ch 14 mpodozis , cornejo , 2012, cenozoic tectonics and porphyry copper systemes of the chilean...

32
329 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 Perelló, 2005; Richards, 2009, 2011a; Tosdal et al., 2009). Some Cenozoic porphyry copper deposits are known to have formed during or shortly after continent-continent, conti- nent-island arc, or island arc-island arc collisions in the Hi- malayas-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 (Perelló et al., 2001). In contrast, other large porphyry deposits such as Bingham Canyon in the western United States formed during the earliest stages of 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 sub- duction-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 Ti- tley, 1998; Valencia-Moreno et al., 2007; Perelló 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 Perelló, 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 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 (60-50 Ma) in southern Peru, and mid-Eocene-early Oligocene (43-32 Ma) and late Miocene-Pliocene (10-6 Ma) in north and central Chile. Although the tectonic setting of the Paleocene porphyry deposits is still poorly understood, those of the northern Chile Eocene- Oligocene 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 conver- gence rates recorded for the Eocene-Oligocene suggest, however, that strong interplate coupling existed dur- ing that time. This permitted the transfer of horizontal stresses and large-scale deformation of the Andean mar- gin, 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 (Aban- ico 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. Tec- tonic inversion of the basin and compressional deformation along with crustal thickening and mountain build- ing 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 fore- arc subduction erosion. In both cases, the largely structurally controlled, syn- to post-tectonic porphyry copper deposits are associ- ated 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 multi- ple events since the Cretaceous. Porphyry copper mineralization is linked to multistage, amphibole-bearing in- trusions of intermediate composition derived from hydrous, oxidized magmas with adakitic geochemical sig- natures. 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 en- hanced subduction erosion during peaks of compressional deformation. Corresponding author: e-mail, [email protected] © 2012 Society of Economic Geologists, Inc. Special Publication 16, pp. 329–360

Upload: georamone

Post on 13-Apr-2016

154 views

Category:

Documents


5 download

DESCRIPTION

cenozoic tectonics and porphyry copper systemes of the chilean andes,

TRANSCRIPT

329

IntroductionTHE STUDY of the tectonic setting of porphyry copper depositsis fundamental to understanding their genesis (e.g., Sillitoe,1998; Kay and Mpodozis, 2001; Cooke et al., 2005; Sillitoeand Perelló, 2005; Richards, 2009, 2011a; Tosdal et al., 2009).Some Cenozoic porphyry copper deposits are known to haveformed during or shortly after continent-continent, conti-nent-island arc, or island arc-island arc collisions in the Hi-malayas-Tibet, the Kerman arc in Iran, and Papua NewGuinea (Solomon, 1990; Zenqiang et al., 2003; Shaifei et al.,2009). A Paleozoic example of this type of deposit may beOyu Tolgoi in Mongolia (Perelló et al., 2001). In contrast,other large porphyry deposits such as Bingham Canyon in thewestern United States formed during the earliest stages of

Basin and Range extension in the Eocene, far inland from thePacific margin of North America (Kloppenburgh et al., 2010).

Noncollisional porphyry copper deposit examples in sub-duction-related arc settings include those from the Chagaibelt in Pakistan, the Laramide porphyry copper province ofthe western United States and northern Mexico (Lang and Ti-tley, 1998; Valencia-Moreno et al., 2007; Perelló et al., 2008)and the Central Andes province, which host some of thelargest known porphyry copper deposits in the world (Camus,2003; Cooke et al., 2005; Sillitoe and Perelló, 2005).

The Andes has long been considered as the type example ofa noncollisional orogenic system (e.g., Jordan et al., 1983),where subduction of Pacific oceanic crust beneath SouthAmerica has been active for the past 570 m.y. (Cawood,2005). Nevertheless, the largest porphyry copper deposits arethe result of anomalous magmatic systems that developed

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

AbstractSubduction under South America has been active for the past 550 m.y. but large porphyry copper deposits

were essentially emplaced during the Paleocene (60−50 Ma) in southern Peru, and mid-Eocene-earlyOligocene (43−32 Ma) and late Miocene-Pliocene (10−6 Ma) in north and central Chile. Although the tectonicsetting of the Paleocene porphyry deposits is still poorly understood, those of the northern Chile Eocene-Oligocene belt were emplaced along the margin-parallel Domeyko fault system, where active compressionaland/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 alonga mechanically weak zone of the Central Andean crust inherited from the Paleozoic. Deformation occurredduring an episode of accelerated westward absolute motion of the South American plate, which coincided withvery high rates of oceanic crust production in the eastern Pacific. The slow South American-Farallon conver-gence rates recorded for the Eocene-Oligocene suggest, however, that strong interplate coupling existed dur-ing that time. This permitted the transfer of horizontal stresses and large-scale deformation of the Andean mar-gin, creating a favorable scenario for the generation and emplacement of porphyry copper magmas along theDomeyko fault system.

The younger, Miocene-Pliocene porphyry copper deposits of central Chile-Argentina were emplaced in adifferent setting, after the initiation of compressional deformation within a volcano-tectonic depression (Aban-ico basin) that evolved during another, late Oligocene to early Miocene, period of increased East Pacificoceanic crust production. Nevertheless, in contrast to the Eocene-Oligocene situation in northern Chile, therelatively stationary position of the South American plate compared to the mantle reference frame and weakinterplate coupling that permitted rapid subduction, increased volcanism, and overriding plate extension. Tec-tonic inversion of the basin and compressional deformation along with crustal thickening and mountain build-ing began at around 20 m.y. ago as interplate coupling increased when the westward motion of South Americaaccelerated and the Nazca-South America convergence velocity decreased in the mid-Miocene. Compressionwas accompanied, as during the Eocene-Oligocene in northern Chile, by slab shallowing and increased fore-arc subduction erosion.

In both cases, the largely structurally controlled, syn- to post-tectonic porphyry copper deposits are associ-ated with long-lived magmatic systems that were active for more than 10 m.y. In northern Chile, the depositsoccur as parts of discrete intrusive clusters that comprise a suite of precursor plutons emplaced during multi-ple events since the Cretaceous. Porphyry copper mineralization is linked to multistage, amphibole-bearing in-trusions of intermediate composition derived from hydrous, oxidized magmas with adakitic geochemical sig-natures. These intrusions appeared when crustal thickness increased to a critical threshold in the course ofdeformation. Production of magmas with high metal-carrying capacity was fostered as fluids were liberatedwhen amphibole became unstable and was destroyed as the crust thickened. At the same time, source regionswithin the mantle were contaminated by hydrated fragments of fore-arc continental crust, as the result of en-hanced subduction erosion during peaks of compressional deformation.

† Corresponding author: e-mail, [email protected]

© 2012 Society of Economic Geologists, Inc.Special Publication 16, pp. 329–360

during short periods at specific locations within the Andeanorogen. These include the Paleocene to early Eocene (66−52Ma) and middle Eocene to early Oligocene (43−32 Ma) beltsin southern Peru and northern Chile, and the late Miocene toearly Pliocene (10−5 Ma) porphyry systems in central Chileand contiguous Argentina (Perelló et al., 2003a; Sillitoe andPerelló, 2005). In this contribution, with emphasis on theChilean belts, we will try to demonstrate how major Cenozoictectonic events along the central Andean convergent margin,prompted by large-scale reorganizations of the global tectonicsystem, were the main triggers for the formation of large por-phyry copper deposits.

Pre-Andean History: From Rodinia Dispersal to Pangea Breakup

The western margin of South America underwent mag-matic and tectonic activity at least since the late Neoprotero-zoic breakup of Rodinia (800−700 Ma), when the separationof Laurentia from Gondwana produced the opening of theproto-Pacific (Iapetus) ocean (Dalziel, 1997). East-directedsubduction of newly formed ancestral Pacific crust belowwestern Gondwana began at ~570 Ma and was fully activealong the proto-Andean margin by 485 to 465 Ma (Pankhurstet al., 1998; Cawood, 2005; Chew et al., 2007). Plate conver-gence in the Central Andes region during the Ordovician toDevonian included the progressive collision and accretion ofa group of tectonostratigraphic terranes of Laurentian and/orGondwanan affinities (e.g., Ramos et al., 1986; Astini et al.,1995) against the western South American margin. Terraneamalgamation contributed to the formation of the accre-tionary Terra Australis orogen, which extended for more than18,000 km along the Pacific margin of Gondwana from Aus-tralia to South America (Cawood, 2005). The accretionarystage was followed, in the Central Andes, by the buildup of alate Carboniferous to Early Permian (320? -280 Ma) supra-subduction magmatic arc on top of the newly accreted terranes,as well as the development of an outboard fore-arc subduc-tion complex that extended for more than 1,000 km along theChilean segment of the Gondwana margin south of 27° S(Mpodozis and Kay, 1992; Hervé, 1988; Willner et al., 2005;Chew et al., 2007).

Magmatism continued from the Permian to the Middle Triassic (280−240 Ma), when great volumes of intrusive andmostly felsic volcanic rocks, including the Choiyoi large ig-neous province in Chile and Argentina and the Mitu Group insouthern Peru (Kay et al., 1989, Sempere et al., 2002), wereemplaced along the western South American margin. Althoughgeochronologic and geochemical data are still incomplete,several competing hypotheses, such as normal or oblique sub-duction, postcollision extension-driven crustal melting, slabbreakoff or slab shallowing, have been proposed to explainthe prevailing tectonic regime along different segments of theAndean margin at that time (Mpodozis and Kay, 1992; Kleimanand Japas, 2009; Ramos and Folguera, 2009, and referencestherein). From the Middle Triassic to earliest Jurassic (240−190 Ma), rifting associated with the incipient stages of Pangeadispersal (e.g., Veevers, 1989), accompanied by a decreasingvolume of bimodal magmatism, seems to have occurredalong the western margin of South America (e.g., Ramos andKay, 1991; Franzese and Spalletti, 2001; Rosas et al., 2007).

Diverse, yet basically subeconomic, porphyry copper depositsformed during these events in northern Chile and along theFrontal Cordillera in west-central Argentina (Sillitoe, 1977;Sillitoe and Perelló, 2005; Cornejo et al., 2006; Munizaga etal., 2008).

Jurassic to Early Eocene Tectonics and Metallogeny of the Central Andes

After the Triassic rifting event, subduction was reestab-lished in northern Chile and southern Peru during the EarlyJurassic when a new magmatic arc developed west of the extinct late Paleozoic arc front. Since then, subduction hasproceeded uninterrupted to date. Initial Jurassic to EarlyCretaceous arc magmatism occurred under extensional con-ditions that permitted the formation of a series of intercon-nected back-arc basins to the east of the main arc, which wereprogressively filled with marine and continental sedimentarystrata (Mpodozis and Ramos, 1989, 2008). Transpressionaldeformation along the arc axis created the intra-arc Atacamafault system in northern Chile (Scheuber and González,1999) and was accompanied in the Early Cretaceous by theemplacement, in northern Chile, of some porphyry copperdeposits at ~140 to 130 Ma (e.g., Antucoya-Buey Muerto,141−139 Ma; Puntillas-Galenosa, 135−132 Ma; Perelló et al.,2003b; Maksaev et al., 2006, 2010). Fast convergence ratesduring the global mid Cretaceous superplume event (Larson,1991) produced an upsurge in volcanism along the Andeanmargin, accompanied by intra-arc extension and transtensionwhich fostered iron oxide-copper-gold (IOCG)−type mineral-ization between 120 and 100 Ma in northern Chile and south-ern Peru (Marschik and Fontboté, 2001; Sillitoe, 2003; Silli-toe and Perelló, 2005; Chen et al., 2010). Small, low-grade,gold-rich porphyry copper deposits such as Andacollo (104Ma), Domeyko-Dos Amigos (108−104 Ma), and Pajonales (97Ma) were emplaced under extensional conditions during thesame general period in north-central Chile (Sillitoe andPerelló 2005; Maksaev et al., 2010).

The extensional and transtensional conditions that domi-nated early Andean subduction ended in the early Late Cre-taceous, when the back-arc basins were tectonically inverted(Mpodozis and Ramos, 1989; Tomlinson et al., 2001a). Theshift to a more contractional, subduction-related regime oc-curred together with the accelerated westward drift of SouthAmerica, in response to the final opening of the Atlantic(Russo and Silver, 1996; Somoza and Zaffarana, 2008). Sub-sequently, the coastal magmatic arc was abandoned (Fig. 1b)and the magmatic front jumped to the east in the Late Creta-ceous, where it remained relatively stationary until the earlyEocene. Abrupt shifts in the magmatic front such as this hasbeen accompanied throughout the Andean history, by tran-sient 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 etal. (2005) suggested that changes of this type reflect mantlecontamination from fore-arc crust removed during enhancedsubduction erosion processes associated with major contrac-tional events along the Andean margin.

In northern Chile, the Cretaceous-Tertiary boundary wasmarked by another short pulse of contractional deformation,the K-T event of Cornejo et al. (2003). Paleocene to early

330 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 330

Eocene magmatism included the emplacement of a porphyrycopper belt between southern Peru (Cerro Verde, 61 Ma;Quellaveco, 54 Ma; Cuajone, 52 Ma) and Cerro Colorado (52Ma), Spence (57 Ma), and Relincho (61 Ma) in northern Chile(Sillitoe and Perelló, 2005). Structural and geochemical datagathered by the authors suggest that these deposits in north-ern Chile formed in a neutral stress to mildly extensional arc

built on a relatively thin crust over a steep subduction zone.This is consistent with the very low rates of and oblique con-vergence between the South American and Farallon plates(Pardo-Casas and Molnar, 1987), and also with within-plategeochemical signatures of Paleocene to early Eocene rocks inthe Inca de Oro-El Salvador region (Cornejo and Matthews,2001).

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 331

0361-0128/98/000/000-00 $6.00 331

Late CretaceousPeruvian Event KT

Event

Sm/Yb

Eoc-E Olig.Incaic intrusions

Post Incaic intrusionsIncaic Event

90 80 70 60 50 40 30 20 10 0AGE (Ma)

26 Ma

Porphyry Cudeposits

IgnimbritesANTOFAGASTA

TRANSEC T(20 - 23ºS)

w10

20

30

40

50

60

E

15º

20º

25º

70º 65º

PERU

Arica

Antofagasta

CHILE

Salar deAtacama

La Paz

Altiplano

Puna

ARGEN

TINA

WC EC SA

SA

WC

EC

SP

Late Cretaceous-Eocene arc

Mid - Cretaceous arc

Jurassic - Early Cretaceousarcs

0

20

40

60

80

100

120

140

160

180

200(b)

AG

E (

Ma)

AG

E (

Ma)

w EModern CVZ Arc

Distal tuffs/reset ages

Longitude ( ºW )

72 71 70 69 68 67 66 65 64

Longitude ( ºW )

Volcanic andintrusive rocks

71 70 69 68 67 66 65(a)

(c) (d)

FA

FIG. 1. (a). Main morphotectonic units of the central Andes, between 15° and 30° S. FA = modern fore-arc zone, in-cluding 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 volcaniczone (CVZ); EC = Eastern Cordillera; SP = Sierras Pampeanas; SA = sub-Andean fold-and-thrust belt. (b). Relationship be-tween age and longitude (distance to the trench) for <200 Ma volcanic and intrusive rocks of the Central Andes (19°−28° S;Haschke et al., 2002, CVZ = modern Central volcanic zone of the Andes). (c). Longitude vs. age for Cenozoic intrusive andvolcanic rocks at 20° to 23° S (Iquique to Antofagasta transect; Trumbull et al., 2006). The ~32 to 26 Ma gap may be relatedto a transient episode of flat subduction produced as a consequence of the Incaic tectonic episode. The east to west migra-tion of the magmatic front during the Miocene is considered to record late-stage slab steepening (Kay et al., 1999; Kay andCoira, 2009). (d). Geochemical changes since the Late Cretaceous near 26° S (Copiapó-El Salvador region; data fromCornejo and Mathews, 2001). Note the short-lived peaks in the Sm/Yb ratio during major tectonic events, superimposed overa more subdued long-term trend to increased values. Peak values may reflect contamination of the mantle magma source re-gions as a result of massive removal of fore-arc continental crust during periods of enhanced subduction erosion (Kay andMpodozis, 2002; Kay et al., 2011).

Middle Eocene to Early Oligocene Tectonics of the Central Andes

The Domeyko fault system

One of the most relevant tectonic and metallogenicepisodes in the Central Andes correlates with the middleEocene to early Oligocene (45−33 Ma) Incaic tectonic event(Noble et al., 1979; Maksaev and Zentilli, 1988; Mpodozis andPerelló, 2003; Sillitoe and Perelló, 2005); during this period,bending of the continental margin generated the Bolivianorocline and the Domeyko fault system along the Pre-cordillera of northern Chile. The Domeyko fault system (Fig.2) is a >1,000-km-long, 40- to 60-km-wide, orogen-parallelzone of deformation composed of a complex array of strike-slip, normal, and reverse faults, together with thin- and thick-skinned folds and thrusts, which extends along the Cordillerade Domeyko (also known as Precordillera) in northern Chilebetween 20° and 27° S (e.g., Reutter et al., 1991, 1996;Cornejo et al., 1997). Some authors (e.g., Amilibia andSkarmeta, 2003; Amilibia et al., 2008) proposed that most ofthese faults and folds initiated during Late Cretaceous as aconsequence of the inversion of normal faults inherited fromthe Mesozoic back-arc extension. However, others (e.g., Tom-linson et al., 2001a; Mpodozis et al., 2005) interpreted thatthe Andean back-arc basins were first inverted during theearly Late Cretaceous to form a proto-Cordillera deDomeyko, while a second main tectonic pulse along theDomeyko fault system, coincident with the Incaic event, pro-duced its final uplift (Reutter et al., 1991, 1996; Scheuber andReutter 1992; Tomlinson et al., 1993; Maksaev and Zentilli,1999). Parts of the Domeyko fault system were subsequentlyreactivated during the Oligocene and the Quaternary (Tom-linson and Blanco, 1997a, b; Audin et al., 2003; Soto et al.,2005).

The kinematics of the middle Eocene to early Oligocenedeformation along the Domeyko fault system is a matter ofcontroversy; evidence for both left- and right-lateral displace-ments, including reversal in the sense of shear, has been re-ported along different parts of the faulted domain (Reutter etal., 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 Domeykowas exhumed between 40 and 30 m.y. ago (Maksaev and Zentilli, 1999; Nalpas et al., 2005) in association with surfacetectonic uplift and profound erosion, the products of whichaccumulated in syntectonic basins east and west of the area ofdeformation (Mpodozis et al., 2005; Hong et al., 2007; Wot-zlaw et al., 2011).

Origin of the Domeyko fault system

At first glance, the Domeyko fault system could be consid-ered as a trench-linked fault system (Woodcock, 1986) thatnucleated in the thermally weakened crust of the middleEocene to early Oligocene magmatic arc of northern Chileduring a period of suggested fast Eocene oblique conver-gence between the Farallon and South America plates (Pardo-Casas and Molnar, 1987; Somoza, 1998). However, when re-cent paleomagnetic and structural studies are taken intoaccount it becomes apparent that tectonic activation of theDomeyko fault system during the Incaic episode is essentially

a consequence of the formation of the sharp bend of the west-ern South American margin, known as the Arica elbow or Bo-livian orocline (Fig. 3). Paleomagnetic studies have been es-sential in obtaining a more constrained view of thedeformational history of this segment of the Central Andesand support the tectonic model for orocline formation firstproposed by Isacks (1988).

Figure 3a is a simplified regional map showing the distrib-ution of paleomagnetic (declination) vectors measured for theCentral Andes. Importantly, independent of age, Mesozoicand Paleogene rocks have been rotated up to 50°. In contrast,rotations measured in Miocene and younger rocks (<18−11Ma in northern Chile; <20 Ma in southern Peru) are negligi-ble, suggesting that most of the rotations were acquired dur-ing 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 re-gion (domain C) between Antofagasta and Arica (Taylor et al.,2005; Arriagada et al., 2008). The magnitude of the counter-clockwise rotations decreases significantly at the Abancay De-flection in central Peru (domain A; Fig. 3a), the latter breakbeing interpreted as a zone of intense Eocene-early Oligoceneleft-lateral shear along the boundary between the Arequipaand Paracas basement terranes (Ramos, 2009; Roperch et al.,2011; Fig. 4b). In northern Chile, clockwise rotations de-crease progressively south of Antofagasta and essentially dis-appear near Vallenar (28°30' S) upon entering the essentiallynonrotated domain E (Fig. 3a), which extends southward tothe latitude of Santiago (33° S).

In his landmark paper, Isacks (1988) proposed that the ob-served paleomagnetic rotations and the formation of the sea-ward concave Bolivian orocline are related to along-strikevariations in the amount of late Cenozoic shortening pro-duced during contractional deformation focused along a mechanically and thermally weakened zone located in theoverriding South American plate. Recent structural studiesindicate, however, that the bulk of the shortening (~60%), atand near the Arica bend (13°−19° S; Fig. 3), is pre-Neogenein age and occurred between 40 and 20 Ma (Lamb, 2001;Müller et al., 2002; Kley et al., 2005; McQuarrie, 2006). In-caic shortening concentrated within the Eastern Cordillera ofBolivia where >12 km of terrigenous sedimentary strata accu-mulated in a Paleozoic marine basin above highly attenuatedcontinental crust (Figs. 1, 3). The amount of horizontal short-ening reaches a maximum near the axis of the orocline, wherethe Paleozoic sedimentary sequence is thickest, and decreasessymmetrically along-strike (Oncken et al., 2006; Gotberg etal., 2010, and references therein, Fig. 4a) as the Paleozoicsedimentary rocks of the Eastern Cordillera become thinnerand give way to metamorphic and crystalline rocks in bothPeru (Cordillera de Marañón) and northwest Argentina (Sier-ras Pampeanas, see Fig. 3).

