You are on page 1of 25

©2023 Society of Economic Geologists, Inc.

Economic Geology, v. XXX, no. XX, pp. X–X

A Model for the Lithospheric Architecture of the Central Andes and the
Localization of Giant Porphyry Copper Deposit Clusters
Alexander D. Farrar,1,2,† David R. Cooke,1 Jon M.A. Hronsky,3, 4 David G. Wood,2,* Sebastian B. Benavides,2,**
Matthew J. Cracknell,1 James F. Banyard,2 Santiago Gigola,2,5 Tim Ireland,2 Simon M. Jones,2 and José Piquer6
1 Centre for Ore Deposits and Earth Sciences (CODES), University of Tasmania, Private Bag 79, Hobart, Tasmania 7001, Australia
2 First Quantum Minerals, Apoquindo 4445 – piso 6, Las Condes, Santiago, Chile
3 Western Mining Services Pty. Ltd., West Perth, Western Australia 6005, Australia
4 Centre for Exploration Targeting, School of Earth Sciences, University of Western Australia, 35 Stirling Highway,
Crawley, Western Australia 6009, Australia
5 Independent Consultant, Los Pimientos 1107, Las Heras, Mendoza 5539, Argentina
6 Eduardo Morales Miranda Ave., Emilio Pugín Building, Isla Teja Campus, Universidad Austral de Chile, Valdivia, Chile

Abstract
In the central Andes, giant porphyry copper deposits of similar ages group into discrete geographic clusters
that are regularly spaced and aligned within orogen-parallel belts. This clustering highlights how exceptional
geologic processes affected localized regions of the lithosphere during mineralization and that the spatial
and temporal distribution of giant porphyry deposits is nonrandom. Development of favorable regions of
lithosphere for significant metal concentration are linked to the overlap of structural pathways that focus fluid
and magma flow from the mantle to upper crust during high-horizontal-compressive-strain events. These
structural pathways are notoriously difficult to identify in the field due to their often-subtle surficial mani-
festations and continental scale. Field mapping at multiple scales in northwest Argentina and southern Peru,
as well as regional structural traverses throughout the central Andes, indicates the presence of regional-scale
structural corridors 5 to 25 km wide and hundreds of km long that consist of myriad fault planes. The variable
width and diffuse surface expression of these corridors is interpreted to reflect the upward propagation of
underlying zones of basement weakness through younger supracrustal sequences in the overriding plate. Such
structural corridors are (1) apparent at multiple scales of investigation, (2) long-lived, (3) preferentially reac-
tivated though time, and (4) evident in geophysical data sets. This structural architecture formed in response
to the interplay of pre-Cenozoic tectonics and the orientation of inherited structural weaknesses. These fault
systems persist in the upper crust as steep zones of enhanced permeability that can preferentially reactivate as
pathways for ascending hydrous magmas and fluids during major deformation events. Linear orogen-parallel
structural belts cogenetic with the magmatic arc provide the first-order control to giant porphyry copper
deposit distribution. The second-order control is the intersection of orogen-oblique structural corridors with
the orogen-parallel belts, localizing deposit clusters at these intersections. Such regions are inferred to have
been zones of deep permeability, with vertical translithospheric pathways activated during high-strain tectonic
events that affected the intra-arc stress field.

Introduction geographic areas underwent favorable structural and geody-


The central Andes, between 14°S and 35°S latitude, is the namic preconditioning for porphyry copper deposit genesis
most important copper province on the planet, accounting for is of obvious benefit to mineral explorers. This knowledge
approximately 40% of the world’s annual copper mine pro- is particularly valuable in regions such as the central Andes,
duction (U.S. Geological Survey [USGS], 2021). The decades where postmineralization cover may obscure undiscovered
leading up to the year 2000 saw numerous greenfields porphy- mineral deposits.
ry copper discoveries in the central Andes (Sillitoe and Perel- Giant porphyry copper deposits (>3.1 Mt of contained
ló, 2005); however, since 2000, few significant new greenfields copper; Clark, 1993) in the central Andes align within linear,
porphyry discoveries have been reported (e.g., Sillitoe, 2010a; orogen-parallel belts that spatially and temporally coincide
Rode et al., 2015; Schodde, 2019; Sillitoe et al., 2019). Declin- with the magmatic arc (Fig. 1; Cooke et al., 2005; Sillitoe and
ing discovery rates in the central Andes are likely a function Perelló, 2005; Sillitoe, 2010b) and have been noted to exhibit
of increasing exploration maturity and extensive postminer- marked along-arc spatial periodicity that is not explained by
alization cover, issues that affect many other mineral belts upper crustal or stochastic processes alone (Yáñez and Mak-
worldwide (McCuaig and Hronsky, 2014). Predicting which saev, 1994; Richards, 2003; Tomlinson and Cornejo, 2012;
Hayward et al., 2018). At the district scale, giant deposits tend
to occur as discrete, multideposit clusters of similar age (Fig.
†Corresponding author: e-mail, alex.farrar@utas.edu.au 1B-D; Sillitoe, 2010b; Hayward et al., 2018). A “giant por-
*Present address: Anglo American, 201 Charlotte St., Brisbane, Queensland phyry copper deposit cluster” is spatially defined as a grouping
4000, Australia.
**Present address: Anglo American, Calle Esquilache 371 – piso 10, San of porphyry copper deposits and prospects, belonging to the
Isidro, Lima 15073, Peru. same metallogenic epoch, that contains at least one deposit

ISSN 0361-0128; doi:10.5382/econgeo.5010; 26 p. 1 Submitted: August 25, 2022/Accepted: January 29, 2023

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
2 FARRAR ET AL.

Fig. 1. A) Simplified geologic map of the central Andes, colored by the tectonic cycles defined by Mpodozis and Ramos (1989)
and Charrier et al. (2007). The locations of interpreted suture zones between basement terranes are shown as dashed lines
(see text for discussion). Giant porphyry copper deposits are colored by metallogenic age (after Sillitoe and Perelló, 2005).
B) Middle Miocene-Pliocene metallogenic belt of central Chile; five deposit clusters as defined by Hayward et al. (2018) are
separated by 90 ± 15 km. C) Middle Eocene-early Oligocene metallogenic belt of northern Chile; four deposit clusters are
separated by 115 ± 10 km. Two clusters of Paleocene-early Eocene deposits are separated by 70 km. D) Middle Eocene-
early Oligocene Andahuaylas-Yauri belt; four deposit clusters are separated by 75 ± 10 km. Two clusters of Paleocene-early
Eocene deposits are separated by 120 km. Background fault interpretations from Segemar (1997), Ingemmet (1999), and
Sernageomin (2002).

that is giant (e.g., Hayward et al., 2018; Fig. 1B-D). In the 2012; Hayward et al., 2018), and, in southern Peru, clusters of
central Andes, giant deposits within the same cluster are typi- giant porphyry and skarn deposits in the middle Eocene-early
cally located <10 km from the cluster centroid (Fig. 1B-D); Oligocene Andahuaylas-Yauri belt display a periodicity of ap-
sometimes this distance can extend to 30 km due to postmin- proximately 70 km (Fig. 1D; Perelló et al., 2003).
eralization fault displacement (e.g., Chuquicamata cluster; Emplacement of porphyry copper deposits into regularly
Dilles et al., 1997; Fig. 1C). Within a cluster, the location of spaced clusters along arc-parallel structural belts has previ-
ore deposition is controlled by upper crustal structural, rheo- ously been proposed to reflect the role of regional-scale arc-
logical, and physiochemical processes that also determine the oblique structural corridors (Salfity, 1985; Richards et al.,
tonnage and grade of individual porphyry deposits (Sillitoe, 2001; Gow and Walshe, 2005; Piquer et al., 2016; Yáñez and
2010b). The middle Miocene-Pliocene deposits of central Rivera, 2019). This study draws upon the mineral system con-
Chile exhibit a cluster periodicity of 80 to 100 km (Fig. 1B; cept (Wyborn et al., 1994; McCuaig et al., 2010; McCuaig and
Hayward et al., 2018), the middle Eocene-early Oligocene de- Hronsky, 2014) to interrogate the role of these arc-oblique
posits of northern Chile exhibit a periodicity of 110 to 120 km structures as second-order controls on porphyry emplace-
(Fig 1C; Yáñez and Maksaev, 1994; Tomlinson and Cornejo, ment. Hereafter, we use the terms “lineament” to describe

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 3

regional-scale (>50-km-long) linear features inferred from arc-parallel “metallogenic belt” paradigm as the first-order
geophysical or geologic data and “structural corridor” when control to deposit distribution (e.g., Camus and Dilles, 2001;
field evidence supports numerous parallel faults that spatially Sillitoe and Perelló, 2005; Acosta et al., 2009).
coincide with the trace of a lineament along its strike (e.g., Recent studies continue to emphasize the importance of
Hayward et al., 2018). The central Andes contains 49 known arc-oblique fault zones as inherited orientations of weakness
giant porphyry copper deposits within 25 discrete clusters. in basement rocks (Allmendinger et al., 1983; Charrier et al.,
The region is characterized by a long magmatic and orogenic 2007) that controlled the distribution of pre-Andean sedimen-
history, active tectonics, and a lack of vegetation. Coupled tary basins (Ramos, 1994; Jacques, 2003; Amilibia et al., 2008;
with high-quality remote sensing data sets and a rich body Espinoza et al., 2019) and volcanic units (Cembrano and Lara,
of geologic research, the area is an ideal site for investigating 2009; Lavallée et al., 2009; Gioncada et al., 2010; Lanza et
crustal-scale tectonic controls on porphyry copper emplace- al., 2013). Though these characteristics of arc-oblique fault
ment. We present new mapping data that reveal how first- and zones are widely accepted, their role in the localization of por-
second-order controls on porphyry emplacement—continen- phyry copper deposits is controversial (Gilluly, 1976; Paterson
tal-scale lineaments—can manifest as subtle structural corri- and Schmidt, 1999; Richards, 2000; Tomlinson and Cornejo,
dors in the field (Salfity, 1985; Richards et al., 2001; Lanza et 2012). In the central Andes, this stems from the sometimes
al., 2013; Piquer et al., 2016). This study provides important subtle and at other times total lack of surficial geologic evi-
insight into how these structural corridors create the focused dence for segments of arc-oblique fault systems, which leads
permeability necessary for the localization of giant porphyry to uncertainty regarding their position and existence. Defining
copper deposit clusters in the central Andes and elsewhere in a causal link between such enigmatic fault systems and min-
the world. eralization, in a way comparable to well-defined arc-parallel
basement structures, such as the Domeyko and Incapuquio
Arc-Oblique Structural Corridors in the Central fault zones (e.g., Reutter et al., 1996; Carlotto et al., 2010), is
Andes—An Ongoing Controversy therefore challenging. Discrepancies between studies regard-
The development of fertile upper-crustal magmas requires ing the trace of some arc-oblique structural corridors, a lack of
hydrated and oxidized mafic melts generated above the discussion around levels of uncertainty and inconsistent ter-
subducted slab that accumulate in the lower crustal MASH minology may accentuate skepticism (e.g., Salfity, 1985; Behn
(melting, assimilation, storage, homogenization) zone. Melts et al., 2001; Richards et al., 2001; Jacques, 2003; Hayward
subsequently ascend though the crust via staged ponding et al., 2018; Yáñez and Rivera, 2019; Espinoza et al., 2021).
chambers, promoting magma fractionation and the delivery Discrepancies between structural models likely stem from
of oxidized magmatic volatile phases enriched in Cu and S to the use of differing data sets, spatial limitations to study areas
higher levels of the crust (Richards et al., 2003; Tosdal and (e.g., data set extents, country borders), and the intrinsic dif-
Dilles, 2020; Loucks, 2021). This process can eventually lead ficulty in validating cryptic structural corridors at the surface
to porphyry ore formation via fluid release from pipes, dikes, (McCuaig and Hronsky, 2014).
and stocks that are expelled from lower to midcrustal storage The Rayleigh-Taylor gravitational diapirism model is an al-
chambers. The conspicuous grouping of giant porphyry de- ternate mechanism that may explain regularly spaced giant
posits into periodically spaced clusters in central Andean mag- porphyry copper deposit clusters in the central Andes without
matic arcs implies that these stages of magma migration, stall- the requirement of arc-oblique structural corridors (Yáñez
ing, and fluid release are highly localized. This has led many and Maksaev, 1994; Tomlinson and Cornejo, 2012). This
workers to emphasize the importance of regional-scale, arc- model emphasizes the plume-like ascent of magmatic fluids
oblique structural corridors that localize deposits where they produced during dehydration of the downgoing slab that initi-
intersect the arc in different porphyry belts around the world ates from Rayleigh-Taylor gravitational instabilities, with the
(e.g., Glen and Walshe, 1999; Hill et al., 2002; Richards, 2003; equidistant spacing of ascending diapirs controlled by the
Love et al., 2004; Piquer et al., 2016; Wiemer et al., 2022). thickness of the overlying layers and the viscosity difference
Early studies by Mayo (1958) and Anderson (1966) produced of the ascending and descending mixing fluids. Drawbacks of
“lineament maps” in the southwestern United States, where the Rayleigh-Taylor model are that magma ascent through the
deposits were correlated across tracts of postmineralization lithosphere is widely regarded to be along preexisting perme-
cover and the importance of inherited Precambrian structures ability as dike-like fingers, and the model does not adequately
and structural intersections for controlling porphyry copper explain the reoccurrence of magmatism of multiple ages in
deposits were noted. In the central Andes, early studies that the same location, linear trends of coeval intrusions, or arc-
correlated arc-oblique structural trends with the localization oblique trends of volcanism, phenomena that are adequately
of porphyry copper deposits were conducted by Segerstrom accounted for in the arc-oblique structural corridor model
(1970), Thomas (1970), Ricci and Figueroa (1971), and Salfity (Richards and Villeneuve, 2002; Tomlinson and Cornejo,
(1985). Lowell (1974) and Sillitoe (1988) carried out pioneer- 2012; Hayward et al., 2018).
ing studies emphasizing the localization of porphyry copper
deposits into orogen-parallel belts, some of which centered Tectonic History of the Central Andean Margin
around arc-parallel fault systems, such as the Domeyko fault The western margin of South America has been an ocean-
system in Chile. The model of Sillitoe (1988) grouped por- facing continental margin since the late Neoproterozoic
phyry deposits into narrow, elongate structural belts of similar breakup of Rodinia, when the proto-Pacific Ocean opened
age, which has proven robust and repeatable, with subsequent (Merdith et al., 2021). A series of para-allochthonous and al-
dating of mineral deposits confirming the significance of the lochthonous terranes, likely sourced from Laurentia, were ac-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
4 FARRAR ET AL.

