Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Journal of South American Earth Sciences 111 (2021) 103435

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


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

Dynamic and kinematic characterization of the basement structures of the


Mocho-Choshuenco Volcanic Complex, Southern Andes, Chile
Carolina Peña *, Catalina Cabello, Andrés Tassara
Departamento Ciencias de La Tierra, Facultad de Ciencias Químicas, Universidad de Concepción; Casilla 160-C, Barrio Universitario s/n, Concepción, Chile

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

Keywords: Mocho-Choshuenco Volcanic Complex (MCVC) is one of the most hazardous volcanoes in Chile’s Southern
Mocho-Choshuenco volcanic complex Volcanic Zone (SVZ). It is located at 40◦ S 72◦ W in Los Ríos Region, Chile, where the SVZ coexists with the
Liquiñe-Ofqui fault system margin-parallel Liquiñe-Ofqui Fault System (LOFS) and with NW-striking Andean Transverse Faults (ATF). The
Southern Andes
LOFS is 1200 km-long intra-arc strike-slip fault system located between 38◦ and 46◦ S, and partially controls the
Strain analysis
Stress field
occurrence, spatial distribution and geometry of major stratovolcanoes. The fault-fracture networks induced by
crustal deformation plays an essential role in the generation of pathways, facilitating the transport and
emplacement of magmas. Thus, the Chilean Andes provides one of the best natural laboratories in the world to
assess the link between tectonics and volcanism. In this study, we provide new structural, kinematic and dynamic
insights, which would condition the formation and evolution of the MCVC and laying foundations for future
works.
Structural field data was collected and analyzed from regional and local scales. The strain analysis shows
heterogeneous deformation. The stress field analysis at a regional scale indicates a strike-slip dominated trans­
pressional regime with N61◦ E-trending σ1 and N26◦ W-trending σ3, which is consistent with the actual oblique
convergence. Deformation driven by this oblique convergence is partially partitioned, where the margin-parallel
component originates the dextral strike-slip in LOFS. Results are consistent with reactivation of preexisting NW-
striking faults under the regional stress field, facilitating hydrothermal fluid flow and magma migration, asso­
ciated with volcanoes, dikes and minor eruptive centers present in the study area. As a final result, we propose a
simplified, three-dimensional model for the MCVC’s basement, contributing to the understanding of the system
and the way in which the different factors interact.

1. Introduction deformation imposed by oblique plate motion is accommodated in the


intra-arc by Liquiñe–Ofqui Fault System (LOFS) (e.g. Arancibia et al.,
One of the fundamental problems in active tectonic margins is the 1999; Cembrano et al., 1996; Pérez-Flores et al., 2016), a 1200 km-long
interplay between tectonics and volcanism. Traditionally, the nature strike-slip fault system located between 33 and 46◦ S. The LOFS, in
and transport of magma through the lithosphere has been studied from a addition to anisotropies inherited from the basement, partially controls
petrological or geochemical view, largely neglecting the key role of the occurrence, geometry and spatial distribution of major eruptive
deformation (Cembrano and Lara, 2009). However, in the last years, vents (e.g. Lara et al., 2008; Cembrano and Lara, 2009; Sielfeld et al.,
several authors have contributed to clarify the importance of this rela­ 2016).
tionship, postulating that stress fields and structures linked to brittle However, in the southern Andes there are many poorly studied
deformation play a fundamental role in controlling the morphology of volcanic centers from the point of view of this link at a local scale. The
volcanic edifices, the spatial distribution and geometry of minor erup­ local nature of the volcanic arc basement acts as a second-order factor in
tive centers, and their associated feeder dikes (e.g. Corazzato and this complex interaction between tectonic and magmatic processes
Tibaldi, 2006; Lara et al., 2006; Pérez-Flores et al., 2016; Sielfeld et al., (Cembrano and Lara, 2009) and may even suggest shallow magma
2016; Tibaldi et al., 2017). The Southern Volcanic Zone (SVZ) is one of pathways be more sensitive to the presence of upper crustal weakness
the best natural laboratories for this kind of studies, because part of the zones than to the regional stress (Tibaldi et al., 2017). Therefore, a real

* Corresponding author.
E-mail address: capena@udec.cl (C. Peña).

https://doi.org/10.1016/j.jsames.2021.103435
Received 31 May 2020; Received in revised form 6 March 2021; Accepted 16 June 2021
Available online 24 June 2021
0895-9811/© 2021 Elsevier Ltd. All rights reserved.
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

understanding of this relationship requires local studies that reflect the (MCVC) is presented as an ideal site to find new clues about this link. The
key role of minor structures. Knowledge of the dynamic of local struc­ study area corresponds to the surroundings of the Mocho-Choshuenco
tures could be used to complement the regional analysis and improve Volcanic Complex (MCVC), a compound stratovolcano with a domi­
insight on the characteristics and behavior of long-lived structures and nant andesite composition (Moreno and Lara, 2007). It is located in the
their relationship with different volcanic systems. central part of the SVZ (López-Escobar et al., 1995) at 39.9◦ S, sur­
With all this background, the Mocho-Choshuenco volcanic complex rounded by important NW-elongated lakes and just 20 km W from the

Fig. 1. 1:5,000,000 geological map of the Southern Andes Volcanic Zone (SVZ), modified from SERNAGEOMIN (2003). Regional map showing the location and
extent of the LOFS (red lines), ATF (blue lines), Faults (black lines) and main Holocene volcanoes (green triangles). Convergence velocity vector of Nazca Plate after
Angermann et al. (1999). A blue square shown MCVC location. (For interpretation of the references to color in this figure legend, the reader is referred to the Web
version of this article.)