Arriagada et al. (2008) attempted to remove the combinedeffects of accumulated horizontal shortening and block rota-tions. Figure 3b shows their preferred solution for the restoredshape of the continental margin (Peru-Chile trench) at 45Ma, before the formation of the Bolivian orocline. As shownin Figure 4b, the ca. 30° E azimuth of the Farallon-South

332 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 332

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 333

0361-0128/98/000/000-00 $6.00 333

70º30' 67º30'

21º

22º30'

24º

25º30'

27º

70º30' 67º30'69º

27º

25º30'

24º

22º30'

21º

B O L I V I A

AR

GE

NT

INA

Chimborazo (41)

La Planada

CopaquireRosario (Collahuasi) (36-35)

Ujina (36-35)Quebrada Blanca (37-35)

Salar de Ascotán

El Abra (38-37)ConchiRadomiro Tomic (36-34)

Chuquicamata (36-32)Alejandro Hales (39-36)

Miranda (38-37)Toki (39)Opache (38-37)

CALAMA

Esperanza-Telégrafo (42-40)Caracoles (42-41)

Centinela (45-44)

Polo Sur (42-41)

Salarde

Atacama

Gaby (42)

Zaldívar (38-37)Escondida (38-37)

Salar de PuntaNegra

Sierra Juncal (40)

Sierra del Jardín (42)

El Salvador (42-41)

Exploradora (35)

Salar dePedernales

Potrerillos(36)

Salar deMaricunga

Eocene plutons (48-38 Ma)

Paleozoic basement

Reverse fault

Normal or strike-slip fault

N

ANTOFAGASTA

QC

CA

CE

LE

SE

PS

RF

YP

COPIAPO

Escondida Este (36-35)

0 100km

Salars

Middle Eocene to early Oligo-cene porphyry Cu Mo Au de-posits

+ +

FIG. 2. Sketch map of the Cordillera de Domeyko (or Precordillera) and the Domeyko fault system, showing main faultstraces, 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 Ex-ploradora-Juncal; PS = Potrerillos- El Salvador; RF = Río Figueroa) and the locations and ages (in parentheses, Ma) ofEocene-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).

334 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 334

Low

er P

aleo

zoic

sed

imen

tary

roc

ks

Low

er P

aleo

zoic

(Neo

pro

tero

zoic

?)m

etam

orp

hic

and

intr

usiv

e ro

cks

Mid

dle

-Eoc

ene

to e

arly

Olig

ocen

ep

orp

hyry

cop

per

bel

t

Pal

eom

agne

tic d

eclin

atio

n ve

ctor

Pal

eom

agne

tic d

omai

n

10º

18º

14º

Lim

aLa

s B

amb

as

Aric

a

La P

azS

anta

Cru

z

Chu

qui

cam

ata

La E

scon

did

a

Juju

y

Cop

iap

óS

ierr

asP

amp

eana

s

Ant

ofag

asta

Peru - Chile Trench

PE

RU

BOLIVI

A

ARG

ENTI

NA

Peru - Chile Trench

Tota

l dis

pla

cem

ent

vect

ors

(45-

0 M

a)

Restored posit

ion of

trench at 4

5 Ma

10º

18º

14º

22º

26º

El M

orro

Cor

dill

era

de

Mar

añón

Col

lahu

asi

AB

AN

CA

Y D

EFL

EC

TIO

N

E

D

C

B

A

D

BRAZILIAN

SHIELD

(a)

(b)

74º

70º

66º

62º

78º

22 º

BRAZILIAN

SHIELD

74º

70º

66º

62º

78º

250

km25

0 km

FIG

. 3.

(a).

Pale

omag

netic

dec

linat

ion

data

for

the

Cen

tral

And

es in

nor

ther

n C

hile

and

sou

ther

n Pe

ru th

at in

dica

te th

e C

entr

al A

ndea

n ro

tatio

n pa

tter

n. D

ata

from

Arr

iaga

da e

t al

. (20

06, 2

008)

, Rop

erch

et

al. (

2006

, 201

1), T

aylo

r et

al.

(200

5), a

nd r

efer

ence

s ci

ted

ther

ein.

Dis

trib

utio

n of

low

er P

aleo

zoic

sed

imen

ts o

f th

e B

oliv

ian

basi

n an

d cr

ysta

lline

and

met

amor

phic

roc

ks m

odifi

ed fr

om R

amos

and

Dal

la S

alda

(201

1; F

ig. 1

). A

lso

show

n ar

e th

e m

ain

Eoc

ene-

Olig

ocen

e po

rphy

ry c

oppe

r de

posi

tsof

the

Cen

tral

And

ean

belt

(Per

elló

et a

l., 2

003a

). R

otat

ion

dom

ains

des

igna

ted

A to

E a

re d

iscu

ssed

in th

e te

xt. (

b). P

ositi

on o

f the

pla

te b

ound

ary

(Per

u-C

hile

tren

ch)

at 4

5 M

a an

d to

tal (

Pale

ogen

e to

pre

sent

) di

spla

cem

ent v

ecto

rs fo

r m

ater

ial p

oint

s on

the

Sout

h A

mer

ican

mar

gin

acco

rdin

g to

the

two-

dim

ensi

onal

res

tora

tion

mod

elof

Arr

iaga

da e

t al.

(200

8).

America Eocene to Oligocene convergence vector calculatedby Somoza (1998, Fig. 4b) was nearly orthogonal to the (re-stored) NW-SE−trending section of the margin along thecenter and northern limb of the Bolivian orocline. Neverthe-less plate convergence was, at the same time, highly obliquealong the southern limb of the orocline fostering margin-parallel shear, and (theoretically dextral) strike-slip faulting innorthern Chile where the Domeyko fault system was formed(Fig. 4b). Bending of the margin seems to have been accom-panied at the southern limb of the Bolivian orocline by whole-sale NE-directed crustal flow (Fig. 3b), which is also requiredto explain the excess orogenic volume and crustal thicknessbelow the Altiplano-Puna reported by Kley and Monaldi(1998) and Hindle et al. (2005).

Mass transfer, likely associated with lower crustal flow to-ward the center of the orocline (Hindle et al., 2005), increasesthe possibility of strike-slip faulting along the Domeyko faultsystem. Continued displacement was likely initially blockednear the orocline axis where deformation was dominated bymargin-normal contraction (Figs. 3b, 4b). As suggested byMcQuarrie (2002) and Boutelier and Oncken (2010), masstransfer toward the core of the orocline implies crustal thin-ning and stretching in central Chile, which may have inhib-ited mountain building south of 28° S. Crustal thinning and amore 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 contrastshelp to explain the termination of the Incaic porphyry copper

belt at approximately 30° S and the possibly segmented na-ture of the belt along its strike length between southern Peruand central Chile (Mpodozis and Perelló, 2003; see below).

Middle Eocene to Early Oligocene Porphyry CopperProvince of Northern Chile

Overview

Middle Eocene to early Oligocene porphyry copper de-posits of the Central Andes were emplaced contemporane-ously with the Incaic tectonic event, when the entire Andeanmargin was being reshaped during the formation of the Aricabend. Mineralized centers occur in Peru near the eastern endof the Abancay Deflection (Andahuaylas-Yauri cluster; Perellóet al., 2003a; Fig. 3a), although the vast majority are locatedin Chile along the Domeyko fault system (Sillitoe and Perelló,2005; Figs. 2, 3a). El Morro, the southernmost porphyry copper-gold deposit of economic importance (Perelló et al.,1996), is located where the rotated and thickened southernlimb of the Bolivian orocline (domain D; Fig. 3a) terminatesand merges with the nonrotated central Chile domain E (seebelow). Most porphyry copper deposits were emplaced alongthe Domeyko fault system at long-lived zones of focused mag-matism, which in some cases were active well before theEocene. They occur as parts of discrete intrusive clusters separated by large barren areas where only the Paleozoicbasement and back-arc basin sedimentary cover is exposed(Fig. 2).

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 335

0361-0128/98/000/000-00 $6.00 335

Initial (non rotated)Domeyko fault

system

45 Ma Trench

PARACAS

TER

RAN

E

AREQUIPA

TERRANE

0 Ma Trench

BrazilianShield

Shear zone along

terrrane boundary

Mechanicallyweak crust(Paleozoicsedimentaryrocks

Fold andThrust

Belt

Domeykofault system

Orthogonal

convergence

Orthogonal

convergence

Oblique

convergence

(b)

?

21ºS

(a)

N S

Distance from orocline axis

1500 1000 500 0 500 1000 1500 km

Tecto

nic

sh

ort

en

ing

(km

)

Latitude (ºS)

0º 10º 15º 20º 25º 30º 35º

400

300

200

100

1

2

4

3b

3a

FIG. 4. (a). Different horizontal tectonic shortening estimates for the Central Andes, showing how values decrease sym-metrically 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 assum-ing initial crustal thickness of 40 km; 3b = same, but considering 35 km as initial thickness (Gotberg et al., 2010); 4 = short-ening needed to balance paleomagnetically determined rotations (Arriagada et al., 2008). (b). Tectonic sketch showing howoblique convergence along the southern limb of the orocline may drive strike-slip displacements along the Domeyko faultsystem 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. Noteleft-lateral shear along the Abancay Deflection along the boundary between the Arequipa and Paracas terranes.

The structural control of the porphyry systems is strikinglydiverse (cf. Sillitoe and Perelló, 2005). At Chuquicamata, por-phyry copper-related intrusions were syntectonically emplacedalong an extensional jog in an active dextral strike-slip faultsystem (Lindsay et al., 1995); at Potrerillos, intrusions relatedto porphyry copper deposits ascended along subvertical, left-lateral strike-slip faults and migrated laterally near the surfacealong an active low-angle thrust (Tomlinson et al., 1993;Niemeyer and Munizaga, 2008). In contrast, porphyry stocksrelated to mineralization at Escondida or El Abra-Fortunawere emplaced along premineralization reverse and strike-slipfaults formed during earlier stages of the Incaic event (Dilleset al., 2011; Hervé et al., 2012). Other mineralized intrusionswere cut and displaced by late-stage faults (Esperanza-Telé-grafo, Chuquicamata; Dilles et al., 1997; Tomlinson et al.,1997a, b).

Not all porphyry copper deposits are, however, obviouslyrelated to major Domeyko fault splays (Sillitoe and Perelló,2005), although the deposits of the Collahuasi-QuebradaBlanca cluster, which crop out to the east of the main faultsystem (Fig. 2), seem to have been emplaced over a basementdiscontinuity marked by regional isotopic changes in Neo-gene magmatic rocks (Mamani et al., 2010). At El Salvadorand Polo Sur, 42 to 41 Ma porphyry copper-related intrusionsare encapsulated in preexisting, up to 10-m.y. older, caldera-related rhyolite domes, indicating that the middle Eocene toearly Oligocene magmas effectively reused the same conduits(Cornejo et al., 1997).

These differences in the structural controls of porphyrycopper deposits can be attributed to the changing tectonicconditions along different segments of the Domeyko faultsystem during the >10-m.y. Incaic event. Camus (2003) rec-ognized three temporal groupings of mineralized intrusionsalong the Domeyko fault system at 43 to 42, 39 to 37, and 36to 33 Ma.

In the Chuquicamata, El Abra, and Quebrada Blanca-Col-lahuasi districts, where the three generations of intrusionsoccur, only the two youngest were fertile in copper (Campbellet al., 2006; Maksaev et al., 2009; Dilles et al., 2011), whereasin the Escondida district all three events are related to mineralization (Hervé et al., 2012). In contrast, most of the copper-bearing intrusions recognized in the Centinela districtseem to have been emplaced during the early event. In orderto illustrate the changing nature of the diverse porphyry clus-ters and explore the links between mineralized centers andthe evolution of the Domeyko fault system, the salient geo-logic features of three porphyry copper districts, Chuquica-mata-El Abra, Escondida, and Centinela (Figs. 2, 5), are ex-amined below.

Chuquicamata-El Abra

The Chuquicamata-El Abra intrusive cluster (Fig. 5a) is sit-uated along the northern Domeyko fault system, near theboundary between paleomagnetic domains C and D (Fig. 3a).The geology of the region records superimposed tectonic andmagmatic events beginning in the Mesozoic. In this region,the Precordillera (Cordillera de Domeyko) is formed by twofault-bounded, N-S−trending basement ranges. The westernrange, Sierra de Moreno, is a thick-skinned block uplifted inthe early Late Cretaceous (~85−84 Ma), during the inversion

of the Mesozoic back-arc basin (Ladino et al., 1997; Tomlin-son et al., 2001a). It is bounded to the west by high-angle,west-vergent reverse faults that place Neoproterozoic to Pa-leozoic basement units on top of Jurassic sedimentary strata(Fig. 5a). The east-vergent Arca reverse fault that bounds theeastern side of the Sierra de Moreno (Fig. 5a) has been in-terpreted as a reactivated normal fault, which formed nearthe eastern edge of the Mesozoic back-arc basin (Tomlinsonet al., 2001a; Fig. 5a). Syn- to post-tectonic red beds (Tolarand Tambillos Formations; Fig. 5a) were shed to the east andwest of the uplifting block during the Late Cretaceous. Afterdeformation, volcanic rocks, which unconformably cover thebasement units and Mesozoic sedimentary strata at Sierra deMoreno, developed in two separate episodes during the latestCretaceous (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) wasuplifted between 43 and 38 Ma during the Incaic event (Tom-linson et al., 1997a) along a new set of west- and east-vergent,high-angle reverse faults. As in other regions of northernChile, volcanism in the Chuquicamata-El Abra region sharplydiminished at this time; however, a syntectonic sedimentarysequence (Sichal Formation; Fig. 5a) that accumulated in anarrow basin between Sierra de Moreno and Sierra delMedio contains a few intercalations of tuffs and volcanic brec-cias that yield K-Ar and Ar/Ar ages between 43 and 36 Ma(Tomlinson et al., 2001a). The Incaic deformation also includesa strike-slip component that produced segmented N-NE− toNE-trending, dextral strike-slip faults (e.g., the Mesabi faultin Fig. 5a; Tomlinson et al., 1997a). These faults becamemore important in the southern part of the area nearChuquicamata (Fig. 5a), in accordance with increasing plateconvergence obliquity to the south along the Andean marginduring the Eocene (Fig. 4b).

As shown in Figure 5, the region between Chuquicamataand El Abra is one of the anomalous zones along the Cor -dillera de Domeyko (Fig. 2), where magmatism was recurrentsince the Late Cretaceous. Despite lacking evidence for largevolumes of middle Eocene to early Oligocene volcanic prod-ucts, 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 thesouthern termination of the Sierra de Moreno (Tomlinson etal., 2001a, 2010; Dillles et al., 2011). Figure 5b shows the re-stored shape of the Fortuna-El Abra batholith before beingseverely dismembered by left-lateral displacements along theWest fault (Tomlinson et al., 2007a; Dilles et al., 2011). Fieldrelationships indicate that the batholith was intruded alongthe trace of the Quetena reverse fault, an inverted normalfault inherited from Mesozoic back-arc extension. A flatteningfoliation in metaclastic rocks in the contact aureole is consis-tent with emplacement during regional Incaic E-W shorten-ing (Tomlinson and Blanco, 1997a).

The Fortuna-El Abra batholith, described by Dilles et al.(1997) as a porphyry copper batholith, is a long-lived, com-posite magmatic system that contains intrusive phases em-placed during different stages of the Incaic event; it is similarto the Andahuaylas-Yauri batholith of southern Peru, de-scribed by Perelló et al. (2003a). The batholith comprises an

336 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 336

older group of intrusions, namely the Los Picos complex andPajonal diorite, emplaced between ~45 and 42 Ma (Tomlin-son et al., 2001a; Campbell et al., 2006), which include pre-dominantly mafic pyroxene- and biotite-bearing quartz mon-zodiorites, monzodiorites, and quartz monzonites that arecopper barren (Fig. 5b). The second intrusive group includesthe Fortuna-El Abra granodiorite complex and is largely com-posed of two hornblende-bearing granodioritic phases (An-tena, 39.5−39.0 Ma and Fiesta, 38−37.5 Ma) as well as por-phyritic components (San Lorenzo porphyries; Fig. 5b). TheFortuna-El Abra Complex intrusions, derived from hydrous,oxidized, and sulfur-rich magmas (Dilles et al., 2011), are as-sociated with large porphyry copper deposits with U-Pb agesbetween 39 and 36 Ma (El Abra, 38−37 Ma; the Toki clusterincluding Toki, 39 Ma and Opache, 38−37 Ma; AlejandroHales, 39−36 Ma; Conchi, 36 Ma; Perelló, 2003; Campbell et

al., 2006; Boric et al., 2009; Marquardt et al., 2009; Barra,2011; Fig. 5b).

The youngest intrusive event in the Chuquicamata clusterwas related to the emplacement, beyond the southern limits ofthe Fortuna-El Abra batholith, of the Chuquicamata (36−32Ma) and Radomiro Tomic (36−34 Ma) porphyry copper de-posits (Lindsay et al., 1995; Reutter et al., 1996; Ballard et al.,2001; Ossandón et al., 2001; Campbell et al., 2006; Barra, 2011;Fig. 5). Sibson (1987), who relied on maps of Perry (1952), in-terpreted the Chuquicamata porphyry copper deposit to havebeen emplaced syntectonically, following an extensionalstepover linking two separate but overlapping, parallel, dextralstrike-slip faults. Alternatively, Lindsay et al. (1995) favored amodel in which the required space for magma emplacementcoincided with a releasing bend (cf. Cunningham and Mann,2007) along a single, continuously linked dextral fault.

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 337

0361-0128/98/000/000-00 $6.00 337

WEST FAULT

21º30'

22º

22º30'

CALAMA

69º

El Abra

Conchi Viejo

Clara Granodiorite

Llareta Granodiorite

FutureWest Fault

Porphyry Cu depositsunder post mineral cover

porphyrydikes

Small copperprospects

Opache

69º

22º00'

(Toki Cluster)

Fiesta Granodiorite(38-37.5 Ma)

Los PicosComplex

(45-42 Ma)

Upper Cretaceous volcanic rocks(Quebrada Mala fm.)

Upper Cretaceoussedimentary rocks

(Tolar fm)

Lower to MiddleEocene volcanic

rocks(Icanche fm)

Arca fault

Co.

Jas

pe

R. Tomic

Alejandro Hales

Chuquicamata

Mesabi Fault

Toki

El Abra

SIE

RR

A D

EL

ME

DIO

Cretaceous reverse fault

SIE

RR

A D

E M

OR

ENO

Upper Cretaceous volcanosedimentary sequences (Cerro Empexaand Quebrada Mala formation)

Lower to Middle Eocene volcanic rocks (Icanche formation)

Upper Eocene-Oligocene, syntectonic, sedimentary rocks (Sichaland Calama formations)

Eocene (45-37 Ma) monzodiorites to granodiorites (includingFortuna El Abra batholith)

Chuquicamata Porphyry Complex (36-33 Ma)

Cretaceous intrusive rocks

Cretaceous reverse faultwith Eocene reactivation

Eocene reverse faultrelated to EW shortening

(a) (b)

San Lorenzo

Early Abra-AntenaGanodiorite

Alejandro Hales

GenovevaTokiQuetena

Miranda

Quetenafault

AntenaGranodiorite

(39.5-39 Ma)

Pajonal Diorite

+++

++++

+

++

++

++++ +++

+

++

+

++ +

+

+ +++

+ + ++ + +

+ +

++ +

+ +

++

+

++ +++ + ++

+ +++ ++

++ ++ ++

+

+ ++++ +++

++ ++

+ ++ ++

+++

+

+

++++++

++

+

++

+

+ +

+ +

++

+

+++

++ + ++

+

+++

+

++

+

+

Quetena Fault

+

++

Paleocene intrusive rocks

Jurassic to Lower Cretaceous, back-arc, sedimentary sequences

0 10 20km

0 5 10km

+ +++

++

+

+

++ +

+

++

+ ++ +

+

+

+

+

+

++

++

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

Sierra del Medio basement block (Upper Paleozoic)

Triassic volcanic and sedimentary rocks

Sierra de Moreno basement block (Neoproterozoic to Paleozoic)

Upper Cretaceous, syntectonic, red beds (Tambillos and Tolarformations)

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

After porphyry emplacement at the Chuquicamata andRadomiro Tomic, to the south of their present location, thedeposits were displaced as a consequence of late Oligocene toearly Miocene (31.0−16.3 Ma) left-lateral movements alongthe throughgoing regional West fault or West Fissure. Dis-placement along the West fault, which includes, at Chuquica-mata, according to McInnes et al. (1999), a 600 ± 100 m westsideup vertical displacement component (contested by Tom-linson et al., 2001b), was able to offset the Fortuna-El Abrabatholith 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 dis-placement component was transferred to a set of normalfaults that bound the Cenozoic Calama extensional and/ortranstensional basin (Blanco, 2008; Blanco and Tomlinson,2009). Seismic data reveal that a buried normal fault with 1-to 1.5-km down-to-the-east displacement limits the north-western margin of the basin, where up to 2,500 m of silici-clastic continental strata accumulated between the Oligoceneand Miocene (Jordan et al., 2004; Blanco, 2008).

