You are on page 1of 23

Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research


journal homepage: http://ees.elsevier.com

F
OO
Pyroclastic deposits and eruptive heterogeneity of Volcán Antuco (37°S; Southern
Andes) during the Mid to Late Holocene (<7.2 ka)
Jorge E. Romero a ,⁎ , Victoria Ramírez a , Mohammad Ayaz Alam a , Jorge Bustillos b , Alicia Guevara c ,

PR
Roberto Urrutia d , Alessandro Pisello e , Daniele Morgavi e , Evelyn Criollo c
a
Departamento de Geología, Facultad de Ingeniería, Universidad de Atacama, Avenida Copayapu 485, Copiapó, Chile
b
Carrera de Ingeniería en Geología, Facultad de Ingeniería en Geología, Minas, Petróleos y Ambiental, Universidad Central de Ecuador, Quito, Ecuador
c
Departamento de Metalurgia Extractiva Escuela Politécnica Nacional, Quito, Ecuador
d Centro de Ciencias Ambientales (EULA), Universidad de Concepción, Concepción, Chile
e Department of Physics and Geology, University of Perugia, Piazza dell’ Università, 06123 Perugia, Italy

ARTICLE INFO

Article history:
D ABSTRACT

Volcán Antuco (37°24′ S, 71°22′W; 2979 m asl) is the 13th ranked high threat volcano in Chile with 27 recorded
TE
Received 17 July 2019 historical eruptions, mostly (~96%) with volcanic explosivity indices (VEI) of ~1–2. An older eruptive record has
Received in revised form 18 December 2019
been reconstructed from sections exposed on the western flank and is intimately related to a well-documented
Accepted 18 December 2019
catastrophic sector collapse at ~7.2 cal ka BP. However, very little is known about Antuco's post-collapse erup-
Available online xxx
tive history in other sectors, especially on the eastern flanks where prevalent westerly winds favor optimal east-
ward tephra transport and deposition. Our study reveals a more complete record of activity that has already
Keywords been indicated from previous work with at least 23 tephra-forming explosive eruptions, most of them within the
EC

Active volcanism last c. 7.2 ka, including 4 events that have generated pyroclastic density currents that have widely inundated
Tephrostratigraphy the lower eastern flanks. Tephra from these eruptive events are typically composed of scoria, free crystals and
Strombolian lithics, with occasional pumice. The composition of juvenile fragments varies between basalt and trachyandesite
Southern Andes
(50.2–62.2 wt% SiO2) and show phenocrysts of plagioclase, olivine and pyroxene. Our results show that most
of the eruptions of Antuco (c. 79%) are Strombolian to violent Strombolian. These eruptions have an estimated
longer repose times (c. 200 year) and are likely higher in magnitude than those registered during historical times.
This study also shows that the composition, style and magnitude may change from one eruptive episode to the
RR

next. This eruptive variability seems in complete accord with recent findings from other centers in the Southern
Volcanic Zone exhibiting similar temporal eruptive diversity and ultimately, has significant implications with re-
spect to hazard assessment.
© 2019
CO

1. Introduction duce highly vesicular tephra deposits larger in volume than lava flows
(e.g., Heiken, 1978; Taddeucci et al., 2004; Pioli et al., 2008; Ruth
Mafic (i.e. basaltic to basaltic andesite) volcanism generally have and Calder, 2014; Barsotti et al., 2015). In some cases, short-lived,
Hawaiian or Strombolian eruption style, but can have a combination of high-altitude (~2 km) lava fountains may also produce sustained tephra
or transition in between both end-member styles (Parfitt, 2003) with columns (e.g., Romero et al., 2014; Bonaccorso et al., 2014; An-
typical discharge rates from 103 to 105 kg·s−1 (Simkin and Siebert, dronico et al., 2014; Romero et al., 2018). In a few cases basaltic to
1994). Most Strombolian activity consists of short-lived intermittent ex- basaltic-andesite eruptions are of Sub-plinian and Plinian styles or types
UN

plosions with ejection of pyroclastic rocks (e.g., Chouet et al., 1974; (e.g., Sable et al., 2006; Costantini et al., 2009, 2011; Romero
McGetchin et al., 1974; Blackburn et al., 1976; Heiken, 1978; et al., 2016; Arzilli et al., 2019) with mass discharge rates (MDR)
Francis and Oppenheimer, 2004; Patrick et al., 2007). However, up to 0.7–2.0 × 108 kg·s−1 (Costantini et al., 2009). The most impor-
sustained eruption plumes during paroxysmal events could rise to 10 km tant controlling factor in this eruption heterogeneity is the viscosity of
above the vent being fed by closely spaced individual explosions the basaltic magma (e.g. Pioli et al., 2012; Dominguez et al., 2016;
(Parfitt, 2003), developing the so-called “violent Strombolian” erup- Spina et al., 2019). The explosive behavior of mafic volcanoes can
tions (Walker, 1973), which are responsible to pro be fully substantiated through detailed tephrostratigraphic survey and
analysis of their explosive products.
Antuco is a compound stratovolcano located 90 km east of Los An-
geles city in the central part of the Southern Andes of Chile (Fig.

Corresponding author.
E-mail address: jorge_eduardorm@hotmail.com (J.E. Romero) 1A). It has exclusively erupted basaltic products (50.9–53.7% SiO2 since

https://doi.org/10.1016/j.jvolgeores.2019.106759
0377-0273/© 2019.
2 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

F
OO
PR
D
TE
EC

Fig. 1. Location and geology of Volcán Antuco. A: Location map showing the regional capital (star), main cities (black circles) and active volcanoes (orange triangles), where Volcán
RR

Antuco is highlighted (red triangle). B: Local map of Volcán Antuco, showing the natural protected area (green) of Laguna del Laja National Park, the lake (blue), rivers (blue lines) and the
Q-45 international road (red line). The international border is represented by a black dotted line. The sites (S) and points (P) of field observation are reported. C: Geological map, modified
from Martínez et al. (2018). The map has been updated with the current level of the lake. (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)

the Last Glacial Maxima (Martínez et al., 2018). Antuco lavas are typ- sary to characterize and evaluate the threat to the visitors and
ically richer in Na2O than the average of circum-Pacific basalts (Vergara tourism-related facilities located within Laguna del Laja National Park,
CO

et al., 1985), and are among the most isotopically primitive Holocene as well as to the adjacent critical infrastructure, namely, hydroelectric
lavas in the Southern Volcanic Zone (Hildreth and Moorbath, 1988; power stations and an international route to Argentina.
Lohmar, 2000; Hickey-Vargas et al., 2016). At least 27 historical In this study, we present the results of a Late Holocene (i.e. the last
eruptions are reported for the period 1624–1939, mostly VEI 1–2 with c. 7 kyr) tephrostratigraphic survey carried out between March 2015
Strombolian style, which produced basaltic lava flows and restricted sco- and November 2018. Both tephra fall and pyroclastic density currents
riaceous tephra fallout (Petit-Breuilh, 1994; Lohmar, 2000) with low have been studied in terms of their stratigraphy, grain size, componen-
potential of preservation. The latest major eruptive episode at Volcán try, juvenile morphology, petrology and geochemistry, in order to com-
UN

Antuco occurred during 1852–1853 (Smith, 1915) and was character- prehend the style and magnitude of their parental eruptions as well as
ized by combined central-vent and parasitic cone-forming activity result- associated volcanic hazards.
ing in the emission of lava flows that inundated the western flanks and
the eastward distribution of predominantly black-colored tephra fall- 2. Volcán Antuco
out (Fig. 1B). This eruption was later characterized as VEI 3 (Smith,
1915). Despite high-eruptive frequency during historical times (15th Located in the Biobio Region of southern Chile (37°24′S-71°22′W;
to 19th Centuries), its current dormancy for nearly a century and its 2979 m above the sea level), Volcán Antuco is a composite stratovol-
13th position in the ranking of high risk volcanoes in Chile (www. cano which has been active during the last 150.4 ka, and has an esti-
sernageomin.cl), very few tephrostratigraphic studies have been previ- mated volume to 62 km3 (Martínez et al., 2018). The main cone is
ously conducted for this volcano (e.g., Lohmar et al., 1999). The eval- 1 km high, with a volume of 4 km3, and is located in the center of a sec-
uation of the explosive behavior of Volcán Antuco is therefore neces tor collapse scar triggered by a phreatic eruption at an ancestral edifice
(Moreno et al., 2000), which collapsed at 6.2 ka (~7.2 cal ka BP,
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 3

Lohmar, 2000). The evolution of Antuco has been divided into three in these points in order to elucidate their origin, and samples with c.
stages by Martínez et al. (2018) as follows: (1) Early Antuco (Ea), 200 g of bulk material were collected for lithology and grain-size analy-
relates to initial edifice construction at c. 150 ka and continued spo- ses. Sampling sites were selected in order to minimize the potential ef-
radically up to the end of the Last Glacial Maxima at c. 17 ka; (2) fects of water and/or wind erosion, and redeposition. Most described
Late Antuco edifice (La), comprising the volcanic history between 17 sites are outcrops with natural exposures, and in some cases they consti-
and c. 7.2 cal ka BP and (3) a post-collapse stage (PCa), during the last tute road-cuts along Q-45 international road, connecting Chile and Ar-

F
<7.2 ka. gentina (Fig. 2). A lacustrine sediment core from Urrutia et al. (2010)
During the Ea stage, basaltic andesite to dacite lava flows were ef- has been also used to enhance tephra correlation and chronology be-
fused (50.9 to 64.5 wt% of SiO2), in some cases in a subglacial environ- tween proximal to distal correlations.

OO
ment (Martínez et al., 2018). Lava flow deposits are typically interca-
lated with volcanic breccias and form continuous successions, which are 3.2. Dating
then covered by younger lavas (Thiele et al., 1998; López-Escobar et
al., 1981; Moreno et al., 2000; Martínez et al., 2018). More primi- Three radiocarbon ages were acquired as part of this study and in-
tive compositions than those observed in Ea are reported the lavas from clude a horizon of leaves and wood in section S27 (M3734), and two
La stage (<55 wt% of SiO2; Martínez et al., 2018). The ancestral An- organic paleosols in section S17 (S17-A and S17-B) (Table 1). In order
tuco edifice is estimated to have reached a height of 2 km with basal di- to better constrain our tephrochronology, we have utilized five already
ameter of 12 km prior to its westward sector collapse, which generated published radiocarbon dates - three of which come from a sediment core

PR
a voluminous (5 km3) debris avalanche deposit (DAD) (Moreno et al., retrieved from nearby Lago del Laja (Urrutia et al., 2010), and the re-
1984; Thiele et al., 1998; Lohmar, 2000). This DAD extends for at maining two dates are from charcoal within a debris avalanche deposit
least 20 km westward, confined primarily within the Laja River Valley, exposed on the southwest shore of the Laguna del Laja (Lohmar, 2000)
and includes proximal avalanche facies comprised of toreva mega-blocks .
(i.e., rotated and slid parts of the collapsed edifice such as Cerro Con- For the radiocarbon dating, at least 100 g of organic material was
dor), as well as medial-to-distal avalanche facies comprising hummocks initially sampled with no evident adhering modern roots to avoid con-
and intervening diamicts (Thiele et al., 1998). tamination. Samples were then dried at low temperature with sample

D
Following the sector collapse, a series of pyroclastic density cur- splits sent to Direct AMS laboratories, Bothell, Washington, USA for Ac-
rents (PDC), composed of basaltic fine to coarse ash and angular to celerator Mass Spectrometry (AMS) dating. All results have been cor-
subrounded lithics, covered a surface area of about 300 km2, with the rected for isotopic fractionation with an unreported δ13C value mea-
thicknesses varying between 13 and 40 m and an estimated total vol- sured on the prepared carbon by the accelerator. All the reported ages,
TE
ume >5 km3, producing the so-called “Arenas Negras de Trupán-Laja” including those of Lohmar (2000) and Urrutia et al. (2010) were
deposit (Thiele et al., 1998; Moreno et al., 2000, 2018). These calibrated using CALIB rev. 7.0.4 (Stuiver and Reimer, 1993), using
PDC deposits have been interpreted as produced by phreatomagmatic the SHCal13 calibration curve for the Southern Hemisphere (Hogg et
wet base surges (Moreno et al., 2000). Two unusually thick and far al., 2013). Ages are given as calibrated ranges of years before present
reaching basaltic andesite lava flows were erupted soon after the depo- (y BP, present considered to be 1950) according to 2σ confidence level
sition of these PDCs (Watt, 2019). The Laja River was dammed with (95.4%), while ages in literature are expressed also as Anna Domini (y
EC

the debris that catastrophically collapsed producing a voluminous flood AD).


event, which formed a prominent 50 km3 alluvial fan in the central
valley, which was composed of “black sands” (Vergara and Katsui, 3.3. Grain size, componentry and density analyses
1969; Thiele et al., 1998; Moreno et al., 2000, 2018). The subse-
quent effusion of 'a'ā and pahohehoe lavas, more mafic than those pre- Samples were dried at 40 °C for 48 h. A total of 26 grain-size analy-
viously erupted (<53 wt% of SiO2), built a new 400 m high cone in- ses were carried out by manual sieving in the range from −1 to 6 phi
RR

side the collapse scar (PCa), with a total volume of 4 km3 (Lohmar, (Ф), at steps of 1 phi in order to obtain their grain size distributions
2000; Martínez et al., 2018; Watt, 2019). Some of these lavas have (GSD). Typically, lithologic componentry analyses were conducted on
dammed the waters of the current Laguna del Laja (Melnik et al., grain size fractions between −1 and 2 phi, however, in some cases this
2006). The Holocene eruptive activity has been sourced mainly from was not possible due to the small size fraction of those tephras (i.e.,
the main crater, as well as from lateral fissures and parasitic cones asso- largest particles up to 1 phi). In such cases, the 1 phi fraction alone
ciated with the re-emergent post-avalanche cone. These eruptions have was used. Each fraction was quartered, and >100 grains were separated
CO

produced tephra deposits (both PDCs and tephra fallouts) that have been to quantify their relative abundances using a digital microscope. The
primarily directed eastwards, and are exposed in sections adjacent to the density of these clasts was obtained from 98 samples of scoria, pumice,
frontier with Argentina, either overlying glacial till deposits or interbed- and lava, ranging from 1 to 3 cm in diameter, using a pycnometer filled
ded within lacustrine sediments (González-Ferrán, 1995; Moreno et with distilled water (Supplementary Fig. 2 and Supplementary Table 2).
al., 2000; Torres et al., 2008; Urrutia et al., 2010). Porous particles were covered with hot paraffin in order to make them
waterproof. The particles were added to the pycnometer, which was
3. Methods then weighed using a Pocket Scale MH-200 (measures between 0.01 and
UN

200 g), giving the weight of the particle. The weight of the displaced liq-
3.1. Field stratigraphy uid was then determined, and hence the relative density of the particle.

Tephra thickness and particle size data were obtained from nineteen 3.4. Petrography, texture and mineralogy
field sites, and stratigraphic data was obtained from another 30 sites.
From these 49 descriptive sites, 22 composite stratigraphic columns In order to describe their crystallinity (mineral assemblage, crys-
were constructed (Fig. 1B; see Supplementary Table 1 for location, dis- tal size and their relative abundance), 2-D texture and vesicularity, 10
tance from the main cone, etc.). All observations occur at distances be- thin sections were prepared from selected scoria and pumice samples,
tween 0.9 and 21 km east-southeast of Antuco crater. Thickness, struc- with size ranging from coarse lapilli to bomb. Slides were analyzed un-
ture, grading and particle size of identified tephras were registered der polarizing microscope at the Universidad de Atacama. 3-D textures
4 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

F
OO
PR
Fig. 2. Field observations of tephra deposits. A: Proximal facies at P15. The clasts show some degree of agglutination. TF = tephra fallout. B and C: Medial facies in S5 and S14. Well

D
stratified deposits were accumulated in flat areas and exposed by subsequent erosion. D and E: Medial-to-distal facies exposed due to the retreat of the lake, showing the interplay between
clastic sequences (gravels, sands, beach deposits and lacustrine sediments) and tephra units at sites 27 (D) and 23 (E). F: Terraced sediments next to Río Petroquines showing lacustrine,
pyroclastic and moraine deposits in S21. G: Pyroclastic sequence deposited above a moraine deposit (MR) in S20, which in fact overlies lacustrine sediments (LS). The tephra deposits are
interrupted by aeolian sand unit (S) and their top portions have been eroded and reworked (RW). H: Lacustrine sediments and paleosols interrupted by several ash fall layers at S18. I:
TE
Terraced pyroclastic deposits and alluvial sediments found in S28. H and I correspond to distal facies.Photographs B and I were taken by Luis Cortes and Hector Moyano, while the rest
were taken by J.E. Romero.

