You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/248816712

Crustal-scale structural architecture in central Chile based on seismicity and


surface geology: Implications for Andean mountain building

Article in Tectonics · June 2010


DOI: 10.1029/2009TC002480

CITATIONS READS

167 991

8 authors, including:

Marcelo Farias Diana Comte


University of Chile University of Chile
93 PUBLICATIONS 2,411 CITATIONS 164 PUBLICATIONS 4,154 CITATIONS

SEE PROFILE SEE PROFILE

Reynaldo Charrier J. Martinod


University of Chile Université Savoie Mont Blanc
197 PUBLICATIONS 6,831 CITATIONS 181 PUBLICATIONS 10,249 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by J. Martinod on 29 May 2015.

The user has requested enhancement of the downloaded file.


TECTONICS, VOL. 29, TC3006, doi:10.1029/2009TC002480, 2010
Click
Here
for
Full
Article
Crustal‐scale structural architecture in central Chile based
on seismicity and surface geology: Implications for Andean
mountain building
Marcelo Farías,1,2,3 Diana Comte,2 Reynaldo Charrier,1 Joseph Martinod,3
Claire David,1,2,3 Andrés Tassara,2,4 Felipe Tapia,1 and Andrés Fock1,5
Received 26 February 2009; revised 21 October 2009; accepted 23 December 2009; published 29 May 2010.
[1] We document a crustal‐scale structural model for geology: Implications for Andean mountain building, Tectonics,
the central Chile Andes based on seismicity and sur- 29, TC3006, doi:10.1029/2009TC002480.
face geology, which consists in a major east verging
ramp‐detachment structure connecting the subduction 1. Introduction
zone with the cordillera. The ramp rises from the sub- [2] The Andes are the Earth’s longest and highest active
ducting slab at ∼60 km depth to 15–20 km below the mountain chain formed in an ocean‐continent subduction
western edge of the cordillera, extending eastward as a margin. It is widely recognized that the elevation and crustal
10 km depth flat detachment. This structure plays a thickness of this mountain range have been mainly produced
fundamental role in the Andean orogenesis because by crustal shortening because of the interaction between the
most of the shortening has been accommodated by subducting oceanic Nazca plate and the overriding conti-
structures rooted in it and allows the distribution of nental South American plate. In the shallow levels of the
crustal thickening in a “simple shear deformation continental crust, most of the shortening has been accom-
mode.” Indeed, despite shortening distribution being modated in the eastern flank of the range in several fold‐
and‐thrust belts (Figure 1). Therefore, deformation occurs in
very asymmetric (∼16 km versus ∼70 km in the western
the hanging walls of large east verging detachments in
and eastern side, respectively), the western side is which ramps are predominantly east verging, showing that
higher and thicker than what is expected. Yield strength tectonic transport in the Andes preferentially goes to the east
envelopes show strong rheological control on this [e.g., Isacks, 1988; Allmendinger et al., 1990; Allmendinger
structure. Vp and Vp/Vs variations in the upper mantle and Gubbels, 1996; Ramos et al., 1996; Allmendinger and
and in the deepest limit of the seismogenic interplate Zapata, 2000; Cristallini and Ramos, 2000; Giambiagi
contact mark the intersection of the ramp with the slab, and Ramos, 2002; McQuarrie, 2002; Giambiagi et al.,
which coincides with the blueschist‐eclogite transition. 2003a; Ramos et al., 2004; Arriagada et al., 2006; Vergés
Therefore, subduction processes would control the et al., 2007; McQuarrie et al., 2008]. Therefore, detach-
depth where the major east verging structure may ments related to eastern peripheral deformation belts appear
merge with the slab. Such a ramp‐flat structure is to be the most relevant structures controlling the Andean
orogeny, as occurs in many modern and old mountain
observed in other parts of the Chilean margin; hence,
chains [e.g., Cook and Varsek, 1994].
it seems to be a first‐order feature in the Andean sub- [3] Shortening accommodated in the eastern flank of the
duction zone. This structure delimitates upward the Andes has been widely studied, usually favored by the
rocks, transmitting part of the plate convergence stress sedimentary constitution of the rocks involved in the fold
from the plate interface, and controls mountain‐building and thrust belts, as well as by the great database of geo-
tectonics, thus playing a key role in the Andean orogeny. physical data and core drilling performed by oil exploration
Citation: Farías, M., D. Comte, R. Charrier, J. Martinod, C. David, companies. These factors have allowed determining the
A. Tassara, F. Tapia, and A. Fock (2010), Crustal‐scale structural depth and geometry of detachment levels, as well as
architecture in central Chile based on seismicity and surface performing accurate temporal and geometrical reconstruc-
tion of deformation. The relevance of determining the depth
and geometry of detachments resides in the fact that they
1
Departamento de Geología, FCFM, Universidad de Chile, Santiago, define the boundary conditions for structural restoration.
Chile.
2
The lack of a well‐constrained determination of the detach-
Departamento de Geofísica, FCFM, Universidad de Chile, Santiago, ment (in which superficial structures are rooted) may lead to
Chile.
3
LMTG, CNRS‐IRD‐Université de Toulouse, Toulouse, France. overestimate or underestimate the amount of shortening;
4
Now at Departamento de Ciencias de la Tierra, Universidad de therefore, the location of this rooting structure is fundamental.
Concepción, Concepción, Chile. [4] The westward prolongation of detachments into the
5
Now at Sociedad Química y Minera S.A., Antofagasta, Chile. arc and fore‐arc region, in which plate interaction occurs, is
poorly understood. This situation is mainly produced by the
Copyright 2010 by the American Geophysical Union. fact that shortening is very much smaller here, thrusts verge
0278‐7407/10/2009TC002480
mainly to the west (in northern Chile, from Muñoz and

TC3006 1 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 1. (a) Tectonic framework of the Andean margin. (b) Maximum elevation and Moho depth from
32°C to 37°S. Elevations calculated from SRTM90m DEM, and crustal thickness is after Tassara et al.
[2006]. (c) Main tectonic and morphological features of the Andes of central Chile and western Argentina.
Seismologic stations of the permanent network of the University of Chile (white inverted triangles) and
temporary network deployed during January–April 2004 (dark inverted triangles) are shown. Grid in
Figure 1c corresponds to the region and cells in which tomography was performed. Absolute plate motion
velocity is after Gripp and Gordon [2002]. Focal mechanisms of the two greatest shallow crust earth-
quakes of the last years are those calculated by Harvard CMT.

2 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Charrier [1996], Victor et al. [2004], Farías et al. [2005], Andes (Figure 1): the Aroma earthquake (24 July 2001, Mw =
and García and Hérail [2005] and in central Chile, from 6.3) at 19°30′S [Farías et al., 2005; Legrand et al., 2007],
Ramos et al. [1996], Charrier et al. [2005], and Fock et al. the Altos del Teno event (28 August 2004, Mw = 6.5) at
[2006]), numerous intrusive bodies perturb the structure and 35°10′S [Farías et al., 2006; Comte et al., 2008], and the
the lithological continuity, and there is no evidence for the more recent Aysén event (21 April 2007, ML = 6.2) at
downdip prolongation of these structures nor their rooting 45°15′S. These earthquakes, as well as most of the detected
detachments. These circumstances have led to a poor knowl- shallow seismicity, occurred in association with structural
edge on the structural connection between the mountain belt systems very relevant during the Cenozoic mountain
and the subduction zone, as well as on the mechanisms of building [Barrientos et al., 2004; Farías, 2007; Farías et
stress‐strain transfer from the interplate contact toward the al., 2005, 2006; Comte et al., 2008; Lange et al., 2008].
continent. Although some studies have attempted to balance Shallow crustal seismicity occurring in relationship to great
the forces released in the subduction zone and those trans- fault systems emphasizes the relevance of this kind of
ferred to the orogen [e.g., Isacks, 1988; Kono et al., 1989; earthquakes in order to explore the relevance of these
Lamb and Davis, 2003; Yáñez and Cembrano, 2004; Lamb, structures during mountain building, their subsurface pro-
2006], these models did not considered the geometry of the longation into the continental lithosphere, and their associ-
structures involved in the resulting strain transfer. ation with main detachments and links to the subduction
[5] In the Andes of central Chile (33–35°S), the con- interface.
struction of well‐constrained crustal‐scale cross sections has [8] In this study, we integrate seismological and geolog-
presented several obstacles, such as the difficulties to ical data obtained in the Andes of central Chile (33–35°S) in
identify stratigraphic markers and to determine ages of order to address two main questions: (1) which is the fore‐
deposition and deformation. Among other reasons, this sit- arc–arc structural array that relates tectonics in the subduc-
uation resides in the fact that outcropping sequences in this tion zone to the mountain belt, and (2) which has been the
region are predominantly of an igneous composition, they contribution of this structural array for the construction of
exhibit large lateral variations in facies and thickness, and the mountain belt. This study is based on the analysis of
there is a pervasive low‐grade metamorphism affecting them seismicity recorded by the permanent network of the Seis-
[e.g., Levi et al., 1989; Vergara et al., 1993; Muñoz et al., mologic Survey at the Universidad de Chile and a temporary
2006]. network deployed along one of the largest and longest
[6] However, the knowledge on the tectonic evolution of structural systems in the high Andes of central Chile. We
the western side of the range in this region has increased correlate these data with surface geology in order to infer the
during the last years due to a more accurate mapping and crustal‐scale structural architecture across the whole
dating of tectonic and depositional events. This has been mountain range. Based on our findings, we show a major
possible due to a more systematic search for unaltered levels detachment running below the entire mountain belt, which
[Fuentes, 2004; Muñoz et al., 2006], the use of more pen- would be connected to the subduction zone through a ramp
etrative radioisotopic dating [Charrier et al., 1996, 2002, that cuts across the fore‐arc lithosphere wedge. We finally
2005; Fock et al., 2006; Montecinos et al., 2008; Oliveros et propose a qualitative approach to the mechanisms control-
al., 2008], the discovery of abundant Cenozoic mammal ling the transference of deformation from the interplate
fauna [Wyss et al., 1990; Flynn et al., 1995; Croft et al., interface toward the mountain belt, which could be very
2003; Flynn et al., 2003], and the use of quantitative geo- significant for understanding orogenic mechanisms not only
morphology [Farías et al., 2008a] and low‐temperature in the study region, but also along the Chilean subduction
thermochronology [Farías et al., 2008a; Maksaev et al., margin and even in other subduction orogens.
2009]. These new studies have allowed to establish that
the contractive development of the Andes in central Chile 2. Tectonic and Geological Settings
occurred fundamentally since the latest Oligocene–earliest
Miocene [e.g., Charrier et al., 2002, 2005] (similar to the 2.1. Generalities
central Argentinean Andes [Ramos et al., 1996; Giambiagi [9] The subduction of the oceanic Nazca (former Farallon)
and Ramos, 2002; Giambiagi et al., 2003a; Ramos et al., plate beneath the South American continent is the main
2004]), and that the main stage of surface uplift occurred tectonic process along the Andean margin since Jurassic
during the late Miocene–early Pliocene [Farías et al., times [e.g., see Mpodozis and Ramos, 1989; Charrier et al.,
2008a]. 2007, and references therein]. Late Cenozoic evolution of
[7] Another source of data that has been used to under- convergence rates shows a strong diminution to its current
stand the fore‐arc evolution comes from the seismology. rate (N78°E, ∼8 cm yr−1 relative plate motion; Figure 1a).
Although most of the studies in this discipline performed in Current absolute plate motion relative to hot spots frame for
the Andean region have paid attention to the subduction the South American and Nazca plates are 4.8 and 3.2 cm yr−1,
zone (nucleation and propagation of megathrust earthquakes respectively [Gripp and Gordon, 2002].
[M ≥ 8]), last years improvements of permanent seismo- [10] The Chilean‐Pampean flat‐slab subduction region
logical networks and the deployment of temporary networks (27°S–33°S), where slab dip is <10° between 100 and 150 km
have permitted detecting abundant shallow (<20 km depth) depth [e.g., Jordan et al., 1983; Pardo et al., 2002],
crustal seismicity beneath the Chilean fore‐arc–arc region. represents a major along‐strike change on the Andean
On the recorded data, three events stand out in the Chilean subduction system. North and south of this segment, the dip