The West fault forms, at present, the sharp western limit ofthe Chuquicamata orebody. Nevertheless, structural inter-pretations by Sibson (1987) and Lindsay et al. (1995) consid-ered that the internal architecture of second-order faults andveins indicates that the deposit is essentially intact. Accordingto Lindsay et al. (1995), the Oligocene to early Miocene Westfault propagated northward, along the western edge of theporphyry complex (previously emplaced at 36−33 Ma), with-out displacing the mineralized intrusive units within the for-mer dextral releasing bend and/or stepover.

Escondida

Despite also being along the Domeyko fault system, thetectonic history of the Escondida region (Fig. 2) is markedlydifferent, highlighting the significant changes in tectonicstyles along the >1,000-km strike of the fault system. Defor-mation in this area seems to have been dominated by tectonicescape linked to passive rotation and transport of brittleupper crustal blocks over hot and ductile lower crust in a waysimilar to the so-called orogenic float or clutch tectonics mod-els discussed by Oldow et al. (1990), Lamb (1994), and Tikoffet al. (2002; see below). At this latitude, the Cordillera deDomeyko appears as a discontinuous mountain range formedby a group of discrete basement blocks bounded to the westby a 150-km-long shear lens (Escondida shear lens) devel-oped between the regional Sierra de Varas and Escondidafaults (Figs. 6a, 7). To the east, the Domeyko range abuts theSalar de Atacama depression, which is a deep subsiding basinfilled by >9 km of Cretaceous to Tertiary continental sedi-mentary strata (Pananont et al., 2004; Mpodozis et al., 2005).The basin was built on top of a large positive gravimetricanomaly (Central Andean Gravity High; Götze and Krausse,2002; Fig. 6b), which indicates the occurrence, at depth, ofdense crustal rocks that may help to explain its long-lived sub-sidence basin history, recorded at least since the Cretaceous.

The isolated basement blocks that form the core of therange are separated by small, triangular basins, with interiordrainage (Fig. 6a). The southern rhomboid-shaped blocks(e.g., San Carlos and Imilac; Fig. 6a) are bounded along theirnorthwestern margins by high-angle, SE-dipping reverse faults.

The blocks at Quimal, Los Morros, and Mariposas are lim-ited 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 arethrust over the sedimentary fill of the Salar de Atacama basin(Fig. 6a), which includes, among other units, a 2,500-m-thicksequence of Eocene to early Oligocene continental con-glomerates and poorly consolidated gravels. Internal pro-gressive unconformities and Ar/Ar ages between 44 and 43Ma from a tuffaceous horizon just above the base of this se-quence (Loma Amarilla Formation) indicate that thesestrata-accumulated syntectonically during the regional Incaicdeformation (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 ofa 250- × 50-km basement sliver that was transported north-ward during the Incaic deformation. According to Mpodoziset al. (1993a, b), the continuous northward shift of the dis-placed block was impeded by a buttress located to the northof the moving sliver as the displacement was transferred tothe east by tectonic escape (cf. Mann, 1997) toward thedeeply subsiding Salar de Atacama basin. In this model, theSalar de Punta Negra depression (Fig. 6c) would have formedas an extensional basin at the trailing edge of the displacedblock. Displacement transfer seems to have occurred byclockwise rotation of small detached blocks, which in turngenerated the local triangular-shaped extensional basins be-tween the rotating blocks as well as contractional deformationin their northeastern corners where basement was thrust overthe Salar de Atacama basin fill (Fig. 6c). Mpodozis et al.(1993a, b) located this buttress at Sierra de Limón Verde,which is a N-plunging basement half dome that attains one ofthe highest elevations (3,500 m.a.s.l.) in the Cordillera deDomeyko (Fig. 6a). However, if along-strike changes in localstresses resulting from the formation of the Bolivian oroclineare considered, the buttressing effect may have been pro-vided by the nonrotated paleomagnetic domain C, locatednorth of Calama (Fig. 3), where initial Eocene deformationwas 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 set-ting and distribution of the barren and mineralized intrusionsthat form part of the Escondida cluster (labeled “LE,” Fig. 2).The area encompasses the widest part of the regional Escon-dida shear lens, which is separated to the east from the SierraImilac and Sierra San Carlos basement blocks by the Escon-dida fault (the Panadero-Portezuelo fault is an alternativename used by Hervé et al., 2012). These two blocks were thenseparated by the intervening triangular Salar de Hamburgodepression (Fig. 7). Drilling shows that the Salar de Ham-burgo fill includes >1,200 m of red beds, lahars, and pyro-clastic rocks with U-Pb zircon ages of 38 Ma (San CarlosStrata, Fig. 7; Marinovic et al., 1995; Urzúa, 2009; Hervé etal., 2012); therefore, these units are equivalent to the upperportion of the syntectonic Loma Amarilla Formation in theSalar de Atacama. The Hamburgo fault is a NE-trending,high-angle reverse fault that places the late Paleozoic base-ment of the San Carlos block over the sedimentary sequencesof the Salar de Hamburgo (Fig. 7).

338 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 338

A protracted, >40-m.y. Cretaceous to Tertiary history ofmagmatism is recorded in the Escondida region. The oldestintrusive events produced Late Cretaceous (81−71 Ma; Fig.7) tholeiitic to alkaline pyroxene gabbros and diorites as wellas hornblende-pyroxene monzodiorites and diorites, in addi-tion to early Paleocene (66−64 Ma) pyroxene diorites. Theserocks intruded the sedimentary strata of the Mesozoic back-arc basin in the Escondida shear lens to the south and west ofEscondida (Fig. 7); Paleocene to early Eocene volcanic rocks(59−53 Ma) are also present in the area (Marinovic et al.,1995; Richards et al., 2001; Urzúa, 2009). All of this focusedand recurrent magmatic activity took place east of the Andeanarc front, which during the Late Cretaceous to early Pale-ocene (85−50 Ma) was located farther west, in the Centraldepression of the Antofagasta region (Boric et al., 1990).

Incaic magmatism in the Escondida cluster began in themiddle Eocene (~44 Ma), when left-lateral displacementsalong the Escondida fault and differential rotation betweenthe San Carlos and Imilac blocks created the triangular Salarde Hamburgo depression (Fig. 7). The oldest intrusions are agroup of 44 to 41 Ma pyroxene-biotite monzodiorites and pyroxene-hornblende granodiorites, emplaced within the Es-condida shear lens to the north of Escondida (Marinovic et al.,1995; Richards et al., 2001; Urzúa, 2009). Together, they likelyrepresent the roof of an underlying, partially eroded plutonnearly 20 km in diameter (Fig. 7). Most of these rocks, like theLos Picos complex and Pajonal diorite of the Chuquicamata-El Abra region, are barren, although geochronologic data andrelationships between intrusive phases in the vicinity of theChimborazo porphyry copper deposit indicate, according to

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 339

0361-0128/98/000/000-00 $6.00 339

Salar de AtacamaEl

Bo

rdo

Esca

rpam

en

Bla

ncas

faul

Salar de los Morros

Salar Verónica

Barr

anca

s

Salar dePunta Negra

N

Cen

tinel

a F

ault

Cor

don

de L

ila

Sier

raLim

ón Verde

Salar deHamburgo

Sie

rra

de

Va

ras

faul

Esc

ond

ida

faul

t

Sier

ra d

e A

lmei

da

Salarde Elvira

(a)

I

M

LM

Q

SC

ESCONDIDA

GABY

ESPERANZA-TELEGRAFO

Los

Toro

s fa

ult

Cen

tinel

a fa

ult

23º

24º

69º

ESL

Fortuna-El Abrabatholith

CALAMA

Sierra Limón Verde

Paleozoicbasement

blocks

ANTOFAGASTA

Sie

rra

de

Var

asf a

ult

Sie

rra

de

lMed

io ?

?

N

24º

70º 68º

Non-rotatedDomain C

Extensional(rotational)

basins

BUTTRESS ZONE

Free face

GravityHigh

RotatedDomain D

Bolivia

Arg

entin

a

Boundaryfault

Salarde

Atacama

Salarde

Punta Negra

(b) (c)

Dee

p B

asin

ThrustsA'

La Escondida shear lens (ESL )

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

Mesozoic to Paleocene strata ofthe Centinela District

Outcrops of Lower Cretaceousto Upper Oligocene continentalclastic strata of the Salar deAtacama basin

Cordón de Lila-Sierra deAlmeida stable domain

Strike-Slip faults Reverse faults Normal faults0 20 40 km

25º

A

rr'

α

0 50 km

Transition zone betwenC and D domains

N

FIG. 6. (a). Main structural elements of the Cordillera de Domeyko (between Escondida and Sierra Limón Verde andSalar de Punta Negra (22° 30°−25° S; see location in Fig. 2). Note the large shear lens (Escondida shear lens) flanked by theEscondida 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 clockwise-rotated 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 alonga left-lateral strike-slip fault system. Displacement is impeded, as shown, by a buttress at the leading edge of the block, andtransferred toward the right by means of clockwise block rotations. Note extensional basins created between the rotatingblocks. A-A' and r-r = position of points and lines before and after rotation (α = rotation angle). Adapted from Beck et al.(1993).

Hervé et al. (2012 ), that an early phase of copper mineral-ization probably occurred at ~41 Ma.

The second episode of Eocene-Oligocene magmatismbegan with the emplacement of a closely spaced group ofsmall 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;Urzúa, 2009), intrude both the late Paleocene to earlyOligocene volcanic rocks of the Escondida shear lens and thelate Paleozoic basement units of the Imilac block (Fig. 7).Their distribution suggests that they could be apophyses of alarger pluton at depth that intruded along the Escondidafault. The slightly younger group of porphyry copper stocksinclude a series of multiphase, NE- to N-NE−trending, dike-like intrusions that were emplaced at or near the Escondidafault at 38 to 37 Ma; these include the deposits at Zaldívar,Escondida Norte, Escondida, and Pinta Verde, and, fartheraway, at Baker (Richards et al., 2001; Urzúa, 2009; Hervé etal., 2012; Fig. 7).

The final event of Incaic magmatism in the Escondida clus-ter was related to the emplacement of the Escondida Esteand Pampa Escondida deposits, immediately to the east ofthe Escondida fault (Fig. 7), between 36.0 and 34.5 Ma(Hervé et al., 2012). The mineralized porphyries of the Es-condida cluster, with the exception, perhaps, of Chimborazo,postdate the earlier phase of sinistral faulting and block rota-tions along this segment of the Cordillera de Domeyko. De-formation seems to have begun at ~42 Ma (age of the base ofthe Loma Amarilla Formation; see above), although the lackof offset on any of the porphyry intrusions across the localfault strands (Panadero-Portezuelo fault) of the larger Escon-dida fault indicates that major along-strike fault activity hadceased by 38 Ma, as shown by the across-fault 38−37 Ma por-phyry dikes (Hervé et al., 2012).

Centinela

Porphyry copper mineralization in the Centinela district (la-beled “CE,” Fig. 2) occurs within a 25-km-wide, fault-bounded

340 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 340

Sie

rra

de

Var

asfa

ult

24º30'

69º

24º

La E

scon

did

a fa

ult

Hambu

rgo f

ault

Imila

c fa

ult

Pinta Verde

Sie

rra

de Im

ilac

Sier

raSa

n Car

los

Chimborazo (41)

Zaldivar (38-37) Baker (38-37)

Escondida(38-37)

Escondida Norte (38-37)

Salar deHamburgo

Salar dePunta Negra

Pampa Escondida (36-34.5)

0 5 10 km

Upper Paleocene to Lower Eocene (59-53 Ma) volcanicsequences

Eocene (44-41 Ma) px-bt monzodiorites and px-hbgranodiorites

Upper Eocene (42?-36 Ma) sedimentary-volcanicsequence (San Carlos strata)

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

Upper Eocene (38-35 Ma) dacitic to granodioriticmineralized porphyry intrusions (Escondida cluster)

Upper Triassic-Lower Cretaceous sedimentary sequences

Upper Paleozoic (300-270 Ma) basement

Upper Cretaceous (81-72 Ma) gabbro-diorites and hb-pxdiorites

Upper Cretaceous (74-70 Ma) rhyolitic ignimbrites

Lower Paleocene (66-64 Ma) px diorites

Triassic (240-220 Ma) intrusive rocks

Escondida Este(36-34.5)

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

belt of Late Cretaceous to early Eocene volcanic rocks, lo-cated between the Paleozoic basement exposures of theCordillera de Domeyko to the east and an early Cretaceousvolcanic sequence in the Coastal Range to the west (Fig. 8).The Centinela district hosts one of the more recently discov-ered porphyry copper clusters in northern Chile. Althoughthe occurrence of exotic copper mineralization at El Tesorowas known for a long time, the full potential of the districtonly began to be assessed in the mid-1990s (Perelló 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 theCentinela cluster comprise a group of Early Cretaceousolivine-pyroxene gabbros and hornblende-bearing quartz dior-ites, with U-Pb zircon and K-Ar ages between 124 and 100Ma (Mpodozis et al., 1993b; Marinovic and García, 1999).

These rocks, emplaced within Jurassic marine limestones ofthe northern Chile back-arc basin (Fig. 8), are almost 100 kmeast of the Early Cretaceous magmatic front, which, as notedabove, was situated at that time in the Coastal Range (Boricet al., 1990). Younger volcanic and intrusive events occurredin the Late Cretaceous, when a volcanic sequence (QuebradaMala Formation) and a group of coeval 70 to 66 Ma pyroxenediorites to rhyolite porphyries and flow domes, dated (U-Pbzircon) between 70 and 66 Ma, were generated after the An-dean arc front migrated eastward into the Centinela region(see Figs. 1b, 8). Volcanism continued after the Cretaceous-Tertiary boundary deformation event, with eruptions fromstratovolcanoes and small collapse calderas that were activebetween the early Paleocene (64 Ma) and the early Eocene(53 Ma). During this interval, a diverse group of epizonal in-trusions composed mainly of pyroxene-biotite quartz dioriteto monzodiorite (60 Ma) and hornblende- biotite granodiorite

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 341

0361-0128/98/000/000-00 $6.00 341

Upper Paleozoic (290-270 Ma) basement

Upper Triassic (210-200 Ma) volcanic and sedimentaryrocks

Jurassic to Lower Cretaceous back-arc sedimentaryand volcanic rocks

Lower Cretaceous (124-100 Ma) ol-px gabbros todiorites and hb granodioritic porphyry intrusions

Lower Cretaceous (?) volcanic rocks

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

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

Paleocene to Lower Eocene (64-53 Ma) volcanic rocks(Cinchado formation )

Paleocene (60-56 Ma) px-bt monzodiorites andrhyolitic porphyry intrusions

Eocene (44-40 Ma) px-hb monzodioritic to hb-btgranodioritic stocks and mineralized dacitic porphyryintrusions (Esperanza-Telégrafo and Centinela-Polo Sur)

Upper Eocene (44-40 Ma) syntectonic, sedimentary andvolcanic rocks

Lower Paleocene (65-64 Ma) diorites and dacitic porphyryintrusions

Sie

rra

del B

uitr

e fa

ult

23º

69º00

Esperanza (42-40)

Sierra Limón Verde

Las Lomasduplex

Los

Toro

s fa

ult

Cen

tinel

a fa

ult

Llano (41)Mirador (41-39)

Orión (44-41)

Telégrafo (42-40)

Caracoles (42-41)

Centinela (45-44)

Penacho Blanco (42)

Pilar (43)

Polo Sur (42-41)

Sierra

Agua Dulce

Undifferentiated Cretaceous granitoids

0 5 10 km

as

Las Lomas fault

Esperanza fault

Coronado fault

Llano fault

Sherezade (44-43)

FIG. 8. Regional geologic map of the Centinela cluster area. Note the 35-km-long NNE trend of 42−40 Ma porphyry cop-per deposits emplaced during earlier stages of the Incaic event. Multiple superimposed intrusive pulses and volcanicepisodes between 120 and 40 Ma show a remarkable recurrence of magmatic events for >80 m.y. All ages are based on re-cently acquired U-Pb zircon data. (p = prospects, ol = olivine, bt = biotite, px = pyroxene).

(58−57 Ma) plus andesitic to dioritic porphyritic intrusions,were emplaced into the Mesozoic units and Paleogene vol-canic edifices.

Incaic magmatism and mineralization in the Centinela dis-trict occurred between 45 and 39 Ma (Mpodozis et al., 2009a;Perelló et al., 2010) and began, as revealed by numerous newU-Pb zircon ages in the intrusive rocks and Re-Os ages inmolybdenite, ~12 to 10 m.y. after the termination of volcan-ism in the early Eocene.

This event coincides with the mostly copper-barren earlyphase of Incaic intrusions at Chuquicamata-El Abra (45−42Ma) and Escondida (44−41 Ma). The oldest Incaic intrusiverocks include a small group of 45 Ma pyroxene-biotite andquartz diorites yet, in contrast to Chuquicamata-El Abra andEscondida, at Centinela, numerous mineralized porphyrycenters were emplaced between 44 and 39 Ma. They form,together with some barren stocks, a 40-km-long, N- to NE-trending belt, which includes at least 10 discrete intrusivecomplexes (Fig. 8). A syntectonic sequence of conglomeratesand volcaniclastic sandstones, which accumulated at the sametime as porphyry copper emplacement, comprises interbed-ded layers of dacitic block-and-ash deposits and tuffs with U-Pb zircon ages between 42 and 39 Ma.

The oldest porphyry systems (45−43 Ma) occur along thesouthwest end of the belt, and the age decreases systematicallyto the northeast until reaching 39 Ma at the northeast edge ofthe porphyry trend (Fig. 8). The geometry of the porphyrycomplexes is controlled by their position relative to the mainstructural feature of the district, a 3- to 5-km-wide, N-S−trend-ing fault zone that cuts obliquely across the porphyry belt. Thiszone of intense deformation constitute to the northern termi-nation of the Sierra de Varas fault, which stretches for >250km along the western border of the Cordillera de Domeyko(Mpodozis et al., 1993b; Soto et al., 2005; Figs. 2, 6a), and wasactive both during and after porphyry emplacement (Fig. 8).

Porphyry deposits located west and east of this zone of con-centrated deformation are largely undeformed. Copper miner-alization in mineralized porphyry systems, located west of thefault zone, are related to subvertical, hornblende-biotite dacitesdike swarms intruded into Paleocene volcanic/subvolcanic units(e.g., Centinela) or early Eocene rhyolitic dome complexes(Polo Sur, Perelló et al., 2010). The oldest deposits where em-placed at 45 to 44 Ma (Centinela) and 44 to 43 Ma (Sher-erezade) to be followed by the intrusion by several barren py-roxene-hornblende dioritic stocks and lacoliths dated at 43 Ma,although a porphyry copper system with the same age has beenalso recognized at Pilar. A new pulse of copper-bearing intru-sions occurred, finally, between 42 to 41 Ma, at Polo Sur, whiledacitic porphyries with a similar age (42 Ma) but apparentlybarren have been documented at Penacho Blanco (Fig. 8).Mirador, the youngest porphyry deposit recognized so far in thedistrict (41−39 Ma; Mora et al., 2009) and located east of thefault zone (Fig. 8), is also structurally undisturbed. The coppermineralization, hosted within Jurassic marine limestones andevaporites (Mora et al., 2009) is associated with a group ofmultiphase, W- to NW-trending intrusions that, as in the olderCentinela 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, 42−41 Ma;Telégrafo, 42−40 Ma; Esperanza, 42−40 Ma; Llano, 41 Ma;

Perelló et al., 2004, 2010; Bisso et al., 2009; Münchmeyer andValenzuela, 2009; Swaneck et al., 2009); they are all associ-ated with tilted porphyry dike swarms. These deposits areemplaced into moderately to steeply dipping strata, which aredisrupted by major postmineral faults that exhibit reverse,normal, and strike-slip displacement components (Figs. 8, 9).

Although the widespread gravel cover makes it difficult tosatisfactorily resolve the structural relationships within thewhole district, the regional structure around the Esperanzaand Telégrafo deposits (Fig. 8) comprises a long-wavelength,asymmetric, basement-cored anticline, bounded to the westby a moderately E-dipping yet unexposed thrust fault thatwas discovered during exploration drilling (Telégrafo fault;Perelló et al., 2004, 2010; Bisso et al., 2009; Münchmeyerand Valenzuela, 2009; Fig. 9). The hinge zone of the anticlineis, in turn, sliced by two subvertical faults (Coronado andLlano faults; Figs. 8, 9) linked to the N-S−trending regionalfault zone. Figure 9 includes a west-east structural sectionacross the Esperanza deposit, where mineralization is associ-ated with a group of easterly inclined porphyry dikes em-placed within the ~40° to 50° W-dipping Triassic to UpperCretaceous strata that form the frontal limb of the anticline.This panel, containing in part mineralized and altered hostrocks to the porphyry deposits, is upthrown to the west,along the Telégrafo fault, over barren, unaltered mid-Eocene(42−39 Ma) sedimentary and volcanic rocks that accumu-lated 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 Creta-ceous strata. The rectilinear fault trace and the mismatch ofthe lithology and age of the Late Cretaceous volcanic rocksacross the Esperanza fault show it includes an importantcomponent of strike-slip movement, although the preciseage of deformation and genetic links between both faults re-main to be determined. The Llano and Coronado faults are,however, as shown in Figure 9, younger faults that are su-perimposed over the Telégrafo-Esperanza system, which ex-hibits both left-lateral and large, down-to-the-east compo-nents of displacement, part of which has a late Miocene oryounger (<10 Ma) age.