2
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

main line of the LOFS. The MCVC volcanic products evidence contin­ (e.g. Lange et al., 2008; Pérez-Estay et al., 2020) verifying its recent
uous volcanic activity from the middle Pleistocene to historical erup­ movement. On February 11 (2021) a Mw 4.7 earthquake took place in
tions (Rawson et al., 2015). Moreover, in this area there are an the study area 37 km E of Panguipulli locality. Due to its strike-slip focal
important number of flank and monogenetic cones, making it an mechanism, it was associated with the Neltume fault (Centro Sis­
excellent place for the local scale studies of the relationship between mológico Nacional; CSN). This earthquake triggered landslides and rock
tectonism and volcanism. Unfortunately, only few works carried out in falls, followed by several smaller aftershocks. This event highlights that
the area mention this link (Esparza, 2006; Gardenweg and Sellés, 2012; this fault system and the associated secondary faults present a great
Rawson et al., 2016) and none of them in a deep and specific way. This seismogenic potential.
work seeks to clarify the nature of the basement structures in the study The SVZ is characterized by different types of volcanism depending
area and the possible role they play in the development and evolution of on the structures involved. Cembrano and Lara (2009) proposed a
the volcanic complex. conceptual model to explain the spatial relationship between the
As a final result, a structural model of the MCVC basement was built, first-order geochemical characteristics of volcanic systems and the
providing new structural, kinematic and dynamic insights that charac­ structural configuration of the basement for the SVZ. Stratovolcanoes in
terize the study area, which would condition the formation and evolu­ a (1) NE-trending system represent tension fractures, extensional-shear
tion of the MCVC. fractures or tail cracks oriented subparallel to the maximum horizon­
tal stress and are dominantly basaltic to andesitic (e.g.
2. Geological and tectonic background Callaqui-Copahue-Mandolehue; López-Escobar et al., 1995). Mean­
while, the (2) NW-trending systems have a compressive regime, and
2.1. Structural setting facilitate the permanence of magma in horizontal reservoirs for long
periods of time, favoring its magmatic differentiation and frequently
At Southern Andes magmatic arc, known as Southern Volcanic Zone include dacitic to rhyolitic rocks (e.g. Puyehue-Cordón Caulle; Lara
(SVZ) between 33◦ and 46◦ S (López-Escobar et al., 1995), the stress et al., 2004). These NW-SE systems correspond to co-seismic misor­
distribution and deformation pattern are governed by the iented fissural volcanism, however there are cases spatially and
dextral-oblique convergence between Nazca and South America plates, temporally associated with ancient structures that may provide direct
at a rate of ca. 66 mm/yr that has prevailed for the last 20 Ma (e.g. pathways for the ascent of magmas (e.g. Villarrica–Quetrupillán–Lanín;
Pardo-Casas and Molnar, 1987; Jarrard, 1986; Sielfeld et al., 2019), with Cembrano and Lara, 2009). This “contradiction” evidences the role of
a subduction angle of ca. 16◦ and a convergence angle (α) of ca. 78◦ regional and local structures which condition the behavior and
(Angermann et al., 1999) (Fig. 1). It should be noted that at 38◦ S sig­ composition of volcanic systems.
nificant changes take place in the basement architecture, intra-arc tec­
tonics and volcanic nature (Sielfeld et al., 2016). Exhumation rates are 2.2. Local geology
limited to less than 0.1 mm/year north of this latitude and more than 1
mm/year south of this latitude (Glodny et al., 2008). From 38◦ S Qua­ The Mocho-Choshuenco Volcanic Complex (MCVC) sits on top of the
ternary volcanism is spatially and genetically linked to the Liqui­ basement’s intrusive rocks belonging to the North Patagonian Batolith,
ñe-Ofqui Fault System (LOFS) (Fig. 1; Cembrano and Lara, 2009), and partially covered by Pleistocene and Holocene volcanic rocks.
resulting in a different causal link between arc volcanism and crustal Furthermore, in the study area there are minor outcrops of Mesozoic
deformation. sedimentary rocks and low-grade metamorphic rocks assigned to the
Part of the deformation generated by the oblique plate motion is Upper Paleozoic (Di Biase, 1976) (Fig. 2).
accommodated in the intra-arc LOFS (Fig. 1; e.g. Dewey and Lamb, 1992; The plutonic basement is conformed by intrusive rocks generated by
Arancibia et al., 1999; Cembrano and Lara, 2009; Cembrano and Hervé, different magmatic pulses ranging in age from the Carboniferous to the
1993), with variations in the nature, time and degree of partial parti­ Miocene (Munizaga, 1988; Rodríguez et al., 1999). These plutonic rocks
tioning of the deformation (Cembrano et al., 2000). are grouped in Futrono-Riñihue Batolith (Carboniferous-Lower
The LOFS is 1200 km-long, N10◦ E striking, intra-arc strike-slip fault Permian), Panguipulli Pluton (Lower-Middle Jurassic), Paimún (Creta­
system, that dominates the region between 38◦ and 46◦ S (Cembrano and ceous) and Miocene Granitoids, Miocene Porphyry and their respective
Lara, 2009; Rosenau et al., 2006), which is associated with NE-striking equivalents (Rodríguez et al., 1999; Lara and Moreno, 2004; Moreno and
normal-dextral faults that splay off NNE-striking faults (Pérez-Flores Lara, 2007) (Fig. 2), mostly with dominant differentiated compositions.
et al., 2016). LOFS is spatially related to Quaternary volcanic activity, Panguipulli Pluton cuts a Triassic sedimentary sequence at the east of
controlling the architecture of the volcanic arc and the occurrence, the study area (Panguipulli Formation; Rodríguez et al., 1999). Towards
spatial distribution and geometry of major stratovolcanoes (Fig. 1; the south, a Miocene volcanic-sedimentary sequence (Lago Ranco
Melnick et al., 2006; Cembrano and Lara, 2009; Sielfeld et al., 2016). Layers) is intruded by Miocene Granitoids (Campos et al., 1998). Near
Present-day seismicity has been detected throughout the fault system the Liquiñe locality, metamorphic rocks, belonging to Liquiñe Meta­
(Lange et al., 2008), registering even larger earthquakes (e.g. Mw 6.2 in morphic Complex (Paleozoic-Triassic) were observed at the E of LOFS
Aysén Fjord) (Legrand et al., 2011; Mora et al., 2010). (Hervé, 1977; Moreno and Lara, 2007).
In turn, NW- striking second order structures (Fig. 1) interpreted by In the study area, the NW-striking morphological features stand out
several authors as inherited faults from subsequently reactivated pre- (Moreno and Lara, 2007) Roquer et al. (2019) studied an exposed part of
Andean geological processes (e.g. López-Escobar et al., 1995; Melnick the ATF just north of the study area (39◦ S), however around the MCVC
et al., 2006; Stanton-Yonge Sesnic, 2016; Veloso et al., 2019), are there are no NW-striking faults documented. The only evidence of the
recognized as part of a fault system called Andes Transversal Faults ATF in the area is the alignment of glacial valleys, river courses and lake
(ATF) (Pérez-Flores et al., 2016; Sielfeld et al., 2019). ATF include a basins. The main faults in this zone are the LOFS and Pirehueico-Reigolil
group of sinistral and sinistral-reverse NW to WNW-striking faults and Fault, located east of the study area, which is responsible for
morphological lineaments (Roquer et al., 2019). At regional scale these NNE-striking morphologic features and putting rocks of different ages in
structures play a key role in the subduction earthquake segmentation contact with each other (Fig. 2). Other important faults in this area,
(Melnick et al., 2009) and are spatially and genetically associated with recognized by other authors, are Los Baños, Enco and Neltume faults.
the occurrence of several ore deposits (e.g. Piquer et al., 2016), Cenozoic The first one is a NW-striking sinestral-slip fault, located on the north
basins (e.g. Radic, 2010) and placement and alignment of volcanic shore of Pirehueico Lake, which interrupts and separates the
chains (e.g. Lara et al., 2006; Cembrano y Lara, 2009). Pirehueico-Reigolil fault (Lara and Moreno, 2004). The Enco fault is a
Several authors have recorded and studied seismicity along the LOFS minor NE-striking normal fault, marked by Enco river basin and it

3
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

Fig. 2. Geological-structural map of the study area, between 39◦ S and 40◦ S. Geology was modified from Rodríguez et al. (1999) and Lara and Moreno (2004), and
structures was modified from Esparza (2006); Rodríguez et al. (1999); Moreno and Lara (2007); Lara and Moreno (2004); Di Biase (1976).

connects the drainage of Panguipulli and Riñihue lakes (Rodríguez et al., 1976; Moreno and Lara, 2007). Rawson et al. (2015) identified 27 vol­
1999). Finally, the Neltume fault (Campos et al., 1998) is a NE-striking canic deposits; these units were interpreted as products of a separate
dextral-slip fault, which could be the northern extension of the Maihue explosive eruption.
fault (Rosenau, 2004). Additionally, there are ca. 45 minor scoria cones on the flanks
(Moreno and Lara, 2007). The cones are subdivided into four main
2.3. Mocho-Choshuenco Volcanic Complex groups: Chanchán (4 cones to the NE of the volcanic complex), Fui (ca.
25 cones to the NE of the edifice; subdivided into Fui Sur and Fui Alto),
The MCVC is a compound stratovolcano located at SVZ (39.9◦ S): Alto Caunahue (ca. 10 cones at the SW flank) and Ranquil (4 cones at the
Mocho volcano has a 4 km-wide caldera, infilled by a glacier and with a NE flank) (Fig. 3). Both the orientation of these groups of cones and the
young scoria cone in the center, called Mocho; and Choshuenco volcano, volcanic complex itself suggest a spatial and genetic relationship with
which forms part of the north-western rim of the caldera, has only been the feeder fractures and surrounding structures (e.g. Di Biase, 1976;
partially affected by the collapse so it retains its conical form (Rawson Gardenweg and Sellés, 2012).
et al., 2015; Moreno and Lara, 2007). The volcanic product composition
ranges from basaltic andesite to rhyolite (Rawson et al., 2016), 3. Methodology
evidencing continuous volcanic activity from the middle Pleistocene
(~350 ka) to historical eruptions. MCVC is one of the most productive 3.1. Field structural mapping
(ca. 1 km3/kyr) and active volcanoes in the SVZ during post-glacial
times (<18 ka), in explosiveness terms, with ca. 75 explosive erup­ This study is based on the results obtained from the field work car­
tions and an eruptive frequency of one explosive eruption every ~220 ried out in the surroundings of the MCVC. Prior to the field work major
years (Rawson et al., 2015). sites of interest were established based on photointerpretation of Digital
The volcanic units present in this area are represented by effusions Elevation Models (DEM) and QuickBird ® imagery (freely available
from the caldera, Choshuenco volcano, Mocho volcano and parasite from Google Earth) in combination with information from previous
activity originated by a series of minor eruptive centers and mono­ works. In these areas five key structural sites were defined: Lago Pan­
genetic cones, located on the flanks of the volcanic complex (Di Biase, guipulli, Liquiñe, Lago Neltume, Lago Pirehueico and Enco (Fig. 2 and

