Вы находитесь на странице: 1из 32

2012 Society of Economic Geologists, Inc.

Special Publication 16, pp. 329360

Chapter 14
Cenozoic Tectonics and Porphyry Copper Systems of the Chilean Andes
CONSTANTINO MPODOZIS AND PAULA CORNEJO
Antofagasta Minerals, Apoquindo 4001, Piso 18, Santiago, Chile

Abstract
Subduction under South America has been active for the past 550 m.y. but large porphyry copper deposits
were essentially emplaced during the Paleocene (6050 Ma) in southern Peru, and mid-Eocene-early
Oligocene (4332 Ma) and late Miocene-Pliocene (106 Ma) in north and central Chile. Although the tectonic
setting of the Paleocene porphyry deposits is still poorly understood, those of the northern Chile EoceneOligocene belt were emplaced along the margin-parallel Domeyko fault system, where active compressional
and/ortranspressional deformation and block rotations took place during the formation of the Bolivian orocline.
Eocene-early Oligocene oroclinal bending was a consequence of differential tectonic shortening focused along
a mechanically weak zone of the Central Andean crust inherited from the Paleozoic. Deformation occurred
during an episode of accelerated westward absolute motion of the South American plate, which coincided with
very high rates of oceanic crust production in the eastern Pacific. The slow South American-Farallon convergence rates recorded for the Eocene-Oligocene suggest, however, that strong interplate coupling existed during that time. This permitted the transfer of horizontal stresses and large-scale deformation of the Andean margin, creating a favorable scenario for the generation and emplacement of porphyry copper magmas along the
Domeyko fault system.
The younger, Miocene-Pliocene porphyry copper deposits of central Chile-Argentina were emplaced in a
different setting, after the initiation of compressional deformation within a volcano-tectonic depression (Abanico basin) that evolved during another, late Oligocene to early Miocene, period of increased East Pacific
oceanic crust production. Nevertheless, in contrast to the Eocene-Oligocene situation in northern Chile, the
relatively stationary position of the South American plate compared to the mantle reference frame and weak
interplate coupling that permitted rapid subduction, increased volcanism, and overriding plate extension. Tectonic inversion of the basin and compressional deformation along with crustal thickening and mountain building began at around 20 m.y. ago as interplate coupling increased when the westward motion of South America
accelerated and the Nazca-South America convergence velocity decreased in the mid-Miocene. Compression
was accompanied, as during the Eocene-Oligocene in northern Chile, by slab shallowing and increased forearc subduction erosion.
In both cases, the largely structurally controlled, syn- to post-tectonic porphyry copper deposits are associated with long-lived magmatic systems that were active for more than 10 m.y. In northern Chile, the deposits
occur as parts of discrete intrusive clusters that comprise a suite of precursor plutons emplaced during multiple events since the Cretaceous. Porphyry copper mineralization is linked to multistage, amphibole-bearing intrusions of intermediate composition derived from hydrous, oxidized magmas with adakitic geochemical signatures. These intrusions appeared when crustal thickness increased to a critical threshold in the course of
deformation. Production of magmas with high metal-carrying capacity was fostered as fluids were liberated
when amphibole became unstable and was destroyed as the crust thickened. At the same time, source regions
within the mantle were contaminated by hydrated fragments of fore-arc continental crust, as the result of enhanced subduction erosion during peaks of compressional deformation.

Introduction
THE STUDY of the tectonic setting of porphyry copper deposits
is fundamental to understanding their genesis (e.g., Sillitoe,
1998; Kay and Mpodozis, 2001; Cooke et al., 2005; Sillitoe
and Perell, 2005; Richards, 2009, 2011a; Tosdal et al., 2009).
Some Cenozoic porphyry copper deposits are known to have
formed during or shortly after continent-continent, continent-island arc, or island arc-island arc collisions in the Himalayas-Tibet, the Kerman arc in Iran, and Papua New
Guinea (Solomon, 1990; Zenqiang et al., 2003; Shaifei et al.,
2009). A Paleozoic example of this type of deposit may be
Oyu Tolgoi in Mongolia (Perell et al., 2001). In contrast,
other large porphyry deposits such as Bingham Canyon in the
western United States formed during the earliest stages of
Corresponding

author: e-mail, cmpodozis@aminerals.cl

Basin and Range extension in the Eocene, far inland from the
Pacific margin of North America (Kloppenburgh et al., 2010).
Noncollisional porphyry copper deposit examples in subduction-related arc settings include those from the Chagai
belt in Pakistan, the Laramide porphyry copper province of
the western United States and northern Mexico (Lang and Titley, 1998; Valencia-Moreno et al., 2007; Perell et al., 2008)
and the Central Andes province, which host some of the
largest known porphyry copper deposits in the world (Camus,
2003; Cooke et al., 2005; Sillitoe and Perell, 2005).
The Andes has long been considered as the type example of
a noncollisional orogenic system (e.g., Jordan et al., 1983),
where subduction of Pacific oceanic crust beneath South
America has been active for the past 570 m.y. (Cawood,
2005). Nevertheless, the largest porphyry copper deposits are
the result of anomalous magmatic systems that developed

329

330

MPODOZIS AND CORNEJO

during short periods at specific locations within the Andean


orogen. These include the Paleocene to early Eocene (6652
Ma) and middle Eocene to early Oligocene (4332 Ma) belts
in southern Peru and northern Chile, and the late Miocene to
early Pliocene (105 Ma) porphyry systems in central Chile
and contiguous Argentina (Perell et al., 2003a; Sillitoe and
Perell, 2005). In this contribution, with emphasis on the
Chilean belts, we will try to demonstrate how major Cenozoic
tectonic events along the central Andean convergent margin,
prompted by large-scale reorganizations of the global tectonic
system, were the main triggers for the formation of large porphyry copper deposits.
Pre-Andean History:
From Rodinia Dispersal to Pangea Breakup
The western margin of South America underwent magmatic and tectonic activity at least since the late Neoproterozoic breakup of Rodinia (800700 Ma), when the separation
of Laurentia from Gondwana produced the opening of the
proto-Pacific (Iapetus) ocean (Dalziel, 1997). East-directed
subduction of newly formed ancestral Pacific crust below
western Gondwana began at ~570 Ma and was fully active
along the proto-Andean margin by 485 to 465 Ma (Pankhurst
et al., 1998; Cawood, 2005; Chew et al., 2007). Plate convergence in the Central Andes region during the Ordovician to
Devonian included the progressive collision and accretion of
a group of tectonostratigraphic terranes of Laurentian and/or
Gondwanan affinities (e.g., Ramos et al., 1986; Astini et al.,
1995) against the western South American margin. Terrane
amalgamation contributed to the formation of the accretionary Terra Australis orogen, which extended for more than
18,000 km along the Pacific margin of Gondwana from Australia to South America (Cawood, 2005). The accretionary
stage was followed, in the Central Andes, by the buildup of a
late Carboniferous to Early Permian (320? -280 Ma) suprasubduction magmatic arc on top of the newly accreted terranes,
as well as the development of an outboard fore-arc subduction complex that extended for more than 1,000 km along the
Chilean segment of the Gondwana margin south of 27 S
(Mpodozis and Kay, 1992; Herv, 1988; Willner et al., 2005;
Chew et al., 2007).
Magmatism continued from the Permian to the Middle
Triassic (280240 Ma), when great volumes of intrusive and
mostly felsic volcanic rocks, including the Choiyoi large igneous province in Chile and Argentina and the Mitu Group in
southern Peru (Kay et al., 1989, Sempere et al., 2002), were
emplaced along the western South American margin. Although
geochronologic and geochemical data are still incomplete,
several competing hypotheses, such as normal or oblique subduction, postcollision extension-driven crustal melting, slab
breakoff or slab shallowing, have been proposed to explain
the prevailing tectonic regime along different segments of the
Andean margin at that time (Mpodozis and Kay, 1992; Kleiman
and Japas, 2009; Ramos and Folguera, 2009, and references
therein). From the Middle Triassic to earliest Jurassic (240
190 Ma), rifting associated with the incipient stages of Pangea
dispersal (e.g., Veevers, 1989), accompanied by a decreasing
volume of bimodal magmatism, seems to have occurred
along the western margin of South America (e.g., Ramos and
Kay, 1991; Franzese and Spalletti, 2001; Rosas et al., 2007).
0361-0128/98/000/000-00 $6.00

Diverse, yet basically subeconomic, porphyry copper deposits


formed during these events in northern Chile and along the
Frontal Cordillera in west-central Argentina (Sillitoe, 1977;
Sillitoe and Perell, 2005; Cornejo et al., 2006; Munizaga et
al., 2008).
Jurassic to Early Eocene Tectonics and
Metallogeny of the Central Andes
After the Triassic rifting event, subduction was reestablished in northern Chile and southern Peru during the Early
Jurassic when a new magmatic arc developed west of the
extinct late Paleozoic arc front. Since then, subduction has
proceeded uninterrupted to date. Initial Jurassic to Early
Cretaceous arc magmatism occurred under extensional conditions that permitted the formation of a series of interconnected back-arc basins to the east of the main arc, which were
progressively filled with marine and continental sedimentary
strata (Mpodozis and Ramos, 1989, 2008). Transpressional
deformation along the arc axis created the intra-arc Atacama
fault system in northern Chile (Scheuber and Gonzlez,
1999) and was accompanied in the Early Cretaceous by the
emplacement, in northern Chile, of some porphyry copper
deposits at ~140 to 130 Ma (e.g., Antucoya-Buey Muerto,
141139 Ma; Puntillas-Galenosa, 135132 Ma; Perell et al.,
2003b; Maksaev et al., 2006, 2010). Fast convergence rates
during the global mid Cretaceous superplume event (Larson,
1991) produced an upsurge in volcanism along the Andean
margin, accompanied by intra-arc extension and transtension
which fostered iron oxide-copper-gold (IOCG)type mineralization between 120 and 100 Ma in northern Chile and southern Peru (Marschik and Fontbot, 2001; Sillitoe, 2003; Sillitoe and Perell, 2005; Chen et al., 2010). Small, low-grade,
gold-rich porphyry copper deposits such as Andacollo (104
Ma), Domeyko-Dos Amigos (108104 Ma), and Pajonales (97
Ma) were emplaced under extensional conditions during the
same general period in north-central Chile (Sillitoe and
Perell 2005; Maksaev et al., 2010).
The extensional and transtensional conditions that dominated early Andean subduction ended in the early Late Cretaceous, when the back-arc basins were tectonically inverted
(Mpodozis and Ramos, 1989; Tomlinson et al., 2001a). The
shift to a more contractional, subduction-related regime occurred together with the accelerated westward drift of South
America, in response to the final opening of the Atlantic
(Russo and Silver, 1996; Somoza and Zaffarana, 2008). Subsequently, the coastal magmatic arc was abandoned (Fig. 1b)
and the magmatic front jumped to the east in the Late Cretaceous, where it remained relatively stationary until the early
Eocene. Abrupt shifts in the magmatic front such as this has
been accompanied throughout the Andean history, by transient geochemical changes during and after arc migration
(e.g., Cornejo and Matthews, 2001; Haschke et al., 2006: Fig.
1d). Stern (1991, 2011), Kay and Mpodozis (2002), and Kay et
al. (2005) suggested that changes of this type reflect mantle
contamination from fore-arc crust removed during enhanced
subduction erosion processes associated with major contractional events along the Andean margin.
In northern Chile, the Cretaceous-Tertiary boundary was
marked by another short pulse of contractional deformation,
the K-T event of Cornejo et al. (2003). Paleocene to early

330

331

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES


0

15

WC

EC

SA

20

Modern CVZ Arc

La Paz

40

PERU
Al

tip
lan
o

80
SA

CHILE

100

Salar de
Atacama

EC

Mid - Cretaceous arc

120

Antofagasta
Puna

25

140

EN
TI
NA

FA

Late CretaceousEocene arc

60

20

AGE (Ma)

Arica

Jurassic - Early Cretaceous


arcs

160

Distal tuffs/reset ages

SP
70

AR

WC

65

180

Longitude ( W )

(a)

(b)

200

71

70

69

68

67

66

65

Post Incaic intrusions

10

Incaic Event

Sm/Yb

20
26 Ma
30
AGE (Ma)

40

Late Cretaceous
Peruvian Event

Porphyry Cu
deposits
Volcanic and
intrusive rocks
Ignimbrites

50
60

72

ANTOFAGASTA
TRANSEC T
(20 - 23S)

71

70

Eoc-E Olig.
Incaic intrusions

KT
Event

(c)

Longitude ( W )
69

68

67

66

65

(d)
90

64

80

70

60

50

40

30

20

10

AGE (Ma)

FIG. 1. (a). Main morphotectonic units of the central Andes, between 15 and 30 S. FA = modern fore-arc zone, including the Coastal Range and, farther to the east, the Precordillera (or Cordillera de Domeyko) shown in Figure 2; WC =
Western Cordillera which, north of 27 S, is essentially formed by the active magmatic arc of the Central Andean volcanic
zone (CVZ); EC = Eastern Cordillera; SP = Sierras Pampeanas; SA = sub-Andean fold-and-thrust belt. (b). Relationship between age and longitude (distance to the trench) for <200 Ma volcanic and intrusive rocks of the Central Andes (1928 S;
Haschke et al., 2002, CVZ = modern Central volcanic zone of the Andes). (c). Longitude vs. age for Cenozoic intrusive and
volcanic rocks at 20 to 23 S (Iquique to Antofagasta transect; Trumbull et al., 2006). The ~32 to 26 Ma gap may be related
to a transient episode of flat subduction produced as a consequence of the Incaic tectonic episode. The east to west migration of the magmatic front during the Miocene is considered to record late-stage slab steepening (Kay et al., 1999; Kay and
Coira, 2009). (d). Geochemical changes since the Late Cretaceous near 26 S (Copiap-El Salvador region; data from
Cornejo and Mathews, 2001). Note the short-lived peaks in the Sm/Yb ratio during major tectonic events, superimposed over
a more subdued long-term trend to increased values. Peak values may reflect contamination of the mantle magma source regions as a result of massive removal of fore-arc continental crust during periods of enhanced subduction erosion (Kay and
Mpodozis, 2002; Kay et al., 2011).

Eocene magmatism included the emplacement of a porphyry


copper belt between southern Peru (Cerro Verde, 61 Ma;
Quellaveco, 54 Ma; Cuajone, 52 Ma) and Cerro Colorado (52
Ma), Spence (57 Ma), and Relincho (61 Ma) in northern Chile
(Sillitoe and Perell, 2005). Structural and geochemical data
gathered by the authors suggest that these deposits in northern Chile formed in a neutral stress to mildly extensional arc
0361-0128/98/000/000-00 $6.00

built on a relatively thin crust over a steep subduction zone.


This is consistent with the very low rates of and oblique convergence between the South American and Farallon plates
(Pardo-Casas and Molnar, 1987), and also with within-plate
geochemical signatures of Paleocene to early Eocene rocks in
the Inca de Oro-El Salvador region (Cornejo and Matthews,
2001).

331

332

MPODOZIS AND CORNEJO

Middle Eocene to Early Oligocene Tectonics


of the Central Andes
The Domeyko fault system
One of the most relevant tectonic and metallogenic
episodes in the Central Andes correlates with the middle
Eocene to early Oligocene (4533 Ma) Incaic tectonic event
(Noble et al., 1979; Maksaev and Zentilli, 1988; Mpodozis and
Perell, 2003; Sillitoe and Perell, 2005); during this period,
bending of the continental margin generated the Bolivian
orocline and the Domeyko fault system along the Precordillera of northern Chile. The Domeyko fault system (Fig.
2) is a >1,000-km-long, 40- to 60-km-wide, orogen-parallel
zone of deformation composed of a complex array of strikeslip, normal, and reverse faults, together with thin- and thickskinned folds and thrusts, which extends along the Cordillera
de Domeyko (also known as Precordillera) in northern Chile
between 20 and 27 S (e.g., Reutter et al., 1991, 1996;
Cornejo et al., 1997). Some authors (e.g., Amilibia and
Skarmeta, 2003; Amilibia et al., 2008) proposed that most of
these faults and folds initiated during Late Cretaceous as a
consequence of the inversion of normal faults inherited from
the Mesozoic back-arc extension. However, others (e.g., Tomlinson et al., 2001a; Mpodozis et al., 2005) interpreted that
the Andean back-arc basins were first inverted during the
early Late Cretaceous to form a proto-Cordillera de
Domeyko, while a second main tectonic pulse along the
Domeyko fault system, coincident with the Incaic event, produced its final uplift (Reutter et al., 1991, 1996; Scheuber and
Reutter 1992; Tomlinson et al., 1993; Maksaev and Zentilli,
1999). Parts of the Domeyko fault system were subsequently
reactivated during the Oligocene and the Quaternary (Tomlinson and Blanco, 1997a, b; Audin et al., 2003; Soto et al.,
2005).
The kinematics of the middle Eocene to early Oligocene
deformation along the Domeyko fault system is a matter of
controversy; evidence for both left- and right-lateral displacements, including reversal in the sense of shear, has been reported along different parts of the faulted domain (Reutter et
al., 1996; Dilles et al., 1997; Tomlinson and Blanco 1997a, b;
Hoffman-Rothe et al., 2004; Niemeyer and Urrutia, 2009).
Fission-track age data show that the Cordillera de Domeyko
was exhumed between 40 and 30 m.y. ago (Maksaev and
Zentilli, 1999; Nalpas et al., 2005) in association with surface
tectonic uplift and profound erosion, the products of which
accumulated in syntectonic basins east and west of the area of
deformation (Mpodozis et al., 2005; Hong et al., 2007; Wotzlaw et al., 2011).
Origin of the Domeyko fault system
At first glance, the Domeyko fault system could be considered as a trench-linked fault system (Woodcock, 1986) that
nucleated in the thermally weakened crust of the middle
Eocene to early Oligocene magmatic arc of northern Chile
during a period of suggested fast Eocene oblique convergence between the Farallon and South America plates (PardoCasas and Molnar, 1987; Somoza, 1998). However, when recent paleomagnetic and structural studies are taken into
account it becomes apparent that tectonic activation of the
Domeyko fault system during the Incaic episode is essentially
0361-0128/98/000/000-00 $6.00

a consequence of the formation of the sharp bend of the western South American margin, known as the Arica elbow or Bolivian orocline (Fig. 3). Paleomagnetic studies have been essential in obtaining a more constrained view of the
deformational history of this segment of the Central Andes
and support the tectonic model for orocline formation first
proposed by Isacks (1988).
Figure 3a is a simplified regional map showing the distribution of paleomagnetic (declination) vectors measured for the
Central Andes. Importantly, independent of age, Mesozoic
and Paleogene rocks have been rotated up to 50. In contrast,
rotations measured in Miocene and younger rocks (<1811
Ma in northern Chile; <20 Ma in southern Peru) are negligible, suggesting that most of the rotations were acquired during a single Paleogene episode of deformation (Roperch et al.,
2006, 2011; Arriagada et al., 2008, and references therein).
Rotations are counterclockwise in southern Peru (Domain B;
Fig. 3a), clockwise in northern Chile south of Antofagasta
(domain D), and almost nonexistent in the intermediate region (domain C) between Antofagasta and Arica (Taylor et al.,
2005; Arriagada et al., 2008). The magnitude of the counterclockwise rotations decreases significantly at the Abancay Deflection in central Peru (domain A; Fig. 3a), the latter break
being interpreted as a zone of intense Eocene-early Oligocene
left-lateral shear along the boundary between the Arequipa
and Paracas basement terranes (Ramos, 2009; Roperch et al.,
2011; Fig. 4b). In northern Chile, clockwise rotations decrease progressively south of Antofagasta and essentially disappear near Vallenar (2830' S) upon entering the essentially
nonrotated domain E (Fig. 3a), which extends southward to
the latitude of Santiago (33 S).
In his landmark paper, Isacks (1988) proposed that the observed paleomagnetic rotations and the formation of the seaward concave Bolivian orocline are related to along-strike
variations in the amount of late Cenozoic shortening produced during contractional deformation focused along a
mechanically and thermally weakened zone located in the
overriding South American plate. Recent structural studies
indicate, however, that the bulk of the shortening (~60%), at
and near the Arica bend (1319 S; Fig. 3), is pre-Neogene
in age and occurred between 40 and 20 Ma (Lamb, 2001;
Mller et al., 2002; Kley et al., 2005; McQuarrie, 2006). Incaic shortening concentrated within the Eastern Cordillera of
Bolivia where >12 km of terrigenous sedimentary strata accumulated in a Paleozoic marine basin above highly attenuated
continental crust (Figs. 1, 3). The amount of horizontal shortening reaches a maximum near the axis of the orocline, where
the Paleozoic sedimentary sequence is thickest, and decreases
symmetrically along-strike (Oncken et al., 2006; Gotberg et
al., 2010, and references therein, Fig. 4a) as the Paleozoic
sedimentary rocks of the Eastern Cordillera become thinner
and give way to metamorphic and crystalline rocks in both
Peru (Cordillera de Maran) and northwest Argentina (Sierras Pampeanas, see Fig. 3).
Arriagada et al. (2008) attempted to remove the combined
effects of accumulated horizontal shortening and block rotations. Figure 3b shows their preferred solution for the restored
shape of the continental margin (Peru-Chile trench) at 45
Ma, before the formation of the Bolivian orocline. As shown
in Figure 4b, the ca. 30 E azimuth of the Farallon-South

332

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

7030'

La Planada

YP

6730'

N
Copaquire
Rosario (Collahuasi) (36-35)
Ujina (36-35)
Quebrada Blanca (37-35)

QC

21

21

Salar de Ascotn

El Abra (38-37)
Conchi
Radomiro Tomic (36-34)

CA

Chuquicamata (36-32)
Alejandro Hales (39-36)
Miranda (38-37)

Toki (39)
Opache (38-37)

2230'

IV
OL

IA

2230'

CALAMA

EsperanzaTelgrafo (42-40)
Caracoles (42-41)

Gaby (42)

CE
Salar
de
Atacama

Centinela (45-44)
Polo Sur (42-41)

ANTOFAGASTA

Chimborazo (41)
Zaldvar (38-37)
Escondida (38-37)

24

LE

24

Escondida Este (36-35)

Sierra Juncal (40)

2530'

SE

PS

2530'

Exploradora (35)
Sierra del Jardn (42)

El Salvador (42-41)

AR

GE

NT

IN

Salar de Punta
Negra

Salar de
Pedernales

Potrerillos
(36)

Middle Eocene to early Oligocene porphyry Cu + Mo+ Au deposits


Eocene plutons (48-38 Ma)
Paleozoic basement
Reverse fault
Normal or strike-slip fault

Salar de
Maricunga

27

COPIAPO

Salars

27

100km

RF
7030'

69

6730'

FIG. 2. Sketch map of the Cordillera de Domeyko (or Precordillera) and the Domeyko fault system, showing main faults
traces, exposures of Paleozoic basement, clusters of Eocene plutonic rocks (YP = Yabricoya-La Planada intrusive cluster;
QBC = Quebrada Blanca-Collahuasi; CA = Chuquicamata-El Abra; CE = Centinela, LE = La Escondida; SE = Sierra Exploradora-Juncal; PS = Potrerillos- El Salvador; RF = Ro Figueroa) and the locations and ages (in parentheses, Ma) of
Eocene-early Oligocene porphyry copper deposits. More detailed maps of CA, LE, and CE clusters are shown in Figures 5,
7, and 8. Based on the 1:1,000,000 geologic map of Chile (SERNAGEOMIN, 2002).
0361-0128/98/000/000-00 $6.00

333

333

10

Pe

0361-0128/98/000/000-00 $6.00

334

h
nc
Tre

(a)

Sierras
Pampeanas

26

250 km

Total displacement vectors


(45-0 Ma)

70

66

FIG. 3. (a). Paleomagnetic declination data for the Central Andes in northern Chile and southern Peru that indicate the Central Andean rotation pattern. Data from
Arriagada et al. (2006, 2008), Roperch et al. (2006, 2011), Taylor et al. (2005), and references cited therein. Distribution of lower Paleozoic sediments of the Bolivian
basin and crystalline and metamorphic rocks modified from Ramos and Dalla Salda (2011; Fig. 1). Also shown are the main Eocene-Oligocene porphyry copper deposits
of the Central Andean belt (Perell et al., 2003a). Rotation domains designated A to E are discussed in the text. (b). Position of the plate boundary (Peru-Chile trench)
at 45 Ma and total (Paleogene to present) displacement vectors for material points on the South American margin according to the two-dimensional restoration model
of Arriagada et al. (2008).