Structural relationships at Caracoles, in spite of being only10 km to the south, are entirely different and difficult tomatch with those observed at Esperanza-Telégrafo. In thesouthern part of the district (Fig. 8), the fault zone is dis-placed to the west, out of strike with the fault zone to thenorth, and includes a strike-slip duplex (Las Lomas duplex)bounded by two N-S−trending subvertical faults (Las Lomasand Centinela faults; Fig. 8). Kinematic indicators show evi-dence for an episode of left-lateral displacement, even if theinternal geometry of the duplex is compatible with an earlier(Late Cretaceous?) event of dextral shear (Marinovic andGarcía, 1999; Mpodozis et al., 2009a). The Caracoles deposit(42−40 Ma, Swaneck et al., 2009) is hosted in early Paleocene(64−60 Ma) volcaniclastic rocks beneath the gravel cover.Copper mineralization is linked to a steeply SE-dippingswarm of thin dacite porphyry dikes that follow the internalstructural trends of the Las Lomas duplex; this is consistentwith emplacement of the porphyry along a zone of (earlier orsynmineral?) strike-slip deformation.

342 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 342

Long-lived intrusive clusters and porphyry deposits: What is the relationship?

As discussed above, Eocene to Oligocene porphyry copperdeposits in northern Chile are located in districts of focusedand long-lived magmatic activity distributed along the Cor -dillera de Domeyko (Fig. 2). In some districts, magmatismwas active for 50 m.y. (Chuquicamata-El Abra, Escondida) oreven 80 m.y. (Centinela). Magmatism began in a back-arc set-ting while the arc front was located much farther west in theCoastal Range or Central depression (Boric et al., 1990); todate, the cause of these zones of long-term magmatism is un-certain. Yañez and Maksaev (1994) suggested that porphyryspacing may be related to Rayleigh-Taylor diapirism along themid Eocene-early Oligocene arc. Behn et al. (2001) notedthat a correlation exists between the location of porphyry cop-per deposits and regional E-W−trending negative magneticanomalies that extend from the coast to the modern Andeanarc, and which may represent crustal structures favorable formagma ascent as the arc migrated eastward since the Jurassic.Richards (2003) speculated that porphyry copper depositswere emplaced at the intersections of the Domeyko fault sys-tem with NW-trending transcordilleran crustal-scale struc-tures, yet with the exception of Potrerillos (Fig. 2), none ofthese has been documented on the western (Chilean) side ofthe Andes. To address these observations, Tomlinson andCornejo (2012) proposed a hybrid model to explain porphyryspacing, which combines Rayleigh-Taylor diapirism withcrustal-scale structural control.

Even if these models explain porphyry spacing, they do notelucidate why clustered magmatism began long before theEocene. As more data are gathered, an even more strikingspatial relationship is emerging in several districts (Escon-dida, Centinela, Chuquicamata, Quebrada Blanca-Collahuasi,Sierra Exploradora; Figs. 5, 7, 8), between extended, Creta-ceous and younger magmatic activity, porphyry copper de-posits, and Triassic intrusions which, near Escondida and Col-lahuasi, show evidence of weakly developed porphyry-stylecopper mineralization (Cornejo et al., 2006; Munizaga et al.,2008). In addition, indistinguishable geochemical signaturesreported in a pilot study by Wilson et al. (2011), which com-pared Triassic granodiorites to mineralized Eocene porphyrydeposits at Alejandro Hales (Fig. 5), may suggest tapping of acommon 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 centralChile (31°−34° S) and contiguous Argentina differs from thatof northern Chile in that the main deformation events areyounger, the tectonic style is different, and the principal ageof porphyry copper mineralization is much younger (lateMiocene to early Pliocene). At present, a zone of flat subduc-tion (Central Chile or Pampean flat-slab region; Cahill andIsacks, 1992; Kay and Mpodozis, 2002; Ramos et al., 2002;

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 343

0361-0128/98/000/000-00 $6.00 343

Coronadofault

Esperanzafault

Llano faultPalMi-Pl

Telégrafofault

Jur

Trv

KsKs

Ks

49200 49400 49600

3200

2800

2400

2000

1600

1200

800

Ks

Eo

Ol Ep

0 0.5 1 km

W E

Mi-Pl Ks

Lower Paleocene dacitic domes (64-63 Ma)

49000

Miocene to Pliocene gravelsMi-Pl

Eocene sedimentary and volcanic strataEo

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

Ep Esperanza copper porphyry intrusions (41-40 Ma)

Pal

Upper Cretaceous volcanic rocks (Quebrada Malaformation, 70-66 Ma)Ks

Trv

Jurassic marine sedimentary sequences

Upper Triassic volcanic rocks (210-200 Ma)

Jur

Left-lateral displacement on faults 0.5% CuT

Oligocene sedimentary rocks

Ol

Ol

FIG. 9. Schematic structural section across the Esperanza porphyry Cu deposit, Centinela district. The Esperanza ore-body is part of a W-dipping sliver of Jurassic and Late Cretaceous strata intruded by Eocene porphyry dikes (41−40 Ma),thrust to the west (Telégrafo fault) on top of Eocene (42−37 Ma) sedimentary and pyroclastic sequences. The Esperanza andCoronado faults have a complex and younger displacement history, including strike-slip components. The down-to-the-eastdisplacement shown along the Coronado and Llano faults corresponds only to the youngest (late Miocene-Pliocene?) episodeof deformation. Location of section shown in Figure 8.

Fig. 10) that was formed by progressive slab shallowing, begin-ning in the early Miocene (Kay et al., 1987), extends from 27°to 33° S. As in northern Chile, the Coastal Range is formed bythe volcanic and intrusive remnants of a Jurassic to Cretaceousarc system. From 32° S southward, a thick sequence of coevalmarine and terrestrial sedimentary rocks exposed along theeastern slope of the Cordillera Principal corresponds to sedi-ments that accumulated within the Mesozoic Neuquén back-arc basin (Mpodozis and Ramos, 2008; Figs. 10−12).

The most notable geologic feature, however, is a severalkilometer-thick volcano-sedimentary sequence that formsmost of the western part of the Cordillera Principal between32° and 37° S (Figs. 10b, 11, 12), which has been traditionally

assigned to the Abanico, Coya-Machalí, and Cura-Mallín For-mations (e.g., Charrier et al., 1996, 2002; Jordan et al., 2001;Kay et al., 2005; Farías et al., 2008). These sequences accu-mulated in extensional volcano-tectonic depressions or intra-arc basins that are referred to as the Abanico basin; Ar/Arages at the latitude of Santiago (33° S) range from latestEocene to early Miocene (35−21 Ma; Muñoz et al., 2006).Volcanic rocks are calc-alkaline to tholeiitic in composition,and the overall geochemical signature suggests that volcanicactivity occurred over a relatively thin crust (<35 km; Kay etal., 2005; Muñoz et al., 2006).

Equivalent units extend for more than 1,500 km to thesouth along the crest of the Andean range into the northern

344 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 344

Puerto Montt

Aconcagua

Perú-C

hile trench

Concepción

AR

GEN

TIN

A

SomuncuráPlateau

Figure 11

33º

35º

41º

69º

Coa

stal

Ran

ge

Santiago

Liqu

iñe

- O

fqui

fau

lt

0 100 km

Cen

tral

Val

ley

Maipo Orocline

37º

39º

43º

Middle and upper Miocene volcanicrocks (with major sedimentaryparticipation in the south)

Oligocene to middle Miocene volcanicsequences and interbeddedsedimentary strata

Late Oligocene to middle Miocenecontinental strata

Late Oligocene to middle MioceneSomuncurá plateau volcanic rocks

Eocene volcanic and intrusive rocks

Oligocene to Miocene intrusive rocks

IslaMagdalena

71º73º(b)

Bariloche

32º

80º

20º

30º

75º

Juan Fernández Ridge

Per

ú- C

hile

Tre

nch

PERU

BOLIVIA

AR

GE

NTI

NA

Altip

lano-Puna

30 Ma

40 Ma

3.4 cm/yr

Peru-C

hile trench

40º

(a)

SVZ

CVZ

e

Chile

tre

20 Ma

Flat-slabR

egion

Antofagasta

Santiago

Concepción

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 eleva-tions >3 km. (b). Simplified geologic map showing the distribution of Oligocene-Miocene volcanic and sedimentary strata insouth-central Argentina and Chile from 32° to 45° S. Based on the compilation by Jordan et al. (2001) and the 1:1,000,000geologic 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).

Patagonian Andes (Fig. 10b). Even though the erosion levelincreases and the volume of preserved volcanic products decreases southward, these sequences define a progressivelysouthward-widening belt that extends from the CoastalRange to the eastern slope of the Andes at the latitude ofPuerto Montt (41° S, Fig. 10b). The nature of interbeddedsedimentary rocks changes along strike, from continental tolacustrine facies in the northern part of the belt to marine facies toward the south; geochemical signatures also displayevidence for crustal thinning and extension increasing to thesouth. Muñoz et al. (2000) indicated that 37 to 20 Ma maficlavas 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

asthenospheric upwelling during a transient Oligocene toearly Miocene event of very rapid subduction. Deep mantleupwelling has also been suggested by Kay et al. (2007a) to ex-plain the origin of the 33 to 17 Ma basaltic flows of the largeback-arc Somuncurá volcanic plateau in Patagonia (Fig. 10b).Oligocene to Miocene submarine basaltic pillow lavas withMORB geochemical features in the Aysén region indicate ex-treme crustal thinning along the arc farther south, between43° and 46° S (Silva et al., 2003).

Miocene compressional failure of the Abanico basin

The late Eocene to early Miocene extensional period ter-minated at ~20 to 18 Ma, followed by compressional defor-mation leading to the emergence of the modern Andes in

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 345

0361-0128/98/000/000-00 $6.00 345

Mendoza

Rancagua

Curicó

Valparaíso

Tupungato

San José

Maipo MAIPO OROCLINESYMMETRY AXIS

CHILE ARGENTINA71º 69º

33º

34º

35º

FRONTAL

CORDILLERA

P R

I N C

I P A

L C

O D

I L L E R

A

Cretaceous intrusive rocksLate Paleozoic to Jurassic batholiths(Coastal Cordillera) Eocene? - Early Miocene volcano-

sedimentary sequences (Coya - Machalí/Abanico formations)

Miocene volcanic rocks (Farellonesformation)

CO

AS

T A L R

A N

G E

M

Early Cretaceous sedimentary strata(Neuquén basin)Triassic to Jurassic sedimentary strata(Neuquén basin)

Quaternary stratovolcanoes

Quaternary volcanic rocks

Plio-Pleistocene tuffs

Fore arc Miocene sedimentary strata

Miocene intrusive rocks (PrincipalCordillera)

Fontal Cordillera Paleozoic basement

0 25 50 km

Los Andes

EL TENIENTE

Upper Cenozoic foreland basin strata

Jurassic to Cretaceous volcano-sedimentarysequences (Coastal Codillera)

Santiago

RIO BLANCO -LOS BRONCES

A

FIG. 11. Geologic map of the area around the Maipo orocline from 32°30' to 35° 30' S (location in Fig. 10). Note changein the structural trend of the Coastal Range and Principal Cordillera across the Maipo orocline (Farías et al., 2008; Arriagadaet al., 2009), from N-S, to the north, to N-NE, to the south. A = Aconcagua fold-and-thrust belt, M = Malargüe fold-and-thrust belt (adapted from Farías et al., 2008). See text for more details.

346 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 346

32º

33º

34º

Piuquenes (11)

Los Azules (10-8)

Rincones de Araya (9)

El Altar (12-10)

Los Pelambres (14-10)

Cerro Bayo del Cobre(12-10)

El Yunque (15)

Aconcagua

Amos- Andrés (9-8)

Vizcachitas (11-10)Morro Colorado Pimentón (11-9)

San Felipe

West Wall (11-9)

Novicio (15-13)

Rio Blanco-Los Bronces(7.7-4.7)

Los Sulfatos (7-6)

Tupungato

San José

Maipo

Santiago

El Teniente(6.5-4.6)

Rancagua

Valparaiso

0 50 km

El Pachón (9-8)

Los Machos (15-14)Los Piches (14-12.5)

Cerro Mercedario

Mercedario ( 13)

Quaternary stratovolcanoesQuaternary volcanic rocksUpper Cenozoic foreland sedimentarysequencesMiocene intrusive rocksVolcanic rocks of the Abanico basin (Eocene? to Miocene) include the Abanico, Coya-Machali and Farellones formationsJurassic to Cretaceous sedimentarysequences of the Neuquén basinCoastal Range block (Paleozoic to earlyCretaceous)Frontal Cordillera Paleozoic basement

71º 70º

Chile Argentina

70º

Poc

uro

faul

tLR

A

M

FIG. 12. Tectonic sketch of the northern end of the Abanico intra-arc basin (31°−34° S), showing the location and age(Ma, in parentheses) of Miocene to early Pliocene porphyry copper deposits of central Chile and contiguous Argentina andthe composite fold-and-thrust belt developed along the eastern margin of the basin (LR = La Ramada, A = Aconcagua, M =Malargüe fold-and-thrust belts).

central and southern Chile and contiguous Argentina (e.g.,Giambiagi et al., 2003). Between 31° and 34° S, much of thisdeformation was focused along the eastern border of theAbanico basin, which finally collapsed and was inverted by com-pressional deformation in response to the collision between

the colder and rigid western Coastal Range and easternFrontal Cordillera blocks (Fig. 13). Under these conditions,the late Eocene to early Miocene volcanic sequences weretectonically transported to the east and juxtaposed over thesedimentary sequences of the Neuquén back-arc basin to

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 347

0361-0128/98/000/000-00 $6.00 347

Rapid ConvergenceWeak intraplate

coupling

Intra-arcAbanico basin Mesozoic

sedimentary wedge

Lithosphere

Asthenosphere

Coastal Range Block

Steep slab35-21 Ma

Slow ConvergenceStrong intraplate

coupling

Subduction erosion

Farellones volcanismCollapsed& inverted

Abanico basin

Zone of partial melting

12-5 Ma

Shallowing slab

Strong intraplatecoupling

Syntectonic intrusions

El TenienteLos Bronces Los Pelambres

Lower crust melting andmixing with mantle-derived magmas

Delaminatedlower crustal blocks?

10-5 Ma

Moho

W E

Slow Convergence

AcceleratingSAM plate

Aconcagua-La Ramada

FTB

Aconcagua-La Ramada

FTB

Frontal Cordillera block

Frontal Cordillera

FIG. 13. Schematic diagram near 33° to 34° S, showing the evolution of the Abanico intra-arc basin during the Oligoceneand Miocene (SAM = South America, FTB = fold and thrust belt). Major porphyry copper deposits began to be formed at10 Ma when the deformation front migrated to the east and the Frontal Cordillera was uplifted as a consequence of shal-lowing of the subducted Nazca plate. Even though the diagrams combine geologic relationships observed at different lati-tudes it shows the tectonic position of the Los Pelambres intrusions along the boundary thrusts of the deformed Abanicobasin and the Río Blanco-Los Bronces and El Teniente intrusions in the less deformed rocks near the center of the formerbasin farther to the west. Red arrows below the Abanico basin show hypothetical magma paths. (SAM = South Americanplate, FTB= fold and thrust belt)

form the La Ramada, Aconcagua, and Malargüe fold-and-thrust belts (Ramos et al., 1996; Figs. 11−13). In central Chile,a sharp decrease in the volume of volcanism ensued, the arcfront migrated eastward, and the geochemical and isotopicsignatures of younger, middle to late Miocene volcanic sequences (e.g., Farellones Formation; Figs. 10−11) indicateprogressive crustal thickening as a consequence of increasedhorizontal shortening (Ramos et al., 1996; Kay and Mpodozis,2002; Stern et al., 2010).

Between 33° and 35° S, the amount of Miocene andyounger back-arc shortening decreased along strike fromnorth to south, as the tectonic style in the deformed back-arcsequences changed from the narrow, thin-skinned Aconcaguafold-and-thrust belt to the wider, mixed thin- and thick-skinned style of the Malargüe fold-and-thrust belt to thesouth (Ramos et al., 1996, 2004; Giambiagi et al., 2011; Figs.11−12). The transition zone, at 34° S, coincides with the sym-metric axis of the W-NW−trending Maipo orocline (Farías etal., 2008; Arriagada et al., 2009), and is revealed by thechange in orientation of both the Chile trench and the mainstructural trends of the Principal Cordillera, from N-S to N-NE (Fig. 11). The Maipo orocline, a more subtle feature thanthe Bolivian orocline (see above), is also shown in paleomag-netic studies, as the magnitude of paleomagnetic block rota-tions determined in all rocks older than 10 Ma changes from4°clockwise north of the orocline to 32° clockwise to thesouth (Arriagada et al., 2009; see Fig. 10b). These changes co-incide with a rapid fall of the absolute elevation of the Andeanrange, as well as an overall decrease of crustal thickness, from50 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 themore contractional conditions prevailing during the Neogenenorth of latitude 33° S, as shallowing of the subducting Nazcaslab progressed. This, in turn, lead to the establishment of themodern Chilean or Pampean flat-slab region between 28°and 33° S (e.g., Kay and Mpodozis, 2002). Slab shallowing isgenerally attributed to the buoyancy effect introduced by thesubduction of the E-NE−trending Juan Fernández Ridgeduring the late Miocene (Yañez et al., 2001; Ramos et al.,2002, and references therein), although recent analogue andnumerical experiments (e.g., Martinod et al., 2005) show thatmoderate-sized, buoyant ridges that impinge on a trench arenot able to alone induce formation of flat-slab segments of thedimensions observed in central Chile and contiguous Ar-gentina. Other authors (Manea et al., 2012) suggest that acombination of trenchward motion of thick cratonic lithos-phere accompanied by trench retreat may better explain theformation of the Chilean-Pampean flat slab during theMiocene.

Late Miocene to Early Pliocene Porphyry Copper Deposits of Central Chile and Contiguous Argentina

Overview

The major porphyry copper deposits in central Chile andcontiguous Argentina are preferentially located at or near thetransition zone between the Chilean (or Pampean) flat-slabzone and the steeper subduction zone beneath southernChile. This region contains three of the world’s largest copper

deposits, at Los Pelambres, Río Blanco-Los Bronces, and ElTeniente (Camus, 2003; Cooke et al., 2005; Sillitoe and Perelló,2005), as well as numerous smaller deposits and prospectsthat straddle the Chile-Argentina border region from 31° to35° S (Fig. 12). Although the first evidence of mineralizationthat postdates the beginning of Miocene deformation is as oldas 15 to 13 Ma (Novicio, 15–13 Ma; Los Machos, 14−12.5 Ma;Los Piches, 14−12.5 Ma; Maksaev et al., 2009; Toro et al.,2009; Fig. 12), most of the deposits north of 33° S (LosAzules, Rincones de Araya, El Altar, Los Pelambres, ElPachón, Vizcachitas, Amos-Andrés, Pimentón, Novicio, plusWest Wall in the San Felipe cluster) were emplaced between12 and 8 Ma (Sillitoe and Perelló, 2005; Maksaev et al., 2009;Toro et al., 2009; Maygadán et al., 2011; Perelló et al., 2012).

At this time, the E-NE−trending fragment of the Juan Fer-nández Ridge began to subduct below the Los Pelambres re-gion (Yañez et al., 2001; Kay and Mpodozis, 2002), causingthe Andean deformation front to move eastward; uplift of theFrontal Cordillera (Fig. 11) commenced and was accommo-dated by regional thick-skinned faults (Ramos et al., 1996,2004; Pérez, 2001; Giambiagi et al., 2003; Fig. 12). By con-trast, south of 33° S, the youngest and largest deposits, in-cluding Río Blanco-Los Bronces and El Teniente (Maksaev etal., 2004; Deckart et al., 2005; Maksaev et al., 2009; Fig. 12),were emplaced from 7 to 4 Ma during a period (7−5 Ma) oftransient shallowing of the subducting slab below theNeuquén basin, when volcanism with arc-like geochemicalsignatures expanded up to 500 km east of the present dayPerú-Chile trench (Ramos and Folguera, 2005; Kay et al.,2006).

From a structural point of view, porphyry copper depositsof central Chile and contiguous Argentina include a group ofstocks that postkinematically intruded regional thrusts alongthe boundary between the former Abanico basin and the thin-skinned La Ramada fold-and-thrust belt (Los Pelambres,Amos-Andrés, Pimentón; Figs. 12, 14a), along the northernpart of the belt. Another, southern group (e.g., Vizcachitas,West Wall, Río Blanco-Los Bronces, and El Teniente; Figs.11−12) comprises intrusive complexes emplaced farther westof the thrust belt in gently folded sequences of the Abanicobasin, with no obvious relationship to regionally significanttectonic structures. A distinctive feature of the porphyry sys-tems at Los Pelambres, Río Blanco-Los Bronces, and El Te-niente is the occurrence of barren and/or mineralized mag-matic hydrothermal breccia complexes (Skewes and Stern,1994).