creted to the western margin of proto-South America during Ghidella, 2012). Separation of South America from Africa in
the Pampean and Famatinian tectonic cycles (Mpodozis and the middle Cretaceous resulted in westward acceleration of
Ramos, 1989; Charrier et al., 2007; Ramos, 2010; Table 1). the South American plate, which significantly increased com-
Final amalgamation of the central Andes was achieved with pressional stress along the trench, reducing the slab subduc-
the accretion of the Chilenia terrane in the Devonian (Álvarez tion angle and shifting the magmatic arc to the east (Somoza
et al., 2011). There are unresolved questions concerning the and Zaffarana, 2008). Fluctuations in the far-field stress re-
Paleozoic history of the central Andes, including the mode gime, possibly due to changing plate convergence conditions
of accretion of allochthonous terranes (strike-slip versus col- controlled by the absolute westward velocity of South Amer-
lisional), whether some terranes accreted separately or as ica (Oncken et al., 2006), manifested as a series of contrac-
larger coherent terranes (e.g., Arequipa-Antofalla, Famatina, tional events known as the Peruvian (Late Cretaceous), K-T
Mejillonia), the locations of suture zones between terranes, (Paleocene), Incaic (middle Eocene-early Oligocene), and
and the structural controls on synorogenic basin evolution Quechua (Late Cenozoic) episodes, which were separated
(Ramos, 2010; Horton, 2018; Franceschinis et al., 2020). by periods of relative quiescence and extensional reactivation
Following the accretionary stage of central Andean evolu- of older structures (e.g., Sempere et al., 1990; Oncken et al.,
tion, a series of punctuated extensional and compressional 2006; Charrier et al., 2007). The Bolivian Orocline (Fig. 1A)
tectonic cycles occurred—the Gondwanan, Pre-Andean, and is interpreted to have developed during these events (Isacks,
First Andean cycles (Table 1; Fig. 1A). These tectonic cycles 1988; Allmendinger et al., 2005); however, estimates of short-
were accompanied by the formation of significant NW- and ening, timing, and the tectonic trigger that led to the com-
N-striking extensional sedimentary basins, whose extensional mencement of oroclinal bending are debated (e.g., Arriagada
axes utilized the inherited structural architecture associated et al., 2008; Eichelberger and McQuarrie, 2015; Schellart,
with earlier tectonic cycles and the suture zones of basement 2017; Wiemer et al., 2022).
terranes (Allmendinger et al., 1983; Ramos 1994; Charrier et
al., 2007; Espinoza et al., 2019). Crustal extension during the Methods
early Silurian to late Carboniferous was sufficiently severe
to rift part of the Arequipa-Antofalla basement terrane away Geology preparation
from southwest Pangea (Bahlburg and Hervé, 1997; Ramos, Geology map sheets from Peru (Ingemmet, 1999), Chile
2010). Subsequent Triassic extension was associated with (Sernageomin, 2002), Argentina (Segemar, 1997), and Bo-
transform plate convergence (Müller et al., 2019; Young et al., livia (Sergeomin, 2020) were digitized into a coherent, cross-
2019). During this time, the initiation of rifting and breakup border digital geologic map of the central Andes as attributed
of Pangea produced a series of NW- and N-elongated failed polygons. Several 1:100k and 1:250k map sheets were utilized
rifts and sedimentary basins along the margin of southwest for site-specific studies, such as those referenced in the case
Pangea (Espinoza et al., 2021). The Second and Final Andean studies. We created a Geographic Information System (GIS)
phases were associated with periods of oblique plate conver- workspace that incorporated the digitized geology, geophysi-
gence and variations in the convergence rate (Somoza and cal data discussed below, suture zone information, and our

Table 1. Tectonic Cycles that Affected the Central Andes (after Coira et al. [1982], Mpodozis and Ramos [1989], and Charrier et al. [2007])

Tectonic cycle From To Key features


Final Andean phase Upper Eocene Present Present configuration of the Andes adopted; fluctuating subduction convergence rate
and convergence obliquity; flat slab subduction events in Peru and Chile affects upper
plate stress regimes; three temporally distinct giant porphyry events

Second Andean phase Upper Cretaceous Lower Eocene Full separation of South America from Africa causes tectonic inversion of the back arc,
and compression in forearc shifts volcanic eastward

First Andean phase Lower Jurassic Mid-Cretaceous La Negra volcanic arc extends from southern Peru to southern Chile along present-day
coastline; back-arc basins form in Peru, Chile, and Argentina deposited on top of the
Gondwanan magmatic arc; Atacama fault zone active

Pre-Andean Upper Permian Lower Jurassic Subduction slows/stops due to consolidation of Pangea; rifting of Pangea and oblique
subduction produces an extensional regime; voluminous felsic magmatism (e.g.,
Choiyoi, Carabaya); NW segmented extensional basins in Chile, Argentina, and Peru
(Marañon, Mitu, Domeyko, Cuyo, Neuquén)

Gondwanan Upper Devonian Lower Permian Extension, rifting off part of the Arequipa-Antofalla terrane, followed by passive margin;
from late Carboniferous N-S magmatic arc established along the cordillera of Chile
and western Argentina; NW and NE truncated back arc basins

Famatinian Upper Cambrian Lower Devonian Accretion of Arequipa-Antofalla, Famatinia (?), Cuyania, Chilenia terranes; volcanic
arc in northwest Argentina, thick back-arc sedimentary sequences deposited in Peru,
Bolivia, and Argentina controlled by NW- and NE-trending basin-bounding faults

Pampean Neoproterozoic Lower Cambrian Breakup of Rodinia, opening of proto-Pacific Ocean, onset of subduction, accretion of
Pampia; thick sedimentation in northwest Argentina

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 5

field mapping (First Quantum Minerals, unpub. data, 2020). 2010), isotopic data (Carlier et al., 2005; Mamani et al., 2008),
We assigned each lithostratigraphic unit to one of a simplified and geophysical data (Peri et al., 2015; Franceschinis et al.,
set of formational modes: intrusion, volcanogenic, shallow 2020). The interpretation of terrane sutures (Figs. 1A, 2A)
marine, deep marine, and metamorphic, according to geolog- represents a best fit to published location with some fine tun-
ic provenance. This reduced edge effects between individual ing from the regional gravimetry and geology data sets.
map sheets and mitigated mismatches caused by the change
in scale of mapping between countries and map sheets. Ad- Tectono-lithologic interpretation
ditionally, each lithostratigraphic unit was assigned to one of Published regional-scale faults were incorporated in our cen-
seven tectonic cycles according to age (Table 1; Fig. 1A). tral Andes tectono-lithologic interpretation (Ramos, 1977;
Allmendinger et al., 1983; Jordan et al., 1983; Salfity, 1985;
Geophysical interpretations Scheuber and Andriessen, 1990; Martinez et al., 1995; Cem-
The gridded isostatic residual gravity model and the first verti- brano et al., 1996; Astini and Thomas, 1999; Richards et al.,
cal derivative filter of the isostatic residual gravity model (Fig. 2001; Bouzari and Clark, 2002; Perelló et al., 2003; Carlier et
2A) are based on commercial and public ground station data al., 2005; Amilibia et al., 2008; Audin et al., 2008; Lavallée et
provided by First Quantum Minerals. These data were con- al., 2009; Tibaldi et al., 2009; Roperch et al., 2011; Carlotto,
verted from observed gravity to a Bouguer anomaly using the 2013; Lanza et al., 2013; Piquer et al., 2016, 2017, 2021b; Es-
elevation measured at each station with a Bouguer reduction pinoza et al., 2019, 2021). We conducted a structural interpre-
density of 2.67 g/cm³. The isostatic residual gravity anomaly tation of geology for the tectonic cycles outlined in Table 1 at
was calculated based on the Airy approximation using a re- a scale of 1:1M by visualizing rocks with ages of each tectonic
duction density of 2.67 g/cm³ onshore and 2.2 g/cm³ offshore, cycle. As a proxy for potential basement fault systems that
a density contrast at the Moho of 0.5 g/cm³ and a reference might be reactivated in different ways through geologic time,
Moho depth of 31 km. A Canny filter with an upward continu- attention was paid to identifying the margins and truncations
ation of 10,000 m was applied to the first vertical derivative of pre-Andean sedimentary basins (Ramos, 1994; Amilibia et
of the isostatic residual gravity model (Fig. 2A; Xiaodi et al., al., 2008; Espinoza et al., 2019), axes of rifts, axes of volcanic
2018). Lineament interpretation was carried out by hand at arcs, aligned plutons, regions that underwent rapid crustal
a scale of approximately 1:3,000,000 (1:3M) and focused on thickening or thinning, as well as upper Cenozoic aligned vol-
identifying continuous lineaments that follow abrupt contrasts canic centers inboard of the main volcanic arc (Salfity, 1985;
in bulk density (e.g., Yáñez and Rivera, 2019). The Canny fil- Chernicoff et al., 2002; Yáñez and Rivera, 2019). Due to the
ter edge model was used as an independent visual guide to possibility of assigning young structural features to older tec-
assist with the manual lineament interpretation, with non- tonic cycles where young fault movement offset older rocks,
continuous, strike-parallel Canny edges often extrapolated to additional confidence weighting was applied to interpreted
produce coherent lineaments. A drawback of the Canny filter fault systems that underwent demonstrable offset during
on N-S– and E-W–oriented pixelated data is that it assigns the tectonic cycle in question, as well as those that displayed
orientations that broadly align to W, NW, N, NE, and E trend- alignments of magmatic activity of the age being examined.
ing. Lineaments tended to range in length from 50 to 200 km. Lineaments from each tectonic cycle were compared to one
Regional-scale government and commercial airborne mag- another to investigate reactivation during different tectonic
netic data sets provided by First Quantum Minerals were used cycles. Recurrence of a lineament in multiple tectonic cycles
to conduct a magneto-structural interpretation of the cen- was recorded as a proxy for the confidence of interpretation.
tral Andes (Fig. 2B). Lineament interpretations of airborne Tectono-lithologic lineaments tended to be 50 to 300 km long.
magnetic data were carried out using gridded data on four
processing products: reduction to the pole (RTP) transform, Lineament map
analytic signal transform, the first vertical derivative filter of The spatial correlations of GIS-interpreted tectono-lithologic,
RTP data to enhance high frequency features, and an upward magnetic, and gravimetric lineaments were assessed manu-
continuation filter to highlight low-frequency anomalies. In- ally. Lineaments defined by two or more data sets advanced
terpretations were conducted at scales between 1:1M and to a “lineament map” to be validated against field-based
1:3M, dependent on flight line spacing of each survey, using evidence. Fieldwork was conducted to collect structural in-
the methods outlined in Isles and Rankin (2013). Interpreta- formation on selected lineaments and to characterize their
tions were conducted by hand on A0 maps, utilizing trans- surficial expressions. Published mapping was incorporated
parent overlays, which were then georeferenced and digitized to provide a field-based, independent component to our in-
into a GIS workspace. terpretations. Interpreted lineaments with no surficial expres-
sions were eliminated, and those with no field data available
Suture zones were preserved but assigned a lower confidence. The iterative
Suture zones of basement terranes in western South America process of desktop interpretation followed by field validation
accreted since the Neoproterozoic were taken from the lit- produced a more refined product at each iteration (Fig. 3).
erature and refined using regional gravimetry data (Fig. 2A). Fault segments, of which structural corridors are composed,
Some adjustment of the inferred positions of the suture zones represent two or more spatially correlated input data sets.
was necessary due to spatial inconsistencies in both the loca-
tion and orientation of sutures between different studies. The Fieldwork
sources examined include those based largely on geologic data Field mapping conducted at multiple scales was undertaken
(Mpodozis and Ramos, 1989; Charrier et al., 2007; Ramos, in selected areas of the central Andes to evaluate whether sur-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
6 FARRAR ET AL.

PERU PERU

Fig. 11
BOLIVIA BOLIVIA

AR
EQ
UI
PA
-A
NT
OF
Isostatic residual AL Reduced to Pole Fig. 6
gravity LA
(nT)
(mGal)

426 -295.5
130 -150
Magnetic
Terrane suture lineaments
zone
Aeromagnetic
Canny edge filter CHILE survey outlines CHILE
PAM
PIA
CHILENIA

CUYA
NIA

ARGENTINA ARGENTINA
0 250 0 250

kilometers kilometers

Fig. 2. A) Histogram-equalized isostatic residual gravity anomaly map with Canny edge filter and interpreted Paleozoic
suture zones of terranes (see “Methods” section in text; Mpodozis and Ramos, 1989; Carlier et al., 2005; Charrier et al., 2007;
Mamani et al., 2008; Ramos, 2010; Peri et al., 2015; Franceschinis et al., 2020). B) Histogram-equalized, reduced-to-pole
magnetic map from the World Digital Magnetic Anomaly Map (WDMAM; Lesur et al., 2016). Red lines define areas where
the regional aeromagnetic data sets (not shown) were utilized to carry out the magneto-structural interpretation (black lines).
Outlines showing locations of Figures 6 and 11 are displayed.

face features associated with interpreted lineaments can be AF, supported by First Quantum Minerals exploration teams,
recognized. Field mapping in northwest Argentina was un- between November 2018 and March 2020. In total, 647 fault
dertaken by DW, SG, JB, SB, AF, and First Quantum Miner- measurements were recorded in southern Peru, Chile, and
als company geologists during campaigns in 2014, 2015, and northwest Argentina. Strike and dip were recorded for each
2018 (Figs. 4B-D; 5). Field mapping in southern Peru was un- structure as well as sense of movement criteria where pre-
dertaken by AF, SB, DW, SG, First Quantum Minerals com- served. Regional traverses involved transects, up to 200 km
pany geologists, and consultants during mapping campaigns long, across preselected lineaments at observation stations
in 2014 and 2016. Regional field traverses were conducted by spaced approximately 5 km along their strike, with a particu-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 7