Table 1
Antuco 14C ages. All the analyses were carried out at DirectAMS (USA). All the ages are given in yBP (present considered to be 1950 by convention). S17A and S17B were sampled at S17,
while M3734 was sampled at S27.
EC

Radiocarbon Calibrated
age age yBP Relative
yBP ± 1σ (2σ, area under
Sample error 95.4%) distribution Method Source Material

210199-6c 270 ± 60 238–453 0.580 – Lohmar Charcoal


(2000)
RR

M3734 1247 ± 30 1055–1185 0.950 AMS This Organic


study matter
1490 ± 55 1271–1435 0.967 AMS Urrutia Sediment
et al.
(2010)
1915 ± 50 1702–1929 1.000 AMS Urrutia Sediment
et al.
(2010)
CO

S17-A 2978 ± 26 2967–3179 0.999 AMS This Paleosol


study
S17-B 3022 ± 28 3055–3248 0.941 AMS This Paleosol
study
3015 ± 50 3065–3357 1.000 AMS Urrutia Sediment
et al.
(2010)

(See Fig. 3 and Section 3.2 for details.)


UN

of 10 selected juvenile particles of scoria and pumice (fine to coarse tometer (XRD) D8-Advance at DEMEX-EPN. The qualitative and quan-
lapilli in size; same tephras than those analyzed under 2-D observa- titative identification of mineral phases were carried out with the soft-
tions), were deciphered through the use of a Tescan-Vega scanning elec- ware Diffracplus (EVA) and these phases were compared with the XRD
tron microscope with energy dispersive X-Ray spectroscopy (SEM-EDS) spectrums of the International Center of Diffraction Data (ICDD) data-
operated at 15.0 kV and with a zoom of 20–250× at the Department base.
of Extractive Metallurgy of the National Polytechnic School of Ecuador
(DEMEX-EPN).
Mineralogical characterization was undertaken on pulverized sam-
ples of pumice and scoria (lapilli-to-bomb in size) using X-Ray diffrac
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 5

3.5. Bulk, glass and crystal geochemistry 4. Results

Geochemistry was obtained from different sources. For bulk compo- 4.1. Stratigraphy
sitions, 12 scoria and pumice samples, with weights >7.5 g, were an-
alyzed using X-Ray Fluorescence (XRF). Samples were crushed in an 4.1.1. Tephrostratigraphic framework
agate mortar (1 g) and subsequently mixed with fluxes such as lithium Preservation of tephra and the overall completeness of the broader

F
metaborate. Melting was carried out in a platinum melting pot in or- tephrostratigraphic framework are dependent on the distance from the
der to make a melted bead in the Katanax fusion fluxer. The glass beads vent and strongly influenced by the interaction of physical processes as-
were analyzed using a Bruker S8 Tiger XRF instrument and Spectra Plus sociated with slope, fluvial, glacial and lake environments. Most of these

OO
software. Loss on ignition (LOI) was obtained from continuous burn- detrital deposits that have interacted with the fresh volcanic products
ing using a SNOL muffle at 950 °C for a period of 30 min, and the LOI were described and mapped in a series of previous works (e.g., Moreno
values never exceeded 1.46%. Identified major oxides correspond to et al., 1984; Lohmar, 2000; Melnik et al., 2006; Urrutia et al.,
SiO2, Al2O3, TiO2, MgO, CaO, Na2O, K2O, P2O5 and iron oxide (as to- 2010) and the interpretations of their origin are used along with the re-
tal FeO). Surface geochemical composition of 8 selected juvenile parti- sults of this study.
cles (pumice and scoria) was obtained by SEM-EDS with a Bruker Quan- According to a facies analysis and interpretation of pyroclastic de-
tax EDS instrument with Secondary Electron Detector (SE). Quantitative posits (see Cas and Wright, 1987 for details), we were able to discrim-
analysis was carried out through the software Espirit 1.8. Errors for indi- inate between individual pyroclastic units (produced by tephra fallouts

PR
vidual elements were always below 1%. This technique provides “bulk” or PDCs) and composite sequences of deposits. These eruptive sequences
information on the glass composition, crystallinity and degree of sec- frequently show i) individual pyroclastic units interrupted by thin layers
ondary alteration of the particle (Taddeucci et al., 2007). The analyt- of other clastic sediments (indicating short periods of time with paucity
ical error associated with spectrum quantification is within 10% of the between eruptions), or ii) they are composed of a series of tephra fall
measured value (5% for oxides with abundances >10 wt%), while par- and PDC deposits. In both cases, their facies indicate similar erupting
ticle roughness and orientation relative to the EDS detector causes an conditions. Internally bedded tephra fall deposits that do not show in-
error of 15% (Lautze et al., 2012). terbedded layers of soils or other clastic deposits, are interpreted as be-

D
For major element analyses acquired through electron microprobe, ing produced by pulsatory eruptions (see Houghton and Carey, 2015).
samples were prepared by putting them into epoxy resin and left into Only El Pino Tephra (EPT) and Correntoso Eruptive Sequence have been
vacuum for c. 12 h to let the resin enter within the porous clasts. Epoxy previously described in the field by Lohmar (2000), but not yet strati-
chips were afterwards cut and polished, and a 13-nm carbon coating was graphically defined. Once properly described, tephras were correlated
TE
applied to the polished surface of the chips to be analyzed. Major ele- and named according to the local names of the closest creeks and rivers
ments compositions were determined with a Cameca SX100 electron mi- to the typo locations (stratotype) of each unit.
croprobe analyzer (EMPA) at the Department of Earth and Environment, The proximal tephra facies, located in the upper flanks of Antuco
University of Munich. The analyses were carried out using a 15 kV accel- (<4 km from the crater), are characterized by meter-thick pyroclastic
eration voltage and 20 nA beam current. A defocused 10-μm beam was units (Fig. 2A), and the deposits are mainly composed of coarse-grained
used for the investigation of groundmass/glass; while for the crystalline tephra (i.e., bombs and blocks) that exhibit variable degrees of welding
EC

phases, a 1-μm focused beam was used. The standards used for calibra- (from slightly to densely welded according to Houghton and Carey,
tions were: orthoclase for Al and K, albite for Si and Na, wollanstonite 2015) and strong dips (>40°) in the slope direction. At lower al-
for Ca, periclase for Mg, illmenite for Ti, iron oxide for Fe and chromium titudes, medial deposits, identified close to the foothill of the cone
oxide for Cr. The uncertainty is between 1 and 5% for major elements (4–6 km from the crater), consist of planar surfaces and terraces ex-
and up to 10% for elements at concentration <0.2 wt%. posed by fluvial erosion that preserve and register a record of primary
bedded, medium-to-coarse grained tephras (lapilli to bomb-sized) (Fig.
RR

3.6. Isopach and isopleth mapping 2B). However, in these areas, no intervening soils have formed that
might separate and distinguish each tephra unit, but instead other col-
Despite the large number of field observations, low preservation of luvium or alluvium layers roughly mark their stratigraphic limits. In
tephra fall deposits and limited access to some areas allowed us to con- some cases, the surface of these deposits appears terraced and is cov-
struct only 3 isopach maps, based on a minimum of 10 and a maximum ered by colluvium deposits (Fig. 2C). In areas adjacent to Laguna del
of 21 control points each map. The limitations of our results are as- Laja, primary pyroclastic beds are often separated by many meters of
fluvial and lacustrine sediments (Fig. 2D). At the site S27, a radiocar-
CO

sessed in the Discussion section. From the isopach hand-drawing, bulk


tephra volumes were calculated by integration, using the exponential bon date obtained within organic-rich lacustrine sediment below tephra
thinning (Pyle, 1989; Pyle, 1995) and Weibull methods (Bonadonna beds yielded an age of 1055–1185 y BP (Table 1). At the same local-
and Costa, 2012). With those volumes, the deposit mass and magni- ity, beach deposits formed by medium size detrital rocks (fine polymic-
tude were calculated (Pyle, 2000). For isopleth reconstruction, field tic gravel) and also laminated silt and sand packages are the most com-
measurements of the three largest scoria or pumice clasts with the geo- mon sedimentary features, in addition to volcaniclastic and pyroclas-
metrical mean of their 3 axes were used, as recommended in the liter- tic deposits (Fig. 2D and E). At distances >8 km east from the main
crater, well preserved sedimentary sequences (Fig. 2F) consist of la-
UN

ature (Supplementary Table 1; e.g., Barberi et al., 1995; Bonadonna


and Costa, 2013). This methodology has also been used with other custrine sediments, which have been buried by diamict deposits (5–6 m
scoria tephra falls (e.g., Andronico et al., 2014). The column heights thickness) that are interpreted to represent glacial till deposited during
were calculated using the methodology described by Carey and Sparks a glacial advance during the Holocene (~10 ka; Melnik et al., 2006;
(1986). Eruption duration was subsequently estimated by using the to- Penna et al., 2011). These till deposits are in turn overlain by tephra
tal mass erupted and the MDR from the equations of Sparks et al. beds identified in this study (Fig. 2G). As soil formation in this area
(1997a, 1997b) and Mastin et al. (2009). occurs at low rates, or is impeded by both the suppression of vegeta-
tion by volcanism and associated aridity, just a few paleosols are iden-
tified at about 17 km east from the crater, and they are intercalated
by thin (up to a few cm) pyroclastic units (Fig. 2H), most of them
fine-to-medium in grain size (ash to fine lapilli). Two of these paleosols
6 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

have been dated (Fig. 3; S17-A and S17-B), and they gave calibrated of five layers named α, β, γ, δ and ε from the base to top. Unit α is com-
radiocarbon ages of 2967–3179 y BP and 3055–3248 y BP, respectively posed of a basal layer with c. 3 cm of fine-to-medium grained iridescent
(Table 1). Near the international border between Chile and Argentina, scoria lapilli, directly overlain by 30 cm of gray-blue massive ash (Fig.
at a distance of >20 km from the crater, the pyroclastic deposits turn 5A) supporting individual scoria lapilli fragments. This architecture is
thicker and coarser again (there is some degree of redeposition in PDCs, interpreted as a tephra fallout (shower bedded) sequence, followed by
which turns them thicker) and form outcrops of many meters (Fig. 2I). a subsequent PDC deposit. The occurrence of 17 cm of finely laminated

F
These distal PDC deposits lack of sufficient information to be correlated silt overlying this unit (Fig. 4) suggests lake encroachment (transgres-
with proximal units, thus they are not counted in the present report of sion) following deposition of the tephra.
units/eruptive events. A general stratigraphy is shown in Fig. 3, while Another eruption is registered by unit β, which consists of at least six

OO
stratigraphic correlations are presented in Fig. 4. tephra fall layers composed of medium-to-coarse grained scoria lapilli
with a total thickness of 87 cm (Fig. 5B). The basal layer is oscilla-
4.1.2. Units α to ε tory graded, while subsequent deposited layers are, from bottom to
Units α to ε (Fig. 3) have been identified exclusively at the site top, inversely graded, ungraded and normally graded and their indi-
S23 (Fig. 4), about 8.6 km southeast from the main crater, and consist vidual thicknesses decrease progressively towards the upper part of

PR
D
TE
EC
RR
CO
UN

Fig. 3. Integrated stratigraphic column, with the ages of each deposit (in red), the names of the units (in black) and the type of pyroclastic deposit: TF corresponds to tephra falls and
PDC to pyroclastic density currents. Numbers between brackets are the number of individual deposits which compose each unit. Units with blue color (such as α and ESES) have a large
proportion of basaltic glassy ash in the matrix, which give them that distinctive “bluish” color. Thicknesses are the maximum observed in the field. Paleosols are not shown, as they are
discontinuous between different stratigraphic sections and they are properly shown in Fig. 4. (For interpretation of the references to color in this figure legend, the reader is referred to
the web version of this article.)
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 7

F
OO
PR
D
TE
EC
RR
CO
UN

Fig. 4. Stratigraphic correlation of pyroclastic units found in several locations east of Volcán Antuco. Location of each site is reported on top of each stratigraphic section, and thicknesses
are on the left side. Radiocarbon samples are labeled with red color, at their original position of sampling. The green box shows a zoom into the upper part of S23 stratigraphy, which
is compared to the core sampled in lacustrine sediments by Urrutia et al. (2010). (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)

the unit. These bedding characteristics are interpreted to represent pul- Unit γ lies over c. 6 cm of lacustrine silt, and corresponds to a 14-cm
satile eruptive behavior, while shorter episodes or strong shifts in the thick tephra fall, composed of fine-to-medium size lapilli, oscillatory
dispersal axis of the plume may also explain this thickness variation. graded, with high content of hydrothermally altered scoria (
8 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

F
OO
PR
D
TE
EC
RR
CO

Fig. 5. Field observations of the different pyroclastic units identified within this study. All the units are in stratigraphic order, from oldest to youngest. Pictures from A to F, and also I
and R, correspond to S23. Black shovel has ~40 cm in length. In A, the αS represents a basal scoria fall layer, while αF is a PDC (or massive flow) unit. Pictures F, G and H, show field
exposures of EPT (note the color grading to the top). The latter two correspond to S28 and S29. The best exposure of CES is seen at S16 (J), but its uppermost portion has been eroded,
UN

and covered by unit PT. The black line in I indicates a body composed of sand between silt. In K, EVES unit (S27) has a medium lapilli tephra fallout to the base, and several fine ash
layers to the top. Inserts L, M, N, O correspond to PT tephra at its respective S1, S8, S7 and S20 locations. This unit presents a lower pumice unit overlaid by a black scoria unit, however
in different sites a basal, lithic rich layer (M), or a top, black ash unit (N) are identified. Shovels in L, M, Q and S are 53 cm in length. Pictures P, Q and S are found in locality S13, while
T, U, V, W and X correspond to sites S7, S22, S10, S5 and P7. The unit ESES is observed in R, S and T, and from U to X units APDC, LBT, EAES, and BT are listed. All the localities can be
tracked in Fig. 1 and Supplementary Table 1.All the photographs were taken by J.E. Romero.

Fig. 5C). This deposit may be interpreted as produced during duce this type of tephra fall deposits (see Houghton and Carey, 2015).
phreatomagmatic/Vulcanian activity. More than c. 34 cm of sand con- This deposit is overlain by 23 cm of well-sorted bedded fine sands of
taining reworked and dispersed volcanic particles divide this unit from low-energy fluvial origin.
unit δ. A strong change occurs in unit ε, which is exclusively composed
Unit δ has 13 cm of thickness (Fig. 5D) and consists of medium size of pumice. This unit consists of three individual tephra fall layers,
scoria lapilli, with no grading, diffusely bedded, with centimeter-thick two of them (bottom and top) composed of clast-supported fine-to-
ash layers. Strombolian eruptions with pulsatile behavior typically pro
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 9