Los Pelambres

The late Miocene Los Pelambres porphyry copper-molyb-denum deposit and its smaller gold-bearing satellite, theFrontera deposit (Fig. 12), are located in a narrow belt of in-tense deformation that involves Oligocene to early Miocene(33−18 Ma) volcanic rocks of the Abanico basin; these vol-canic rocks, previously described as the Los Pelambres For-mation, form part of the northern termination of the La Ra-mada fold-and-thrust belt at 31°42' S (Mpodozis et al., 2009b;Perelló et al., 2012 ; Fig. 14a). The deposit is formed by mul-tiple magmatic-hydrothermal centers with ages of 12.3 to 10.8Ma, hosted by a precursor quartz diorite pluton (Los Pelam-bres stock, 13.6−13.0 Ma) and adjacent andesitic (Abanico

348 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 348

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 349

0361-0128/98/000/000-00 $6.00 349

Con

del

l

70º1

8'70

º12'

33º6

'

70º2

4'

Río

Tot

oral

Los Pelambres fault

Río Santa Cruz

Cordillera de Santa Cruz

70º3

0'

Up

per

Pal

eozo

ic b

asem

ent

33º1

2'

Cun

cum

én

Mio

cene

vol

cani

c ro

cks

(Far

ello

nes

Form

atio

n)

0

2

4km

?

RíoSanFrancis

co

EsteroYerbaLoca

31º4

0'

31º5

0'

32º0

0'

70º2

0'

Pre

tect

onic

px

dio

rites

and

gra

nod

iorit

es (2

3-21

Ma)

Pos

t te

cton

ic m

onzo

dio

rites

and

hb

-bea

ring

gran

dio

rites

and

dac

ites

(16-

15 M

a)

Cha

ling

a In

trus

ive

Co

mp

lex

and

rel

ated

intr

usiv

es (2

3-15

Ma)

Pocuro fault

LOS

PE

LAM

BR

ES

0

1

0 km

Chi

le

CH

ALI

NG

AIN

TR

US

IVE

CO

MP

LEX

Mondaca fault El Y

unq

ueO

rtig

a Los

Pic

hes

Los

Sul

fato

sLa

Pal

oma

Don

oso

Bre

cha

Sur

San

Enr

ique

-Mon

olito

Infie

rnill

o

RIO

BLA

NC

O-L

OS

BR

ON

CE

S

Río

Bla

nco

El P

lom

o

La A

mer

ican

a

Sur

Sur

San

Man

uel

Los

Pel

amb

res

El

Pac

hón

Arg

entin

a

SA

N F

RA

NC

ISC

OB

AT

HO

LITH

Low

er M

ioce

ne (2

1-18

Ma)

vol

cani

c se

que

nces

Up

per

Cre

tace

ous

to E

ocen

e in

trus

ive

rock

s

Cre

tace

ous

to P

aleo

cene

vol

cani

c an

d s

edim

enta

ry u

nits

Jura

ssic

-Ear

ly C

reta

ceou

s, b

ack-

arc

sed

imen

tary

seq

uenc

esS

ynte

cton

ic (?

) gab

bro

s to

px

mon

zogr

anite

s (1

8 M

a)

La C

opa

rhyo

lite-

bre

ccia

com

ple

x (4

Ma)

Mag

mat

ic-h

ydro

ther

mal

bre

ccia

s

Dac

itic

por

phy

ry in

trus

ions

(5 M

a)

Qz

dio

ritic

por

phy

ry in

trus

ions

(ca

7 M

a)

San

Fra

ncis

co B

atho

lith

(14?

-8 M

a)

(a)

(b)

Intr

usiv

e ro

cks

asso

ciat

ed w

ith c

opp

er m

iner

aliz

atio

nat

Los

Pel

amb

res

and

El P

achó

n (1

4-8

Ma)

> 0

.5%

Cu

Str

ongl

y d

efor

med

, Olig

ocen

e to

Low

er-M

ioce

ne(3

3-18

Ma)

vol

cani

c un

it ("

Los

Pel

amb

res

Form

atio

n")

Olig

ocen

e to

Low

er M

ioce

ne (2

5-21

Ma)

vol

cani

c se

que

nces

Tria

ssic

rift

str

ata

Fron

tera

FIG

. 14.

Com

pari

son

betw

een

the

geol

ogic

set

ting

of (a

) the

Los

Pel

ambr

es-E

l Pac

hón

and

(b) R

ío B

lanc

o-L

os B

ronc

es d

istr

icts

. Des

pite

the

diff

eren

ce in

sca

le, b

oth

porp

hyry

com

plex

es a

re a

ssoc

iate

d w

ith N

-NW

−tre

ndin

g be

lts o

f int

rusi

ons

and

mag

mat

ic-h

ydro

ther

mal

bre

ccia

s em

plac

ed d

urin

g th

e w

anin

g st

ages

of l

ong-

lived

ear

-lie

r m

agm

atic

cen

ters

(C

halin

ga in

trus

ive

com

plex

, San

Fra

ncis

co b

atho

lith)

. Map

of t

he R

ío B

lanc

o-L

os B

ronc

es r

egio

n is

from

Ira

rráz

aval

et a

l. (2

010)

.

Formation) country rocks. The precursor stock and its con-tained porphyry copper mineralization were postkinematicallyemplaced along the high-angle, Los Pelambres reverse fault,which constitutes the eastern limit of the above-mentionedzone of concentrated deformation (Mpodozis et al., 2009b;Perelló et al., 2009, 2012). At a more regional scale, however,the Los Pelambres stock appears to be a satellite intrusivebody of a much larger (>250 km2) and long-lived, multistagepluton located a short distance to the west (Chalinga intrusivecomplex; Fig. 14), which was active for at least 8 m.y. between23 and 15 Ma. The complex includes a pretectonic (with ref-erence to regional deformation) suite of gabbros, pyroxenediorites, and granodiorites, with U-Pb zircon ages between 23and 21 Ma; a group of 18 Ma syntectonic olivine gabbro-dior-ites and granodiorites; and a younger and larger group ofpost-tectonic, 16 to 15 Ma hornblende-bearing granodiorites.Rocks of this latter group form the eastern margin of theChalinga intrusive complex, where they cut the traces of re-gional thrust faults (Fig. 14a).

Several small granodiorite to quartz diorite stocks andhornblende-bearing dacite porphyry intrusions, dated at 15 to13 Ma (Fig. 14a), form a NW-SE−trending string that extendsfrom the eastern Chalinga intrusive complex across the inter-national frontier to Cerro Mercedario, ~70 km to the south-east in Argentina showing a southeastward propagation of intrusive magmatism from the Chalinga complex during themiddle to late Miocene (Figs. 12, 14a). Some of these intru-sions are associated with large porphyry-related hydrothermalalteration zones, such as El Yunque (Fig. 14a). Porphyry mag-matism and copper mineralization at Los Pelambres evolvedalong this trend between ~14 and 10 Ma, whereas much morelimited data suggest that El Pachón was active between 9.2and 8.4 Ma, and the Cerro Mercedario porphyry copper de-posit at ~13 Ma (Sillitoe, 1977; Bertens et al., 2006; Perelló etal., 2012).

Río Blanco-Los Bronces

The world’s largest copper district at Río Blanco-Los Bronces(Serrano et al., 1996; Skewes et al., 2003; Frikken et al., 2005;Irarrázaval et al., 2010; Toro et al., 2012), is located, farthersouth, near the center of the former Abanico basin (Figs.11−12). Although younger than Los Pelambres, it also formedduring the final stages of evolution of a long-lived, >10-m.y,magmatic system, which includes a large premineral intrusivecomplex (San Francisco batholith; Fig, 14b); this is remarkablysimilar to the Chalinga intrusive complex of the Los Pelambresarea and was emplaced within gently folded, early Miocene(18−15 Ma) volcanic rocks (Farellones Formation; Fig. 14b).

Deckart et al. (2005, 2012) recognized that the San Fran-cisco batholith includes three main intrusive phases, emplacedbetween12 and 8 Ma, although an older U-Pb zircon age (14.7Ma) for pyroxene monzodiorites (Jerez, 2007; Deckart 2012)indicates that magma emplacement began during the middleMiocene. The main mineralization at Río Blanco-Los Broncesoccurred along an 8-km-long, N-NW−trending corridor,which commences in the San Francisco batholith and extendsfrom Río Blanco in the northwest to Los Sulfatos in the south-east (Fig. 14b). Multiple magmatic-hydrothermal brecciacomplexes associated with porphyry copper-bearing, amphi-bole-rich quartz diorite porphyry intrusions were emplaced

along this trend, beginning at 7.7 Ma, during the waningstages of the San Francisco batholith. Intrusive activity termi-nated with the emplacement of late-mineral dacitic porphyryintrusions at ~5 Ma, and the postmineral La Copa rhyolitebreccia complex at 4.7 Ma (Skewes et al., 2003; Deckart et al.,2005, 2012; Irarrázaval et al., 2010; Toro et al., 2012; Fig.14b).

El Teniente

El Teniente, 100 km south of Río Blanco-Los Bronces, islocated near the axis of the Maipo orocline, at the center ofthe deformed Coya-Machalí (Abanico) basin and 30 km westof the most internal thrust sheets of the Aconcagua fold-and-thrust belt (Fig. 15). The deposit is hosted by gentlyfolded volcanic rocks of the El Teniente volcanic complex(informal unit equivalent to the Farellones Formation),which unconformably overlies Oligocene to early Miocenevolcanic rocks of the Coya-Machalí Formation (Kay et al.,2005; Stern et al., 2010). Mineralization at El Teniente is as-sociated with a magmatic-hydrothermal center that recordsat least 6 m.y. of continuous activity between ~9 and 3 Ma.As at Los Pelambres and Río Blanco-Los Bronces, magma-tism at El Teniente includes a large premineral intrusivecomplex containing older mafic facies (andesite intrusivesills and olivine-pyroxene gabbros), described as the Te-niente Mafic Complex, and a younger core of hydrous, horn-blende-bearing intrusions (Sewell Tonalite; Cannell et al.,2005; Stern et al., 2010; Vry et al., 2010; Fig. 15). Ages forthis group of intrusions are still not well constrained (Te-niente Mafic Complex: 8.4 Ma K-Ar whole-rock; SewellTonalite: 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 witha >2-km-long, N-NW−trending body of dacite porphyry anda series of small magmatic-hydrothermal breccias (Vry et al.,2010 ), which in contrast to Los Pelambres or Río Blanco-LosBronces, were emplaced within the earlier intrusions (Te-niente Mafic Complex and Sewell Tonalite; Fig. 15). DetailedU-Pb zircon, Ar/Ar, and Re-Os geochronology (Maksaev etal., 2004) has allowed recognition of five pulses of felsic in-trusions linked to mineralization emplaced between 6.5 and4.6 Ma. Subsequently, the late-mineral 4.81 Ma Braden brec-cia pipe and a family of narrow, E-NE−oriented olivine-horn-blende lamprophyre dikes; ages from 3.9 to 2.9 Ma (Stern etal., 2011) mark the last magmatic pulses in the district beforethe magmatic front migrated ~50 km eastward to the Chile-Argentine frontier to form the northernmost active volcanoesof the modern Andean Southern volcanic zone (Fig. 12).

Origin of Miocene-Pliocene porphyry copper alignments

One of the most striking features of porphyry copper de-posits in central Chile is their relationship to N-NW−trendingalignments of multiple magmatic-hydrothermal centers, em-placed during the final stages of long-lived magmatic systemsthat include breccia complexes whose formation has beenconsidered to be triggered by decompression during rapid ex-humation and tectonic uplift (Warnaars et al., 1985; Skewesand Stern, 1994).

When the overall geometry of the porphyry alignments ofthe three main centers described in this paper are considered,

350 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 350

their orientation is similar and almost perpendicular to the di-rection of the Miocene and Pliocene plate convergence (So-moza and Ghidella, 2005). In contrast, the late lamprophyredikes at El Teniente (Fig. 15) are nearly parallel to theMiocene plate convergence vector (see Fig. 16). Recent stud-ies on the relationships between volcanism and tectonics inthe modern Southern volcanic zone of the Andes (Sepúlvedaet al., 2005; Cembrano and Lara, 2009) have shown thatprimitive mantle-derived basalts ascend along NE-trendingfractures and faults, parallel to σHmax, which is regionallyclose to the orientation of plate convergence. Nevertheless,during the recent eruptions of southern Andes volcanoes likePuyehue-Cordón Caulle (40°30' lat S, 1960, 2011) evolvedrhyolites erupted along NW-trending basement faults thatare, in theory, severely disoriented in relationship to re-gional stresses to allow magma ascent. To overcome thisparadox, Sepúlveda et al. (2005) and Lara et al. (2006) havesuggested that coseismic or postseismic stress relaxationduring large subduction-zone earthquakes can producetransient episodes of extension that allow ascent of evolvedmagmas that otherwise would remain trapped in crustalreservoirs. Similarly, decompression during seismic events(see Sibson, 1987, 1994) appears to be a plausible mecha-nism to explain repeated cycles of breccia formation andporphyry intrusion along now-concealed N-NW− to NE-ori-ented faults that may have tapped the roofs of overpres-sured, deeper seated magma chambers during the evolutionof the central Chile porphyry coppers systems.

Discussion

Linking geochemistry and tectonics: the suggested adakite connection

Both the middle Eocene to early Oligocene and late Mioceneto early Pliocene porphyry copper-bearing intrusions includeintermediate rocks (SiO2 >56%) with abundant hydrous min-eralogy, dominated by hornblende-bearing granodiorites anddacites; these intrusions show geochemical and isotopic sig-nature (SiO2 >56%, Sr >400 ppm, high Sr/Y ratios, lowHREE contents, high La /Yb ratios, 87Sr/86Sr <0.704) similarto those described for typical adakitic rocks (cf. Defant andDrumond, 1990; Castillo, 2012). Concave middle REE pat-terns in Chilean porphyry coppers indicate hornblende frac-tionation, whereas the lack of negative europium anomaliesdenotes a high oxidation state of the magmas (see Kay et al.,2005).

A relationship between adakites and mineralized porphyrysystems was proposed by Thiéblemont et al. (1997) and Kayand Mpodozis (2001), while some authors (e.g., Sun et al.,2011) have suggested that the adakitic signatures of copper-rich magmas are indicative of direct melting of subductedoceanic crust. In accord with these views Oyarzún et al.(2001) proposed that the middle Eocene to early Oligoceneporphyry copper belt of northern Chile may have beenformed when fast, oblique convergence led to flat subductionand direct melting of the downgoing plate, whereas Reich etal. (2003) proposed that the porphyry copper intrusions at

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 351

0361-0128/98/000/000-00 $6.00 351

Teniente Volcanic Complex

Braden Pipe (4.81±0.10 Ma)

Hydrothermal brecciasMafic Complex (8.9±14.4 Ma)

Sewell Tonalite (7.05 0.14 Ma)±

Laguna La Negra

Laguna La Huifa

BradenPipe

6232

6228

6224

N Mine

1600

800

00

800

MarginalBreccia

1200

Level Teniente 5(2.284m)

Teniente Dacite Porphyry5.28±0.10 Ma

Latite Dike4.82±0.09 Ma

Porphyry A5.67±0.19 Ma

Daciticporphyries

6.09±0.18 Ma

Braden Pipe4.81±0.10 Ma

Sewell Tonalite7.05±0.14 Ma

Mafic Complex8.9±1.4 Ma?

400

Biotite breccia Igneous breccia Anhydrite breccia Tourmaline breccia

Mafic Complex Sewell Tonalite Porphyry "A" Dacite Porphyry

Braden Pipe

Supergene zone

(a)(b)

Lamprophyre dykes

0 2 km

0 400 km

Unconsolidated deposits

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

Los Pelambres formed by melting of the subducting east-northeast arm of the Juan Fernández Ridge under flat-slabconditions. However, as noted by Kay and Kay (2002) andCastillo (2012), thermal models for subducting plates showthat sufficiently high pressure-temperature conditions forslab fusion can be reached only in exceptional circumstances,including subduction of very young and hot oceanic crust, asprevailed during emplacement of the 12 Ma Cerro Pampaadakites near the Chile Triple Junction in Patagonia (Kay andKay, 2002).

Richards and Kerrich (2007) and Richards (2011b) stronglyargued against slab melts being a necessary ingredient in porphyry copper-gold mineralization, pointing out that thecritical factors for adakite genesis include elevated water andsulfur contents as well as high oxidation state of the magmas,which together result in hornblende fractionation and sup-pression of plagioclase crystallization. Richards (2011b) con-sidered that such hydrous, oxidized conditions are typical innormal arc settings. Nevertheless, this view is inconsistentwith the fact that giant porphyry copper deposits are notwidespread throughout the Andean history; as we have shown

they formed during and/or after the most important tectonicepisodes that reshaped the whole Andean margin leading tomountain building and crustal thickening.

Kay et al. (1999) and Kay and Mpodozis (2001) argued thatpart of the water content of mineralizing Andean magmas canbe derived from the exsolution of fluids during the transfor-mation of hydrous lower crustal amphibolite to dry garnet-bearing eclogite during crustal thickening. When the crust ofthe arc thickens to a critical value of ~45 km, amphibole andplagioclase break down, water is liberated, and eclogite forms.These relationships explain why the largest porphyry copperdeposits preferentially form during contractional events, suchas the Incaic episode of southern Peru and northern Chile,and the late Miocene to early Pliocene event in central Chileand contiguous Argentina.

Another process that can also contribute to generation ofwater-rich adakitic magmas during periods of deformation issubduction erosion (von Heune and Scholl, 1991; Stern, 1991,2011; Kay et al., 2005). Subduction of oceanic crust, pelagicand terrigenous sediments, and crust tectonically erodedfrom the edge of the continent into the mantle-source region

352 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 352

140 130 120 110 100 80 70 60 50 40 30 20 1090

East PacificWest PacificSouth Pacific

10

7

6

5

4

3

2

1

0

Averagepreserved

half-spreadingrate

(cm/yr)

9

0

Age (Ma)

(a)

8

80 70 60 50 40 30 20 10 0

2

1

0

Velo

city

(cm

/yr)

E W velocity

S N velocity

Age (Ma)

(d)0

0(E-W)

-40º

-20º

+20º

15

10

5

4

3

4

1

Convergencevelocity(cm/yr)

Angle betweenconvergence vector

and present-daySouth American

margin

(b)

(c)

Age (Ma)

2

60 50 30 20 0 Ma1040

FIG. 16. Temporal changes of critical plate parameters that can be linked to adjustments in the tectonic regime along theAndean margin. (a). Ocean crust production (average half-spreading rates since 140 Ma) for different regions of the Pacificbasin (Conrad and Lithgow-Bertelloni, 2007). (b). Absolute velocity of the South American plate since 80 Ma, treated as anangular velocity vector decomposed into its ~ E to W and S to N components (Silver et al., 1998). (c). Cenozoic convergencerates between the Farallon-Nazca and South American plates. Data from: 1 = Sdrolias and Muller (2006), 2 = Pardo-Casasand 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 andMiocene to Pliocene.

of Andean magmas may provide large amounts of water (e.g.,Stern et al., 2010). Transient geochemical changes, such asthose depicted in Figure 1d, showing adakitic signatures(Haschke et al., 2002; Kay et al., 2005) are consistent with theloss of slivers of fore-arc crust by subduction erosion (Fig. 13)during, or immediately following, major Mesozoic to Ceno-zoic deformation events. Intermittent and massive loss offore-arc crust is a possibility that has been recognized in re-cent numerical models of subduction erosion reported byKeppie et al. (2009) and may explain the abrupt eastwardshifts of the magmatic front that punctuated the tectonic his-tory of the Central Andes. Without ignoring other models foradakite formation (e.g., Castillo, 2012), high Sr/Y hydrousmagmas linked to the middle Eocene to early Oligocene andlate Miocene to early Pliocene porphyry copper deposits ofnorthern and central Chile, respectively, are most likely at-tributed to a combination of melting of mantle derived mag-mas, including asthenosphere contaminated with pieces offore-arc crust that entered the mantle through subductionerosion, that mixed with fluids derived from dehydration ofthe base of thickened amphibole-rich lower crust duringthese two main periods of Andean deformation (e.g., Kay etal., 2007b).

Plate tectonics and Andean tectonic regimes

As discussed above, giant Andean porphyry copper de-posits were emplaced during regional deformation eventsthat reshaped the entire Andean margin. Regional tectonicand tectonomagmatic events are, ultimately, the result ofchanges in plate interaction parameters that could influencethe state of stress in the overriding plate. Among thesechanges on sea-floor spreading and plate convergence rates,hot-spot activity, absolute plate velocity, and shifts in the po-sition of the plate contact (trench roll-back), together withvariations in slab age, width, and/or dip has been consideredto strongly influence the dynamics of convergent marginsworldwide (e.g., Uyeda and Kanamori, 1979; Heuret andLallemand, 2005; Schellart and Rawlinson, 2010, and refer-ences therein). In recent years, variations in absolute platevelocities (Russo and Silver, 1996; Silver et al., 1998) as wellas variations on the degree of mechanical interplate cou-pling (Yañez and Cembrano, 2004; Luo and Liu 2009a, b;Iaffaldano et al., 2012) have been proposed as some of thefundamental controls on Andean orogeny. Lamb and Davis(2003) even suggested that differences in plate coupling arepossibly linked to along-strike differences in climate thatmodulated the supply of sediments to the trench along theAndean margin which, upon subduction, lubricated theplate interface and, as a result, determined the degree ofmechanical coupling.