Data compilation and processing


Northwest Argentina case study
Regional data sets
Geology and porphyry deposits: The area incorporating So-
compa volcano and the Salar de Arizaro in northwest Argen-
Gravity Tectono- tina is dominated by igneous rocks that span several tectonic
Magnetic
interpretation lithologic cycles (Fig. 4A). The oldest rocks, considered basement in the
interpretation
interpretation area, are Ordovician-Silurian felsic plutonic rocks and NNE-
Multidisciplinary correlation striking aplite dikes composing the Taca Taca Formation
(Zappettini et al., 2001). The Taca Taca Formation exhibits
NW- and NNE-striking shear zones that are interpreted to
be associated with its intrusion and cooling history (Orts et
No Evidence in 2 or
more data sets?
al., 2011). The Taca Taca Formation is intruded by a Permian
NE-trending, coarse-grained granite batholith assigned to the
Llullaillaco plutonic complex, and contemporaneous steeply
Yes
dipping N- and NE-trending felsic dikes (Zappettini et al.,
Lineament Map 2001; Fig. 4B). The Taca Taca Alto Cu-Mo porphyry pros-
pect (Fig. 4A, B) occurs within the Permian granite batholith
and consists of structurally controlled advanced argillic altera-
Field tion and residual quartz ledges interpreted as the roots of a
validation Fieldwork
Reject lithocap (Benavides, 2017). Taca Taca Alto has an unreliable
lineament age of 323 ± 5 Ma by K-Ar magmatic biotite that has likely
Government and been affected by Ar loss (Sillitoe, 1977). The genesis of Taca
published mapping Taca Alto has been linked separately to the Permian Llullail-
Yes
laco plutonic complex and the Eocene-Oligocene Santa Ines
Complex (Zappettini et al., 2001; Rubinstein et al., 2021). We
Field data favor a Permian age for Taca Taca Alto based on observed
No Field evidence
available? for lineament? crosscutting field relationships within the Permian batholith
and abundant Permian dike arrays in the immediate vicinity
low N (Fig. 4B).
con o Yes
fide
nce
Six kilometers southeast of Taca Taca Alto, steeply dipping,
NNE-striking Oligocene porphyritic rhyodacite dikes intrude
Regional-scale Not adequately the Taca Taca Formation, which, based on drilling data, co-
features
captured? - check and refine alesce at shallow depths at the giant Oligocene Taca Taca
Bajo porphyry Cu-Au-Mo deposit (1,759 Mt @ 0.44% Cu,
Yes 0.012% Mo, 0.13 g/t Au: Gray et al., 2020; Fig. 4C). These
dikes were emplaced between 30.5 ± 0.3 to 29.0 ± 0.3 Ma (U-
Structural corridor Pb SHRIMP; Ince et al., 2022), with mineralization occurring
synchronously (29.1 ± 0.1 Ma Re-Os molybdenite; Ince et al.,
Fig. 3. Flow diagram outlining the iterative process to produce the structural
corridor map.
2022). Taca Taca Bajo occupies an anomalous “rear-arc” po-
sition relative to the coeval N-trending middle Eocene-early
Oligocene metallogenic belt 140 km to the west (Sillitoe and
lar focus on basement outcrops. A global positioning system Perelló, 2005; Fig. 1).
(GPS)-equipped tablet with the geologic and geophysical lay- Oligocene-lower Miocene lava, tuff, conglomerate, and
ers in a GIS workspace accompanied all fieldwork, which fa- felsic domes of the Quebrada de Agua Complex crop out in
cilitated the correlation of outcrop-scale observations and re- the central and western parts of the study area (Fig. 4A; Zap-
gional-scale features. Stereonet plots of fault data were made pettini et al., 2001). East of Socompa volcano, middle Mio-
using mplstereonet in python. Density contours are of poles cene Cu-Au–bearing porphyry dikes intruded the Quebrada
to faults; contours use the Schmidt 1% method, and units are de Agua Complex (Fig. 4D) and are associated with banded
in points per 1% area. quartz-magnetite “Maricunga-style” porphyry vein stock-
works (Muntean and Einaudi, 2000). These zones of porphyry
Results mineralization are ringed by magmatic-hydrothermal brec-
To demonstrate the characteristics of arc-oblique structural cias, which are crosscut by later NW-trending porphyry dikes.
corridors and their relevance to exploration, we included the Middle Miocene to Quaternary basalt to rhyolite lavas, py-
results of field mapping from two exploration sites as well as roclastic deposits, volcanic debris avalanche deposits, alluvial
data from regional field traverses. We compared our district- and colluvial gravels, and salt pan deposits are the youngest
scale geologic field mapping with published regional-scale units in the region (Zappettini et al., 2001).
government maps. We also compared field observations District-scale structural mapping—1:5k, 1:10k, 1:25k scales:
with geophysical analysis of the same regions to demonstrate A series of NNE- to NE-striking faults control the distribu-
how geophysical features and structural corridors manifest in tion of most geologic units adjacent to the Salar de Arizaro,
the field at different scales. including the orientation of Ordovician-Silurian shear zones

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
8 FARRAR ET AL.

Fig. 4. A) Regional-scale geology map of northwest Argentina and eastern Chile simplified from 1:250,000-scale (Zappettini
et al., 2001) and 1:100,000-scale (Ramírez et al., 1991) maps, respectively. B) District-scale geology map of the Taca Taca
Alto and Taca Taca Bajo areas from 1:25,000 mapping. Surface extents of magmatic-hydrothermal alteration for each deposit
are shown. C) Geologic map of the Taca Taca Bajo porphyry deposit from 1:5,000-scale mapping. The 0.6% Cu cutoff grade
shell, projected vertically to surface, is shown (Gray et el., 2020). D) Geologic mapping near Socompa Volcano at 1:10,000.
Black inset boxes represent the area of each subfigure and are labeled accordingly. Inset map of the central Andes shows the
outline of (A).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 9

Fig. 5. Lower-hemisphere equal-area stereonets of poles to fault planes (black points) from mapping in northwest Argentina
(Fig. 4B-D), overlain on density contours of the poles. Contours use the Schmidt 1% method; units are in points per 1%
area. A) Taca Taca Bajo mapping at 1:5k records a dominant population of steeply dipping WNW- to NW-striking faults with
secondary populations of steeply dipping N- and NE- striking faults. B) 1:25,000 mapping in the region around Taca Taca
Bajo demonstrates a dominant fault population of steeply dipping NE-striking faults and a secondary population of steeply
dipping NW-striking faults. C) Data for the Socompa volcano area demonstrates a dominant, steeply dipping NW-striking
fault population.

within the Taca Taca Formation (Orts et al., 2011), Permian by a nearby fault zone. A steep, NW-striking cleavage is de-
intrusions, mineralized Oligocene porphyry dikes, and min- veloped within the NW-striking fault zone, defined by prefer-
eralized veins (Figs. 4, 5). There are several locations where ential alignment of quartz and phyllosilicates within Paleozoic
NE-trending basins filled with volcanic-derived gravel termi- volcanic and intrusive rocks. Outside of the NW-striking fault
nate abruptly against Ordovician plutonic rocks, suggestive of zone, a pervasive cleavage was not detected.
a northeast fault control to these basins (Fig. 4B). The Quebrada de Agua Complex (Fig. 4D) is located 40 km
North-striking faults, prominent at 1:5,000 scale at Taca northwest of Taca Taca Alto, across a 25-km span of Miocene-
Taca Bajo, are subparallel to postmineralization fault brec- Holocene volcanic cover (Fig. 4A). The complex hosts early,
cias (Fig. 5A). A 2-km-long, NNW-striking structure, the TK2 Cu-Au–mineralized, NE-trending and later NW-striking por-
fault, occurs on the western margin of Taca Taca Bajo (Fig. phyry dikes, and faults strike predominantly northwest and
4C). It separates a deeply leached profile on the east from a north-northwest (Fig. 5C).
shallowly leached profile on the west, indicating recent move- Regional-scale structural mapping—1:100k and 1:250k:
ment. Abundant pyrite in the fault suggests it was a pathway Numerous 3- to 20-km-long NW-striking faults that controlled
for hydrothermal fluids. The timing and sense of movement the orientation of continental sedimentary basins in Argentina
along the TK2 fault have not been resolved, but it appears to and Chile are evident in government maps (Fig. 4A; Ramírez
have been active multiple times. et al., 1991; Zappettini et al., 2001). In many places, regional-
Northwest-striking faults occur regionally but are most pro- scale NW-striking faults are spatially coincident with NW-
nounced and concentrated within a 7-km-wide corridor that striking faults that we mapped at smaller scales (described
intersects the Taca Taca Bajo deposit (Fig. 4A-C). Ordovician- above) and link outcropping faults, mapped at smaller scales,
Silurian NW-striking basement shear zones occur within the beneath volcanic cover sequences (Fig. 4). Mapped regional
host batholith at Taca Taca Bajo (Orts et al., 2011). In the NW-striking faults intersect the Taca Taca Formation and the
northeast part of Taca Taca Bajo, several NW- to WNW-strik- Llullaillaco plutonic complex and continue northwest in the
ing faults truncate Oligocene mineralized veins and dikes and Miocene volcanic cover (Fig. 4A; Zappettini et al., 2001). Far-
separate domains of jarosite-rich, gypsum-poor granite from ther along strike to the northwest, within the Quebrada de
gypsum-rich, jarosite-absent granite. Where visible, kinemat- Agua Complex, a series of NW-striking faults continues into
ic indicators on NW-striking faults record sinistral movement, Chile (Fig. 4A) and is spatially coincident with our mapping of
defined by slickensides and offset of some Permian dike ar- NW-striking faults in this area (Fig. 4D).
rays (Fig. 4C). The magmatic-hydrothermal alteration cell The multiscale, pervasive, and long-lived NW-striking faults
associated with Taca Taca Alto is bound between two NW- described above require a NW-striking basement fault system
trending faults, the northernmost of which projects beneath (e.g., Orts et al., 2011) that we have named the Socompa struc-
2 km of Miocene-Holocene volcanic gravel cover and inter- tural corridor, which, within our study area, extends 80 km
sects the southern extent of Taca Taca Bajo (Fig. 4B). Within from Taca Taca Bajo, via Taca Taca Alto, through the mineral-
the Permian granite, at the projected intersection with the ized porphyry dikes (Fig. 4D) and into Chile (Fig. 4A). The
NW-trending fault zone, there is a change in the strike and an Socompa structural corridor passes mostly through outcrop of
increase in the apparent width of the Permian granite body. Miocene-Holocene volcanic and gravel cover, which obscures
Permian felsic dikes within the granite located northeast and its surface expression. However, along its strike where older
southwest of the NW-trending fault zone are steeply dipping, rocks crop out within the corridor, NW-striking faults, shears,
but within the fault zone, the dikes are moderately northwest and dikes are prevalent, as well as multiple epochs of mineral-
dipping, with curved outcrop patterns indicative of rotation ized porphyry formation.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
10 FARRAR ET AL.

A sinuous, N-striking structural trend is present along the Continental-scale geology—1:3M: The Domeyko basin
Argentina-Chile border, composed of high-angle faults 2 to comprises Late Triassic-mid-Cretaceous volcanosedimen-
20 km long that bound the Oligocene-Miocene Quebrada de tary rocks (Fig. 6D), which are distributed along a sinuous,
Agua Complex (Fig. 4A). We interpret these faults to bound 800-km-long, N-trending mesh of Mesozoic extensional faults
graben and half-graben basins that were depocenters for the in northern Chile (Amilibia et al., 2008; Espinoza et al., 2021;
Quebrada de Agua Complex. Subsequent inversion of the Fig. 6C-E). Subbasin depocenters (Prinz et al., 1994) and
graben uplifted these rocks to the surface. In the Socompa basement highs were controlled by a series of NW-striking
area, the graben changes strike from north-northwest to faults (Richards et al., 2001; Espinoza et al., 2019) and NE-
north-northeast, the Quebrada de Agua Complex unit thins striking faults (e.g., Antofagasta-Calama lineament; Thomas,
(Fig. 4A), and Paleozoic basement rocks are exposed on the 1970; Palacios et al., 2007; Amilibia et al., 2008) that coincide
Chilean side (Ramírez et al., 1991), implying this kink may with many of the magnetic and gravimetric lineaments de-
be a transfer zone. This location coincides with the intersec- scribed above. The NE-striking Salinas Grandes lineament is
tion of the N-trending graben with the NW-striking Socompa the regional manifestation of the Taca Taca lineament, which
structural corridor and the middle-Miocene Cu-Au bearing is interpreted to have controlled the emplacement of Meso-
porphyry dikes described above (Fig. 4A). zoic igneous rocks along its strike (Salfity, 1985; Zappettini et
The NE-striking Taca Taca Lineament has been proposed al., 2001; Chernicoff et al., 2002; Fig. 7).
as the regional-scale manifestation of NE-striking faults and Volcanism in the Neogene-Quaternary arc extends up
the NE-striking distribution of Siluro-Ordovician, Permian, to 300 km into the back-arc in Argentina along four NW-
Oligocene, and Miocene magmatic rocks that extend along trending belts that are separated from each other by ~150
the western shoreline of the Salar de Arizaro (Zappettini km (Figs. 6C; 7). These back-arc volcanic belts have been
et al., 2001). Our field mapping supports the presence of a recognized as conspicuous features in this part of Argentina
pervasive northeast fault trend that spatially corresponds to (Salfity, 1985; Matteini et al., 2002; Richards and Villeneuve,
the Taca Taca lineament, manifest in outcrop as NE-striking 2002; Gioncada et al., 2010; Lanza et al., 2013) and are
faults, shear zones, dikes, and mineralized veins as described spatially coincident with the NW-trending magnetic (Salf-
above, which occur throughout different geologic epochs ity, 1985; Richards et al., 2001; Chernicoff et al., 2002; Fig.
(Figs. 4B-C, 5A-B). 6B) and gravity lineaments described above (Fig. 6A). Lavas
Continental-scale gravimetric data—1:3M: Lineament in­ erupted along these belts include some of the most primi-
terpretation of gravimetric data utilizing Canny edge detec- tive magma compositions of the central Andes magmatic arc
tion demonstrates that this region of northwest Argentina (Gioncada et al., 2010), and the spatial distribution of fore-
and contiguous Chile is dominated by NW-, N-, and NE- land volcanoes is controlled by the intersection of NW- and
striking lineaments, with a minor E-trending population NE-striking continent-scale fault systems (Richards and Vil-
(Fig. 6A). NW- and NE-striking gravimetric lineaments leneuve, 2002; Gioncada et al., 2010). To the east of the NW-
spatially coincide with the Socompa fault system and Salinas trending volcanic belts, a NW-trending structural corridor
Grandes lineament (Fig. 6A). The significance and utility of separates the southern portion of the Cretaceous-Paleogene
gravimetric data to the interpretation of deep crustal struc- Salta basin from the Paleozoic metamorphic basement rocks
tures in this region is well documented (Rivera and Cerda, of the northern Pampeanas Ranges (Jordan et al., 1983;
2012; Yáñez and Rivera, 2019). Parallel gravity-derived Jacques, 2003; Fig. 6C).
lineaments exhibit a degree of regularity to their spacing,
which is roughly 70 to 90 km for NW-striking lineaments, Southern Peru case study
40 to 80 km for N-striking lineaments, and 70 to 100 km for Geology and porphyry occurrences: The 50-km-long, NNW-
NE-striking lineaments. trending Rio Tambo valley in southern Peru is located within
Continental-scale magnetic data—1:3M: Magnetic linea- the Central volcanic zone of the central Andes, 40 km north
ments display west-northwest, northwest, north, northeast, of a prolific cluster of Paleocene-early Eocene giant porphyry
and east trends, including trends broadly coincident with the copper deposits (Figs. 8, 9; Sillitoe and Perelló, 2005; Sim-
strike of the Socompa fault system and Salinas Grandes lin- mons et al., 2013). The Tambo River has incised a 3,000-m-
eament (Fig. 6B). Lineament interpretation of aeromagnetic deep, steep-sided valley that exposes a cross section through
data in the region was reported by Behn et al. (2001) and the local geology, which has been described at a district scale
Chernicoff et al. (2002). As for the gravity lineaments, there by García (1978) and Lavallée et al. (2009). Field mapping
appears to be a periodicity to the spacing of parallel linea- within the study area (Fig. 9A) was conducted at 1:12,500 and
ments of between 50 and 100 km. Several prominent trends 1:5,000 scales.
include 300-km-long, 40-km-wide, regularly spaced NW- The stratigraphically lowest unit comprises marls and
trending magnetic highs that extend from the present-day vol- limestones belonging to the Jurassic Arcurquiña Forma-
canic arc into Argentina and are coincident with NW-trending tion (García, 1978), which are unconformably overlain by
gravity features (Fig. 6B). These features have been described plagioclase-phyric volcanic and pyroclastic rocks and volca-
by other studies in this region (Richards et al., 2001; Cher- nogenic sedimentary deposits that belong to the Late Cre-
nicoff et al., 2002; Yáñez and Rivera, 2019). In Chile, regu- taceous Toquepala Group (García, 1978). Within the study
larly spaced, E-oriented magnetic lineaments were noted by area, these sequences were intruded by three discrete Mio-
Behn et al. (2001), who termed them “transverse magnetic cene felsic plutonic complexes (Fig. 9A). The two southern
anomalies” and interpreted the magnetic highs as subsurface plutonic complexes each crop out over an area of 20 km² and
“transbatholiths.” are composed of hornblende-biotite monzonite intruded by