El Morro

22

74
62

(b)

AN
LI
ZI LD
A
IE
BR SH

250 km

Paleomagnetic declination vector

Copiap

INA

NT

GE

AR

Jujuy

Middle-Eocene to early Oligocene


porphyry copper belt

La Escondida

18

Lower Paleozoic (Neoproterozoic?)


metamorphic and intrusive rocks

Lower Paleozoic sedimentary rocks

Antofagasta

Chuquicamata

Collahuasi

Santa Cruz

14

10

78

f
no
o
i
sit a
po 45 M
d
re at
sto nch
e
R tre

D Paleomagnetic domain

22

hil

C
u-

Arica

AN
LI
ZI LD
A
E
I
BR SH

18

La Paz

VI

LI

BO

62

ONLas Bambas

66

Pe

FL
Y DE
NCA

ECTI

70

hil

C
u-

ABA

Lima

PERU

74

Cordillera de
Maran

nc

Tre

14

78

334
MPODOZIS AND CORNEJO

335

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

(a)

10

15

20

25

30

35

(b)
S
CA E
RA AN
PA E RR
T

Latitude (S)
400

st

aT

re

nc

200

3a
100

Initial (non rotated)


Domeyko fault
system

1500

1000

500

500

1000

Mechanically
weak crust
(Paleozoic
sedimentary
rocks

al
on
og ence
h
t
Or verg
n
co

3b

Distance from orocline axis

Be lt

300

Fo
ld

A
TE RE
R R QU
AN IPA
E

ru
Th

Brazilian
Shield

d
an

l
na e
go
tho rgenc
r
O ve
n
co

h
nc
Tre
Ma
45

Tectonic shortening (km)

ng
alo
ne ndary
o
z
u
ear bo
Sh rane
terr

1500 km

21S

Domeyko
fault system

e
liqu ce
Ob ergen
v
n
co

FIG. 4. (a). Different horizontal tectonic shortening estimates for the Central Andes, showing how values decrease symmetrically north and south of the orocline axis. 1 = Neogene deformation in the Subandean belt (Oncken et al., 2006); 2 =
total shortening (Peloegene + Neogene, Oncken et al., 2006); 3a = shortening needed to accommodate crustal area assuming initial crustal thickness of 40 km; 3b = same, but considering 35 km as initial thickness (Gotberg et al., 2010); 4 = shortening needed to balance paleomagnetically determined rotations (Arriagada et al., 2008). (b). Tectonic sketch showing how
oblique convergence along the southern limb of the orocline may drive strike-slip displacements along the Domeyko fault
system at the beginning of the Incaic event. The general scheme comes from figures taken from Isacks (1988) and Lamb
(2001). Reverse faults prevailed north of 21 S as a consequence of the original N-NW trend of the continental margin. Note
left-lateral shear along the Abancay Deflection along the boundary between the Arequipa and Paracas terranes.

America Eocene to Oligocene convergence vector calculated


by Somoza (1998, Fig. 4b) was nearly orthogonal to the (restored) NW-SEtrending section of the margin along the
center and northern limb of the Bolivian orocline. Nevertheless plate convergence was, at the same time, highly oblique
along the southern limb of the orocline fostering marginparallel shear, and (theoretically dextral) strike-slip faulting in
northern Chile where the Domeyko fault system was formed
(Fig. 4b). Bending of the margin seems to have been accompanied at the southern limb of the Bolivian orocline by wholesale NE-directed crustal flow (Fig. 3b), which is also required
to explain the excess orogenic volume and crustal thickness
below the Altiplano-Puna reported by Kley and Monaldi
(1998) and Hindle et al. (2005).
Mass transfer, likely associated with lower crustal flow toward the center of the orocline (Hindle et al., 2005), increases
the possibility of strike-slip faulting along the Domeyko fault
system. Continued displacement was likely initially blocked
near the orocline axis where deformation was dominated by
margin-normal contraction (Figs. 3b, 4b). As suggested by
McQuarrie (2002) and Boutelier and Oncken (2010), mass
transfer toward the core of the orocline implies crustal thinning and stretching in central Chile, which may have inhibited mountain building south of 28 S. Crustal thinning and a
more extensional tectonic regime prevailed in that region
(south of 28 S) during the Eocene to early Miocene (e.g.,
Jordan et al., 2001; Charrier et al., 2002). These contrasts
help to explain the termination of the Incaic porphyry copper
0361-0128/98/000/000-00 $6.00

belt at approximately 30 S and the possibly segmented nature of the belt along its strike length between southern Peru
and central Chile (Mpodozis and Perell, 2003; see below).
Middle Eocene to Early Oligocene Porphyry Copper
Province of Northern Chile
Overview
Middle Eocene to early Oligocene porphyry copper deposits of the Central Andes were emplaced contemporaneously with the Incaic tectonic event, when the entire Andean
margin was being reshaped during the formation of the Arica
bend. Mineralized centers occur in Peru near the eastern end
of the Abancay Deflection (Andahuaylas-Yauri cluster; Perell
et al., 2003a; Fig. 3a), although the vast majority are located
in Chile along the Domeyko fault system (Sillitoe and Perell,
2005; Figs. 2, 3a). El Morro, the southernmost porphyry
copper-gold deposit of economic importance (Perell et al.,
1996), is located where the rotated and thickened southern
limb of the Bolivian orocline (domain D; Fig. 3a) terminates
and merges with the nonrotated central Chile domain E (see
below). Most porphyry copper deposits were emplaced along
the Domeyko fault system at long-lived zones of focused magmatism, which in some cases were active well before the
Eocene. They occur as parts of discrete intrusive clusters
separated by large barren areas where only the Paleozoic
basement and back-arc basin sedimentary cover is exposed
(Fig. 2).

335

336

MPODOZIS AND CORNEJO

The structural control of the porphyry systems is strikingly


diverse (cf. Sillitoe and Perell, 2005). At Chuquicamata, porphyry copper-related intrusions were syntectonically emplaced
along an extensional jog in an active dextral strike-slip fault
system (Lindsay et al., 1995); at Potrerillos, intrusions related
to porphyry copper deposits ascended along subvertical, leftlateral strike-slip faults and migrated laterally near the surface
along an active low-angle thrust (Tomlinson et al., 1993;
Niemeyer and Munizaga, 2008). In contrast, porphyry stocks
related to mineralization at Escondida or El Abra-Fortuna
were emplaced along premineralization reverse and strike-slip
faults formed during earlier stages of the Incaic event (Dilles
et al., 2011; Herv et al., 2012). Other mineralized intrusions
were cut and displaced by late-stage faults (Esperanza-Telgrafo, Chuquicamata; Dilles et al., 1997; Tomlinson et al.,
1997a, b).
Not all porphyry copper deposits are, however, obviously
related to major Domeyko fault splays (Sillitoe and Perell,
2005), although the deposits of the Collahuasi-Quebrada
Blanca cluster, which crop out to the east of the main fault
system (Fig. 2), seem to have been emplaced over a basement
discontinuity marked by regional isotopic changes in Neogene magmatic rocks (Mamani et al., 2010). At El Salvador
and Polo Sur, 42 to 41 Ma porphyry copper-related intrusions
are encapsulated in preexisting, up to 10-m.y. older, calderarelated rhyolite domes, indicating that the middle Eocene to
early Oligocene magmas effectively reused the same conduits
(Cornejo et al., 1997).
These differences in the structural controls of porphyry
copper deposits can be attributed to the changing tectonic
conditions along different segments of the Domeyko fault
system during the >10-m.y. Incaic event. Camus (2003) recognized three temporal groupings of mineralized intrusions
along the Domeyko fault system at 43 to 42, 39 to 37, and 36
to 33 Ma.
In the Chuquicamata, El Abra, and Quebrada Blanca-Collahuasi districts, where the three generations of intrusions
occur, only the two youngest were fertile in copper (Campbell
et al., 2006; Maksaev et al., 2009; Dilles et al., 2011), whereas
in the Escondida district all three events are related to
mineralization (Herv et al., 2012). In contrast, most of the
copper-bearing intrusions recognized in the Centinela district
seem to have been emplaced during the early event. In order
to illustrate the changing nature of the diverse porphyry clusters and explore the links between mineralized centers and
the evolution of the Domeyko fault system, the salient geologic features of three porphyry copper districts, Chuquicamata-El Abra, Escondida, and Centinela (Figs. 2, 5), are examined below.
Chuquicamata-El Abra
The Chuquicamata-El Abra intrusive cluster (Fig. 5a) is situated along the northern Domeyko fault system, near the
boundary between paleomagnetic domains C and D (Fig. 3a).
The geology of the region records superimposed tectonic and
magmatic events beginning in the Mesozoic. In this region,
the Precordillera (Cordillera de Domeyko) is formed by two
fault-bounded, N-Strending basement ranges. The western
range, Sierra de Moreno, is a thick-skinned block uplifted in
the early Late Cretaceous (~8584 Ma), during the inversion
0361-0128/98/000/000-00 $6.00

of the Mesozoic back-arc basin (Ladino et al., 1997; Tomlinson et al., 2001a). It is bounded to the west by high-angle,
west-vergent reverse faults that place Neoproterozoic to Paleozoic basement units on top of Jurassic sedimentary strata
(Fig. 5a). The east-vergent Arca reverse fault that bounds the
eastern side of the Sierra de Moreno (Fig. 5a) has been interpreted as a reactivated normal fault, which formed near
the eastern edge of the Mesozoic back-arc basin (Tomlinson
et al., 2001a; Fig. 5a). Syn- to post-tectonic red beds (Tolar
and Tambillos Formations; Fig. 5a) were shed to the east and
west of the uplifting block during the Late Cretaceous. After
deformation, volcanic rocks, which unconformably cover the
basement units and Mesozoic sedimentary strata at Sierra de
Moreno, developed in two separate episodes during the latest
Cretaceous (Cerro Empexa and Quebrada Mala Formations)
and early to middle Eocene (Icanche Formation; Fig. 5a;
Tomlinson et al., 2001a, 2010).
The eastern, Sierra del Medio basement block (Fig. 5a) was
uplifted between 43 and 38 Ma during the Incaic event (Tomlinson et al., 1997a) along a new set of west- and east-vergent,
high-angle reverse faults. As in other regions of northern
Chile, volcanism in the Chuquicamata-El Abra region sharply
diminished at this time; however, a syntectonic sedimentary
sequence (Sichal Formation; Fig. 5a) that accumulated in a
narrow basin between Sierra de Moreno and Sierra del
Medio contains a few intercalations of tuffs and volcanic breccias that yield K-Ar and Ar/Ar ages between 43 and 36 Ma
(Tomlinson et al., 2001a). The Incaic deformation also includes
a strike-slip component that produced segmented N-NE to
NE-trending, dextral strike-slip faults (e.g., the Mesabi fault
in Fig. 5a; Tomlinson et al., 1997a). These faults became
more important in the southern part of the area near
Chuquicamata (Fig. 5a), in accordance with increasing plate
convergence obliquity to the south along the Andean margin
during the Eocene (Fig. 4b).
As shown in Figure 5, the region between Chuquicamata
and El Abra is one of the anomalous zones along the Cordillera de Domeyko (Fig. 2), where magmatism was recurrent
since the Late Cretaceous. Despite lacking evidence for large
volumes of middle Eocene to early Oligocene volcanic products, intrusive magmatism of this age is well recorded. A large
(>500 km2), composite intrusive complex (Fortuna-El Abra)
was emplaced syntectonically between 45 and 38 Ma near the
southern termination of the Sierra de Moreno (Tomlinson et
al., 2001a, 2010; Dillles et al., 2011). Figure 5b shows the restored shape of the Fortuna-El Abra batholith before being
severely dismembered by left-lateral displacements along the
West fault (Tomlinson et al., 2007a; Dilles et al., 2011). Field
relationships indicate that the batholith was intruded along
the trace of the Quetena reverse fault, an inverted normal
fault inherited from Mesozoic back-arc extension. A flattening
foliation in metaclastic rocks in the contact aureole is consistent with emplacement during regional Incaic E-W shortening (Tomlinson and Blanco, 1997a).
The Fortuna-El Abra batholith, described by Dilles et al.
(1997) as a porphyry copper batholith, is a long-lived, composite magmatic system that contains intrusive phases emplaced during different stages of the Incaic event; it is similar
to the Andahuaylas-Yauri batholith of southern Peru, described by Perell et al. (2003a). The batholith comprises an

336

337

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

(a)
Cretaceous reverse fault

69

Los Picos
Complex
(45-42 Ma)

Cretaceous reverse fault


with Eocene reactivation

Lower to Middle
Eocene volcanic
rocks
(Icanche fm)

Pajonal Diorite

(b)

+
+ +
+
+ ++

+ + +
+ +

Upper Cretaceous
sedimentary rocks
(Tolar fm)

+
+ + + +
+ +
+

+ +
El Abra
+
+ + +
+
+
+
+
+
+
+
+
+ +
+ + + + + + + + + + + + +
+
Conchi Viejo
+ + + + + + +
+ + +
+
+ + +
+
+ +
+ + + + +
+ + + + +
+ + + +
2200'
+ +
+
+ + + + + ++
Clara Granodiorite
+ +
+
+ + + +
+

+ +

+ + +

Eocene reverse fault


related to EW shortening
DEL M
EDIO

10

20km

San Lorenzo
porphyry
dikes

DE
SIE
RRA

+
+

2130'
Antena
Granodiorite
(39.5-39 Ma)

SIERR
A

MO
R

ENO

Small copper
prospects

Arca fault

+
+
+
+
+

+ + + +
+ + + +
+ + + +

+ +

Early Abra-Antena
Ganodiorite
Upper Cretaceous volcanic rocks
(Quebrada Mala fm.)
Alejandro Hales

Toki

Miranda

Co. Jaspe

Quetena
fault

Opache

Porphyry Cu deposits
under post mineral cover
(Toki Cluster)
0
Future
West Fault

10km

Chuquicamata Porphyry Complex (36-33 Ma)


Eocene (45-37 Ma) monzodiorites to granodiorites (including
Fortuna El Abra batholith)
Upper Eocene-Oligocene, syntectonic, sedimentary rocks (Sichal
and Calama formations)
Lower to Middle Eocene volcanic rocks (Icanche formation)

El Abra
22

+ +

Paleocene intrusive rocks


Upper Cretaceous volcanosedimentary sequences (Cerro Empexa
and Quebrada Mala formation)
Cretaceous intrusive rocks
Upper Cretaceous, syntectonic, red beds (Tambillos and Tolar
formations)
Jurassic to Lower Cretaceous, back-arc, sedimentary sequences

R. Tomic
Mesabi Fault

Chuquicamata

+
+ +
+ + +
+ +
+

Llareta Granodiorite

+
+ +
+ + +
+ +

Fiesta Granodiorite
(38-37.5 Ma)

WEST FAULT

Quetena
Genoveva

+ +

Triassic intrusive rocks (ca. 230 Ma; Elena granodiorite)

Alejandro Hales

Triassic volcanic and sedimentary rocks

Toki

Quetena Fault

Sierra del Medio basement block (Upper Paleozoic)

CALAMA

69

Sierra de Moreno basement block (Neoproterozoic to Paleozoic)

2230'

FIG. 5. (a). Simplified geologic map of the El Abra-Chuquicamata region, indicating age of faults (based on Tomlinson et
al., 2001). (b). Restored map of the El Abra-Fortuna batholith after removal of 35 km of Oligocene-early Miocene left-lateral motion on the West fault, showing main intrusive phases and mineralized centers (Dilles et al., 2011). Names of intrusive units east of the future trace of the West fault follow the nomenclature that has been employed for intrusive phases near
El Abra.

older group of intrusions, namely the Los Picos complex and


Pajonal diorite, emplaced between ~45 and 42 Ma (Tomlinson et al., 2001a; Campbell et al., 2006), which include predominantly mafic pyroxene- and biotite-bearing quartz monzodiorites, monzodiorites, and quartz monzonites that are
copper barren (Fig. 5b). The second intrusive group includes
the Fortuna-El Abra granodiorite complex and is largely composed of two hornblende-bearing granodioritic phases (Antena, 39.539.0 Ma and Fiesta, 3837.5 Ma) as well as porphyritic components (San Lorenzo porphyries; Fig. 5b). The
Fortuna-El Abra Complex intrusions, derived from hydrous,
oxidized, and sulfur-rich magmas (Dilles et al., 2011), are associated with large porphyry copper deposits with U-Pb ages
between 39 and 36 Ma (El Abra, 3837 Ma; the Toki cluster
including Toki, 39 Ma and Opache, 3837 Ma; Alejandro
Hales, 3936 Ma; Conchi, 36 Ma; Perell, 2003; Campbell et
0361-0128/98/000/000-00 $6.00

al., 2006; Boric et al., 2009; Marquardt et al., 2009; Barra,


2011; Fig. 5b).
The youngest intrusive event in the Chuquicamata cluster
was related to the emplacement, beyond the southern limits of
the Fortuna-El Abra batholith, of the Chuquicamata (3632
Ma) and Radomiro Tomic (3634 Ma) porphyry copper deposits (Lindsay et al., 1995; Reutter et al., 1996; Ballard et al.,
2001; Ossandn et al., 2001; Campbell et al., 2006; Barra, 2011;
Fig. 5). Sibson (1987), who relied on maps of Perry (1952), interpreted the Chuquicamata porphyry copper deposit to have
been emplaced syntectonically, following an extensional
stepover linking two separate but overlapping, parallel, dextral
strike-slip faults. Alternatively, Lindsay et al. (1995) favored a
model in which the required space for magma emplacement
coincided with a releasing bend (cf. Cunningham and Mann,
2007) along a single, continuously linked dextral fault.