Tectonic regime: Northern Chile

Figure 16 shows time-dependent variations for some platetectonic parameters that can be compared with the geologicrecord of the Andean margin. A correlation can be made be-tween the Incaic compressional event and an episode of veryrapid oceanic crust production in the eastern Pacific, whichpeaked in the middle Eocene at ~40 Ma (Fig. 16a; Conradand Lithgow-Bertelloni, 2007). At the same time the east-west component of the absolute velocity vector of the South

American plate (Silver et al., 1998) sharply increased (Fig.4b). However, as shown in Figure 4c these effects were notbalanced, as should have occurred, by an associated increasein the Farallon-South America convergence rate. Accordingto Somoza (1998) and Somoza and Ghidella (2005), the Far-allon-South America convergence velocity was rather low(6−7 cm/yr) or, as alternatively suggested by Pardo-Casas andMolnar (1987) and Sdrolias and Muller (2006), was rapidlydecreasing (Fig. 16c).

This apparent paradox can be resolved if a high degree ofcoupling existed at the plate interface at that time. If this wasthe case for northern Chile and southern Peru during theEocene, the mechanically weak margin of the Central Andes,pushed from the east and west, may have started to bend andcontract to form the Bolivian orocline and the Domeyko faultsystem. High interplate coupling may also explain the forma-tion of a flat-slab zone and the subsequent Oligocene mag-matic lull recorded in southern Peru and northern Chile (e.g.,Sandeman et al., 1995; James and Sacks, 1999; Kay et al., 1999;Perelló 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, inturn, would promote eastward migration of the shortening andmagmatic fronts toward the Andean foreland (e.g., Espurt etal., 2008; Keppie et al., 2009; Martinod et al., 2010). Conver-gence rates increased during the Oligocene and Miocene,leading to steepening of the Incaic flat slab; accumulation ofrelated mafic magmas below the already thickened crust ledto the production of crustal melts and the eruption of largevolumes of ignimbrites during the inception of volcanismalong the modern Central Andean volcanic zone during theMiocene (Kay and Coira, 2009).

Tectonic regime: Central Chile

Although the tectonic evolution of central and south-cen-tral Chile appears to be different from that of northern Chile,it is in fact complementary and preconditioned by the forma-tion of the Bolivian orocline. As noted above, crustal flux to-ward the north along the southern limb of the orocline duringthe middle Eocene to early Oligocene Incaic contraction mayhave caused stretching and thinning of the upper crust in cen-tral Chile, thereby facilitating the opening of the Abanicobasin (McQuarrie, 2002; Arriagada et al., 2008; Boutelier andOncken, 2010). The along-strike differences in tectonic stylealso agree with the numerical simulations by Schellart (2008),which show that higher compression occurs near the center ofwide subduction zones where the trench remains stationaryor advances toward the continent. Rapid trench retreat (roll-back) along the lateral slab edges may explain extension in theoverriding plate and be compatible with the opening of theAbanico basin in central Chile.

The late Eocene to Oligocene opening of the Abanico basinoccurred during a period of steady westward displacement ofthe South America plate (Silver et al., 1998), a period that alsocoincided with another transient episode of very rapid gener-ation of oceanic crust in the eastern Pacific, which com-menced immediately after the formation of the Nazca plate(Conrad and Lithgow-Bertelloni, 2007; Fig. 16a). In contrastto the situation during the Incaic event, the convergence ratebetween the Nazca and South American plates reached, at

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 353

0361-0128/98/000/000-00 $6.00 353

this time, a record high (15 cm/yr; Somoza, 1998; Sdrolias andMuller, 2006; Fig. 16b). Such a disparity may be explained ifweak plate coupling permitted rapid subduction and, as aconsequence, the generation of large volumes of magma andextension in the overriding plate during the Oligocene toearly Miocene in central Chile and contiguous Argentina.

The beginning of contraction in central Chile and contigu-ous 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 decreasein oceanic crust production in the eastern Pacific, and a dropin plate convergence rates between the Nazca and SouthAmerican plates (Somoza, 1998; Conrad and Lithgow-Bertel-loni, 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 etal., 2004) and enhanced subduction erosion. Contaminationof the asthenosphere through subduction of fore-arc crustcreated favorable conditions to produce water-rich maficmelts with high sulfur and metal contents; these melts hadthe capacity to ascend and evolve within an upper crustalmagma chamber to generate large porphyry copper deposits.

Concluding RemarksThere are few studies that consider the relationships be-

tween the regional-scale tectonic evolution of the Andes andthe formation of giant Cenozoic porphyry copper deposits.However, it is apparent that these deposits formed duringcritical moments in the tectonic evolution of the Andean mar-gin. The emplacement of the large middle Eocene to earlyOligocene porphyry copper intrusions in northern Chileseems to be associated with the formation of the Bolivian orocline during the Incaic event, which was the result of anunusual combination of factors. One critical factor was the ac-celeration of the absolute westward motion of South Americaconcurrent with strong mechanical coupling between theSouth American and Farallon plates at a time when the rateof ocean-crust production in the eastern Pacific was veryhigh. Bending of the Chilean margin during the Incaic eventactivated the Domeyko fault system in northern Chile andtriggered the accompanying crustal thickening, slab shallow-ing, and increased subduction erosion. Volcanism virtuallyceased and favorable tectonomagmatic conditions (i.e. en-hanced, subduction erosion, crustal thickening, lower crustdehydration) permitted the formation of fertile hydrous mag-mas, while the transpressional and/or compressional upperplate tectonic regime contributed to the establishment oflong-lived, upper-crustal magma chambers from which con-centrating copper evolved, mostly below the Domeyko faultsystem.

Younger, late Miocene to early Pliocene porphyry copperdeposits of central Chile and contiguous Argentina were emplaced after inversion and collapse of the extensional,intra-arc Abanico basin. The basin evolved between the lateEocene and early Miocene when a relatively stable position ofSouth America over the mantle, linked to weak interplatecoupling, permitted fast subduction of the Nazca plate underthe Andean margin. Acceleration of the westward motion ofSouth America relative to the mantle reference frame at 20Ma induced contractional deformation, accompanied by

crustal thickening and eastward migration of the magmaticfront. At the same time, the subduction angle shallowed,leading to the formation of a flat-slab region between 27° and33° S as the Juan Fernández Ridge was being subducted be-neath the western edge of South America. These changesagain created favorable conditions for the formation of fertilehydrous magmas.

The relationships described above demonstrate that theconcentration of huge porphyry copper deposits in theChilean Andes resulted directly from the tectonic evolutionof the margin and indicate that a tectonic trigger is essentialfor the formation of giant porphyry coppers systems.

AcknowledgmentsThis contribution is the result of long years of work with

many colleagues at the Chilean Geological Survey, AntofagastaMinerals, and various universities both in Chile and abroad.We are especially grateful to Sue Kay, Andy Tomlinson, TerryJordan, Cesar Arriagada, Moyra Gardeweg, Rick Allmendinger,Victor Ramos, Pierrick Roperch, Francisco Camus, StephenMatthews, Nicolás Blanco, Francisco Hervé, Reynaldo Char-rier, Carlos Münchmeyer, Ricardo Muhr, José Cembrano,Carlos Arévalo, and many others who, for lack of memory, weare here unable to mention. We thank Jeff Hedenquist, DickSillitoe, José Perelló, and Francisco Camus for pushing usthrough this endeavor, and Antofagasta Minerals for provid-ing time and support for the writing. Francisco Moraleshelped with the preparation of the figures. Victor Ramos, SueKay, José Perello, Jeff Hedenquist, and Dick Sillitoe carefullyedited the manuscript and made numerous suggestions thathelped to greatly improve earlier versions of the manuscript.

REFERENCESAmilibia, A., and Skarmeta, J., 2003, La inversión tectónica en la Cordillera

de Domeyko en el norte de Chile y su relación con la intrusión de sistemasporfíricos de Cu-Mo [ext.abs.]: X Congreso Geológico Chileno, Concep-ción, Extended Abstracts, v. 2, p. 1−7.

Amilibia, A., Sabat, F., McClay, K.R., Muñoz, J.A., and Chong, G., 2008, Therole of inherited tectono-sedimentary architecture in the development ofthe central Andean mountain belt: Insights from the Cordillera deDomeyko: Journal of Structural Geology, v. 30, p. 1520−1539.

Anderson, M., Alvarado, P., Zandt, G., and Beck, S., 2007, Geometry andbrittle deformation of the subducting Nazca Plate, Central Chile and Ar-gentina: Geophysical Journal International, v. 171, p. 419–434.

Arriagada, C., Roperch, P., Mpodozis, C., and Cobbold, P.R., 2008, Paleo-gene building of the Bolivian orocline: Tectonic restoration of the CentralAndes in 2-D map view: Tectonics, v. 27, TC6014, 14 p., doi:10.1029/2008TC002269.

Arriagada, C., Mpodozis, C., Yañez, G., Charrier, R., Farías, M., and Ro-perch, P., 2009, Rotaciones tectónicas en Chile central: El oroclino de Va-llenar y el “megakink” del Maipo [ext.abs.]: XII Congreso Geológico Chi-leno Santiago, Extended Abstracts (CD), 4 p.

Astini, R.A., Benedetto, J.L., and Vaccari, N.E., 1995, The early Paleozoicevolution of the Argentine Precordillera as a Laurentian rifted, drifted, andcollided terrane: A geodynamic model: Geological Society of America Bul-letin, v. 107, p. 253–273.

Audin, L., Hérail, G., Riquelme, R., Darrozes, J, Martinod, J., and Font, E.,2003, Geomorphological markers of faulting and neotectonic activity alongthe western Andean margin, northern Chile: Journal of Quaternary Sci-ence, v. 18, p. 681–694.

Ballard, J., Palin, J.R., Palin, M., Williams, I., and Campell, I., 2001, Two agesof porphyry intrusion resolved for the super-giant Chuquicamata copperdeposit of northern Chile by ELA-ICP-MS: Geology, v. 29, p. 383–386.

Barra, F., 2011, Assessing the longevity of porphyry Cu-Mo deposits: Exam-ples from the Chilean Andes: SGA Biennial Meeting, 11th, Antofagasta,Proceedings, p. 399−401.

354 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 354

Beck, M., Rojas, C., and Cembrano, J., 1993, On the nature of buttressing inmargin parallel strike-slip fault systems: Geology, v. 21, p. 755−758.

Behn, G., Camus, F., Carrasco P., and Ware, H., 2001, Aeromagnetic signa-ture of porphyry copper systems in northern Chile and its geologic impli-cations: Economic Geology, v. 96, p. 239−248.

Bertens, A., Clark, A.H., Barra, F., and Deckart, K., 2006, Evolution of theLos Pelambres-El Pachón porphyry copper-molybdenum district, Chile/Argentina [ext. abs.]: XI Congreso Geológico Chileno, Antofagasta, Ex-tended Abstracts, v. 2, p. 179–181.

Bisso, C., Lazcano, E., Guzmán, J., and González, S., 2009, Geología y desarro-llo del yacimiento Esperanza, Distrito Sierra Gorda, Antofagasta [ext. abs.]:XII Congreso Geológico Chileno, Santiago, Extended Abstracts (CD), 4 p.

Blanco, N., 2008, Estratigrafía y evolución tectono-sedimentaria de la cuencacenozoica de Calama (Chile, 22° S): Unpublished M.Sc. thesis, Universityof Barcelona, 68 p.

Blanco, N., and Tomlinson, A.J., 2009, Carta Chiu-Chiu, Región de Antofa-gasta: Santiago, Servicio Nacional de Geología y Minería, Carta Geológicade Chile, 117, 1:50,000, 54 p.

Boric, R., Díaz, F., and Maksaev, V., 1990, Geología y yacimientos metalífe-ros de la región de Antofagasta: Santiago, Servicio Nacional de Geología yMinería, Boletín 40, 246 p.

Boric, R., Díaz, J., Becerra, H., and Zentilli, M., 2009, Geology of the Min-istro Hales mine (MMH), Chuquicamata district, Chile [ext. abs.]: XIICongreso Geológico Chileno, Santiago, Extended Abstracts (CD), 4 p.

Boutelier, D.A., and Oncken, O., 2010, Role of the plate margin curvature inthe plateau buildup: Consequences for the central Andes: Journal of Geo-physical Research, v 115, B04402, 27 p., doi:10.1029/2009JB006296.

Cahill, T. A. and Isacks, B. L, 1992, Seismicity and shape of the subductedNazca Plate: Journal of Geophysical Research, v. 97, p. 503–517.

Campbell, I.H., Ballard J.R., Palin J.M., Allen C., and Faunes, A., 2006, U-Pb zircon geochronology of granitic rocks from the Chuquicamata-El Abraporphyry copper belt of northern Chile: Excimer laser ablation ICP-MSanalysis: Economic Geology, v. 101, p. 1327−1344.

Camus, F, 2003, Geología de los sistemas porfíricos en los Andes de Chile:Santiago, Servicio Nacional de Geología y Minería, 267 p.

Cannell, J., Cooke, D.R., Walshe, J.L., and Stein, H., 2005, Geology, miner-alization, alteration, and structural evolution of the El Teniente porphyryCu-Mo deposit: Economic Geology, v. 100, p. 979–1003.

Castillo, P.R., 2012, Adakite petrogenesis: Lithos, v. 134−135, p. 304−316.Cawood, P.A, 2005 Terra Australis orogen: Rodinia breakup and develop-

ment of the Pacific and Iapetus margins of Gondwana during the Neopro-terozoic and Paleozoic: Earth-Science Reviews, v. 69, p. 249–279.

Cembrano, J., and Lara, L., 2009, The link between volcanism and tectonicsin the southern volcanic zone of the Chilean Andes: A review: Tectono-physics, v. 472, p. 96−113.

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 Meso-zoic-early Cenozoic evolution of the Chilean Andes in the upper Tinguirir-ica valley (35º S), Central Chile: Journal of South American Earth Sciences,v. 9, p. 1−30.

Charrier, R., Baeza, O., Elgueta, S., Flynn, J.J., Gans, P., Kay, S.M., Muñoz,N., Wyss, A.R., and Zurita, E., 2002, Evidence for Cenozoic extensionalbasin development and tectonic inversion south of the flat-slab segment,southern Central Andes, Chile (33°–36° SL): Journal of South AmericanEarth Sciences, v. 15, p. 117–139.

Chen, H., Clark, A.H., Kyser, T.K., Ullrich, T.D., Baxter, R., Chen, Y., andMoody, T.C., 2010, Evolution of the giant Marcona-Mina Justa iron oxide-copper-gold district, south-central Peru: Economic Geology, v. 105, p.155–185.

Chew, D.M., Schaltegger, U., Košler, J., Whitehouse, M., Gutjahr, M., Spik-ings, R., and Miškovíc, A., 2007, U-Pb geochronologic evidence for the evo-lution of the Gondwanan margin of the north-central Andes: Geological So-ciety of America Bulletin, v. 119, p. 697−711.

Conrad, C., and Lithgow-Bertelloni, C., 2007, Faster seafloor spreading andlithosphere production during the mid-Cenozoic: Geology, v. 35, p. 29–32.

Cooke, D.R., Hollings, P., and Walshe, J.L., 2005, Giant porphyry deposits:Characteristics, distribution, and tectonic controls: Economic Geology, v.100, p. 801–818.

Cornejo, P., and Matthews, S., 2001, Evolution of magmatism from the up-permost Cretaceous to Oligocene, and its relationship to changing tectonicregime in the Inca de Oro-El Salvador area (northern Chile) [ext. abs.]:South American Symposium on Isotope Geology, 3rd, Pucón, Chile, Ex-tended Abstracts (CD), p. 558−561.

Cornejo, P., Tosdal, R.M., Mpodozis, C., Tomlinson, A., Rivera, O., and Fan-ning, M.C., 1997, El Salvador, Chile, porphyry copper deposit revisited:Geologic and geochronologic framework: International Geology Review, v.39, p. 22–54.

Cornejo, P., Matthews, S., and Pérez de Arce, C., 2003, The “K-T” compres-sive deformation event in northern Chile (24º−27º S) [ext. abs.]: X Con-greso Geológico Chileno, Concepción, Extended Abstracts (CD), The-matic Session ST1, p. 1−13.

Cornejo, P., Matthews, S., Marinovic, N., Pérez de Arce, C., Basso, M., Al-faro J., and Navarro, M., 2006, Alteración hidrotermal y mineralización re-currente de Cu y Cu-Mo durante el Pérmico y el Triásico en la Cordillerade Domeyko (Zona de Zaldívar-Salar de los Morros): Antecedentes geo-cronológicos U-Pb, 40Ar-39Ar y Re-Os [ext. abs.]: XI Congreso GeológicoChileno, Antofagasta, Extended Abstracts, v. 2, p. 219−222.

Cunningham, W.D., and Mann, P., 2007, Tectonics of strike-slip restrainingand releasing bends: Geological Society Special Publication 290, p. 1–12.

Dalziel, I.W.D., 1997, Neoproterozoic-Paleozoic geography and tectonics:Review, hypothesis, environmental speculation: Geological Society ofAmerica Bulletin, v. 109, p. 16−42.

Deckart, K., Clark, A.H., Aguilar, A.C., Vargas, V.R., Bertens, A.N.,Mortensen, J.K., and Fanning, M., 2005, Magmatic and hydrothermalchronology of the giant Rio Blanco porphyry copper deposit, central Chile:Implications of an integrated U-Pb and 40Ar-39Ar database: Economic Ge-ology, v. 100, p. 905–934.

Deckart, K., Clark, A.H., Aguilar, C., Cuadra, P., and Fanning, M., 2012, Re-finement of the time-space evolution of the giant Mio-Pliocene Río Blanco-Los Bronces porphyry Cu-Mo cluster, Central Chile: New U-Pb (SHRIMPII) and Re-Os geochronology and 40Ar-39Ar thermochronology data: Min-eralium Deposita (in press, available online)

Defant, M.J., and Drummond, M.S., 1990.Derivation of some modern arc mag-mas by melting of young subducted lithosphere: Nature, v. 347, p. 662−665.

Dilles, J., Tomlinson, A., Martin, M., and Blanco, N., 1997, El Abra and For-tuna complexes: A porphyry copper batholith sinistrally offset by the FallaOeste [ext. abs.]: VIII Congreso Geológico Chileno, Antofagasta, ExtendedAbstracts, v. 3, p. 1883−1887.

Dilles, J., Tomlinson, A., García, M., and Alcota, H., 2011, The geology of theFortuna Granodiorite Complex, Chuquicamata district, northern Chile:Relation to porphyry copper deposits: SGA Biennial Meeting, 11th, Antofa-gasta, Proceedings, p. 399−401.

Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna,C., and Brusset, S., 2008, Flat subduction dynamics and deformation of theSouth American plate: Insights from analog modeling: Tectonics, v. 27,TC3011, doi:10.1029/2007TC002175, 2008.

Farías, M., Charrier, R., Carretier, S., Martinod, J., Fock, A., Campbell, D.,Cáceres, J., and Comte, D., 2008, Late Miocene high and rapid surface up-lift and its erosional response in the Andes of central Chile (33°–35° S):Tectonics, v. 27, TC1005, doi:10.1029/2006TC002046.

Franzese, J., and Spalleti, L., 2001, Late Triassic-Early Jurassic continentalextension in southwestern Gondwana, tectonic segmentation and pre-breakup rifting: Journal of South American Earth Sciences, v. 14, p.257−270.

Frikken, P.H., Cooke, D.R., Walshe, J.L., Archibald, D., Skarmeta, J., Ser-rano, L., and Vargas, R., 2005, Mineralogical and isotopic zonation in theSur-Sur tourmaline breccia, Río Blanco-Los Bronces Cu-Mo deposit,Chile: Implications for ore genesis: Economic Geology, v. 100, p. 935−961.

Giambiagi, L., Ramos, V., Godoy, E., Alvarez, P., and Orts, S., 2003, Cenozoicdeformation and tectonic style of the Andes, between 33° and 34° south lat-itude: Tectonics, v. 22 1041, doi:10.1029/2001TC001354.

Giambiagi, L., Mescua, J., and Bechis, F., 2011, El Frente orogénico del sec-tor sur de los Andes centrales.Variaciones latitudinales en acortamiento, to-pografía, levantamiento estructural y denudación [ext. abs.]: XVIII Con-greso Geológico Argentino, Neuquén, Extended Abstracts (CD), 2 p.

Gilbert, H., Beck, S., and Zandt, G., 2006, Lithospheric and upper mantlestructure of central Chile: Geophysical Journal International, v. 165, p.383–398.

Gotberg, N., McQuarrie, N., and Carlotto, V., 2010, Comparison of crustalthickening budget and shortening estimates in southern Peru (12°–14° S):Implications for mass balance and rotations in the “Bolivian orocline:” Ge-ological Society of America Bulletin, v. 122, p. 727–742.

Götze, H.J., and Krausse, S., 2002, The Central Andean gravity high. A relicof an old subduction complex?: Journal of South American Earth Sciences,v. 14, p. 799−811.

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 355

0361-0128/98/000/000-00 $6.00 355

Hammerschmidt, K., Dobel, R., and Friedrichsen, H., 1992.Implication of40Ar-39Ar dating of early Tertiary volcanic-rocks from the North-ChileanPrecordillera: Tectonophysics, v. 202, p. 55–81.