3 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

of the slab is nearly 30°E, being the typical example of the [14] The eastern Principal Cordillera domain is charac-
“Chilean‐type subduction” (in the sense of Uyeda and terized by the presence of Mesozoic sequences of predom-
Kanamori [1979]). This segmentation has been interpreted inant sedimentary composition. These units were deposited
as the result of buoyancy forces exerted by the subduction of into the northern sector of the Neuquén back‐arc basin,
the Juan Fernández ridge at 32.5°S (Figure 1) [Pilger, 1981; which developed in this zone during most of the Mesozoic
Nur and Ben‐Avraham, 1981; Gutscher et al., 2000; Yáñez [Uliana et al., 1989; Giambiagi et al., 2003b, and references
et al., 2001]. In addition, some distinctive features appear at therein]. This part of the Principal Cordillera has accom-
approximately 33°S (Figure 1): (1) strike change (Maipo modated most of the shortening in this region since ∼16 Ma
Orocline in this work) of the trench, coast, and mountain belt [Giambiagi and Ramos, 2002; Giambiagi et al., 2003a;
from N–S, northward, to NNE–SSW, southward; (2) decrease Ramos et al., 2004] when shortening almost ended in western
of magmatic activity in the late Miocene until its extinction sectors.
at 1.9 Ma in the flat‐slab region [Kay et al., 1991; Ramos et [15] After 8.5 Ma, just east from the eastern Principal
al., 2002]; (3) decrease in mountain elevation and crustal Cordillera, the uplift of Proterozoic–lower Triassic basement
thickness south of 33°S; (4) increase in flexural rigidity by the means of the inversion of rift‐related high‐angle faults
south of 34°S along the mountain belt [e.g., Pérez‐Gussinyé concluded in the raise of the Frontal Cordillera [Giambiagi
et al., 2007; Tassara et al., 2007]; (5) southward segmen- and Ramos, 2002; Giambiagi et al., 2003a; Ramos et al.,
tation of the mountain range by a longitudinal valley 2004]. Simultaneously or shortly after, high‐angle out‐of‐
(Central Depression); and (6) absence of great basement‐ sequence reverse faults deformed the eastern Abanico basin
involved ranges as the Frontal Cordillera, Pampean Ranges, and the eastern Principal Cordillera [Giambiagi and Ramos,
and Precordillera south of 34°15′S. 2002; Giambiagi et al., 2003a; Fock et al., 2006]. At circa
4 Ma, shortening migrated farther east to the foreland
[Giambiagi et al., 2003a], and the Chilean belt reached most
2.2. Geologic Evolution
of its present‐day elevation, diminishing drastically the
[11] The Andean region between 33°S and 35°S comprises uplift rates in the Chilean belt since then in about 1 order of
five major continental morphostructural units (Figures 1 magnitude from 1 to 2 mm yr−1 during the late Miocene to
and 2), from west to east: Coastal Cordillera, Central 0.1 mm yr−1 [Farías et al., 2008a].
Depression, Principal Cordillera (in this work, subdivided [16] Magmatic history in the central Chile Andes is
into a western, central and eastern Principal Cordillera, see mainly related to an almost continuous eastward migration
Figure 2), Frontal Cordillera (not present south of 34°15′S), of the arc since the Jurassic [e.g., Kay et al., 2005; Charrier
and active the foreland. et al., 2007]. Coeval to the beginning of contractive tec-
[12] The Coastal Cordillera consists of a late Paleozoic‐ tonics in the early Miocene, many granitic intrusions
Triassic basement in its western flank and east dipping emplaced in the westernmost Principal Cordillera (lower
Jurassic to Cretaceous intra‐arc sequences in its eastern Miocene intrusive belt, Figure 2) [e.g., Kay et al., 2005;
flank, whose series extends to the middle of the Central Charrier et al., 2007]. Shortly after, the arc migrated
Depression Thomas [1958] (see Figure 2). slightly to the east, as evidenced by the volcanic rocks of the
[13] From the eastern half of the Central Depression to the Farellones Formation, which represent the locus of the arc
central Principal Cordillera, an extensional basin developed until the Langhian in a zone that did not accommodate
during late Eocene to late Oligocene times (Abanico basin), significant deformation. At the end of the volcanic pulses
which began to be inverted in the latest Oligocene–earliest related to this formation, the magmatic arc migrated again to
Miocene [Godoy et al., 1999; Charrier et al., 2002]. This basin the east, intruding the eastern flank of the Farellones and
was filled by the predominantly volcanic‐volcanoclastic Abanico formations. This magmatic activity formed a long
Abanico Formation. In early stages of inversion, folding and intrusive chain that was active in the late Miocene between
high‐angle reverse faulting mainly affected both basin edges 13 and 7 Ma [e.g., Kurtz et al., 1997; Kay et al., 2005;
[Fock et al., 2006]. In the middle zone of the former basin, Charrier et al., 2007] (hereafter we will refer to this chain as
the Farellones Formation, predominantly volcanic, was “late Miocene intrusive belt”). After this event, volcanic/
deposited during early to middle Miocene times. This unit is magmatic activity declined, but some pulses shifted to the
generally mildly folded, excepting near the edges of the west, forming the porphyry copper deposits of the El Teniente
Abanico basin, where overlays the Abanico Formation either and Río Blanco–Los Bronces. This magmatic arc was active
unconformably or developing growth strata in its lower between ∼9 and ∼4 Ma [e.g., Maksaev et al., 2004; Deckart
layers (older than 16 Ma, see Figure 3) [Charrier et al., 2002, et al., 2005] being coeval with the out‐of‐sequence event
2005; Fock et al., 2006]. After ∼16 Ma, contractive defor- and the uplift of the cordillera [Farías et al., 2008a]. After
mation migrated toward the eastern Principal Cordillera. this time, magmatism migrated again eastward to the eastern

Figure 2. (a) Simplified geological map and (b) cross sections of the Andes of central Chile and western Argentina. Only
main inverse faults active during the Neogene are plotted. El Diablo and Las Leñas faults belong to the El Fierro fault sys-
tem. Modified after Servicio Geológico Minero Argentino [1997], Godoy et al. [1999], Charrier et al. [2002], Servicio
Nacional de Geología y Minería [2002], Giambiagi et al. [2003a], and Fock et al. [2006]. Cross sections only show
what can be observed in the field. AFTB, Aconcagua fold‐and‐thrust belt; MFTB, Malargüe fold‐and‐thrust belt.

4 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 2

5 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 3

6 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Principal Cordillera, place in which the present‐day volcanic that exhumation has been very low and not capable of raise
arc is located. the 100–120°C isotherm since the Abanico Formation
deposition (that is, no more than 3–4 km considering a
thermal gradient of 25–40°C km−1) (see dating in the work
3. Structural Features of the Mountain Belt in by Farías et al. [2008a]).
Central Chile and Western Central Argentina [20] Just east from the mountain front, a series of syn-
clines and anticlines were developed between circa 22 and
[17] We present two cross section (Figure 2b) in which we 16 Ma, which is evidenced by growth strata developed in
integrate previous works (Thiele [1980], Giambiagi et al. the lower layers of the Farellones Formation [Fock et al.,
[2003a], and Fock et al. [2006] for the Santiago section, 2006] (Figures 2b, 3a, and 3b). To the east, some folds
P1 in Figure 2b; Charrier [1981], Godoy et al. [1999], and and faults deform subtly the Cenozoic sequence, showing a
Charrier et al. [2002, 2005] for the Cachapoal section, P2 in predominant east vergence in both sections (Figure 2b).
Figure 2b) with new data obtained in this study. In the Particularly, the San José fold (Figures 2b and 3b) stands out
following paragraph, we describe the major structural and because in its core it is developed an east vergent duplex
geological features along these sections that can be con- involving sedimentary layers of the Abanico Formation.
sidered representatives of the study region because of their Above this zone, the Farellones Formation presents growth
location. strata developed prior 16 Ma [Fock et al., 2006].
3.1. Eastern Central Depression and Western Principal
Cordillera 3.2. Central Principal Cordillera
[18] Only Cenozoic deposits crop out in this sector [21] As in the previous sector, only Cenozoic deposits
(Abanico and Farellones formations, and intrusive bodies). crop out here. These units are bounded to the east by east
The western edge of this region is characterized by east verging faults that produce the overriding of the Abanico
dipping partially inverted normal faults (Los Angeles fault Formation on the Mesozoic sequences of the eastern Prin-
system according to Carter and Aguirre [1965], Infiernillo cipal Cordillera. These faults would constitute the eastern
fault according to Fock et al. [2006]) (Figure 2) in which edge of the Abanico basin, and they can be traced over more
Cenozoic units override Mesozoic sequences. This fault than 300 km along strike (Figures 2a and 3c). The main
system would be the western edge of the Abanico Basin faults of this system received different names according to
[Fock et al., 2006]. their latitude: El Diablo fault at 33°45′S [Fock et al., 2006],
[19] The western edge of the Principal Cordillera is defined Las Leñas–Espinoza fault at 34°30′S [Charrier et al., 2002],
by a west vergent reverse fault system (San Ramón–Pocuro and El Fierro fault at approximately 35°S [Davidson and
fault and its southward prolongation [Thiele, 1980; Charrier Vicente, 1973]. Hereafter, we will name these faults as El
et al., 2005; Fock et al., 2006; Rauld et al., 2006; Armijo et Fierro fault system.
al., 2010] (Figure 2). On the basis of geomorphologic [22] In the central Principal Cordillera, very significant
markers, the cordilleran front structural system would have folds and thrusts affect the Abanico Formation. Folds
accommodated a maximum vertical throw of 0.7–1.1 km exhibit maximum amplitudes over 2–3 km. Folding in this
since the late Miocene at the latitude of Santiago, and 600– sector can be mostly related to the development of east
800 m at 35°S [Farías et al., 2008a]). This structural system verging fault‐bend and fault‐propagation folds (Figures 2b,
has not accommodated much more throw since basin 3c, and 3d). In this domain, the geometry of folds and
inversion than the present‐day elevation of this zone. This related faults suggest tectonic inversion of both Cenozoic
because apatite fission tracks dating coincides with strati- and Mesozoic extensional basins (Figures 3c and 3e). Folds
graphic ages in the rocks of the Abanico Formation and are usually cut by break‐through reverse faults, as well as by
subsidence related to the orogenic load in the Central out‐of‐sequence faults and back thrusts likely related to the
Depression has been almost negligible (no more than 500 m late Miocene–early Pliocene event [Fock et al., 2006] or at
in Santiago [Farías et al., 2008b]. The former reason means least posterior to the basin inversion. Among them, the

Figure 3. (a) Growth strata developed in the eastern limb of the San Ramón anticline. (b) East vergent fold and thrusts
related to the San José anticline, immediately south of Maipo river. Deformation in less competent layers of the Abanico
Formation (sedimentary layers) is produced by progressive development of east vergent thrusts, which also produced east-
ward growth strata in the Farellones Formation during the lower Miocene and the progressive increase of slope in the west-
ern limb of the fold. Asterisk indicates zircon SHRIMP U‐Pb age [Fock et al., 2006]. (c) Structure in the eastern flank of the
Abanico basin in the Volcán valley (upper course of the Maipo drainage basin). Lo Valdés Formation is a Tithonean‐
Neocomian marine unit and the Colimapu Formation is an early Cretaceous clasic unit. Double asterisk indicates zircon
SHRIMP U‐Pb unpublished age obtained in this work. (d) View to the structure along the Las Leñas valley (central
Principal Cordillera in profile P2, Figure 2b). (e) View to the southern slope of the Las Leñas valley immediately east of the
El Fierro thrust (eastern Principal Cordillera in profile P2, Figure 2b). The Río Damas Formation is a Kimmeridgian unit,
the Leñas‐Espinoza Formation is a Callovian volcanic‐marine unit, and the Termas del Flaco Formation is a Tithonian unit
(equivalent to the Lo Valdés Formation).

7 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

structures would be rooted in a detachment fault located at


midcrustal levels.