F
OO
PR
D
TE
Fig. 5. Continued

medium white pumice lapilli, while the middle layer consist of fine ash 4.1.4. Units ζ, η and θ
(Fig. 5E). El Pino tephra is covered by 8 cm of coarse sand at S23, which is
EC

overlain by 22 cm of tephra fallout (Fig. 5I). This unit (ζ) consists of


4.1.3. El Pino tephra (EPT) 3 layers, which correspond to 10 cm of inversely graded lithic-rich fine
This unit has been recognized at four locations (S20, S23, S28 and lapilli and normally graded 4 cm of coarse ash and 8 cm fine-to-medium
S29; see Figs. 2 and 3). In its type locality at S23, it is rather more com- lapilli scoria.
plex than other pyroclastic units, and consists of four different tephra The unit η consists of 23 cm of normally graded, fine to medium sco-
fall layers (Fig. 5F): The basal layer corresponds to 13-cm inversely ria lapilli (Fig. 5I). Between layers η and θ, 30 cm of lacustrine silt show
graded medium size pumice lapilli fall deposit, texturally identical to the
RR

a striking sedimentary structure of highly deformed beds (Fig. 5I), in-


one observed in unit ε. Above this, >50 cm of coarse lapilli to bomb terpreted as soft-sediment deformation. Tephra fallout θ is composed of
size pumices show inverse grading. These pumices are yellow to orange medium size, normally graded scoria lapilli. This set of pyroclastic units,
in color, and their shape is subangular. The sequence continues with an- composed of fine pumice lapilli intercalated with black ash layers, is ten-
other 50 cm of medium lapilli pumice with no grading. Both lower and tatively correlated with those beds occurring within S17 based on their
middle subunits are composed of low vesicularity pumices. At the top, a position closely bellow Petroquines Tephra (PT; see below) (also present
lithic-rich layer of 20 cm of fine to medium lapilli has angular and juve- in S23), and a radiocarbon age (3025–3248 y BP).
CO

nile scoria clasts with low-vesicularity. This deposit is also observed at


S20, with a total thickness of 82 cm, and at the Pino River stratigraphic 4.1.5. Correntoso eruptive sequence (CES)
section of Lohmar (2000), where it is 85 cm thick. Two distal outcrops This unit has been recognized in several sections along the Pichachen
of this tephra are found in S28 and S29. This same tephra is also rep- international road, at S16 (A and B), S17, S18, S20, S28 and S29 (Fig.
resented in the Pichachén stratigraphic section (Lohmar, 2000) where 4). The deposit corresponds to a pyroclastic density current, consti-
an associated 35 cm thick PDC deposit is composed of 60% gray-color tuted by c. 70% matrix and 30% clasts including lithics and scoria,
ash matrix and 40% of scoria, lithics and pumice clasts. Above this, a and has been correlated to the “Pichachen” stratigraphic section re-
UN

tephra fallout deposit consists of a 3 cm layer of medium lapilli scoria,


trieved by Lohmar (2000). The architecture of the deposit is vari-
which turns into 17 cm of pumice towards the top. The pumice is white
able, including its thickness. For example, at S20 (8.2 km SE from the
colored and shows inverse grading (Fig. 5G). In both S28 and S29, the
crater), it consists of a basal PDC with a thickness of 45 cm, com-
vesicular pumice turns dense (poorly vesicular) and blocky to the top,
posed of a matrix of medium size brown-to-blue ash, which supports
and its thickness varies from 12 to 10 cm (Fig. 5H). This deposit is over-
lapilli-size scoria. The upper 20 cm are laminated; however, a second
lain by another 80-cm thick PDC unit which is composed of 75% of gray
unit on the top, with a similar thickness, is also laminated, but com-
ash matrix and 25% of pumice, scoria and lithic fragments.
posed of fine ash. Section S16 is characterized by a massive deposit (cf.
laminations of S20) supporting scoria lapilli clasts, with a thickness of
>70 cm, which show diffuse lamination to the top with discontinuous
mm- (or cm-) thick fine lapilli lenses (Fig. 5J). Near the bottom, a sco
10 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

ria lapilli tephra fall deposit is found (Fig. 5J), and this feature is 4.1.9. El Soldado eruptive sequence (ESES)
also observed in S17 and S18, were total deposit thicknesses are much This pyroclastic unit has been identified at 6 sections along the east-
smaller (6 to 9 cm). At a distance of >20 km from the vent, this deposit ern flank of Volcán Antuco (S6, S12, S13, S14, S23 and S27; Fig. 4), and
has a total thickness ranging from 55 to 67 cm, but is represented by can be correlated to the unit found 231–253 cm deep within the lacus-
a massive ash matrix supporting clasts, with no evidence of lamination, trine sediment core of Urrutia et al. (2010). It consists of a basal, in-
clastic lenses or basal tephra fallout. versely or non-graded lapilli scoria fall deposit, up to 34 cm thick, which

F
is directly overlaid by two matrix-supported PDCs, that contain respec-
4.1.6. El Volcán eruptive sequence (EVES) tively >80% of ashy matrix, and 40% ash matrix supporting lapilli sco-
El Volcán eruptive sequence consists of a basal black lapilli scoria fall ria clasts and lithics, both with a total thickness up to 70 cm. Proximal

OO
deposit with inverse grading and c. 23 cm thickness, which is covered outcrops in the rim of the collapse scar consist of approximately 10 m
by at least four coarse ash fall deposits separated by silt layers of <1 cm of massive PDC with abundant bombs at the base of the deposit (Fig.
thick (Fig. 5K). Thus, this record means an initial paroxysmal eruption 2A). Most of the sections mentioned here correspond to medial sites,
followed by closely-spaced set of minor eruptive events. about 8 km distant from the vent, showing important variability in their
internal architecture and total thickness. For example, at S23, the PDC
4.1.7. Petroquines tephra (PT) is composed of abundant cauliflower-shaped bombs and blocks, some
Petroquines tephra comprises a lower pumice fall deposit immedi- of them reddish in color, and crossed lamination is observed to the top
ately overlain by a black scoriaceous unit. This distinct tephra couplet (Fig. 5R). At S28, the succession is composed of a basal medium lapilli

PR
is one of the most readily recognizable units preserved on Volcán An- scoria with reverse grading, 50 cm-thick medium ash matrix supporting
tuco and therefore is a key stratigraphic marker. This unit is clearly rec- cauliflower shaped bombs and blocks, and a upper cross-stratified de-
ognized at S7, S8, S14, S16 (A and B), S17, S18, S20, S23, S27, S28 posit with blocks and bombs. In the case of S13, the sequence consist of
and S29 (Fig. 4), and at many other sections. A proximal outcrop of a basal coarse ash to fine lapilli scoria fall layer with reverse grading,
this tephra fall deposit in S1 (3.17 km northeast from crater) shows both which is directly overlaid by c. 23 cm of strongly bedded, cross-stratified
pumice and scoria units. At that site, the yellow pumice fallout deposit is fine to coarse ash with some insights of bomb sags. This layer is overlaid
>20 cm thick (non-visible base) and composed of clasts lapilli to bomb/ by a package of silt and ash beds, c. 6 cm-thick, which in some cases is

D
block, with maximum diameters of ~10–15 cm (Fig. 5L). Above this deformed showing convoluted lamination and is affected by fluid escape
layer, there is a 120 cm thickness slightly welded spatter deposit, with structures (Fig. 5S) originated from the unit below. A second flow unit
maximum clasts of 50 cm and varying from dark gray to reddish in color (11 cm-thick) rests over those laminated silt and ash sediments, and it is
(Fig. 5L). Bedding is not observed, but in the lower section (0–30 cm) massive and composed of a fine to medium ash matrix supporting c. 35%
TE
the deposit develops inverse grading and contains only black-gray ju- of fine lapilli particles which include scoria, gray pumice and lithics.
venile material, followed by a medium section (30–80 cm) showing the The upper unit (c. 22 cm-thick) is full of lapilli scoria and deformed
largest clast fragments formed mostly by reddish bombs, which abun- rip-up clasts of laminated sediments (Fig. 5S) that may represent sub-
dance is between 30 and 40% vol. The upper section (80–120 cm) is strate lithic clasts entrained during the PDC emplacement. A 4-cm thick
characterized by the size reduction of the reddish spatter, having a nor- layer can be distinguished to the top, which is composed of a fine to
medium ash matrix supporting c. 15% gray pumice particles. Due to its
EC

mal grading. In contrast, the gray bombs are ungraded. The full expo-
sures are found at S8 and S23, where they consist of a basal lithic-rich, distinct sedimentary characteristics, this deposit is further discussed in
medium sized lapilli tephra fall, composed of dense hydrothermally al- Section 5. At S6 (Fig. 5T), located 8.4 km from the eruptive source,
tered porphyritic clast overlaid by a normally graded pumice fall de- this unit is thinner and shows evidence of changes in the flow regime
posit, which has abundant hydrothermally altered scoria and a top unit because of lapilli scoria lenticular (lensoidal) beds. According to Urru-
composed of normally graded scoria lapilli (Fig. 5M). The contact be- tia et al. (2010), this unit is constrained to an age of 1000 to 1100 y
tween these two layers is often gradational (Fig. 4N and O). At S27 the AD, and from our M3734 radiocarbon dating (Fig. 4), obtained below
RR

top scoria fall shows three different sub-layers with inverse grading, and PT unit (which is younger than ESES), a calibrated age of 1055–1185 y
bomb-sized scoria are also found. At this site, organic matter found in BP was obtained (895–765 y AD), indicating that the age of Urrutia et
sediment 3 m below PT has yielded a radiocarbon age of 1055–1185 y al. (2010) is consistent with our stratigraphy and radiocarbon dating.
BP, thus suggesting a much lower age for the PT unit.
4.1.10. Arriero pyroclastic density current (APDC)
4.1.8. Zorrito tephras (ZT) This unit has been exclusively found at S23, overlying ESES and con-
CO

The Zorrito tephras have been found at S5, S13 and S14. At S14, sists of a massive, 50-cm thick pyroclastic flow deposit containing c.
typically consists of three discrete layers: a basal layer 24 cm thick 60% of scoria lapilli clasts, c. 35% of accidental and accessory lithics
composed of normally graded scoria lapilli with largest particles up to (mainly dense porphyritic rocks, hydrothermally altered clasts and frag-
3.5 cm diameter. A thin (1-cm thick) paleosol divides this unit from ments of the former ground), and c. 5% of reddish scoria (Fig. 5U).
the upper fallout layer, which consists of 2.5 cm of vesicular fine scoria The supporting matrix is light gray in color, and the clasts are nor-
lapilli inversely graded. A similar stratigraphy has been found at S13, mally graded. To the top, this unit has developed an incipient soil.
where the basal, non-graded scoria lapilli deposit thickness is 23 cm; Through stratigraphic correlation, this unit is observed at a depth of 130
UN

however, the tephra has been strongly weathered by ground-water cir- to 179 cm in the lacustrine core of Urrutia et al. (2010), and its age
culation (Fig. 5P). At the top, and separated by 18 cm of silt sediments should be close to 1380 y AD. A similar PDC was reported by Lohmar
and 44 cm of sand, three tephra layers are found, the lower composed of (2000) in the southernmost Laguna del Laja, where charcoal fragments
coarse scoria lapilli, the middle layer of lapilli-sized scoria and the upper within the deposit yielded a calibrated age between 238 and 453 BP
of highly weathered inversely graded scoria lapilli (Fig. 5Q). Similarly, (1712 and 1497 y AD). Because of the age discrepancies between these
at S5, the deposit consists of a basal layer (4 cm thickness) composed of two dated PDC deposits, it is conceivable that the unit described by
non-graded scoria lapilli reaching up to 1.5 cm in size. This is followed Lohmar (2000) might represent a separate younger event not recog-
by a middle layer of 2 cm thickness, composed of coarse scoria ash, and nized in the present study. Clearly, further work is required to clarify
the top layer consists of 4 cm of inversely graded scoria lapilli, with the this.
largest clasts up to 1.5 cm.
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 11

4.1.11. Los Barros tephra (LBT) 26 cm) is composed of bombs (<14 cm diameter), some of them slightly
This unit corresponds to a tephra fall deposit distributed in the south- welded, then grading to fine lapilli at the top layer (26–32 cm). At Site
east flank of Antuco (S4, S5, S9, S10, S11, S13, S23 and S14), and is 5 (4.7 km from the main crater), 3 layers and a total thickness of 26 cm
clearly recognizable, because it is composed of highly vesicular irides- are observed. The base layer (0–4 cm) is formed by dense lithics and
cent scoria (i.e. density <0.9 g/cm3). The architecture of this unit is scoria showing inverse grading. From 4 to 11 cm, a second and un-
well represented at S10 and consists of three layers (Fig. 5V): a in- graded layer consists of fine iridescent scoria, while towards the top

F
versely graded deposit of 4 cm thickness with largest scoria particles up (11–26 cm), one is formed of inversely graded scoria lapilli. At site 4,
to 1.5 cm at the base, overlaid by a coarse ash layer of 3 cm thickness, located southeast of the crater (4.53 km distance), BT unit is thinner (c.
and normally graded scoria layer at the top, with the largest scoria frag- 10 cm thickness) deposit and composed only of two layers. The basal

OO
ments up to 2.5 cm and some of the largest particles having platy shape. layer (0–5 cm) comprises inversely graded lapilli sized scoria, while the
upper layer (5–10 cm) is characterized by ungraded lapilli sized sco-
4.1.12. El Arenal eruptive sequence (EAES)
ria. Due to its morphological expression and stratigraphic position above
This unit is well represented at S5 and S23, but also observed in the
EAES unit, this tephra must be younger than AD 1800. Considering that
lacustrine core described by Urrutia et al. (2010). It consists of at
the last major eruptive event at Volcán Antuco occurred in AD 1852–53
least 4 low-to-moderately vesicular scoria fall layers, interrupted by la-
(Smith, 1915) with tephra [widely] dispersed to the east, we therefore
custrine silt layers (Fig. 5U) or paleosols (Fig. 5W). At S5, the sequence
conclude that Bueyes tephra likely originated during the 1852–53 erup-
starts to the base with a scoria deposit of 22 cm thickness, which is in-

PR
tion. It is interesting to note that the lacustrine sediment core outside
versely graded and has a higher proportion (c. 25%) of hydrothermally
of this eastward directed dispersal axis, has not registered the 1852–53
altered (reddish) scoria to the bottom, while to the top is full of unal-
tered, highly vesicular scoria. Another layer, with a thickness of 17 cm, tephra or its BT correlative.
mostly consists of ungraded moderately vesicular scoria, and few red-
dish scorias (c. 5%). The two upper layers are 8 and 5-cm thickness re- 4.2. Grain size, lithology and texture of tephras
spectively, and correspond to ungraded medium size lapilli. Similarly, at
S23, the basal layer is the thicker of the sequence, and it reaches up to The integrated stratigraphic column, showing all recognized units
of this study, in addition to the GSD and lithologic components of the

D
15-cm thickness and it is normally graded and composed of medium to
coarse lapilli. The following deposits are relatively thin (2 to 6-cm thick- tephras, is reported in Fig. 6A. GSD from tephra fall deposits have their
ness each layer) and they are also composed of scoria lapilli. This inter- mode at 1 phi (particles larger than coarse ash), while PDCs display
nal architecture of the EAES may indicate that it was produced during their mode at 3 phi (fine ash) (Fig. 6A). According to the classification
TE
four different eruptions, shortly spaced in time. Stratigraphic correla- scheme of Folk and Ward (1957), from a total of 29 GSDs, 3.5% are
tions to tephra layers found in a coring site in the northernmost edge of considered as moderately well sorted, 3.5% as moderately poorly sorted,
Laguna del Laja (Urrutia et al., 2010) are possible as using the stratig- while 7% are moderately sorted and the majority (86%) is poorly sorted.
raphy of the core, plus textural properties of the tephra and geochemical Tephra fallout deposits have a median grain size (Mdφ) ranging from
features of the EAES glassy groundmass (i.e. SiO2 vs Alkali). These four −1.46 to 0.07 φ (fine lapilli to very coarse ash), while PDCs range from
tephra fall layers are found between 30 and 59 cm depth in the lacus- 0.63 to 2.39 φ (medium to very coarse ash) (Fig. 6B). The sorting coef-
EC

trine sediment core from Urrutia et al. (2010), and they have thick- ficient (σ1) of fallout deposits is constrained to a range from 0.57 to 2.03
nesses of 4, 1, 1 and 3 cm from base to top, being constrained to an age φ (moderately well sorted to very poorly sorted), and the PDCs show a
between 1750 and 1850 y AD. The top layer corresponds to the Bueyes narrower sorting spectrum from 1.04 to 0.95 φ (poorly to very poorly
Tephra (BT deposit; see Section 4.1.13), and the subsequent three lay- sorted) (Fig. 6B). According to the classification of Cas and Wright
ers are correlated to EAES. LBT deposit is not expected to be observed in (1987) for pyroclastic deposits, all of these samples cover a range from
this lacustrine record according to its dispersal (see Section 4.6). These well sorted to poorly sorted.
RR

thicknesses are thinner than those of EAES seen in S23 (c. 8.6 km from The classification of Houghton and Wilson (1989) has been used
the crater), and consistent with the position of the coring site (13 km to qualitatively describe the vesicle fraction of pyroclastic particles:
northeast from the crater), out of the dispersal axis. According to the age non-vesicular (0%), very poorly vesicular (0–20%), poorly vesicular
span of all EAES tephras and the BT unit (100 years), this period of ac- (20–40%), moderately vesicular (40–60%), very vesicular (60–80%),
tivity was modulated by one eruption every an average repose time of c. and extremely vesicular (>80%). Vesicularity was visually estimated
20 years. from selected juvenile grains using optical microscope and SEM. We
CO

have distinguished several types of lithologic components within the


4.1.13. Bueyes tephra (BT) tephra deposits, and they are listed below:
This is the youngest unit recognized in the stratigraphic record; how-
ever, on the top surface of many of the studied sections (i.e. current a) Pumices (Fig. 7A and B), which consist of glassy particles with clear
ground level), the ground has medium size scoria lapilli, which certainly color (light brown to white), moderately to very poorly vesicular,
represents the latest small eruptive episodes originating from Antuco. with blocky shape and jagged to broken perimeter. Their vesicles are
This deposit has been studied at 15 sections, which are then reported ex- sub-spherical, elongated or irregular.
tensively in the section of distribution, volume and eruption source para-
UN

b) Scoria (Fig. 7C to E), comprising blocky to sub-angular glassy par-


meters. Medial outcrops (>3.5 to 10 km distance from the main crater) ticles, with color varying from black, to dark gray and even golden
consist of a 10 to 60 cm thick tephra fall deposit, characterized by sev- or iridescent. Their vesicularity goes from moderately to very poorly
eral layers differentiated by their internal grading and componentry. At vesicular, and vesicles are irregular rather than sub-spherical.
Site 2 (4.35 km from the main crater), 3 layers are identified and a total c) Shards (Fig. 7F), made of transparent to semi-transparent, uncol-
deposit thickness of 32 cm was measured (Fig. 5X). The sequence is de- ored or slightly black glassy particles, without vesicularity, exhibit-
posited above an old laharitic deposit and consists of a basal layer with ing sharp or jagged perimeters as original constituents of inter-vesic-
inverse grading (0–14 cm), which is composed of dark gray lapilli sized ular walls. Filamentous shapes are commonly observed. Their surface
scoria. An intermediate layer (14– is usually smooth or “molten” (i.e. representative of rapid cooling of
plastic magma).
12 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