4
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

Fig. 3. Digital Elevation Model showing the preferential orientation of flank and monogenetic cones: Alto Caunahue Group (yellow asterisk), Ranquil Group (red
asterisk), Fui Volcanic Group (green asterisk) and Chanchán Volcanoes (blue asterisk) (modified from Moreno and Lara, 2007; Gardeweg and Sellés, 2012). (For
interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

Table 1). A detailed structural mapping was performed at each structural uniform deformation field, only when P and T axes are presented in
site, characterizing the geometry and kinematics of the fault planes. The defined clusters, the fault data population will be the result of homo­
strike and slip of the all identified fault planes were measured and in­ geneous faulting and the kinematics axes will have a geological sense.
formation such as the width of the fault, texture of the fault, striations The table summarized the main structural elements mapping at each
and crosscutting relationship was collected. The sense of movement was structural site and their strain and stress solutions. The age of the host
established using kinematic criteria for brittle faults (e.g. Petit, 1987). rock from data of Munizaga et al. (1988), Rodríguez et al. (1999) and
The type of kinematic indicator was also recorded, mostly R-type frac­ Lara and Moreno (2004).
tures and mineral slickenfibers.
3.2.2. Dynamic analysis
3.2. Fault-slip data inversion The “Multiple Inverse Method” (MIM) (Yamaji, 2000) was used to
determine the stress field. This method calculates the orientation of
All data collected in the field was kinematically and dynamically paleo-stress tensors from the inversion of fault-slip data and allows
analyzed, from regional and local scales. The regional scale includes all isolation of homogeneous stress fields recorded on polyphasic hetero­
fault-slip data measured in the study area, whereas, for local scale the geneous fault-slip data. The stress state is defined by four parameters:
data inversion analyses were done for each structural site the orientation of the three principal stresses axes and the stress ratio ϕ
independently. (ϕ = σ2 - σ3/σ1 - σ3). To obtain this information, the software requires
the strike and dip of the fault planes and direction and sense of
3.2.1. Kinematic analysis displacement. Based on this data set, the program generates a combi­
The method of fault-slip analysis was used for the determination of natorial number of sub-groups of k-faults (Yamaji, 2000) and calculates
strain field (Marrett and Allmendinger, 1990; Allmendinger et al., 2012) the optimal stress tensor for each of the groups of k-fault planes.
by means of FaultKin 7 software. The software calculates and plots the For calculating the stress axes, the software uses classical stress
orientation of the compression (P) and tension (T) axes for each indi­ inversion methods based on the Wallace-Bott hypothesis (Wallace,
vidual fault plane and the average kinematic axes (shortening, stretch­ 1951; Bott, 1959), according to which the displacement in a fault plane
ing and intermediate axes) for different fault populations, based on the occurs parallel to the orientation of the maximum shear stress, under a
orientation of the fault plane and the displacement vector resolved on defined stress tensor. For an optimal result, it attempts to minimize the
such fault plane (slickenside) (Marrett and Allmendinger, 1990). The misfit angle, defined as the angle between the observed and theoretical
distribution of the P and T axes can show if the data represents or not a slip direction on a fault plane (for an assumed stress state). An optimal

5
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

stress state is defined when the misfit angle is less than 30◦ for each of

The table summarized the main structural elements mapping at each structural site and their strain and stress solutions. The age of the host rock from data of Munizaga et al. (1988), Rodríguez et al. (1999) and Lara &
the k-fault planes contained in the group. Finally, the values for σ1 and

Tectonic regime

Transtensional
σ3, calculated for each generated sub-group of k-faults, are plotted onto

Strike-slip
two different stereograms. Solutions of principal axes are color-coded

strike-slip
according to their stress ratio (ϕ) and have an attached tail pointing
toward the orientation of the complementary principal axis (e.g. σ1’s tail


points toward σ3, and vice versa). If the principal axes are in a single

0,7

0,5
ϕ cluster, with similar stress ratio colors and tail points orientation, it


Dynamic secondary stress

indicates a homogeneous stress tensor, whereas, if the fault-slip data are

330/08

078/19
heterogeneous, every stress state will be represented in a different
cluster (Yamaji, 2000). The distribution of ϕ-values may be a calculation
σ3


tool, for example when the ϕ-histogram shows a Gaussian type distri­
067/38 bution, the major value will be the more representative, but in the case

347/03
of a bimodal distribution, one stress field solution is calculated for each
σ1

representative ϕ-value. In some cases, the orientation of the principal



axes solutions may display widespread distribution, hindering selection
of the representative stress field. Thus, the tectonic regimes proposed by
Tectonic regime

Transpressional
Transtensional

Transtensional

Ritz (1994) were used, in which the orientation of the principal axes are
Strike-slip
strike-slip

strike-slip

strike-slip

determined by the tectonic regimes as extensional, compressional or


strike-slip. Depending on the ϕ-value, the strike-slip regime could be
either transpressional (ϕ < 0.5) or transtensional (ϕ > 0.5).

In this study, the dynamic analysis was carried out considering a


0.7

0,4

0,2

0,5

minimum of 10 faults. For the most of the analysis 5 was used as k-value,
ϕ


Dynamic principal stress

recommended by Yamaji (2000) for databases with less than 100 fail­
283/54

330/22

354/40

245/66

ures. Only for Liquiñe structural site a k-value equal to 3 was used due to
σ3

the low number of available data.


4. Results
028/10

067/17

254/12

034/21
σ1

4.1. Architecture and geometry of fault-fracture networks in the study


area
304/25

323/15

349/19

147/46

094/04
T-axe

The principal structural elements of each structural site, based on


field work, are summarized in Table 1, and the spatial distribution of the
Kinematic

036/04

055/09

080/02

049/08

190/60

main faults are illustrated in Fig. 4. The described structural elements


P-axe

are observed in metamorphic, igneous and volcanic-sedimentary base­


ment rocks, ranging in age from Triassic to Miocene (Moreno and Lara,
NS to NE-SW EW

2007) (Fig. 2).


to ENE NE-SW

NW-SE NE-SW
NS to NE-SW

NS to NE-SW

NS to NE-SW

In the MCVC’s surrounding area, different fault sets were observed


orientation

ENE-WSW

and inferred (Fig. 4), which can be divided into (1) NNE-striking first
Principal

NW-SE

order faults, (2) NW-striking second order faults and (3) local NE-
striking and N–S striking third order faults.
The master (first order) faults are represented by two km-long NNE-
striking subparallel and subvertical faults: Pirehueico-Reigolil fault and
Main structural

andesitic dikes
sinestral faults

sinestral faults

the main branch of LOFS, in addition to Neltume fault, a synthetic NE-


Dextral faults

Dextral faults

Dextral faults

Dextral faults

Dextral faults
normal faults

normal faults

striking fault (Fig. 4). The Liquiñe-Ofqui fault is strongly marked by


elements

the alignment of valleys and escarpments, cutting dominantly granitoid


basement and bringing together Triassic and Cretaceous rocks, of the
Liquiñe Metamorphic Complex and Paimún granitoids respectively,
Miocene granitoids (Miocene, 18-5 Ma).