4. Seismology
4.1. Data Acquisition
[25] We used the seismologic data recorded by the per-
manent network of the Seismologic Survey at the Uni-
versidad de Chile between 1980 and 2004 complemented
with a temporary network deployed from January to April
2004 (Figures 1 and 2). The permanent network has
24 seismologic stations in the study region and the temporary
network consisted in seven short‐period three‐component
stations. The final database includes 23,449 events, from
which 217 earthquakes were detected by the temporary
network.

4.2. Tomographic Inversion and Hypocentral


Relocalization Procedure
[26] Once arrival times were picked, hypocenters were
Figure 4. Initial 1‐D P wave velocity models. first determined using the HYPOINVERSE program [Klein,
1978] with a 1‐D P wave velocity model based on that of
Thierer et al. [2005] (Figure 4). One‐dimensional S wave
velocity model was determined using 1.75 as Vp/Vs. Each
high‐angle west verging Chacayes‐Yesillo fault (see profile event was located with different trial depths in order to
P1 in Figures 2b and 3c) stands out because of its >2 km of minimize the effect of the initial conditions on the final
vertical throw [Charrier et al., 2005] cutting across the hypocentral determination. Trial depths were varied between
previously deformed series. 0 and 250 km with an increment of 5 km. The location with
the lowest root mean square misfit and with the maximum
number of body waves first arrivals was selected for each
3.3. Eastern Principal Cordillera and Frontal
event. This procedure ended with ∼140,000 and ∼95,000 P
Cordillera
and S arrival times, respectively. Because this procedure
[23] The eastern Principal Cordillera extends east of the El determines preliminary hypocentral locations, these results
Fierro fault system and consists of Mesozoic marine and are considered preliminary travel times.
continental sedimentary deposits including an Oxfordian [27] In order to obtain better hypocentral locations, we
gypsum level (Figures 2b, 3c, and 3d) and some volcanic constructed a simple 3‐D velocity structure that was used to
layers. Neogene syntectonic foreland basin deposits (Altos relocate the earthquakes. This procedure consists in using
de Tunuyán Conglomerates) overlying the Mesozoic the preliminary hypocenters and seismic wave travel times
sequence record the beginning of shortening and the east- to construct a 3‐D velocity structure using the SPHYPIT90/
ward migration of deformation to the eastern Principal SPHREL3D90 program (see details in the work by Roecker et
Cordillera shortly after 16 Ma [Giambiagi et al., 2003a]. al. [1993] and at http://gretchen.geo.rpi.edu/roecker/manuals/
Shortening in this region (∼47 km) has been mainly sphypit90/Sphypit90.html). Inversion was performed on a
accommodated by east verging thrusts and related folds region divided into 6 × 7 blocks with a grid spacing of 30 ×
[Giambiagi and Ramos, 2002]. Structural restoration based 30 km2 (see Figures 1 and 6) and 12 layers of 10 km thick,
on balanced cross sections [Giambiagi and Ramos, 2002] except the shallowest one, which is 13 km thick. P wave and
suggests that these thrusts would be rooted in a ∼10 km S wave velocities were independently inverted.
depth east verging detachment fault along the latitude of the [28] The resulting Vp and Vs models were used to relocate
profile P1 (Figure 2b). the hypocenters and adjust travel times. Hypocenters were
[24] North of 34°15′S, the crystalline basement of the classified and filtered according to the following criteria
Frontal Cordillera crops out east of the Principal Cordillera. (used by Abers and Roecker [1991]): (1) standard deviation
There, several basement blocks were uplifted by high‐angle of the residual travel times <0.7 s, (2), maximum spatial
east verging reverse faults accommodating at least 15 km of standard error < 0.2 s, and (3) maximum spatial correction
shortening between 9 and 6 Ma [Giambiagi and Ramos, the hypocenter <0.2 s, and total change in location from start
2002]. Giambiagi et al. [2003a] and Ramos et al. [2004] to finish < 40 km. These filtered events were used to gen-
interpreted this thick‐skinned contractive style as a result erate a new inversion. Iterations continued until resulting
of the inversion of faults related to the Mesozoic extensional models converged to a solution: three iterations were needed
basins. According to these authors, the reactivation of these to achieve convergence. Inversion resulted in a final model
consisting in 1008 final blocks with 877 blocks considered

8 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 5. Standard deviation calculated for each event using the nine velocity models. Data filtering
was made considering the most reliable earthquakes.

as reliable (those having >20 rays hits, however most of [32] In order to use the most reliable hypocenters, we only
blocks are hit by >1000 rays). The resulting velocity model considered those having standard deviations smaller than
was used to relocate the hypocenters, which were classified ±1 km in latitude and longitude and depth deviations minor
and filtered again. than ±2.5 km. In addition, we used those events having
procedural RMS < 0.3 s. This course of action resulted in
4.3. Model Validation and Filtering 7077 events that will be used in the rest of this work. The
[29] It is necessary to point out that the tomography was filtering treatment used in the earthquakes detected by the
performed with the aim of determining better hypocentral temporary network (217 events) was different: we consider
locations rather than obtaining an accurate velocity field that they are a priori better constrained because of the
because of its rough resolution (30 × 30 × 10 km each cell). proximity of stations to the quakes (less than 30 km); this
However, the following tests show that this final model has filtering consisted in selecting those event having RMS <
a relevant regional validity that can give insight about the 0.35 s, resulting in 149 events.
real velocity field in the study region.
[30] Tomography testing consisted in the development of 4.4. Results and Analysis of Seismologic Data
eight additional models using 1‐D velocity structures ran- 4.4.1. General Results
domly perturbed from the initial model of Thierer et al. [33] The final distribution of hypocenters shows that most
[2005] (Figure 4). Comparison among the nine resulting of the crustal seismicity is located beneath the Principal
velocity models shows a minimal deviation at each cell Cordillera and eastern Central Depression at depths shal-
(<3%). Because of this reason, we took the mean velocity of lower than 20 km (Figure 6). Superficial seismicity in the
the nine models as the final Vp and Vs models that will be offshore and coastal fore arc mainly corresponds to events
used in the rest of this work (considering the deviation as the occurring in the interplate contact, except some earthquakes
involved error at each cell). located in the overriding plate. As previously suggested by
[31] In order to test the sensibility on hypocenter reloca- Barrientos et al. [2004] and Charrier et al. [2005], the most
lization, we compare the hypocenters determined by the relevant cluster of shallow crustal seismicity is located close
nine models (Figure 5). This comparison consisted in the to the Chile‐Argentina boundary aligned with the El Fierro
determination of the standard deviation in latitude, longitude fault system (Figure 6).
and depth for each event. Figure 5 illustrates that more than 4.4.2. Ramp Flat Seismologic Structure
the half of the events has minimal standard deviation, [34] In order to analyze the relationship between seis-
showing that the different models tend to a similar hypo- micity, seismic velocity fields and lithospheric structure,
central relocalization neglecting the deviation of the initial Figure 7 shows three profiles crossing perpendicularly the
velocity models. Actually, the events that exhibited minimal orogen strike and covering most of the study region (note
standard deviation coincide with those events having mini- that these profiles have different azimuths because the
mal RMS derived from the procedure (<0.3 s). orogen strike changes in the Maipo orocline, see Figure 2).

9 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 6. Regional distribution of final hypocenters. Solid lines correspond to the locations of cross
sections in Figure 7. Grid corresponds to that used as the inversion region (each cell has a 30 × 30 km
planar size).

At first sight, the geometry displayed by seismicity can be [38] In spite of the former features related to the ramp or
interpreted as a ramp‐flat crustal‐scale structure. its inferred southward prolongation, it is also likely that
[35] Seismicity associated with the flat geometry is seismicity defining the ramp in the mantle could be related
located beneath the Chilean Principal Cordillera. This to dehydrating associated with fluid releasing from the slab.
structure dips ∼10°W in the western Principal Cordillera, In this case, faulting into the lithosphere mantle associated
where it is located at 15–10 km depth. In the central eastern with the ramp structure would not occur, and thus this zone
Principal Cordillera, the structure is located at 10–5 km of the fore arc would only transmit the stress that is released
depth and dips ∼5°W. In this sector, the flat structure in deformation along and above the crustal ramp‐flat
coincides fairly well with the geometry and depth proposed structure that is very well observed seismically in the study
by Giambiagi et al. [2003a] for the detachment that has region (see section 8.2).
controlled shortening in the eastern Principal Cordillera.
[36] The ramp segment dips ∼40°W and extends down-
ward from the western edge of the Principal Cordillera to the 5. Rheological Analysis
Moho below the Central Depression. Although seismicity in [39] The strength of the continental lithosphere is con-
the ramp segment is not present in the lithosphere mantle in trolled by its depth‐dependent rheological structure. This is
the southern sections (possibly reflecting the aseismic mainly depending on the thickness and composition of
behavior of this part of the lithosphere), it is well detected crustal layers, the thickness of the lithosphere mantle, the
in the section across the Central Depression at 33.2°S temperature structure, the strain rate, and the presence or
(Figure 7a). In this profile, the seismic ramp intersects the absence of fluids [e.g., Carter and Tsenn, 1987; Kirby and
Wadatti‐Benioff zone at ∼60 km depth. Kronenberg, 1987; Burov and Diament, 1995, 1996;
[37] The seismic ramp, and its likely location in zones in Cloetingh et al., 2005].
which it is not seismically present, can be correlated with [40] In order to analyze the probable rheological control on
slight discontinuities on Vp and Vp/Vs within the lithosphere the ramp‐flat structure, we constructed four one‐dimensional
mantle wedge (Figure 7). Discontinuity on the P wavefield columns of compressive yield strength envelopes (Figure 8)
consists in an eastward velocity increase from 7.3–7.7 to across the profile BB′ (Figure 7b). They are based on the
7.9–8.2 km s−1. Discontinuities on Vp/Vs are not much 3‐D lithospheric compositional and geometrical model of
obvious; however, the ramp (or what could be its southward Tassara et al. [2006] (which consists in two crustal layers
prolongation in Figures 7b and 7c) is surrounded by high and a lithosphere mantle; Figure 8a), a 2‐D geothermal
Vp/Vs regions (>1.80) (Figure 7). gradient (based on the work by Oleskevich et al. [1999] and