F
OO
PR
D
Fig. 6. Grain size and components of selected tephras. A: Grain size of tephras from dry mechanic sieving and lithological components of studied fractions are reported to the right. B: Plot
TE
of the grain size of tephra fallout and PDC samples, according to the classification of Walker (1971).
EC
RR

Fig. 7. Lithologic components of tephra deposits. A: Brown, low-density (vesicular) pumice. B: White, high-density (poorly vesicular) pumice. C, D and E: Blocky vesicular scoria, angular
CO

dense scoria and iridescent scoria, respectively. F: Glassy shards. G: Free crystals (mainly ol, plg and px). H, I and J: Fresh black lithics, yellow hydrothermally-altered lithics and reddish
oxidized lithics, respectively. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

d) Free crystals (Fig. 7G), which involves idiomorphic to xenomorphic In all the studied tephra deposits, black dense scoria is a major con-
individual crystals of plagioclase, olivine and occasionally pyroxene, stituent. However, at units ε and the base of EPT, there is a clear pre-
sometimes reaching up to 2 mm size. dominance of pumice, especially in the coarser grain size fractions (Fig.
e) Lithics (Fig. 7H to J) are recycled particles, angular to sub-an- 6A). Nevertheless, there is a considerable change in the middle EPT,
UN

gular, often porphyritic when fresh. They have variable vesicular- where all the fractions have similar amounts of dense scoria and lithics,
ity, and hydrothermal alteration is frequent and confers them white while towards the top, an increasing amount of both dense and irides-
(clay-covered), yellowish (jarosite and sulphur coverage) or reddish cent scoria is better observed (Fig. 6A). The amount of recycled mater-
color (goethite and hematite coverage). Their surface may show pit- ial is often >25% (Fig. 6A); however, in some cases the altered scoria
ting, which corresponds to dissolution due to the exposure to acid is only representative of post-depositional weathering of tephra, such as
hydrothermal fluids. in γ, ZT and ESES tephras.
f) Aggregates, composed of numerous particles, without any preferen- The juvenile materials of PT are composed of scoria and pumice.
tial origin (crystals, glass shards, lithics, etc.) adhered through elec- Scoria fragments are blocky, their vesicles are sub-spherical to irreg-
trostatic forces or hydrothermal cement. ular and surfaces are smooth or curved (Fig. 8A). Rarely, vesicles
have been filled by zeolite. Pumices are also blocky and poorly vesicu
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 13

F
OO
PR
Fig. 8. Morphology and petrography of different particles. A: Blocky scoria with smooth texture (S7-PT Base), B: Blocky pumice with sub-spherical vesicles (S7-PT Top). C: Scoria with
spongy texture and ropy skin texture (S13-ESES PDC). D: Scoria with polygonal shape and smooth-curved depressions (S13-ZT Base). E: Blocky scoria with ragged perimeter (S5-LBT). F:
Blocky scoria with polygonal to spherical vesicles (S5-BT). The units where these juvenile fragments were collected are mentioned in the upper right corner, while scales are in red. (For
interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

D
lated (30%), with sub-spherical vesicles (Fig. 8B) and phenocrysts of are blocky particles with scarce and irregular vesicles and contain abun-
plagioclase and augite. The juvenile scoria of ESES unit has spongy tex- dant microlites. In agreement with these authors, the samples from units
ture, extremely vesicular (>80%), and the particle surface is generally EPT, ESES and BT (ALV53) are described here as tachylite, and samples
TE
smooth, curved or ropy skin (Fig. 8C). Blocky particles are not common, from units PT-base, PT-top, ZT, LBT and BT (BE1853 and AE1853(2)) as
but they have higher phenocryst fraction. Most of the grains show in- sideromelane. Particularly, EPT is extremely microlite-rich (c. 85% of the
ter-vesicular walls with fluidal texture, and weak bulges are commonly groundmass).
observed as the result of quick quenching of the glassy particles prior to EPT tephra is formed by vitrophyric scoria; however, the most im-
the bursting of gas bubbles. Scoria from the basal ZT unit is very vesic- portant features correspond to its banded and trachytic textures (i.e.,
ular (70%), with polygonal shape vesicles and in some cases their col- oriented crystals). These scorias are poorly-to-moderately vesicular
EC

lapse is responsible of smooth-curved depressions limited by inter-vesic- (35–45%), with groundmass of about 48.5 to 58.5%, and often con-
ular walls (Fig. 8D). Weak bulges are also very common in these sam- tain plagioclase-porphyritic, angular andesitic enclaves. In the case of
ples. Scoria from the upper ZT is extremely vesicular (80–90%), but in PT unit, both base and top samples (Fig. 8A and B) are very similar
addition to blocky shapes, fusiform clasts are very common too. At LBT, in terms of their textures and crystallinity, but orthopyroxene is exclu-
the scoria fragments are blocky and their vesicles are spherical or ovoid, sively found at PT-base and are characterized by their intergranular and
while the particle surfaces are fluidal, smooth and ropy skinned (Fig. glomeroporphyritic textures without any sign of disequilibrium. In the
RR

8E). The perimeter of these particles is ragged. Finally, the scoria of BT ESES unit, a dense scoria (Fig. 8C) has only 7% vesicularity and mainly
unit has blocky morphology, moderate vesicularity (>45%), with polyg- vitrophyric, intergranular and hialophytic textures. A sample from ZT at
onal to spherical vesicles and smooth and ropy skin surfaces (Fig. 8F). S5 (Fig. 8D) and an LBT sample at S10 are characterized by extremely
vesicularity (86%). Finally, two samples of the BT unit and one from the
4.3. Petrography 1852–53 lava flow were studied. These two airfall samples correspond
to a cauliflower-shaped bomb (BE1853) found at S4, and a scoria bomb
The scoria fragments are composed of 5.0 to 14.5% of crystals from S5 (AE1853 (2)). BE1853 has up to 85% (Fig. 8E), while AE1853
CO

and a mineral assemblage comprising subhedral to euhedral plagio- (2) has 74% (Fig. 8F) of vesicularity and the lava sample (ALV53) has
clase (3–10%), subhedral to euhedral olivine (ol, 0.2–1.0%) and occa- higher crystallinity (35%). All the samples from ZT to BT show disequi-
sionally subhedral clinopyroxene, orthopyroxene and anhedral opaques. librium textures, including sieve texture in phenocrysts and resorption
The size of plagioclase phenocrysts ranges between 0.05 and 1.0 mm, rims in the plagioclase. In some cases, glassy groundmass is palagonized.
with the exception of ALV53 sample, which contains phenocrysts up Due to the high glass fraction of pulverized samples, it was very difficult
to 2.5 mm. Olivine occurs in sizes ranging from 0.02 to 0.15 mm; to recognize mineral phases through XRD analyses, and only plagioclase
however, in ALV53, it reaches 1.5 mm. Pyroxenes are generally small was confidently distinguished from spectrums (Supplementary material
UN

(clinopyroxene from 0.05 to 0.6 mm and orthopyroxene from 0.1 to 2).


0.3 mm when they occur) and opaques range from 0.05 to 2 mm;
however, the largest op crystals are found in PT unit. The vesicular- 4.4. Geochemistry
ity ranges between 7 and 86%, while the groundmass corresponds to
5.5 to 88%. The petrographic characteristics of every sample described The bulk geochemical composition of all the analyzed tephra fall
here are available in Table 2. According to Taddeucci et al. (2015), samples (Table 3) spans between basalts and andesites/trachyandesites
basaltic glassy particles may be described as sideromelane, if they con- (volatile free basis, normalized values; Fig. 9A; 50.18 to 62.22 wt% of
tain brown to golden glass with spherical vesicles, fluidal overall par- SiO2). Similar bulk compositions were reported from the tephras stud-
ticle morphology and few or no microlites; or as tachylite, when they ied by Lohmar (2000), covering a range from 50.67 to 61.79 wt%
of SiO2 (Fig. 9A); however, lava compositions from Ea and La stages
14 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

Table 2
Summary of petrographic features of selected pyroclastic samples from different units at Volcán Antuco.

Sample A2273 APS1N1 AE1853 APS13PF AES5N5 AES10 BE1853 AE1853 (2) ALV53
Unit EPT-middle PT-base PT-top ESES ZT LBT BT BT BT
Site Site 23 Site 1 Site 1 Site 13 Site 5 Site 10 Site 3 Site 5 Site 4
Phenocrysts (%) – microlites not considered
Plagioclase 3.0 4.0 3.0 4.0 3.5 10.0 4.0 3.5 26.0

F
Olivine 1.0 1.0 1.0 1.0 1,0 0.5 0.2 0.3 6.0
Clinopyroxene 1.0 1.0 2.0 0.0 1,0 2.0 1.0 1.0 3.0
Ortopyroxene 0.0 0.0 0.0 0.0 0,0 0.5 0.3 0.0 0.0

OO
Opaques 0.5 0.0 2.0 0.0 0,5 1.5 1.0 0.2 0.0
Vesicles 50.0 74.0 32.0 7.0 86,0 80.0 35–45 85.0 25.0
Groundmass 45.5 20.0 60.0 88.0 8,0 5.5 58.5–48.5 10.0 40.0
Norm. phenocrysts (%) – vesicle and groundmass free basis
Plagioclase 54.5 66.7 37.5 80.0 58.3 69.0 61.5 70.0 74.3
Olivine 18.2 16.7 12.5 20.0 16.7 3.4 3.1 6.0 17.1
Clinopyroxene 18.2 16.7 25.0 0.0 16.7 13.8 15.4 20.0 8.6
Orthopyroxene 0.0 0.0 0.0 0.0 0.0 3.4 4.6 0.0 0.0
Opaques 9.1 0.0 25.0 0.0 8.3 10.3 15.4 4.0 0.0

PR
Textures
Vitrophyric X X X X X X X
Sieve X
Glomeroporphyritic X X X X X X
Vesicular X X X X X X X X
Intergranular X X X X X X X X X
Trachytic X X
Hialophylitic X X
Banded X

D
TE
EC
RR
CO

Fig. 9. Geochemistry of juvenile products. A and B: Plot of Alkali v/s Silica from Le Bas et al. (1986) using bulk rock (XRF) glassy groundmass (EMPA) and surface (EDS) compositions
of tephra samples from Volcán Antuco. In A, the light-blue dotted circle show the compositions reported by Torres et al. (2008). C: FeO*/MgO vs. SiO2 discrimination diagram (after
Miyashiro, 1974) for bulk rock compositions of Antuco. FeO* refers to total FeO. Harker diagrams for representative major elements from bulk rock, glassy groundmass and surface geo-
chemical analyses are shown below. (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)Data sources for Antuco are
UN

included in the color key. Data from Chillán, Callaqui and Copahue volcanoes was obtained from Polanco (1998), Dixon et al. (1999), Naranjo and Polanco (2004) and Daga et al.
(2016).

also display trachydacitic compositions. The most mafic compositions basaltic trachyandesite (Fig. 9A; 44.1 to 50.64 wt% of SiO2). Glassy
(basaltic) are represented by the units BT and β, while the most fel- groundmasses of 4 samples, analyzed by EMPA (Table 5), show a dom-
sic extreme is represented by PT and EPT units (andesite). All these inant basaltic–andesite (units BT, EAES and LBT) and a trachytic (PT)
samples remain in the field of subalkaline (calk-alkaline) series; how- composition (Fig. 9B; SiO2 wt% from to 54.97 to 60.95). This is typi-
ever, bulk particle surface composition measured trough EDS (Table cal for glassy groundmasses, which often display higher SiO2 contents
4) yields an alkaline composition ranging between tephrite and than the bulk composition, as the crystals are less silicic than the melt
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 15

F
OO
PR
D
TE
Fig. 10. Isopach maps (cm) of PT (A), LBT (B) and BT (C) tephra fallout deposits. In D, the plot of thickness (m) vs. square root area (km) shows the regression of these deposits and
compares them with different basaltic to basaltic andesite eruptions of the Southern Andes: Navidad 1989–90 (Lonquimay Volcanic Complex), Pómez los Baños (Planchón Peteroa Volcanic
EC

Complex) and Calbuco 2015 eruption.Sources are: ⁎Naranjo et al. (1991), ⁎⁎Naranjo and Haller (2002) and ⁎⁎⁎Romero et al. (2016).

Table 3
Major element geochemistry (XRF) of selected bulk pyroclastic samples from different units at Volcán Antuco. Total FeO is reported as Fe2O3.

Sample A2265 A2268 A2269 A2272 A2273(1) A2273(2) A2274 A2275 APP02 APS1N1 ABE1853 APS13PF AB1853(1)
Site Site 23 Site 23 Site Site 23 Site 23 Site 23 Site 23 Site 23 P02 Site 1 Site 2 Site 13 P09
RR

23
Strat. Unit β Unit δ Unit ε EPT- EPT- EPT- EPT- EPT- EPT- PT-base PT-top ESES BT
unit base middle middle middle top pich
Na2O 2.92 3.75 4.78 5.23 5.17 5.27 5.21 4.87 4.93 4.9 5.01 4.09 3.26
MgO 5.22 2.76 1.11 1.05 1.18 1.24 1.2 2.21 1.29 1.36 1.68 2.62 5.16
Al2O3 19.07 17.02 15.9 16.29 16.26 16.28 16.41 16.61 15.99 15.88 16.45 17.49 18.29
SiO2 50.87 53.95 61.2 61.41 61.09 61.08 60.7 58.32 60.17 59.69 60.01 55.14 51.58
P2O5 0.18 0.31 0.35 0.34 0.34 0.35 0.35 0.33 0.35 0.34 0.38 0.29 0.24
CO

K2O 0.56 1.01 1.99 1.68 1.68 1.63 1.61 1.38 1.65 1.6 1.46 1.17 0.85
CaO 12.72 9.37 5.23 5.22 5.35 5.48 5.45 6.78 5.66 5.72 6.49 8.91 11.38
TiO2 0.99 1.54 0.97 0.88 0.91 0.95 0.95 1.04 0.96 0.91 1.04 1.1 1.11
Mn2O3 0.19 0.24 0.19 0.24 0.26 0.24 0.3 0.24 0.2 0.2 0.25 0.19 0.19
Fe2O3 8.66 10.24 6.63 6.56 6.61 6.77 6.73 7.52 6.9 6.66 7.47 7.77 9.02
LOI 0.59 0.69 1.46 0.92 0.92 0.67 0.94 0.64 0.86 0.71 0.03 0.88 0.13
Total 101.97 100.86 99.82 99.82 99.78 99.96 99.85 99.92 98.96 97.97 100.28 99.65 101.2
UN

from which they crystalized (Cashman and Rust, 2016). All the py- tions from anorthite to andesine (Supplementary Fig. 1): PT-phenocrysts
roclastic samples, and lavas from the ancestral Volcán Antuco, display correspond to andesine, while ZT display a range from bytown-
a tholeiitic affinity (Fig. 9). Major element oxides, such as CaO, Al2O3, ite-labradorite (similarly to EAES), LBT are exclusive bytownite and BT
and MgO decrease with increasing SiO2, while P2O5, TiO2 and FeOtot is exclusively labradorites.
show increasing trends, with an inflection point at ~55 to ~57 wt%
of SiO2, then followed by a decreasing trend (Fig. 9). In the case of
Na2O and K2O, they increase with increasing SiO2 (Fig. 9). Crystal chem-
istry derived through EMPA on plagioclase phenocrysts reveal composi
16 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

Table 4
Surface geochemistry (EDS) of selected juvenile grains from different units at Volcán Antuco. FeO and Fe2O3 are reported as total FeO.