with Miocene granitoids (Lara and Moreno, 2004). The same applies to
Granitoid, Panguipulli Pluton (Middle-

Granitoid, Panguipulli Pluton (Middle-


Tonalite, Panguipulli Pluton (Middle-

Lutite, Panguipulli Formation (Upper

Pirehueico-Reigolil fault with Jurassic granitoids and Paleozoic meta­


Lower Jurassic, 180-160 Ma) and

morphic rocks. The main branch of LOFS, despite being a well-marked


Gneisses, Liquiñe Metamorphic
Complex (Triassic protholite)

Lower Jurassic, 180-160 Ma)

Lower Jurassic, 180-160 Ma)

lineament at a regional scale, is difficult to find good outcrops with


exposed faults. In Liquiñe structural site, faults with cm-to m-scale thick
Triassic, 230-208 Ma)

are observed (Fig. 5b), mostly oriented on N–S or E-W, but the largest
structures observed contains water flows and a thick vegetation cover,
Host rock (age)

making it difficult to register. Further south, at Lago Pirehueico struc­


tural site, the rectilinear NNE-striking valleys are observed cutting the
mountains (Fig. 5a). These valleys are the morphological expression of
LOFS and their intrusive rocks are cut by a set of NE-striking andesite
dikes (Fig. 5c and Table 1), reaching a maximum thickness of 3.5 m. The
Lago Panguipulli
Lago Pirehueico
Moreno (2004).

Lago Neltume structural site is a synthetic fault zone (Fig. 5a), in this
Structural site

Lago Neltume

place the faults are mostly randomly oriented (Fig. 5d), noting the
absence of NW-striking structures. Several faults show mm-cm scale
Liquiñe
Table 1

Enco

thick and subvertical dip (Fig. 6d). In the vicinity of MCVC there are a set
of minor eruptive centers with an evident volcanic alignment (Fig. 3).

6
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

NNE-striking Fui Alto Volcanic Group is subparallel to Liquiñe-Ofqui calcite slickenfibres (Fig. 6b). NE-striking and NW-striking faults mainly
fault, meanwhile Chanchán Volcanoes, Fui Sur Volcanic Group and Alto present a dextral-slip and left-slip respectively. The Ranquil group,
Caunahue Group present a NE-SW orientation. located at the N-flank of the MCVC (Fig. 3), consists of a set of
Second order faults are km-long NW-striking left-slip to left-reverse NW-striking cones, sub-parallel to the orientation of the volcanic com­
oblique slip faults, mostly correspond to faults inferred from morpho­ plex, this alignment would be directly related to this fault zone.
logic lineaments with the exception of the Panguipulli structural site The local third order faults correspond to cm to m-scale structures of
where faults were measured. This type of faults associated with ATF varied orientations and sense of movement, mainly with NE- and N–S
corresponds to pre-Andean anisotropies inherited from the basement (e. strike and cores that do not exceed 3 cm thick. Enco structural site is
g. Pérez-Flores et al., 2016; Piquer et al., 2016) as well as LOFS, are representative of this type of faults, but the high extent of vegetation
strongly evidenced in the study area by the alignment of valleys, lakes cover did not allow a large amount of data to be documented (Fig. 6c).
and even the volcanic complex itself. The representative outcrop of However, there is a trend in a NE-SW direction, which coincides with the
second order faults is Lago Panguipulli structural site, characterized by valley of the Enco River (Fig. 6a).
m-long faults with cores that go up to 20 cm thick, some of them with

Fig. 4. (a) Study area final geological-structural map, showing LOFS (black lines) and the NW-Striking faults belonging to ATF (red lines). The white squares show
the different structural sites (Liquiñe, Lago Pirehueico, Lago Neltume, Lago Panguipulli and Enco). (b) The rossette diagram corresponds to the direction of the dikes
measured in the field. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

7
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

4.2. Kinematics and dynamics of fault-fracture networks analyzed by structural sites, in an effort to find and differentiate
local-scale homogeneous strain fields recorded within the regionally
4.2.1. Regional-scale analysis heterogeneous fault-data set.
For the regional kinematic and dynamic analysis, the entire fault Regional stress analysis shows several trends in the ϕ-histogram,
population were used (n = 75), collected from the five structural sites with main modes at 0.3 and 0.6 (Fig. 7d). Whereas, clustering of stress
and few of them from others parts of the study area. Inversion proced­ axes solutions display a well-marked trend: NE-SW orientation for σ1
ures were applied to the database, obtaining different parameters axis and NW-SE in the σ3 axis case (Fig. 7c). From this, the calculated
(Fig. 7). stress field reflects two stress axes solutions: (1) a main transpressional
The regional strain analysis shows P and T axes widely distributed, regime associated with a ϕ-value of 0.3 yields a N61◦ E-trending sub­
forming poorly defined clusters, (Fig. 7b). This suggests that in the study horizontal σ1 axis and a N26◦ W-trending subhorizontal σ3 axis (Fig. 7c)
area there is not homogeneous kinematic information, hinting at a and (2) a secondary transtensional regime linked to ϕ-value of 0.6, with
complex geological evolution. Considering this, the strain analysis a N42◦ W-trending subvertical σ1 axis and a N45◦ W subvertical σ3.
would not be representative without an arbitrary data selection (Marrett
and Allmendinger, 1990; Veloso et al., 2015; Pérez-Flores et al., 2016; 4.2.2. Local-scale analysis
Yamaji, 2000). To overcome this problem, the fault-slip data was A summary of the strain and stress field results of fault-slip data at

Fig. 5. (a) Google Earth image of the study area’s northeast zone showing major faults (black lines) and inferred faults (dotted line) at Liquiñe and Lago Pirehueico
structural sites. (b) Normal fault Liquiñe structural site (N8◦ E/86◦ W rake 76◦ ). (c) 60–80 cm wide dike at the Lago Pirehueico structural site (N35◦ E/83◦ W). (d)
Reverse fault at Lago Pirehueico structural site (N5◦ E/43◦ E rake 135◦ ).

8
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

each structural site is shown in Fig. 8 and Table 1. Second order Andes transverse fault system is represented by the
The NNE-trending fault zones associated with the LOFS (Lago Pir­ Lago Panguipulli structural site (Fig. 6a and Table 1). Its fault-slip data
ehueico and Liquiñe structural sites) (Fig. 5a and Table 1) show P and T reflects a preferential NW-trend, the clustering of P and T axes show a
axes defined clusters with a minor scattering. The principal stress axes wide distribution. This heterogeneous fault-slip data analyzed suggests
indicate NE-trending shortening and NW-trending stretching and the polyphasic deformations and/or the reactivation of pre-existing anisot­
strain field solutions show ENE-to N-trending subvertical σ1 and WNW- ropies. The stress analysis evidence two stress field solutions; the ob­
to W-trending subvertical σ3, corresponding to strike-slip tectonic tained transpressional (principal) and strike-slip (secondary) local stress
regime (ϕ = 0.4–0.7). In Lago Neltume structural site, the fault-slip data fields share a common ϕ-value (ϕ = 0.5). The principal solution presents
shows dispersion of P and T axis, with trends near the E-W and N–S NNE-trending subhorizontal σ1 and subvertical σ3. Meanwhile, the
orientations respectively. The principal stress axes solutions indicate secondary strain field solution shows a subhorizontal N-trending σ1 with
two solutions for the stress field, the first one with an ENE-trending E-W trending σ3 axis.
horizontal σ1 and NNW-trending horizontal σ3, indicating a trans­ The kinematic analysis of Enco structural site (Fig. 6a and Table 1)
pressive tectonic regime (ϕ = 0.2). The secondary solution shows an shows moderate dispersion of P and T axes, keeping sub-vertical
ENE-trending σ1 and NNW-trending horizontal σ3 in a transtensional trending for P axis. Unfortunately, the limited fault data collected at
regime (ϕ = 0.7). The orientations of kinematic and dynamics axes in the site precludes a dynamic analysis. The heterogeneity of this site
this structural site show a good agreement for these N–S striking fault could be attributed to the reasons mentioned above or to the little fault-
sets. slip information in this area.