Haschke, M., Siebel W., Günther A., and Scheuber, E., 2002, Repeatedcrustal thickening and recycling during the Andean orogeny in North Chile(21°–26° S): Journal of Geophysical Research, v. 107, doi 10.1029/2001JB000328.

Haschke, M., Günther, A., Melnick, D., Echtler, H., Reutter, K.J., Scheuber,E., and Oncken, O., 2006, Central and southern Andean tectonic evolutioninferred from arc magmatism, in Oncken, O., et al., eds., The Andes: Fron-tiers in Earth Sciences, Pt. III: Springer-Verlag, p. 337−353,

Hervé, F., 1988, Late Paleozoic subduction and accretion in southern Chile:Episodes, v. 11, p. 183–188.

Hervé, M., Sillitoe, R.H., Wong, C., Fernández, P., Crignola, F., Ipinza, M.,and Urzúa, F., 2012, Geologic overview of the Escondida porphyry copperdistrict, northern Chile: Society of Economic Geologists Special Publica-tion, 16, p. 55–78.

Heuret, A., and Lallemand, S., 2005.Plate motions, slab dynamics and back-arc deformation: Physics of the Earth and Planetary Interiors, v. 149, p.31–51.

Hindle, D., Kley, J., Oncken, O., and Sobolev, S., 2005.Crustal balance andcrustal flux from shortening estimates in the Central Andes: Earth andPlanetary Science Letters, v. 230, p. 113–124.

Hoffmann-Rothe, A., Ritter, O., and Janssen, C., 2004, Correlation of elec-trical conductivity and structural damage at a major strike-slip fault innorthern 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-EasternCordillera transition (23°–26° S): Control by preexisting heterogeneities onthe pattern of initial Andean shortening: Geology, v. 35, p. 271–274.

Iaffaldano, G., Di Giuseppe, E., Corbi, F., Funiciello, F., Faccenna, C., andBunge, H., 2012, Varying mechanical coupling along the Andean margin:Implications for trench curvature, shortening and topography: Tectono-physics, v. 526–529, p. 16−23.

Introcaso, A., Pacino, M.C., and Fraga, A.,1992, Gravity, isostasy and Andeancrustal shortening between latitudes 30° and 35° S: Tectonophysics, v. 205,p. 31−38.

Irarrázaval, 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 porphyrycopper-molybdenum deposit at Los Sulfatos, Los Bronces-Río Blanco dis-trict, central Chile: Society of Economic Geologists Special Publication 15,p. 253−269.

Isacks, B.L., 1988, Uplift of the Central Andean plateau and bending of theBolivia orocline: Journal of Geophysical Research, v. 93, p. 3211–3231.

James, D., and Sacks, I.S., 1999, Cenozoic formation of the central Andes: Ageophysical perspective: Society of Economic Geologists Special Publica-tion 7, p. 1–25.

Jerez, C, 2007, Contribución a la geocronología y geoquímica de los intrusi-vos Yerba Loca y San Francisco, Cordillera de Chile central (33° S): Unpu-blished 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. 341–361.

Jordan, T.E., Burns, W., Veiga, R., Pángaro, F., Copeland, P., Kelley, S., andMpodozis, C., 2001, Extension and basin formation in the Southern Andescaused by increased convergence rate: A Mid-Cenozoic trigger for theAndes: Tectonics, v. 20, p. 308–324.

Jordan, T.E., Mpodozis, C., Blanco, N., Pananont, P., and Dávila, F., 2004,Extensional basins in a convergent margin: Oligocene-Miocene Salar deAtacama and Calama basins, Central Andes [abs.]: EOS, v. 85(47), FallMeeting 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. 303−311.

Kay, S.M., and Coira, B., 2009, Shallowing and steepening subduction zones,continental lithosphere loss, magmatism and crustal flow under the centralAndean Altiplano-Puna plateau: Geological Society America Memoir 204,p. 229–260.

Kay, S.M., and Copeland, P., 2006, Early to middle Miocene back-arc mag-mas of the Neuquén basin of the southern Andes: Geochemical conse-quences of slab shallowing and the westward drift of South America(35°–39° S lat.): Geological Society of America Special Paper 407, p.185–213.

Kay, S.M., and Mpodozis, C., 2001, Central Andean ore deposits linked toevolving shallow subduction systems and thickening crust: GSA Today, v.11(3), p. 4−9.

——2002, Magmatism as a probe to the Neogene shallowing of the Nazcaplate beneath the modern Chilean flat-slab: Journal of South AmericanEarth Sciences, v. 15, p 39–57.

Kay, S.M., Maksaev, V.A., Moscoso, R., Mpodozis, C. and Nasi, C., 1987, Pro-bing the evolving Andean lithosphere: Mid-late Tertiary magmatism inChile (29°–30°30’ S) over the modern zone of subhorizontal subduction:Journal of Geophysical Research, v. 92, p. 6173–6189.

Kay, S.M., Ramos, V.A., Mpodozis, C., and Sruoga, P., 1989, Late Paleozoicto Jurassic silicic magmatism at the Gondwanaland margin: Analogy to theMiddle Proterozoic in North America?: Geology, v. 17, p. 324−328.

Kay, S.M., Mpodozis, C., and Coira, B., 1999, Magmatism, tectonism andmineral deposits of the Central Andes (22°–33°S latitude) : Society of Eco-nomic Geologists Special Publication 7, p. 27–59.

Kay, S.M., Godoy, E., and Kurtz, A., 2005, Episodic arc migration, crustalthickening, subduction erosion, and magmatism in the south-centralAndes: Geological Society of America Bulletin, v. 117, p. 67−88.

Kay, S.M., Mancilla, O., and Copeland, P., 2006, Evolution of the backarcChachahuén volcanic complex at 37°S latitude over a transient Mioceneshallow subduction zone under the Neuquén basin: Geological Society ofAmerica Special Paper 407, p. 215−246.

Kay, S.M., Ardolino, A., Gorring, M., and Ramos, V.A., 2007a, The So-muncura large igneous province in Patagonia: Interaction of a transientmantle thermal anomaly with a subducting slab: Journal of Petrology, v. 48,p. 43−77.

Kay, S.M., Goss, A., Kay, R., Mpodozis, C., 2007b, Frontal arc migration,forearc subduction erosion and adakitic-like magmas: Examples from theAndean margin and implications for other arcs [abs.]: EOS, American Geo-physical 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. 241–254.

Kleiman, L., and Japas, M., 2009, The Choiyoi volcanic province at34°S–36°S (San Rafael, Mendoza, Argentina): Implications for the late Pa-leozoic evolution of the southwestern margin of Gondwana: Tectono-physics, v. 473, p. 283–299.

Kley, J., and Monaldi, C., 1998, Tectonic shortening and crustal thickness inthe Central Andes: How good is the correlation?: Geology, v. 26, p. 723–726.

Kley, J., Monaldi, C., Rossello, E., and Ege, H., 2005, The Eastern Cordilleraof the Central Andes: Inherited mechanical weakness as a first-order con-trol on the Cenozoic orogeny [ext. abs.]: International Symposium on An-dean Geodynamics, 6th, Barcelona, Extended Abstracts, p. 432−435.

Kloppenburg, A., Grocott, J., and Hutchinson, D., 2010, Structural settingand synplutonic fault kinematics of a Cordilleran Cu-Au-Mo porphyry min-eralization system, Bingham mining district, Utah: Economic Geology, v.105, p. 743–761.

Ladino, M., Tomlinson, A.J., and Blanco, N., 1997, Nuevos antecedentespara la edad de la deformación cretácica en Sierra de Moreno, II Regiónde Antofagasta-Norte de Chile [ext. abs.]: VIII Congreso Geológico Chi-leno, Antofagasta, Extended Abstracts, p. 103−107.

Lamb, S.H., 1994, Behavior of the brittle crust in wide zones of deformation:Journal of Geophysical Research, v. 99, p. 4457–4483.

——2001, Vertical axis rotation in the Bolivian orocline, South America, 2.Kinematic and dynamical implications: Journal of Geophysical Research, v.106, p. 2633−2653.

Lamb, S., and Davis, P., 2003, Cenozoic climate change as a possible causefor the rise of the Andes: Nature, v. 425, p. 792−797.

Lang, J.R., and Titley, S.R., 1998, Isotopic and geochemical characteristics ofLaramide magmatic systems in Arizona and implications for the genesis ofporphyry copper deposits: Economic Geology, v. 93, p. 138−170.

Lara, L., Lavenu, A., Cembrano, J., and Rodríguez, C., 2006, Structural con-trols of volcanism in transversal chains: Resheared faults and neotectonicsin the Cordón Caulle-Puyehue area (40.5° S), Southern Andes: Journal ofVolcanology and Geothermal Research, v. 158, p. 70−86.

Larson, R.L., 1991, Geological consequences of superplumes: Geology, v. 19,p. 963–966.

Lindsay, D., Zentilli, M., and Rojas de la Rivera, J., 1995, Evolution of an ac-tive ductile to brittle shear system controlling mineralization at theChuquicamata porphyry copper deposit, northern Chile: International Ge-ology Review, v. 37, p. 945−958.

356 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 356

Luo, G., and Liu, M., 2009a, Why short-term crustal shortening leads tomountain building in the Andes, but not in Cascadia?: Geophysical Re-search Letters, v. 36, L08301, 5 p., doi:10.1029/2009GL037347.

——2009b, How does trench coupling lead to mountain building in theSubandes?: A viscoelastoplastic finite element model: Journal of Geophys-ical Research, v. 114, B03409, 16 p., doi:10.1029/2008JB005861, 2009.

Maksaev, V., and Zentilli, M., 1988, Marco metalogénico regional de los me-gadepósitos de tipo pórfido cuprífero del Norte Grande de Chile [ext. abs.]:V Congreso Geológico Chileno, Santiago, Extended Abstracts, v. 1(B), p.181–212.

——1999, Fission track thermochronology of the Domeyko Cordillera,northern Chile: Implications for Andean tectonics and porphyry coppermetallogenesis: Exploration and Mining Geology, v. 8, p. 65−89.

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 theevolution of a supergiant porphyry Cu-Mo deposit: Society of EconomicGeologists Special Publication 11, p. 15–54.

Maksaev, V., Munizaga, F., Fanning, M., Palacios, C., and Tapia, J., 2006,SHRIMP U-Pb dating of the Antucoya porphyry copper deposit: New evi-dence for an early Cretaceous porphyry-related metallogenic epoch in theCoastal Cordillera of northern Chile: Mineralium Deposita, v. 41, p.637−644.

Maksaev, V., Munizaga, F., Mathur, R., Barra, F., Fanning, M., McWilliams,M., Ruiz, J., and Sanhueza, A., 2009, Geochronology of the Collahuasi por-phyry Cu-Mo district, north Chilean Andes [ext. abs.]: XII Congreso Ge-ológico Chilen, Santiago, Extended Abstracts (CD), 4 p.

Maksaev, V., Almonacid, T., Munizaga, F., Valencia, V., McWilliams, M., andBarra, F, 2010, Geochronological and thermochronological constraints onporphyry copper mineralization in the Domeyko alteration zone, northernChile: Andean Geology, v. 37, p. 144−176.

Mamani, M., Wörner, G., and Sempere, T., 2010, Geochemical variations inigneous rocks of the Central Andean orocline (13°S to 18°S): Tracingcrustal thickening and magma generation through time and space: Geolog-ical Society of America Bulletin, v. 122, p. 162–182.

Manea, V., Pérez-Gussinyé, M., and Manea, M., 2012, Chilean flat slab sub-duction controlled by overriding plate thickness and trench rollback: Geol-ogy, v. 40, p. 35−38.

Mann, P., 1997, Model for the formation of large transtensional basins inzones of tectonic escape: Geology, v. 25, p. 211–214.

Marinovic, N., and García, M., 1999, Hoja Pampa Unión, Región de Antofa-gasta: Santiago, Servicio Nacional de Geología y Minería, Mapas Geológi-cos, N° 9 (escala 1:100,000).

Marinovic, N., Smoje, I., Maksaev, V., Hervé, M., and Mpodozis, C., 1995,Hoja Aguas Blancas, Región de Antofagasta: Santiago, Servicio Nacional deGeología y Minería, Carta Geológica de Chile N˚ 70 (1:250,000), 150 p.

Marquardt, J.C., Rojo, J., Rivera, S., and Pizarro, J., 2009 Geología del pór-fido cuprífero Miranda: nuevo descubrimiento en el Cluster Toki, Distritode Chuquicamata, Chile [ext. abs.]: XII Congreso Geológico Chileno, San-tiago, Extended Abstracts (CD), 4 p.

Marschik, R., and Fontboté, L., 2001, The Candelaria-Punta del Cobre ironoxide Cu-Au(-Zn-Ag) deposits, Chile: Economic Geology, v. 96, p. 1799–1826.

Martinod, J., Funiciello, F., Faccenna, C., Labanieh, S., and Régard, V., 2005,Dynamical effects of subducting ridges: Insights from 3-D laboratory mod-els: Geophysical Journal International, v. 163, p. 1137–1150.

Martinod, J., Husson, L, Roperch, P., Guillaume, B., and Espurt, N., 2010,Horizontal subduction zones, convergence velocity and the building of theAndes: Earth and Planetary Science Letters, v. 299, p. 299–309.

Maydagán, L., Franchini, M., Chiaradia, M., Pons, J., Impiccini, A., Toohey,J., and Rey, R., 2011, Petrology of the Miocene igneous rocks in the Altarregion, main Cordillera of San Juan, Argentina. A geodynamic modelwithin the context of the Andean flat-slab segment and metallogenesis:Journal of South American Earth Sciences, v. 32, p. 38−40.

McInnes, B.I.A., Farley, K.A., Sillitoe, R.H., and Kohn B.P., 1999, Applica-tion of apatite (U-Th)/He thermochronometry to the determination of thesense and amount of vertical fault displacement at the Chuquicamata por-phyry copper deposit, Chile: Economic Geology, v. 94, p. 937–948.

McQuarrie, N., 2002, Initial plate geometry, shortening variations, and evo-lution of the Bolivian orocline: Geology, v. 30, p. 867–870.

——2006, Revisiting shortening estimates along the Bolivian orocline: Im-plications of thermal heating, erosion and crustal flow on the developmentof a high elevation plateau [abs.]: Geological Society of America Abstractswith Programs, Specialty Meeting 2, Mendoza, p. 86.

Mora, R, Alfaro, J., Salazar, G., and González, R., 2009, Pórfido cuprífero Mi-rador: el descubrimiento más reciente en el Distrito Centinela [ext. abs.]:XII Congreso Geológico Chileno, Santiago, Extended Abstracts (CD), 4 p.

Mpodozis, C., and Kay, S.M., 1992, Late Paleozoic to Triassic evolution of theGondwana margin: Evidence from Chilean Frontal Cordilleran batholiths:Geological Society of America Bulletin, v. 104, p. 999−1014.

Mpodozis, C., and Perelló, J., 2003, Porphyry copper metallogeny of the mid-dle-Eocene-early Oligocene arc of western South America. Relationshipswith volcanism and arc segmentation [ext. abs.]: X Congreso GeológicoChileno, Concepción, 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 SciencesSeries, v. 11, p. 59−90.

——2008, Tectónica jurásica en Argentina y Chile: extensión, subducciónoblicua, rifting, deriva y colisiones?: Revista de la Asociación Geológica Ar-gentina, v. 63, p. 481−497.

Mpodozis, C., Marinovic, N., and Smoje, I., 1993a, Eocene left lateral strikeslip faulting and clockwise block rotations in the Cordillera de Domeyko,west of the Salar de Atacama, northern Chile: International Symposium onAndean Geodynamics, 2nd, Oxford, Proceedings, p. 225–228.

Mpodozis, C., Marinovic, N., Smoje, I., and Cuitiño, L., 1993b, Estudio ge-ológico-estructural de la Cordillera de Domeyko entre Cerro Limón Verdey Sierra Mariposas, Región de Antofagasta: Servicio Nacional de Geologíay Minería [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 Atacamabasin, Antofagasta, northern Chile: Implications for the tectonic evolutionof the central Andes: Tectonophysics, v. 399, p. 125–154.

Mpodozis, C., Cembrano, J., and Mora, R., 2009a, Deformación compresiva-oblicua polifásica y pórfidos cupríferos eocenos en el Sistema de Fallas deDomeyko: la región de Esperanza-Caracoles (Distrito Centinela) [ext. abs.]:XII Congreso Geológico Chileno, Santiago, Extended Abstracts (CD), 4 p.

Mpodozis, C., Brockway, H., Marquardt, C., and, Perelló, J, 2009b, Geocro-nología U/Pb y tectónica de la región de Los Pelambres-Cerro Mercedario:Implicancias para la evolución cenozoica de los Andes del centro de Chiley Argentina [ext. abs.]: XII Congreso Geológico Chileno, Santiago, Exten-ded Abstracts (CD), 4 p.

Müller, J., Kley, J., and Jacobshagen, V., 2002, Structure and Cenozoic kine-matics of the Eastern Cordillera, southern Bolivia (21º S): Tectonics, v. 21,1037, 24 p., doi:10.1029/2001TC001340.

Münchmeyer, C., and Valenzuela, D., 2009, El yacimiento Telégrafo: un pór-fido cuprífero en etapa de exploración avanzada en el Distrito Centinela[ext. abs.]: XII Congreso Geológico Chileno, Santiago, Extended Abstracts(CD), 4 p.

Munizaga, F., Maksaev, V., Fanning, C.M., Giglio, S., Yaxley, G., and Tassi-nari, C.G., 2008, Late Paleozoic-Early Triassic magmatism on the westernmargin of Gondwana: Collahuasi area, northern Chile: Gondwana Re-search, v. 13, p. 407−427.

Muñoz, 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-cen-tral Chile to the late Oligocene increase in plate convergence rate: RevistaGeológica de Chile, v. 27, p. 177–203.

Muñoz, M., Fuentes, F., Vergara, M., Aguirre, L., Nyström, J.O., and Féraud,G., 2006, Abanico East Formation: Petrology and geochemistry of volcanicrocks behind the Cenozoic arc front in the Andean Cordillera, central Chile(33°50' S): Revista Geológica de Chile, v. 33, p. 109−140.

Nalpas, T., Hérail, G., Mpodozis, C., Riquelme, R., Clavero, J., and Dabard,M P., 2005, Thermochronological data and denudation history along a tran-sect between Chañaral and Pedernales (~26º S), North Chilean Andes:Orogenic implications [ext. abs.]: International Symposium on Andean Ge-odynamics, 6th, Barcelona, Extended Abstracts, p. 548−551.

Niemeyer, H., and Munizga, R., 2008, Structural control of the emplacementof the Portrerillos porphyry copper, central Andes of Chile: Journal ofSouth American Earth Sciences, v. 26, p. 261−270.

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 deChile: Andean Geology, v. 36, p. 37−49.

Noble, D., McKee, E., and Mégard, F., 1979, Early Tertiary ‘’Incaic’’ tecton-ism, uplift, and volcanic activity, Andes of central Peru: Geological Societyof America Bulletin, v. 90, p. 903−907.

Oldow, J.S., Bally, A.W., and Avé Lallemant, H.G., 1990, Transpression, oro-genic float and lithospheric balance: Geology, v. 18, p. 991–994.

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 357

0361-0128/98/000/000-00 $6.00 357

Oncken, O., Hindle, D., Kley, J., Elger, P., Victor, P., and Schemmann, K.,2006, Deformation of the Central Andean upper plate system: Facts, fic-tion, and constraints for plateau models, in Oncken, O., et al., eds., TheAndes: Frontiers in Earth Sciences, Pt. III, Springer-Verlag, p. 3–28.

Ossandón, G., Fréraut, R., Gustafson, L., Lindsay, D., and Zentilli, M., 2001,Geology of the Chuquicamata mine: A progress report: Economic Geology,v. 96, p. 240−270.

Oyarzún, R., Márquez, A., Lillo, J., López, I., and Rivera, S., 2001, Giant ver-sus small porphyry copper deposits of Cenozic age in northern Chile:Adakitic versus normal calc-alkaline magmatism: Mineralium Deposita, v.32, p. 794−798.

Pananont, P., Mpodozis, C., Jordan, T., Blanco, N., and Brown, L., 2004,Cenozoic evolution of the northwestern Salar de Atacama basin, northernChile: 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 centralSierras Pampeanas: An Early to Mid-Ordovician continental arc on theGondwana margin: Geological Society Special Publication 142, p. 343–367.

Pardo-Casas, F., and Molnar, P., 1987, Relative motion of the Nazca (Faral-lon) and South American plates since late Cretaceous time: Tectonics, v. 6,p. 233–248.

Perelló, J., 2003, Conchi porphyry copper deposit, Antofagasta region, nort-hern Chile [ext. abs.]: X Congreso Geológico Chileno, Concepción, Exten-ded Abstracts (CD), 1 p.

Perelló, J., Urzúa, F., Cabello, J., and Ortiz, F., 1996, Clustered gold-bearingOligocene porphyry copper and associated epithermal mineralization at LaFortuna, Vallenar region, Chile: Society of Economic Geologists SpecialPublication 5, p. 81−90.