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 11

Fig. 6. Maps of northern Chile, northwest Argentina,


and southern Bolivia, showing data sets (A-C). Red box
in (A-C), (E) is the area of Fig. 4A. Location of (A-C)
shown in inset map of the central Andes within (B). A)
Histogram-equalized isostatic residual gravity anomaly,
overlain with Canny edge detection contours and lin-
eaments interpreted from these gravity data. B) Histo-
gram-equalized reduced-to-pole (RTP) magnetic data
from northern Chile and Argentina overlying the first
vertical derivative of the RTP magnetic; magnetic of
World Digital Magnetic Anomaly Map (WDMAM) with
an RTP filter is in background; black lines are lineaments
interpreted from magnetic data, which are ranked as
major and minor. C) Geology, highlighting Paleozoic
metamorphic basement, the Mesozoic Domeyko basin,
the Cretaceous-Paleogene Salta basin, and the Neogene
volcanic arc. Brown inset denotes the area of subfigure
(D). Mapped faults from Jordan et al. (1983), Ser-
nageomin (2002), Amilibia et al. (2008), Espinoza et al.
(2019), and Segemar (1997). D) The Domeyko rift basin
is composed of Triassic to mid-Cretaceous sedimentary
rocks that unconformably overlie Carboniferous-Perm-
ian igneous basement rocks. Faults from (C) highlight
the importance of inherited N-, NW-, and NE-striking
structural architecture for controlling the distribution of
geologic units. E) Triassic-Jurassic rift faults modified
from Ramos (1994), Charrier et al. (2007), McGroder
et al. (2015), and Espinoza et al. (2021) and Cretaceous
rift faults from McGroder et al. (2015).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
12 FARRAR ET AL.

Fig. 7. Structural corridors, giant porphyry copper deposits, and Neogene-Quaternary volcanic arc of central Chile. Segments
of structural corridors that are well constrained with multiple layers of evidence for their trace can be clearly distinguished
from fault systems that are less defined and therefore more speculative. Fault strands reflect observations from surface map-
ping and magnetic and gravimetry data. All named fault systems have previously been identified and follow the nomenclature
of Richards et al. (2001), except for the Socompa structural corridor, which we introduce here. Modified from Richards et al.
(2001) and Gioncada et al. (2010).

diorite porphyry dikes that are tens of meters long. Centered Regional-scale structural mapping—1:100k: We mapped a
around the diorite dike outcrops is km-scale porphyry-style 25-km-long, NNW-striking white mica-pyrite ± pyrophyllite-
magmatic-hydrothermal alteration (Fig. 9B). These previous- alunite–altered fault zone that extends along the Tambo River
ly unreported porphyry systems are spatially associated with Valley floor (Fig. 9A). We interpret that this fault zone forms
some small-scale historic Cu workings. Capping these units part of a NNW-trending Neogene graben, which has previ-
at the highest levels of the valley are flat-lying volcanic rocks ously been interpreted to control the spatial distribution of
of the Pliocene-Quaternary Barroso Group, which in turn are volcanoes in the region (Fig. 9A; Lavallée et al., 2009). The
overlain by Holocene-age ash and pumice from the Ubinas, marked change in orientation of the Tambo River from north-
Huaynaputina, and Ticsani volcanoes. northeast to northwest in the center of the case study area
District-scale structural mapping—1:5k, 1:12.5k: Field (Fig. 9C) corresponds to a NW-striking sinistral fault, inferred
mapping at 1:12.5k (Fig. 9A) and 1:5k (Fig. 9B) scales docu- to have dislocated the graben by 5 km during the Pleistocene
mented moderately to steeply dipping faults that strike pre- (Lavallée et al., 2009). The intersection of this NW-striking
dominantly north-northwest and northeast to east-northeast sinistral fault with the NNW-striking white mica-pyrite ±
(Fig. 10), which in some instances controlled the orientation pyrophyllite-alunite–altered fault zone corresponds to the
of porphyry dikes and truncated hydrothermal alteration location of the porphyry-magmatic-hydrothermal alteration
within the Toquepala Group volcanic rocks (Fig. 9A, B). Re- system (Fig. 9A, B). In the eastern part of the study area,
cent uplift has produced fault scarps with numerous associ- the Puno Group continental sedimentary rocks and Barosso
ated E-striking faults. A 25-km-long, 3-km-wide NNW-strik- Group volcanic rocks are separated by a sharp, NNW-trend-
ing fault system containing hydrothermal breccias with white ing contact parallel to the strike of the graben (Fig. 9A).
mica-pyrite ± pyrophyllite-alunite alteration extends parallel Regional-scale magnetic data—1:100k: Here, structural in-
to the strike of the valley from Ubinas volcano and intersects terpretation of magnetic data (Fig. 9D) highlights prominent,
the porphyry system described above (Fig. 9A). NW- and NE-trending magnetic lineaments throughout the

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 13

magnetic grain to much of southern Peru (Fig. 11B), which


is consistent with NW-striking sinistral faults in field mapping
(Fig. 9A, B). Also apparent are ENE-trending magnetic linea-
ments >100 km long that extend from the coastline into the
interior that are consistent with magnetic lineaments we in-
terpreted at a smaller scale (Fig. 9C).
Continental-scale geology—1:3M: A 1,000-km-long, NW-
trending belt of Jurassic-Cretaceous intrusive rocks of the
Arequipa segment of the Coastal batholith are spatially corre-
lated with NW-striking magnetic and gravimetric lineaments,
the Incapuquio fault zone and Paleocene-Eocene giant por-
phyry copper deposits (Fig. 11; Cobbing, 1982; Simmons et
al., 2013). A series of NW-trending Mesozoic sedimentary
basins were inverted along NW-striking reverse faults dur-
ing Late Cretaceous compressional tectonism (Vicente, 1989;
Perez et al., 2016), and ENE- and NNW-trending faults con-
trolled the deposition of Cenozoic continental sedimentary
basins (Fig. 11C; Ingemmet, 1999; Carlotto, 2013).
Regional traverses
Regional traverses (Fig. 12A) were undertaken as part of the
iterative process of interpretation and validation of the base-
ment fault systems map and to characterize the surficial ex-
pression of regional- and continental-scale geophysical fea-
tures. Though there is a compelling trench-parallel structural
orientation to many faults, including prominent fault systems
such as the Atacama and Domeyko faults (Fig. 7), there is a
high proportion of faults that strike oblique to the trench (Fig.
12B).
Continental-scale gravimetric and magnetic lineaments
(Fig. 6A, B) were studied in the field by conducting traverses
along their strike for up to several hundred km (Fig. 12A). We
found that coherent and coincident magnetic and gravimetric
lineaments repeatedly manifest on the surface as a preferred
Fig. 8. Geology map of southern Peru simplified from the map of Ingemmet
orientation to surface faults, whose strikes broadly align to
(1999). The giant porphyry cluster of Cuajone, Quellaveco, and Toquepala as those of the coincident geophysical lineaments. These struc-
well as faults and Quaternary volcanoes are shown. Blue rectangle is the area tural corridors are typically composed of small-scale synthetic
of the Rio Tambo valley in Fig. 9A. Inset shows location within the central faults with little offset. Occasionally, fault systems will include
Andes. important regional-scale faults with tens of km of offset, such
as the West Fissure (Tomlinson and Blanco, 1997; Fig. 13A,
valley, some of which are spatially coincident with those rec- B), but as a proportion of the total fault population, these are
ognized in field mapping. Large magnetic bodies visible in the the exception, not the rule.
data possibly represent the subsurface extents of the felsic- Of the faults that yielded reliable kinematic indicators, a
intermediate magnetite-series intrusions mapped at surface quarter recorded more than one sense of movement. Fault
reactivation is well documented throughout the central Andes
and their contact metamorphism aureoles. In outcrop, intru-
(e.g., Allmendinger et al., 1983; Scheuber and Andriessen,
sive contacts have similar trends to NW- and ENE-striking
1990; Palacios et al., 2007; Piquer et al., 2017). Figure 13A
magnetic lineaments (Fig. 9D).
shows a segment of the Domeyko fault zone south of Calama,
Continental-scale gravimetric data—1:3M: The major mapped as a series of thrust faults of Paleozoic volcanic rocks,
gravimetric feature of southern Peru consists of NW-trending whereas north of Calama, a segment of the same fault system
gravity highs that strike parallel to the Peruvian coastline (Fig. is interpreted to have undergone multiple transcurrent reac-
11A). These are intersected by NE-trending lineaments that tivation events (Tomlinson and Blanco, 1997; Fig. 13B). Pre-
are apparent in the Canny edge contours (Fig. 11A). A N- cambrian metamorphic basement is exposed in locally faulted
striking lineament is present in the southern section of our contact with Paleozoic and Mesozoic sequences adjacent to
field mapping area, and immediately to the northwest of our major faults such as the multiply reactivated Incapuquio fault
mapping area, a north-northwest striking lineament is present. in southern Peru (Fig. 13E; Shackleton et al., 1979) and the
These lineaments broadly spatially coincide with the NNW- Atacama and Domeyko faults in Chile (Scheuber and An-
striking graben in our field mapping area (Figs. 9A, 11A). driessen, 1990; Charrier et al., 2007).
Continental-scale magnetic data—1:3M: Airborne mag- Within the volcanic arc, trends of volcanoes and volca-
netics demonstrate a dominant trench-parallel, NW-striking nic vents align along both arc-parallel and NW-striking arc-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
14 FARRAR ET AL.

Fig. 9. A) Geology map of the Tambo River valley from 1:12.5k- and 1:5k-scale mapping, including faults from Lavallée et
al. (2007). The red box shows the study area, and the black box shows the location of (B). Geology outside of the study area
from García (1978). Lithology legend applies to both (A) and (B). Regional white mica-pyrite ± pyrophyllite-alunite alteration
is interpreted to be volcanic. B) Geology from 1:5k mapping, including the extent of quartz-sericite-pyrite (phyllic) hydro-
thermal alteration associated with porphyry systems, which merges with regional volcanic alteration. C) 5m digital eleva-
tion model (red box) overlain on histogram-equalized stretch of bands 4-3-2 LANDSAT 8 imagery from October 2020. D)
Heli-magnetic histogram-equalized reduced-to-pole (RTP) with a semitransparent first vertical derivative filter of the RTP.
Magneto-structural interpretations of this survey are the black lines.

oblique structural corridors (Fig. 7; Richards and Villeneuve, rocks and basement sequences, regional-scale fault systems
2002). In Chile, we observed NW-striking faults along the are comparatively narrower, more defined, and with a signifi-
projected strike of NW-trending volcanic “breakouts” in Ar- cantly higher fault density (Fig. 14A).
gentina (e.g., Figs. 4, 5; Richards et al., 2001; Lanza et al.,
2013; El Toro, Archibarca, Culampaja, Socompa structural Discussion
corridors). Where structural corridors coincide with Quater-
nary volcanic cover sequences, their surficial manifestation is Levels of confidence in the interpretation of
diffuse, consisting of widely spaced faults that strike parallel to structural corridors
the structural corridor. These zones can be up to 30 km wide, Although the best available geologic and geophysical data sets
which we interpret to represent the upward propagation of have been used to produce our structural corridor interpreta-
the basement fault zone through the overlying, younger cover tion of the central Andes (Fig. 14A), it is not a “fact map.”
sequences (Coleman et al., 2019; Fig. 13C, D). Within older Owing to the scale, multiple reactivation episodes and sur-

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 15

n = 62 n = 53

Fig. 10. Lower-hemisphere equal-area stereonets of poles to fault planes (black points) from mapping in the Rio Tambo val-
ley (Fig. 9A-B), overlain on density plots of the poles. A) At 1:5k scale, steeply dipping ENE-striking faults are dominant; a
secondary population of moderately dipping NNW-striking faults is also present. B) At 1:12.5k scale, steeply dipping ENE-
striking faults dominate.