Sample S7P17 S7E17 S6E17 S13E17-B S13E17-M S5L3 S5L2 1 S5L1


Site Site 7 Site 7 Site 6 Site 13 Site 13 Site 5 Site 5 Site 5
Strat. unit PT - base PT - top ESES ZT ZT LBT EAES BT
Na2O 4.24 4.49 4.98 5.53 4.17 4.33 3.41 4.35
MgO 2.46 4.0 4.39 5.66 4.97 3.86 3.24 4.52

F
Al2O3 28.19 18.59 16.8 20.18 13.83 19.08 23.9 23.52
SiO2 47.47 50.64 48.19 55.08 45.83 48.62 46.04 44.1
P2O5 3.85 – – – – 4.69 5.16 9.38

OO
K2O 0.82 0.81 0.65 0.76 0.74 0.46 0.36 0.57
CaO 2.55 6.93 8.83 3.62 7.6 6.6 4.41 4.25
TiO2 3.01 2.06 0.82 1.66 1.69 1.07 1.9 2.14
MnO – – 0.39 – – – – –
FeOT 7.42 12.5 14.96 7.52 21.17 11.3 11.58 7.16
Total 100 100 100 100 100 100 100 100

PR
Table 5 Table 6
Mean groundmass geochemistry (EMPA) of selected juvenile grains from different units at Summary of eruption parameters obtained from the integration of isopachs using two dif-
Volcán Antuco. FeO and Fe2O3 are reported as total FeO. ferent methods. Exponential volumes are estimated as V = 13, 08T0bt 2, where Tₒ = e
ln(Tₒ) is the maximum thickness and bt 2 = ln(2)/k ∗ √Π is the distance from the vent
Sample S7 S5-L4 S5-L3 S5-L2 S5-L1 where thickness halfs; Weibull volumes are estimated as where initial ranges
Site Site 7 Site 5 Site 5 Site 5 Site 5
for ϴ, λ, n are 0.1–5000 cm, 0.1–1000 km and 0.2–2, respectively (Pyle, 1989, 1995;
Strat. unit PT ZT LBT EAES BT
Measurements n=3 n=4 n=4 n = 14 n=5 Bonadonna and Costa, 2012). The total mass of these deposits was estimated as the
Na2O 5.55 0.04 4.01 4.06 3.97 product of the bulk density and the volume. MDR values were calculated solving the
MgO 1.44 39.93 4.43 4.6 4.68 equations of Sparks et al. (1997a, 1997b) and Mastin et al. (2009) for column

D
SiO2 60.95 39.64 54.97 55.18 55.02 height: HT = 0.220 ∗ MER 0,2590,259 and HT = 0.304 ∗ MER 0,241, for heights of <15 and
Al2O3 16.84 0.02 15.06 16.17 14.96 >15 km, respectively. Duration of eruptions is given by the quotient between the mean
P2O5 0.38 0.04 0.31 0.28 0.31 TM and the mean MDR (it is expressed in hours). An andesite magma density of 2750 kg/
SO3 0.05 0.03 0.12 0.05 0.01 m 3 was used for DRE estimation of PT, and a basaltic magma density of 2850 2750 kg/m 3
was used for LBT and BT. We assumed a 5% wt. of lithics when obtaining DRE volumes.
TE
Cl 0.09 0.02 0.07 0.03 0.06
K2O 1.68 0.01 1.16 0.93 1.25 Magnitude is defined as M = log10(mean TM) − 7 and Intensity is =log10(mean MDR) + 3,
CaO 4.44 0.22 7.55 8.34 7.16 following Pyle (2000).
TiO2 1.01 0.04 1.55 1.45 1.74
FeOT 5.85 22.09 9.65 9.62 9.81 Method Unit
Cr2O3 0 0.02 0.03 0.05 0.01
MnO 0.14 0.38 0.2 0.19 0.19 PT LBT BT
Total 98.44 102.47 99.04 100.95 99.19
EC

Exponential (km 3) 0.1749 0.0151 0.0826


Isopachs (cm) 5, 15, 25, 45, 100 4, 8, 12, 13 4, 15, 20, 25, 50
MRSE 0.1299 0.001417 0.002368
Segment 1 (km 3) 0.00287 0.0151 0.04533
c 6.466 0.3071 0.7567
4.5. Density m 0.3369 0.2017 0.1258
Segment 2 (km 3) 0.172 0.03727
c 6.711 2.649
RR

Only the eruptions with a wider representation in the field (i.e., EPT, m 0.277 0.1997
PT, LBT and BT) have been described here in terms of their density (Sup- Weibull (km 3) 0.1269 0.0150 0.0797
plementary Fig. 2 and Supplementary Table 2), in order to assess the MRSE 0.002368 0.001563 0.01098
θ 1.123 0.1111 0,377
eruption dynamics in Section 4.6. Pumices from the EPT unit have den-
λ 8.935 10.04 13.88
sities varying from 346 to 1787 kg/m3, but most of them are near the
η 1.414 1493 1822
mean value of 1187 kg/m3. The scoria from LBT has a more variable Bulk density (kg/m 3) 1413 820 807
density, between 423 and 2203 kg/m3, with a mean value of 995 kg/
CO

Column height (km) 8.22 7.49 19.30


m3. Density of BT scoria samples range from 866 to 2188 kg/m3, with TME 2.47E+11 1.24E+10 6.67E+10
a mean value of 865 kg/m3. The lava samples also show a wide range TMW 1.79E+11 1.23E+10 6.43E+10
mean MDR (kg/s) 1.15E+06 1.05E+06 6.68E+06
density, from 1996 to 2953 kg/m3, which mostly depends on the vesicle mean DREV (m 3) 0.078 0.004 0.023
fraction. Thus, an average density for lava is 2351 kg/m3. Vmagma (m 3) 0.074 0.004 0.022
Duration (h) 51.5 3 3
4.6. Distribution, volume and eruption source parameters Mean I 9 9 10
a
Mean M 4.33 3.09 3.82–4.12
UN

VEI 3 3 3 to 4
Three isopach maps were drawn based on field observations. PT de-
posit have been mapped to the east through 23 km from the vent and a Considerate the mass of lava flows.
is dispersed to the southeast up to 10 km downwind, with a shift to
the east in distal areas (Fig. 10A). LBT deposit is thinner and its dis-
persal axis has southeast direction (Fig. 10B), and BT deposit is well persed to the northeast (Fig. 10C). The integration all isopachs was
constrained at a distance of ~3 to ~8 km from the crater and is dis done following the exponential thinning method (Pyle, 1989, 1995),
with two (when available) straight-line segment, and by the Weibull
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 17

method (Bonadonna and Costa, 2012). Thus, the bulk volumes According to Westgate and Gorton (1981), the attributes of tephra
(Table 6) are 0.127–0.175 km3 for PT, about 0.015 for LBT km3 and beds that are readily discernible in the field (such as color, degree
0.080–0.083 km3 for BT. Using the bulk density of those deposits re- of weathering, thickness, upper/lower contacts, texture, lithic/juvenile
ported in Section 4.5, it was possible to obtain their total masses componentry, sedimentary structures, distribution and stratigraphic as-
for each using integration method (Table 6), which are constrained sociation) may provide sufficient control for identification and correla-
to a range of 1010–1011 kg. The column heights were interpreted from tion in areas close to source vents, but detailed petrographic and petro-

F
isopachs maps (Supplementary Fig. 2), following the methodology of chemical studies are usually required to identify distal tephra units en-
Carey and Sparks (1986), and yielded mean values of 7.49, 8.22 closed within non-volcanic sedimentary sequences. Generally, tephra
and 19.30 for each LBT, PT and BT eruptions. The mass discharge rate within our record were mostly correlated by their lithostratigraphic at-

OO
(MDR) values were estimated from the mean total masses obtained tributes. In the case of PT deposit, this unit was identified by a de-
from both the methods, using the equations of Sparks et al. (1997a, termined arrangement of tephra falls with contrasting compositions
1997b) for the columns <15 km high and Mastin et al. (2009) for the whereas EPT unit was correlated using bulk rock composition of pumice
columns >15 km high, following the observations made by Romero et fragments in the surroundings of Antuco, as seen in Fig. 9. However,
al. (2013) for southern Andes eruptions using both the equations. In all distal correlations may be done to EPT considering the thick proximal
the cases, these eruptions are constrained to MER values around 106 kg/ outcrops it has (it is reasonable to figure that the thickness of this de-
s, which allows the calculation of eruption durations, which are 51.5 h posit decays towards the southeast into Argentina). According to per-
sonal communications with Dr. Gustavo Villarosa, studying the sediment

PR
for PT and 3 h for LBT and BT. Using the equations of Pyle (2000),
the intensities (I) and magnitudes (M) for these eruptions are 9–10 and cores from Lake Portezuelo (62 km south-southeast from Antuco, Ar-
3.09–4.33 respectively, and VEI (Newhall and Self, 1982) values from gentina), they have found a tephra fallout layer with andesite to dacite
glass composition very similar to that observed for EPT tephra, but with
3 to 4 are assigned using the column heights and erupted volumes. Ac-
lower amounts of alkali (Fig. 9A). This composition is also different
cording to the observations made during the present study, BT unit was
from that of glass shards of other volcanoes in the area (e.g., Copahue
deposited during the 1852–53 AD eruption, thus some rough size esti-
and Callaqui). It is mostly composed of highly vesicular blocky pumice
mates on the lava flow of that eruption (at least 7.5 km2 and a mean
with color ranging from clear to dark brown, but also denser blocky sco-
thickness of ~3.75 m) yields a volume of c. 0.028 km3, which is equiv-

D
ria and black glass shards, similar to that seen in EPT tephra. This de-
alent to 6.61 × 108 kg (using density reported in Section 4.5). When
posit has an age span of 3.5–4.0 cal. kyr BP (99.7% probability), thus
lava flows are considered, M increases to 4.12 and VEI attains level 4.
well constrained to our geochronology. In fact, we point to this site as a
very likely candidate of distal record of EPT eruption.
TE
5. Discussion
Radiocarbon dating on older tephras has been difficult due to the
widespread vegetation-free surface, or scarce high Andean steppe
5.1. Tephra correlation and age constrains
around the volcano, and the inherent lack of charcoal or organic remains
within pyroclastic deposits. This condition extends to the entire histori-
Although the stratigraphy established in this work was obtained from
cal time, as Antuco means “sunny” or “water to the sun” and comes from
subaerial outcrops, an important part of the record was originally de-
Acluen, which may mean “dry and warm”, both words from the native
EC

posited in Laguna del Laja, located immediately adjacent to Volcán An-


tuco. McNamara et al. (2019) have demonstrated that the thickness Araucanos language (Petit-Breuilh, 2004). Moreover, these climate con-
and grain size of tephra fall deposits in lake cores and nearby terres- ditions have persisted since c. 1500 AD, and only small periods of wetter
trial sections are very similar; however, cores sampled close to fluvial conditions have interrupted this scenario over the last 2800 y BP (Tor-
inflows are affected by sediment deposition from the lake's catchment res et al., 2008; Urrutia et al., 2010). Consequently, it is difficult to
and they differ from primary deposits not only in their greater thickness date the older pyroclastic deposits, and only age constrains from pale-
and organic content, but also in poor sorting and lack of grading. In the osols and sediments are presented in this study. Based on distal strati-
RR

case of the stratigraphic sections presented in this study, tephra deposits graphic correlations, EPT unit must to be constrained to an age span of
still preserve their original features of grading and lateral continuity, 3 to 4 kyr BP, and tephras below it (i.e., α to ε) are older than 4 kyr. To
not showing evidence of reworking or mixing with other sediments, thus solve the difficult task of assigning a maximum age to our deposits, we
erosional processes may be considered negligible for those sections. In have used an erosive unconformity, which has been identified by Mel-
fact, the thickness is likely minimum due to compactness of tephra, as a nik et al. (2006) in the central and northern sectors of the lake, and
consequence of overlying sediment load. it is correlated to the outburst flood generated during the catastrophic
CO

The younger units studied here (BT and EAES) are stratigraphically opening of Laguna del Laja (see Moreno et al., 2018), that shortly fol-
and geochemically correlated to tephras found in lacustrine sediments lowed the sector collapse of the ancestral Volcán Antuco at (<7 ky BP).
(Torres et al., 2008; Urrutia et al., 2010). In fact, these units show Relicts of this outburst flood have been also identified to the west (Los
identical major element glass geochemistry, which is a critical parame- Cipreses Plain), and correspond to a 5 to 40 m thick erosive alluvial de-
ter to distinguish minor geochemical variations in tephras (Westgate posit lying above the volcanic debris avalanche, composed of a sandy
and Gorton, 1981); thus many tephra beds can be distinguished on the matrix and volcanic clasts (basaltic or andesitic) that reach 2 or 3 m size
basis of a few major element oxides in volcanic glass, such as FeOt and (Moreno et al., 1984). Neither the erosive unconformity nor the mega
UN

CaO (wt%) (Alloway et al., 2013). However, post-depositional chem- flood were recognized through our field observations in S23 (which is
ical instability of volcanic glass, especially basaltic, has been indicated the most complete, and older stratigraphic section deposited in lacus-
as a source of potential miscorrelation of tephras (Pollard et al., 2003; trine sediments), thus indicating that all the tephra units from α to EPT
Blockley et al., 2005). Considering this, and the composition of An- are at least, younger than this mega-flood event, which post-dates c. 7 ka
tuco eruptive products, EMPA glass geochemistry has been used here for BP.
correlation of the younger (i.e., poorly weathered) tephras. Correlations
may be benefited from a detailed geochemical study using trace element 5.2. Source of eruptive activity
geochemistry or by mineral composition data, particularly those of bi-
otites and/or Fe Ti oxides could be helpful (e.g., de Silva and Fran- Even though this study has been carried out in the vicinity of Volcán
cis, 1989; Alloway et al., 2013). Antuco, the source of the tephras has to be carefully veri
18 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

fied. There are at least three particular elements in Antuco that delin- these volcanoes over the study area during the <7.2 kyr BP period, and
eates the strategy for correlating its products to the source: (1) Antuco to conclude that all the tephras studied here have Volcán Antuco as their
has a particularly mafic and Na-enriched composition when compared genetic source.
to other basaltic volcanoes from the circum-Pacific area (Vergara et
al., 1985); (2) its products have been essentially basaltic after the LGM 5.3. Eruptive behavior of Antuco
(Martínez et al., 2018); (3) There are three nearby volcanoes within a

F
distance of 50–60 km (Nevados de Chillán, Callaqui and Copahue; Fig. Distinguishing the eruptive styles leading to the deposits studied here
1) and a few others at <200 km, that may deliver tephra to the Antuco is not straightforward, as these parental eruptions represent a contin-
area during infrequent episodes of atypical wind conditions (southerlies uum of extreme styles (Hawaiian to Plinian) rather than well-defined

OO
and northerlies). eruption styles (Parfitt and Wilson, 1994), thus classification schemes
60 km northward of Antuco, four main Holocene pumice fall de- (e.g., Walker, 1973; Cas and Wright, 1988) should be used with some
posits (H1–H4; younger than 5.9 ka BP) are recognized sourced from caution (Parfitt, 2003). However, there are a few textural, petrological
Nevados de Chillán, and were dispersed to the east and south with and dispersal characteristics that may be used to support the interpreta-
thickness rapidly decreasing with distance from the source (to 5 cm tions.
at around 10 km to the east of Volcán Viejo) with no current evi- Tephra samples mostly contain vesicular ash fragments and glass
dence of tephra deposition 70 km eastward in Argentina (Dixon et al., shards that preserve former melt interstices between adjacent bubbles
(e.g., ESES, ZT and LBT units), thus indicating the fluidal fragmenta-

PR
1999). These eruptions have been interpreted as Subplinian and Vul-
canian from their field characteristics of limited dispersal (Dixon et al., tion of a low viscosity (crystal-poor) hot basaltic magma (Cashman and
1998). Major elements from these tephras do not correspond with the Rust, 2016). Observations at Mt. Etna, Italy (Taddeucci et al., 2015)
field of Antuco's chemical compositions (Fig. 9), similar to that ob- suggest that sideromelane is a typical product of Hawaiian and Strom-
served when their lavas are compared (e.g., Déruelle and López-Es- bolian eruptions, meanwhile tachylite is a product commonly found in
cobar, 1999). Far away, at 160 km to the north, Laguna del Maule ash-rich explosions (small, impulsive explosions at intervals from min-
volcanic field has been highly active and explosive (Plinian) in the last utes to hours, simultaneously ejecting limited volumes of ash-sized par-
14 ky, but erupting exclusively rhyodacitic and rhyolitic tephras (Fier- ticles and blocky bombs). Based on micro-scale observations of the ju-