337

338

MPODOZIS AND CORNEJO

After porphyry emplacement at the Chuquicamata and


Radomiro Tomic, to the south of their present location, the
deposits were displaced as a consequence of late Oligocene to
early Miocene (31.016.3 Ma) left-lateral movements along
the throughgoing regional West fault or West Fissure. Displacement along the West fault, which includes, at Chuquicamata, according to McInnes et al. (1999), a 600 100 m west
sideup vertical displacement component (contested by Tomlinson et al., 2001b), was able to offset the Fortuna-El Abra
batholith 35 to 37 km in a left-lateral sense (Dilles et al., 1997;
Tomlinson and Blanco, 1997b; Tomlinson at al., 2001a; Fig.
5). South of Chuquicamata, at least part of the strike-slip displacement component was transferred to a set of normal
faults that bound the Cenozoic Calama extensional and/or
transtensional basin (Blanco, 2008; Blanco and Tomlinson,
2009). Seismic data reveal that a buried normal fault with 1to 1.5-km down-to-the-east displacement limits the northwestern margin of the basin, where up to 2,500 m of siliciclastic continental strata accumulated between the Oligocene
and Miocene (Jordan et al., 2004; Blanco, 2008).
The West fault forms, at present, the sharp western limit of
the Chuquicamata orebody. Nevertheless, structural interpretations by Sibson (1987) and Lindsay et al. (1995) considered that the internal architecture of second-order faults and
veins indicates that the deposit is essentially intact. According
to Lindsay et al. (1995), the Oligocene to early Miocene West
fault propagated northward, along the western edge of the
porphyry complex (previously emplaced at 3633 Ma), without displacing the mineralized intrusive units within the former dextral releasing bend and/or stepover.
Escondida
Despite also being along the Domeyko fault system, the
tectonic history of the Escondida region (Fig. 2) is markedly
different, highlighting the significant changes in tectonic
styles along the >1,000-km strike of the fault system. Deformation in this area seems to have been dominated by tectonic
escape linked to passive rotation and transport of brittle
upper crustal blocks over hot and ductile lower crust in a way
similar to the so-called orogenic float or clutch tectonics models discussed by Oldow et al. (1990), Lamb (1994), and Tikoff
et al. (2002; see below). At this latitude, the Cordillera de
Domeyko appears as a discontinuous mountain range formed
by a group of discrete basement blocks bounded to the west
by a 150-km-long shear lens (Escondida shear lens) developed between the regional Sierra de Varas and Escondida
faults (Figs. 6a, 7). To the east, the Domeyko range abuts the
Salar de Atacama depression, which is a deep subsiding basin
filled by >9 km of Cretaceous to Tertiary continental sedimentary strata (Pananont et al., 2004; Mpodozis et al., 2005).
The basin was built on top of a large positive gravimetric
anomaly (Central Andean Gravity High; Gtze and Krausse,
2002; Fig. 6b), which indicates the occurrence, at depth, of
dense crustal rocks that may help to explain its long-lived subsidence basin history, recorded at least since the Cretaceous.
The isolated basement blocks that form the core of the
range are separated by small, triangular basins, with interior
drainage (Fig. 6a). The southern rhomboid-shaped blocks
(e.g., San Carlos and Imilac; Fig. 6a) are bounded along their
northwestern margins by high-angle, SE-dipping reverse faults.
0361-0128/98/000/000-00 $6.00

The blocks at Quimal, Los Morros, and Mariposas are limited to the west and north by left-lateral strike-slip faults
(Mpodozis et al., 1993a, b). Along the El Bordo Escarpment,
the eastern margin of the Imilac and Mariposas blocks are
thrust over the sedimentary fill of the Salar de Atacama basin
(Fig. 6a), which includes, among other units, a 2,500-m-thick
sequence of Eocene to early Oligocene continental conglomerates and poorly consolidated gravels. Internal progressive unconformities and Ar/Ar ages between 44 and 43
Ma from a tuffaceous horizon just above the base of this sequence (Loma Amarilla Formation) indicate that these
strata-accumulated syntectonically during the regional Incaic
deformation (Hammerschmidt et al., 1992; Mpodozis et al.,
2005).
The tectonics of this segment of the Cordillera de Domeyko
(Fig. 6a) can be interpreted as a result of the displacement of
a 250- 50-km basement sliver that was transported northward during the Incaic deformation. According to Mpodozis
et al. (1993a, b), the continuous northward shift of the displaced block was impeded by a buttress located to the north
of the moving sliver as the displacement was transferred to
the east by tectonic escape (cf. Mann, 1997) toward the
deeply subsiding Salar de Atacama basin. In this model, the
Salar de Punta Negra depression (Fig. 6c) would have formed
as an extensional basin at the trailing edge of the displaced
block. Displacement transfer seems to have occurred by
clockwise rotation of small detached blocks, which in turn
generated the local triangular-shaped extensional basins between the rotating blocks as well as contractional deformation
in their northeastern corners where basement was thrust over
the Salar de Atacama basin fill (Fig. 6c). Mpodozis et al.
(1993a, b) located this buttress at Sierra de Limn Verde,
which is a N-plunging basement half dome that attains one of
the highest elevations (3,500 m.a.s.l.) in the Cordillera de
Domeyko (Fig. 6a). However, if along-strike changes in local
stresses resulting from the formation of the Bolivian orocline
are considered, the buttressing effect may have been provided by the nonrotated paleomagnetic domain C, located
north of Calama (Fig. 3), where initial Eocene deformation
was taken up by pure east-west shortening (Tomlinson et al.,
2001a; see Figs. 3, 4).
Figure 7 is a more detailed map showing the geologic setting and distribution of the barren and mineralized intrusions
that form part of the Escondida cluster (labeled LE, Fig. 2).
The area encompasses the widest part of the regional Escondida shear lens, which is separated to the east from the Sierra
Imilac and Sierra San Carlos basement blocks by the Escondida fault (the Panadero-Portezuelo fault is an alternative
name used by Herv et al., 2012). These two blocks were then
separated by the intervening triangular Salar de Hamburgo
depression (Fig. 7). Drilling shows that the Salar de Hamburgo fill includes >1,200 m of red beds, lahars, and pyroclastic rocks with U-Pb zircon ages of 38 Ma (San Carlos
Strata, Fig. 7; Marinovic et al., 1995; Urza, 2009; Herv et
al., 2012); therefore, these units are equivalent to the upper
portion of the syntectonic Loma Amarilla Formation in the
Salar de Atacama. The Hamburgo fault is a NE-trending,
high-angle reverse fault that places the late Paleozoic basement of the San Carlos block over the sedimentary sequences
of the Salar de Hamburgo (Fig. 7).

338

339

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

er

Si e
r ra

fault
Toro
s

Fortuna-El Abra
batholith
pa m

El Bo rdo Esca
r

Bo

CALAMA

Thrusts
r'

Salar de
Atacama

Paleozoic
basement
blocks

24

Rotated
Domain D

s f
au l

Salar de
Hamburgo

nc

rra

de

Bar r a

Sie

25

in

tina

Free face
Salar
de
Punta Negra

Boundary
fault

La Escondida shear lens (ESL )


Cordillera de Domeyko clockwise rotated
basement blocks (Q: Quimal, LM: Los Morros,
M: Mariposas, I: Imilac, SC: San Carlos)

A lm
eid
a

as

Salar de
Punta Negra

50 km

Gravity
High

de V ar as fault

24

Sie r r a

n de L
ila
Cordo

ANTOFAGASTA

ESCONDIDA

ESL

Extensional
(rotational)
basins

Salar
de
Atacama

Deep Bas

Salar de los
Morros

SC

Sierra Limn Verde

Argen

Salar
de Elvira

Blanc
a

fault

A'

a
ivi

Transition zone betwen


C and D domains

faul
Escondida

en

Salar
Vernica

Va r as

de

BUTTRESS ZONE

GABY

LM

Sierra

(c)

23

Lo s

Cen
tine
Ce
ntilanefaul
la tFault

68

Non-rotated
Domain C

de

ESPERANZATELEGRAFO

70

(b)

imn V

Sier r a d e l Medio

69

(a)

20 40 km

Cordn de Lila-Sierra de
Almeida stable domain

Mesozoic to Paleocene strata of


the Centinela District
Reverse faults

Strike-Slip faults

Outcrops of Lower Cretaceous


to Upper Oligocene continental
clastic strata of the Salar de
Atacama basin

Normal faults

FIG. 6. (a). Main structural elements of the Cordillera de Domeyko (between Escondida and Sierra Limn Verde and
Salar de Punta Negra (22 3025 S; see location in Fig. 2). Note the large shear lens (Escondida shear lens) flanked by the
Escondida and Sierra de Varas strike-slip faults along the western edge of the range and the discontinuous basement blocks
(labeled with letters) forming the core of the range. (b). Tectonic sketch of the Cordillera de Domeyko between 21 and 25
S, indicating major Eocene-Oligocene Incaic structures (Tomlinson and Blanco, 1997a). Note contrast between clockwiserotated blocks in rotated domain D (Fig. 3) and deformation associated with reverse faults in nonrotated domain C. (c).
Model of lateral transfer of displacement of a tectonic sliver bounded by a buttress and a free face moving northward along
a left-lateral strike-slip fault system. Displacement is impeded, as shown, by a buttress at the leading edge of the block, and
transferred toward the right by means of clockwise block rotations. Note extensional basins created between the rotating
blocks. A-A' and r-r = position of points and lines before and after rotation ( = rotation angle). Adapted from Beck et al.
(1993).

A protracted, >40-m.y. Cretaceous to Tertiary history of


magmatism is recorded in the Escondida region. The oldest
intrusive events produced Late Cretaceous (8171 Ma; Fig.
7) tholeiitic to alkaline pyroxene gabbros and diorites as well
as hornblende-pyroxene monzodiorites and diorites, in addition to early Paleocene (6664 Ma) pyroxene diorites. These
rocks intruded the sedimentary strata of the Mesozoic backarc basin in the Escondida shear lens to the south and west of
Escondida (Fig. 7); Paleocene to early Eocene volcanic rocks
(5953 Ma) are also present in the area (Marinovic et al.,
1995; Richards et al., 2001; Urza, 2009). All of this focused
and recurrent magmatic activity took place east of the Andean
arc front, which during the Late Cretaceous to early Paleocene (8550 Ma) was located farther west, in the Central
depression of the Antofagasta region (Boric et al., 1990).
0361-0128/98/000/000-00 $6.00

Incaic magmatism in the Escondida cluster began in the


middle Eocene (~44 Ma), when left-lateral displacements
along the Escondida fault and differential rotation between
the San Carlos and Imilac blocks created the triangular Salar
de Hamburgo depression (Fig. 7). The oldest intrusions are a
group of 44 to 41 Ma pyroxene-biotite monzodiorites and
pyroxene-hornblende granodiorites, emplaced within the Escondida shear lens to the north of Escondida (Marinovic et al.,
1995; Richards et al., 2001; Urza, 2009). Together, they likely
represent the roof of an underlying, partially eroded pluton
nearly 20 km in diameter (Fig. 7). Most of these rocks, like the
Los Picos complex and Pajonal diorite of the ChuquicamataEl Abra region, are barren, although geochronologic data and
relationships between intrusive phases in the vicinity of the
Chimborazo porphyry copper deposit indicate, according to

339

340

MPODOZIS AND CORNEJO

lt

69

Sie rra

Sie

de

rra

V aras

de
Imi
lac

f au

24

Chimborazo (41)

Escondida Norte (38-37)


Baker (38-37)

Zaldivar (38-37)

Imila

c f au

lt

Pampa Escondida (36-34.5)


Escondida Este
(36-34.5)
t
Salar de
l
u
Hamburgo
fa
o
rg
bu
am
H

Sa S
n ierr
Ca a
rlo
s

La Escondida fault

Pinta Verde
Escondida
(38-37)

Salar de
Punta Negra
Lower Paleocene (66-64 Ma) px diorites

10 km

2430'

Upper Eocene (42?-36 Ma) sedimentary-volcanic


sequence (San Carlos strata)
Upper Eocene (38-35 Ma) dacitic to granodioritic
mineralized porphyry intrusions (Escondida cluster)

Upper Cretaceous (74-70 Ma) rhyolitic ignimbrites


Upper Cretaceous (81-72 Ma) gabbro-diorites and hb-px
diorites
Upper Triassic-Lower Cretaceous sedimentary sequences

Upper Eocene (39-38 Ma) hb dioritic porphyry intrusions


Eocene (44-41 Ma) px-bt monzodiorites and px-hb
granodiorites
Upper Paleocene to Lower Eocene (59-53 Ma) volcanic
sequences

Triassic (240-220 Ma) intrusive rocks


Upper Paleozoic (300-270 Ma) basement

FIG. 7. Simplified geologic map of the area around Escondida, highlighting major regional faults and the different intrusive phases that form part of the Escondida intrusive cluster. Compiled and adapted from Marinovic et al. (1995),
Richards et al. (2001), Urza (2009), Herv et al. (2012), and field data from the authors (px = pyroxene, hb = hornblende,
bt = biotite).

Herv et al. (2012 ), that an early phase of copper mineralization probably occurred at ~41 Ma.
The second episode of Eocene-Oligocene magmatism
began with the emplacement of a closely spaced group of
small intrusions distributed across the Escondida fault (Fig.
7). These more evolved, amphibole-bearing dioritic stocks,
with U-Pb zircon ages of 39 to 38 Ma (Richards et al., 2001;
Urza, 2009), intrude both the late Paleocene to early
Oligocene volcanic rocks of the Escondida shear lens and the
late Paleozoic basement units of the Imilac block (Fig. 7).
Their distribution suggests that they could be apophyses of a
larger pluton at depth that intruded along the Escondida
fault. The slightly younger group of porphyry copper stocks
include a series of multiphase, NE- to N-NEtrending, dikelike intrusions that were emplaced at or near the Escondida
fault at 38 to 37 Ma; these include the deposits at Zaldvar,
Escondida Norte, Escondida, and Pinta Verde, and, farther
away, at Baker (Richards et al., 2001; Urza, 2009; Herv et
al., 2012; Fig. 7).
0361-0128/98/000/000-00 $6.00

The final event of Incaic magmatism in the Escondida cluster was related to the emplacement of the Escondida Este
and Pampa Escondida deposits, immediately to the east of
the Escondida fault (Fig. 7), between 36.0 and 34.5 Ma
(Herv et al., 2012). The mineralized porphyries of the Escondida cluster, with the exception, perhaps, of Chimborazo,
postdate the earlier phase of sinistral faulting and block rotations along this segment of the Cordillera de Domeyko. Deformation seems to have begun at ~42 Ma (age of the base of
the Loma Amarilla Formation; see above), although the lack
of offset on any of the porphyry intrusions across the local
fault strands (Panadero-Portezuelo fault) of the larger Escondida fault indicates that major along-strike fault activity had
ceased by 38 Ma, as shown by the across-fault 3837 Ma porphyry dikes (Herv et al., 2012).
Centinela
Porphyry copper mineralization in the Centinela district (labeled CE, Fig. 2) occurs within a 25-km-wide, fault-bounded

340

341

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

belt of Late Cretaceous to early Eocene volcanic rocks, located between the Paleozoic basement exposures of the
Cordillera de Domeyko to the east and an early Cretaceous
volcanic sequence in the Coastal Range to the west (Fig. 8).
The Centinela district hosts one of the more recently discovered porphyry copper clusters in northern Chile. Although
the occurrence of exotic copper mineralization at El Tesoro
was known for a long time, the full potential of the district
only began to be assessed in the mid-1990s (Perell et al.,
2010, and references therein).
The district records again, a lengthy history (almost 80 m.y.)
of magmatic activity, from the Early Cretaceous to the Eocene.
The oldest plutonic rocks emplaced within the confines of the
Centinela cluster comprise a group of Early Cretaceous
olivine-pyroxene gabbros and hornblende-bearing quartz diorites, with U-Pb zircon and K-Ar ages between 124 and 100
Ma (Mpodozis et al., 1993b; Marinovic and Garca, 1999).

These rocks, emplaced within Jurassic marine limestones of


the northern Chile back-arc basin (Fig. 8), are almost 100 km
east of the Early Cretaceous magmatic front, which, as noted
above, was situated at that time in the Coastal Range (Boric
et al., 1990). Younger volcanic and intrusive events occurred
in the Late Cretaceous, when a volcanic sequence (Quebrada
Mala Formation) and a group of coeval 70 to 66 Ma pyroxene
diorites to rhyolite porphyries and flow domes, dated (U-Pb
zircon) between 70 and 66 Ma, were generated after the Andean arc front migrated eastward into the Centinela region
(see Figs. 1b, 8). Volcanism continued after the CretaceousTertiary boundary deformation event, with eruptions from
stratovolcanoes and small collapse calderas that were active
between the early Paleocene (64 Ma) and the early Eocene
(53 Ma). During this interval, a diverse group of epizonal intrusions composed mainly of pyroxene-biotite quartz diorite
to monzodiorite (60 Ma) and hornblende-biotite granodiorite
Sierra Limn Verde

6900

Orin (44-41)
as

Mirador (41-39)
Llano (41)
Esperanza (42-40)
Llano fault

Telgrafo (42-40)
Caracoles (42-41)

Esperanza fault

Centinela (45-44)

ra e
er lc
Si Du
ua
Ag

23

Coronado fault

Penacho Blanco (42)

Los

Toro
s

ult

fault

Las Lomas fault

Centinela

Sierr

a de

Sherezade (44-43)

l Buit

re fa

Las Lomas
duplex

fault

Pilar (43)
Polo Sur (42-41)

10 km

Upper Eocene (44-40 Ma) syntectonic, sedimentary and


volcanic rocks

Undifferentiated Cretaceous granitoids


Upper Cretaceous (78-66 Ma) sedimentary and
volcanic sequences (Quebrada Mala fomation)

Eocene (44-40 Ma) px-hb monzodioritic to hb-bt


granodioritic stocks and mineralized dacitic porphyry
intrusions (Esperanza-Telgrafo and Centinela-Polo Sur)
Paleocene (60-56 Ma) px-bt monzodiorites and
rhyolitic porphyry intrusions
Paleocene to Lower Eocene (64-53 Ma) volcanic rocks
(Cinchado formation )
Lower Paleocene (65-64 Ma) diorites and dacitic porphyry
intrusions

Lower Cretaceous (?) volcanic rocks


Lower Cretaceous (124-100 Ma) ol-px gabbros to
diorites and hb granodioritic porphyry intrusions
Jurassic to Lower Cretaceous back-arc sedimentary
and volcanic rocks
Upper Triassic (210-200 Ma) volcanic and sedimentary
rocks
Upper Paleozoic (290-270 Ma) basement

Upper Cretaceous (78-68 Ma) px diorites and (minor)


rhyolitic porphyry intrusions

FIG. 8. Regional geologic map of the Centinela cluster area. Note the 35-km-long NNE trend of 4240 Ma porphyry copper deposits emplaced during earlier stages of the Incaic event. Multiple superimposed intrusive pulses and volcanic
episodes between 120 and 40 Ma show a remarkable recurrence of magmatic events for >80 m.y. All ages are based on recently acquired U-Pb zircon data. (p = prospects, ol = olivine, bt = biotite, px = pyroxene).
0361-0128/98/000/000-00 $6.00

341

342

MPODOZIS AND CORNEJO

(5857 Ma) plus andesitic to dioritic porphyritic intrusions,


were emplaced into the Mesozoic units and Paleogene volcanic edifices.
Incaic magmatism and mineralization in the Centinela district occurred between 45 and 39 Ma (Mpodozis et al., 2009a;
Perell et al., 2010) and began, as revealed by numerous new
U-Pb zircon ages in the intrusive rocks and Re-Os ages in
molybdenite, ~12 to 10 m.y. after the termination of volcanism in the early Eocene.
This event coincides with the mostly copper-barren early
phase of Incaic intrusions at Chuquicamata-El Abra (4542
Ma) and Escondida (4441 Ma). The oldest Incaic intrusive
rocks include a small group of 45 Ma pyroxene-biotite and
quartz diorites yet, in contrast to Chuquicamata-El Abra and
Escondida, at Centinela, numerous mineralized porphyry
centers were emplaced between 44 and 39 Ma. They form,
together with some barren stocks, a 40-km-long, N- to NEtrending belt, which includes at least 10 discrete intrusive
complexes (Fig. 8). A syntectonic sequence of conglomerates
and volcaniclastic sandstones, which accumulated at the same
time as porphyry copper emplacement, comprises interbedded layers of dacitic block-and-ash deposits and tuffs with UPb zircon ages between 42 and 39 Ma.
The oldest porphyry systems (4543 Ma) occur along the
southwest end of the belt, and the age decreases systematically
to the northeast until reaching 39 Ma at the northeast edge of
the porphyry trend (Fig. 8). The geometry of the porphyry
complexes is controlled by their position relative to the main
structural feature of the district, a 3- to 5-km-wide, N-Strending fault zone that cuts obliquely across the porphyry belt. This
zone of intense deformation constitute to the northern termination of the Sierra de Varas fault, which stretches for >250
km along the western border of the Cordillera de Domeyko
(Mpodozis et al., 1993b; Soto et al., 2005; Figs. 2, 6a), and was
active both during and after porphyry emplacement (Fig. 8).
Porphyry deposits located west and east of this zone of concentrated deformation are largely undeformed. Copper mineralization in mineralized porphyry systems, located west of the
fault zone, are related to subvertical, hornblende-biotite dacites
dike swarms intruded into Paleocene volcanic/subvolcanic units
(e.g., Centinela) or early Eocene rhyolitic dome complexes
(Polo Sur, Perell et al., 2010). The oldest deposits where emplaced at 45 to 44 Ma (Centinela) and 44 to 43 Ma (Shererezade) to be followed by the intrusion by several barren pyroxene-hornblende dioritic stocks and lacoliths dated at 43 Ma,
although a porphyry copper system with the same age has been
also recognized at Pilar. A new pulse of copper-bearing intrusions occurred, finally, between 42 to 41 Ma, at Polo Sur, while
dacitic porphyries with a similar age (42 Ma) but apparently
barren have been documented at Penacho Blanco (Fig. 8).
Mirador, the youngest porphyry deposit recognized so far in the
district (4139 Ma; Mora et al., 2009) and located east of the
fault zone (Fig. 8), is also structurally undisturbed. The copper
mineralization, hosted within Jurassic marine limestones and
evaporites (Mora et al., 2009) is associated with a group of
multiphase, W- to NW-trending intrusions that, as in the older
Centinela and Polo Sur deposits, appear to be subvertical.
By contrast, deposits emplaced along the fault zone (Fig. 8)
have intermediate ages of 42 to 40 Ma (Caracoles, 4241 Ma;
Telgrafo, 4240 Ma; Esperanza, 4240 Ma; Llano, 41 Ma;
0361-0128/98/000/000-00 $6.00

Perell et al., 2004, 2010; Bisso et al., 2009; Mnchmeyer and


Valenzuela, 2009; Swaneck et al., 2009); they are all associated with tilted porphyry dike swarms. These deposits are
emplaced into moderately to steeply dipping strata, which are
disrupted by major postmineral faults that exhibit reverse,
normal, and strike-slip displacement components (Figs. 8, 9).
Although the widespread gravel cover makes it difficult to
satisfactorily resolve the structural relationships within the
whole district, the regional structure around the Esperanza
and Telgrafo deposits (Fig. 8) comprises a long-wavelength,
asymmetric, basement-cored anticline, bounded to the west
by a moderately E-dipping yet unexposed thrust fault that
was discovered during exploration drilling (Telgrafo fault;
Perell et al., 2004, 2010; Bisso et al., 2009; Mnchmeyer
and Valenzuela, 2009; Fig. 9). The hinge zone of the anticline
is, in turn, sliced by two subvertical faults (Coronado and
Llano faults; Figs. 8, 9) linked to the N-Strending regional
fault zone. Figure 9 includes a west-east structural section
across the Esperanza deposit, where mineralization is associated with a group of easterly inclined porphyry dikes emplaced within the ~40 to 50 W-dipping Triassic to Upper
Cretaceous strata that form the frontal limb of the anticline.
This panel, containing in part mineralized and altered host
rocks to the porphyry deposits, is upthrown to the west,
along the Telgrafo fault, over barren, unaltered mid-Eocene
(4239 Ma) sedimentary and volcanic rocks that accumulated when porphyry intrusions were being emplaced at depth.
The tilted, frontal-limb panel of the anticline is, in turn,
bounded to the east by the subvertical Esperanza fault (Fig.
9) that places Jurassic limestones over Late Upper Cretaceous strata. The rectilinear fault trace and the mismatch of
the lithology and age of the Late Cretaceous volcanic rocks
across the Esperanza fault show it includes an important
component of strike-slip movement, although the precise
age of deformation and genetic links between both faults remain to be determined. The Llano and Coronado faults are,
however, as shown in Figure 9, younger faults that are superimposed over the Telgrafo-Esperanza system, which exhibits both left-lateral and large, down-to-the-east components of displacement, part of which has a late Miocene or
younger (<10 Ma) age.
Structural relationships at Caracoles, in spite of being only
10 km to the south, are entirely different and difficult to
match with those observed at Esperanza-Telgrafo. In the
southern part of the district (Fig. 8), the fault zone is displaced to the west, out of strike with the fault zone to the
north, and includes a strike-slip duplex (Las Lomas duplex)
bounded by two N-Strending subvertical faults (Las Lomas
and Centinela faults; Fig. 8). Kinematic indicators show evidence for an episode of left-lateral displacement, even if the
internal geometry of the duplex is compatible with an earlier
(Late Cretaceous?) event of dextral shear (Marinovic and
Garca, 1999; Mpodozis et al., 2009a). The Caracoles deposit
(4240 Ma, Swaneck et al., 2009) is hosted in early Paleocene
(6460 Ma) volcaniclastic rocks beneath the gravel cover.
Copper mineralization is linked to a steeply SE-dipping
swarm of thin dacite porphyry dikes that follow the internal
structural trends of the Las Lomas duplex; this is consistent
with emplacement of the porphyry along a zone of (earlier or
synmineral?) strike-slip deformation.