face cover, the level of inherent uncertainty to the trace of Sinistral offset along some NW-striking fault systems is con-
a continental-scale fault system will always be higher than sistent with our field observations in northwest Argentina and
for faults mapped at outcrop scale. This makes continental- southern Peru (Figs. 4B, C; 9A, B) and has also been widely
scale fault maps speculative and open to reinterpretation as reported by other workers (e.g., Lavallée et al., 2009; Lanza et
new data sets become available. Nevertheless, little scientific al., 2013; Perez et al., 2016; Espinoza et al., 2021).
insight or exploration benefit is to be gained from the sole Dextral reactivation of NE-trending fault systems, such as
reliance on demonstrably true factors, which are ultimately the Antofagasta-Calama lineament, has been suggested dur-
limited to surface geologic evidence and drill data. Abundant ing orocline formation, consistent with the orientations of the
surficial geologic evidence is present from field mapping for deforming mesh presented here (Palacios et al., 2007; Amilib-
the arc-parallel Atacama and Domeyko fault systems (Fig. 7; ia et al., 2008). Reactivation of N- to NNE-trending faults is
Scheuber and Andriessen, 1990; Reutter et al., 1996). In con- complex, with both sinistral and dextral offset depending on
trast, surface evidence for arc-oblique structures is less abun- the distance from the orocline core (Fig. 14B). The strike-slip
dant due to their diffuse and cryptic nature. Many arc-oblique reactivation of these systems is interpreted to be influenced
structures are obscured beneath Neogene and Quaternary by the net clockwise rotation invoked by orocline formation
cover sequences, leading us to rely comparatively more on on N-striking structures south of 20°S, balanced against fluc-
geophysical evidence in such regions and reducing the level tuating plate convergence orientations and velocities during
of confidence in the interpretation of the structural corridor much of the Cenozoic that manifest as multiple reactivation
(Fig. 14A). events along the N-trending fault systems (Tomlinson and
Blanco, 1997; Eichelberger and McQuarrie, 2015).
Spatial distribution of structural corridors At the continental scale (Fig. 14B), not all fault systems
The proposed structural corridors (Fig. 14A) define a mesh of presented here display obvious offset. This might be because
orthorhombic symmetry (Sagy et al., 2003). At the continental intersecting fault zones that are reactivated synchronously but
scale, parallel structural corridors display a spatial periodicity with opposing kinematics display complex interference be-
of approximately 60 to 90 km. North of approximately 20°S, havior where they intersect. The resultant observed offset is
structural corridors appear to have been rotated in an anti- the result of the sum of the mutual offsets produced by each
clockwise manner relative to the structural corridors south incremental slip and can be distributed across the width of
of this latitude, e.g., N- to NNE-trending faults in Chile are the fault zones.
north-northwest striking in Peru (Fig. 14C). This rotation is
interpreted to reflect oroclinal bending, thought to have com- Genesis and development of arc-oblique structural corridors
menced during the Eocene, with upper crustal responses con- Basement fault systems in the central Andes reflect multiply
tinuing to the present day (e.g., Allmendinger et al., 2005; Ar- reactivated preexisting fabrics associated with the Paleozoic
riagada et al., 2008). During orocline bending, the preexisting accretional history of the margin and the Mesozoic extension-
structural mesh would have absorbed tectonic rotational strain al events (e.g., Fig. 13A, B, E, F; Ramos, 1994; Bahlburg and
via strike slip reactivation of fault systems, largely preserving Hervé, 1997; Perez et al., 2016; Espinoza et al., 2021; Wiemer
the original mesh form (e.g., Amilibia et al., 2008; Arriagada et al., 2022). Suture zones provide steeply dipping zones of
et al., 2008; Eichelberger and McQuarrie, 2015). Much of the translithospheric weakness that are available to be reactivated
oroclinal deformation was accommodated in the fault mesh as strike-slip faults or major extensional faults, around which
via sinistral reactivation of NW-striking fault systems and both subsequent fault zones propagate when tectonic strain is ap-
sinistral and dextral reactivation of N- to NNE-striking fault plied (Holdsworth et al., 1997; Begg et al., 2009; Wiemer et
systems in the southern limb of the orocline (Figs. 14B, 15C). al., 2022).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
16 FARRAR ET AL.

Fig. 11. Regional-scale maps of southern Peru and northernmost Chile showing the same area with different data sets. The
pink box in each image is the area of Figure 9A. A) Histogram-equalized isostatic residual gravity anomaly with Canny filter
and lineaments interpreted from gravity data. B) Histogram-equalized reduced-to-pole (RTP) aeromagnetic data overlain on
the first vertical derivative of RTP magnetic data from the World Digital Magnetic Anomaly Map (WDMAM) survey. Black
lines are lineaments interpreted from these magnetic data. C) Geology map showing rift-related Permian to Jurassic rocks
(Mitu Group), Jurassic to Cretaceous intrusions (Coastal Batholith), Miocene to Quaternary volcanic rocks, and structural
corridors (this study). Annotations highlight structural orientations to sedimentary basin distribution; the Mitu Group is con-
trolled by NW-striking normal faults that were inverted during Andean deformation (Perez et al., 2016), whereas Cenozoic
sedimentary basins can exhibit north-northwest, northwest, and east-northeast controls to basin orientation (Carlotto, 2013).
FZ = fault zone.

Examples of tectonic events that may have reactivated Zamora and Gil, 2018; Espinoza et al., 2021). This extension
and enhanced inherited structural architecture include Late phase was accompanied by rift basins inboard of the margin
Permian to early Triassic extension, which produced a series that might have connected to extensional basins that devel-
of rift basins proximal to the southwest Pangea continental oped along the margin (Espinoza et al., 2021). Our analysis
margin (Figs. 6E, 15A; Jacques, 2003). This period is associat- supports the role that regional-scale, penetrative fault systems
ed with sinistral transform plate convergence along southwest had in controlling the distribution of Paleozoic to Mesozoic
Pangea (Young et al., 2019; Fig. 15A), which favors the reacti- sedimentary basins (Figs. 6, 11).
vation of deep segments of preexisting arc-oblique basement Preexisting structural corridors were further enhanced
fault systems (Umhoefer, 2011; Cao and Neubauer, 2016; when, in the Early Cretaceous, the opening of the south At-
Lutz et al., 2022). During this extensional epoch, the Mara- lantic Ocean produced a series of extensional basins in Argen-
ñón, Mitu, Domeyko, Cuyo, and Neuquén rift basins initi- tina, Bolivia, and Peru (Figs. 6E, 15B; McGroder et al., 2015)
ated, all of which were truncated by inherited, trench-oblique and significantly increased compressional stress along the
basement fault systems that controlled the orientation of ba- western margin of the central Andes that reactivated preexist-
sin margins and subbasin depocenters (Figs. 6E, 15A; Ramos, ing structural architecture (Sempere et al., 2002; Charrier et
1994, Jacques, 2003; Charrier et al., 2007; Perez et al., 2016; al., 2007; Somoza and Zaffarana, 2008). We infer segments of

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 17

270°

n = 340

270°

n = 647

Fig. 12. A) Simplified lithology map of the central Andes with fault measurements taken as part of the regional traverses. Field
transects follow numerous interpreted lineaments for hundreds of km. Southern Peru and northwest Argentina case study
areas are within the purple rectangles. B) Lower-hemisphere equal-area stereonet of poles to fault planes (black points) mea-
sured as part of regional traverses are overlain on a density plot of the poles. Steeply dipping NE-, NW-, and NNE-striking
faults are the dominant populations. C) Field traverses combined with southern Peru and northwest Argentina case studies
stereonet plot as per (B) define dominant fault populations of steeply dipping NE-, NW-, and N-striking faults.

the structural corridors we present were further reactivated bal et al., 2016; Lutz et al., 2022) and have been imaged in
and enhanced during Late Cretaceous-Cenozoic contraction- geophysical investigations proximal to ore deposits (Lezaeta,
al events and orocline formation (Fig. 15C; Sempere et al., 2001; Koulakov et al., 2006; Haugaard et al., 2021). The me-
1990; Oncken et al., 2006; Arriagada et al., 2008; Carlotto, chanics of a vertical connection between upper crust and up-
2013; Eichelberger and McQuarrie, 2015; Puigdomenech per mantle, through the ductile lower crust, have been cap-
et al., 2020), which includes sinistral offset along reactivated tured by the Jelly Sandwich lithospheric model (e.g., Burov
NW-striking faults in our case studies (Figs. 4B, C; 9A, B). and Watts, 2006) and the lithospheric feedback model (Chat-
zaras et al., 2015). During high rates of strain, the upper crust
Lithospheric coupling and activation of translithospheric and lithospheric mantle brittly deform, are vertically linked,
pathways and together act as a coupled system with the ductile lower
Key to understanding the relevance of long-lived arc-oblique crust localizing stress and acting as a fluid conduit between
structural corridors for localizing giant porphyry copper de- the two mechanically stronger layers (Sawyer, 1985; Stuwe,
posit clusters is to appreciate their role in distributing strain 2007; Inbal et al., 2016; Fig. 15C). During high-horizontal-
and vertical fluid migration through the lithosphere. Focused compressional-stress events, the intersection of two basement
vertical deformation can extend as a coherent and coupled fault systems provides a vertical channel of highly anomalous
system from the upper mantle to the upper crust as steep permeability in the upper crust, which is further enhanced by
lithosphere shear zones (Ford et al., 2014; Li et al., 2021) highly pressurized fluids and magmas ascending from the up-
that can offset the lithosphere-asthenosphere boundary (In- per mantle along the magmatic arc (Fig. 15C; Richards, 2003;

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
18 FARRAR ET AL.

Fig. 13. Examples of the surface expression of structural corridors. A) Domeyko structural corridor, northern Chile, looking
north, showing thrusting of Carboniferous-Permian volcanic rocks; 508518 mE, 7426516 mN. B) Domeyko structural cor-
ridor, West Fissure fault; Cretaceous conglomerates in contact with Miocene conglomerate; inferred dextral and sinistral
movement sense (Tomlinson and Blanco, 1997); 512896mE, 7572072mN. C) Looking west at NE-trending Antofagasta-
Calama structural corridor (blue lines) and NW-trending Lipez fault system (red lines). Quaternary normal faults affect the
topography of pampa cover. A 2-km-long, northeast lateral jog of the Loa River in the distance coincides with the trace of
the Antofagasta-Calama lineament. D) Photo looking northwest along a 10-km-long Quaternary normal fault in the volca-
nic cover showing 5 m of vertical displacement; geologist outlined in green for scale on far right; 548689mE, 7583054mN.
E) Incapuquio structural corridor, southern Peru, showing Precambrian basement in contact with Cretaceous granite;
386458mE, 8052040mN. F) Secondary faults within the Incapuquio fault zone showing striations with both subvertical and
subhorizontal rake, indicating at least two movement events. Structural measurements are given as dip/dip direction; coordi-
nates are in WGS84 UTM zone 19S; colored lines highlight structural features.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
-12 ° LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 19

-72 ° -68 ° -64 ° -80 ° -70 ° -60 °


BR BR
A AZ
IL
IA
N
B AZ
IL
IA
N
CR CR
AT AT
O O
N N

Earliest recorded Mesozoic


to Cenozoic activation

-20°
Neogene
-16 °

Late Cretaceous to
Paleogene
Early to mid-Cretaceous
Triassic-Late Jurassic

Strike slip offset

Structural corridors
-20°

confidence level
Poorly defined

-30°
Semiconstrained
Well constrained

Proterozic 0 200
basement
kilometers
Basement
terrane suture
-80 ° -70 ° -60 °
Giant porphyry deposit
-24 °

contained Cu equiv (Mt) BR

>25
C AZ
IL
IA
N
CR
AT
15 to 25 O
N
10 to 15
5 to 10 Strike of structural
corridor
-20 °

3 to 5
N to NNE
-28°

NNW
NW
NW to NNW
NNE
NE
-32°

-30 °

0 200 0 200

kilometers
kilometers

Fig. 14. A) Structural corridors, colored by confidence level. They define a rhomb-shaped mesh that abuts the rigid cratonic
basement of Brazil and Argentina. Interpreted sutures of Paleozoic and Precambrian terranes are shown, as are giant por-
phyry copper deposits. Blue inset box is the location of Figure 9A. B) Structural corridors colored by the earliest Mesozoic
to Cenozoic extensional event that affects them. Strike-slip offsets, possibly associated with Cenozoic oroclinal formation,
are annotated based on the fault mesh pattern and age of fault systems. C) Structural corridors colored by strike allow for
enhanced discrimination of the effect of oroclinal bending.