10 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 7

11 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Yáñez and Cembrano [2004]) (Figure 8b), and experimental correlates well with the western edge of this low‐viscosity
rheological parameters. ductile zone (columns II and III). The ramp segment extends
[41] The 2‐D geothermal gradient used in this analysis from there downward to the top of the lower crustal low‐
was mainly constructed using a heat diffusion model in viscosity ductile zone near the Moho (column II). This
which the downgoing slab produces the downward isotherm “shift” from an upper to a lower detachment level is usual in
advection [Oleskevich et al., 1999; Yáñez and Cembrano, ramp structures, which generally connect weak zones across
2004]. The effects of the volcanic arc on the thermal gra- a rigid layer [see Cook and Varsek, 1994, and references
dient were minimized: we consider that the long‐term therein]. In spite of the lack of direct evidence for the lower
effects of the slab are more relevant than the continuously ductile zone, the activity of its top as a deep east verging
migrating magmatic/volcanic arc. detachment could explain the reactivation of the deeply
[42] The rheological parameters used in this work are rooted basement structures that controlled the uplift of the
exactly those used by Tassara et al. [2006]. They assumed a Frontal Cordillera in the late Miocene [Giambiagi and
quartzite rheology for a granitic upper crust, a dry diabase Ramos, 2002; Giambiagi et al., 2003a; Ramos et al., 2004],
rheology for a quartz‐diorite lower crust, and a wet dunite as well as those that are now controlling the uplift of the San
rheology for a continental harzburgite lithosphere mantle, Rafael block farther east [Ramos et al., 2004].
whose values are those determined by Carter and Tsenn [46] Yield strength envelopes predict that the rocks
[1987] and Burov and Diament [1995] (see Figure 8 caption located above the detachment beneath the Principal Cordillera
for details). should prevail very rigid. However, this zone concentrates
[43] Resulting yield strength envelopes (Figure 8c) illus- abundant seismicity normally aligned with some structures
trate that in the western flank of the eastern Principal Cor- observed at the surface. Because this zone has been widely
dillera (column IV), low‐viscosity ductile rocks should deformed during the Neogene (Figure 2b), it is likely that
prevail between 8 and 20 km depth (upper crust) and seismic activity in this region is related to the reactivation of
between 35 and 60 km depth (lower crust). These low‐ ancient discontinuities and fractures. Indeed, Charrier et al.
viscosity ductile zones are confined by three “rigid” (high‐ [2002, 2005] proposed that the deformation within the
strength brittle and/or high‐viscosity ductile) layers located Abanico Formation is mostly related to the inversion of the
between 12 and 3 km depth, between 20 and 35 km depth, normal faults that controlled the development of the Eocene‐
and immediately below the Moho. Oligocene extensional Abanico basin. Likewise, Neogene
[44] The geothermal gradient diminishes to the west. Its deformation in the eastern Principal Cordillera and Frontal
effect is clearly observed in column III (western edge of the Cordillera is related to the reactivation of older normal faults
central Principal Cordillera), showing that the upper low‐ as well as fracturing along less competent levels such as
viscosity ductile zone wedges out toward the west and dis- gypsum layers, old Paleozoic sutures and basement fabrics
appears beneath the western Principal Cordillera (column II). [Ramos et al., 1996; Giambiagi and Ramos, 2002; Giambiagi
This analysis predicts mechanic coupling between the upper et al., 2003a]. Therefore, seismicity above the detachment
and lower crusts beneath the westernmost Principal Cor- would be related to the reactivation of older structures and
dillera and Central Depression (column I). Yield strength discontinuities rather than new faults.
envelopes also suggest that the viscosity in the base of the [47] Yield strength envelopes analysis shows that the fore
crust would remain low beneath the western edge of the arc beneath the Coastal Cordillera should be very rigid, even
Principal Cordillera. However, farther west, strong coupling though some earthquakes related to the ramp segment are
between the crust and the upper mantle is observed (column I). located in this zone (Figure 7a). This fact suggests that the
This is produced by the cooling exerted by the subduction of rheology of this zone may differ from what is expected,
a cold oceanic lithosphere that deflects downward the iso- which could be explained by the presence of fluid released
therms by advection. Such a kind of strong and cold con- from the subducting slab or lithosphere discontinuities
tinental lithosphere has been proposed for the northern Chile derived from the fore‐arc evolution. We will discuss this
fore arc, which would furthermore behave as a strong subject in section 8.
eastward intender [Tassara, 2005]. In central Chile, this
characteristic also appears in this analyze; that is, the lith-
osphere fore arc west of the Central Depression is expected 6. Integrating Seismologic Data and Surface
to be highly rigid and cold. Geology
[45] Beneath the Principal Cordillera (columns III and IV),
the top of the upper crustal low‐viscosity ductile zone is 6.1. Construction of Balanced Cross Sections
fairly well correlated with the flat segment of the seismic [48] In spite of the good correlation existing between yield
structure (Figure 8c). The western edge of the flat segment strength envelopes and the locus of seismicity, both are

Figure 7. Crustal‐scale cross sections perpendicular to the orogen strike showing velocity structures and relocated hypo-
centers. Location and orientation of sections are indicated in Figure 6. Moho depth is after Tassara et al. [2006]. In all
sections, earthquakes were projected using a 40 km width box (+20/−20 km from the profile). White circles are the events
recorded by the permanent network, and red circles are those obtained from the temporary network. The ramp‐flat structure
has been drawn in Vp/Vs sections. All sections are at the same scale, using the same color palette for velocities, and there is
no vertical exaggeration.

12 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 8. Compressional yield strength envelopes analysis. (a) Figure 6b used as reference to indicate
the location of the columns where yield strength envelopes were calculated. Moho and intracrustal density
discontinuity (ICD) depth are after Tassara et al. [2006]. The 400°C isotherm is reported for reference.
(b) Geothermal gradient for the four columns used for yield strength envelop calculation. (c) Resulting
yield strength envelops. Geothermal gradient is approximated from Oleskevich et al. [1999] and Yáñez
and Cembrano [2004]. The ICD delimitates an upper crust with quartzite composition (H = 1.9 × 105
[J mol−1], A = 5 × 10−12 [N−3 m6 s−1] [Burov and Diament, 1995]) from a lower crust with quartz‐diorite
composition (H = 2.12 × 105 [J mol−1], A = 5.1 × 10−15[N−2.4 m5.76 s−1] [Burov and Diament, 1995]).
Mantle has been considered with a wet dunite composition (H = 4.44 × 105 [J mol−1], A = 7.94 × 10−17
[N−3.35 m11.22 s−1] [Carter and Tsenn, 1987]). Maximum deviatoric compressive stress in the continental
lithosphere (∼100 MPa) is according to England and Molnar [1991]. Locations of columns are as follows:
column I is below the Central Depression, column II is below the western edge of the Principal Cordillera,
column III is below the central Principal Cordillera, and column IV is beneath the Chilean side of the
eastern Principal Cordillera.

13 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Figure 9. Structural cross sections and shallow seismicity. (a) Maipo profile (P1 in Figure 2b)
(b) Cachapoal profile (P2 in Figure 2b). Structural restoration shows that less than 20 km (∼16 km of
shortening minimum) has been accommodated within the Cenozoic sequences (Abanico and Farellones
formations). This shortening is distributed almost equitably on both flanks of the former extensional
basin. Note the features of the bordering fault systems, which present evidence for tectonic inversion.

present‐day features, and hence, they cannot be directly constructed two upper crustal cross sections integrating
associated with a long‐term situation. However, as we will surface geology and seismicity (Figure 9). These sections
expose in the following paragraphs, they are also well cor- display the structure only in the Chilean side of the cordil-
related with surface geology and structural reconstruction lera: the Argentinean side of the belt has been already
made on the eastern flank of the Andes [Giambiagi and studied in detail by Giambiagi and Ramos [2002] and
Ramos, 2002; Giambiagi et al., 2003a]. Giambiagi et al. [2003a] and the resolution of recorded
[49] At depth, the seismic flat structure coincides fairly seismicity does not extent farther east to Argentina.
well with the detachment proposed by Giambiagi and [50] Cross sections are constrained by down‐plug pro-
Ramos [2002], Giambiagi et al. [2003a], and Ramos et al. jection of surface structure and its correlation with seis-
[2004] based on balanced cross sections. In order to ana- micity, geometrical constraints, and the age of deformation.
lyze the role of this structure on mountain building, we Despite the precise downward prolongation of some par-

14 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Table 1. Shortening Across the Andes at 33.8°Sa


First Stage Second Stage Third Stage Fourth Stage
22–15 Ma 15–8.5 Ma 8.5–4 Ma 4–0 Ma Total
b
Abanico basin 16 16
Eastern Principal Cordillera 6 24 17 47
Frontal Cordillera 15 15
Foreland 6 6
Total 22 24 32 6 84
a
Approximated shortening in km. The values are according to Giambiagi and Ramos [2002].
b
Estimated in this work.

ticular faults may be debated, these cross sections show the [53] After circa 16 Ma, shortening migrated from the
geometry of the major orogen‐scale structures. It is neces- Abanico basin to the east. Initially, shortening affected the
sary to point out that this restoration could underestimate the eastern Principal Cordillera, propagating to the Frontal
shortening because of minimization of fault slip when they Cordillera at circa 9 Ma, and returning into the eastern
are not related to folding or where the cut across the folds. It Principal Cordillera during an out‐of‐sequence event (8–4 Ma)
is likely that this minimization will not modify significantly in the study region. It is necessary to point out that timing
the actual shortening because erosion‐exhumation related to and shortening magnitude present wide variation along
vertical movements of main faults are not very relevant and strike: shortening has clear decreasing trending to the south
even fission track dating did not record exhumation in rocks and being slightly younger to the south [Ramos et al., 1996;
of the hanging walls [Farías et al., 2008a], at least in the Giambiagi et al., 2008, 2009]. Finally, shortening migrated
western Principal Cordillera. to the foreland at circa 4 Ma [Giambiagi et al., 2003a].
[51] We also expect that the use of seismicity to determine Deformation in the eastern side of the mountain range has
the depth of the detachment fault controlling the deforma- been predominantly accommodated by east verging thrusts
tion observed at the surface would better constrain the and some backthrusts. Shortening in the eastern Principal
structural construction. In fact, it is widely recognized that Cordillera, Frontal Cordillera, and foreland is at least ∼62 km
the depth of the detachment respect to the surface structures at 33.8°S [Giambiagi and Ramos, 2002], which represents
does determine the amount of shortening and that assuming about 80% of the total shortening across the chain at this
a certain depth without more constraints than those observed latitude (see Table 1).
at the surface may lead to underestimating or overestimating
the total deformation along a cross section. Moreover,
6.3. Synthesis of Deformation Events
performing of balanced cross sections usually consider that
thickness of sequences involved in deformation is constant. [54] Considering the timing of deformation, most of the
However, the Abanico Formation reports important changes shortening occurred during three major events at 33.8°S
in its thickness [e.g., Charrier et al., 2002]. Moreover, the (Table 1): (1) Abanico basin inversion (∼16 km of short-
Mesozoic sequences also present great difference in thick- ening between 22 and 16 Ma, and 6 km of shortening in the
ness, from about 30 km between the Coastal Cordillera and eastern Principal Cordillera before 15 Ma), (2) thin‐skinned
Central Depression (intra‐arc domain) to about 10 km in the fold‐and‐thrust belt development in the eastern Principal
eastern Principal Cordillera (Neuquén basin). Cordillera (24 km of shortening between 16 and 8.5 Ma
[Giambiagi and Ramos, 2002]), and (3) uplift of the Frontal
Cordillera (15 km of shortening between 8.5 and 6 Ma) and
6.2. Shortening Magnitude and Timing out‐of‐sequence thrusting in the central eastern Principal
[52] In the zone where the Abanico basin developed Cordillera (∼17 km of shortening between 8.5 and 4 Ma
(eastern Central Depression, western and central Principal [Giambiagi and Ramos, 2002]).
Cordillera), the orogen approximately displays a symmetric [55] The out‐of‐sequence thrusting event in the eastern
double‐vergency system of faults, preserving a central Principal Cordillera represents a disruption of the eastward
portion that remains scarcely deformed (Figure 9). A mini- migration of shortening. Shortening migrated eastward
mum estimation of shortening in this part of the chain is along the detachment until the high‐angle basement faults
16 km (∼20% of the total shortening across the mountain rooted in the Frontal Cordillera were reactivated. This
belt, see Table 1). At this place, most of the contractive reactivation caused the return of the deformation to the axis of
deformation occurred in the lower Miocene, resulting in the the mountain belt as out‐of‐sequence thrusting [Cristallini
partial inversion of the Abanico basin. Deformation is dis- and Ramos, 2000; Giambiagi et al., 2003a; Ramos et al.,
tributed almost equitably on both flanks. Thus, the resulting 2004]. Considering that this event was coeval with the
geometry is consistent with the inversion of an extensional migration of the magmatic activity to west [e.g., Kay et al.,
ramp‐flat listric fault system [McClay, 1995] (Figure 9), as 2005] and that in this new arc developed the porphyry
well as local structural geometry evidences basin inversion copper deposits, it is very likely a relationship. Preliminar-
(see also Figure 3). During the basin inversion, Mesozoic ily, we think that the westward migration of shortening and
sequence in the eastern Principal Cordillera accommodated the high crustal thickening reached at this time would alter
∼6 km of shortening [Giambiagi and Ramos, 2002]. the magmatic ascent and differentiation. After 4 Ma, the