342

343

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

49000

49200

49400

UTM (E )
m.a.s.l.

49600

Esperanza
fault
Coronado
fault

Llano fault
Mi-Pl

Pal

Ol

3200

2800
Mi-Pl

Ks
Jur

Ep

2400
Ks

2000

Ks
Trv

1600
Ks

Eo

1200

Ep
Pal
Ks

Esperanza copper porphyry intrusions

(41-40 Ma)

Jur

Jurassic marine sedimentary sequences

Trv

Upper Triassic volcanic rocks (210-200 Ma)

1 km

Telgrafo
fault

Lower Paleocene dacitic domes (64-63 Ma)


Upper Cretaceous volcanic rocks (Quebrada Mala
formation, 70-66 Ma)

0.5

Mi-Pl

800

Miocene to Pliocene gravels

Ol

Oligocene sedimentary rocks

Eo

Eocene sedimentary and volcanic strata

Left-lateral displacement on faults

0.5% CuT

FIG. 9. Schematic structural section across the Esperanza porphyry Cu deposit, Centinela district. The Esperanza orebody is part of a W-dipping sliver of Jurassic and Late Cretaceous strata intruded by Eocene porphyry dikes (4140 Ma),
thrust to the west (Telgrafo fault) on top of Eocene (4237 Ma) sedimentary and pyroclastic sequences. The Esperanza and
Coronado faults have a complex and younger displacement history, including strike-slip components. The down-to-the-east
displacement shown along the Coronado and Llano faults corresponds only to the youngest (late Miocene-Pliocene?) episode
of deformation. Location of section shown in Figure 8.

Long-lived intrusive clusters and porphyry deposits:


What is the relationship?
As discussed above, Eocene to Oligocene porphyry copper
deposits in northern Chile are located in districts of focused
and long-lived magmatic activity distributed along the Cordillera de Domeyko (Fig. 2). In some districts, magmatism
was active for 50 m.y. (Chuquicamata-El Abra, Escondida) or
even 80 m.y. (Centinela). Magmatism began in a back-arc setting while the arc front was located much farther west in the
Coastal Range or Central depression (Boric et al., 1990); to
date, the cause of these zones of long-term magmatism is uncertain. Yaez and Maksaev (1994) suggested that porphyry
spacing may be related to Rayleigh-Taylor diapirism along the
mid Eocene-early Oligocene arc. Behn et al. (2001) noted
that a correlation exists between the location of porphyry copper deposits and regional E-Wtrending negative magnetic
anomalies that extend from the coast to the modern Andean
arc, and which may represent crustal structures favorable for
magma ascent as the arc migrated eastward since the Jurassic.
Richards (2003) speculated that porphyry copper deposits
were emplaced at the intersections of the Domeyko fault system with NW-trending transcordilleran crustal-scale structures, yet with the exception of Potrerillos (Fig. 2), none of
these has been documented on the western (Chilean) side of
the Andes. To address these observations, Tomlinson and
Cornejo (2012) proposed a hybrid model to explain porphyry
spacing, which combines Rayleigh-Taylor diapirism with
crustal-scale structural control.
0361-0128/98/000/000-00 $6.00

Even if these models explain porphyry spacing, they do not


elucidate why clustered magmatism began long before the
Eocene. As more data are gathered, an even more striking
spatial relationship is emerging in several districts (Escondida, Centinela, Chuquicamata, Quebrada Blanca-Collahuasi,
Sierra Exploradora; Figs. 5, 7, 8), between extended, Cretaceous and younger magmatic activity, porphyry copper deposits, and Triassic intrusions which, near Escondida and Collahuasi, show evidence of weakly developed porphyry-style
copper mineralization (Cornejo et al., 2006; Munizaga et al.,
2008). In addition, indistinguishable geochemical signatures
reported in a pilot study by Wilson et al. (2011), which compared Triassic granodiorites to mineralized Eocene porphyry
deposits at Alejandro Hales (Fig. 5), may suggest tapping of a
common source region at depth. The problem remains open.
Neogene Tectonic Province of Central Chile
and Argentina
Late Eocene (?) to early Miocene extension and
the Abanico intra-arc basin
The tectonic evolution of the Andean orogen in central
Chile (3134 S) and contiguous Argentina differs from that
of northern Chile in that the main deformation events are
younger, the tectonic style is different, and the principal age
of porphyry copper mineralization is much younger (late
Miocene to early Pliocene). At present, a zone of flat subduction (Central Chile or Pampean flat-slab region; Cahill and
Isacks, 1992; Kay and Mpodozis, 2002; Ramos et al., 2002;

343

344

MPODOZIS AND CORNEJO

Fig. 10) that was formed by progressive slab shallowing, beginning in the early Miocene (Kay et al., 1987), extends from 27
to 33 S. As in northern Chile, the Coastal Range is formed by
the volcanic and intrusive remnants of a Jurassic to Cretaceous
arc system. From 32 S southward, a thick sequence of coeval
marine and terrestrial sedimentary rocks exposed along the
eastern slope of the Cordillera Principal corresponds to sediments that accumulated within the Mesozoic Neuqun backarc basin (Mpodozis and Ramos, 2008; Figs. 1012).
The most notable geologic feature, however, is a several
kilometer-thick volcano-sedimentary sequence that forms
most of the western part of the Cordillera Principal between
32 and 37 S (Figs. 10b, 11, 12), which has been traditionally

80

(a)

PERU

75

assigned to the Abanico, Coya-Machal, and Cura-Malln Formations (e.g., Charrier et al., 1996, 2002; Jordan et al., 2001;
Kay et al., 2005; Faras et al., 2008). These sequences accumulated in extensional volcano-tectonic depressions or intraarc basins that are referred to as the Abanico basin; Ar/Ar
ages at the latitude of Santiago (33 S) range from latest
Eocene to early Miocene (3521 Ma; Muoz et al., 2006).
Volcanic rocks are calc-alkaline to tholeiitic in composition,
and the overall geochemical signature suggests that volcanic
activity occurred over a relatively thin crust (<35 km; Kay et
al., 2005; Muoz et al., 2006).
Equivalent units extend for more than 1,500 km to the
south along the crest of the Andean range into the northern
73

(b)

71

IA

V
LI

h
Per- Chile Trenc

Oroc
lin

e
35

Per-Ch

INA
ENT

INA
ENT

39

e - Ofq

ui fau
lt

ARG

Ma

ile trench

30

ARG

Flat-slab
Region

Cen

tral

Val
le

Peru-C
hile tre
nc

37

Range

Ma

Figure 11

Concepcin

ge

Santiago

Concepcin

Somuncur
Plateau

20

Liqui

SVZ
40

Maip
o

32

CVZ

ez Rid

ernnd

Juan F

33

Santiago

Coastal

40

Antofagasta

3.4 cm/yr
30

una
Altiplano-P

BO

20

69
Aconcagua

a
M

Bariloche

Oligocene to Miocene intrusive rocks

41

Puerto Montt

Middle and upper Miocene volcanic


rocks (with major sedimentary
participation in the south)
Oligocene to middle Miocene volcanic
sequences and interbedded
sedimentary strata

43

Late Oligocene to middle Miocene


continental strata
0

Late Oligocene to middle Miocene


Somuncur plateau volcanic rocks

100 km

Isla
Magdalena

Eocene volcanic and intrusive rocks

FIG. 10. (a). Map showing position of the Chilean-Pampean flat-slab region and the distribution of Quaternary volcanoes
(CVZ = Central Andean volcanic zone, SVZ = Southern Andes volcanic zone). Light-colored area shows region with elevations >3 km. (b). Simplified geologic map showing the distribution of Oligocene-Miocene volcanic and sedimentary strata in
south-central Argentina and Chile from 32 to 45 S. Based on the compilation by Jordan et al. (2001) and the 1:1,000,000
geologic map of Chile (SERNAGEOMIN, 2002). Arrows north and south of the Maipo orocline (or Maipo mega kink;
Arriagada, et al., 2009) indicate the average values of paleomagnetically determined block rotations (Arriagada et al., 2009).
0361-0128/98/000/000-00 $6.00

344

345

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

CHILE

71

ARGENTINA
L A
TA R
N LE
O IL
FR RD
O
C

Los Andes

33

69

Valparaso

Mendoza

RIO BLANCO LOS BRONCES

CO
A
NG
E

Rancagua

P R
I N
C
I P
A L

ST
AL
RA

34

Tupungato

Santiago

A
San Jos

EL TENIENTE
Maipo

ORO
CL IN
SYM
E
MET
RY A
XIS

C O
D I
L
L E
R A

Curic

35

MA I
PO

M
0

Early Cretaceous sedimentary strata


(Neuqun basin)
Triassic to Jurassic sedimentary strata
(Neuqun basin)
Jurassic to Cretaceous volcano-sedimentary
sequences (Coastal Codillera)
Late Paleozoic to Jurassic batholiths
(Coastal Cordillera)
Fontal Cordillera Paleozoic basement

Upper Cenozoic foreland basin strata


Miocene intrusive rocks (Principal
Cordillera)
Miocene volcanic rocks (Farellones
formation)
Cretaceous intrusive rocks
Eocene? - Early Miocene volcanosedimentary sequences (Coya - Machal/
Abanico formations)

25

50 km

Quaternary stratovolcanoes
Quaternary volcanic rocks
Plio-Pleistocene tuffs
Fore arc Miocene sedimentary strata

FIG. 11. Geologic map of the area around the Maipo orocline from 3230' to 35 30' S (location in Fig. 10). Note change
in the structural trend of the Coastal Range and Principal Cordillera across the Maipo orocline (Faras et al., 2008; Arriagada
et al., 2009), from N-S, to the north, to N-NE, to the south. A = Aconcagua fold-and-thrust belt, M = Malarge fold-andthrust belt (adapted from Faras et al., 2008). See text for more details.

Patagonian Andes (Fig. 10b). Even though the erosion level


increases and the volume of preserved volcanic products
decreases southward, these sequences define a progressively
southward-widening belt that extends from the Coastal
Range to the eastern slope of the Andes at the latitude of
Puerto Montt (41 S, Fig. 10b). The nature of interbedded
sedimentary rocks changes along strike, from continental to
lacustrine facies in the northern part of the belt to marine
facies toward the south; geochemical signatures also display
evidence for crustal thinning and extension increasing to the
south. Muoz et al. (2000) indicated that 37 to 20 Ma mafic
lavas exposed at the latitude of Puerto Montt (41 S; Fig. 10b)
and erupted in an extensional setting possess island-arc
geochemical affinities and were likely produced by deep
0361-0128/98/000/000-00 $6.00

asthenospheric upwelling during a transient Oligocene to


early Miocene event of very rapid subduction. Deep mantle
upwelling has also been suggested by Kay et al. (2007a) to explain the origin of the 33 to 17 Ma basaltic flows of the large
back-arc Somuncur volcanic plateau in Patagonia (Fig. 10b).
Oligocene to Miocene submarine basaltic pillow lavas with
MORB geochemical features in the Aysn region indicate extreme crustal thinning along the arc farther south, between
43 and 46 S (Silva et al., 2003).
Miocene compressional failure of the Abanico basin
The late Eocene to early Miocene extensional period terminated at ~20 to 18 Ma, followed by compressional deformation leading to the emergence of the modern Andes in

345

346

MPODOZIS AND CORNEJO


70

Quaternary stratovolcanoes
Quaternary volcanic rocks
Upper Cenozoic foreland sedimentary
sequences
Miocene intrusive rocks
Volcanic rocks of the Abanico basin (Eocene?
to Miocene) include the Abanico,
Coya-Machali and Farellones formations
Jurassic to Cretaceous sedimentary
sequences of the Neuqun basin
Coastal Range block (Paleozoic to early
Cretaceous)
Frontal Cordillera Paleozoic basement

Chile

Argentina

Los Azules (10-8)


Rincones de Araya (9)

Piuquenes (11)
El Altar (12-10)
El Pachn (9-8)

Los Pelambres (14-10)


El Yunque (15)
Mercedario ( 13)
Cerro Mercedario

32

Cerro Bayo del Cobre


(12-10)
Amos- Andrs (9-8)
LR

Vizcachitas (11-10)
Morro Colorado
West Wall (11-9)

Pocuro fault

San Felipe

Aconcagua
Novicio (15-13)

Rio Blanco-Los Bronces


(7.7-4.7)

Valparaiso

33

Pimentn (11-9)

Los Machos (15-14)


Los Piches (14-12.5)

Los Sulfatos (7-6)

Tupungato

Santiago

San Jos

34

Rancagua
El Teniente
(6.5-4.6)
0

Maipo

50 km

M
71

70

FIG. 12. Tectonic sketch of the northern end of the Abanico intra-arc basin (3134 S), showing the location and age
(Ma, in parentheses) of Miocene to early Pliocene porphyry copper deposits of central Chile and contiguous Argentina and
the composite fold-and-thrust belt developed along the eastern margin of the basin (LR = La Ramada, A = Aconcagua, M =
Malarge fold-and-thrust belts).
0361-0128/98/000/000-00 $6.00

346

347

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

central and southern Chile and contiguous Argentina (e.g.,


Giambiagi et al., 2003). Between 31 and 34 S, much of this
deformation was focused along the eastern border of the
Abanico basin, which finally collapsed and was inverted by compressional deformation in response to the collision between

the colder and rigid western Coastal Range and eastern


Frontal Cordillera blocks (Fig. 13). Under these conditions,
the late Eocene to early Miocene volcanic sequences were
tectonically transported to the east and juxtaposed over the
sedimentary sequences of the Neuqun back-arc basin to

Rapid Convergence

Intra-arc
Abanico basin

Weak intraplate
coupling
Coastal Range Block

Mesozoic
sedimentary wedge

Frontal Cordillera block


Moho

St

35-21 Ma

Lithosphere

ee

sl

Asthenosphere

ab

Slow Convergence
Strong intraplate
coupling

Collapsed
& inverted
Abanico basin

Farellones volcanism
Syntectonic intrusions

AconcaguaLa Ramada
FTB

Accelerating
SAM plate

Subduction erosion

Zone of partial melting

Sh

allo

12-5 Ma

win
g

sla
b

Slow Convergence
Strong intraplate
coupling

El Teniente
Los Bronces

Los Pelambres

AconcaguaLa Ramada
FTB

Frontal Cordillera

Lower crust melting and


mixing with mantle-derived magmas
Delaminated
lower crustal blocks?

10-5 Ma

FIG. 13. Schematic diagram near 33 to 34 S, showing the evolution of the Abanico intra-arc basin during the Oligocene
and Miocene (SAM = South America, FTB = fold and thrust belt). Major porphyry copper deposits began to be formed at
10 Ma when the deformation front migrated to the east and the Frontal Cordillera was uplifted as a consequence of shallowing of the subducted Nazca plate. Even though the diagrams combine geologic relationships observed at different latitudes it shows the tectonic position of the Los Pelambres intrusions along the boundary thrusts of the deformed Abanico
basin and the Ro Blanco-Los Bronces and El Teniente intrusions in the less deformed rocks near the center of the former
basin farther to the west. Red arrows below the Abanico basin show hypothetical magma paths. (SAM = South American
plate, FTB= fold and thrust belt)
0361-0128/98/000/000-00 $6.00

347

348

MPODOZIS AND CORNEJO

form the La Ramada, Aconcagua, and Malarge fold-andthrust belts (Ramos et al., 1996; Figs. 1113). In central Chile,
a sharp decrease in the volume of volcanism ensued, the arc
front migrated eastward, and the geochemical and isotopic
signatures of younger, middle to late Miocene volcanic
sequences (e.g., Farellones Formation; Figs. 1011) indicate
progressive crustal thickening as a consequence of increased
horizontal shortening (Ramos et al., 1996; Kay and Mpodozis,
2002; Stern et al., 2010).
Between 33 and 35 S, the amount of Miocene and
younger back-arc shortening decreased along strike from
north to south, as the tectonic style in the deformed back-arc
sequences changed from the narrow, thin-skinned Aconcagua
fold-and-thrust belt to the wider, mixed thin- and thickskinned style of the Malarge fold-and-thrust belt to the
south (Ramos et al., 1996, 2004; Giambiagi et al., 2011; Figs.
1112). The transition zone, at 34 S, coincides with the symmetric axis of the W-NWtrending Maipo orocline (Faras et
al., 2008; Arriagada et al., 2009), and is revealed by the
change in orientation of both the Chile trench and the main
structural trends of the Principal Cordillera, from N-S to NNE (Fig. 11). The Maipo orocline, a more subtle feature than
the Bolivian orocline (see above), is also shown in paleomagnetic studies, as the magnitude of paleomagnetic block rotations determined in all rocks older than 10 Ma changes from
4clockwise north of the orocline to 32 clockwise to the
south (Arriagada et al., 2009; see Fig. 10b). These changes coincide with a rapid fall of the absolute elevation of the Andean
range, as well as an overall decrease of crustal thickness, from
50 km at 32 S to <40 km at 36 S (Introcaso et al., 1992;
Gilbert et al., 2006; Anderson et al., 2007).
These along-strike differences seem to reflect the effect of the
more contractional conditions prevailing during the Neogene
north of latitude 33 S, as shallowing of the subducting Nazca
slab progressed. This, in turn, lead to the establishment of the
modern Chilean or Pampean flat-slab region between 28
and 33 S (e.g., Kay and Mpodozis, 2002). Slab shallowing is
generally attributed to the buoyancy effect introduced by the
subduction of the E-NEtrending Juan Fernndez Ridge
during the late Miocene (Yaez et al., 2001; Ramos et al.,
2002, and references therein), although recent analogue and
numerical experiments (e.g., Martinod et al., 2005) show that
moderate-sized, buoyant ridges that impinge on a trench are
not able to alone induce formation of flat-slab segments of the
dimensions observed in central Chile and contiguous Argentina. Other authors (Manea et al., 2012) suggest that a
combination of trenchward motion of thick cratonic lithosphere accompanied by trench retreat may better explain the
formation of the Chilean-Pampean flat slab during the
Miocene.
Late Miocene to Early Pliocene Porphyry Copper
Deposits of Central Chile and Contiguous Argentina
Overview
The major porphyry copper deposits in central Chile and
contiguous Argentina are preferentially located at or near the
transition zone between the Chilean (or Pampean) flat-slab
zone and the steeper subduction zone beneath southern
Chile. This region contains three of the worlds largest copper
0361-0128/98/000/000-00 $6.00

deposits, at Los Pelambres, Ro Blanco-Los Bronces, and El


Teniente (Camus, 2003; Cooke et al., 2005; Sillitoe and Perell,
2005), as well as numerous smaller deposits and prospects
that straddle the Chile-Argentina border region from 31 to
35 S (Fig. 12). Although the first evidence of mineralization
that postdates the beginning of Miocene deformation is as old
as 15 to 13 Ma (Novicio, 1513 Ma; Los Machos, 1412.5 Ma;
Los Piches, 1412.5 Ma; Maksaev et al., 2009; Toro et al.,
2009; Fig. 12), most of the deposits north of 33 S (Los
Azules, Rincones de Araya, El Altar, Los Pelambres, El
Pachn, Vizcachitas, Amos-Andrs, Pimentn, Novicio, plus
West Wall in the San Felipe cluster) were emplaced between
12 and 8 Ma (Sillitoe and Perell, 2005; Maksaev et al., 2009;
Toro et al., 2009; Maygadn et al., 2011; Perell et al., 2012).
At this time, the E-NEtrending fragment of the Juan Fernndez Ridge began to subduct below the Los Pelambres region (Yaez et al., 2001; Kay and Mpodozis, 2002), causing
the Andean deformation front to move eastward; uplift of the
Frontal Cordillera (Fig. 11) commenced and was accommodated by regional thick-skinned faults (Ramos et al., 1996,
2004; Prez, 2001; Giambiagi et al., 2003; Fig. 12). By contrast, south of 33 S, the youngest and largest deposits, including Ro Blanco-Los Bronces and El Teniente (Maksaev et
al., 2004; Deckart et al., 2005; Maksaev et al., 2009; Fig. 12),
were emplaced from 7 to 4 Ma during a period (75 Ma) of
transient shallowing of the subducting slab below the
Neuqun basin, when volcanism with arc-like geochemical
signatures expanded up to 500 km east of the present day
Per-Chile trench (Ramos and Folguera, 2005; Kay et al.,
2006).
From a structural point of view, porphyry copper deposits
of central Chile and contiguous Argentina include a group of
stocks that postkinematically intruded regional thrusts along
the boundary between the former Abanico basin and the thinskinned La Ramada fold-and-thrust belt (Los Pelambres,
Amos-Andrs, Pimentn; Figs. 12, 14a), along the northern
part of the belt. Another, southern group (e.g., Vizcachitas,
West Wall, Ro Blanco-Los Bronces, and El Teniente; Figs.
1112) comprises intrusive complexes emplaced farther west
of the thrust belt in gently folded sequences of the Abanico
basin, with no obvious relationship to regionally significant
tectonic structures. A distinctive feature of the porphyry systems at Los Pelambres, Ro Blanco-Los Bronces, and El Teniente is the occurrence of barren and/or mineralized magmatic hydrothermal breccia complexes (Skewes and Stern,
1994).
Los Pelambres
The late Miocene Los Pelambres porphyry copper-molybdenum deposit and its smaller gold-bearing satellite, the
Frontera deposit (Fig. 12), are located in a narrow belt of intense deformation that involves Oligocene to early Miocene
(3318 Ma) volcanic rocks of the Abanico basin; these volcanic rocks, previously described as the Los Pelambres Formation, form part of the northern termination of the La Ramada fold-and-thrust belt at 3142' S (Mpodozis et al., 2009b;
Perell et al., 2012 ; Fig. 14a). The deposit is formed by multiple magmatic-hydrothermal centers with ages of 12.3 to 10.8
Ma, hosted by a precursor quartz diorite pluton (Los Pelambres stock, 13.613.0 Ma) and adjacent andesitic (Abanico

348

349

(a)

3200'

3150'

ora
Tot

7030'

El Pachn

Frontera

Chile

7020'

(b)

3312'

336'

4km

Los Piches

Ortiga

Infiernillo

Donoso

7018'

Lower Miocene (21-18 Ma) volcanic sequences

Pretectonic px diorites and granodiorites (23-21 Ma)

Los Sulfatos

La Americana

Sur Sur

Ro Blanco

Condell

7012'

Miocene volcanic rocks (Farellones


Formation)
> 0.5% Cu

San Francisco Batholith (14?-8 Ma)

Qz dioritic porphyry intrusions (ca 7 Ma)

Dacitic porphyry intrusions (5 Ma)

La Copa rhyolite-breccia complex (4 Ma)


Magmatic-hydrothermal breccias

La Paloma

San Enrique-Monolito

Brecha Sur

ncisc o

SAN FRANCISCO
BATHOLITH

San Manuel

El Plomo

RIO BLANCO-LOS BRONCES

7024'

FIG. 14. Comparison between the geologic setting of (a) the Los Pelambres-El Pachn and (b) Ro Blanco-Los Bronces districts. Despite the difference in scale, both
porphyry complexes are associated with N-NWtrending belts of intrusions and magmatic-hydrothermal breccias emplaced during the waning stages of long-lived earlier magmatic centers (Chalinga intrusive complex, San Francisco batholith). Map of the Ro Blanco-Los Bronces region is from Irarrzaval et al. (2010).

