N flank 773525 9842805 2.154 4.03 3.6820 164 ± 5 163 ± 5
4.34 3.6574 163 ± 4
16EQ10 Lava flow, N flank 773156 9842420 2.073 7.09 3.5358 163 ± 3 163 ± 3
7.81 3.5441 164 ± 3
16EQ11 Lava flow, N flank 772791 9841758 1.484 4.10 2.2092 143 ± 4 145 ± 4
4.15 2.2810 147 ± 4

41
ACCEPTED MANUSCRIPT

40
Longitude Latitude K Ar* 40
Ar* x 1011 Age ± 1σ Mean age
Sample Location
(m) (m) (%) (%) (at/g) (ka) (ka)

Licto and Calpi cones


16EQ47 Lava flow, Licto cone 766283 9802345 1.275 2.24 2.4787 186 ± 9 183 ± 9
(Cerro Tulabug) 2.15 2.4042 181 ± 9
17EQ114 Distal lava flow, 750964 9817521 1.494 1.58 1.0111 65 ± 4 62 ± 4
Calpi, under the DAD 1.50 1.0043 64 ± 4
of Chimborazo 1.30 0.8785 56 ± 4
16EQ34 Lava flow, Calpi, NW 752666 9818239 1.492 0.27 0.2477 16 ± 6 9 ± 3*

PT
cone 0.20 0.1861 12 ± 6
0.13 0.1263 8±6
0.11 0.0953 6±6

RI
0.07 0.0693 4±6
16EQ35 Lava flow, Calpi, SW 752117 9817907 1.417 0.09 0.1561 11 ± 12 8 ± 5*
cone 0.09 0.1536 10 ± 12

SC
0.07 0.1391 9 ± 13
0.04 0.0805 5 ± 13
NU 0.02 0.0385 3 ± 12
MA

Table 1:
D

New groundmass K-Ar ages obtained in this study. Column headings indicate sample name,
E

detail of the outcrop location, sample coordinates projected using the Universal Transverse
PT

Mercator (UTM) coordinate system (Zone 17), potassium concentration in percent, radiogenic

argon content in percent, and in atoms per gram (x1011), ages, and weighted mean age in ka,
CE

with 1-sigma uncertainty. Ages marked with an asterisk (*) were calculated using reciprocal
AC

variances weighted average (see text).

42
ACCEPTED MANUSCRIPT

Table 2 (Bablon et al.)

Sagoatoa Igualata
16EQ03 16EQ40 16EQ41* 16EQ14 16EQ22* 16EQ23 16EQ24 16EQ27 16EQ29 16EQ30
wt.%
SiO2 57.54 63.98 58.13 63.09 53.02 56.32 54.40 61.23 57.81 65.87
TiO2 0.84 0.90 0.81 0.59 1.09 0.96 1.04 0.62 0.93 0.54
Al2O3 15.56 15.32 16.87 16.94 18.32 18.19 16.59 17.41 16.93 16.29
Fe2O3 7.49 6.04 7.28 5.17 8.88 7.38 8.99 6.03 7.33 4.36
CaO 6.99 4.59 6.88 5.29 8.16 7.01 7.96 5.81 6.74 4.20

PT
MgO 6.05 2.55 4.66 2.35 5.03 4.00 5.57 2.35 4.18 1.80
MnO 0.11 0.08 0.10 0.09 0.12 0.09 0.12 0.10 0.11 0.06
K2O 1.25 2.49 1.19 1.77 1.22 1.66 1.32 1.69 1.79 2.38

RI
Na2O 3.95 3.76 3.89 4.51 3.86 4.10 3.71 4.49 3.89 4.32
P2O5 0.23 0.29 0.20 0.19 0.30 0.29 0.30 0.28 0.29 0.18

SC
L.O.I. 0.06 0.80 1.84 0.13 -0.17 1.22 -0.04 0.20 0.59 1.33
ppm
Sc 18.30 11.88 14.90 9.47 15.80 13.87 16.34 8.52 15.50 6.93
NU
V 187.92 158.16 160.56 106.22 190.51 176.31 210.68 118.90 177.04 90.21
Cr 256.62 49.20 123.16 24.41 124.88 57.29 163.31 21.52 100.32 17.94
Co 29.01 17.22 23.56 12.90 27.64 23.10 29.98 13.73 23.62 11.06
Ni 144.22 47.49 66.01 16.71 58.54 46.54 95.46 17.03 58.89 13.52
MA