15 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

magmatic activity ended in the porphyry copper deposits thickening should have occurred in order to compensate the
[Maksaev et al., 2004; Deckart et al., 2005], coeval with the elevation of the westernmost Andes. This implies that the
migration of shortening to the foreland. During this new amount of crustal shortening inferred from geophysical data
stage, about 6 km of shortening has been accommodated (supported by determining of crustal thickness and its his-
[Giambiagi and Ramos, 2002]. At this time, the high cor- tory of thickening [e.g., Introcaso et al., 1992; Pose et al.,
dillera did not accommodate significant shortening in the 2005]) matches the amount of shortening derived from
study region. In turn, strike‐slip deformation is presently structural cross sections, as well as the uplift timing and
reported by seismic activity in earthquakes M > 5 [Farías et magnitude [Farías et al., 2008a].
al., 2006; Farías, 2007]. According to Farías et al. [2008a], [59] The advance to west of the crust beneath the detach-
the migration of contractive tectonics to the foreland and the ment would be opposed by the westward increasing of
beginning of a strike‐slip regime in the mountain belt would the lithosphere rigidity in the fore arc (Figures 8 and 10).
be related to the opposition of shortening exerted by the This opposition seems to be evidenced by the seismic cluster
high elevation reached by the mountain belt in the Pliocene. located in the ramp immediately above the Moho (Figure 9a),
which presents compressive focal mechanisms [Pardo et al.,
2008]. Decreasing westward movement of the crust beneath
7. Discussion: Implications for Mountain the ramp‐flat structure means that thickening occurred in a
Building in Subduction Zones mixed pure and simple shear mode. Thus, east vergent
crustal wedging would have controlled the accommodation
7.1. Simple Shear and Pure Shear Mountain Building of the thick mountain root beneath the High Andes. This
[56] Before 22 Ma, crustal thickness would have been situation can finally explain the decreasing crustal thickness
moderate (<35 km thick) and approximately uniform across and mountain elevation to the west (Figure 10), as well as
the Abanico basin according to studies on the geochemistry the minimal subsidence occurring in the Central Depression.
of the basin‐related volcanic deposits [e.g., Fuentes, 2004; Moreover, decreasing westward advance of the crust below
Kay et al., 2005; Muñoz et al., 2006; Montecinos et al., the detachment could explain the fact that seismicity related
2008]. In fact, there is no structural evidence of relevant to the ramp contacting the slab or occurring in the upper
shortening after the Permian and before the late Oligocene mantle is not be observed everywhere. If the Moho has been
in this Andean region [Charrier et al., 2002; Giambiagi et al., displaced along the ramp fault will remain unsolved until
2003a; Ramos et al., 2004; Fock et al., 2006; Giambiagi et al., better accurate geophysical images are obtained, even
2009]. though receiver function derived images along transects at
[57] Although in both flanks of the Abanico basin crustal ∼30°S and ∼36°S do not show this situation [Gilbert et al.,
thickness was approximately similar before inversion and 2006]. It is likely that if displacements occur in the Moho,
surface shortening accommodated in both sides has been ductile behavior in the base of the crust would allow the
approximately the same amount, the present‐day crust is redistribution of the displaced material.
∼20 km thicker below the eastern boundary of the former [60] A major east verging ramp‐flat structure with a
basin (central Principal Cordillera) than beneath the western geometry also controlled by the rheology of the continental
Principal Cordillera (Figure 10). This suggests that in situ plate has been already proposed for the northern Chile
surface deformation cannot explain the thickening of the margin (at 19°30′S) by Farías et al. [2005] and Tassara
crust in a “pure shear mode” (in the sense of Allmendinger [2005] based on the works of Isacks [1988] and Lamb et
and Gubbels [1996]). Therefore, other thickening mechan- al. [1997], among others. These authors proposed that this
isms would have driven the construction of the Chilean structure would be connected to the detachment fault that
flank of the belt. extends through the Altiplano to the Eastern Cordillera and
[58] We propose that some of the ∼70 km of surface Subandean zone where most of the shortening has been
shortening accommodated east of the Abanico basin has accommodated [e.g., McQuarrie, 2002; see also Ramos et
been transferred to the west beneath the detachment in a al., 2004]. According to this model, the ramp encloses
“simple shear mode” rather than in a “pure shear mode” west and upward a rigid fore arc acting as a pseudoindenter
(both in the sense Allmendinger and Gubbels [1996] and see that resists the westward advance of the crustal mass located
Figure 10), similarly to the proposition of Ramos et al. beneath the detachment (see Figure 10). It results in crustal
[2004]. It is necessary to point out that this “simple shear thickening in the Precordillera and Western Cordillera, in a
mode” of deformation does not mean that all shortening is zone where the surface shortening has been moderate (less
transferred to the west, but only partially. The importance of than 20 km) and essentially older than uplift [e.g., García,
deep shortening in the western part of the chain is also 2002; Victor et al., 2004; Farías et al., 2005; Hoke et al.,
supported by the fact that most of the surface and rock uplift 2007; Riquelme et al., 2007], and where subsidence in the
of the western and central Principal Cordillera occurred Central Depression can be also considered negligible
between 10 and 4 Ma, even though most of the surface [Farías et al., 2005, 2008b].
shortening was accommodated before 16 Ma [Farías et al., [61] Our model for the structural architecture in which the
2008a]. Furthermore, the fact that subsidence is almost Andean Cordillera developed is consistent not only with
negligible in the peripheral Central Depression [Farías et evidence from the central Chile–central Argentina Andes,
al., 2008a, 2008b], despite the lack of enough shortening but also from the Altiplano‐Puna central Andes. This model
in the westernmost Principal Cordillera supporting this part is based on the fact that reported shortening along this
of the chain [Farías et al., 2008a], implies that deep Andean region has been mostly accommodated by east

16 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

vergent thrusts, in which west vergent thrusts correspond to 2006; Rauld et al., 2006; Armijo et al., 2010]). Neverthe-
back thrust. In spite of that, relevant west vergent thrusts less, these structures are not much relevant in the context of
occur along the western cordillera front (west vergent thrust the Andean shortening. In fact, west vergent structures do
system in northernmost Chile [e.g., Muñoz and Charrier, not exhibit more than 2–3 km of vertical throw along the
1996; García, 2002; Victor et al., 2004; Farías et al., 2005; Chilean flank of the belt (which is evidenced by thermo-
García and Hérail, 2005] and the San Ramón‐Pocuro fault chronological data in central Chile [Farías et al., 2008a] and
in central Chile [Charrier et al., 2002, 2005; Fock et al., by direct structural observations in northern Chile [Victor et

Figure 10

17 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

al., 2004; Farías et al., 2005]) and its accommodated coincides with the deepest limit of the seismogenic contact
shortening has been only a minimal fraction of the total along the Chilean subduction zone (i.e., along the place in
shortening across the belt [cf. García, 2002; Victor et al., which thrust interplate earthquakes propagate [Suarez and
2004; Farías et al., 2005; this work]. Furthermore, consid- Comte [1993]) (Figure 10). In addition, the continental
ering that subsidence in the peripheral Central Depression mantle in this zone presents relevant variations on Vp and
has been almost negligible, the relevance of west vergent Vp/Vs respect to the initial one‐dimensional model (Figure 7).
thrusts along the western cordillera flank on crustal thicken- In spite of the uncertainties related to the velocity fields,
ing (and therefore on Andean building) is therefore minimal. these variations could be related to mantle serpentinization
Consequently, evidence supports our proposition of crustal because the referred intersection would also coincide with
deep thickening related to the east vergent ramp‐detachment the 400°–500°C isotherm (Figures 7 and 9), which is the
structure; on the contrary, evidence does not support west upper limit of serpentinite stability [Carlson and Miller,
vergent model for the Andean development proposed by 2003]. Empirical relationships between the degree of upper
Armijo et al. [2010]. continental mantle serpentinization and Vp [Carlson and
Miller, 2003] predict that the observed discontinuity would
7.2. Stress‐Strain Transfer From the Interplate correspond to a change from ∼0% to 20% of serpentinization
Contact at the conditions of pressure‐temperature expected for this
[62] The east vergent ramp emerging from the interplate zone. In addition, high Vp/Vs ratios at depths minor than
contact area or the downward prolongation of its trace from 60 km in the mantle wedge (Figure 7) are also consistent with
the crust to the mantle has been visualized at different lati- serpentinization [e.g., Kamiya and Kobayashi, 2000].
tudes along the margin, intersecting the slab approximately [64] Nonetheless, the fact that in northern and central
at the same depth (∼60 km). Hypocentral location of Chile the depth at which the ramp would intersect the slab is
earthquakes within the overriding plate show that such a approximately the same despite the different thermal gradient
structure is active at 19°S [Comte et al., 1999; David et al., expected in both regions suggests that not only the mainly
2002] and 27°S [Comte et al., 2002; Pardo et al., 2002], and temperature‐controlled serpentinization would determine
seismic images visualize a strong west dipping reflector this feature. Therefore, it is likely that pressure also exerts a
immediately above the slab at ∼38°S [Gross et al., 2007]. very significant control. In fact, approximately at this depth a
Furthermore, comparable features at a similar depth have strong slab dewatering occurs due to the metamorphic
been reported in other subduction zones (Central America reactions related to the blueschist‐eclogite transition. In order
[e.g., Dinc‐Akdogan et al., 2007; Syracuse et al., 2008], to test whether metamorphic reactions, serpentinization, or
northwestern Pacific [e.g., Zhao et al., 2000; Mishra et al., the interplay between both control this feature, thermo-
2003], and Cascadia [e.g., Zhao et al., 2001]). Furthermore, mechanical modeling is required.
in northern Chile, David et al. [2002] determined that [65] It is clear that an important shift on the mechanic
earthquakes along the ramp immediately above the slab behavior along the interplate contact occurs in the inter-
have compressive mechanisms. Nevertheless, we agree that section of the ramp with the slab at nearly 60 km depth.
there is no enough evidence supporting the real activity of Following Lamb and Davis [2003] and Lamb [2006], the
the ramp in the lithosphere mantle because of the absence of stress produced by the plate convergence is mostly trans-
seismicity in all the sections. Therefore, it is likely that ferred to the overriding plate along the seismogenic inter-
faulting extents into this deep region only in some places, plate contact, which also occurs at this depth (Figure 10).
probably in relationship with ancient discontinuities and/or Moreover, Tassara [2005] proposed that the rigid behavior
with mineral reactions associated with slab dehydration and of the fore arc (which is expected to be strongly coupled
uppermost continental mantle hydration (serpentinization), with the slab) would promote more effectively the stress
which downdip limit should be related to the transformation transfer to the continent. In this context, the ramp structure
from blueschist to eclogite facies (see below). in the lithosphere mantle or its geometrical prolongation into
[63] If the intersection of the ramp or its deep prolongation this zone control the strain transfer and delimitate upward
into the mantle with the slab is located everywhere at similar the rocks that transmit part of the plate convergence forces
depths in the Chilean fore arc, it is likely that subduction toward the continental crust. In this way, this zone is crucial
processes control this feature. In fact, this intersection in controlling mountain building.

Figure 10. Model for crustal growth and the relevance of the ramp‐detachment structure at approximately 33°50′S.
(a) Two different models of mountain building (models modified after Allmendinger and Gubbels [1996]). In “pure
shear mode” of deformation, the crust thickens in the same place where surface shortening occurs. In turn, in “simple shear
mode,” thickening occurs far from the place where surface shortening occurs due to the presence of a detachment that
transports the deep crust to the left. (b) Initial setting before shortening at 22 Ma. (c) Present‐day (shortened) crustal
configuration of the continental plate. Shortening in the eastern Principal Cordillera, Frontal Cordillera, and foreland
according to Giambiagi and Ramos [2002]. Moho depth is after Tassara et al. [2006]. The 400°C isotherm is based on the
work by Oleskevich et al. [1999] and Yáñez and Cembrano [2004]. Note that the deepest limit of the seismogenic contact
also coincides with the blueschist‐eclogite transition. It can be observed that despite the differences of shortening at the
surface, crustal thickening has been distributed to the west in a partial simple shear mode of deformation. LVZ is a low‐
velocity zone expected in this area because of magmatic plume. Cuyania Terrane is according to Ramos et al. [2004].