D
stein et al., 2014), and these compositions are not found in the sur- venile groundmass (i.e., microlite abundance), but also vesicle shape
roundings of Antuco. To the south, Copahue volcano (50 km southeast and particle overall morphology, we may correlate units PT, ZT and
from Antuco) was considerably explosive between 8.7 y 5.9 ka BP; how- LBT to Hawaiian or Strombolian eruption styles, while EPT and ESES
ever, it has only one pyroclastic fall deposit during that time span, well should be typical of ash-rich eruptions. However, in the current study,
TE
recognized but with proximal distribution, and most of the PDC em- we have not identified deposits containing elongated fibers (Pele's hairs)
placed to the east (Polanco, 1998). Another proximal tephra fall de- or drop-shaped lapilli (Pele's tears), which are diagnostic for Hawaiian
posit recently recognized by Jorquera et al. (2019) is composed of eruptions (Tilling et al., 2010), thus inferring that these eruptions are
about 60% of juvenile lapilli and 40% hydrothermally altered lithics. mostly Strombolian. Particularly, plate scoria has been observed in LBT
The deposit is dispersed towards the southeast and its age ranges from unit at S5, and thus suggests the development of a lava fountain con-
5.5 and 12 ka BP (C. Jorquera, pers. comm.), thus its dispersal, age trolled by a slug flow during violent-Strombolian eruptions (Ruth and
EC

and facies are not correlated with tephras recognized near Volcán An- Calder, 2014). On the other hand, the progressive decrease in the abun-
tuco. Tephra fallouts from the last 4 ka found in lacustrine cores sam- dance of aerodynamically shaped fluidal pyroclasts and an increase in
pled 13 km southeast from Copahue (G. Villarosa, pers. comm.), show the abundance of ragged scoria or pumice reflect the relative viscos-
tephra glass major element compositions clearly distinctive with com- ity of the erupted magmas (Houghton and Gonnermann, 2008), in
positions seen in Antuco glass, with the only exception of PT unit (Fig. which is observed for EPT and BT units. According to MDR rate values
9A), which is sourced from Antuco as isopachs testify (Fig. 10). Re- obtained for BT, this eruption may be classified as Sub-Plinian in the
RR

cent tephra fall deposits produced by this volcano during both 2000 and sense of Bonadonna and Costa (2013), and its magnitude is 4.12, if
2012-to-present eruptive cycles have had dispersal axes to the northeast the erupted lava is considered.
and southeast (e.g., Naranjo and Polanco, 2004; Petrinovic et al., The case of EPT is unique within the context of present
2014; Daga et al., 2016), not affecting the area of Antuco, constituting tephra-stratigraphy, as it is the thickest tephra fall unit that we have
fine-grained ash layers (not as seen in Antuco), and with an overall com- recognized, and it is fully composed of pumice. Our field observations
position enriched in K2O but lower Fe2O3/MgO ratios when compared (thickness, internal architecture and componentry) suggest a sub-Plinian
eruption style. In fact, contemporaneous eruptions with similar eruption
CO

to Antuco products (Fig. 9A and B). Callaqui volcano (60 km S from


Antuco) has experienced six eruptions between 9.95 and 0.23 ka BP; style and magma composition (i.e., basaltic-andesite Subplinian erup-
however, its most explosive period occurred between 2.63 and 2.30 ka tions) have similar characteristics of vesicularity, size and density grad-
BP (VEI 4; Polanco, 1998). These explosive deposits of Callaqui vol- ing within their tephra fall deposits (e.g., Calbuco in 2015, Romero et
cano mostly correspond to PDCs, and pumice fallouts do not exceed al., 2016). Additionally, the exceptional microlite-rich groundmass of
30 cm in thickness at proximal outcrops. Both Copahue and Callaqui EPT sample (c. 85%) has also been observed in clasts of Plinian-type
have erupted high-K andesites (Polanco, 1998), which are easily dis- eruptions from Mt. Etna, Italy 122 BC and Tarawera, New Zealand 1886
tinguishable from the Antuco tephras using major element chemistry (Sable et al., 2006). Distal glass compositions with lower amounts of
UN

(Fig. 9). Other volcanoes to the south, such as Lonquimay have pro- alkali, and wider compositions of SiO2 than those observed in bulk rock
duced tephra fall deposits during the last 4800 y BP, which have had 0.1 geochemistry (Fig. 9), may be explained as follows. Hydrous basaltic
to 0.3 km3 of non-DRE volume but are dispersed to the east (Polanco, and andesitic magmas (50–60 wt% SiO2) are especially susceptible to
1998, 2010; Bustamante, 2013), thus any possible occurrence in the ascent-driven crystallization, which creates a large range of glass com-
area of Antuco is unlikely. In addition, the Plinian deposits of Llaima positions for a narrow range in bulk composition (Cashman and Rust,
and Sollipulli volcanoes (150 and 175 km south of Antuco, respectively) 2016). Higher SiO2 amounts (i.e., dacite compositions) of the products
are unlikely to have been deposited northwards in the vicinity of An- from distal areas is also consistent the abundant glassy fragments (Cash-
tuco, since these eruptions (8.8 and 2.9 ka BP respectively) were pri- man and Rust, 2016).
marily dispersed southeast and east-northeast (Naranjo and Moreno, Although PDCs are usually related to silicic eruptions, many of
1991; Naranjo et al., 1993). This leads to ruling out any influence of them have been recently recognized at basaltic volcanoes and
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 19

they were emplaced at high temperatures and very high speed (e.g., Ya- that paleo-lake (i.e., level before 1980) was as much as 7.5 to 8.5 km.
mamoto et al., 2005; Miyabuchi et al., 2006; Amigo et al., 2013; There are really few cases present in the literature where demonstrable
Di Roberto et al., 2014; Alloway et al., 2017; Andronico et al., subaqueous pyroclastic density currents have occurred (e.g., Cas and
2018). Units α, CES, ESES and APDC consist of PDC deposits. Some of Wright, 1991; Soriano et al., 2012); however, the short distance of
their outcrops are found as far as 20.6 km east from the vent, suggesting these deposits to the source and the strong altitude difference (about
that PDCs are important phenomena in the eruptive behavior of Antuco. 1500 m) may have facilitated the long run-out of such deposits. Experi-

F
Particularly, the most distal deposits of PDCs (e.g., two massive PDCs mental observations suggest that the entrance of PDCs to the water de-
found at sites 28 and 29) may have been produced by moderate-to-large pends on the temperature of the current (“cold” currents are able to
explosive eruptions soon after the sector collapse of the PCa edifice, and displace underwater, while the hot, i.e. >250 °C, are able to transport

OO
correlations to the west flank of the volcano are highly recommended, above the water surface) rather than the density of the PDC (Freundt,
as several PDCs with similar characteristics have been found there (e.g., 2003). PDCs entering the water may develop indistinguishable facies,
Thiele et al., 1998; Moreno et al., 2000, 2018). when compared to subaerial deposits, and may not show any water cool-
ESES unit has typical lithofacies of both cross-stratified and massive ing or mixing (Mandeville et al., 1996). When hot PDCs entrain water
ignimbrite deposits (see Brown and Andrews, 2015). Sub-units identi- in subaqueous environment, they cause fluidization in underlying sedi-
fied within ESES should be interpreted as different flow units of a single ments, which results in the formation of large-scale load casts at the base
vent-derived current, initiated by a discrete eruption, collapse pulse or of the flows (Kokelaar, 1982; Howells et al., 1985). Similar soft-sed-
major fluctuation in during a sustained eruption interval (Wilson and iment deformation structures have been observed in the silt middle of

PR
Hildreth, 1997; Hildreth and Fierstein, 2012). In fact, the overall two PDC units of ESES at S13, and their size range from 5 to 7 cm in
unit stratigraphy is defined by a basal scoria fallout, a cross-stratified ig- their amplitude (Fig. 11A). However, their origin is not clear as they
nimbrite (PDC 1; Fig. 11), a period of short-lived quiescence with depo- may also constitute seismites. Juvenile components at the matrix of the
sition of ash interbedded with silt, and finally, two subsequent massive S13 deposit are fresh in texture, without any sign of rework, thus is
ignimbrites on top (PDCs 2 and 3; Fig. 11). Particularly, the existence of interpreted as a primary depositional features. In fact, the features ob-
many clasts supported in the ash matrix of the upper PDC unit (3; Fig. served in S13, and their contrast to the rest of the sites, may indicate an
11A and B) may be interpreted as cannibalisation of previous deposits, underwater deposition environment, but more detailed study is required

D
which is a typical feature of PDCs that may erode the substrate, pulling to validate this hypothesis.
up accidental lithics and transporting them within the current (Sparks Most of historical eruptions with VEI < 2 are not present in the
et al., 1997a, 1997b; Calder et al., 2000; Roche, 2015). stratigraphic records, probably due to their small volume, restricted dis-
The ESES unit is interbedded between lacustrine sediments at S6, persal and rapid erosion, even when cold-dry climate conditions, likely
TE
S12, S13 and S23, thus indicating i) a subaqueous emplacement or ii) a favorable for tephra preservation at Antuco were almost stable between
changing level of the lake, that allows silt deposition bellow and above 1500 and 1900 CE in lake Laja, which roughly coincides with the Eu-
the volcanic deposits at different times. The distance of the vent to ropean Little Ice Age (Urrutia et al., 2010). We suggest that all the
identified deposits had sufficient size (in terms of their volume, col
EC
RR
CO
UN

Fig. 11. Brittle and ductile deformation of soft sediments in the surrounding area of Volcán Antuco. A and B: Architecture of unit ESES at S13, with bomb sags (BS) and soft-sediment
clasts within different PDC units. Tephra fall deposit is indicated with TF. The white dotted line marks the limit between the upper PDC (3) and a matrix supported PDC that should
be considered that a top sub-unit of the latter. C and D: Seismites in S10 and S27. E and F: Normal faults (i.e., graben structures) in S18 and 200 m southwest of S29. In E, faults are
affecting tephra deposits and silt layers, while in F faults affect Pleistocene to Holocene volcanic conglomerates interbedded by medium to coarse grained sandstones. Insert F is courtesy
of Amapola Albornoz.
20 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

umn height and intensity) to be preserved efficiently at distances >4 km tion of PDCs to the north-northeastern flanks of Antuco, where the lake
within the volcaniclastic basins east of Antuco, thus probably VEI > 2. is situated.
According to our dating, 15 units are constrained to the last 3 kyr At least three soft-sediment deformed units (Fig. 11C and D), and
(65% of the field-recognized tephras), implying a mean repose time of a normal fault (Fig. 11E) have been identified in the stratigraphic sec-
200 years. However, this repose time is subject to biases, mainly re- tions presented here (Fig. 3). These units fit well with the six field crite-
garding erosion and dispersal trends not studied within this investiga- ria for their identification as seismites (Melnik et al., 2006, and refer-

F
tion thus must be considered as maximum. For example, historical erup- ences therein): (1) they correspond to poorly consolidated sediments; (2)
tions from Antuco show two main repose times: 1) a 4.5 to 5.0 year a cyclic repetition of similar structures is observed; (3) the places where
repose time related to high eruption rates, and 2) repose from 20.3 to they are identified correspond to flat areas not influenced by landslides;

OO
63.3 years at low eruption rates (Supplementary Fig. 3A and B). In the (4) their position is enveloped by undeformed beds; (5) there is a lat-
second case, the repose time has been largely exceeded since the last eral continuity and regional abundance and; (6) they share similarities
eruption in (80 years). Considering these limitations, it is better to esti- to structures reported elsewhere. In fact, observed structures correspond
mate the total frequency of each type of eruption style in terms of the to recumbent folds in lacustrine varved silt, up to 20 cm in amplitude,
total number of eruptions, rather than a time-dependent periodicity. In limited by sand to the base and top (Fig. 11C). These can be classified as
fact, from a total number of 19 tephra units, 14 represent tephra fall seismo-slumps, which are common in lacustrine chalky deposits which
episodes (c. 77.8%) and 4 are PDCs (22.2%), which represent at least display regular and delicate varved lamination, and are caused by liq-
23 individual eruptions, as some of those units are composed of sev- uefaction (Montenat et al., 2007 and references therein). Melnik et

PR
eral individual deposits. According to the criteria applied for identifying al. (2006) described the Laguna del Laja Fault System (LLFS) as a series
eruptive styles, we can count 15 eruptions with a Strombolian to Violent of fragile and ductile deformation structures grouped in three segments,
Strombolian behavior (78.9%); 3 Sub-Plinian (15.8%) and just 1 Plinian from which the northern segments cumulates a meter-scale slip faulting,
eruption (5.26%). The inter-relationship between the repose times and while the southern segment has a poorly defined fault geometry masked
the eruption style should be assessed in further investigations. by the Antuco volcanics and integrates seismites, such as the ones iden-
Sieve texture has been attributed to partial dissolution of Na-plagio- tified in this work. However, the normal faults here described are not
clase reacting to a calcic melt due to magma mingling (Tepley III et al., part of the LLFS, as they are associated to a NE-SW trending normal fault

D
1999), to the mixing between mafic and felsic magmas (Eichelberger, system assumed as the eastern part of two hemi-graben structures (Al-
1980) or due to decompression of water-saturated magmas that would bornoz, 2019), which also integrates the LLFS to the west. These faults
crystallize plagioclase as volatiles are exsolved, while resorption is pos- are sub-parallel to the NE-SW alignment of volcanic structures and dykes
sible during ascent prior to water saturation (Blundy and Cashman, of the Antuco-Sierra Velluda volcanoes, and are interpreted as an exten-
TE
2001). Petrologic studies on these disequilibrium textures in lavas from sional system that feeds the magmatic activity (Albornoz, 2019). Despite
Antuco suggest processes of magmatic differentiation within an upper this, research carried out in similar volcano-tectonic backgrounds (e.g.,
magmatic reservoir (5 km depth) continuously affected by the recharge Tongariro Graben, New Zealand; Gómez-Vasconcelos et al., 2017)
of deep, mafic magmas, which cause depressurization and convection suggest that careful attention should be paid to confirm the origin of the
(Norambuena, 2016). Although detailed studies are required to better faulting, whether tectonic, magmatic or volcano-tectonic, as this may af-
understand the reservoir dynamics, it can be suggested that primitive fect seismic (and volcanic) hazard assessments.
EC

magma intrusions in the reservoir of Antuco may trigger sequences of The structures recognized in the current study are constrained to
explosive and effusive eruptions associated with these textures, which the last 3 kyr BP, nevertheless a precise age is not established. De-
are constantly observed in ZT, LBT and BT tephras. Similar eruption spite this, a minimum recurrence of one earthquake every one thou-
mechanisms have been proposed for the explosive eruptions of other sand years can be estimated, considering the fact that earthquake of
basaltic volcanoes (e.g., Bertagnini et al., 2003; Landi et al., 2004; magnitudes >Mw 5 are assumed necessary for triggering liquefaction
Romero et al., 2018). of varved lamination sequences (Marco and Agnon, 1995), while at
RR

magnitudes >5.5 these soft sediments turn into brittle deformation (Ro-
5.4. Volcanic and seismic hazards driguez-Pascua et al., 2000) such as one seen in Fig. 11D. This im-
plies that the study area is subject to shallow seismogenic processes gen-
The main volcanic hazards related to the explosive activity of An- erating >Mw 5 earthquakes, which may produce liquefaction. In ad-
tuco are the generation of both PDCs and tephra fallouts. Most of ex- dition, a relationship between local Mw ≥ 4.8 earthquakes and VEI ≥ 0
plosive eruptions identified here show textural and compositional fea- eruptions have been statistically documented to occur above the back-
tures consistent with Strombolian eruptions. This is in agreement with ground noise 6–10 days before and after the earthquake (see details in
CO

previous works in the study area (Lohmar, 2000; Moreno et al., Lemarchand and Grasso, 2007). Thus, local earthquakes may be cor-
2000; Moreno, 2016); however, 3 VEI 3–4 eruptions (and a probably related to volcanic activity in the area of Antuco. This seismicity must
VEI > 4 eruption older than 3 kyr) have been identified as the most ex- be also taken into account for the stability of Antuco edifice, and its re-
plosive events in the last 3 kyr. Such eruptions are an actual threat for lation to the sector collapse at about 7 ka BP must be studied in detail.
critical infrastructure, such as an international road and hydroelectric
power stations, in addition to touristic facilities. This type of eruptive 6. Conclusions
heterogeneity has been previously observed in other mafic volcanoes
UN

of the southern Andes (e.g., Naranjo and Moreno, 1991; Polanco, Tephrostratigraphic records found east of Antuco account for at
1998; Naranjo and Haller, 2002; Amigo et al., 2013; Bustamante, least 23 explosive eruptions in the last <7.2 kyr BP. Any likely influ-
2013; Rawson et al., 2015; Alloway et al., 2017; Naranjo et al., ence of other eruptive source(s) within a distance of 200 km south and
2017; Romero et al., 2018) and demands the attention of the risk north from Antuco has been properly evaluated using criteria, such as
management agencies. In addition, theoretical investigations points to age, dispersal and major element geochemistry of these tephras, and
PDCs as the main cause of tsunami generation by pyroclastic flows suggest that these deposits likely belong to Antuco. Even though we
(Watts and Waythomas, 2003), and this consideration must be taken have not obtained any absolute age for deposits older than 4 kyr (ob-
into account for future eruptions, especially under the possible genera served in Section S23) studied in this contribution, we infer they are
younger than 7.2 ka BP, as we have not recognized a local erosive un
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 21

conformity related to the catastrophic flooding of Laguna del Laja after Drs. Tatiana Izquierdo, Manuel Abad, Hugo Moreno, Silke Lohmar,
the sector collapse of La edifice, which exists below the lacustrine sedi- Gabriel Orozco, Daniel Melnick, Gustavo Villarosa and Edmundo
ments studied here and has the maximum age of 7.2 ka BP. Thus, all of Polanco significantly contributed to some interpretations. In memory of
our stratigraphy is constrained to the post-collapse history (PCa stage) 44 young conscripts and a sergeant, who lost their lives on May 18, 2005
of Antuco. due to a sudden violent snow-storm during an irresponsible military ex-
Textural and compositional analyses of tephra deposits recognized ercise on the lower flanks of Volcán Antuco.