Upper Paleozoic basement

Triassic rift strata

Jurassic-Early Cretaceous, back-arc sedimentary sequences

Cretaceous to Paleocene volcanic and sedimentary units

10 km

Chalinga Intrusive Complex and related intrusives (23-15 Ma)


Intrusive rocks associated with copper mineralization
at Los Pelambres and El Pachn (14-8 Ma)
Post tectonic monzodiorites and hb-bearing
grandiorites and dacites (16-15 Ma)
Syntectonic (?) gabbros to px monzogranites (18 Ma)

Argentina

El Yunque

ruz

Strongly deformed, Oligocene to Lower-Miocene


(33-18 Ma) volcanic unit ("Los Pelambres Formation")
Oligocene to Lower Miocene (25-21 Ma) volcanic sequences
Upper Cretaceous to Eocene intrusive rocks

R o

Cuncumn

Pocuro fault

Los Pelambres

aca fa
ult

3140'

ult
res fa

Ro Santa Cru z

Mond

CHALINGA
INTRUSIVE
COMPLEX

elamb
Los P

Fra

0361-0128/98/000/000-00 $6.00
Ro Sa n

S
de
ro Y r b a Loca
e

C
ra
ille
ord
aC
ant
Est
e

LOS PELAMBRES

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

349

350

MPODOZIS AND CORNEJO

Formation) country rocks. The precursor stock and its contained porphyry copper mineralization were postkinematically
emplaced along the high-angle, Los Pelambres reverse fault,
which constitutes the eastern limit of the above-mentioned
zone of concentrated deformation (Mpodozis et al., 2009b;
Perell et al., 2009, 2012). At a more regional scale, however,
the Los Pelambres stock appears to be a satellite intrusive
body of a much larger (>250 km2) and long-lived, multistage
pluton located a short distance to the west (Chalinga intrusive
complex; Fig. 14), which was active for at least 8 m.y. between
23 and 15 Ma. The complex includes a pretectonic (with reference to regional deformation) suite of gabbros, pyroxene
diorites, and granodiorites, with U-Pb zircon ages between 23
and 21 Ma; a group of 18 Ma syntectonic olivine gabbro-diorites and granodiorites; and a younger and larger group of
post-tectonic, 16 to 15 Ma hornblende-bearing granodiorites.
Rocks of this latter group form the eastern margin of the
Chalinga intrusive complex, where they cut the traces of regional thrust faults (Fig. 14a).
Several small granodiorite to quartz diorite stocks and
hornblende-bearing dacite porphyry intrusions, dated at 15 to
13 Ma (Fig. 14a), form a NW-SEtrending string that extends
from the eastern Chalinga intrusive complex across the international frontier to Cerro Mercedario, ~70 km to the southeast in Argentina showing a southeastward propagation of
intrusive magmatism from the Chalinga complex during the
middle to late Miocene (Figs. 12, 14a). Some of these intrusions are associated with large porphyry-related hydrothermal
alteration zones, such as El Yunque (Fig. 14a). Porphyry magmatism and copper mineralization at Los Pelambres evolved
along this trend between ~14 and 10 Ma, whereas much more
limited data suggest that El Pachn was active between 9.2
and 8.4 Ma, and the Cerro Mercedario porphyry copper deposit at ~13 Ma (Sillitoe, 1977; Bertens et al., 2006; Perell et
al., 2012).
Ro Blanco-Los Bronces
The worlds largest copper district at Ro Blanco-Los Bronces
(Serrano et al., 1996; Skewes et al., 2003; Frikken et al., 2005;
Irarrzaval et al., 2010; Toro et al., 2012), is located, farther
south, near the center of the former Abanico basin (Figs.
1112). Although younger than Los Pelambres, it also formed
during the final stages of evolution of a long-lived, >10-m.y,
magmatic system, which includes a large premineral intrusive
complex (San Francisco batholith; Fig, 14b); this is remarkably
similar to the Chalinga intrusive complex of the Los Pelambres
area and was emplaced within gently folded, early Miocene
(1815 Ma) volcanic rocks (Farellones Formation; Fig. 14b).
Deckart et al. (2005, 2012) recognized that the San Francisco batholith includes three main intrusive phases, emplaced
between12 and 8 Ma, although an older U-Pb zircon age (14.7
Ma) for pyroxene monzodiorites (Jerez, 2007; Deckart 2012)
indicates that magma emplacement began during the middle
Miocene. The main mineralization at Ro Blanco-Los Bronces
occurred along an 8-km-long, N-NWtrending corridor,
which commences in the San Francisco batholith and extends
from Ro Blanco in the northwest to Los Sulfatos in the southeast (Fig. 14b). Multiple magmatic-hydrothermal breccia
complexes associated with porphyry copper-bearing, amphibole-rich quartz diorite porphyry intrusions were emplaced
0361-0128/98/000/000-00 $6.00

along this trend, beginning at 7.7 Ma, during the waning


stages of the San Francisco batholith. Intrusive activity terminated with the emplacement of late-mineral dacitic porphyry
intrusions at ~5 Ma, and the postmineral La Copa rhyolite
breccia complex at 4.7 Ma (Skewes et al., 2003; Deckart et al.,
2005, 2012; Irarrzaval et al., 2010; Toro et al., 2012; Fig.
14b).
El Teniente
El Teniente, 100 km south of Ro Blanco-Los Bronces, is
located near the axis of the Maipo orocline, at the center of
the deformed Coya-Machal (Abanico) basin and 30 km west
of the most internal thrust sheets of the Aconcagua foldand-thrust belt (Fig. 15). The deposit is hosted by gently
folded volcanic rocks of the El Teniente volcanic complex
(informal unit equivalent to the Farellones Formation),
which unconformably overlies Oligocene to early Miocene
volcanic rocks of the Coya-Machal Formation (Kay et al.,
2005; Stern et al., 2010). Mineralization at El Teniente is associated with a magmatic-hydrothermal center that records
at least 6 m.y. of continuous activity between ~9 and 3 Ma.
As at Los Pelambres and Ro Blanco-Los Bronces, magmatism at El Teniente includes a large premineral intrusive
complex containing older mafic facies (andesite intrusive
sills and olivine-pyroxene gabbros), described as the Teniente Mafic Complex, and a younger core of hydrous, hornblende-bearing intrusions (Sewell Tonalite; Cannell et al.,
2005; Stern et al., 2010; Vry et al., 2010; Fig. 15). Ages for
this group of intrusions are still not well constrained (Teniente Mafic Complex: 8.4 Ma K-Ar whole-rock; Sewell
Tonalite: 7.05 Ma total gas, Ar/Ar; Maksaev et al., 2004,
Stern et al., 2010, and references therein).
The copper mineralization at El Teniente is associated with
a >2-km-long, N-NWtrending body of dacite porphyry and
a series of small magmatic-hydrothermal breccias (Vry et al.,
2010 ), which in contrast to Los Pelambres or Ro Blanco-Los
Bronces, were emplaced within the earlier intrusions (Teniente Mafic Complex and Sewell Tonalite; Fig. 15). Detailed
U-Pb zircon, Ar/Ar, and Re-Os geochronology (Maksaev et
al., 2004) has allowed recognition of five pulses of felsic intrusions linked to mineralization emplaced between 6.5 and
4.6 Ma. Subsequently, the late-mineral 4.81 Ma Braden breccia pipe and a family of narrow, E-NEoriented olivine-hornblende lamprophyre dikes; ages from 3.9 to 2.9 Ma (Stern et
al., 2011) mark the last magmatic pulses in the district before
the magmatic front migrated ~50 km eastward to the ChileArgentine frontier to form the northernmost active volcanoes
of the modern Andean Southern volcanic zone (Fig. 12).
Origin of Miocene-Pliocene porphyry copper alignments
One of the most striking features of porphyry copper deposits in central Chile is their relationship to N-NWtrending
alignments of multiple magmatic-hydrothermal centers, emplaced during the final stages of long-lived magmatic systems
that include breccia complexes whose formation has been
considered to be triggered by decompression during rapid exhumation and tectonic uplift (Warnaars et al., 1985; Skewes
and Stern, 1994).
When the overall geometry of the porphyry alignments of
the three main centers described in this paper are considered,

350

351

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

6232

Laguna La Negra

( a)

(b)

Teniente Dacite Porphyry N Mine


5.280.10 Ma

Level Teniente 5
(2.284m)

1600

Laguna
La Huifa

400 km

8.91.4 Ma?

Dacitic
porphyries
6.090.18 Ma

Braden
Pipe

6228

Mafic Complex

800

Supergene zone

Braden Pipe
4.810.10 Ma
00

Marginal
Breccia

Porphyry A
5.670.19 Ma

2 km

6224

Sewell Tonalite (7.050.14 Ma)


Mafic Complex (8.914.4 Ma)

Braden Pipe (4.810.10 Ma)

Teniente Volcanic Complex

Unconsolidated deposits

Sewell Tonalite

Hydrothermal breccias
800

Latite Dike
4.820.09 Ma
400

Mafic Complex
Braden Pipe

Sewell Tonalite
Biotite breccia

Porphyry "A"

7.050.14 Ma
1200

Dacite Porphyry

Igneous breccia

Anhydrite breccia

Tourmaline breccia

Lamprophyre dykes

FIG. 15. Intrusive complexes in the area of El Teniente porphyry copper deposit (Stern et al., 2010). Note the N-NW
trend of mineralized porphyry intrusions and magmatic-hydrothermal breccias, hosted within the precursor Teniente mafic
complex and Sewell Tonalite. Late-stage lamprophyre dikes are perpendicular to the trend of the copper-bearing porphyry
deposits and breccias.

their orientation is similar and almost perpendicular to the direction of the Miocene and Pliocene plate convergence (Somoza and Ghidella, 2005). In contrast, the late lamprophyre
dikes at El Teniente (Fig. 15) are nearly parallel to the
Miocene plate convergence vector (see Fig. 16). Recent studies on the relationships between volcanism and tectonics in
the modern Southern volcanic zone of the Andes (Seplveda
et al., 2005; Cembrano and Lara, 2009) have shown that
primitive mantle-derived basalts ascend along NE-trending
fractures and faults, parallel to Hmax, which is regionally
close to the orientation of plate convergence. Nevertheless,
during the recent eruptions of southern Andes volcanoes like
Puyehue-Cordn Caulle (4030' lat S, 1960, 2011) evolved
rhyolites erupted along NW-trending basement faults that
are, in theory, severely disoriented in relationship to regional stresses to allow magma ascent. To overcome this
paradox, Seplveda et al. (2005) and Lara et al. (2006) have
suggested that coseismic or postseismic stress relaxation
during large subduction-zone earthquakes can produce
transient episodes of extension that allow ascent of evolved
magmas that otherwise would remain trapped in crustal
reservoirs. Similarly, decompression during seismic events
(see Sibson, 1987, 1994) appears to be a plausible mechanism to explain repeated cycles of breccia formation and
porphyry intrusion along now-concealed N-NW to NE-oriented faults that may have tapped the roofs of overpressured, deeper seated magma chambers during the evolution
of the central Chile porphyry coppers systems.
0361-0128/98/000/000-00 $6.00

Discussion
Linking geochemistry and tectonics:
the suggested adakite connection
Both the middle Eocene to early Oligocene and late Miocene
to early Pliocene porphyry copper-bearing intrusions include
intermediate rocks (SiO2 >56%) with abundant hydrous mineralogy, dominated by hornblende-bearing granodiorites and
dacites; these intrusions show geochemical and isotopic signature (SiO2 >56%, Sr >400 ppm, high Sr/Y ratios, low
HREE contents, high La /Yb ratios, 87Sr/86Sr <0.704) similar
to those described for typical adakitic rocks (cf. Defant and
Drumond, 1990; Castillo, 2012). Concave middle REE patterns in Chilean porphyry coppers indicate hornblende fractionation, whereas the lack of negative europium anomalies
denotes a high oxidation state of the magmas (see Kay et al.,
2005).
A relationship between adakites and mineralized porphyry
systems was proposed by Thiblemont et al. (1997) and Kay
and Mpodozis (2001), while some authors (e.g., Sun et al.,
2011) have suggested that the adakitic signatures of copperrich magmas are indicative of direct melting of subducted
oceanic crust. In accord with these views Oyarzn et al.
(2001) proposed that the middle Eocene to early Oligocene
porphyry copper belt of northern Chile may have been
formed when fast, oblique convergence led to flat subduction
and direct melting of the downgoing plate, whereas Reich et
al. (2003) proposed that the porphyry copper intrusions at

351

352

MPODOZIS AND CORNEJO


10

East Pacific
West Pacific
South Pacific

(a)

9
8
7

Average
preserved
half-spreading
rate
(cm/yr)

6
5
4
3
2
1

Age (Ma)
140 130 120 110 100 90

80

70

60

50

40

30

20

10

(c)

0
15

10

(b)
4

W velocity
2

N velocity
1

Velocity (cm/yr)

1
3

(d)
Angle between
convergence vector
and
present-day
-20
South American
margin
0(E-W)
-40

Age (Ma)
80

70

60

50

40

30

20

10

Convergence
velocity
(cm/yr)

0
4

+20

Age (Ma)
60

50

40

30

20

10

0 Ma

FIG. 16. Temporal changes of critical plate parameters that can be linked to adjustments in the tectonic regime along the
Andean margin. (a). Ocean crust production (average half-spreading rates since 140 Ma) for different regions of the Pacific
basin (Conrad and Lithgow-Bertelloni, 2007). (b). Absolute velocity of the South American plate since 80 Ma, treated as an
angular velocity vector decomposed into its ~ E to W and S to N components (Silver et al., 1998). (c). Cenozoic convergence
rates between the Farallon-Nazca and South American plates. Data from: 1 = Sdrolias and Muller (2006), 2 = Pardo-Casas
and Molnar (1987), 3 = Somoza (2008), 4 = Somoza and Ghidella (2005). (d). Direction of the Farallon-Nazca plate motion,
shown as the deviation angle from the E-W path (0, north = positive, south = negative). Adapted from Somoza and Ghidella
(2005). Shaded areas in all graphics indicate the time of porphyry copper emplacement during the Eocene to Oligocene and
Miocene to Pliocene.

Los Pelambres formed by melting of the subducting eastnortheast arm of the Juan Fernndez Ridge under flat-slab
conditions. However, as noted by Kay and Kay (2002) and
Castillo (2012), thermal models for subducting plates show
that sufficiently high pressure-temperature conditions for
slab fusion can be reached only in exceptional circumstances,
including subduction of very young and hot oceanic crust, as
prevailed during emplacement of the 12 Ma Cerro Pampa
adakites near the Chile Triple Junction in Patagonia (Kay and
Kay, 2002).
Richards and Kerrich (2007) and Richards (2011b) strongly
argued against slab melts being a necessary ingredient in
porphyry copper-gold mineralization, pointing out that the
critical factors for adakite genesis include elevated water and
sulfur contents as well as high oxidation state of the magmas,
which together result in hornblende fractionation and suppression of plagioclase crystallization. Richards (2011b) considered that such hydrous, oxidized conditions are typical in
normal arc settings. Nevertheless, this view is inconsistent
with the fact that giant porphyry copper deposits are not
widespread throughout the Andean history; as we have shown
0361-0128/98/000/000-00 $6.00

they formed during and/or after the most important tectonic


episodes that reshaped the whole Andean margin leading to
mountain building and crustal thickening.
Kay et al. (1999) and Kay and Mpodozis (2001) argued that
part of the water content of mineralizing Andean magmas can
be derived from the exsolution of fluids during the transformation of hydrous lower crustal amphibolite to dry garnetbearing eclogite during crustal thickening. When the crust of
the arc thickens to a critical value of ~45 km, amphibole and
plagioclase break down, water is liberated, and eclogite forms.
These relationships explain why the largest porphyry copper
deposits preferentially form during contractional events, such
as the Incaic episode of southern Peru and northern Chile,
and the late Miocene to early Pliocene event in central Chile
and contiguous Argentina.
Another process that can also contribute to generation of
water-rich adakitic magmas during periods of deformation is
subduction erosion (von Heune and Scholl, 1991; Stern, 1991,
2011; Kay et al., 2005). Subduction of oceanic crust, pelagic
and terrigenous sediments, and crust tectonically eroded
from the edge of the continent into the mantle-source region

352

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES

of Andean magmas may provide large amounts of water (e.g.,


Stern et al., 2010). Transient geochemical changes, such as
those depicted in Figure 1d, showing adakitic signatures
(Haschke et al., 2002; Kay et al., 2005) are consistent with the
loss of slivers of fore-arc crust by subduction erosion (Fig. 13)
during, or immediately following, major Mesozoic to Cenozoic deformation events. Intermittent and massive loss of
fore-arc crust is a possibility that has been recognized in recent numerical models of subduction erosion reported by
Keppie et al. (2009) and may explain the abrupt eastward
shifts of the magmatic front that punctuated the tectonic history of the Central Andes. Without ignoring other models for
adakite formation (e.g., Castillo, 2012), high Sr/Y hydrous
magmas linked to the middle Eocene to early Oligocene and
late Miocene to early Pliocene porphyry copper deposits of
northern and central Chile, respectively, are most likely attributed to a combination of melting of mantle derived magmas, including asthenosphere contaminated with pieces of
fore-arc crust that entered the mantle through subduction
erosion, that mixed with fluids derived from dehydration of
the base of thickened amphibole-rich lower crust during
these two main periods of Andean deformation (e.g., Kay et
al., 2007b).
Plate tectonics and Andean tectonic regimes
As discussed above, giant Andean porphyry copper deposits were emplaced during regional deformation events
that reshaped the entire Andean margin. Regional tectonic
and tectonomagmatic events are, ultimately, the result of
changes in plate interaction parameters that could influence
the state of stress in the overriding plate. Among these
changes on sea-floor spreading and plate convergence rates,
hot-spot activity, absolute plate velocity, and shifts in the position of the plate contact (trench roll-back), together with
variations in slab age, width, and/or dip has been considered
to strongly influence the dynamics of convergent margins
worldwide (e.g., Uyeda and Kanamori, 1979; Heuret and
Lallemand, 2005; Schellart and Rawlinson, 2010, and references therein). In recent years, variations in absolute plate
velocities (Russo and Silver, 1996; Silver et al., 1998) as well
as variations on the degree of mechanical interplate coupling (Yaez and Cembrano, 2004; Luo and Liu 2009a, b;
Iaffaldano et al., 2012) have been proposed as some of the
fundamental controls on Andean orogeny. Lamb and Davis
(2003) even suggested that differences in plate coupling are
possibly linked to along-strike differences in climate that
modulated the supply of sediments to the trench along the
Andean margin which, upon subduction, lubricated the
plate interface and, as a result, determined the degree of
mechanical coupling.
Tectonic regime: Northern Chile
Figure 16 shows time-dependent variations for some plate
tectonic parameters that can be compared with the geologic
record of the Andean margin. A correlation can be made between the Incaic compressional event and an episode of very
rapid oceanic crust production in the eastern Pacific, which
peaked in the middle Eocene at ~40 Ma (Fig. 16a; Conrad
and Lithgow-Bertelloni, 2007). At the same time the eastwest component of the absolute velocity vector of the South
0361-0128/98/000/000-00 $6.00

353

American plate (Silver et al., 1998) sharply increased (Fig.