18 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

[66] On the basis of the evidence presented in this work, area at ∼60 km depth and controls the structuring of the
the ramp‐detachment structure seems to be the first‐order Andes has been proposed for northern Chile. Inferences
feature controlling the transference of strain and stress from made on the south central Chile region also suggest such a
the subduction zone to the mountain belt not only in the structure there, as well as observations made in different
Andes of central Chile, but also probably along the entire subduction zones around the Pacific basin. The intersection
Andean margin. Because this structure should be strongly of the ramp with the slab coincides with the deepest limit of
controlled by subduction processes and similar features are the seismogenic interplate contact. It is also marked by sharp
observed in other subduction regions, it is likely that the variations in the overriding mantle seismic velocities that
model presented in this work holds for other mountain belts can be interpreted as a result of serpentinization of the
formed in a subduction regime. lithospheric mantle wedge probably in relationship with
blueschist‐eclogite reactions in the slab.
8. Conclusion [70] Therefore, we propose that subduction controls the
tectonic behavior of the fore arc, weakening the zones by
[67] Using the seismicity recorded in central Chile by which strain stress is transferred to the mountain range from
permanent and temporary networks, we performed a 3‐D the plate interface. Likewise, the major east verging ramp
tomography inversion that led to the relocalization of the delimitates upward rigid rocks that transmit part of the plate
most reliable hypocenters. We showed the presence of a convergence forces toward the continental lithosphere.
crustal‐scale ramp‐flat structure that connects the subduc- [71] This model suggests that the east verging ramp‐flat
tion zone at ∼60 km depth with the mountain belt at ∼10 km. structure is the first‐order structure in the Andean mountain
The flat segment crosses the entire mountain belt and cor- belt orogeny. Because this structural architecture is strongly
relates with the east verging detachment that accommodated controlled by the subduction factory and similar features can
most of the upper crust shortening during the Neogene be observed in other regions around the Pacific basin, this
mountain building. model could be potentially applicable to other subduction
[68] Geological cross sections show that in the Chilean margins.
side of the belt, upper crustal shortening was much smaller
than in the Argentinean fold‐and‐thrust belts (1/5 versus 4/5
of the total shortening). In the western part of the central
[72] Acknowledgments. This work was supported by FONDECYT
Chile Andes, neither the present‐day crustal thickness nor grants 1030965 and 11085022, PBCT ANILLO ACT 18 and PDA‐07
the uplift of this side of the mountain belt can be explained Project, INSU grant “Relief de la Terre. Impact du climat sur la dynamique
by surface shortening. In fact, most of the Neogene crustal du relief des Andes: quantification et modélisation,” and an IRD Phd schol-
thickening and uplift of the western part of the central Chile arship to M. Farías. The authors particularly recognize the labor made by
the Seismologic Survey at the University of Chile. We acknowledge Steven
Andes would result from the shortening accommodated Roecker for providing the SPHREL90/SPHYPIT programs. Useful discus-
beneath the detachment. sion with C. Arriagada, M. Pardo, G. Hérail, A. Reynaldo, and G. Yáñez
[69] Despite huge latitudinal contrasts in the morpholog- helped develop and clarify our ideas. Some figures were made using
ical and tectonic evolution of the Chilean Andes, a similar GMT 4 [Wessel and Smith, 1998] and GRASS 6.3 programs; GTOPO30,
general lithosphere structural scheme in which a major east SRTM90 topographic data, and the 2 min gridded ocean bathymetry of
Smith and Sandwell [1997] were used in some figures.
verging fault system emerges from the interplate contact

References
Abers, G. A., and S. Roecker (1991), Deep structure of Cretaceous, Tectonics, 25, TC1008, doi:10.1029/ und ihre tektonische, magmatische und paläogeo-
an arc‐continent collision: Earthquake relocation 2004TC001770. graphische Entwicklung, Ph.D. thesis, 270 pp.,
and inversion for upper mantle P and S wave veloc- Barrientos, S., E. Vera, P. Alvarado, and T. Monfret Freie Univ. Berlin, Berlin, Germany.
ities beneath Papua New Guinea, J. Geophys. Res., (2004), Crustal seismicity in central Chile, J. South Charrier, R., A. R. Wyss, J. J. Flynn, C. C. Swisher,
96, 6379–6401, doi:10.1029/91JB00145. Am. Earth Sci., 16, 759–768, doi:10.1016/j. M. A. Norell, F. Zapatta, C. McKenna, and M. J.
Allmendinger, R. W., and T. Gubbels (1996), Pure jsames.2003.12.001. Novacek (1996), New evidence for late Mesozoic–
and simple shear plateau uplift, Altiplano‐Puna, Burov, E. B., and M. Diament (1995), The effective early Cenozoic evolution of the Chilean Andes in
Argentina and Bolivia, Tectonophysics, 259, 1–14, elastic thickness (Te) of continental lithosphere: the upper Tinguiririca valley (35°S), central Chile,
doi:10.1016/0040-1951(96)00024-8. What does it really mean?, J. Geophys. Res., 100, J. South Am. Earth Sci., 9, 393–422, doi:10.1016/
Allmendinger, R. W., and T. R. Zapata (2000), The 3905–3927, doi:10.1029/94JB02770. S0895-9811(96)00035-1.
footwall ramp of the Subandean decollement, north- Burov, E., and M. Diament (1996), Isostasy, equivalent Charrier, R., O. Baeza, S. Elgueta, J. J. Flynn, P. Gans,
ernmost Argentina, from extended correlation of elastic thickness, and inelastic rheology of continents S. M. Kay, N. Muñoz, A. R. Wyss, and E. Zurita
seismic reflection data, Tectonophysics, 321, 37– and oceans, Geology, 24, 419–422, doi:10.1130/ (2002), Evidence for Cenozoic extensional basin
55, doi:10.1016/S0040-1951(00)00077-9. 0091-7613(1996)024<0419:IEETAI>2.3.CO;2. development and tectonic inversion south of the
Allmendinger, R. W., D. Figueroa, D. Snyder, J. Beer, Carlson, R. L., and D. J. Miller (2003), Mantle wedge flat‐slab segment, southern central Andes, Chile
C. Mpodozis, and B. L. Isacks (1990), Foreland water contents estimated from seismic velocities in (33°–36°S.L.), J. South Am. Earth Sci., 15, 117–
shortening and crustal balancing in the Andes at partially serpentinized peridotites, Geophys. Res. 139, doi:10.1016/S0895-9811(02)00009-3.
30°S latitude, Tectonics, 9(4), 789–809, Lett., 30(5), 1250, doi:10.1029/2002GL016600. Charrier, R., M. Bustamante, D. Comte, S. Elgueta, J. J.
doi:10.1029/TC009i004p00789. Carter, W. D., and L. Aguirre (1965), Structural geolo- Flynn, N. Iturra, N. Muñoz, M. Pardo, R. Thiele,
Armijo, R., R. Rauld, R. Thiele, G. Vargas, J. Campos, gy of the Aconcagua province and its relationship to and A. R. Wyss (2005), The Abanico Extensional
R. Lacassin, and E. Kausel (2010), The West the Central Valley graben, Chile, Geol. Soc. Am. Basin: Regional extension, chronology of tectonic
Andean Thrust, the San Ramón Fault, and the seis- Bull., 76, 651–664, doi:10.1130/0016-7606(1965) inversion, and relation to shallow seismic activity
mic hazard for Santiago, Chile, Tectonics, 29, 76[651:SGOAPA]2.0.CO;2. and Andean uplift, Neues Jahrb. Geol. Palaeontol.
TC2007, doi:10.1029/2008TC002427. Carter, N. L., and M. C. Tsenn (1987), Flow properties Abh., 236, 43–47.
Arriagada, C., P. R. Cobbold, and P. Roperch (2006), of continental lithosphere, Tectonophysics, 136, 27– Charrier, R., L. Pinto, and M. P. Rodriguez (2007), Tec-
Salar de Atacama basin: A record of compres- 63, doi:10.1016/0040-1951(87)90333-7. tonostratigraphic evolution of the Andean Orogen in
sional tectonics in the central Andes since the mid‐ Charrier, R. (1981), Geologie der chilenischen Haupt- Chile, in The Geology of Chile, edited by T. Moreno
kordillere zwischen 34° und 34°30′ südlicher Breite and W. Gibbons, pp. 21–114, Geol. Soc., London.