Fig. 6. (a) Google Earth image of the study area’s northwest zone showing major faults (black lines) and inferred faults (dotted line) at Lago Panguipulli, Enco and
Lago Neltume Structural site. (b) Reverse fault at Lago Panguipulli structural site (N75◦ E/85◦ W rake 130◦ ). (c) Dike with a great alteration to clay, 50 cm thick at the
Enco structural site (N34◦ E/24◦ SE). (d) Normal-dextral slip fault at Lago Neltume Structural site (N46◦ E/43◦ SE rake 130◦ ).

9
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

Fig. 7. Lower hemisphere, equal-area projections showing gathered fault-slip data and results of strain and stress analyses, at regional scale. (a) Mesoscopic fault-slip
data. (b) Orientation of T (red dots) and P (blue dots) axes (e.g. Marrett and Allmendinger, 1990). (c) Orientation of calculated stress axes (σ1 and σ3) for principal
(red triangle and star) and secondary (blue triangle and star) stress (MIM. Yamaji, 2000). (d) ϕ-Histogram. (For interpretation of the references to color in this figure
legend, the reader is referred to the Web version of this article.)

4.3. Feeder dikes and minor eruptive centers the frequency, magnitude and characteristics of the volcanic activity
presented by MCVC, the changes in the stress field due to the last great
As a first-order approach, regional stress field controls the spatial glaciation, its effect on the magmatic reservoir, and the relationship of
distribution and geometry of dikes and minor eruptive centers (e.g. the flank and monogenetic cones with the volcanic complex. The
Nakamura, 1977; Sielfeld et al., 2016; Tibaldi et al., 2017), nonetheless structural characterization provided in this work are consistent with that
favorably oriented pre-existing weakness zones may reactivate and be proposed by the previous authors.
the loci for magma migration (Sielfeld et al., 2016).
In field, mostly andesitic variable-thicknesses dikes were observed. A
total of 15 dikes orientations were measured in rocks of Panguipulli 5.1. Results analysis
Pluton, Miocene granitoids and Lago Ranco Layers, 12 of them in
Jurassic rocks located in the Lago Pirehueico structural site. A prefer­ The presence of LOFS is clearly reflected in the geometry and strike-
ential NE-SW bimodal arrangement is obtained (Fig. 4). slip kinematics of the structural sites located in its vicinity (Liquiñe and
The spatial distribution of minor eruptive centers (Fig. 3) shows a Lago Pirehueico) (Fig. 4). The same applies to the other structural sites,
clear volcanic alignment in 3 main orientations: (1) NE-trending, pre­ where each one is consistent with what is expected for its location: the
sented by Chanchán volcanoes, Fui Sur and Alto Caunahue Group, (2) Neltume structural site is associated with NE-trending strike-slip faults,
NNE-trending Fui Alto Group is located in the vicinity of the main the Lago Panguipulli structural site reflects NW-striking left-slip struc­
branch of the LOFS and (3) NW-trending Ranquil Group is aligned in tures, which limit the edge of the Panguipulli lake; and the Enco struc­
accordance with the trend of the inherited structures of the basement. tural site, related to a normal structure that controls the Enco river basin
(Figs. 4 and 6a).
5. Discussion Since the study area is under a transpressional deformation regime, it
is possible to compare the structural configuration with the model
Several authors have carried out studies in the area around the proposed by Riedel (1929). At first glance, it would appear that the
Mocho-Choshuenco volcanic complex (e.g. Hervé, 1976; Pérez, 2005; geometry and kinematics of the fault-fracture network at a regional scale
Moreno and Lara, 2007; Gardenweg and Sellés, 2012; Rawson et al., (Fig. 4) assert that the strain is concordant with the model, however the
2015), highlighting the marked presence of the LOFS and the volcanic angles of the structures do not match completely and there are tempo­
activity of MCVC. Esparza (2006), in his structural work, reveals the rality problems. Tentatively, if the structures in the area were considered
occurrence of several local structures associated to the volcanic system, to fit this model, the main fault would correspond to the LOFS. The
whereas the tephrostratigraphic and geochemical studies (e.g. Rawson Neltume fault could be interpreted as an R structure of the main fault,
et al., 2015; Gardenweg and Sellés, 2012; Rawson et al., 2016) discuss even though it has a higher angle than expected. Meanwhile the ge­
ometry and the normal movement of the Enco fault would be consistent

10
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

Fig. 8. Lower hemisphere, equal-area projections showing gathered fault-slip data and results of strain and stress analyses at each structural site. (a) Mesoscopic
fault-slip data and their kinematics indicators. (b) Solution of P and T axes whit their orientation of T (red dots) and P (blue dots) axes (e.g. Marrett and Allmendinger,
1990). (c) Orientation of calculated stress axes (σ1 and σ3) for principal (red triangle and star) and secondary (blue triangle and star) stress (MIM. Yamaji, 2000), (d)
Histogram of phi frequency. (For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)

with Riedel’s model. Fault reactivation of pre-existing faults depends geometry of the local data presented here in a better way.
mainly on depth, rock mechanical properties, pore fluid pressure and the Unfortunately, the faults’ age in the study area is not available.
orientation of faults with respect to the maximum principal stress axis (e. Methodological difficulties do not allow us to determine directly or
g. Sibson, 1996, 2003; Rosenau, 2006). For this reason, we speculated indirectly the age of the faults, due to the absence of datable minerals in
that the NW-trending faults (Panguipulli and Los Baños faults) being the fault planes and observable cross-cutting relationships. However, the
inherited structures with favorable orientation regarding the current results obtained from the dynamic analysis of the data give stresses
stress regime were reactivated as a kind of R′ structures, accommodating consistent with a single predominant tectonic regime which is the result
the deformation in spite of not being originated directly by the stress of the oblique convergence. Therefore, regardless of the age of the faults
regime. The presence of tension cracks with NE-orientation could be it can be assumed that most of the structures were active simultaneously
reflected by the dominating NE-trending dikes and several NE-alignment in recent times under this stress field, superimposed on past tectonic
minor eruptive cones. However, in a strict sense, in order to define an regimes that may have been recorded. Changes in the local stress field
ideal Riedel’ model, a western boundary strike-slip fault is required. indicate that the orientation of regional stress axes is disturbed by the
Neither field work nor satellite data show the existence of any major intersection of fault zones (e.g. Pérez-Flores et al., 2016). This hetero­
fault towards the west of the study area, making it probable that towards geneity of stress fields associated with different fault-fracture systems is
the W-limit, the system will gradually fade, becoming more marked in what allows the emplacement and transport of magma and, therefore,
the proximity of the main line of the LOFS. In the future, it would be the existence of volcanic activity in the MCVC surroundings.
desirable to produce structural model that matches the distribution and