4b). However, as shown in Figure 4c these effects were not
balanced, as should have occurred, by an associated increase
in the Farallon-South America convergence rate. According
to Somoza (1998) and Somoza and Ghidella (2005), the Farallon-South America convergence velocity was rather low
(67 cm/yr) or, as alternatively suggested by Pardo-Casas and
Molnar (1987) and Sdrolias and Muller (2006), was rapidly
decreasing (Fig. 16c).
This apparent paradox can be resolved if a high degree of
coupling existed at the plate interface at that time. If this was
the case for northern Chile and southern Peru during the
Eocene, the mechanically weak margin of the Central Andes,
pushed from the east and west, may have started to bend and
contract to form the Bolivian orocline and the Domeyko fault
system. High interplate coupling may also explain the formation of a flat-slab zone and the subsequent Oligocene magmatic lull recorded in southern Peru and northern Chile (e.g.,
Sandeman et al., 1995; James and Sacks, 1999; Kay et al., 1999;
Perell et al., 2003a; Hasckhe et al., 2006; Kay and Coira,
2009). Slab flattening intensifies interplate frictional forces,
which increases the possibility of subduction erosion; this, in
turn, would promote eastward migration of the shortening and
magmatic fronts toward the Andean foreland (e.g., Espurt et
al., 2008; Keppie et al., 2009; Martinod et al., 2010). Convergence rates increased during the Oligocene and Miocene,
leading to steepening of the Incaic flat slab; accumulation of
related mafic magmas below the already thickened crust led
to the production of crustal melts and the eruption of large
volumes of ignimbrites during the inception of volcanism
along the modern Central Andean volcanic zone during the
Miocene (Kay and Coira, 2009).
Tectonic regime: Central Chile
Although the tectonic evolution of central and south-central Chile appears to be different from that of northern Chile,
it is in fact complementary and preconditioned by the formation of the Bolivian orocline. As noted above, crustal flux toward the north along the southern limb of the orocline during
the middle Eocene to early Oligocene Incaic contraction may
have caused stretching and thinning of the upper crust in central Chile, thereby facilitating the opening of the Abanico
basin (McQuarrie, 2002; Arriagada et al., 2008; Boutelier and
Oncken, 2010). The along-strike differences in tectonic style
also agree with the numerical simulations by Schellart (2008),
which show that higher compression occurs near the center of
wide subduction zones where the trench remains stationary
or advances toward the continent. Rapid trench retreat (rollback) along the lateral slab edges may explain extension in the
overriding plate and be compatible with the opening of the
Abanico basin in central Chile.
The late Eocene to Oligocene opening of the Abanico basin
occurred during a period of steady westward displacement of
the South America plate (Silver et al., 1998), a period that also
coincided with another transient episode of very rapid generation of oceanic crust in the eastern Pacific, which commenced immediately after the formation of the Nazca plate
(Conrad and Lithgow-Bertelloni, 2007; Fig. 16a). In contrast
to the situation during the Incaic event, the convergence rate
between the Nazca and South American plates reached, at

353

354

MPODOZIS AND CORNEJO

this time, a record high (15 cm/yr; Somoza, 1998; Sdrolias and
Muller, 2006; Fig. 16b). Such a disparity may be explained if
weak plate coupling permitted rapid subduction and, as a
consequence, the generation of large volumes of magma and
extension in the overriding plate during the Oligocene to
early Miocene in central Chile and contiguous Argentina.
The beginning of contraction in central Chile and contiguous Argentina at ~20 Ma coincides (as shown in Fig. 16d)
with an acceleration of the absolute motion of South America
(see discussion in Kay and Copeland, 2006), drastic decrease
in oceanic crust production in the eastern Pacific, and a drop
in plate convergence rates between the Nazca and South
American plates (Somoza, 1998; Conrad and Lithgow-Bertelloni, 2007; Fig. 16). Inversion of the Abanico basin resulted,
north of 35 S in continued deformation during the Miocene,
causing an increase in crustal thickness to >50 km (Ramos et
al., 2004) and enhanced subduction erosion. Contamination
of the asthenosphere through subduction of fore-arc crust
created favorable conditions to produce water-rich mafic
melts with high sulfur and metal contents; these melts had
the capacity to ascend and evolve within an upper crustal
magma chamber to generate large porphyry copper deposits.
Concluding Remarks
There are few studies that consider the relationships between the regional-scale tectonic evolution of the Andes and
the formation of giant Cenozoic porphyry copper deposits.
However, it is apparent that these deposits formed during
critical moments in the tectonic evolution of the Andean margin. The emplacement of the large middle Eocene to early
Oligocene porphyry copper intrusions in northern Chile
seems to be associated with the formation of the Bolivian
orocline during the Incaic event, which was the result of an
unusual combination of factors. One critical factor was the acceleration of the absolute westward motion of South America
concurrent with strong mechanical coupling between the
South American and Farallon plates at a time when the rate
of ocean-crust production in the eastern Pacific was very
high. Bending of the Chilean margin during the Incaic event
activated the Domeyko fault system in northern Chile and
triggered the accompanying crustal thickening, slab shallowing, and increased subduction erosion. Volcanism virtually
ceased and favorable tectonomagmatic conditions (i.e. enhanced, subduction erosion, crustal thickening, lower crust
dehydration) permitted the formation of fertile hydrous magmas, while the transpressional and/or compressional upper
plate tectonic regime contributed to the establishment of
long-lived, upper-crustal magma chambers from which concentrating copper evolved, mostly below the Domeyko fault
system.
Younger, late Miocene to early Pliocene porphyry copper
deposits of central Chile and contiguous Argentina were
emplaced after inversion and collapse of the extensional,
intra-arc Abanico basin. The basin evolved between the late
Eocene and early Miocene when a relatively stable position of
South America over the mantle, linked to weak interplate
coupling, permitted fast subduction of the Nazca plate under
the Andean margin. Acceleration of the westward motion of
South America relative to the mantle reference frame at 20
Ma induced contractional deformation, accompanied by
0361-0128/98/000/000-00 $6.00

crustal thickening and eastward migration of the magmatic


front. At the same time, the subduction angle shallowed,
leading to the formation of a flat-slab region between 27 and
33 S as the Juan Fernndez Ridge was being subducted beneath the western edge of South America. These changes
again created favorable conditions for the formation of fertile
hydrous magmas.
The relationships described above demonstrate that the
concentration of huge porphyry copper deposits in the
Chilean Andes resulted directly from the tectonic evolution
of the margin and indicate that a tectonic trigger is essential
for the formation of giant porphyry coppers systems.
Acknowledgments
This contribution is the result of long years of work with
many colleagues at the Chilean Geological Survey, Antofagasta
Minerals, and various universities both in Chile and abroad.
We are especially grateful to Sue Kay, Andy Tomlinson, Terry
Jordan, Cesar Arriagada, Moyra Gardeweg, Rick Allmendinger,
Victor Ramos, Pierrick Roperch, Francisco Camus, Stephen
Matthews, Nicols Blanco, Francisco Herv, Reynaldo Charrier, Carlos Mnchmeyer, Ricardo Muhr, Jos Cembrano,
Carlos Arvalo, and many others who, for lack of memory, we
are here unable to mention. We thank Jeff Hedenquist, Dick
Sillitoe, Jos Perell, and Francisco Camus for pushing us
through this endeavor, and Antofagasta Minerals for providing time and support for the writing. Francisco Morales
helped with the preparation of the figures. Victor Ramos, Sue
Kay, Jos Perello, Jeff Hedenquist, and Dick Sillitoe carefully
edited the manuscript and made numerous suggestions that
helped to greatly improve earlier versions of the manuscript.
REFERENCES
Amilibia, A., and Skarmeta, J., 2003, La inversin tectnica en la Cordillera
de Domeyko en el norte de Chile y su relacin con la intrusin de sistemas
porfricos de Cu-Mo [ext.abs.]: X Congreso Geolgico Chileno, Concepcin, Extended Abstracts, v. 2, p. 17.
Amilibia, A., Sabat, F., McClay, K.R., Muoz, J.A., and Chong, G., 2008, The
role of inherited tectono-sedimentary architecture in the development of
the central Andean mountain belt: Insights from the Cordillera de
Domeyko: Journal of Structural Geology, v. 30, p. 15201539.
Anderson, M., Alvarado, P., Zandt, G., and Beck, S., 2007, Geometry and
brittle deformation of the subducting Nazca Plate, Central Chile and Argentina: Geophysical Journal International, v. 171, p. 419434.
Arriagada, C., Roperch, P., Mpodozis, C., and Cobbold, P.R., 2008, Paleogene building of the Bolivian orocline: Tectonic restoration of the Central
Andes in 2-D map view: Tectonics, v. 27, TC6014, 14 p., doi:10.1029/2008
TC002269.
Arriagada, C., Mpodozis, C., Yaez, G., Charrier, R., Faras, M., and Roperch, P., 2009, Rotaciones tectnicas en Chile central: El oroclino de Vallenar y el megakink del Maipo [ext.abs.]: XII Congreso Geolgico Chileno Santiago, Extended Abstracts (CD), 4 p.
Astini, R.A., Benedetto, J.L., and Vaccari, N.E., 1995, The early Paleozoic
evolution of the Argentine Precordillera as a Laurentian rifted, drifted, and
collided terrane: A geodynamic model: Geological Society of America Bulletin, v. 107, p. 253273.
Audin, L., Hrail, G., Riquelme, R., Darrozes, J, Martinod, J., and Font, E.,
2003, Geomorphological markers of faulting and neotectonic activity along
the western Andean margin, northern Chile: Journal of Quaternary Science, v. 18, p. 681694.
Ballard, J., Palin, J.R., Palin, M., Williams, I., and Campell, I., 2001, Two ages
of porphyry intrusion resolved for the super-giant Chuquicamata copper
deposit of northern Chile by ELA-ICP-MS: Geology, v. 29, p. 383386.
Barra, F., 2011, Assessing the longevity of porphyry Cu-Mo deposits: Examples from the Chilean Andes: SGA Biennial Meeting, 11th, Antofagasta,
Proceedings, p. 399401.

354

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES


Beck, M., Rojas, C., and Cembrano, J., 1993, On the nature of buttressing in
margin parallel strike-slip fault systems: Geology, v. 21, p. 755758.
Behn, G., Camus, F., Carrasco P., and Ware, H., 2001, Aeromagnetic signature of porphyry copper systems in northern Chile and its geologic implications: Economic Geology, v. 96, p. 239248.
Bertens, A., Clark, A.H., Barra, F., and Deckart, K., 2006, Evolution of the
Los Pelambres-El Pachn porphyry copper-molybdenum district, Chile/
Argentina [ext. abs.]: XI Congreso Geolgico Chileno, Antofagasta, Extended Abstracts, v. 2, p. 179181.
Bisso, C., Lazcano, E., Guzmn, J., and Gonzlez, S., 2009, Geologa y desarrollo del yacimiento Esperanza, Distrito Sierra Gorda, Antofagasta [ext. abs.]:
XII Congreso Geolgico Chileno, Santiago, Extended Abstracts (CD), 4 p.
Blanco, N., 2008, Estratigrafa y evolucin tectono-sedimentaria de la cuenca
cenozoica de Calama (Chile, 22 S): Unpublished M.Sc. thesis, University
of Barcelona, 68 p.
Blanco, N., and Tomlinson, A.J., 2009, Carta Chiu-Chiu, Regin de Antofagasta: Santiago, Servicio Nacional de Geologa y Minera, Carta Geolgica
de Chile, 117, 1:50,000, 54 p.
Boric, R., Daz, F., and Maksaev, V., 1990, Geologa y yacimientos metalferos de la regin de Antofagasta: Santiago, Servicio Nacional de Geologa y
Minera, Boletn 40, 246 p.
Boric, R., Daz, J., Becerra, H., and Zentilli, M., 2009, Geology of the Ministro Hales mine (MMH), Chuquicamata district, Chile [ext. abs.]: XII
Congreso Geolgico Chileno, Santiago, Extended Abstracts (CD), 4 p.
Boutelier, D.A., and Oncken, O., 2010, Role of the plate margin curvature in
the plateau buildup: Consequences for the central Andes: Journal of Geophysical Research, v 115, B04402, 27 p., doi:10.1029/2009JB006296.
Cahill, T. A. and Isacks, B. L, 1992, Seismicity and shape of the subducted
Nazca Plate: Journal of Geophysical Research, v. 97, p. 503517.
Campbell, I.H., Ballard J.R., Palin J.M., Allen C., and Faunes, A., 2006, UPb zircon geochronology of granitic rocks from the Chuquicamata-El Abra
porphyry copper belt of northern Chile: Excimer laser ablation ICP-MS
analysis: Economic Geology, v. 101, p. 13271344.
Camus, F, 2003, Geologa de los sistemas porfricos en los Andes de Chile:
Santiago, Servicio Nacional de Geologa y Minera, 267 p.
Cannell, J., Cooke, D.R., Walshe, J.L., and Stein, H., 2005, Geology, mineralization, alteration, and structural evolution of the El Teniente porphyry
Cu-Mo deposit: Economic Geology, v. 100, p. 9791003.
Castillo, P.R., 2012, Adakite petrogenesis: Lithos, v. 134135, p. 304316.
Cawood, P.A, 2005 Terra Australis orogen: Rodinia breakup and development of the Pacific and Iapetus margins of Gondwana during the Neoproterozoic and Paleozoic: Earth-Science Reviews, v. 69, p. 249279.
Cembrano, J., and Lara, L., 2009, The link between volcanism and tectonics
in the southern volcanic zone of the Chilean Andes: A review: Tectonophysics, v. 472, p. 96113.
Charrier, R., Wyss, A.R., Flynn, J.J., Swisher, C.C., Norell, M.A., Zapatta, F.,
McKenna, M.C., and Novaceck, M.J., 1996, New evidence for late Mesozoic-early Cenozoic evolution of the Chilean Andes in the upper Tinguiririca valley (35 S), Central Chile: Journal of South American Earth Sciences,
v. 9, p. 130.
Charrier, R., Baeza, O., Elgueta, S., Flynn, J.J., Gans, P., Kay, S.M., Muoz,
N., Wyss, A.R., and Zurita, E., 2002, Evidence for Cenozoic extensional
basin development and tectonic inversion south of the flat-slab segment,
southern Central Andes, Chile (3336 SL): Journal of South American
Earth Sciences, v. 15, p. 117139.
Chen, H., Clark, A.H., Kyser, T.K., Ullrich, T.D., Baxter, R., Chen, Y., and
Moody, T.C., 2010, Evolution of the giant Marcona-Mina Justa iron oxidecopper-gold district, south-central Peru: Economic Geology, v. 105, p.
155185.
Chew, D.M., Schaltegger, U., Koler, J., Whitehouse, M., Gutjahr, M., Spikings, R., and Mikovc, A., 2007, U-Pb geochronologic evidence for the evolution of the Gondwanan margin of the north-central Andes: Geological Society of America Bulletin, v. 119, p. 697711.
Conrad, C., and Lithgow-Bertelloni, C., 2007, Faster seafloor spreading and
lithosphere production during the mid-Cenozoic: Geology, v. 35, p. 2932.
Cooke, D.R., Hollings, P., and Walshe, J.L., 2005, Giant porphyry deposits:
Characteristics, distribution, and tectonic controls: Economic Geology, v.
100, p. 801818.
Cornejo, P., and Matthews, S., 2001, Evolution of magmatism from the uppermost Cretaceous to Oligocene, and its relationship to changing tectonic
regime in the Inca de Oro-El Salvador area (northern Chile) [ext. abs.]:
South American Symposium on Isotope Geology, 3rd, Pucn, Chile, Extended Abstracts (CD), p. 558561.
0361-0128/98/000/000-00 $6.00

355

Cornejo, P., Tosdal, R.M., Mpodozis, C., Tomlinson, A., Rivera, O., and Fanning, M.C., 1997, El Salvador, Chile, porphyry copper deposit revisited:
Geologic and geochronologic framework: International Geology Review, v.
39, p. 2254.
Cornejo, P., Matthews, S., and Prez de Arce, C., 2003, The K-T compressive deformation event in northern Chile (2427 S) [ext. abs.]: X Congreso Geolgico Chileno, Concepcin, Extended Abstracts (CD), Thematic Session ST1, p. 113.
Cornejo, P., Matthews, S., Marinovic, N., Prez de Arce, C., Basso, M., Alfaro J., and Navarro, M., 2006, Alteracin hidrotermal y mineralizacin recurrente de Cu y Cu-Mo durante el Prmico y el Trisico en la Cordillera
de Domeyko (Zona de Zaldvar-Salar de los Morros): Antecedentes geocronolgicos U-Pb, 40Ar-39Ar y Re-Os [ext. abs.]: XI Congreso Geolgico
Chileno, Antofagasta, Extended Abstracts, v. 2, p. 219222.
Cunningham, W.D., and Mann, P., 2007, Tectonics of strike-slip restraining
and releasing bends: Geological Society Special Publication 290, p. 112.
Dalziel, I.W.D., 1997, Neoproterozoic-Paleozoic geography and tectonics:
Review, hypothesis, environmental speculation: Geological Society of
America Bulletin, v. 109, p. 1642.
Deckart, K., Clark, A.H., Aguilar, A.C., Vargas, V.R., Bertens, A.N.,
Mortensen, J.K., and Fanning, M., 2005, Magmatic and hydrothermal
chronology of the giant Rio Blanco porphyry copper deposit, central Chile:
Implications of an integrated U-Pb and 40Ar-39Ar database: Economic Geology, v. 100, p. 905934.
Deckart, K., Clark, A.H., Aguilar, C., Cuadra, P., and Fanning, M., 2012, Refinement of the time-space evolution of the giant Mio-Pliocene Ro BlancoLos Bronces porphyry Cu-Mo cluster, Central Chile: New U-Pb (SHRIMP
II) and Re-Os geochronology and 40Ar-39Ar thermochronology data: Mineralium Deposita (in press, available online)
Defant, M.J., and Drummond, M.S., 1990.Derivation of some modern arc magmas by melting of young subducted lithosphere: Nature, v. 347, p. 662665.
Dilles, J., Tomlinson, A., Martin, M., and Blanco, N., 1997, El Abra and Fortuna complexes: A porphyry copper batholith sinistrally offset by the Falla
Oeste [ext. abs.]: VIII Congreso Geolgico Chileno, Antofagasta, Extended
Abstracts, v. 3, p. 18831887.
Dilles, J., Tomlinson, A., Garca, M., and Alcota, H., 2011, The geology of the
Fortuna Granodiorite Complex, Chuquicamata district, northern Chile:
Relation to porphyry copper deposits: SGA Biennial Meeting, 11th, Antofagasta, Proceedings, p. 399401.
Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna,
C., and Brusset, S., 2008, Flat subduction dynamics and deformation of the
South American plate: Insights from analog modeling: Tectonics, v. 27,
TC3011, doi:10.1029/2007TC002175, 2008.
Faras, M., Charrier, R., Carretier, S., Martinod, J., Fock, A., Campbell, D.,
Cceres, J., and Comte, D., 2008, Late Miocene high and rapid surface uplift and its erosional response in the Andes of central Chile (3335 S):
Tectonics, v. 27, TC1005, doi:10.1029/2006TC002046.
Franzese, J., and Spalleti, L., 2001, Late Triassic-Early Jurassic continental
extension in southwestern Gondwana, tectonic segmentation and prebreakup rifting: Journal of South American Earth Sciences, v. 14, p.
257270.
Frikken, P.H., Cooke, D.R., Walshe, J.L., Archibald, D., Skarmeta, J., Serrano, L., and Vargas, R., 2005, Mineralogical and isotopic zonation in the
Sur-Sur tourmaline breccia, Ro Blanco-Los Bronces Cu-Mo deposit,
Chile: Implications for ore genesis: Economic Geology, v. 100, p. 935961.
Giambiagi, L., Ramos, V., Godoy, E., Alvarez, P., and Orts, S., 2003, Cenozoic
deformation and tectonic style of the Andes, between 33 and 34 south latitude: Tectonics, v. 22 1041, doi:10.1029/2001TC001354.
Giambiagi, L., Mescua, J., and Bechis, F., 2011, El Frente orognico del sector sur de los Andes centrales.Variaciones latitudinales en acortamiento, topografa, levantamiento estructural y denudacin [ext. abs.]: XVIII Congreso Geolgico Argentino, Neuqun, Extended Abstracts (CD), 2 p.
Gilbert, H., Beck, S., and Zandt, G., 2006, Lithospheric and upper mantle
structure of central Chile: Geophysical Journal International, v. 165, p.
383398.
Gotberg, N., McQuarrie, N., and Carlotto, V., 2010, Comparison of crustal
thickening budget and shortening estimates in southern Peru (1214 S):
Implications for mass balance and rotations in the Bolivian orocline: Geological Society of America Bulletin, v. 122, p. 727742.
Gtze, H.J., and Krausse, S., 2002, The Central Andean gravity high. A relic
of an old subduction complex?: Journal of South American Earth Sciences,
v. 14, p. 799811.