19 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Cloetingh, S., P. A. Ziegler, F. Beekman, P. A. M. estado cortical y crecimiento de los Andes central its superimposed Miocene magmatic arc, Chilean
Andriessen, N. Hardebol, and P. Dèzes (2005), Australes, paper presented at XI Congreso Geológico central Andes: First seismic and gravity evidence,
Intraplate deformation and 3D rheological structure Chileno, Univ. Católica del Norte, Antofagasta, Tectonophysics, 306, 217–236, doi:10.1016/
of the Rhine Rift System and adjacent areas of the Chile. S0040-1951(99)00046-3.
northern Alpine foreland, Int. J. Earth Sci., 94, Farías, M., R. Charrier, S. Carretier, J. Martinod, Gripp, A. E., and R. G. Gordon (2002), Young tracks of
758–778, doi:10.1007/s00531-005-0502-3. A. Fock, D. Campbell, J. Cáceres, and D. Comte hotspots and current plate velocities, Geophys. J.
Comte, D., L. Dorbath, M. Pardo, T. Monfret, H. Haessler, (2008a), Late Miocene high and rapid surface uplift Int., 150, 321–361, doi:10.1046/j.1365-
L. Rivera, M. Frogneux, B. Glass, and C. Meneses and its erosional response in the Andes of central 246X.2002.01627.x.
(1999), A double‐layered seismic zone in Arica, Chile (33° –35°S), Tectonics, 27, TC1005, Gross, K., S. Buske, S. Shapiro, and P. Wigger (2007),
northern Chile, Geophys. Res. Lett., 26, 1965– doi:10.1029/2006TC002046. Seismic imaging of the subduction zone in southern
1968, doi:10.1029/1999GL900447. Farías, M., S. Carretier, R. Charrier, J. Martinod, central Chile, paper presented at 20th Colloquium on
Comte, D., H. Haessler, L. Dorbath, M. Pardo, T. Monfret, A. Tassara, A. Encinas, and D. Comte (2008b), Latin American Earth Science, Dtsch. Forschungs-
A. Lavenu, B. Pontoise, and Y. Hello (2002), Seis- No subsidence in the development of the Central gem., Kiel, Germany.
micity and stress distribution in the Copiapo, north- Depression along the Chilean margin, paper pre- Gutscher, M. A., R. Maury, J. P. Eissen, and E. Bourdon
ern Chile subduction zone using combined on‐ and sented at 7th International Symposium on Andean (2000), Can slab melting be caused by flat subduc-
off‐shore seismic observations, Phys. Earth Planet. Geodynamics, Inst. de Rech. pour le Dév., Nice, tion?, Geology, 28, 535–538, doi:10.1130/0091-
Inter., 132, 197–217, doi:10.1016/S0031-9201(02) France. 7613(2000)28<535:CSMBCB>2.0.CO;2.
00052-3. Flynn, J. J., A. R. Wyss, R. Charrier, and C. C. Swisher Hoke, G. D., B. L. Isacks, T. E. Jordan, N. Blanco, A. J.
Comte, D., M. Farías, R. Charrier, and A. González (1995), An early Miocene anthropoid skull from the Tomlinson, and J. Ramezani (2007), Geomorphic
(2008), Active tectonics in the central Chilean Chilean Andes, Nature, 373, 603–607, doi:10.1038/ evidence for post‐10 Ma uplift of the western flank
Andes: 3D tomography based on the aftershock 373603a0. of the central Andes 18°30′–22°S, Tectonics, 26,
sequence of the 28 August 2004 shallow crustal Flynn, J. J., A. R. Wyss, D. A. Croft, and R. Charrier TC5021, doi:10.1029/2006TC002082.
earthquake, Eos Trans, AGU, 89(53), Fall Meet. (2003), The Tinguiririca fauna, Chile: Biochronol- Introcaso, A., M. C. Pacino, and H. Fraga (1992),
Suppl., Abstract T41A–1932. ogy, paleoecology, biogeography, and a new earliest Gravity, isostasy and Andean crustal shortening
Cook, F. A., and J. L. Varsek (1994), Orogen‐scale decol- Oligocene South American land mammal ‘age’, between latitudes 30° and 35°S, Tectonophysics,
lements, Rev. Geophys., 32, 37–60, doi:10.1029/ Palaeogeogr. Palaeoclimatol. Palaeoecol., 195, 205, 31–48, doi:10.1016/0040-1951(92)90416-4.
93RG02515. 229–259, doi:10.1016/S0031-0182(03)00360-2. Isacks, B. L. (1988), Uplift of the central Andean
Cristallini, E. O., and V. A. Ramos (2000), Thick‐ Fock, A., R. Charrier, M. Farías, and M. Muñoz (2006), Plateau and bending of the Bolivian Orocline,
skinned and thin‐skinned thrusting in La Ramada Fallas de vergencia oeste en la Cordillera Principal J. Geophys. Res., 93, 3211–3231, doi:10.1029/
fold and thrust belt: Crustal evolution of the High de Chile central: Inversión de la cuenca de Abanico JB093iB04p03211.
Andes of San Juan, Argentina (32° SL), Tectono- (33°–34°S), Ser. Publ. Espec. 6, pp. 48–55, Asoc. Jordan, T. E., B. L. Isacks, R. W. Allmendinger, J. A.
physics, 317, 205–235, doi:10.1016/S0040-1951 Geol. Argent., Buenos Aires. Brewer, V. A. Ramos, and C. J. Ando (1983),
(99)00276-0. Fuentes, F. (2004). Petrología y metamorfismo de muy Andean tectonics related to geometry of subducted
Croft, D. A., J. J. Flynn, and A. R. Wyss (2003), Diver- bajo grado de unidades volcánicas Oligoceno‐ Nazca plate, Geol. Soc. Am. Bull., 94, 341–361,
sification of mesotheriids (Mammalia:Notoungu- Miocenas en la ladera occidental de los Andes de doi:10.1130/0016-7606(1983)94<341:ATRTGO>
lata:Typotheria) in the middle latitudes of South Chile central (33°S). Ph.D. thesis, 407 pp. Univ. 2.0.CO;2.
America, J. Vertebr. Paleontol., 23(3, Suppl), 43A. de Chile, Santiago. Kamiya, S., and Y. Kobayashi (2000), Seismological
David, C., J. Martinod, D. Comte, G. Hérail, and García, M. (2002), Évolution oligo‐néogène de l’Alti- evidence for the existence of serpentinized wedge
H. Haessler (2002), Intracontinental seismicity and plano Occidental (Arc et Avant‐Arc du Nord du mantle, Geophys. Res. Lett., 27, 819–822,
Neogene deformation of the Andean forearc in the Chili, Arica): Tectonique, volcanisme, sédimenta- doi:10.1029/1999GL011080.
region of Arica (18.5°S–19.5°S), paper presented tion, géomorphologie et bilan érosion‐sédimentation, Kay, S. M., C. Mpodozis, V. A. Ramos, and F. Munizaga
at 5th International Symposium on Andean Geody- Ph.D. thesis, Univ. Joseph Fourier, Grenoble, France. (1991), Magma source variations for mid‐late Ter-
namics, Inst. de Rech. pour le Dév., Toulouse, García, M., and G. Hérail (2005), Fault‐related folding, tiary magmatic rocks associated with a shallowing
France. drainage network evolution and valley incision dur- subduction zone and thickening crust in the central
Davidson, J., and J.‐C. Vicente (1973), Características ing the Neogene in the Andean Precordillera of Andes, in Andean Magmatism and Its Tectonic Set-
paleogeográficas y estructurales del área fronteriza northern Chile, Geomorphology, 65(3–4), 279– ting, edited by R. S. Harmon and C. W. Rapela,
de las Nacientes del Teno (Chile) y Santa Elena 300, doi:10.1016/j.geomorph.2004.09.007. Spec. Pap. Geol. Soc. Am., 265, 113–137.
(Argentina) (Cordillera Principal, 35° a 35°15′ de Giambiagi, L. B., and V. A. Ramos (2002), Structural Kay, S. M., E. Godoy, and A. Kurtz (2005), Episodic
latitud sur), Actas Congr. Geol. Argent., V, 11–55. evolution of the Andes between 33°30′ and 33°45′ arc migration, crustal thickening, subduction ero-
Deckart, K., A. H. Clark, C. Aguilar, R. Vargas, A. Bertens, S, above the transition zone between the flat and sion, and magmatism in the south‐central Andes,
J. K. Mortensen, and M. Fanning (2005), Magmatic normal subduction segment, Argentina and Chile, Geol. Soc. Am. Bull., 117, 67–88, doi:10.1130/
and hydrothermal chronology of the giant Río J. South Am. Earth Sci., 15, 101–116, B25431.1.
Blanco porphyry copper deposit, central Chile: doi:10.1016/S0895-9811(02)00008-1. Kirby, S. H., and A. K. Kronenberg (1987), Rheology
Implications of an integrated U‐Pb and 40Ar/39Ar Giambiagi, L. B., V. A. Ramos, E. Godoy, P. P. of the lithosphere: Selected topics, Rev. Geophys.
database, Econ. Geol., 100, 905–934, doi:10.2113/ Alvarez, and S. Orts (2003a), Cenozoic deformation Space Phys., 25, 1219–1244, doi:10.1029/
100.5.905. and tectonic style of the Andes, between 33° and 34° RG025i006p01219.
Dinc‐Akdogan, N., M. Thorwart, Y. Dzierma, W. Rabbel, south latitude, Tectonics, 22(4), 1041, doi:10.1029/ Klein, F. W. (1978), Hypocenter location program
E. R. Flüh, J. Goßler, W. Taylor, and G. Alvarado 2001TC001354. HYPOINVERSE, U.S. Geol. Surv. Open File
(2007), Seismicity of southern Nicaragua and north- Giambiagi, L., P. P. Alvarez, E. Godoy, and V. A. Rep., 78‐694, 113 pp.
ern Costa Rica: A combined off‐shore and onshore Ramos (2003b), The control of pre‐existing exten- Kono, M., Y. Fukao, and A. Yamamoto (1989), Moun-
study, paper presented at EGU General Assembly, sional structures in the evolution of the southern tain building in the central Andes, J. Geophys. Res.,
Vienna, Austria. sector of the Aconcagua fold and thrust belt, Tecto- 94, 3891–3905, doi:10.1029/JB094iB04p03891.
England, P., and P. Molnar (1991), Inferences of devia- nophysics, 369, 1–19, doi:10.1016/S0040-1951(03) Kurtz, A. C., S. M. Kay, R. Charrier, and E. Farrar
toric stress in actively deforming belts from simple 00171-9. (1997), Geochronology of Miocene plutons and
physical models, Philos. Trans. R. Soc. London, Giambiagi, L., F. Bechis, V. García, and A. H. Clark exhumation history of the El Teniente region, central
Ser. A, 337, 151–164, doi:10.1098/rsta.1991.0113. (2008), Temporal and spatial relationships of thick‐ Chile (34–35°S), Rev. Geol. Chile, 24(1), 75–90.
Farías, M. (2007), Tectónica y erosión en la evolución and thin‐skinned deformation: A case study from Lamb, S. (2006), Shear stresses on megathrusts: Impli-
del relieve de los Andes de Chile central durante the Malargüe fold‐and‐thrust belt, southern central cations for mountain building behind subduction
el Neógeno/Tectonique, érosion et évolution du Andes, Tectonophysics, 459(1–4), 123–139, zones, J. Geophys. Res., 111, B07401, doi:10.1029/
relief dans les Andes du Chili central au cours du doi:10.1016/j.tecto.2007.11.069. 2005JB003916.
Néogene, Ph.D. thesis, 194 pp., Univ. de Chile, Giambiagi, L., M. Ghighlione, E. Cristallini, and Lamb, S., and P. Davis (2003), Cenozoic climate
Santiago, and Univ. de Toulouse III, Toulouse, G. Bottesi (2009), Características estructurales del change as a possible cause for the rise of the Andes,
France, 30 Nov. sector sur de la faja plegada y corrida de Malargüe Nature, 425, 792–797, doi:10.1038/nature02049.
Farías, M., R. Charrier, D. Comte, J. Martinod, and (35°–36°S): Distribución del acortamiento e influ- Lamb, S., L. Hoke, L. Kennan, and J. Dewey (1997),
G. Hérail (2005), Late Cenozoic deformation and encia de estructuras previas, Asoc. Geol. Argent. Cenozoic evolution of the central Andes in Bolivia
uplift of the western flank of the Altiplano: Evidence Rev., 65(1), 140–153. and northern Chile, in Orogeny Through Time, edi-
from the depositional, tectonic, and geomorphologic Gilbert, H., S. Beck, and G. Zandt (2006), Lithospheric ted by J.‐P. Burg and M. Ford, Geol. Soc. Spec.
evolution and shallow seismic activity (northern and upper mantle structure of central Chile and Publ., 121, 237–264.
Chile at 1 9°3 0′S), Tectonics , 24, TC40 01, Argentina, Geophys. J. Int., 165, 383–398, Lange, D., J. Cembrano, A. Rietbrock, C. Haberland,
doi:10.1029/2004TC001667. doi:10.1111/j.1365-246X.2006.02867.x. T. Dahm, and K. Bataille (2008), First seismic record
Farías, M., D. Comte, and R. Charrier (2006), Sismici- Godoy, E., G. Yáñez, and E. Vera (1999), Inversion of for intra‐arc strike‐slip tectonics along the Liquiñe‐
dad superficial en Chile central: Implicancias para el an Oligocene volcano‐tectonic basin and uplift of Ofqui fault zone at the obliquely convergent plate