11
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

may drive to build a volcano edifice with an elongated anatomy. Thus,


dikes directly underneath the MCVC probably present a predominant
NW-trend.
The 3 main orientations of minor eruptive centers could be associ­
ated with principal structures. NE-trending orientations (Chanchán
volcanoes, Fui Sur and Alto Caunahue Group) are favorable for the
development of tension cracks, as they are consistent with the σ1-trend
obtained for the regional stress field, optimal for the occurrence of
monogenetic volcanic fields (Corazzato and Tibaldi, 2006).
NNE-trending minor eruptive centers (Fui Alto Group) located
sub-parallel to LOFS suggest that this fault system plays an important
role in the transport of magmatic fluids through the crust. Finally,
NW-trending cones (Ranquil Group) aligned sub-parallel to the elon­
gation of the volcanic complex, hint at a genetic link of its feeder
structures.
In terms of composition, the major elements of the volcanic products
indicate a consanguinity of the magmas of the MCVC, independent of the
issuing vent, highlighting the absence of basalts (McMillan et al., 1989;
Moreno and Lara, 2007). However, certain differences are found by
comparing these compositions with the surrounding monogenetic cones.
Gardeweg and Sellés (2012) assert that the volcanic products of Fui Sur
Group present basaltic to dacitic compositions, similar to MCVC,
Fig. 9. MCVC Schematic volcano-tectonic 3D model, showing the orientation
of flank and monogenetic cones: Alto Caunahue Group (yellow points), Ranquil whereas the cones of Fui Alto Group are constituted exclusively by
Group (red points), Fui Volcanic Group (light blue points) and Chanchán Vol­ poorly evolved basalts, with high content of MgO, Cr, Ni and incom­
canoes (blue points). Blue tones represent LOFS principal planes, purple NW- patible trace-elements (Ba, Nb, Th y La), achieving values 2 or 3 times
Striking structures and green tension cracks. Orange and red lines represent the higher than the volcanic complex. These basalts are very similar to the
principal feeder pathways. (For interpretation of the references to color in this basalts that compose the Chanchán Volcanoes, although their
figure legend, the reader is referred to the Web version of this article.) geochemistry indicates a provenance from a more primitive magma
(Gardeweg and Sellés, 2012). The basaltic composition of the mono­
5.2. Structural model genetic cones is different to the basalts that originate the magmas of the
MCVC, and because there is no interaction with differentiated magmas,
Our results and the studies of previous authors (e.g. Esparza, 2006; it is inferred that the pathways are independent and mutually isolated.
Moreno and Lara, 2007; Rodríguez et al., 1999) let us build a structural Only Fui Sur Group would be connected to the volcanic complex in
model of the MCVC basement (Fig. 9) that reflects the complexity of depth (Gardeweg and Sellés, 2012). Fui Alto Group and Chanchán
volcano-tectonic interaction and the key role of the intersection of Volcanoes are located within the surroundings of the main line of LOFS,
structural systems. which might favor the ascent of deep magmas, avoiding the interaction
The different structures (of first, second and third order) described in with the crust.
the study area are represented in a generalized and schematic way, The magmatic reservoir is indicated in a schematic way, since the
showing how the LOFS interacts with the NW-striking ATF and the NE- absence of seismic and geophysical data in the area prevents the
trending structures. A series of N–S to NNE-aligned scarps and valleys determination of its location and shape. However, the NW-elongated
are controlled by LOFS (e.g. Hervé, 1976), we propose that the main line anatomy of the volcanic edifice indicates that the reservoir presents an
of this structural system is part of a half-positive flower structure with elongated morphology in the same direction, embracing the ramp-and-
the Pirehueico-Reigolil fault, located towards the east of the study area flat model postulated by Sánchez et al. (2013) for systems dominated by
(Fig. 4), since both have kinematics with a reverse component, similar oblique-to-the-arc fault systems. Therefore, we propose that the
dip and are sub-parallel to each other. The orientation of the lake valleys magmatic reservoir is probably formed at the flat portions of the reverse
and the NW-elongated morphology of the MCVC are some of the evi­ faults similar to that suggested by González et al. (2009) for Pleistocene
dence of NW-striking inherited structures belonging to the ATF (e.g. volcanoes in the magmatic arc of the central Andes. The model is
Sielfeld et al., 2019). The NE-trending structures represented in the consistent with the volcanic products composite since it is perpendicular
model correspond to tension cracks, which owe their origin to the pre­ to the maximum-shortening axis orientation, impeding a direct path of
sent convergence angle, since this direction coincides with the value of the magmatic fluids towards the surface, consequently increasing their
σ1 calculated at regional scale for the study area. Based on field obser­ residence time and allowing their differentiation, and thus explaining
vations and proposals by other authors (Esparza, 2006; Lara and Mor­ the absence of basaltic compositions.
eno, 2004), suggested crosscuts relationships between structures with
different orientation. The presence and interaction of these 6. Conclusions
fault-fracture networks allows the circulation of magmatic and hydro­
thermal fluids, controlling the occurrence, spatial distribution and ge­ The analysis of structural and volcanic elements reflects the impor­
ometry of both the MCVC and its associated volcanic features. tance and complexity of their interaction at a regional and local scale, in
The existing regional stress field affects the orientation of dikes, and the domain of the LOFS and ATF near the MCVC. Our conclusions are
the predominant direction of feeder dyke trends is parallel or close-to- summarized as follows:
parallel to the maximum horizontal shortening at distal zones of the
main volcano edifice (Andersonian dykes) (e.g. Nakamura, 1977). The 1. The new structural insights provided in this work are concordant and
dikes measured in the area are dominantly NE-striking, parallel to the adapt quite well to those proposed by the previous authors and offer
σ1-trend calculated in this work, suggesting that dikes migrate through a global perspective that makes it possible to integrate and comple­
tension cracks. Meanwhile, dikes can also propagate through favorably ment the knowledge already available, establishing the structural
oriented inherited basement faults beneath the volcano edifice (non-­ basis for future work in the area.
Andersonian dykes) (Sielfeld et al., 2016). Parallel dike propagation