355

356

MPODOZIS AND CORNEJO

Hammerschmidt, K., Dobel, R., and Friedrichsen, H., 1992.Implication of


40Ar-39Ar dating of early Tertiary volcanic-rocks from the North-Chilean
Precordillera: Tectonophysics, v. 202, p. 5581.
Haschke, M., Siebel W., Gnther A., and Scheuber, E., 2002, Repeated
crustal thickening and recycling during the Andean orogeny in North Chile
(2126 S): Journal of Geophysical Research, v. 107, doi 10.1029/2001
JB000328.
Haschke, M., Gnther, A., Melnick, D., Echtler, H., Reutter, K.J., Scheuber,
E., and Oncken, O., 2006, Central and southern Andean tectonic evolution
inferred from arc magmatism, in Oncken, O., et al., eds., The Andes: Frontiers in Earth Sciences, Pt. III: Springer-Verlag, p. 337353,
Herv, F., 1988, Late Paleozoic subduction and accretion in southern Chile:
Episodes, v. 11, p. 183188.
Herv, M., Sillitoe, R.H., Wong, C., Fernndez, P., Crignola, F., Ipinza, M.,
and Urza, F., 2012, Geologic overview of the Escondida porphyry copper
district, northern Chile: Society of Economic Geologists Special Publication, 16, p. 5578.
Heuret, A., and Lallemand, S., 2005.Plate motions, slab dynamics and backarc deformation: Physics of the Earth and Planetary Interiors, v. 149, p.
3151.
Hindle, D., Kley, J., Oncken, O., and Sobolev, S., 2005.Crustal balance and
crustal flux from shortening estimates in the Central Andes: Earth and
Planetary Science Letters, v. 230, p. 113124.
Hoffmann-Rothe, A., Ritter, O., and Janssen, C., 2004, Correlation of electrical conductivity and structural damage at a major strike-slip fault in
northern Chile: Journal of Geophysical Research, v. 109, B10101, doi:10.
1029/2004JB003030.
Hong, F., Del Papa, C., Powell, J., Petrinovic, I., Mon, R., and Veraco, V.,
2007, Middle Eocene deformation and sedimentation in the Puna-Eastern
Cordillera transition (2326 S): Control by preexisting heterogeneities on
the pattern of initial Andean shortening: Geology, v. 35, p. 271274.
Iaffaldano, G., Di Giuseppe, E., Corbi, F., Funiciello, F., Faccenna, C., and
Bunge, H., 2012, Varying mechanical coupling along the Andean margin:
Implications for trench curvature, shortening and topography: Tectonophysics, v. 526529, p. 1623.
Introcaso, A., Pacino, M.C., and Fraga, A.,1992, Gravity, isostasy and Andean
crustal shortening between latitudes 30 and 35 S: Tectonophysics, v. 205,
p. 3138.
Irarrzaval, V., Sillitoe, R.H., Wilson, A.J., Toro, J.C., Robles, W., and Lyall,
G.D., 2010, Discovery history of a giant, high-grade, hypogene porphyry
copper-molybdenum deposit at Los Sulfatos, Los Bronces-Ro Blanco district, central Chile: Society of Economic Geologists Special Publication 15,
p. 253269.
Isacks, B.L., 1988, Uplift of the Central Andean plateau and bending of the
Bolivia orocline: Journal of Geophysical Research, v. 93, p. 32113231.
James, D., and Sacks, I.S., 1999, Cenozoic formation of the central Andes: A
geophysical perspective: Society of Economic Geologists Special Publication 7, p. 125.
Jerez, C, 2007, Contribucin a la geocronologa y geoqumica de los intrusivos Yerba Loca y San Francisco, Cordillera de Chile central (33 S): Unpublished M.Sc. thesis, Santiago, Universidad de Chile, 58 p.
Jordan, T.E., Isacks, B., Allmendinger, R., Brewer, V., Ramos, V., and Ando,
C., 1983, Andean tectonics related to geometry of subducted Nazca plate:
Geological Society of America Bulletin, v. 94, p. 341361.
Jordan, T.E., Burns, W., Veiga, R., Pngaro, F., Copeland, P., Kelley, S., and
Mpodozis, C., 2001, Extension and basin formation in the Southern Andes
caused by increased convergence rate: A Mid-Cenozoic trigger for the
Andes: Tectonics, v. 20, p. 308324.
Jordan, T.E., Mpodozis, C., Blanco, N., Pananont, P., and Dvila, F., 2004,
Extensional basins in a convergent margin: Oligocene-Miocene Salar de
Atacama and Calama basins, Central Andes [abs.]: EOS, v. 85(47), Fall
Meeting Supplement, Abstract T35A-0475.
Kay, R.W., and Kay, S.M., 2002, Andean adakites: Three ways to make them:
Acta Petrologica Sinica, v. 18, p. 303311.
Kay, S.M., and Coira, B., 2009, Shallowing and steepening subduction zones,
continental lithosphere loss, magmatism and crustal flow under the central
Andean Altiplano-Puna plateau: Geological Society America Memoir 204,
p. 229260.
Kay, S.M., and Copeland, P., 2006, Early to middle Miocene back-arc magmas of the Neuqun basin of the southern Andes: Geochemical consequences of slab shallowing and the westward drift of South America
(3539 S lat.): Geological Society of America Special Paper 407, p.
185213.
0361-0128/98/000/000-00 $6.00

Kay, S.M., and Mpodozis, C., 2001, Central Andean ore deposits linked to
evolving shallow subduction systems and thickening crust: GSA Today, v.
11(3), p. 49.
2002, Magmatism as a probe to the Neogene shallowing of the Nazca
plate beneath the modern Chilean flat-slab: Journal of South American
Earth Sciences, v. 15, p 3957.
Kay, S.M., Maksaev, V.A., Moscoso, R., Mpodozis, C. and Nasi, C., 1987, Probing the evolving Andean lithosphere: Mid-late Tertiary magmatism in
Chile (293030 S) over the modern zone of subhorizontal subduction:
Journal of Geophysical Research, v. 92, p. 61736189.
Kay, S.M., Ramos, V.A., Mpodozis, C., and Sruoga, P., 1989, Late Paleozoic
to Jurassic silicic magmatism at the Gondwanaland margin: Analogy to the
Middle Proterozoic in North America?: Geology, v. 17, p. 324328.
Kay, S.M., Mpodozis, C., and Coira, B., 1999, Magmatism, tectonism and
mineral deposits of the Central Andes (2233S latitude) : Society of Economic Geologists Special Publication 7, p. 2759.
Kay, S.M., Godoy, E., and Kurtz, A., 2005, Episodic arc migration, crustal
thickening, subduction erosion, and magmatism in the south-central
Andes: Geological Society of America Bulletin, v. 117, p. 6788.
Kay, S.M., Mancilla, O., and Copeland, P., 2006, Evolution of the backarc
Chachahun volcanic complex at 37S latitude over a transient Miocene
shallow subduction zone under the Neuqun basin: Geological Society of
America Special Paper 407, p. 215246.
Kay, S.M., Ardolino, A., Gorring, M., and Ramos, V.A., 2007a, The Somuncura large igneous province in Patagonia: Interaction of a transient
mantle thermal anomaly with a subducting slab: Journal of Petrology, v. 48,
p. 4377.
Kay, S.M., Goss, A., Kay, R., Mpodozis, C., 2007b, Frontal arc migration,
forearc subduction erosion and adakitic-like magmas: Examples from the
Andean margin and implications for other arcs [abs.]: EOS, American Geophysical Union, Spring Meeting 2007, Abstract T42A-05.
Keppie, D.F., Currie, C., and Warren, W., 2009, Subduction erosion modes:
Comparing finite element numerical models with the geological record:
Earth and Planetary Science Letters, v. 287, p. 241254.
Kleiman, L., and Japas, M., 2009, The Choiyoi volcanic province at
34S36S (San Rafael, Mendoza, Argentina): Implications for the late Paleozoic evolution of the southwestern margin of Gondwana: Tectonophysics, v. 473, p. 283299.
Kley, J., and Monaldi, C., 1998, Tectonic shortening and crustal thickness in
the Central Andes: How good is the correlation?: Geology, v. 26, p. 723726.
Kley, J., Monaldi, C., Rossello, E., and Ege, H., 2005, The Eastern Cordillera
of the Central Andes: Inherited mechanical weakness as a first-order control on the Cenozoic orogeny [ext. abs.]: International Symposium on Andean Geodynamics, 6th, Barcelona, Extended Abstracts, p. 432435.
Kloppenburg, A., Grocott, J., and Hutchinson, D., 2010, Structural setting
and synplutonic fault kinematics of a Cordilleran Cu-Au-Mo porphyry mineralization system, Bingham mining district, Utah: Economic Geology, v.
105, p. 743761.
Ladino, M., Tomlinson, A.J., and Blanco, N., 1997, Nuevos antecedentes
para la edad de la deformacin cretcica en Sierra de Moreno, II Regin
de Antofagasta-Norte de Chile [ext. abs.]: VIII Congreso Geolgico Chileno, Antofagasta, Extended Abstracts, p. 103107.
Lamb, S.H., 1994, Behavior of the brittle crust in wide zones of deformation:
Journal of Geophysical Research, v. 99, p. 44574483.
2001, Vertical axis rotation in the Bolivian orocline, South America, 2.
Kinematic and dynamical implications: Journal of Geophysical Research, v.
106, p. 26332653.
Lamb, S., and Davis, P., 2003, Cenozoic climate change as a possible cause
for the rise of the Andes: Nature, v. 425, p. 792797.
Lang, J.R., and Titley, S.R., 1998, Isotopic and geochemical characteristics of
Laramide magmatic systems in Arizona and implications for the genesis of
porphyry copper deposits: Economic Geology, v. 93, p. 138170.
Lara, L., Lavenu, A., Cembrano, J., and Rodrguez, C., 2006, Structural controls of volcanism in transversal chains: Resheared faults and neotectonics
in the Cordn Caulle-Puyehue area (40.5 S), Southern Andes: Journal of
Volcanology and Geothermal Research, v. 158, p. 7086.
Larson, R.L., 1991, Geological consequences of superplumes: Geology, v. 19,
p. 963966.
Lindsay, D., Zentilli, M., and Rojas de la Rivera, J., 1995, Evolution of an active ductile to brittle shear system controlling mineralization at the
Chuquicamata porphyry copper deposit, northern Chile: International Geology Review, v. 37, p. 945958.

356

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES


Luo, G., and Liu, M., 2009a, Why short-term crustal shortening leads to
mountain building in the Andes, but not in Cascadia?: Geophysical Research Letters, v. 36, L08301, 5 p., doi:10.1029/2009GL037347.
2009b, How does trench coupling lead to mountain building in the
Subandes?: A viscoelastoplastic finite element model: Journal of Geophysical Research, v. 114, B03409, 16 p., doi:10.1029/2008JB005861, 2009.
Maksaev, V., and Zentilli, M., 1988, Marco metalognico regional de los megadepsitos de tipo prfido cuprfero del Norte Grande de Chile [ext. abs.]:
V Congreso Geolgico Chileno, Santiago, Extended Abstracts, v. 1(B), p.
181212.
1999, Fission track thermochronology of the Domeyko Cordillera,
northern Chile: Implications for Andean tectonics and porphyry copper
metallogenesis: Exploration and Mining Geology, v. 8, p. 6589.
Maksaev, V., Munizaga, F., McWilliams, M., Fanning, M., Mathur, R., Ruiz,
J., and Zentilli, M., 2004, New chronology for El Teniente, Chilean Andes,
from U-Pb, 40Ar-39Ar , Re-Os, and fission-track dating: Implications for the
evolution of a supergiant porphyry Cu-Mo deposit: Society of Economic
Geologists Special Publication 11, p. 1554.
Maksaev, V., Munizaga, F., Fanning, M., Palacios, C., and Tapia, J., 2006,
SHRIMP U-Pb dating of the Antucoya porphyry copper deposit: New evidence for an early Cretaceous porphyry-related metallogenic epoch in the
Coastal Cordillera of northern Chile: Mineralium Deposita, v. 41, p.
637644.
Maksaev, V., Munizaga, F., Mathur, R., Barra, F., Fanning, M., McWilliams,
M., Ruiz, J., and Sanhueza, A., 2009, Geochronology of the Collahuasi porphyry Cu-Mo district, north Chilean Andes [ext. abs.]: XII Congreso Geolgico Chilen, Santiago, Extended Abstracts (CD), 4 p.
Maksaev, V., Almonacid, T., Munizaga, F., Valencia, V., McWilliams, M., and
Barra, F, 2010, Geochronological and thermochronological constraints on
porphyry copper mineralization in the Domeyko alteration zone, northern
Chile: Andean Geology, v. 37, p. 144176.
Mamani, M., Wrner, G., and Sempere, T., 2010, Geochemical variations in
igneous rocks of the Central Andean orocline (13S to 18S): Tracing
crustal thickening and magma generation through time and space: Geological Society of America Bulletin, v. 122, p. 162182.
Manea, V., Prez-Gussiny, M., and Manea, M., 2012, Chilean flat slab subduction controlled by overriding plate thickness and trench rollback: Geology, v. 40, p. 3538.
Mann, P., 1997, Model for the formation of large transtensional basins in
zones of tectonic escape: Geology, v. 25, p. 211214.
Marinovic, N., and Garca, M., 1999, Hoja Pampa Unin, Regin de Antofagasta: Santiago, Servicio Nacional de Geologa y Minera, Mapas Geolgicos, N 9 (escala 1:100,000).
Marinovic, N., Smoje, I., Maksaev, V., Herv, M., and Mpodozis, C., 1995,
Hoja Aguas Blancas, Regin de Antofagasta: Santiago, Servicio Nacional de
Geologa y Minera, Carta Geolgica de Chile N 70 (1:250,000), 150 p.
Marquardt, J.C., Rojo, J., Rivera, S., and Pizarro, J., 2009 Geologa del prfido cuprfero Miranda: nuevo descubrimiento en el Cluster Toki, Distrito
de Chuquicamata, Chile [ext. abs.]: XII Congreso Geolgico Chileno, Santiago, Extended Abstracts (CD), 4 p.
Marschik, R., and Fontbot, L., 2001, The Candelaria-Punta del Cobre iron
oxide Cu-Au(-Zn-Ag) deposits, Chile: Economic Geology, v. 96, p. 17991826.
Martinod, J., Funiciello, F., Faccenna, C., Labanieh, S., and Rgard, V., 2005,
Dynamical effects of subducting ridges: Insights from 3-D laboratory models: Geophysical Journal International, v. 163, p. 11371150.
Martinod, J., Husson, L, Roperch, P., Guillaume, B., and Espurt, N., 2010,
Horizontal subduction zones, convergence velocity and the building of the
Andes: Earth and Planetary Science Letters, v. 299, p. 299309.
Maydagn, L., Franchini, M., Chiaradia, M., Pons, J., Impiccini, A., Toohey,
J., and Rey, R., 2011, Petrology of the Miocene igneous rocks in the Altar
region, main Cordillera of San Juan, Argentina. A geodynamic model
within the context of the Andean flat-slab segment and metallogenesis:
Journal of South American Earth Sciences, v. 32, p. 3840.
McInnes, B.I.A., Farley, K.A., Sillitoe, R.H., and Kohn B.P., 1999, Application of apatite (U-Th)/He thermochronometry to the determination of the
sense and amount of vertical fault displacement at the Chuquicamata porphyry copper deposit, Chile: Economic Geology, v. 94, p. 937948.
McQuarrie, N., 2002, Initial plate geometry, shortening variations, and evolution of the Bolivian orocline: Geology, v. 30, p. 867870.
2006, Revisiting shortening estimates along the Bolivian orocline: Implications of thermal heating, erosion and crustal flow on the development
of a high elevation plateau [abs.]: Geological Society of America Abstracts
with Programs, Specialty Meeting 2, Mendoza, p. 86.
0361-0128/98/000/000-00 $6.00

357

Mora, R, Alfaro, J., Salazar, G., and Gonzlez, R., 2009, Prfido cuprfero Mirador: el descubrimiento ms reciente en el Distrito Centinela [ext. abs.]:
XII Congreso Geolgico Chileno, Santiago, Extended Abstracts (CD), 4 p.
Mpodozis, C., and Kay, S.M., 1992, Late Paleozoic to Triassic evolution of the
Gondwana margin: Evidence from Chilean Frontal Cordilleran batholiths:
Geological Society of America Bulletin, v. 104, p. 9991014.
Mpodozis, C., and Perell, J., 2003, Porphyry copper metallogeny of the middle-Eocene-early Oligocene arc of western South America. Relationships
with volcanism and arc segmentation [ext. abs.]: X Congreso Geolgico
Chileno, Concepcin, Extended Abstracts (CD), 1 p.
Mpodozis, C., and Ramos, V.A., 1989, The Andes of Chile and Argentina:
Circum-Pacific Council for Energy and Mineral Resources Earth Sciences
Series, v. 11, p. 5990.
2008, Tectnica jursica en Argentina y Chile: extensin, subduccin
oblicua, rifting, deriva y colisiones?: Revista de la Asociacin Geolgica Argentina, v. 63, p. 481497.
Mpodozis, C., Marinovic, N., and Smoje, I., 1993a, Eocene left lateral strike
slip faulting and clockwise block rotations in the Cordillera de Domeyko,
west of the Salar de Atacama, northern Chile: International Symposium on
Andean Geodynamics, 2nd, Oxford, Proceedings, p. 225228.
Mpodozis, C., Marinovic, N., Smoje, I., and Cuitio, L., 1993b, Estudio geolgico-estructural de la Cordillera de Domeyko entre Cerro Limn Verde
y Sierra Mariposas, Regin de Antofagasta: Servicio Nacional de Geologa
y Minera [Chile], Santiago, Informe Registrado IR-93-04, 282 p.
Mpodozis, C., Arriagada, C., Basso, M., Roperch, P., Cobbold, P., and Reich,
M., 2005, Late Mesozoic to Paleogene stratigraphy of the Salar de Atacama
basin, Antofagasta, northern Chile: Implications for the tectonic evolution
of the central Andes: Tectonophysics, v. 399, p. 125154.
Mpodozis, C., Cembrano, J., and Mora, R., 2009a, Deformacin compresivaoblicua polifsica y prfidos cuprferos eocenos en el Sistema de Fallas de
Domeyko: la regin de Esperanza-Caracoles (Distrito Centinela) [ext. abs.]:
XII Congreso Geolgico Chileno, Santiago, Extended Abstracts (CD), 4 p.
Mpodozis, C., Brockway, H., Marquardt, C., and, Perell, J, 2009b, Geocronologa U/Pb y tectnica de la regin de Los Pelambres-Cerro Mercedario:
Implicancias para la evolucin cenozoica de los Andes del centro de Chile
y Argentina [ext. abs.]: XII Congreso Geolgico Chileno, Santiago, Extended Abstracts (CD), 4 p.
Mller, J., Kley, J., and Jacobshagen, V., 2002, Structure and Cenozoic kinematics of the Eastern Cordillera, southern Bolivia (21 S): Tectonics, v. 21,
1037, 24 p., doi:10.1029/2001TC001340.
Mnchmeyer, C., and Valenzuela, D., 2009, El yacimiento Telgrafo: un prfido cuprfero en etapa de exploracin avanzada en el Distrito Centinela
[ext. abs.]: XII Congreso Geolgico Chileno, Santiago, Extended Abstracts
(CD), 4 p.
Munizaga, F., Maksaev, V., Fanning, C.M., Giglio, S., Yaxley, G., and Tassinari, C.G., 2008, Late Paleozoic-Early Triassic magmatism on the western
margin of Gondwana: Collahuasi area, northern Chile: Gondwana Research, v. 13, p. 407427.
Muoz, J., Troncoso, R., Duhart, P., Crignola, P., Farmer, L., and Stern, C.R.,
2000, The relation of the mid-Tertiary coastal magmatic belt in south-central Chile to the late Oligocene increase in plate convergence rate: Revista
Geolgica de Chile, v. 27, p. 177203.
Muoz, M., Fuentes, F., Vergara, M., Aguirre, L., Nystrm, J.O., and Fraud,
G., 2006, Abanico East Formation: Petrology and geochemistry of volcanic
rocks behind the Cenozoic arc front in the Andean Cordillera, central Chile
(3350' S): Revista Geolgica de Chile, v. 33, p. 109140.
Nalpas, T., Hrail, G., Mpodozis, C., Riquelme, R., Clavero, J., and Dabard,
M P., 2005, Thermochronological data and denudation history along a transect between Chaaral and Pedernales (~26 S), North Chilean Andes:
Orogenic implications [ext. abs.]: International Symposium on Andean Geodynamics, 6th, Barcelona, Extended Abstracts, p. 548551.
Niemeyer, H., and Munizga, R., 2008, Structural control of the emplacement
of the Portrerillos porphyry copper, central Andes of Chile: Journal of
South American Earth Sciences, v. 26, p. 261270.
Niemeyer, H., and Urrutia, C., 2009, Transcurrencia a lo largo de la Falla
Sierra de Varas (Sistema de fallas de la Cordillera de Domeyko), norte de
Chile: Andean Geology, v. 36, p. 3749.
Noble, D., McKee, E., and Mgard, F., 1979, Early Tertiary Incaic tectonism, uplift, and volcanic activity, Andes of central Peru: Geological Society
of America Bulletin, v. 90, p. 903907.
Oldow, J.S., Bally, A.W., and Av Lallemant, H.G., 1990, Transpression, orogenic float and lithospheric balance: Geology, v. 18, p. 991994.