F
at Antuco reveal mostly Strombolian, ash-rich and violent-Strombolian
styles; however, BT unit (1852–53 CE) may be classified as sub-Plin- References
ian, while EPT, LBT and PT can be categorized as VEI 3 (violent Strom-

OO
bolian?). EPT unit appears to show distinct architectural and composi- Alloway, B.V., Lowe, D.J., Larsen, G., Shane, P.A.R., Westgate, J.A., 2013. Quaternary
stratigraphy| tephrochronology. In: Elias, S.A., Mock, C.J. (Eds.), Encyclopedia of Qua-
tional similarities with Plinian eruptions, moreover it has a wide dis- ternary Science, 2da versión. Elsevier, Edimburgo, Escocia, pp. 277–304.
persal and tephrostratigraphic correlations are possible to distal areas in Alloway, B.V., Moreno, P.I., Pearce, N.J., De Pol-Holz, R., Henríquez, W.I., Pesce, O.H.,
Argentina. The wide spectrum of pyroclastic deposits found in Volcán Sagredo, E., Villarrosa, G., Outes, V., 2017. Stratigraphy, age and correlation of Lepué
Tephra: a widespread c. 11 000 cal a BP marker horizon sourced from the Chaitén
Antuco indicates that the composition, style and magnitude may change Sector of southern Chile. J. Quat. Sci. 32 (6), 795–829.
from one eruptive episode to the next. This eruptive variability seems in Amigo, Á., Lara, L.E., Smith, V.C., 2013. Holocene record of large explosive eruptions from
complete synergy with recent findings from other centers in the South- Chaitén and Michinmahuida Volcanoes, Chile. Andean Geol. 40 (2), 227–248.
Andronico, D., Scollo, S., Cristaldi, A., Lo Castro, M.D., 2014. Representivity of incom-
ern Volcanic Zone exhibiting similar temporal eruptive diversity and ul-

PR
pletely sampled fall deposits in estimating eruption source parameters: a test using the
timately, has significant implications with respect to risk assessment and 12-13 January 2011 lava fountaining deposits from Mt. Etna volcano, Italy. Bull. Vol-
characterization of potential hazards. On the other hand, likely Plinian canol. 76, 861.
Andronico, D., Di Roberto, A., De Beni, E., Behncke, B., Bertagnini, A., Del Carlo, P., Pom-
deposits, such as EPT unit, requires further studies in order to better con-
pilio, M., 2018. Pyroclastic density currents at Etna volcano, Italy: the 11 February
strain the volcanic hazards of exceptionally explosive eruptions at An- 2014 case study. J. Volcanol. Geotherm. Res. 357, 92–105.
tuco. Arzilli, F., La Spina, G., Burton, M.R., Polacci, M., Le Gall, N., Hartley, M.E., Di Genova,
D., Cai, B., Vo, N.T., Bamber, E., Nonni, S., Atwood, R., Llewellin, E.W., Brooker,
Tephra fallouts during moderate-to-large sized eruptions in future
R.A., Mader, H.M., Lee, P., 2019. Magma fragmentation in highly explosive basaltic
may cause relevant impacts on critical infrastructure along the interna- eruptions induced by rapid crystallization. Nat. Geosci. 1–6. 10 p. doi:10.1038/

D
tional route, hydroelectric power generation and touristic facilities. Ad- s41561-019-0468-6.
ditionally, at least four PDCs have been recognized, thus pointing to haz- Barberi, F., Coltelli, M., Frullani, A., Rosi, M., Almeida, E., 1995. Chronology and dispersal
characteristics of recently (last 5000 years) erupted tephra of Cotopaxi (Ecuador): im-
ardous phenomena that may severely affect these infrastructures, and plications for long term eruptive forecasting. J. Volcanol. Geotherm. Res. 69, 217–239.
also produce local tsunamis if they enter the lake, as evidenced from the Barsotti, S., Neri, A., Bertagnini, A., Cioni, R., Mulas, M., Mundula, F., 2015. Dynamics
TE
ESES eruption record. On the other hand, based on soft-sediment defor- and tephra dispersal of Violent Strombolian eruptions at Vesuvius: insights from field
data, wind reconstruction and numerical simulation of the 1906 event. Bull. Volcanol.
mation structures and fault planes identified in the stratigraphic record doi:10.1007/s00445-015-0939-6.
of the present work, it is suggested that the area is prone to be affected Bertagnini, A., Métrich, N., Landi, P., Rosi, M., 2003. Stromboli volcano (Aeolian Archi-
by >Mw5 earthquakes, which may have a return period of the order pelago, Italy): an open window on the deep-feeding system of a steady state basaltic
volcano. J. Geophys. Res. 108 (B7), 2336. doi:10.1029/2002JB002146.
of few thousands of years. These earthquakes may have very shallow Blackburn, E.A., Wilson, L., Sparks, R.S.J., 1976. Mechanisms and dynamics of Strombo-
sources, as they may be triggered by shallow ruptures of the LLFS or the
EC

lian activity. J. Geol. Soc. Lond. 132, 429–440.


volcano-tectonic dynamics of Volcán Antuco. Blockley, S.P.E., Pyne-O’Donnell, S.D.F., Lowe, J.J., Matthews, I.P., Stone, A., Pollard,
A.M., Turney, C.S.M., Molyneux, E.G., 2005. A new and less destructive laboratory
Supplementary data to this article can be found online at https://doi. procedure for the physical separation of distal glass tephra shards from sediments.
org/10.1016/j.jvolgeores.2019.106759. Quat. Sci. Rev. 24 (16–17), 1952–1960.
Blundy, J., Cashman, K., 2001. Ascent driven crystallization of dacite magmas at Mount
St. Helens, 1980-1986. Contributions to Mineral Petrology 140, 631–650.
CRediT authorship contribution statement
Bonaccorso, A., Calvari, S., Linde, A., Sacks, S., 2014. Eruptive processes leading to the
most explosive lava fountain at Etna volcano: the 23 November 2013 episode. Geo-
RR

Jorge E. Romero: Visualization, Investigation, Data curation, Writ- phys. Res. Lett. 41 (14), 4912–4919.
Bonadonna, C., Costa, A., 2012. Estimating the volume of tephra deposits: a new simple
ing - original draft. Victoria Ramírez: Investigation. Mohammad Ayaz
strategy. Geology 40, 415–418.
Alam: Supervision. Jorge Bustillos: Investigation. Alicia Guevara: For- Bonadonna, C., Costa, A., 2013. Plume height, volume, and classification of explosive vol-
mal analysis. Alessandro Pisello: Investigation. Daniele Morgavi: canic eruptions based on the Weibull function. Bull. Volcanol. 75 (8), 742.
Writing - review & editing. Evelyn Criollo: Formal analysis. Brown, R.J., Andrews, G.D., 2015. Deposits of pyroclastic density currents. In: The Ency-
clopedia of Volcanoes. Academic Press, pp. 631–648.
Bustamante, Ó., 2013. Dispersión de tefra de erupciones explosivas holocenas del Complejo
Uncited reference
CO

Volcánico Lonquimay, Región de la Araucanía, Chile. Memoria para optar al Título de


Geólogo. Departamento de Geología, Universidad de Chile. (128 p).
Calder, E.S., Sparks, R.S.J., Gardeweg, M.C., 2000. Erosion, transport and segregation of
Romero et al., 2014 pumice and lithic clasts in pyroclastic flows inferred from ignimbrite at Lascar Vol-
cano, Chile. J. Volcanol. Geotherm. Res. 104 (1–4), 201–235.
Acknowledgements Carey, S., Sparks, R.S.J., 1986. Quantitative models of the fallout and dispersal of tephra
from volcanic eruption columns. Bull. Volcanol. 48 (2–3), 109–125. doi:10.1007/
BF01046546.
We thank to Dr. Brent V. Alloway for his detailed review and lan- Cas, R.A.F., Wright, J.V., 1987. Volcanic Successions: Ancient and Modern. Allen and
guage assistance in the proof version of this manuscript. Authors are Unin, London. (528 p).
UN

Cas, R.A.F., Wright, J.V., 1991. Subaqueous pyroclastic flows and ignimbrites: an assess-
grateful to Mr. Hector Moyano, Mr. Luis Cortes and Dr. Juan Díaz-Al-
ment. Bull. Volcanol. 53, 357–380.
varado for their assistance and participation in the fieldwork. Dr. Ur- Cashman, K., Rust, A.C., 2016. Volcanic ash: generation and spatial variations. In: Mackie,
rutia thanks Conicyt/Fondap/15130015 research funding. The staff of S., Cashman, K., Ricketts, H., Rust, A., Watson, M. (Eds.), Volcanic Ash: Hazard Ob-
servation. Elsevier, pp. 5–22.
Chile's National Forest Corporation (CONAF), Ana Hinojosa, Miguel In-
Chouet, B., Hamisevicz, N., McGetchin, T.R., 1974. Photoballistics of volcanic jet activity
fante, Segundo Necul and Juan Bascur, provided excellent support for at Stromboli, Italy. J. Geophys. Res. 79 (32), 4961–4976.
the fieldwork within the Laguna del Laja National Park. We thank the Costantini, L., Bonadonna, C., Houghton, B.F., Wehrmann, H., 2009. New physical charac-
collaboration of the Chilean Army (Ejército de Chile) and Chilean Po- terization of the Fontana Lapilli basaltic Plinian eruption, Nicaragua. Bull. Volcanol.
71, 337–355.
lice (Carabineros de Chile) to make the access possible to many of Costantini, L., Pioli, L., Bonadonna, C., Clavero, J., Longchamp, C., 2011. A late Holocene
the remote areas and to provide accommodation facilities. Comments explosive mafic eruption of Villarrica volcano, Southern Andes: the Chaimilla deposit.
made by Drs. Silvina Guzmán and Brian Jicha, plus the editorial han- J. Volcanol. Geotherm. Res. 200, 143–158.
Daga, R., Caselli, A., Ribeiro Guevara, S., Agusto, M., 2016. Tefras emitidas durante la fase
dling of Dr. J.L. Macías, greatly improved the quality of this manuscript. inicial hidromagmática (julio de 2012) del ciclo eruptivo 2012-actual (2016) del vol-
This work is dedicated to Mr. Jesús López. Discussions and support of cán Copahue (Andes del sur). Rev. Asoc. Geol. Argent. 74 (2), 191–206.
De Silva, S.L., Francis, P.W., 1989. Correlation of large ignimbrites—two case studies from
the Central Andes of Northern Chile. J. Volcanol. Geotherm. Res. 37 (2), 133–149.
22 J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759

Déruelle, B., López-Escobar, L., 1999. Basaltes, andésites, dacites et rhyolites des strato- sign realistic source parameters to models of volcanic ash-cloud transport and dispersion
volcans des Nevados de Chilian et de l’Antuco (Andes méridionales): Ia remarquable during eruptions. J. Volcanol. Geotherm. Res. 186, 10–21.
illustration d’une différenciation par cristallisation fractionnée. Comptes Rendus de McGetchin, T.R., Settle, M., Chouet, B.A., 1974. Cinder cone growth modeled after North-
l’Académie des Sciences-Series IIA-Earth and Planetary Science 329 (5), 337–344. east Crater, Mount Etna, Sicily. J. Geophys. Res. 79, 3257–3272.
Di Roberto, A., Bertagnini, A., Pompilio, M., Bisson, M., 2014. Pyroclastic density currents McNamara, K., Rust, A.C., Cashman, K.V., Castruccio, A., Abarzúa, A.M., 2019. Compari-
at Stromboli volcano (Aeolian Islands, Italy): a case study of the 1930 eruption. Bull. son of lake and land tephra records from the 2015 eruption of Calbuco volcano, Chile.
Volcanol. 76 (6), 1–14. doi:10.1007/s00445–014–0827–5. Bull. Volcanol. 81 (2), 10.
Dixon, H.J., Murphy, M.D., Sparks, S.J., Chávez, R., Naranjo, J.A., Dunkley, P.N., Young, Melnik, D., Charlet, F., Echtler, H., De Baist, M., 2006. Incipient axial collapse of the Main

F
S.R., Gilbert, J.S., Pringle, M.R., 1999. The geology of Nevados de Chillán volcano, Cordillera and strain partitioning gradient between the central and Patagonian Andes,
Chile. Revista geológica de Chile 26 (2), 227–253. Lago Laja, Chile. Tectonics 25, TC5004. doi:10.1029/2005TC001918.
Dominguez, L., Pioli, L., Bonadonna, C., Connor, C.B., Andronico, D., Harris, A.J.L., Miyabuchi, Y., Watanabe, K., Egawa, Y., 2006. Bomb–rich basaltic pyroclastic flow deposit
Ripepe, M., 2016. Quantifying unsteadiness and dynamics of pulsatory volcanic activ- from Nakadake, Aso Volcano, southwestern Japan. J. Volcanol. Geotherm. Res. 155

OO
ity. Earth Planet. Sci. Lett. 444, 160–168. (1–2), 90–103. doi:10.1016/j.jvolgeores.2006.02.007.
Eichelberger, J.C., 1980. Vesiculation of mafic magma during replenishment of silicic Montenat, C., Barrier, P., Hibsch, C., 2007. Seismites: an attempt at critical analysis and
magma reservoir. Nature 288, 446–450. classification. Sediment. Geol. 196 (1–4), 5–30.
Fierstein, J., Sruoga, P., Amigo, A., Elissondo, M., Rosas, M., 2014. Tephra in Argentina Moreno, 2016. Peligros del volcán Antuco, región del Biobío. In: Servicio Nacional de Ge-
establishes postglacial eruptive history of Laguna del Maule volcanic field in Chile: 36 ología y Minería, Carta Geológica de Chile, Serie Geología Ambiental, 27. 1 mapa a
silicic eruptions in 14 Ka. In: Actas del XIX Congreso Geológico Argentino Volcanes escala 1:50.000, Santiago, Chile.
Activos, Córdoba. (pp. 23–17). Moreno, H., Varela, J., Lahsen, A., Vergara, M., 1984. Estudio Geológico del Grupo Vol-
Folk, R.L., Ward, W.C., 1957. Brazos River bar [Texas]; a study in the significance of grain cánico Antuco-Sierra Velluda. Departamento de Geología y Geofísica. Universidad de
size parameters. J. Sediment. Res. 27 (1), 3–26. Chile, p. 164. (Contrato OICB-03. ENDESA. Informe Inédito).
Francis, P., Oppenheimer, C., 2004. Volcanoes. 2nd ed. Oxford University Press, New York. Moreno, H., Lohmar, S., López-Escobar, L., Petit-Breuilh, M.E., 2000. Contribución a la

PR
Freundt, A., 2003. Entrance of hot pyroclastic flows into the sea: experimental observa- evolución geológica, geoquímica e impacto ambiental del Volcán Antuco (Andes del
tions. Bull. Volcanol. 65 (2–3), 144–164. Sur, 37°25’ S). In IX Congreso Geológico Chileno. Soc. Geol. de Chile, Santiago, Chile.
Gómez-Vasconcelos, M.G., Villamor, P., Cronin, S., Procter, J., Palmer, A., Townsend, D., Moreno, H., Mella, M., Jara, C., 2018. The most catastrophic Late Holocene geologic
Leonard, G., 2017. Crustal extension in the Tongariro graben, New Zealand: insights process in Central-South Chile: the Antuco volcanic debris avalanche-induced Laja
into volcano-tectonic interactions and active deformation in a young continental rift. Lake flooding, 37.4°S. In: XV Congreso Geológico Chileno. Universidad de Concepción,
GSA Bull. 129 (9–10), 1085–1099. Concepción, Chile.
González-Ferrán, O., 1995. Volcanes de Chile. Centro de Investigaciones Volcanológicas, Naranjo, J.A., Haller, M.J., 2002. Erupciones holocenas principalmente explosivas del vol-
Santiago. cán Planchón, Andes del sur (35 15’S). Revista geológica de Chile 29 (1), 93–113.
Heiken, G., 1978. Characteristics of tephra from Cinder Cone, Lassen Volcanic National Naranjo, J.A., Moreno, H., 1991. Actividad explosiva postglacial en el Volcán Llaima, An-
Park, California. Bull. Volcanol. 41 (2), 119–130. des del Sur (38 45’S). Revista Geológica de Chile 18 (1), 69–80.