Cao and Neubauer, 2016; Cox, 2020; Loucks, 2021). The fo- Preexisting permeability pathways, such as long-lived faults,
cused ascent of fluids through the thinned and ductile lower can be reactivated even where they are severely misoriented
crust may be facilitated by pore fluid overpressures that frac- relative to the stress field when high pore-fluid pressures are
ture the ductile rock (Cao and Neubauer, 2016; Inbal et al., present (Sibson, 1996; Holdsworth et al., 1997; Cox, 2020).
2016; Tosdal and Dilles, 2020). In the upper crust, high pore-fluid pressures are provoked

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
20 FARRAR ET AL.

during the fractionation and devolatilization of metal-fertile stress field (Piquer et al., 2021a). In addition to upper crustal
hydrous magmas at the cupola of a pluton (Sillitoe, 2010b). processes, strike-slip activation of shear zones in the lower
In the upper lithospheric mantle, strength is related to hydra- crust and upper lithospheric mantle can occur under high
tion, whereby increasing hydration weakens the upper mantle rates of strain (Cao and Neubauer, 2016). During periods of
(Hirth and Kohlsted, 1996). A feedback loop of enhanced transpressional reactivation, the intersections of “misaligned”
permeability, fault reactivation, and fluid focus is established and arc-parallel structural corridors become deep and focused
when strain-activated preexisting zones of weakness are the strain anomalies through which magmas can preferentially as-
preferred pathways for magma flux through the upper litho- cend (Fig. 15C). During their ascent through the lithosphere,
spheric mantle that continually enrich these zones as well buoyant magmas and fluids can undergo a complex series
as keeping them weak, and therefore permeable (Cao and of chemical enrichment processes (Tosdal and Dilles, 2020;
Neubauer, 2016). We propose that during high-strain events, Loucks, 2021). These processes are thought to be linked, in
steep upper mantle permeability pathways couple with steep part, to the regional stress field, with periods of waning mag-
upper crustal basement fault zones (Burov and Watts, 2006; matic activity, high rates of crustal thickening, and transition
Chatzaras et al., 2015) to produce temporally transient verti- in tectonic regime correlating with the timing of porphyry Cu
cal translithospheric conduits that facilitate the rapid ascent deposit formation (Tosdal and Richards, 2001; Mpodozis and
of pressurized fluids, magma, and energy from the melting Cornejo, 2012; Loucks, 2021).
downgoing slab to the crust (Fig. 15C; Loucks, 2021). We The spatial correspondence of structural corridor intersec-
envisage that as high-strain conditions wane, the translitho- tions with giant porphyry copper deposit clusters leads us to
spheric coupling of the vertical conduit also wanes, diminish- infer that anomalous structural permeability not only permit-
ing its effectiveness as a focused vertical conduit. It is unclear ted but was further enhanced by the focused ascent of mag-
if the vertical permeability that facilitated the ascent of pres- mas and Cu-rich fluids during high-strain events. Whereas
surized fluids and magma is preserved in the ductile lower our structural corridor map (Fig. 14A) does not invalidate the
crust and in the subcontinental lithospheric mantle follow- Rayleigh Taylor diapiric model for the periodic distribution
ing the termination of the high-horizonal-rate-of-strain event of porphyry deposit clusters (Yáñez and Maksaev, 1994; Tom-
(Cao and Neubauer, 2016). linson and Cornejo, 2012), we find that the regular spacing of
structural corridor intersections within the metallogenic belt
Association of arc-oblique structural corridors and giant adequately explains deposit clustering.
porphyry copper deposit clusters There are many structural corridor intersections within
The intersection of NW- and NE-striking structural corridors metallogenic belts that are not associated with a known giant
has been postulated to be a crucial element in the localiza- porphyry copper deposit cluster (Fig. 14A). We have an incom-
tion of “rear-arc” giant porphyry copper deposits in Argentina, plete knowledge of the true distribution of all giant porphyry
such as Bajo de la Alumbrera, Agua Rica (Fig. 7; Ricci and copper deposits in the central Andes, with postmineralization
Figueroa, 1971; Salfity, 1985; Alonso and Viramonte, 1987; cover, exploration maturity, and magnitude of postmineraliza-
Richards et al., 2001), and Taca Taca Bajo (Orts et al., 2011; tion orogenic uplift and erosion all playing a role in concealing
this study). In northwest Argentina, contemporaneous rear- their true distribution. Many structural corridor intersections
arc magmatism that accompanies these ore deposits is often without an associated giant porphyry copper deposit cluster
low volume and subtle. Understanding the locations of these may represent areas that lacked the optimal combination of
transverse structural corridors may highlight new areas for magmatic, tectonic, and upper crustal architecture, including
exploration where important deposits remain undiscovered. but not limited to an absence of localized arc magmatism with
In Chile, our findings support earlier studies that emphasize the spatio-temporal coincidence of a favorable localized stress
the intersection of arc-oblique structural corridors with the regime with fertile magma evolution (Richards, 2013; Tosdal
arc-parallel metallogenic belt as a necessary condition for the and Dilles, 2020; Loucks, 2021). Alternatively, they might
localization of giant porphyry copper deposit clusters (Figs. represent locations containing undiscovered porphyry depos-
14A, 15C; Salfity, 1985; Richards, 2003; Gow and Walshe, its, especially in areas concealed by postmineralization cover.
2005; Piquer et al., 2016; Espinoza et al., 2021).
The importance of arc-oblique fault systems may be ex- Conclusions
plained partly by the strike-slip geometry invoked on fault Our analysis demonstrates that though there is a clear arc-
systems that are “misoriented” with respect to the prevailing parallel structural fabric to the central Andes between 14°S

Fig. 15. Schematic sketches illustrating the evolution of the central Andean lithosphere at latitudes 20° to 30° S. Tectonic
models are fixed to a present-day South America reference frame. SCLM = Subcontinental lithospheric mantle. Cenozoic
margin from Arriagada et al. (2008); tectonic reconstruction models from Müller et al. (2019) and Young et al. (2019); paleo-
topography from the model of Scotese (2016). A) Triassic rifting associated with a combination of sinistral transform plate
boundary and internal Pangea breakup stresses produced NW- and N-oriented rift basins along southwest Pangea utilizing
preexisting Paleozoic lithospheric weaknesses. B) The separation of South America from Africa in the Early Cretaceous
produced a series of rift-related basins in Argentina that developed along preexisting structural architecture. C) Orocline
formation and relative clockwise rotation in the southern limb. Strain is largely absorbed by strike-slip reactivation within the
preexisting structural mesh, maintaining and augmenting its structural form. Strong plate coupling and crustal thickening
shut down volcanism, promoting the activation of segments of vertically coupled translithospheric permeability. The intersec-
tion of obliquely oriented, vertically coupled segments of translithosperic permeability provided an active channel for the
focused ascent of fluids, magma, and energy from the mantle to the upper crust. Adapted from Rhys et al. (2020).

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 21

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
22 FARRAR ET AL.

and 35°S latitude, there is also an arc-oblique structural fab- Alonso, R.N., and Viramonte, J.G., 1987, Geología y metalogenia de la Puna:
ric that exhibits periodic spacing (e.g., Salfity, 1985; Richards, Estudios geológicos, v. 43, p. 393–407.
Álvarez, J., Mpodozis, C., Arriagada, C., Astini, R., Morata, D., Salazar, E.,
2003; Yáñez and Rivera, 2019). The lithospheric architecture Valencia, V., and Vervoort, J., 2011, Detrital zircons from late Paleozoic
important for vertical fluid migration from the lithospheric accretionary complexes in north-central Chile (28–32 S): Possible finger-
mantle to the crust is defined by a mesh consisting of broadly prints of the Chilenia terrane: Journal of South American Earth Sciences,
NW-, NE-, and N-trending structural corridors. Recognizing v. 32, p. 460–476.
Amilibia, A., Sàbat, F., McClay, K., Muñoz, J., Roca, E., and Chong, G., 2008,
these continental-scale fault systems in the field involves the The role of inherited tectono-sedimentary architecture in the develop-
integration of local- and regional-scale data sets with obser- ment of the central Andean mountain belt: Insights from the Cordillera de
vations from large-scale geologic transects. Geologic recon- Domeyko: Journal of Structural Geology, v. 30, p. 1520–1539.
structions of tectonic cycles that define the evolution of the Anderson, C.A., 1966, Areal geology of the Southwest, in Titley, S.R., ed.,
central Andes show evidence that many structural corridors Geology of the porphyry copper deposits, southwestern North America:
University of Arizona, University of Arizona Press, p. 287.
are long lived and multiply reactivated. Regional-scale struc- Arriagada, C., Roperch, P., Mpodozis, C., and Cobbold, P.R., 2008, Paleo-
tural trends measured in surface geology mirror the traces of gene building of the Bolivian orocline: Tectonic restoration of the
regularly spaced regional lineaments observed in geophysical Central Andes in 2-D map view: Tectonics, v. 27, TC6014, 14 p., doi:
data sets. 10.1029/2008TC002269.
The spatial distribution of arc-oblique structural corridors Astini, R.A., and Thomas, W.A., 1999, Origin and evolution of the Precordil-
lera terrane of western Argentina: A drifted Laurentian orphan: Geological
in the central Andes reflects the interplay of the orientation Society of America Special Paper 336, p. 1–20.
of inherited pre-Andean lithospheric weaknesses that were Audin, L., Lacan, P., Tavera, H., and Bondoux, F., 2008, Upper plate defor-
further developed during Cenozoic compression and orocline mation and seismic barrier in front of Nazca subduction zone: The Chololo
formation. Focused, vertical translithospheric permeability is fault system and active tectonics along the Coastal Cordillera, southern
Peru: Tectonophysics, v. 459, p. 174–185.
achieved at the intersection of two or more structural corri- Bahlburg, H., and Hervé, F., 1997, Geodynamic evolution and tectonostrati-
dors during high-horizontal-rate-of-strain events and increas- graphic terranes of northwestern Argentina and northern Chile: Geological
es with increasing strain rate and pore-fluid pressure. Periods Society of America Bulletin, v. 109, p. 869–884.
of high compressional strain rate also favor the production, Begg, G., Griffin, W., Natapov, L., O'Reilly, S.Y., Grand, S.P., O'Neill, C.,
fractionation, and devolatilization of large volumes of fertile Hronsky, J., Djomani, Y.P., Swain, C., and Deen, T., 2009, The lithospheric
architecture of Africa: Seismic tomography, mantle petrology, and tectonic
hydrous magmas. The conjunction of these architectural and evolution: Geosphere, v. 5, p. 23–50.
geodynamic elements represents regions that underwent fa- Behn, G., Camus, F., Carrasco, P., and Ware, H., 2001, Aeromagnetic signa-
vorable preconditioning for the formation and localization of ture of porphyry copper systems in northern Chile and its geologic implica-
giant porphyry copper deposit clusters. tions: Economic Geology, v. 96, p. 239–248.
Benavides, S., 2017, Characterisation of sericitic alteration at Taca Taca
Acknowledgments Bajo, Argentina: M.Sc. thesis, Hobart, Australia, University of Tasmania,
162 p.
This study forms part of a Ph.D. by the primary author at the Bouzari, F., and Clark, A.H., 2002, Anatomy, evolution, and metallogenic
University of Tasmania (CODES). First Quantum Minerals significance of the supergene orebody of the Cerro Colorado porphyry
funded this research and provided permission to publish– copper deposit, I Región, northern Chile: Economic Geology, v. 97, p.
1701–1740.
their sustained support is gratefully acknowledged. This work Burov, E., and Watts, A., 2006, The long-term strength of continental litho-
benefited significantly from discussions with many geologists sphere: "jelly sandwich" or "crème brûlée"?: Geological Society of America
working in the central Andes. AF wishes to acknowledge the Today, v. 16, p. 4–10.
support in carrying out both theoretical and field-based com- Camus, F., and Dilles, J.H., 2001, A special issue devoted to porphyry copper
ponents of this research with First Quantum Minerals col- deposits of northern Chile: Economic Geology, v. 96, p. 233–237.
Cao, S., and Neubauer, F., 2016, Deep crustal expressions of exhumed strike-
leagues past and present: M. Trott, J. Cabrera, H. Munster, slip fault systems: Shear zone initiation on rheological boundaries: Earth-
S. Boucher, S. Arnillas, C. Manners, R. Rivera, S. Sykora, J. Science Reviews, v. 162, p. 155–176.
Burlando, D. Arribasplata, M. Duran, Y. Amezquita, J. Schorr, Carlier, G., Lorand, J.-P., Liégeois, J.-P., Fornari, M., Soler, P., Carlotto, V.,
R. Roberto, G. Cobeñas, G. Almandoz, J. Martínez, S. Galvez, and Cardenas, J., 2005, Potassic-ultrapotassic mafic rocks delineate two
lithospheric mantle blocks beneath the southern Peruvian Altiplano: Geol-
P. Salas, W. Alvarez, M. Hope, C. Wijns, S. Andersson, M.
ogy, v. 33, p. 601–604.
Ponce, M. Lappalainen, and M. Christie. R. Berry and G. Carlotto, V., 2013, Paleogeographic and tectonic controls on the evolution of
Begg are thanked for discussions that improved the manu- Cenozoic basins in the Altiplano and Western Cordillera of southern Peru:
script. We would like to sincerely thank J. Dilles, P. Perelló, A. Tectonophysics, v. 589, p. 195–219.
Tomlinson, R. Sillitoe, R. Tosdal, two anonymous reviewers, Carlotto, V., Acosta, H., Mamani, M., Cerpa, L., Rodriguez, R., Jaimes, F.,
Navarro, P., Cueva, E., and Chacaltana, C., 2010, Los dominios geotectóni-
and associate editor D. John for detailed reviews that signifi- cos del territorio peruano [ext. abs.]: Sociedad Geológica del Perú, Con-
cantly improved and focused the manuscript. greso Peruano de Geología, 15th, Cusco, 2010, Proceedings, p. 47–50.
Cembrano, J., and Lara, L., 2009, The link between volcanism and tectonics
REFERENCES in the southern volcanic zone of the Chilean Andes: A review: Tectonophys-
Acosta, J.G., Rivera Cornejo, R., Valencia Muñoz, M.M., Chirif Rivera, L.H., ics, v. 471, p. 96–113.
Huanacuni Mamani, D., Rodríguez Morante, I., Villarreal Jaramillo, E., Cembrano, J., Hervé, F., and Lavenu, A., 1996, The Liquiñe Ofqui fault
Paico Estrada, D., and Santisteban Angeldonis, A., 2009, Mapa metalo- zone: A long-lived intra-arc fault system in southern Chile: Tectonophysics,
genético del Perú 2009: Memoria: Lima, Peru, Instituto Geológico Minero v. 259, p. 55–66.
y Metalúrgico, 17 p. Charrier, R., Pinto, L., and Rodríguez, M.P., 2007, Tectonostratigraphic evo-
Allmendinger, R.W., Ramos, V.A., Jordan, T.E., Palma, M., and Isacks, B.L., lution of the Andean orogen in Chile, in Moreno, T., and Gibbons, W., eds.,
1983, Paleogeography and Andean structural geometry, northwest Argen- The geology of Chile: Geological Society, London, p. 21–114.
tina: Tectonics, v. 2, p. 1–16. Chatzaras, V., Tikoff, B., Newman, J., Withers, A.C., and Drury, M.R., 2015,
Allmendinger, R.W., Smalley Jr., R., Bevis, M., Caprio, H., and Brooks, B., Mantle strength of the San Andreas fault system and the role of mantle-
2005, Bending the Bolivian orocline in real time: Geology, v. 33, p. 905–908. crust feedbacks: Geology, v. 43, p. 891–894.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 23