Perelló, J., Cox, D., Garamjav, D., Sandorj, S, Diakov, S., Schissel, D.,Munkhbat, T., and Oyun, G., 2001, Oyu Tolgoi, Mongolia: Siluro-Devonianporphyry Cu-Au-(Mo) and high-sulfidation Cu mineralization with a Cre-taceous chalcocite blanket: Economic Geology, v. 96, p. 1407–1428.

Perelló, J., Carlotto, V., Zárate, A., Ramos, P., Posso, H., Neyra, C., Caballero,A., Fuster, N., and Muhr, R., 2003a, Porphyry-style alteration and mineral-ization of the middle Eocene to early Oligocene Andahuaylas-Yauri belt,Cuzco region, Peru: Economic Geology, v. 98, p. 1575−1605.

Perelló, J., Martini, R., Arcos, R., and Muhr, R., 2003b, Buey Muerto: Por-phyry copper mineralization in the Early Cretaceous arc of northern Chile[ext. abs.]: X Congreso Geológico Chileno, Concepción, Extended Ab-stracts (CD), 1 p.

Perelló, J., Brockway, H., and Martini, R., 2004, Discovery and geology ofthe Esperanza porphyry copper-gold deposit, Antofagasta region, north-ern Chile: Society of Economic Geologists Special Publication 11, p.167−186.

Perelló, J., Razique, A., Schloderer, J., and Asad, R., 2008, The Chagai por-phyry copper belt, Baluchistan Province, Pakistan: Economic Geology, v.103, p. 1583−1612.

Perelló, J., Muhr, R., Mora, R., Martínez, E., Brockway, H., Swaneck, T.,Artal, J., Mpodozis, C., Müncheyer, C., Clifford, J., Acuña, E., Valenzuela,D., and Argandoña, R., 2010, Wealth creation through exploration in a ma-ture terrain: The case history of the Centinela district, northern Chile: So-ciety of Economic Geologists Special Publication 15, p. 229−252.

Perelló, J., Sillitoe, R.H., Mpodozis, C., Brockway, H., and Posso, H., 2012,Geologic setting and evolution of the porphyry copper-molybdenum andcopper-gold deposits at Los Pelambres, central Chile: Society of EconomicGeologists Special Publication 16, p. 79–104.

Pérez, D.J., 2001, Tectonic and unroofing history of Neogene Manantialesforeland basin deposits, Cordillera Frontal (32°30' S), San Juan Province,Argentina: Journal of South American Earth Sciences, v. 14, p. 693–705.

Perry, V.D., 1952, Geology of the Chuquicamata orebody: Mining Engineer-ing, v. 4, p. 1166−1168.

Ramos, V.A., 2009, Anatomy and global context of the Andes: Main geologicfeatures and the Andean orogenic cycle: Geological Society of AmericaMemoir 204, p. 31–65.

Ramos, V.A., and Dalla Salda, L., 2011, Occidentalia: Un terreno acrecionadosobre el margen gondwánico? [ext. abs.]: XVIII Congreso Geológico Ar-gentino, Neuquén, Extended Abstracts , p. 222−223.

Ramos, V., and Folguera, A., 2005, Tectonic evolution of the Andes ofNeuquén: Constraints derived from the magmatic arc and foreland defor-mation: Geological Society Special Publication 252, p. 15−35.

——2009, Andean flat-slab subduction through time: Geological SocietySpecial Publication 327, p. 31–54.

Ramos, V., and Kay, S. M, 1991, Triassic rifting and associated basalts in theCuyo Basin, central Argentina, Geological Society of America, Specialpaper 265, p. 79–91.

Ramos, V.A., Jordan, T.E., Allmendinger, R.W., Mpodozis, C., Kay, S.M.,Cortés, J.M., and Palma, M.A., 1986, Paleozoic terranes of the Central Ar-gentine-Chilean Andes: Tectonics, v. 5, p. 855−880.

Ramos, V.A., Cegarra, M., and Cristallini, E., 1996, Cenozoic tectonics of theHigh Andes of west-central Argentina (30–36° S latitude): Tectonophysics,v. 259, p. 185–200.

Ramos, V.A., Cristallini, E., and Pérez, D., 2002, The Pampean flat slab of theCentral Andes: Journal of South American Earth Sciences, v. 15, p. 59−78.

Ramos, V.A., Zapata, T., and Cristallini, E.O., 2004, The Andean thrust sys-tem: Structural styles and orogenic shortening: American Association of Pe-troleum Geologists Memoir 82, p. 30−50.

Reich, M., Parada, M.A., Palacios, C., Dietrich, A., Schultz, F., andLehmann, B., 2003, Adakite-like signature of late Miocene intrusions at theLos Pelambres giant porphyry copper deposit in the Andes of central Chile:Metallogenic implications: Mineralium Deposita, v. 38, p. 876–885.

Reutter, K.-J., Scheuber, E., and Helmcke, D., 1991, Structural evidence oforogen-parallel strike slip displacements in the Precordillera of northernChile: Geologische Rundschau, v. 80, p. 135−153.

Reutter, K.-J., Scheuber, E., and Chong, G., 1996, The Precordilleran faultsystem of Chuquicamata, northern Chile: Evidence for reversals along arc-parallel strike-slip faults: Tectonophysics, v. 259, p. 213–228.

Richards, J.P., 2003, Tectono-magmatic precursors for porphyry Cu-(Mo-Au)deposit formation: Economic Geology, v. 98, p. 1515–1533.

——2009, Post-subduction porphyry Cu-Au and epithermal Au deposits:Products of remelting of subduction-modified lithosphere: Geology, v. 37,p. 247−250.

——2011a, Magmatic to hydrothermal metal fluxes in convergent and col-lided margins: Ore Geology Reviews, v. 40, p. 1–26.

——2011b, High Sr/Y magmas and porphyry Cu ± Mo ± Au deposits. Justadd water: Economic Geology, v. 106, p. 1075–1081.

Richards, J.P., and Kerrich, R., 2007, Adakite-like rocks: Their diverse originsand questionable role in metallogenesis: Economic Geology, v. 102, p.537–576.

Richards, J.P., Boyce, A.J., and Pringle, M.S., 2001, Geologic evolution of theEscondida area, northern Chile: A model for spatial and temporal local-ization of porphyry copper mineralization: Economic Geology, v. 96, p.271–305.

Roperch, P., Sempere, T., Macedo, O., Arriagada, C., Fornari, M., Tapia, C.,García, M., and Laj, C., 2006, Counterclockwise rotation of late Eocene–Oligocene fore-arc deposits in southern Peru and its significance for oro-clinal 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 rota-tions and transcurrent deformation south of the Abancay Deflection in theAndes of southern Peru: Tectonics, v. 30, TC2010, 23 p., doi:10.1029/2010TC002725.

Rosas, S., Fontboté, L., and Tankard, A., 2007, Tectonic evolution and pale-ogeography of the Mesozoic Pucará basin, central Peru: Journal of SouthAmerican Earth Sciences, v. 24, p. 1–24.

Russo, R., and Silver, P., 1996, Cordillera formation, mantle dynamics, andthe Wilson cycle: Geology, v. 24, p. 511–514.

Sandeman, H.A., Clark, A.H., and Farrar, E., 1995, An integrated tectono-magmatic model for the evolution of the southern Peruvian Andes (13−20°S) since 55 Ma: International Geology Review, v. 37, p. 1039–1073.

Schellart, W.P., 2008, Overriding plate shortening and extension above sub-duction zones: A parametric study to explain formation of the Andes Moun-tains: Geological Society of America Bulletin, v. 120, p. 1441–1454.

Schellart, W.P., and Rawlinson, B, 2010, Convergent plate margin dynamics:New perspectives from structural geology, geophysics and geodynamicmodeling: Tectonophysics, v. 483, p. 4–19.

Scheuber E., and González, G., 1999, Tectonics of the Jurassic-Early Creta-ceous magmatic arc of the north Chilean Coastal Cordillera (22–26° S): Astory of crustal deformation along a convergent plate boundary: Tectonics,v. 18, p. 895–910.

Scheuber, E., and Reutter, K.-J., 1992, Magmatic arc tectonics in the centralAndes between 21° and 25° S: Tectonophysics, v. 205, p. 127−140.

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 MPODOZIS AND CORNEJO

0361-0128/98/000/000-00 $6.00 358

Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe,O., Neraudeau, D., Cardenas, J., Rosas, S., and Jimenez, N., 2002, LatePermian-Middle Jurassic lithospheric thinning in Peru and Bolivia, and itsbearing on Andean-age tectonics: Tectonophysics, v. 345, p. 153–158.

Sepúlveda, F., Lahsen, A., Bonvallot, S., Cembrano, J, Alvarado, A., andLetelier, P., 2005, Morpho-structural evolution of the Cordón Caulle geo-thermal region, Southern volcanic zone, Chile: Insights from gravity and40Ar/ 39Ar dating: Journal of Volcanology and Geothermal Research, v. 148,p. 165−189.

SERNAGEOMIN, 2002, Mapa Geológico 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, Thelate Miocene to early Pliocene Rio Blanco-Los Bronces copper deposit,Central Chilean Andes: Society of Economic Geologists Special Publica-tion 9, p. 299–332.

Shafiei, B, Haschke, M., and Shahabpour, J., 2009, Recycling of orogenic arccrust triggers porphyry Cu mineralization in Kerman Cenozoic arc rocks,southeastern Iran: Mineralium Deposita, v. 44, p. 265–283.

Sibson, R.H., 1987, Earthquake rupturing as a mineralizing agent in hydro-thermal systems: Geology, v. 17, p. 701−704.

——1994, Crustal stress, faulting and fluid flow: Geological Society SpecialPublication 78, p. 69−84.

Sillitoe, R.H., 1977, Permo-Carboniferous, Upper Cretaceous and Mioceneporphyry copper-type mineralization in the Argentinian Andes: EconomicGeology, v. 72, p. 99–103.

——1998, Major regional factors favouring large size, high hypogene grade,elevated gold content and supergene oxidation and enrichment of porphyrycopper deposits, in Porter, T.M., ed., Porphyry and hydrothermal copperand gold deposits: A global perspective: Adelaide, Australian MineralFoundation, p. 21−34.

——2003, Iron oxide-copper-gold deposits: An Andean view: MineraliumDeposita, v. 38, p. 787–812.

——2010, Porphyry copper systems: Economic Geology, v. 105, p. 3–41.Sillitoe, R.H., and Perelló, J., 2005, Andean copper province: Tectonomag-

matic settings, deposit types, metallogeny, exploration, and discovery: Eco-nomic Geology 100th Anniversary Volume, p. 845−890.

Silva, C., Herrera, C., and Hervé, F., 2003, Petrogénesis de lavas y diques bá-sicos de la Formación Traiguén, (43° 30'−46° S, Chile) [ext. abs.]: X Con-greso Geológico Chileno, Concepción, Extended Abstracts (CD), 11 p.

Silver, P.G., Russo, R.M., and Lithgow-Bertelloni, C., 1998, Coupling ofSouth American and African Plate motions and plate deformation: Science,v. 279, p. 60–63.

Skewes, M.A., and Stern C.R., 1994, Tectonic trigger for the formation of lateMiocene Cu-rich breccia pipes in the Andes of central Chile: Geology, v.22, p. 551–554.

Skewes, M.A., Holmgren, C., and Stern, C.R., 2003, The Donoso copperrich, tourmaline-bearing breccia pipe in central Chile: Petrologic, fluid in-clusion and stable isotope evidence for an origin from magmatic fluids:Mineralium Deposita, v. 38, p. 2−21.

Solomon, M., 1990, Subduction, arc reversals and the origin of porphyry cop-per-gold deposits in island arcs: Geology, v. 18, p. 630−633.

Somoza, R., 1998, Updated Nazca (Farallon)-South America relative motionsduring the last 40 My: Implications for mountain building in the Central An-dean region: Journal of South American Earth Sciences, v. 11, p. 211–215.

Somoza, R., and Ghidella, M., 2005, Convergencia en el margen occidentalde América del Sur durante el Cenozoico: subducción de las placas deNazca, Farallón y Aluk: Revista de la Asociación Geológica Argentina, v. 60,p. 797−809.

Somoza, R., and Zaffarana, C.B., 2008, Mid-Cretaceous polar standstill ofSouth America, motion of the Atlantic hotspots and the birth of the AndeanCordillera: Earth and Planetary Science Letters, v. 271, p. 267–277.

Soto, R., Martinod, J., Riquelme, R., Hèrail, G., and Audin, L., 2005, Usinggemorphological markers to discriminate Neogene tectonic activity in thePrecordillera of North Chilean forearc (24–25° S): Tectonophysics, v. 411,p. 41–55.

Stern, C.R., 1991, Role of subduction erosion in the generation of Andeanmagmas: Geology, 19, p. 78−81.

——2011, Subduction erosion: Rates, mechanisms, and its role in arc mag-matism and the evolution of the continental crust and mantle: GondwanaReserach, v. 20, p. 284−338.

Stern, C.R., Skewes, M.A., and Arévalo, A., 2010, Magmatic evolution of thegiant El Teniente Cu-Mo deposit, central Chile: Journal of Petrology, v. 52,p. 1591−1617. doi: 10.1093/petrology/egq029.

Stern, C.R., Floody, R., and Espiñeira, D., 2011, Olivine-hornblende-lam-prophyre dykes from Quebrada los Sapos, El Teniente, Central Chile (34°S): Implications for the temporal geochemical evolution of the Andean sub-arc mantle: Andean Geology, v. 38, p. 1−22.

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 de-posits: International Geology Review, v. 53, p. 691–703.

Swaneck, T., Mora, R., and Martínez, E., 2009, Caracoles: un nuevo pórfidode Cu-Au-Mo en el Distrito Centinela [ext. abs.]: XII Congreso GeológicoChileno, Santiago, Extended Abstracts (CD), 4 p.

Taylor, G.K., Dashwood, B., and Grocott, J., 2005, Central Andean rotationpattern: Evidence from paleomagnetic rotations of an anomalous domain inthe forearc of northern Chile: Geology, v. 33, p. 777−780.

Thiéblemont, D., Steins, G., and Lescuyer, J.L., 1997, Gisements epither-maux et porphyriques : la connexion adakite: Comptes Rendues, Académiedes Sciences de Paris, Earth and Planetary Sciences, v. 325, p. 103−109.

Tikoff, B., Russo, R., Teyssier, Ch., and Tommasi, A., 2002, Clutch tectonicsand the partial attachment of lithospheric layers: EGU Stephan MuellerSpecial Publication Series, v. 1, p. 57−73.

Tomlinson, A.J., and Blanco, N., 1997a, Structural evolution and displace-ment history of the West fault system, Precordillera, Chile: Pt. 1. Synmin-eral history [ext. abs.]: VIII Congreso Geológico Chileno, Antofagasta, Ex-tended Abstracts, v. 3, p. 1873−1878.

——1997b, Structural evolution and displacement history of the West faultsystem, Precordillera, Chile: Pt. 2. Postmineral history [ext. abs.]: VIIICongreso Geológico Chileno, Antofagasta, Extended Abstracts, v. 3, p.1878−1882.

Tomlinson, A.J., and Cornejo, P., 2012, Regional distribution of MiddleEocene-early Oligocene porphyry copper centers in northern Chile: Sec-ond order patterns and possible causes: XIII Congreso Geológico Chileno,in press.

Tomlinson, A.J., Mpodozis C., Cornejo, P., and Ramírez, C.F., 1993, Struc-tural geology of the Sierra Castillo-Agua Amarga fault system, Precordilleraof Chile, El Salvador-Potrerillos: II International Symposium on AndeanGeodynamics, Oxford, Proceedings, p. 259−262.

Tomlinson, A.J., Blanco, N., Maksaev, V., Dilles, J.H., Grunder, A.L., and La-dino, M., 2001a, Geología de la Precordillera Andina de Quebrada Blanca-Chuquicamata, Regiones I y II (20º30'−22º30' S): Servicio Nacional de Geología y Minería, 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 andamount of vertical fault displacement at the Chuquicamata porphyry cop-per deposit, Chile—a discussion: Economic Geology, v. 96, p. 1307–1309.

Tomlinson, A.J., Blanco, N., and Dilles, J., 2010, Carta Calama, Región deAntofagasta: Servicio Nacional de Geología y Minería, Santiago, Carta Ge-ológica de Chile, Serie Preliminar, 8 (1:50,000).

Toro, J.C., Ortúzar, J., Maksaev, V., and Barra, F., 2009, Nuevos anteceden-tes geogronológicos franja de pórfidos Cu-Mo del mioceno-plioceno, Chilecentral: Implicancias metalogénicas [ext. abs.]: XII Congreso GeológicoChileno, Santiago, Extended Abstracts (CD), 4 p.

Toro, J.C., Ortúzar, J., Zamorano, J., Cuadra, P., Hermosilla, J., and Spröhnle,C., 2012, Protracted magmatic-hydrothermal history of the Río Blanco-LosBronces district, central Chile: Development of world’s greatest knownconcentration of copper: Society of Economic Geologists, Special Publica-tion 16, p. 105–126.

Tosdal, R., Dilles, J., and Cooke, D.R., 2009, From source to sinks in aurif-erous magmatic-hydrothermal porphyry and epithermal deposits: Ele-ments, v. 5, p. 289−295.

Trumbull, R., Riller, U., Oncken, O., Scheuber, E., Munier, K., and Hong, F.,2006, The time-space distribution of Cenozoic volcanism in the south-cen-tral Andes: A new data compilation and some tectonic implications, in On-cken, O., et al., eds., The Andes: Frontiers in Earth Sciences, Pt. III:Springer-Verlag, p. 29−43.

Urzúa, F., 2009, Geology, geochronology and structural evolution of La Es-condida 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 sub-duction: Journal of Geophysical Research, v. 84, p. B1049–B1061.

Valencia-Moreno, M., Ochoa-Landín, L., Noguez-Alcántara, B., Ruiz, J., andPérez-Ségur, E., 2007, Geological and metallogenetic characteristics of theporphyry copper deposits of México and their situation in the world con-text: Geological Society of America Special Paper 422, p. 438−458.

CENOZOIC TECTONICS AND PORPHYRY COPPER SYSTEMS OF THE CHILEAN ANDES 359

0361-0128/98/000/000-00 $6.00 359

Veevers, J.J., 1989, Mid Triassic (230 ± 5 Ma) singularity in the stratigraphicand magmatic history of the Pangean heat anomaly: Geology, v. 17, p.784−787.

von Huene, R., and Scholl, D.W., 1991, Observations at convergent marginsconcerning sediment subduction, subduction erosion, and the growth ofcontinental crust: Reviews of Geophysics, v. 29, p. 279–316.

Vry, V.H., Wilkinson, J.J., Seguel, J., and Millán, J., 2010, Multistage intru-sion, brecciation, and veining at El Teniente, Chile: Evolution of a nestedporphyry system: Economic Geology, v. 105, p. 119–153.

Warnaars, F.W., Holmgren, C., and Barassi, S., 1985, Porphyry copper andtourmaline breccias at Los Bronces-Río Blanco, Chile: Economic Geology,v. 80, p. 1544–1565.

Willner, A.P., Thomson, S.N., Kröner, A., Wartho, J.A., Wijbrans, J., andHervé, F., 2005, Time markers for the evolution and exhumation history ofa late Paleozoic paired metamorphic belt in central Chile (34°−35°30' S):Journal of Petrology, v. 46, p. 1835−1858.

Wilson, J., Zentilli, M., Boric, R., Díaz, J., and Maksaev, V., 2011, Geochem-istry of the Triassic and Eocene igneous host rocks of the MMH porphyrycopper deposit, Chuquicamata district, Chile: SGA Biennial Meeting, 11th,Antofagasta, Proceedings, p. 426−428.

Woodcock, N.H., 1986, The role of strike-slip fault systems at plate bound-aries: Philosophical Transactions of the Royal Society of London, Series A,v. 317, p. 13–29.

Wotzlaw, J., Decou, A., von Eynatten, H., Wörner, G., and Frei, D., 2011,Jurassic to Palaeogene tectono-magmatic evolution of northern Chile andadjacent Bolivia from detrital zircon U-Pb geochronology and heavy min-eral provenance: Terra Nova, v. 23, p. 399–406.

Yañez, G., and Cembrano, J., 2004, Role of viscous plate coupling in the LateTertiary Andean tectonics: Journal of Geophysical Research, v. 109,B02407, 21p., doi:10.1029/2003JB002494.

Yañez, G., and Maksaev, V., 1994, Sobre la distribución espacial de cuerposintrusivos asociados a diapirismo y la caracterización de la fuente magmá-tica en pórfidos cupríferos: Inestabilidades gravitacionales en un medio vis-coso [ext. abs.]: VII Congreso Geológico Chileno, Antofagasta, ExtendedAbstracts, v. 2, p. 1642−1646.

Yañez, G., Ranero, C., von Heune, R., and Díaz, J., 2001, Magnetic anomalyinterpretation across the southern Central Andes (32°−34° S): the role ofthe Juan Fernández Ridge in the Late Tertiary evolution of the margin:Journal of Geophysical Research, v. 106, p. 6235−6345.

Zengqian, H., Hongwen, M., Zaw, K., Yuquan, Z., Mingjie, W., Zeng, W.,Guitang, P., and Renli, T., 2003, The Himalayan Yulong porphyry copperbelt: Product of large-scale strike-slip faulting in eastern Tibet: EconomicGeology, v. 98, p. 125–145.

0361-0128/98/000/000-00 $6.00 360

360 MPODOZIS AND CORNEJO