12
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

2. The fault-fracture networks allow the transport of magma through Dewey, J.F., Lamb, S.H., 1992. Active tectonics of the Andes. Tectonophys. v 205, 79–95.
Di Biase, F., 1976. Geología del grupo volcánico Mocho-Choshuenco, Provincia de
the crust and its storage in a magmatic reservoir. The unfavorable
Valdivia, Chile. I Congreso Geológico Chileno, Santiago, pp. F1–F17.
orientation of the reservoir improves the magma differentiation by Esparza, E., 2006. Control estructural en la zona del Complejo Volcánico Mocho-
fractional crystallization. Choshuenco, Andes del Sur. Master Thesis. Universidad de Concepción, Facultad de
3. The calculated regional stress field, with N61◦ E-trending σ1 axis and Ciencias Químicas, Chile.
Gardenweg, M., Sellés, D., 2012. Los volcanes Chanchán y el Grupo Volcánico Fui:
a N26◦ W-trending σ3 axis, is consistent with a unique tectonic volcanes monogenéticos al norte del CV Mocho-Choshuenco, Región de los Ríos, sur
regime, associated with the current oblique convergence. de Chile. XIII Congreso Geológico Chileno, Antofagasta, pp. 591–593.
4. Deformation driven by oblique convergence is partially partitioned, Glodny, J., Gräfe, K., Echtler, H., Rosenau, M., 2008. Mesozoic to Quaternary continental
margin dynamics in South-Central Chile (36–42 S): the apatite and zircon fission
where the margin-parallel component originates a major dextral track perspective. Int. J. Earth Sci. 97 (6), 1271–1291.
strike-slip fault system, the LOFS. At the study area, this tectonic González, G., Cembrano, J., Aron, F., Veloso, E.E., Shyu, J.B.H., 2009. Coeval
regime corresponds to a strike-slip-dominated transpression. compressional deformation and volcanism in the central Andes, case studies from
northern Chile (23 S–24 S). Tectonics 28 (6). https://doi.org/10.1029/
5. The main line of LOFS with Pirehueico-Reigolil fault conform a half- 2009TC002538.
positive flower structure, which is cross-cut by NW-trending inheri­ Hervé, M., 1976. Estudio geológico de la falla Liquiñe-Reloncaví en el área de Liquiñe:
ted structures and NE-striking tension cracks, forming the intersec­ Antecedentes de un movimiento transcurrente (Provincia de Valdivia). I Congreso
Geológico Chileno, Santiago, pp. B39–B56.
tion of structures that allows the volcanic activity. Hervé, M., 1977. Geología del área al este de Liquiñe, Provincia de Valdivia, Xa Región.
6. The diverse structural alignment and magma compositions of flank Master Thesis. Universidad de Chile, Departamento de Geología, Chile.
and monogenetic cones reflect the complexity of the plumbing sys­ Jarrard, R.D., 1986. Terrane motion by strike-slip faulting of forearc slivers. Geology 14,
780–783.
tem and the key role of the different structural systems present in the
Lange, D., Cembrano, J., Rietbrock, A., Haberland, C., Dahm, T., Bataille, K., 2008. First
area. seismic record for intra-arc strike-slip tectonics along the Liquiñe–Ofqui fault zone at
7. The NE-striking dike distribution is consistent with the dominant the obliquely convergent plate margin of the southern Andes. Tectonophysics 455,
regional stress field, nonetheless NW-trending pre-existing weakness 14–24. https://doi.org/10.1016/j.tecto.2008.04.014.
Lara, L., Moreno, H., 2004. Geología del área Liquiñe-Neltume, Regiones de La Araucanía
zones were reactivated, contributing to the construction of the vol­ y de Los Lagos. Servicio Nacional de Geología y Minería Chile, Carta Geológica de
canic edifice. Chile. Serie Geología Básica, No. 83, 1–23.
8. The absence of seismic and geophysical knowledge in the study area Lara, L.E., Naranjo, J.A., Moreno, H., 2004. Rhyodacitic fissure eruption in southern
Andes (Cordón Caulle; 40.5◦ S) after the 1960 (Mw: 9.5) Chilean earthquake: a
makes it difficult to characterize the magmatic reservoir, although structural interpretation. J. Volcanol. Geoth. Res. 138, 127–138.
the ramp-and-flat model fits fairly well as a first approach. Lara, L.E., Lavenu, A., Cembrano, J., Rodríguez, C., 2006. Structural controls of
volcanism in transversal chains: resheared faults and neotectonics in the Cordón
Caulle– Puyehue area (40.5◦ S), Southern Andes. J. Volcanol. Geoth. Res. 158, 70–86.
Declaration of competing interest https://doi.org/10.1016/j.jvolgeores.2006.04.017.
Lara, L.E., Cembrano, J., Lavenu, A., 2008. Quaternary vertical displacement along the
Liquiñe–Ofqui fault zone: differential uplift and coeval volcanism in the southern
The authors declare that they have no known competing financial
Andes? Int. Geol. Rev. 50, 975–993. https://doi.org/10.2747/0020-6814.50.11.975.
interests or personal relationships that could have appeared to influence Legrand, D., Barrientos, S., Bataille, K., Cembrano, J., Pavez, A., 2011. The fluid-driven
the work reported in this paper. tectonic swarm of Aysen Fjord, Chile (2007) associated with two earthquakes (Mw =
6.1 and Mw = 6.2) within the Liquiñe–Ofqui fault zone. Continent. Shelf Res. 31,
154–161. https://doi.org/10.1016/j.csr.2010.05.008.
Acknowledgments López-Escobar, L., Cembrano, J., Moreno, H., 1995. Geochemistry and tectonics of the
Chilean southern Andes basaltic quaternary volcanism. Andean Geol. 22, 219–234.
This work was supported by FONDECYT Project 1151175 Active Marrett, R., Allmendinger, R.W., 1990. Kinematic analysis of fault-slip data. J. Struct.
Geol. 12, 973–986.
Tectonism and Volcanism at the Southern Andes (ACT&VO-SA) and McMillan, N., Harmon, R., Moorbath, S., López-Escobar, L., Strong, D.F., 1989. Crustal
Millenium Nucleus NC160025 The Seismic Cycle along Subduction sources involved in continental arc magmatism: a case study of volcan Mocho-
Zones CYCLO. We are greatly appreciative to Ernesto Cofré and Mario Choshuenco, southern Chile. Geology 17, 1152–1156.
Melnick, D., Folguera, A., Ramos, V., 2006. Structural control on arc volcanism: the
González for invaluable assistance in the field. We sincerely thank Caviahue-Copahue complex, Central to Patagonian Andes transition (38◦ S). J. South
Joaquín Cortés, Paulina Mejías, Amapola Albornoz and Francisco García Am. Earth Sci. 22, 66–88. https://doi.org/10.1016/j.jsames.2006.08.008.
for their comments, corrections and suggestions that greatly improved Melnick, D., Bookhagen, B., Strecker, M.R., Echtler, H.P., 2009. Segmentation of
megathrust rupture zones from forearc deformation patterns over hundreds to
the manuscript. millions of years, Arauco peninsula, Chile. J. Geophys. Res. 114, B01407. https://
doi.org/10.1029/2008JB005788.
References Mora, C., Comte, D., Russo, R., Gallego, A., Mocanu, V., 2010. Aysén seismic swarm
(January 2007) in southern Chile: analysis using joint hypocentral determination.
J. Seismol. 14, 683–691. https://doi.org/10.1007/s10950-010-9190-y.
Allmendinger, R.W., Cardozo, N.C., Fisher, D., 2012. Structural Geology Algorithms:
Moreno, H., Lara, L., 2007. Geología del Complejo Volcánico Mocho-Choshuenco, Región
Vectors & Tensors. Cambridge University Press, Cambridge, England.
de Los Ríos. Servicio Nacional de Geología y Minería Chile, Carta Geológica de Chile.
Angermann, D., Klotz, J., Reigber, C., 1999. Space-geodetic estimation of the
Serie Geología Básica, No. 107, 1–27.
Nazca–South America Euler vector. Earth Planet Sci. Lett. 171, 329–334. https://doi.
Munizaga, F., Hervé, F., Drake, R., Pankhurst, R., Brook, M., Snelling, N., 1988.
org/10.1016/S0012-821X(99)00173-9.
Geochronology of the Lake region of south-central Chile (39◦ -42◦ S): preliminary
Arancibia, G., Cembrano, J., Lavenu, A., 1999. Transpresión dextral y partición de la
results. J. South Am. Earth Sci. 1 (3), 309–316.
deformación en la Zona de Falla Liquiñe–Ofqui, Aisén, Chile (44–45◦ S). Rev. Geol.
Nakamura, K., 1977. Volcanoes as possible indicators of tectonic stress orientation —
Chile 26. https://doi.org/10.4067/S0716-02081999000100001.
principle and proposal. J. Volcanol. Geoth. Res. 2, 1–16. https://doi.org/10.1007/
Bott, M.H.P., 1959. The mechanics of oblique slip faulting. Geol. Mag. 96, 109–117.
BF01637099.
Campos, A., Moreno, H., Muñoz, J., Antinao, J., Clayton, J., Martin, M., 1998. Área de
Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca (farellon) and South
Futrono-Lago Ranco, región de Los Lagos. Servicio Nacional de Geología y Minería
American plates since late Cretaceous time. Tectonics 6, 233–248.
Chile. Mapas Geológicos. No. 8, scale 1:100,000.
Pérez, S., 2005. Volcanismo Explosivo Postglacial del Complejo Volcánico Mocho-
Cembrano, J., Hervé, F., 1993. The Liquiñe–Ofqui Fault Zone: A Major Cenozoic Strike
Choshuenco. Master Thesis. Universidad de Concepción, Facultad de Ciencias
Slip Duplex in the Southern Andes. Second ISAG, Oxford, UK, pp. 175–178.
Químicas, Chile.
Cembrano, J., Schermer, E., Lavenu, A., Sanhueza, A., 2000. Contrasting nature of
Pérez-Estay, N., Yáñez, G., Crempien, J., Roquer, T., Cembrano, J., Valdenegro, P.,
deformation along an intra-arc shear zone, the Liquiñe–Ofqui fault zone, southern
Aravena, D., Arancibia, G., Morata, D., 2020. Seismicity in a transpressional volcanic
Chilean Andes. Tectonophysics 319 (2), 129–149. https://doi.org/10.1016/S0040-
arc: the Liquiñe-Ofqui fault system in the Puyuhuapi area, southern Andes, Chile
1951(99)00321-2.
(44◦ S). Tectonics 39 (11), e2020TC006391. https://doi.org/10.1029/
Cembrano, J., Lara, L., 2009. The link between volcanism and tectonics in the southern
2020TC006391.
volcanic zone of the Chilean Andes: a review. Tectonophysics 471, 96–113. https://
Pérez-Flores, P., Cembrano, J., Sánchez, P., Veloso, E., Arancibia, G., Roquer, T., 2016.
doi.org/10.1016/j.tecto.2009.02.038.
Tectonics, magmatism and paleo-fluid distribution in a strike-slip setting: insights
Cembrano, J., Hervé, F., Lavenu, A., 1996. The Liquiñe Ofqui fault zone: a long-lived
from the northern termination of the Liquiñe-Ofqui fault System, Chile.
intra-arc fault system in southern Chile. Tectonophysics 256, 55–56.
Tectonophysics 680, 192–210. https://doi.org/10.1016/j.tecto.2016.05.016.
Corazzato, C., Tibaldi, A., 2006. Fracture control on type, morphology and distribution of
Petit, J.P., 1987. Criteria for the sense of movement on fault surfaces in brittle rocks.
parasitic volcanic cones: an example from Mt. Etna, Italy. J. Volcanol. Geoth. Res.
J. Struct. Geol. 9 (5–6), 597–608.
158, 177–194. https://doi.org/10.1016/j.jvolgeores.2006.04.018.