Chernicoff, C.J., Richards, J.P., and Zappettini, E.O., 2002, Crustal lineament Holdsworth, R., Butler, C., and Roberts, A., 1997, The recognition of reac-
control on magmatism and mineralization in northwestern Argentina: Geo- tivation during continental deformation: Journal of the Geological Society,
logical, geophysical, and remote sensing evidence: Ore Geology Reviews, v. 154, p. 73–78.
v. 21, p. 127–155. Horton, B.K., 2018, Sedimentary record of Andean mountain building:
Clark, A., 1993, Are outsize porphyry copper deposits either anatomically or Earth-Science Reviews, v. 178, p. 279–309.
environmentally distinctive?: Society of Economic Geologists Special Pub- Inbal, A., Ampuero, J.P. and Clayton, R.W., 2016, Localized seismic deforma-
lication 2, p. 213–283. tion in the upper mantle revealed by dense seismic arrays: Science, v. 354,
Cobbing, E., 1982, The segmented coastal Batholith of Peru: Its relation- p. 88–92.
ship to volcanicity and metallogenesis: Earth-Science Reviews, v. 18, p. Ince, M., Hagemann, S., Fiorentini, M., Kemp, T., Rubinstein, N., Ireland,
241–251. T., and Banyard, J., 2022, Tracking magma replenishment through zircon
Coira, B., Davidson, J., Mpodozis, C., and Ramos, V., 1982, Tectonic and petrochronology of the Taca Taca Bajo Cu-Mo-(Au) porphyry deposit, NW
magmatic evolution of the Andes of northern Argentina and Chile: Earth- Argentina: Society of Economic Geologists, SEG 2022 Conference: Miner-
Science Reviews, v. 18, p. 303–332. als For Our Future, Denver, Colorado, 2022, Proceedings, S3.04.
Coleman, A.J., Duffy, O.B., and Jackson, C.A.L., 2019, Growth folds above Ingemmet, 1999, Mapa geológico del Peru, in Lecaros, W., Moncayo, O., Vil-
propagating normal faults: Earth Science Reviews, v. 196, article 102885. chez, L., and Fernández, A., eds., Carta Geológica Nacional: Perú, Ingem-
Cooke, D.R., Hollings, P., and Walshe, J.L., 2005, Giant porphyry deposits: met, 74 p.
Characteristics, distribution, and tectonic controls: Economic Geology, v. Isacks, B.L., 1988, Uplift of the central Andean plateau and bending of the
100, p. 801–818. Bolivian Orocline: Journal of Geophysical Research: Solid Earth, v. 93, p.
Cox, S., 2020, The dynamics of permeability enhancement and fluid flow in 3211–3231.
overpressured, fracture-controlled hydrothermal systems: Reviews in Eco- Isles, D.J., and Rankin, L.R., 2013, Geological interpretation of aeromag-
nomic Geology, v. 21, p. 25–82. netic data: Perth, Society of Exploration Geophysicists and Australian Soci-
Dilles, J.H., Tomlinson, A.J., Martin, M.W., and Blanco, N., 1997, The El ety of Exploration, 365 p.
Abra and Fortuna granodiorite complexes: A porphyry copper batholith Jacques, J.M., 2003, A tectonostratigraphic synthesis of the Sub-Andean
sinistrally displaced by the Falla Oeste [ext. abs.]: Congreso Geológico basins: Implications for the geotectonic segmentation of the Andean belt:
Chileno, 8th, Antofagasta, 1997, Actas, p. 1883–1887. Journal of the Geological Society, London, v. 160, p. 687–701.
Eichelberger, N., and McQuarrie, N., 2015, Kinematic reconstruction of the Jordan, T.E., Isacks, B.L., Allmendinger, R.W., Brewer, J.A., Ramos, V.A., and
Bolivian orocline: Geosphere, v. 11, p. 445–462. Ando, C.J., 1983, Andean tectonics related to geometry of subducted Nazca
Espinoza, M., Montecino, D., Oliveros, V., Astudillo, N., Vásquez, P., Reyes, plate: Geological Society of America Bulletin, v. 94, p. 341–361.
R., Celis, C., González, R., Contreras, J., and Creixell, C., 2019, The synrift Hirth, G., and Kohlstedt, D.L., 1996, Water in the oceanic upper mantle:
phase of the early Domeyko basin (Triassic, northern Chile): Sedimentary, Implications for rheology, melt extraction and the evolution of the litho-
volcanic, and tectonic interplay in the evolution of an ancient subduction‐ sphere: Earth and Planetary Science Letters, v. 144, p. 93–108.
related rift basin: Basin Research, v. 31, p. 4–32. Koulakov, I., Sobolev, S. V., and Asch, G., 2006, P-and S-velocity images of the
Espinoza, M., Oliveros, V., Vásquez, P., Giambiagi, L., Morgan, L., González, lithosphere—asthenosphere system in the Central Andes from local-source
R., Solari, L., and Bechis, F., 2021, Gondwanan inheritance on the build- tomographic inversion: Geophysical Journal International, v. 167, p. 106-126.
ing of the western central Andes (Domeyko Range, Chile): Structural and Lanza, F., Tibaldi, A., Bonali, F., and Corazzato, C., 2013, Space-time varia-
thermochronological approach (U‐Pb and 40Ar/ 39Ar): Tectonics, v. 40, doi: tions of stresses in the Miocene-Quaternary along the Calama-Olacapato-
10.1029/2020TC006475. El Toro fault zone, Central Andes: Tectonophysics, v. 593, p. 33–56.
Ford, H.A., Fischer, K.M., and Lekic, V., 2014, Localized shear in the deep Lavallée, Y., de Silva, S.L., Salas, G., and Byrnes, J.M., 2009, Structural
lithosphere beneath the San Andreas fault system: Geology, v. 42, p. control on volcanism at the Ubinas, Huaynaputina, and Ticsani Volcanic
295–298. Group (UHTVG), southern Peru: Journal of Volcanology and Geothermal
Franceschinis, P.R., Rapalini, A.E., Escayola, M.P., and Piceda, C.R., 2020, Research, v. 186, p. 253–264.
Paleogeographic and tectonic evolution of the Pampia Terrane in the Lesur, V., Hamoudi, M., Choi, Y., Dyment, J., and Thébault, E., 2016,
Cambrian: New paleomagnetic constraints: Tectonophysics, v. 779, article Building the second version of the world digital magnetic anomaly map
228386. (WDMAM): Earth, Planets and Space, v. 68, p. 1–13.
García, W.M., 1978, Geología del cuadrángulo de Puquina, Omate, Huaitire, Lezaeta, P., 2001, Distortion analysis and 3-D modeling of magnetotelluric
Mazo Cruz, Pizacoma, 34-t, 34-u, 34-v, 34-x, 34-y- [Boletín A 29]: Perú, data in the Southern Central Andes: Unpub. doctoral thesis, Freie Univer-
Instituto de Geologia y Minería, 66 p. sität Berlin, Berlin, Germany, 550 p.
Gilluly, J., 1976, Lineaments; ineffective guides to ore deposits: Economic Li, Y., Weng, A., Xu, W., Zou, Z., Tang, Y., Zhou, Z., Li, S., Zhang, Y., and
Geology, v. 71, p. 1507–1514. Ventura, G., 2021, Translithospheric magma plumbing system of intraplate
Gioncada, A., Vezzoli, L., Mazzuoli, R., Omarini, R., Nonnotte, P., and Guil- volcanoes as revealed by electrical resistivity imaging: Geology, v. 49, p.
lou, H., 2010, Pliocene intraplate-type volcanism in the Andean foreland 1337–1342.
at 26° 10′ S, 64° 40′ W (NW Argentina): Implications for magmatic and Loucks, R.R., 2021, Deep entrapment of buoyant magmas by orogenic tec-
structural evolution of the Central Andes: Lithosphere, v. 2, p. 153–171. tonic stress: Its role in producing continental crust, adakites, and porphyry
Glen, R., and Walshe, J., 1999, Cross‐structures in the Lachlan orogen: The copper deposits: Earth-Science Reviews, v. 220, article 103744.
Lachlan transverse zone example: Australian Journal of Earth Sciences, v. Love, D.A., Clark, A.H., and Glover, J.K., 2004, The lithologic, stratigraphic,
46, p. 641–658. and structural setting of the giant Antamina copper-zinc skarn deposit,
Gow, P.A., and Walshe, J.L., 2005, The role of preexisting geologic architec- Ancash, Peru: Economic Geology, v. 99, p. 887–916.
ture in the formation of giant porphyry-related Cu ± Au deposits: Examples Lowell, J.D., 1974, Regional characteristics of porphyry copper deposits of
from New Guinea and Chile: Economic Geology, v. 100, p. 819–833. the Southwest: Economic Geology, v. 69, p. 601–617.
Gray, D., Lawlor, M., and Briggs, A., 2020, Taca Taca project, Salta Province, Lutz, B.M., Axen, G.J., van Wijk, J.W., and Phillips, F.M., 2022, Whole-litho-
Argentina, NI 43-101 Technical Report: Canada, First Quantum Minerals, sphere shear during oblique rifting: Geology, v. 50, p. 412–416.
342 p. Mamani, M., Tassara, A., and Wörner, G., 2008, Composition and structural
Haugaard, R., Justina, F.D., Roots, E., Cheraghi, S., Vayavur, R., Hill, G., control of crustal domains in the central Andes: Geochemistry, Geophysics,
Snyder, D., Ayer, J., Naghizadeh, M., and Smith, R., 2021, Crustal-scale Geosystems, v. 9, Q03006, doi: 10.1029/2007GC001925.
geology and fault geometry along the gold-endowed Matheson transect of Martinez, C., Soria, E., and Uribe, H., 1995, Deslizamiento de cobertura en
the Abitibi greenstone belt: Economic Geology, v. 116, p. 1053–1072. el sinclinorio mesocenozoico de Sevaruyo-Rio Mulato (Altiplano central de
Hayward, N., Doutre, R., and Micklethwaite, S., 2018, Spatial periodicity in Bolivia): Revista Técnica de Yacimientos Petrolíferos Fiscales Bolivianos,
self-organized ore systems: Society of Economic Geologists Special Publi- v. 16, p. 9–26.
cation 21, p. 1–24. Matteini, M., Mazzuoli, R., Omarini, R., Cas, R., and Maas, R., 2002, Geody-
Hill, K.C., Kendrick, R.D., Crowhurst, P.V., and Gow, P.A., 2002, Copper‐ namical evolution of Central Andes at 24 S as inferred by magma composi-
gold mineralisation in New Guinea: Tectonics, lineaments, thermochronol- tion along the Calama-Olacapato-El Toro transversal volcanic belt: Journal
ogy and structure: Australian Journal of Earth Sciences, v. 49, p. 737–752. of Volcanology and Geothermal Research, v. 118, p. 205–228.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
24 FARRAR ET AL.