Rb 25.89 57.70 28.00 43.91 27.20 43.53 21.54 37.16 45.10 71.36
Sr 536.70 376.50 547.10 616.91 815.67 730.12 926.55 948.97 740.76 534.82
Y 16.34 23.83 13.06 11.80 15.31 15.95 14.68 11.64 15.58 11.17
Zr 129.18 246.62 32.41 128.86 121.67 140.26 118.60 116.87 146.06 145.29
Nb 4.45 9.62 4.90 5.23 7.44 6.40 4.92 4.57 6.64 4.65
D

Ba 475.42 809.53 574.08 853.96 625.38 719.38 620.08 901.10 793.83 949.89
La 14.06 24.00 14.14 18.77 20.14 20.66 23.52 21.90 20.89 19.33
E

Ce 31.14 50.40 24.82 37.27 38.82 42.17 47.06 43.74 44.74 41.90
PT

Nd 17.45 27.33 15.55 17.19 24.40 22.99 26.23 21.61 23.34 18.71
Sm 3.62 5.74 2.98 3.40 4.86 4.36 4.75 4.10 4.44 3.66
Eu 1.12 1.36 0.97 0.92 1.37 1.21 1.34 1.19 1.21 0.81
Gd 3.62 5.01 2.77 2.77 4.37 3.92 4.29 3.15 3.84 2.93
CE

Dy 2.86 4.05 2.31 2.16 3.04 2.95 2.72 2.18 2.78 2.06
Er 1.50 2.04 1.02 0.64 1.22 1.41 1.32 0.76 1.52 0.86
Yb 1.45 1.87 0.87 1.04 1.01 1.28 1.07 1.15 1.26 0.87
AC

Th 2.78 6.55 3.16 4.34 3.49 4.31 4.60 5.20 5.06 6.61

43
ACCEPTED MANUSCRIPT

Huisla-Mulmul complex
RIO 107 RIO 111 16EQ04 16EQ05 16EQ07 16EQ08 16EQ09 16EQ10 16EQ11
wt.%
SiO2 54.02 64.66 55.74 55.52 56.92 60.35 60.72 61.08 52.83
TiO2 0.98 0.58 0.93 0.94 0.97 0.77 0.74 0.77 0.97
Al2O3 18.85 16.58 18.39 17.48 17.52 17.81 17.76 17.37 18.47

PT
Fe2O3 8.38 4.69 7.98 8.27 7.92 6.22 6.40 6.31 9.05
CaO 7.24 4.57 6.80 7.44 6.97 5.72 5.43 5.59 8.20
MgO 4.59 2.10 3.73 4.78 3.89 2.49 2.24 2.47 5.23

RI
MnO 0.13 0.08 0.12 0.13 0.11 0.10 0.10 0.11 0.15
K2O 1.41 2.64 1.69 1.38 1.51 2.12 2.09 2.04 1.37

SC
Na2O 4.08 3.94 4.29 3.81 3.92 4.21 4.26 4.03 3.52
P2O5 0.32 0.17 0.33 0.25 0.27 0.23 0.25 0.22 0.22
L.O.I. 0.56 1.20 0.06 0.31 0.51 0.96 0.01 0.06 -0.18
NU
ppm
Sc 16.46 9.02 12.95 18.91 16.90 11.74 9.48 12.45 25.51
V 209.32 89.23 186.12 178.00 196.57 150.78 123.52 140.61 227.85
Cr 90.20 36.43 69.96 76.48 47.81 5.48 7.10 9.55 42.69
MA

Co 26.39 11.43 26.38 27.27 23.51 15.43 15.14 15.03 28.54


Ni 52.98 19.00 49.02 49.75 30.18 5.36 3.99 7.69 24.18
Rb 52.98 69.84 36.10 27.23 32.42 57.85 63.30 62.37 35.60
Sr 1007.07 541.17 976.12 733.36 726.42 662.15 611.23 542.43 632.07
Y 15.58 13.02 15.36 15.08 16.16 15.10 17.52 19.04 24.73
D

Zr 113.89 168.84 121.00 110.44 112.94 132.16 151.88 145.22 112.77


E

Nb 6.79 7.92 6.38 5.22 5.97 5.90 7.02 5.99 5.16


Ba 874.25 1068.14 913.17 705.75 722.53 927.85 870.18 845.11 675.19
PT

La 20.50 22.76 22.42 15.79 18.98 18.85 20.99 20.10 17.95


Ce 37.59 41.38 43.39 34.24 34.30 39.58 42.29 40.38 35.72
Nd 22.17 18.26 23.04 18.44 20.02 18.40 20.88 20.54 20.60
CE