D
Hickey-Vargas, R., Holbik, S., Tormey, D., Frey, F.A., Roa, H.M., 2016. Basaltic rocks from Naranjo, J.A., Polanco, E., 2004. The 2000 AD eruption of Copahue volcano, southern An-
the Andean Southern Volcanic Zone: insights from the comparison of along-strike des. Revista geológica de Chile 31 (2), 279–292.
and small-scale geochemical variations and their sources. Lithos 258, 115–132. Naranjo, J.A., Moreno, H., Gardeweg, M., 1991. Erupción de 1989-1990 del volcán Lon-
doi:10.1016/j.lithos.2016.04.014. quimay, Andes del sur (38°20’S). In: Congreso Geológico Chileno, 1. pp. 445–448. No.
6, Actas. (Viña del Mar).
TE
Hildreth, W., Fierstein, J., 2012. The Novarupta-Katmai eruption of 1912: largest erup-
tion of the twentieth century: centennial perspectives. In: US Geological Survey Pro- Naranjo, J.A., Moreno, H., Emparan, C., Murphy, M., 1993. Volcanismo explosivo reciente
fessional Paper No. 1791. (259 p). en la caldera del volcán Sollipulli, Andes del Sur (39 S). Revista Geológica de Chile 20
Hildreth, W., Moorbath, S., 1988. Crustal contributions to arc magmatism in the Andes of (2), 167–191.
central Chile. Contrib. Mineral. Petrol. 98, 455–489. doi:10.1007/BF00372365. Naranjo, J.A., Singer, B.S., Jicha, B.R., Moreno, H., Lara, L.E., 2017. Holocene tephra
Hogg, A.G., Hua, Q., Blackwell, P.G., Niu, M., Buck, C.E., Guilderson, T.P., Heaton, T.J., succession of Puyehue-Cordón Caulle and Antillanca/Casablanca volcanic complexes,
Palmer, J.G., Reimer, P.J., Reimer, R.W., Turney, C.S.M., Zimmerman, S.R.H., 2013. southern Andes (40–41 S). J. Volcanol. Geotherm. Res. 332, 109–128.
SHCal13 southern hemisphere calibration, 0–50,000 years cal BP. Radiocarbon 55 (4), Newhall, C., Self, S., 1982. The volcanic explosivity index (VEI): an estimate of explosive
EC

1889–1903. magnitude for historical volcanism. J. Geophys. Res. 87, 1231–1238.


Houghton, B.F., Carey, R.J., 2015. Pyroclastic fall deposits. In: The Encyclopedia of Volca- Parfitt, E., 2003. A discussion of the mechanisms of explosive volcanic eruptions. J. Vol-
noes. Academic Press, pp. 599–616. canol. Geophys. Res. 134, 77–107.
Houghton, B.F., Gonnermann, H.M., 2008. Basaltic explosive volcanism: constraints from Parfitt, E.A., Wilson, L., 1994. The 1983 – 86 Pu’ u ‘O’o eruption of Kilauea volcano,
deposits and models. Chemie der Erde-Geochemistry 68 (2), 117–140. Hawaii: a study of dike geometry and eruption mechanisms for a long-lived eruption.
Houghton, B.F., Wilson, C.J.N., 1989. A vesicularity index for pyroclastic deposits. Bull. J. Volcanol. Geotherm. Res. 59, 179–205.
Volcanol. 51, 451–462. Patrick, M.R., Harris, A.J.L., Ripepe, M., Dehn, J., Rothery, D., Calvari, S., 2007. Strombo-
Howells, M.F., Campbell, S.D.G., Reedman, A.J., 1985. Isolated pods of subaqueous lian explosive styles and source conditions: insights from thermal (FLIR) video. Bull.
RR

welded ash-flow tuff: a distal facies of the Capel Curig Volcanic Formation (Ordovi- Volcanol. 69, 769–784.
cian) North Wales. Geol. Mag. 122, 175 180. Penna, I.M., Hermanns, R.L., Niedermann, S., Folguera, A., 2011. Multiple slope failures
Jorquera, C., Flores, F., Amigo, A., 2019. Nuevos antecedentes sobre depósitos de caída associated with neotectonic activity in the Southern Central Andes (37–37 30′ S),
en Alto Biobío, Chile: ¿Un inédito evento explosivo en el registro eruptivo del vol- Patagonia, Argentina. Bulletin 123 (9–10), 1880–1895.
cán Copahue? In: In 1er Congreso de la Asociación Latinoamericana de Volcanología, Petit-Breuilh, M.E., 1994. Actividad volcánica y cronología eruptiva histórica del volcán
Antofagasta, Chile. Antuco (37240S–71220W), Chile. Rev. Geogr. Chile Terra Aust. 39, 79–102.
Kokelaar, B.P., 1982. Fluidisation of wet sediments during the emplacement and cooling Petrinovic, I.A., Villarosa, G., D’elia, L., Guzman, S.R., Paez, G.N., Outes, A.V., Manzoni,
of various igneous bodies. J. Geol. Soc. Lond. 139, 21 33. C., Delménico, A., Balbis, C., Carniel, R., Hernando, I.R., 2014. La erupción del 22 de
Landi, P., Metrich, N., Bertagnini, A., Rosi, M., 2004. Dynamics of magma mixing and diciembre de 2012 del volcán Copahue, Neuquén, Argentina: caracterización del ciclo
CO

degassing recorded in plagioclase at Stromboli (Aeolian Archipelago, Italy). Contrib. eruptivo y sus productos.
Mineral. Petrol. 147 (2), 213–227. Pioli, L., Erlund, E., Johnson, E., Cashman, K., Wallace, P., Rosi, M., Delgado Granados, H.,
Lautze, N.C., Taddeucci, J., Andronico, D., Cannata, C., Tornetta, L., Scarlato, P., 2008. Explosive dynamics of violent Strombolian eruptions: the eruption of Paricutin
Houghton, B., Castro, M.D.L., 2012. SEM-based methods for the analysis of basaltic 1943-1952. Earth Planet. Sci. Lett. 271, 359–368.
ash from weak explosive activity at Etna in 2006 and the 2007 eruptive crisis at Strom- Pioli, L., Bonadonna, C., Azzopardi, B.J., Phillips, J.C., Ripepe, A.M., 2012. Experimen-
boli. Physics and Chemistry of the Earth, Parts A/B/C 45, 113–127. tal constraints on the outgassing dynamics of basaltic magmas. J. Geophys. Res. Solid
Le Bas, M., Le Maitre, R., Streckeisen, A., Zanettin, B., IUGS Subcommission on the Sys- Earth 117 (B3). doi:10.1029/2011JB008392.
tematics of Igneous Rocks, 1986. A chemical classification of volcanic rocks based on Polanco, E., 1998. Volcanismo explosivo postglacial en la cuenca del Alto Biobío, Andes
the total alkali-silica diagram. J. Petrol. 27 (3), 745–750. del Sur (37°45′ – 38°30′). Memoria para optar al Título de Geólogo. Departamento de
UN

Lemarchand, N., Grasso, J.R., 2007. Interactions between earthquakes and volcano activ- Geología, Universidad de Chile.
ity. Geophys. Res. Lett. 34 (24). Polanco, E., 2010. Volcanoestratigrafía, geoquímica y peligro volcánico del volcán Lon-
Lohmar, S., 2000. Estratigrafía, petrografía y geoquimica del Volcán Antuco y sus depósi- quimay (38°30’S), Andes del Sur (Chile). Departamento de Geoquímica, Petrología y
tos (Andes del Sur, 37°25′S). Memoria de Titulo (Unpublished). Universidad de Con- Prospección Geológic. Facultad de Geología. Universidad de Barcelona.
cepción, Departamento de Ciencias de la Tierra, p. 184. Pollard, A.M., Blockley, S.P.E., Ward, K.R., 2003. Chemical alteration of tephra in the de-
López-Escobar, L., Vergara, M., Frey, F.A., 1981. Petrology and geochemistry of lavas from positional environment: theoretical stability modelling. J. Quat. Sci. 18 (5), 385–394.
Antuco volcano, a basaltic volcano of the southern Andes (37°25’S). J. Volcanol. Ge- Pyle, D.M., 1989. The thickness, volume and grain size of tephra fall deposits. Bull. Vol-
otherm. Res. 11, 329–352. canol. 51, 1–15.
Mandeville, C.W., Carey, S., Sigurdsson, H., 1996. Sedimentology of the Krakatau 1883 Pyle, D.M., 1995. Assessment of the minimum volume of tephra fall deposits. J. Volcanol.
submarine pyroclastic deposits. Bull. Volcanol. 57 (7), 512–529. Geotherm. Res. 69, 379–382.
Marco, S., Agnon, A., 1995. Prehistoric earthquake deformations near Masada, Dead Sea Pyle, D.M., 2000. Sizes of Volcanic Eruption. Encyclopedia of Volcanoes. Academy Press.
graben. Geology 23, 695–698. Rawson, H., Naranjo, J.A., Smith, V.C., Fontijn, K., Pyle, D.M., Mather, T.A., Moreno, H.,
Martínez, P., Singer, B.S., Roa, H.M., Jicha, B.R., 2018. Volcanologic and petrologic evo- 2015. The frequency and magnitude of post-glacial explosive eruptions at Volcán Mo-
lution of Antuco-Sierra Velluda, Southern Andes, Chile. J. Volcanol. Geotherm. Res. cho-Choshuenco, southern Chile. J. Volcanol. Geotherm. Res. 299, 103–129.
349, 392–408. Rodriguez-Pascua, M.A., Calvo, J.P., De Vicente, G., Gómez-Gras, D., 2000. Soft-sediment
Mastin, L.G., Guffanti, M., Servranckx, R., Webley, P., Barsotti, S., Dean, K., Durant, A., deformation structures interpreted as seismites in lacustrine sediments of the Prebetic
Ewert, J.W., Neri, A., Rose, W.I., Schneider, D., Siebert, L., Stunder, B., Swanson, Zone, SE Spain, and their potential use as indicators of earthquake magnitudes during
G., Tupper, A., Volentik, A., Waythomas, C.F., 2009. A multidisciplinary effort to as the Late Miocene. Sediment. Geol. 135 (1–4), 117–135.
J.E. Romero et al. / Journal of Volcanology and Geothermal Research xxx (xxxx) 106759 23

Romero, J., Keller, W., Marfull, V., 2014. Short chronological analysis of the 2007-2009
eruptive cycle and its nested cones formation at Llaima volcano. Journal of Techno-
logical Possibilism 2 (3), 1–9.
Romero, J.E., Viramonte, J.G., Scasso, R.A., 2013. Indirect tephra volume estimations us-
ing theorical models for some chilean historical volcanic eruptions with sustained
columns. Bolletino di Geofisica teorica ed applicata 54 (2), 194–197.
Romero, J.E., Morgavi, D., Arzilli, F., Daga, R., Caselli, A., Reckziegel, F., Viramonte, J.,
Díaz-Alvarado, J., Polacci, M., Burton, M., Perugini, D., 2016. Eruption dynamics of

F
the 22–23 April 2015 Calbuco Volcano (Southern Chile): analyses of tephra fall de-
posits. J. Volcanol. Geotherm. Res. 317, 15–29.
Romero, J.E., Vera, F., Polacci, M., Morgavi, D., Arzilli, F., Alam, M.A., Bustillos, J., Gue-
vara, A., Johnson, J.B., Palma, J.L., Burton, M., Cuenca, E., Keller, W., 2018. Tephra

OO
from the 3 March 2015 sustained column related to explosive lava fountain activity at
Volcán Villarrica (Chile). Front. Earth Sci. 6, 98.
Ruth, D.C., Calder, E.S., 2014. Plate tephra: preserved bubble walls from large slug bursts
during violent Strombolian eruptions. Geology 42 (1), 11–14.
Sable, J.E., Houghton, B.F., Wilson, C.J.N., Carey, R., 2006. Complex proximal sedimen-
tation from Plinian plumes; the example of Tarawera 1886 basaltic plinian eruption.
Bull. Volcanol. 69, 89–103.
Simkin, T., Siebert, L., 1994. Volcanoes of the World. Geoscience Press in Association with
the Smithsonian Institution Global Volcanism Program, Tucson AZ. (368 pp).
Smith, E.R., 1915. Los Araucanos. Trad. por R.E. Latcham. Santiago de Chile. pp. 48–212.

PR
Soriano, C., Riggs, N., Giordano, G., Porreca, M., Conticelli, S., 2012. Cyclic growth and
mass wasting of submarine Los Frailes lava flow and dome complex in Cabo de Gata,
SE Spain. J. Volcanol. Geotherm. Res. 231, 72–86.
Sparks, R.S.J., Bursik, M.I., Carey, S.N., Gilbert, J.S., Glaze, L.S., Sigurdsson, H., Woods,
A.W., 1997. Volcanic Plumes. John Wiley and Sons, New York. (574 pp).
Sparks, R.S.J., Gardeweg, M.C., Calder, E.S., Matthews, S.J., 1997. Erosion by pyroclastic
flows on Lascar Volcano, Chile. Bull. Volcanol. 58 (7), 557–565.
Spina, L., Cannata, A., Morgavi, D., Perugini, D., 2019. Degassing behaviour at basaltic
volcanoes: new insights from experimental investigations of different conduit geome-
try and magma viscosity. Earth Sci. Rev. 192, 317–336. doi:https://doi.org/10.1016/

D
j.earscirev.2019.03.010.
Stuiver, M., Reimer, P.J., 1993. Extended 14 C data base and revised CALIB 3.0 14 C age
calibration program. Radiocarbon 35 (1), 215–230.
Taddeucci, J., Pompilio, M., Scarlato, M., 2004. Conduit processes during the July–August
2001 explosive activity of Mt. Etna (Italy): inferences from glass chemistry and crystal
TE
size distribution of ash particles. J. Volcanol. Geotherm. Res. 137, 33–54.
Taddeucci, J., Scarlato, P., Andronico, D., Cristaldi, A., Buettner, R., Zimanowski, B., Kuep-
pers, U., 2007. Advances in the study of volcanic ash. Eos, Transactions, American
Geophysical Union 88 (24), 253–256.
Taddeucci, J., Edmonds, M., Houghton, B., James, M.R., Vergniolle, S., 2015. Hawaiian
and Strombolian eruptions. In: The Encyclopedia of Volcanoes. Academic Press, pp.
485–503.
EC

Tepley, F.J., III, Davidson, J.P., Clynne, M.A., 1999. Magmatic interactions as recorded in
plagioclase phenocrysts of Chaos Crags, Lassen Volcanic Center, California. J. Petrol.
40 (5), 787–806.
Thiele, R., Moreno, H., Elgueta, S., Lahsen, A., Rebolledo, S., Petit-Breuilh, M.E., 1998.
Quaternary geological-geomorphological evolution of the uppermost course of the
R’ıo Laja Val- ley. Rev. Geol. Chile 25, 229–253.
Tilling, R.I., Heliker, C., Swanson, D.A., 2010. Eruptions of Hawaiian Volcanoes-Past, Pre-
sent, and Future (No. 117). (US Geological Survey).
RR

Torres, L., Parra, O., Araneda, A., Urrutia, R., Cruces, F., Chirinos, L., 2008. Vegetational
and climatic history during the late Holocene in Lake Laja basin (central Chile) in-
ferred from sedimentary pollen record. Rev. Palaeobot. Palynol. 149 (1–2), 18–28.
Urrutia, R., Araneda, A., Torres, L., Cruces, F., Vivero, C., Torrejón, F., Barra, R., Fagel,
N., Scharf, B., 2010. Late Holocene environmental changes inferred from diatom, chi-
ronomid, and pollen assemblages in an Andean lake in Central Chile, Lake Laja (36 S).
Hydrobiologia 648 (1), 207–225.
Vergara, M., Katsui, Y., 1969. Contribución a la Geología y Petrología del Volcán Antuco.
Universidad de Chile, Instituto de Geología, Publicación, Cordillera de los Andes, Chile
CO

Central, pp. 25–47. No. 35.


Walker, G.P.L., 1971. Grain-size characteristics of pyroclastic deposits. The Journal of Ge-
ology 79 (6), 696–714.
Walker, G.P.L., 1973. Explosive volcanic eruptions: a new classification scheme. Geol.
Rundsch. 62, 431–446.
Watt, S.F., 2019. The evolution of volcanic systems following sector collapse. J. Volcanol.
Geotherm. Res. doi:10.1016/j.jvolgeores.2019.05.012.
Watts, P., Waythomas, C.F., 2003. Theoretical analysis of tsunami generation by pyroclas-
tic flows. J. Geophys. Res. Solid Earth 108 (B12).
UN

Westgate, J.A., Gorton, M.P., 1981. Correlation techniques in tephra studies. In: Tephra
studies. Springer, Dordrecht, pp. 73–94.
Wilson, C.J.N., Hildreth, W., 1997. The Bishop Tuff; new insights from eruptive stratigra-
phy. J. Geol. 105, 407–439.
Yamamoto, T., Takada, A., Ishizuka, Y., Miyaji, N., Tajima, Y., 2005. Basaltic pyroclastic
flows of Fuji volcano, Japan: characteristics of the deposits and their origin. Bull. Vol-
canol. 67 (7), 622–633.

You might also like