Mayo, E.B., 1958, Lineament tectonics and some ore districts of the south- relationship with magmatic and hydrothermal activity: Solid Earth, v. 12,
west: Mining Engineering, v. 10, p. 1169–1175. p. 253–273.
McCuaig, T.C., and Hronsky, J.M., 2014, The mineral system concept: The Prinz, P., Wilke, H., and Hillebrandt, A., 1994, Sediment accumulation and
key to exploration targeting: Society of Economic Geologists Special Pub- subsidence history in the Mesozoic marginal basin of northern Chile, in
lication 18, p. 153–175. Reutter, K.-J., Scheuber, E., and Wigger, P., eds., Tectonics of the southern
McCuaig, T.C., Beresford, S., and Hronsky, J., 2010, Translating the mineral Central Andes: Springer-Verlag, p. 219–232.
systems approach into an effective exploration targeting system: Ore Geol- Puigdomenech, C., Somoza, R., Tomlinson, A., and Renda, E., 2020, Paleo-
ogy Reviews, v. 38, p. 128–138. magnetic data from the Precordillera of northern Chile: A multiphase rota-
McGroder, M.F., Lease, R.O., and Pearson, D.M., 2015, Along-strike varia- tion history related to a multiphase deformational history: Tectonophysics,
tion in structural styles and hydrocarbon occurrences, Subandean fold-and- v. 791, article 228569.
thrust belt and inner foreland, Colombia to Argentina: Geological Society Ramírez, C., Gardeweg, M., Davidson, J., and Pino, H., 1991, Mapa geológico
of America, Memoir 212, doi: 10.1130/2015.1212(05). del área de los volcanes Socompa y Pular, Región de Antofagasta. Escala
Merdith, A.S., Williams, S.E., Collins, A.S., Tetley, M.G., Mulder, J.A., 1:100,000: Chile, Servicio Nacional de Geología y Minería, documentos de
Blades, M.L., Young, A., Armistead, S.E., Cannon, J., Zahirovic, S., and trabajo, 4.
Müller, R.D., 2021, Extending full-plate tectonic models into deep time: Ramos, V.A., 1977, Basement tectonics from Landsat imagery in mining
Linking the Neoproterozoic and the Phanerozoic: Earth-Science Reviews, exploration: Geologie en Mijnbouw, v. 56, p. 243–252.
v. 214, article 103477. ——1994, Terranes of southern Gondwanaland and their control in the
Mpodozis, C., and Cornejo, P., 2012, Cenozoic tectonics and porphyry cop- Andean structure (30°–33°S Latitude), in Reutter, K.-J., Scheuber, E., and
per systems of the Chilean Andes: Society of Economic Geologists Special Wigger, P.J., eds., Tectonics of the southern central Andes: Berlin, Springer,
Publication 16, p. 329–360. p. 334.
Mpodozis, C., and Ramos, V., 1989, The Andes of Chile and Argentina: Cir- ——2010, The Grenville-age basement of the Andes: Journal of South Amer-
cum-Pacific Council Energy Mineral Resources Earth Science Series, v. 11, ican Earth Sciences, v. 29, p. 77–91.
p. 59–90. Reutter, K.-J., Scheuber, E., and Chong, G., 1996, The Precordilleran fault
Müller, R.D., Zahirovic, S., Williams, S.E., Cannon, J., Seton, M., Bower, system of Chuquicamata, northern Chile: Evidence for reversals along arc-
D.J., Tetley, M.G., Heine, C., Le Breton, E., Liu, S., Russell, S.H.J., Yang, parallel strike-slip faults: Tectonophysics, v. 259, p. 213–228.
T., Leonard, J., and Gurnis, M., 2019, A global plate model including litho- Rhys, D.A., Lewis, P.D., and Rowland, J.V., 2020, Structural controls on ore
spheric deformation along major rifts and orogens since the Triassic: Tec- localization in epithermal gold-silver deposits: A mineral systems approach:
tonics, v. 38, p. 1884–1907. Reviews in Economic Geology, v. 21, p. 83–145.
Muntean, J.L., and Einaudi, M.T., 2000, Porphyry gold deposits of the Refu- Ricci, M., and Figueroa, L., 1971, Fotolineamientos y mineralización en el
gio district, Maricunga belt, northern Chile: Economic Geology, v. 95, p. noroeste Argentino, exploracion minera de la region noroeste Argentina:
1445−1472. Tucuman, United Nations, p. 45.
Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P., and Schemmann, K., Richards, J.P., 2000, Lineaments revisited: Society of Economic Geologists
2006, Deformation of the central Andean upper plate system—Facts, fic- Newsletter 42, p. 1, 14–20.
tion, and constraints for plateau models, in Oncken, O., Chong, G., Franz, ——2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au) deposit
G., Giese, P., Gotze, H.-J., and Ramos, V., eds., The Andes: Berlin, Heidel- formation: Economic Geology, v. 98, p. 1515–1533.
berg, Springer, p. 3–27. ——2013, Giant ore deposits formed by optimal alignments and combina-
Orts, S., Almandoz, G., Gigola, S., Vazquez, P., and Bocanegra, L., 2011, tions of geological processes: Nature Geoscience, v. 6, p. 911–916.
Control estructural en el sistema porfírico de cu-mo-au Taca Taca Bajo: Richards, J.P., and Villeneuve, M., 2002, Characteristics of late Cenozoic
Reactivación polifásica de estructuras previas: XVIII Congreso Geológico volcanism along the Archibarca lineament from Cerro Llullaillaco to Cor-
Argentino, Neuquén, 2011, Actas, p. 1138–1139. rida de Cori, northwest Argentina: Journal of Volcanology and Geothermal
Palacios, C., Ramírez, L.E., Townley, B., Solari, M., and Guerra, N., 2007, Research, v. 116, p. 161–200.
The role of the Antofagasta-Calama lineament in ore deposit deformation Richards, J.P., Boyce, A.J., and Pringle, M.S., 2001, Geologic evolution of the
in the Andes of northern Chile: Mineralium Deposita, v. 42, p. 301–308. Escondida area, northern Chile: A model for spatial and temporal localiza-
Paterson, S., and Schmidt, K., 1999, Is there a close spatial relationship tion of porphyry Cu mineralization: Economic Geology, v. 96, p. 271–305.
between faults and plutons?: Journal of Structural Geology, v. 21, p. Rivera, O., and Cerda, A., 2012, Los Pórfidos Cupríferos de Chile Central:
1131–1142. Significado de estructuras translitosféricas y anomalías gravimétricas en
Perelló, J., Carlotto, V.C., Zárate, A., Ramos, P., Posso, H.C., Neyra, C., la metalogénesis Andina: XIII Congreso Geológico Chileno, Antofagasta,
Caballero, A., Fuster, N.S., and Muhr, R., 2003, Porphyry-style alteration 2012, Actas, p. 2–4.
and mineralization of the middle Eocene to early Oligocene Andahuaylas- Rode, M., Guitart, A., Sanguinetti, M., and Richard, F., 2015, El depósito tipo
Yauri belt, Cuzco region, Peru: Economic Geology, v. 98, p. 1575–1605. pórfido Cu-Au de Los Helados, III Región, Atacama, Chile: XIV Congreso
Perez, N.D., Horton, B.K., McQuarrie, N., Stuebner, K., and Ehlers, T.A., Geológico Chileno, La Serena, 2015, Actas, p. 373–376.
2016, Andean shortening, inversion and exhumation associated with thin- Roperch, P., Carlotto, V., Ruffet, G., and Fornari, M., 2011, Tectonic rota-
and thick-skinned deformation in southern Peru: Geological Magazine, v. tions and transcurrent deformation south of the Abancay deflection in the
153, p. 1013–1041. Andes of southern Peru: Tectonics, v. 30, p. 1–23.
Peri, V.G., Barcelona, H., Pomposiello, M.C., and Favetto, A., 2015, Mag- Rubinstein, N.A., Zappettini, E.O., and Gómez, A.L., 2021, Porphyry Cu
netotelluric characterization through the Ambargasta-Sumampa Range: deposits in the Central Andes of Argentina: An overview: Journal of South
The connection between the northern and southern trace of the Río de La American Earth Sciences, v. 112, article 103543.
Plata craton-Pampean Terrane tectonic boundary: Journal of South Ameri- Sagy, A., Reches, Z., and Agnon, A., 2003, Hierarchic three‐dimensional
can Earth Sciences, v. 59, p. 1–12. structure and slip partitioning in the western Dead Sea pull‐apart: Tecton-
Piquer, J., Berry, R.F., Scott, R.J., and Cooke, D.R., 2016, Arc-oblique fault ics, v. 22, doi: 10.1029/2001TC001323.
systems: Their role in the Cenozoic structural evolution and metallogen- Salfity, J., 1985, Lineamentos transversales al rumbo andino en el Noroeste
esis of the Andes of central Chile: Journal of Structural Geology, v. 89, p. Argentino: IV Congreso Geologico Chileno, Antofagasta, 1985, Actas, v. 1,
101–117. p. 2/119–2/137.
Piquer, J., Hollings, P., Rivera, O., Cooke, D.R., Baker, M., and Testa, F., Sawyer, D.S., 1985, Brittle failure in the upper mantle during extension of
2017, Along-strike segmentation of the Abanico Basin, central Chile: New continental lithosphere: Journal of Geophysical Research: Solid Earth, v.
chronological, geochemical and structural constraints: Lithos, v. 268, p. 90, p. 3021–3025.
174–197. Schellart, W.P., 2017, Andean mountain building and magmatic arc migration
Piquer, J., Sanchez-Alfaro, P., and Perez-Flores, P., 2021a, A new model for driven by subduction-induced whole mantle flow: Nature Communications,
the optimal structural context for giant porphyry copper deposit formation: v. 8, doi: 10.1038/s41467-017-01847-z.
Geology, v. 49, p. 597–601. Scheuber, E., and Andriessen, P.A., 1990, The kinematic and geodynamic
Piquer, J., Rivera, O., Yáñez, G., and Oyarzún, N., 2021b, The Piuquen- significance of the Atacama fault zone, northern Chile: Journal of Structural
cillo fault system: A long-lived, Andean-transverse fault system and its Geology, v. 12, p. 243–257.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user
LITHOSPHERIC ARCHITECTURE OF THE CENTRAL ANDES 25

Schodde, R., 2019, Exploration and discovery of base and precious metal Tosdal, R.M., and Dilles, J.H., 2020, Creation of permeability in the porphyry
deposits in the Pacific Rim over the last 50 years, PACRIM 2019: Auckland, Cu environment: Reviews in Economic Geology, v. 21, p. 173–204.
MinEx Consulting, 47 p., https://minexconsulting.com/exploration-and-dis- Tosdal, R.M., and Richards, J.P., 2001, Magmatic and structural controls on
covery-of-base-and-precious-metal-deposits-in-the-pacific-rim-over-the- the development of porphyry Cu ± Mo ± Au deposits: Reviews in Economic
last-50-years. Geology, v. 14, p. 157–181.
Scotese, C.R., 2016, Tutorial: PALEOMAP paleoAtlas for GPlates and the Umhoefer, P.J., 2011, Why did the southern Gulf of California rupture so rap-
paleoData plotter program, Technical Report, 56, https://www. earthbyte. idly?—Oblique divergence across hot, weak lithosphere along a tectonically
org/paleomap. active margin: GSA today, v. 21, p. 4–10.
Segemar, 1997, Mapa geológico de la República Agentina: Buenos Aires, Sec- United States Geological Survey (USGS), 2021, Mineral commodity sum-
retaría de Industria, Comercio y Minería, Segemar. maries 2021, Mineral Commodity Summaries: Virginia, U.S. Geological
Segerstrom, K., 1970, Apollo 7 photography in Antofagasta Province, Chile Survey, 200 p.
- An interpretation, in United States Geological Survey, ed., Geological sur- Vicente, J.-C., 1989, Early late Cretaceous overthrusting in the western cor-
vey research 1970, 700-D: Washington, United States government printing dillera of southern Peru: Circum-Pacific Council for Energy and Mineral
office, p. D10–D17. Resources Earth Science Series, v. 11, p. 91–117.
Sempere, T., Hérail, G., Oller, J., and Bonhomme, M.G., 1990, Late Oli- Wiemer, D., Hagemann, S.G., Hronsky, J., Kemp, A.I., Thébaud, N., Ire-
gocene-early Miocene major tectonic crisis and related basins in Bolivia: land, T., and Villanes, C., 2022, Ancient structural inheritance explains gold
Geology, v. 18, p. 946–949. deposit clustering in northern Perú: Geology v. 50, p. 1197–1201.
Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe, Wyborn, L.A.I., Heinrich, C.A., and Jaques, A.L., 1994, Australian Protero-
O., Néraudeau, D., Cárdenas, J., and Rosas, S., 2002, Late Permian-Middle zoic mineral systems: Essential ingredients and mappable criteria: Austra-
Jurassic lithospheric thinning in Peru: Tectonophysics, v. 345, p. 153–181. lian Institute of Mining and Metallurgy, The AusIMM Annual Conference,
Servicio Geologico Minero (Sergeomin), 2020, Mapa geologico de Bolivia Darwin, 1994, Proceedings, p. 109–115.
1M: Servicio Geologico Minero. Xiaodi, T., Guoqing, M., and Dailei, Z., 2018, Application of image process-
Servicio Nacional de Geología y Minería (SERNAGEOMIN), 2002, Mapa ing methods in edge detection of potential field data: ASEG Extended
geológico de Chile: Carta Geológica de Chile, Serie Geología Básica No. Abstracts, v. 2018, p. 1–4.
75, 1 map in 3 sheets, 1:1,000,000 scale. Yáñez, G., and Maksaev, V., 1994, Sobre la distribución espacial de cuerpos
Shackleton, R., Ries, A., Coward, M., and Cobbold, P., 1979, Structure, meta- intrusivos asociados a diapirismo y la caracterización de la fuente mag-
morphism and geochronology of the Arequipa Massif of coastal Peru: Jour- mática en pórfidos cupríferos: Inestabilidades gravitacionales en un medio
nal of the Geological Society, London, v. 136, p. 195–214. viscoso [ext. abst.]: Congreso Geológico Chileno, Antofagasta, Chile, 1994,
Sibson, R.H., 1996, Structural permeability of fluid-driven fault-fracture Actas, p. 1642–1646.
meshes: Journal of Structural Geology, v. 18, p. 1031–1042. Yáñez, G., and Rivera, O., 2019, Crustal dense blocks in the fore-arc and arc
Sillitoe, R.H., 1977, Permo-Carboniferous, upper Cretaceous, and Miocene region of Chilean ranges and their role in the magma ascent and compo-
porphyry copper-type mineralization in the Argentinian Andes: Economic sition: Breaking paradigms in the Andean metallogeny: Journal of South
Geology, v. 72, p. 99–103. American Earth Sciences, v. 93, p. 51–66.
——1988, Epochs of intrusion-related copper mineralization in the Andes: Young, A., Flament, N., Maloney, K., Williams, S., Matthews, K., Zahirovic, S.,
Journal of South American Earth Sciences, v. 1, p. 89–108. and Müller, R.D., 2019, Global kinematics of tectonic plates and subduction
——2010a, Exploration and discovery of base-and precious-metal deposits in zones since the late Paleozoic Era: Geoscience Frontiers, v. 10, p. 989–1013.
the circum-Pacific region - a 2010 perspective: Resource Geology, Special Zamora, G., and Gil, W., 2018, The Marañón basin: Tectonic evolution and
Issue 22, 139 p. paleogeography, in Zamora, G., McClay, K.R., and Ramos, V.A., eds., Petro-
——2010b, Porphyry copper systems: Economic Geology, v. 105, p. 3–41. leum basins and hydrocarbon potential of the Andes of Peru and Bolivia,
Sillitoe, R.H., and Perelló, J., 2005, Andean copper province: Tectonomag- 117: United Kingdom, The American Association of Petroleum Geologists,
matic settings, deposit types, metallogeny, exploration, and discovery: Eco- p. 121–144.
nomic Geology, 100th Anniversary Volume, p. 845–890. Zappettini, E.O., Blasco, G., Ramallo, E.E., and González, O.E., 2001, Hoja
Sillitoe, R.H., Devine, F.A., Sanguinetti, M.I., and Friedman, R.M., 2019, geológica 2569-II Socompa. Escala 1:250,000: Servicio Geológico Minero
Geology of the Josemaría porphyry copper-gold deposit, Argentina: Forma- Argentino, Instituto de Geología y Recursos Minerales, Boletín 260, 62 p.
tion, exhumation, and burial in two million years: Economic Geology, v.
114, p. 407–426.
Simmons, A.T., Tosdal, R.M., Wooden, J.L., Mattos, R., Concha, O.,
McCracken, S., and Beale, T., 2013, Punctuated magmatism associated
with porphyry Cu-Mo formation in the Paleocene to Eocene of southern
Peru: Economic Geology, v. 108, p. 625–639.
Somoza, R., and Ghidella, M.E., 2012, Late Cretaceous to recent plate
motions in western South America revisited: Earth and Planetary Science
Letters, v. 331, p. 152–163.
Somoza, R., and Zaffarana, C.B., 2008, Mid-Cretaceous polar standstill of
South America, motion of the Atlantic hotspots and the birth of the Andean
cordillera: Earth and Planetary Science Letters, v. 271, p. 267–277.
Stuwe, K., 2007, Geodynamics of the lithosphere. An introduction: Berlin,
Springer, 493 p.
Thomas, A., 1970, Beitrag zur tektonik Nordchiles: Geologische Rundschau,
Alex Farrar is undertaking a Ph.D. at CODES,
v. 59, p. 1013–1027.
Tibaldi, A., Corazzato, C., and Rovida, A., 2009, Miocene-Quaternary struc- University of Tasmania, where his research involves
tural evolution of the Uyuni-Atacama region, Andes of Chile and Bolivia: using data analytics to investigate the link between
Tectonophysics, v. 471, p. 114–135. structure, tectonics, and the formation of giant por-
Tomlinson, A.J., and Blanco, N., 1997, Structural evolution and displacement phyry copper deposits. He obtained his B.Sc. (Hons)
history of the West Fault System, Precordillera, Chile: Part 1, synmineral from the University of Melbourne (2006) and gained
history [ext. abs.]: Congreso Geológico Chileno, Antofagasta, Chile, 1997, an M.Sc. from the University of Western Australia
Actas, p. 1878–1882. (2015). Before commencing his Ph.D., Alex worked
Tomlinson, A.J., and Cornejo, P., 2012, Regional distribution of Middle with First Quantum Minerals for 11 years leading greenfield porphyry cop-
Eocene-Early Oligocene porphyry copper centers in Northern Chile: per exploration in the central Andes and worked in sediment-hosted copper
Second-order patterns and possible causes: Congreso Geológico Chileno, exploration and mining in the Central African Copperbelt. Prior to this, Alex
Antofagasta, Chile, 2012, Actas, p. 40–42. worked at George Fisher mine, Queensland, with Mount Isa mines.

Downloaded from http://pubs.geoscienceworld.org/segweb/economicgeology/article-pdf/doi/10.5382/econgeo.5010/5939365/5010_farrar+et+al.pdf


by INGEMMET user

You might also like