Sm 4.30 3.31 4.69 4.60 3.87 3.70 4.70 3.75 4.57


Eu 1.27 0.93 1.36 1.19 1.15 1.01 1.16 1.12 1.20
Gd 3.78 2.70 3.82 3.55 3.63 3.40 4.17 3.87 4.52
Dy 2.81 2.22 2.82 2.62 2.78 2.55 3.06 3.12 4.25
AC

Er 1.23 1.05 1.59 1.16 1.35 1.31 1.21 1.75 2.32


Yb 1.25 1.12 1.21 1.25 1.25 1.35 1.52 1.63 2.29
Th 3.69 8.07 4.10 2.92 3.04 5.58 6.90 6.49 4.27

44
ACCEPTED MANUSCRIPT

Carihuairazo Guano lava flow Licto cone Calpi cones


CAR-14 RIO-18 16EQ28 16EQ43 16EQ47 16EQ34 16EQ35 17EQ114
wt.%
SiO2 59.62 54.71 61.73 61.96 53.08 61.07 54.65 60.90
TiO2 0.76 0.76 0.72 0.71 0.97 0.73 0.92 0.73
Al2O3 17.52 16.06 16.39 16.35 15.99 16.02 15.98 16.24
Fe2O3 6.44 8.18 6.00 5.98 8.84 6.15 7.67 6.24
CaO 6.11 9.08 5.66 5.44 8.16 5.74 7.86 5.85
MgO 3.51 5.77 3.21 3.26 7.84 4.23 6.96 4.10
MnO 0.10 0.12 0.09 0.08 0.13 0.08 0.12 0.09
K2O 1.31 1.44 1.79 1.84 1.24 1.58 1.45 1.52

PT
Na2O 4.41 3.61 4.22 4.17 3.50 4.21 4.03 4.14
P2O5 0.22 0.26 0.20 0.20 0.25 0.19 0.36 0.20
L.O.I. 0.12 0.21 0.34 0.16 0.26 0.70 0.12 0.72

RI
ppm
Sc 13.17 26.00 13.11 12.53 20.85 12.79 18.49 12.39

SC
V 146.54 220.00 143.99 139.08 205.34 145.83 197.08 146.83
Cr 53.58 156.00 56.76 64.73 362.66 193.05 283.17 164.81
Co 20.72 28.00 18.07 17.53 35.86 20.79 31.35 23.44
Ni 46.72 41.00 40.57 48.73 172.54 75.66 136.86 74.25
NU
Rb 17.03 20.00 41.88 45.77 19.36 38.16 27.19 34.35
Sr 692.38 1145.00 578.33 557.92 700.83 562.84 930.73 642.81
Y 11.41 16.90 13.54 13.49 14.07 9.84 13.51 12.12
MA

Zr 104.51 102.00 150.12 149.59 98.96 106.90 109.85 110.43


Nb 4.56 4.10 5.16 6.02 5.34 4.03 8.84 5.44
Ba 610.68 945.00 745.61 840.39 595.85 645.29 776.47 696.31
La 13.06 32.00 16.68 17.38 14.68 13.53 20.97 14.31
Ce 26.87 60.00 33.77 36.73 31.43 28.29 43.17 29.28
D

Nd 15.52 30.00 18.42 18.74 17.55 15.16 22.43 15.69


Sm 3.46 6.05 3.64 3.43 3.74 2.88 4.55 3.42
E

Eu 1.02 1.48 0.98 1.08 1.13 0.98 1.22 0.98


PT

Gd 2.87 4.30 3.32 3.32 3.50 2.54 3.67 2.99


Dy 2.06 3.10 2.33 2.34 2.65 1.81 2.44 1.70
Er 0.32 1.70 1.23 1.26 1.50 0.41 1.01 1.10
Yb 0.97 1.64 1.16 1.04 1.06 0.78 1.03 0.77
CE

Th 2.43 7.60 4.17 4.54 2.21 3.30 4.09 3.34


Table 2:
AC

Major and trace element compositions of whole-rock samples. All major element data were

brought down to a total of 100%. Samples marked with an asterisk (*) have not been dated.

45
ACCEPTED MANUSCRIPT

HIGHLIGHTS:

The activity of the southern Quaternary Ecuadorian arc is younger than 800 ka.

A southward migration of the Ecuadorian arc occurred during the last 600 ka.

The major Pallatanga fault is active since at least 600 ka.

The spatial distribution of the volcanoes is related to the slab geometry at depth.

PT
Major active faults in Ecuador could favour volcanic activities and flank collapses.

RI
SC
NU
MA
E D
PT
CE
AC

46
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8

You might also like