20 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

margin of the southern Andes, Tectonophysics, J. Geophys. Res., 104, 14,965–14,992, doi:10.1029/ B. W. Ticherlaar and L. R. Ruff, J. Geophys. Res.,
455(1–4), 14–24, doi:10.1016/j.tecto.2008.04.014. 1999JB900060. 98, 15,825–15,828, doi:10.1029/93JB00234.
Legrand, D., B. Delouis, L. Dorbath, C. David, J. Campos, Oliveros, V., L. Aguirre, D. Morata, A. Simonetti, Syracuse, E. M., G. A. Abers, K. Fischer, L. MacKenzie,
L. Marquéz, J. Thompson, and D. Comte (2007), M. Vergara, M. Belmar, and S. Calderon (2008), C. Rychert, M. Protti, V. González, and W. Strauch
Source parameters of the Mw = 6.3 Aroma crustal Geochronology of very low‐grade Mesozoic (2008), Seismic tomography and earthquake loca-
earthquake of July 24, 2001 (northern Chile), and Andean metabasites; an approach through the tions in the Nicaraguan and Costa Rican upper man-
its aftershock sequence, J. South Am. Earth Sci., K‐Ar, 40Ar/39A and U‐Pb LA‐MC‐ICP‐MS meth- tle, Geochem. Geophys. Geosyst., 9, Q07S08,
24, 58–68, doi:10.1016/j.jsames.2007.02.004. ods, J. Geol. Soc., 165, 579–584, doi:10.1144/ doi:10.1029/2008GC001963.
Levi, B., L. Aguirre, J. Nyström, H. Padilla, and 0016-76492007-113. Tassara, A. (2005), Interaction between the Nazca and
M. Vergara (1989), Low‐grade regional metamor- Pardo, M., D. Comte, and T. Monfret (2002), Seismo- South American plates and formation of the Altiplano‐
phism in the Mesozoic‐Cenozoic volcanic sequences tectonic and stress distribution in the central Chile Puna plateau: Review of a flexural analysis along
of the central Chile, J. Metamorph. Petrol., 7, 487– subduction zone, J. South Am. Earth Sci., 15, 11– the Andean margin (15°–34°S), Tectonophysics,
495, doi:10.1111/j.1525-1314.1989.tb00611.x. 22, doi:10.1016/S0895-9811(02)00003-2. 399, 39–57, doi:10.1016/j.tecto.2004.12.014.
Maksaev, V., F. Munizaga, M. McWilliams, M. Fanning, Pardo, M., E. Vera, T. Monfret, and G. Yáñez (2008), Tassara, A., H.‐J. Götze, S. Schmidt, and R. Hackney
R. Mathur, J. Ruíz, and M. Zentilli (2004), New Crustal seismicity and 3D seismic wave velocity (2006), Three‐dimensional density model of the
chronology for El Teniente, Chilean Andes, from models in the Andes cordillera of central Chile Nazca plate and the Andean continental margin,
U‐Pb, 40Ar/39Ar, Re‐Os, and fission‐track dating: (33°–34.5°S) from local earthquakes, paper pre- J. Geophys. Res., 111, B09404, doi:10.1029/
Implications for the evolution of supergiant porphyry sented at 7th International Symposium on Andean 2005JB003976.
Cu‐Mo deposits, in Andean Metallogeny: New Dis- Geodynamics, Inst. de Rech. pour le Dév., Nice, Tassara, A., C. Swain, R. Hackney, and J. Kirby (2007),
coveries, Concepts and Updates, edited by R. H. France. Elastic thickness of South America estimated using
Sillitoe et al., Spec. Publ. SEPM Soc. Sediment. Pérez‐Gussinyé, M., A. R. Lowry, and A. B. Watts wavelets and satellite‐derived gravity data, Earth
Geol., 11, 15–54. (2007), Effective elastic thickness of South America Planet. Sci. Lett., 253, 17–36, doi:10.1016/j.epsl.2006.
Maksaev, V., F. Munizaga, M. Zentilli, and R. Charrier and its implications for intracontinental deforma- 10.008.
(2009), Fission track thermochronology of Neo- tion, Geochem. Geophys. Geosyst., 8, Q05009, Thiele, R. (1980), Hoja Santiago, Región Metropoli-
gene plutons in the Principal Andean Cordillera doi:10.1029/2006GC001511. tana, Carta Geol. Chile 39, 51 pp., Inst. de Invest.
of central Chile (33–35°S): Implications for tectonic Pilger, R. H. (1981), Plate reconstruction, aseismic Geol., Santiago, Chile.
evolution and porphyry Cu‐Mo mineralization, ridges, and low angle subduction beneath the Thierer, P. O., E. R. Flüh, H. Kopp, F. Tilmann,
Andean Geol., 36(2), 153–171, doi:10.4067/S0718- Andes, Geol. Soc. Am. Bull., 92, 448–456, D. Comte, and S. Contreras (2005), Local earthquake
71062009000200001. doi:10.1130/0016-7606(1981)92<448: monitoring offshore Valparaiso, Chile, Neues Jahrb.
McClay, K. R. (1995), The geometries and kinematics PRARAL>2.0.CO;2. Geol. Palaeontol. Abh., 236, 173–183.
of inverted fault systems: A review of analogue Pose, F. A., M. Spagnuolo, and A. Folguera (2005), Thomas, H. (1958), Geología de la Cordillera de la Costa
model studies, in Basin Inversion, edited by P. G. Modelo para la variación del volumen orogénico entre el Valle de la Ligua y la Cuesta de Barriga, Bol.
Buchanana, Geol. Soc. Spec. Publ., 88, 97–118, andino y acortamientos en el sector 20°–46°S, Asoc. Inst. Invest. Geol. Chile, 2, 86 pp.
doi:10.1144/GSL.SP.1995.088.01.07. Geol. Argent. Rev., 60(4), 724–730. Uliana, M., K. Biddle, and J. Cerdán (1989), Mesozoic
McQuarrie, N. (2002), The kinematic history of the cen- Ramos, V. A., M. Cegarra, and E. Cristallini (1996), extension and the formation of Argentina sedimen-
tral Andean fold‐thrust belt, Bolivia: Implications Cenozoic tectonics of the High Andes of west‐ tary basins, in Extensional Tectonics and Stratigra-
for building a high plateau, Geol. Soc. Am. Bull., central Argentina (30–36°S latitude), Tectonophy- phy of the North Atlantic Margin, edited by A. J.
114, 950–963, doi:10.1130/0016-7606(2002) sics, 259, 185–200, doi:10.1016/0040-1951(95) Tankard and H. R. Balkwill, AAPG Mem., 46,
114<0950:TKHOTC>2.0.CO;2. 00064-X. 599–613.
McQuarrie, N., J. B. Barnes, and T. A. Ehlers (2008), Ramos, V. A., E. O. Cristallini, and D. J. Pérez (2002), Uyeda, S., and H. Kanamori (1979), Back‐arc opening
Geometric, kinematic, and erosional history of the The Pampean flat‐slab of central Andes, J. South and the mode of subduction, J. Geophys. Res., 84,
central Andean Plateau, Bolivia (15–17°S), Tecton- Am. Earth Sci., 15(1), 59–78, doi:10.1016/S0895- 1049–1061, doi:10.1029/JB084iB03p01049.
ics, 27, TC3007, doi:10.1029/2006TC002054. 9811(02)00006-8. Vergara, M., B. Levi, and R. Villarroel (1993), Geothermal‐
Mishra, O. P., D. Zhao, N. Umino, and A. Hasagawa Ramos, V. A., T. Zapata, E. Cristallini, and A. Introcaso type alteration in a burial metamorphosed volcanic
(2003), Tomography of northern Japan forearc and (2004), The Andean thrust system—Latitudinal var- pile, central Chile, J. Metamorph. Geol., 11, 449–
its implications for interplate seismic coupling, iations in structural styles and orogenic shortening, 454, doi:10.1111/j.1525-1314.1993.tb00161.x.
Geophys. Res. Lett., 30(16), 1850, doi:10.1029/ in Thrust Tectonics and Hydrocarbon Systems, Vergés, J., V. A. Ramos, A. Meigs, E. Cristallini, F. H.
2003GL017736. edited by K. R. McClay, AAPG Mem., 82, 30–50. Bettini, and J. M. Cortés (2007), Crustal wedging
Montecinos, P., U. Chärer, M. Vergara, and L. Aguirre Rauld, R., G. Vargas, R. Armijo, A. Ormeño, C. Valderas, triggering recent deformation in the Andean thrust
(2008), Lithospheric origin of Oligocene‐Miocene and J. Campos (2006), Cuantificación de escarpes de front between 31°S and 33°S: Sierras Pampeanas‐
magmatism in central Chile: U‐Pb ages and Sr‐ falla y deformación reciente en el frente cordillerano Precordillera interaction, J. Geophys. Res., 112,
Pb‐Hf isotope composition of minerals, J. Petrol., de Santiago, paper presented at XI Congreso Geoló- B03S15, doi:10.1029/2006JB004287.
49(3), 555–580, doi:10.1093/petrology/egn004. gico Chileno, Univ. Católica del Norte, Antogafasta, Victor, P., O. Oncken, and J. Glodny (2004), Uplift of
Mpodozis, C., and V. Ramos (1989), The Andes of Chile. the western Altiplano plateau: Evidence from the
Chile and Argentina, in Geology of the Andes and Riquelme, R., G. Hérail, J. Martinod, R. Charrier, and Precordillera between 20° and 21°S (northern
Its Relation to Hydrocarbon and Mineral J. Darrozes (2007), Late Cenozoic geomorphologic Chile), Tectonics, 23, TC4004, doi:10.1029/
Resources, Earth Sci. Ser., vol. 11, edited by signal of Andean forearc deformation and tilting 2003TC001519.
G. E. Ericksen et al., pp. 59–90, Circum‐Pac. associated with the uplift and climate changes of Wessel, P., and W. H. F. Smith (1998), New, improved
Counc. for Energy and Miner. Resour., Houston, the southern Atacama Desert (26°S–28°S), Geomor- version of the Generic Mapping Tools released, Eos
Tex. phology, 86, 283–306, doi:10.1016/j.geomorph.2006. Trans. AGU, 79(47), 579, doi:10.1029/98EO00426.
Muñoz, M., F. Fuentes, M. Vergara, L. Aguirre, J. O. 09.004. Wyss, A. R., M. A. Norell, J. J. Flynn, M. J. Novacek,
Nyström, G. Féraud, and A. Demant (2006), Aba- Roecker, S.W., T. M. Sabitova, L. P. Vinnik, Y. A. R. Charrier, M. C. McKenna, D. Frassinetti, P. Salinas,
nico East Formation: Petrology and geochemistry Burmakov, M. I. Golvanov, R. Mamatkanova, and J. Meng (1990), A new early Tertiary mammal
of volcanic rocks behind the Cenozoic arc front in and L. Munirova (1993), Three‐dimensional elastic fauna from central Chile: Implications for stratigra-
the Andean Cordillera, central Chile (33°50′S), wave velocity structure of the western and central phy and tectonics, J. Vertebr. Paleontol., 10, 518–
Rev. Geol. Chile, 33, 109–140, doi:10.4067/ Tien Shan, J. Geophys. Res., 98, 15,779–15,795, 522.
S0716-02082006000100005. doi:10.1029/93JB01560. Yáñez, G., and J. Cembrano (2004), Role of viscous
Muñoz, N., and R. Charrier (1996), Uplift of the west- Servicio Geológico Minero Argentino (1997), Mapa plate coupling in the late Tertiary Andean tectonics,
ern border of the Altiplano on a west‐vergent thrust geológico de la República Argentina, scale J. Geophys. Res., 109, B02407, doi:10.1029/
system, northern Chile, J. South Am. Earth Sci., 9, 1:2,500,000, Buenos Aires, Argentina. 2003JB002494.
171–181, doi:10.1016/0895-9811(96)00004-1. Servicio Nacional de Geología y Minería (2002), Mapa Yáñez, G., C. R. Ranero, R. von Huene, and J. Díaz
Nur, A., and Z. Ben‐Avraham (1981), Volcanic gaps geológico de Chile, scale 1:1,000,000, Map M61, (2001), Magnetic anomaly interpretation across the
and the consumption of aseismic ridges in South Santiago, Chile. southern central Andes (32°–34°S): The role of
America, in Nazca Plate: Crustal Formation and Smith, W. H. F., and D. T. Sandwell (1997), Global sea the Juan Fernández Ridge in the late Tertiary evolu-
Andean Convergence, edited by L. D. Kulm, floor topography from satellite altimetry and ship tion of the margin, J. Geophys. Res., 106, 6325–
Mem. Geol. Soc. Am., 154, 729–740. depth soundings, Science, 277, 1956–1962, 6345, doi:10.1029/2000JB900337.
Oleskevich, D. A., R. D. Hyndman, and K. Wang doi:10.1126/science.277.5334.1956. Zhao, D., K. Asamori, and H. Iwamori (2000), Seismic
(1999), The updip and downdip limits to great sub- Suarez, G., and D. Comte (1993), Comment on “Seismic structure and magmatism of the young Kyushu sub-
duction earthquakes: Thermal and structural models coupling along the Chilean subduction zone” by duction zone, Geophys. Res. Lett., 27, 2057–2060,
of Cascadia, south Alaska, SW Japan, and Chile, doi:10.1029/2000GL011512.

21 of 22
TC3006 FARÍAS ET AL.: CENTRAL CHILE ANDES ARCHITECTURE TC3006

Zhao, D., K. Wang, G. C. Rogers, and S. M. Peacock Chile, Plaza Ercilla 803, Casilla 13518, Correo 21, J. Martinod, LMTG, CNRS‐IRD‐Université de
(2001), Tomographic image of low P wave velocity Santiago, Chile. (mfarias@dgf.uchile.cl) Toulouse, 14 av. Edouard Belin, F‐31400 Toulouse,
anomalies above slab in northern Cascadia subduc- D. Comte, Departamento de Geofísica, FCFM, France.
tion zone, Earth Planets Space, 53(4), 285–293. Universidad de Chile, Blanco Encalada 2002, Casilla A. Tassara, Departamento de Ciencias de la Tierra,
2777, Correo 21, Santiago, Chile. Universidad de Concepción, Campus Concepción,
A. Fock, SQM Salar S.A., Los Militares 4290, Casilla 160‐C, Concepción, Chile. (andres.tassara@
R. Charrier, C. David, M. Farías, and F. Tapia, 7550081, Las Condes, Chile. (andres.fock@sqm.com) udec.cl)
Departamento de Geología, FCFM, Universidad de

22 of 22

View publication stats

You might also like