357

358

MPODOZIS AND CORNEJO

Oncken, O., Hindle, D., Kley, J., Elger, P., Victor, P., and Schemmann, K.,
2006, Deformation of the Central Andean upper plate system: Facts, fiction, and constraints for plateau models, in Oncken, O., et al., eds., The
Andes: Frontiers in Earth Sciences, Pt. III, Springer-Verlag, p. 328.
Ossandn, G., Frraut, R., Gustafson, L., Lindsay, D., and Zentilli, M., 2001,
Geology of the Chuquicamata mine: A progress report: Economic Geology,
v. 96, p. 240270.
Oyarzn, R., Mrquez, A., Lillo, J., Lpez, I., and Rivera, S., 2001, Giant versus small porphyry copper deposits of Cenozic age in northern Chile:
Adakitic versus normal calc-alkaline magmatism: Mineralium Deposita, v.
32, p. 794798.
Pananont, P., Mpodozis, C., Jordan, T., Blanco, N., and Brown, L., 2004,
Cenozoic evolution of the northwestern Salar de Atacama basin, northern
Chile: Tectonics, v. 23, TC6007, 19 p., doi:10.1029/2003TC001595.
Pankhurst, R.J., Rapela, C.W., Saavedra, J., Baldo, E., Dahlquist, J., Pascua,
I., and Fanning, M., 1998, The Famatinian magmatic arc in the central
Sierras Pampeanas: An Early to Mid-Ordovician continental arc on the
Gondwana margin: Geological Society Special Publication 142, p. 343367.
Pardo-Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Farallon) and South American plates since late Cretaceous time: Tectonics, v. 6,
p. 233248.
Perell, J., 2003, Conchi porphyry copper deposit, Antofagasta region, northern Chile [ext. abs.]: X Congreso Geolgico Chileno, Concepcin, Extended Abstracts (CD), 1 p.
Perell, J., Urza, F., Cabello, J., and Ortiz, F., 1996, Clustered gold-bearing
Oligocene porphyry copper and associated epithermal mineralization at La
Fortuna, Vallenar region, Chile: Society of Economic Geologists Special
Publication 5, p. 8190.
Perell, J., Cox, D., Garamjav, D., Sandorj, S, Diakov, S., Schissel, D.,
Munkhbat, T., and Oyun, G., 2001, Oyu Tolgoi, Mongolia: Siluro-Devonian
porphyry Cu-Au-(Mo) and high-sulfidation Cu mineralization with a Cretaceous chalcocite blanket: Economic Geology, v. 96, p. 14071428.
Perell, J., Carlotto, V., Zrate, A., Ramos, P., Posso, H., Neyra, C., Caballero,
A., Fuster, N., and Muhr, R., 2003a, Porphyry-style alteration and mineralization of the middle Eocene to early Oligocene Andahuaylas-Yauri belt,
Cuzco region, Peru: Economic Geology, v. 98, p. 15751605.
Perell, J., Martini, R., Arcos, R., and Muhr, R., 2003b, Buey Muerto: Porphyry copper mineralization in the Early Cretaceous arc of northern Chile
[ext. abs.]: X Congreso Geolgico Chileno, Concepcin, Extended Abstracts (CD), 1 p.
Perell, J., Brockway, H., and Martini, R., 2004, Discovery and geology of
the Esperanza porphyry copper-gold deposit, Antofagasta region, northern Chile: Society of Economic Geologists Special Publication 11, p.
167186.
Perell, J., Razique, A., Schloderer, J., and Asad, R., 2008, The Chagai porphyry copper belt, Baluchistan Province, Pakistan: Economic Geology, v.
103, p. 15831612.
Perell, J., Muhr, R., Mora, R., Martnez, E., Brockway, H., Swaneck, T.,
Artal, J., Mpodozis, C., Mncheyer, C., Clifford, J., Acua, E., Valenzuela,
D., and Argandoa, R., 2010, Wealth creation through exploration in a mature terrain: The case history of the Centinela district, northern Chile: Society of Economic Geologists Special Publication 15, p. 229252.
Perell, J., Sillitoe, R.H., Mpodozis, C., Brockway, H., and Posso, H., 2012,
Geologic setting and evolution of the porphyry copper-molybdenum and
copper-gold deposits at Los Pelambres, central Chile: Society of Economic
Geologists Special Publication 16, p. 79104.
Prez, D.J., 2001, Tectonic and unroofing history of Neogene Manantiales
foreland basin deposits, Cordillera Frontal (3230' S), San Juan Province,
Argentina: Journal of South American Earth Sciences, v. 14, p. 693705.
Perry, V.D., 1952, Geology of the Chuquicamata orebody: Mining Engineering, v. 4, p. 11661168.
Ramos, V.A., 2009, Anatomy and global context of the Andes: Main geologic
features and the Andean orogenic cycle: Geological Society of America
Memoir 204, p. 3165.
Ramos, V.A., and Dalla Salda, L., 2011, Occidentalia: Un terreno acrecionado
sobre el margen gondwnico? [ext. abs.]: XVIII Congreso Geolgico Argentino, Neuqun, Extended Abstracts , p. 222223.
Ramos, V., and Folguera, A., 2005, Tectonic evolution of the Andes of
Neuqun: Constraints derived from the magmatic arc and foreland deformation: Geological Society Special Publication 252, p. 1535.
2009, Andean flat-slab subduction through time: Geological Society
Special Publication 327, p. 3154.
0361-0128/98/000/000-00 $6.00

Ramos, V., and Kay, S. M, 1991, Triassic rifting and associated basalts in the
Cuyo Basin, central Argentina, Geological Society of America, Special
paper 265, p. 7991.
Ramos, V.A., Jordan, T.E., Allmendinger, R.W., Mpodozis, C., Kay, S.M.,
Corts, J.M., and Palma, M.A., 1986, Paleozoic terranes of the Central Argentine-Chilean Andes: Tectonics, v. 5, p. 855880.
Ramos, V.A., Cegarra, M., and Cristallini, E., 1996, Cenozoic tectonics of the
High Andes of west-central Argentina (3036 S latitude): Tectonophysics,
v. 259, p. 185200.
Ramos, V.A., Cristallini, E., and Prez, D., 2002, The Pampean flat slab of the
Central Andes: Journal of South American Earth Sciences, v. 15, p. 5978.
Ramos, V.A., Zapata, T., and Cristallini, E.O., 2004, The Andean thrust system: Structural styles and orogenic shortening: American Association of Petroleum Geologists Memoir 82, p. 3050.
Reich, M., Parada, M.A., Palacios, C., Dietrich, A., Schultz, F., and
Lehmann, B., 2003, Adakite-like signature of late Miocene intrusions at the
Los Pelambres giant porphyry copper deposit in the Andes of central Chile:
Metallogenic implications: Mineralium Deposita, v. 38, p. 876885.
Reutter, K.-J., Scheuber, E., and Helmcke, D., 1991, Structural evidence of
orogen-parallel strike slip displacements in the Precordillera of northern
Chile: Geologische Rundschau, v. 80, p. 135153.
Reutter, K.-J., Scheuber, E., and Chong, G., 1996, The Precordilleran fault
system of Chuquicamata, northern Chile: Evidence for reversals along arcparallel strike-slip faults: Tectonophysics, v. 259, p. 213228.
Richards, J.P., 2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au)
deposit formation: Economic Geology, v. 98, p. 15151533.
2009, Post-subduction porphyry Cu-Au and epithermal Au deposits:
Products of remelting of subduction-modified lithosphere: Geology, v. 37,
p. 247250.
2011a, Magmatic to hydrothermal metal fluxes in convergent and collided margins: Ore Geology Reviews, v. 40, p. 126.
2011b, High Sr/Y magmas and porphyry Cu Mo Au deposits. Just
add water: Economic Geology, v. 106, p. 10751081.
Richards, J.P., and Kerrich, R., 2007, Adakite-like rocks: Their diverse origins
and questionable role in metallogenesis: Economic Geology, v. 102, p.
537576.
Richards, J.P., Boyce, A.J., and Pringle, M.S., 2001, Geologic evolution of the
Escondida area, northern Chile: A model for spatial and temporal localization of porphyry copper mineralization: Economic Geology, v. 96, p.
271305.
Roperch, P., Sempere, T., Macedo, O., Arriagada, C., Fornari, M., Tapia, C.,
Garca, M., and Laj, C., 2006, Counterclockwise rotation of late Eocene
Oligocene fore-arc deposits in southern Peru and its significance for oroclinal bending in the central Andes: Tectonics, v. 25, TC3010, 29 p.,
doi:10.1029/2005TC001882.
Roperch, P., Carlotto, V., Ruffet, G., and Fornari, M., 2011, Tectonic rotations and transcurrent deformation south of the Abancay Deflection in the
Andes of southern Peru: Tectonics, v. 30, TC2010, 23 p., doi:10.1029/
2010TC002725.
Rosas, S., Fontbot, L., and Tankard, A., 2007, Tectonic evolution and paleogeography of the Mesozoic Pucar basin, central Peru: Journal of South
American Earth Sciences, v. 24, p. 124.
Russo, R., and Silver, P., 1996, Cordillera formation, mantle dynamics, and
the Wilson cycle: Geology, v. 24, p. 511514.
Sandeman, H.A., Clark, A.H., and Farrar, E., 1995, An integrated tectonomagmatic model for the evolution of the southern Peruvian Andes (1320
S) since 55 Ma: International Geology Review, v. 37, p. 10391073.
Schellart, W.P., 2008, Overriding plate shortening and extension above subduction zones: A parametric study to explain formation of the Andes Mountains: Geological Society of America Bulletin, v. 120, p. 14411454.
Schellart, W.P., and Rawlinson, B, 2010, Convergent plate margin dynamics:
New perspectives from structural geology, geophysics and geodynamic
modeling: Tectonophysics, v. 483, p. 419.
Scheuber E., and Gonzlez, G., 1999, Tectonics of the Jurassic-Early Cretaceous magmatic arc of the north Chilean Coastal Cordillera (2226 S): A
story of crustal deformation along a convergent plate boundary: Tectonics,
v. 18, p. 895910.
Scheuber, E., and Reutter, K.-J., 1992, Magmatic arc tectonics in the central
Andes between 21 and 25 S: Tectonophysics, v. 205, p. 127140.
Sdrolias, M., and Muller, R.D., 2006, Controls on back-arc basin formation:
Geochemistry, Geophysics, Geosystems v. 7 Q04016, 40 p., doi:10.1029/
2005GC001090.

358

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES


Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe,
O., Neraudeau, D., Cardenas, J., Rosas, S., and Jimenez, N., 2002, Late
Permian-Middle Jurassic lithospheric thinning in Peru and Bolivia, and its
bearing on Andean-age tectonics: Tectonophysics, v. 345, p. 153158.
Seplveda, F., Lahsen, A., Bonvallot, S., Cembrano, J, Alvarado, A., and
Letelier, P., 2005, Morpho-structural evolution of the Cordn Caulle geothermal region, Southern volcanic zone, Chile: Insights from gravity and
40Ar/ 39Ar dating: Journal of Volcanology and Geothermal Research, v. 148,
p. 165189.
SERNAGEOMIN, 2002, Mapa Geolgico de Chile (1:1.000.000 scale).
Serrano, L., Vargas, R., Stambuk, V., Aguilar, C., Galeb, M., Holmgren, C.,
Contreras, A., Godoy, S., Vela, I., Skewes, M.A., and Stern, C.R., 1996, The
late Miocene to early Pliocene Rio Blanco-Los Bronces copper deposit,
Central Chilean Andes: Society of Economic Geologists Special Publication 9, p. 299332.
Shafiei, B, Haschke, M., and Shahabpour, J., 2009, Recycling of orogenic arc
crust triggers porphyry Cu mineralization in Kerman Cenozoic arc rocks,
southeastern Iran: Mineralium Deposita, v. 44, p. 265283.
Sibson, R.H., 1987, Earthquake rupturing as a mineralizing agent in hydrothermal systems: Geology, v. 17, p. 701704.
1994, Crustal stress, faulting and fluid flow: Geological Society Special
Publication 78, p. 6984.
Sillitoe, R.H., 1977, Permo-Carboniferous, Upper Cretaceous and Miocene
porphyry copper-type mineralization in the Argentinian Andes: Economic
Geology, v. 72, p. 99103.
1998, Major regional factors favouring large size, high hypogene grade,
elevated gold content and supergene oxidation and enrichment of porphyry
copper deposits, in Porter, T.M., ed., Porphyry and hydrothermal copper
and gold deposits: A global perspective: Adelaide, Australian Mineral
Foundation, p. 2134.
2003, Iron oxide-copper-gold deposits: An Andean view: Mineralium
Deposita, v. 38, p. 787812.
2010, Porphyry copper systems: Economic Geology, v. 105, p. 341.
Sillitoe, R.H., and Perell, J., 2005, Andean copper province: Tectonomagmatic settings, deposit types, metallogeny, exploration, and discovery: Economic Geology 100th Anniversary Volume, p. 845890.
Silva, C., Herrera, C., and Herv, F., 2003, Petrognesis de lavas y diques bsicos de la Formacin Traigun, (43 30'46 S, Chile) [ext. abs.]: X Congreso Geolgico Chileno, Concepcin, Extended Abstracts (CD), 11 p.
Silver, P.G., Russo, R.M., and Lithgow-Bertelloni, C., 1998, Coupling of
South American and African Plate motions and plate deformation: Science,
v. 279, p. 6063.
Skewes, M.A., and Stern C.R., 1994, Tectonic trigger for the formation of late
Miocene Cu-rich breccia pipes in the Andes of central Chile: Geology, v.
22, p. 551554.
Skewes, M.A., Holmgren, C., and Stern, C.R., 2003, The Donoso copper
rich, tourmaline-bearing breccia pipe in central Chile: Petrologic, fluid inclusion and stable isotope evidence for an origin from magmatic fluids:
Mineralium Deposita, v. 38, p. 221.
Solomon, M., 1990, Subduction, arc reversals and the origin of porphyry copper-gold deposits in island arcs: Geology, v. 18, p. 630633.
Somoza, R., 1998, Updated Nazca (Farallon)-South America relative motions
during the last 40 My: Implications for mountain building in the Central Andean region: Journal of South American Earth Sciences, v. 11, p. 211215.
Somoza, R., and Ghidella, M., 2005, Convergencia en el margen occidental
de Amrica del Sur durante el Cenozoico: subduccin de las placas de
Nazca, Faralln y Aluk: Revista de la Asociacin Geolgica Argentina, v. 60,
p. 797809.
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. 267277.
Soto, R., Martinod, J., Riquelme, R., Hrail, G., and Audin, L., 2005, Using
gemorphological markers to discriminate Neogene tectonic activity in the
Precordillera of North Chilean forearc (2425 S): Tectonophysics, v. 411,
p. 4155.
Stern, C.R., 1991, Role of subduction erosion in the generation of Andean
magmas: Geology, 19, p. 7881.
2011, Subduction erosion: Rates, mechanisms, and its role in arc magmatism and the evolution of the continental crust and mantle: Gondwana
Reserach, v. 20, p. 284338.
Stern, C.R., Skewes, M.A., and Arvalo, A., 2010, Magmatic evolution of the
giant El Teniente Cu-Mo deposit, central Chile: Journal of Petrology, v. 52,
p. 15911617. doi: 10.1093/petrology/egq029.
0361-0128/98/000/000-00 $6.00

359

Stern, C.R., Floody, R., and Espieira, D., 2011, Olivine-hornblende-lamprophyre dykes from Quebrada los Sapos, El Teniente, Central Chile (34
S): Implications for the temporal geochemical evolution of the Andean subarc mantle: Andean Geology, v. 38, p. 122.
Sun, W., Zhang, H., Ling, M.X., Ding, X., Chung, S.L., Zhou, J., Yang, X.Y.,
and Fan, W., 2011, The genetic association of adakites and Cu-Au ore deposits: International Geology Review, v. 53, p. 691703.
Swaneck, T., Mora, R., and Martnez, E., 2009, Caracoles: un nuevo prfido
de Cu-Au-Mo en el Distrito Centinela [ext. abs.]: XII Congreso Geolgico
Chileno, Santiago, Extended Abstracts (CD), 4 p.
Taylor, G.K., Dashwood, B., and Grocott, J., 2005, Central Andean rotation
pattern: Evidence from paleomagnetic rotations of an anomalous domain in
the forearc of northern Chile: Geology, v. 33, p. 777780.
Thiblemont, D., Steins, G., and Lescuyer, J.L., 1997, Gisements epithermaux et porphyriques : la connexion adakite: Comptes Rendues, Acadmie
des Sciences de Paris, Earth and Planetary Sciences, v. 325, p. 103109.
Tikoff, B., Russo, R., Teyssier, Ch., and Tommasi, A., 2002, Clutch tectonics
and the partial attachment of lithospheric layers: EGU Stephan Mueller
Special Publication Series, v. 1, p. 5773.
Tomlinson, A.J., and Blanco, N., 1997a, Structural evolution and displacement history of the West fault system, Precordillera, Chile: Pt. 1. Synmineral history [ext. abs.]: VIII Congreso Geolgico Chileno, Antofagasta, Extended Abstracts, v. 3, p. 18731878.
1997b, Structural evolution and displacement history of the West fault
system, Precordillera, Chile: Pt. 2. Postmineral history [ext. abs.]: VIII
Congreso Geolgico Chileno, Antofagasta, Extended Abstracts, v. 3, p.
18781882.
Tomlinson, A.J., and Cornejo, P., 2012, Regional distribution of Middle
Eocene-early Oligocene porphyry copper centers in northern Chile: Second order patterns and possible causes: XIII Congreso Geolgico Chileno,
in press.
Tomlinson, A.J., Mpodozis C., Cornejo, P., and Ramrez, C.F., 1993, Structural geology of the Sierra Castillo-Agua Amarga fault system, Precordillera
of Chile, El Salvador-Potrerillos: II International Symposium on Andean
Geodynamics, Oxford, Proceedings, p. 259262.
Tomlinson, A.J., Blanco, N., Maksaev, V., Dilles, J.H., Grunder, A.L., and Ladino, M., 2001a, Geologa de la Precordillera Andina de Quebrada BlancaChuquicamata, Regiones I y II (2030'2230' S): Servicio Nacional de
Geologa y Minera, Santiago, Informe Registrado IR-01-20, 444 p.
Tomlinson, A.J., Dilles, J.H., and Maksaev, V., 2001b, Application of apatite
(U-Th)/He thermochronometry to the determination of the sense and
amount of vertical fault displacement at the Chuquicamata porphyry copper deposit, Chilea discussion: Economic Geology, v. 96, p. 13071309.
Tomlinson, A.J., Blanco, N., and Dilles, J., 2010, Carta Calama, Regin de
Antofagasta: Servicio Nacional de Geologa y Minera, Santiago, Carta Geolgica de Chile, Serie Preliminar, 8 (1:50,000).
Toro, J.C., Ortzar, J., Maksaev, V., and Barra, F., 2009, Nuevos antecedentes geogronolgicos franja de prfidos Cu-Mo del mioceno-plioceno, Chile
central: Implicancias metalognicas [ext. abs.]: XII Congreso Geolgico
Chileno, Santiago, Extended Abstracts (CD), 4 p.
Toro, J.C., Ortzar, J., Zamorano, J., Cuadra, P., Hermosilla, J., and Sprhnle,
C., 2012, Protracted magmatic-hydrothermal history of the Ro Blanco-Los
Bronces district, central Chile: Development of worlds greatest known
concentration of copper: Society of Economic Geologists, Special Publication 16, p. 105126.
Tosdal, R., Dilles, J., and Cooke, D.R., 2009, From source to sinks in auriferous magmatic-hydrothermal porphyry and epithermal deposits: Elements, v. 5, p. 289295.
Trumbull, R., Riller, U., Oncken, O., Scheuber, E., Munier, K., and Hong, F.,
2006, The time-space distribution of Cenozoic volcanism in the south-central Andes: A new data compilation and some tectonic implications, in Oncken, O., et al., eds., The Andes: Frontiers in Earth Sciences, Pt. III:
Springer-Verlag, p. 2943.
Urza, F., 2009, Geology, geochronology and structural evolution of La Escondida copper district, northern Chile: Unpublished Ph.D. thesis, Hobart,
Australia, University of Tasmania, 486 p.
Uyeda, S., and Kanamori, H., 1979, Back-arc opening and the mode of subduction: Journal of Geophysical Research, v. 84, p. B1049B1061.
Valencia-Moreno, M., Ochoa-Landn, L., Noguez-Alcntara, B., Ruiz, J., and
Prez-Sgur, E., 2007, Geological and metallogenetic characteristics of the
porphyry copper deposits of Mxico and their situation in the world context: Geological Society of America Special Paper 422, p. 438458.

359

360

MPODOZIS AND CORNEJO

Veevers, J.J., 1989, Mid Triassic (230 5 Ma) singularity in the stratigraphic
and magmatic history of the Pangean heat anomaly: Geology, v. 17, p.
784787.
von Huene, R., and Scholl, D.W., 1991, Observations at convergent margins
concerning sediment subduction, subduction erosion, and the growth of
continental crust: Reviews of Geophysics, v. 29, p. 279316.
Vry, V.H., Wilkinson, J.J., Seguel, J., and Milln, J., 2010, Multistage intrusion, brecciation, and veining at El Teniente, Chile: Evolution of a nested
porphyry system: Economic Geology, v. 105, p. 119153.
Warnaars, F.W., Holmgren, C., and Barassi, S., 1985, Porphyry copper and
tourmaline breccias at Los Bronces-Ro Blanco, Chile: Economic Geology,
v. 80, p. 15441565.
Willner, A.P., Thomson, S.N., Krner, A., Wartho, J.A., Wijbrans, J., and
Herv, F., 2005, Time markers for the evolution and exhumation history of
a late Paleozoic paired metamorphic belt in central Chile (343530' S):
Journal of Petrology, v. 46, p. 18351858.
Wilson, J., Zentilli, M., Boric, R., Daz, J., and Maksaev, V., 2011, Geochemistry of the Triassic and Eocene igneous host rocks of the MMH porphyry
copper deposit, Chuquicamata district, Chile: SGA Biennial Meeting, 11th,
Antofagasta, Proceedings, p. 426428.
Woodcock, N.H., 1986, The role of strike-slip fault systems at plate boundaries: Philosophical Transactions of the Royal Society of London, Series A,
v. 317, p. 1329.

0361-0128/98/000/000-00 $6.00

Wotzlaw, J., Decou, A., von Eynatten, H., Wrner, G., and Frei, D., 2011,
Jurassic to Palaeogene tectono-magmatic evolution of northern Chile and
adjacent Bolivia from detrital zircon U-Pb geochronology and heavy mineral provenance: Terra Nova, v. 23, p. 399406.
Yaez, G., and Cembrano, J., 2004, Role of viscous plate coupling in the Late
Tertiary Andean tectonics: Journal of Geophysical Research, v. 109,
B02407, 21p., doi:10.1029/2003JB002494.
Yaez, G., and Maksaev, V., 1994, Sobre la distribucin espacial de cuerpos
intrusivos asociados a diapirismo y la caracterizacin de la fuente magmtica en prfidos cuprferos: Inestabilidades gravitacionales en un medio viscoso [ext. abs.]: VII Congreso Geolgico Chileno, Antofagasta, Extended
Abstracts, v. 2, p. 16421646.
Yaez, G., Ranero, C., von Heune, R., and Daz, J., 2001, Magnetic anomaly
interpretation across the southern Central Andes (3234 S): the role of
the Juan Fernndez Ridge in the Late Tertiary evolution of the margin:
Journal of Geophysical Research, v. 106, p. 62356345.
Zengqian, H., Hongwen, M., Zaw, K., Yuquan, Z., Mingjie, W., Zeng, W.,
Guitang, P., and Renli, T., 2003, The Himalayan Yulong porphyry copper
belt: Product of large-scale strike-slip faulting in eastern Tibet: Economic
Geology, v. 98, p. 125145.

360

Вам также может понравиться