13
C. Peña et al. Journal of South American Earth Sciences 111 (2021) 103435

Piquer, J., Berry, R.F., Scott, R.J., Cooke, D.R., 2016. Arc-oblique fault systems: their role Sernageomin, S., 2003. Mapa Geológico de Chile: versión digital. Servicio Nacional de
in the Cenozoic structural evolution and metallogenesis of the Andes of central Chile. Geología y Minería, Publicación Geológica Digital 4.
J. Struct. Geol. 89, 101–117. https://doi.org/10.1016/j.jsg.2016.05.008. Sibson, R.H., 1996. Structural permeability of fluid-driven fault-fracture meshes.
Radic, J.P., 2010. Las cuencas cenozoicas y su control en el volcanismo de los Complejos J. Struct. Geol. 18, 1031–1042. https://doi.org/10.1016/0191-8141(96)00032-6.
Nevados de Chillán y Copahue–Callaqui (Andes del Sur, 36–39 ◦ S). Andean Geol. Sibson, R.H., 2003. Brittle-failure controls on maximum sustainable overpressure in
37, 220–246. different tectonic regimes. Am. Assoc. Petrol. Geol. Bull. 87, 901–908. https://doi.
Rawson, H., Naranjo, J.A., Smith, V., Fontijn, K., Pyle, D.M., Mather, T.A., Moreno, H., org/10.1306/01290300181.
2015. The frequency and magnitude of post-glacial explosive eruptions at Volcán Sielfeld, G., Cembrano, J., Lara, L., 2016. Transtension driving volcano-edifice anatomy:
Mocho-Choshuenco, southern Chile. J. Volcanol. Geoth. Res. 299, 103–129, 10 insights from Andean transverse-to-the-orogen tectonic domains. Quat. Int. (2016
.1016/j.jvolgeores .2015 .04 .003. 1–17. https://doi.org/10.1016/j.quaint.2016.01.002.
Rawson, H., Pyle, D.M., Mather, T.A., Smith, V.C., Fontijn, K., Lachwycz, S.M., Sielfeld, G., Lange, D., Cembrano, J., 2019. Intra-arc crustal seismicity: seismotectonic
Naranjo, J.A., 2016. The magmatic and eruptive response of arc volcanoes to implications for the southern Andes volcanic zone, Chile. Tectonics 38 (2), 552–578.
deglaciation: insights from southern Chile. Geology 44 (4), 251–254. https://doi. https://doi.org/10.1029/2018TC004985.
org/10.1130/G37504.1. Stanton-Yonge Sesnic, A., 2016. Tectonic Role of Margin-Parallel and Margin-Transverse
Riedel, W., 1929. Zur mechanik geologischer Brucherscheinungen. Zentral blatt fuer Faults during Oblique Subduction at the Southern Volcanic Zone of the Andes.
Mineralogie, Geologie und Palaeontologie, pp. 354–368 (Stuttgart). https://doi.org/10.1002/2016TC004226.
Ritz, J.-F., 1994. Determining the slip vector by graphical construction: use of a Tibaldi, A., Bonali, F.L., Corazzato, C., 2017. Structural control on volcanoes and magma
simplified representation of the stress tensor. J. Struct. Geol. 16, 737–741. paths from local-to orogen-scale: the central Andes case. Tectonophysics 699, 16–41.
Rodríguez, C., Pérez, Y., Moreno, H., Clayton, J., Antinao, J., Duhart, P., Martin, M., https://doi.org/10.1016/j.tecto.2017.01.005.
1999. Área de Panguipulli-Riñihue, región de Los Lagos. Servicio Nacional de Veloso, E., Gomila, R., González, R., Cembrano, J., Jensen, E., Arancibia, G., 2015. Stress
Geología y Minería Chile. Mapas geológicos, No. 10, scale 1:100,000. fields recorded on large-scale strike-slip fault systems: effects on the tectonic
Roquer, T., Veloso, E.A., Arancibia, G., Molina Piernas, E., Sepúlveda, J., Rowland, J.V., evolution of crustal slivers during oblique subduction. Tectonophysics 664,
Elizalde, J., 2019. Maximum sustainable fluid pressure within a strike-slip tectonic 244–255. https://doi.org/10.1016/j.tecto.2015.09.022.
setting: preliminary results from the Andean Transverse Faults, Southern Volcanic Veloso, E.E., Roquer, T., Arancibia, G., Molina Piernas, E., Sepúlveda, J., Rowland, J.V.,
Zone of the Andes (39◦ S). In: AGU Fall Meeting Abstracts, vol. 2019. MR23G-0187. Morata, D., 2019. Tectonic constraints leading to hydrothermal fluid flow: insights
Rosenau, M., 2004. Tectonics of the Southern Andean Intra-arc Zone (38◦ -42◦ S). Ph.D. from the Liquiñe-Ofqui fault system and the andean transverse faults, southern
thesis. Freie University, Germany. volcanic zone of the Andes (39◦ S). In: AGU Fall Meeting Abstracts, vol. 2019.
Rosenau, M., Melnick, D., Echtler, H., 2006. Kinematic constraints on intra-arc shear and MR23G-0190.
strain partitioning in the southern Andes between 38 ◦ S and 42 ◦ S latitude. Wallace, R.E., 1951. Geometry of shearing stress and relation of faulting. J. Geol. 59,
Tectonics 25, 1–16. https://doi.org/10.1029/2005TC001943. 118–130.
Sánchez, P., Pérez-Flores, P., Arancibia, G., Cembrano, J., Reich, M., 2013. Crustal Yamaji, A., 2000. The multiple inverse method: a new technique to separate stresses from
deformation effects on the chemical evolution of geothermal systems: the intra-arc heterogeneous fault-slip data. J. Struct. Geol. 22, 441–452. https://doi.org/
Liquiñe–Ofqui Fault System, Southern Andes. Int. Geol. Rev. 55, 1384–1400. https:// 10.1016/S0191-8141(99)00163-7.
doi.org/10.1080/00206814.2013.775731.

14

You might also like