Download as pdf or txt
Download as pdf or txt
You are on page 1of 28

BULLETIN OF THE SOCIETY OF ECONOMIC GEOLOGISTS

Vol. 111 January–February No. 1

The Timing of Magmatism and Ore Formation in the El Abra Porphyry


Copper Deposit, Northern Chile: Implications for Long-Lived Multiple-Event
Magmatic-Hydrothermal Porphyry Systems*
Karen J. Correa,1,† Osvaldo M. Rabbia,2 Laura B. Hernández,2 David Selby,3 and Marcelo Astengo4
1 Departamento de Ciencias de La Tierra, Casilla 160-C, Universidad de Concepción, Concepción CP 4030000, Chile
2 Instituto GEA, Casilla 160-C, Universidad de Concepción, Concepción CP 4030000, Chile
3 Department of Earth Sciences, University of Durham, Durham DH1 3LE, UK
4 Superintendencia de Geología, Sociedad Contractual Minera El Abra, Calama CP 1390000, Chile

Abstract
Zircon LA-ICP-MS and CA-TIMS and molybdenite Re-Os geochronology for the El Abra porphyry cop-
per deposit (Chile) document a ~8.6 Ma protracted magmatic-hydrothermal history. Initial magmatic events
occurred at ~45 Ma, and continued to ~36 Ma. An ~1.8 m.y. magmatic lull is recorded in the early stages of
magmatism, which is characterized by equigranular rocks from the Pajonal suite. Magmatism from ~41 to
~36 Ma, corresponding to the El Abra Granodiorite Complex, is more recurrent, ending in a ~1.4 m.y. period of
porphyritic intrusions, which are coeval with Cu-Mo mineralization. Late porphyritic units reveal subtle zircon
inheritance likely associated with early magmatic pulses. The young stages of magmatism indicate system reju-
venation based on plagioclase phenocryst compositions, suggesting that the porphyry system is ultimately linked
to a less evolved magma. Although there are multiple porphyry and veining events, molybdenite Re-Os ages
deine a focused mineralization episode from 36.34 to 36.18 Ma. Molybdenite Re-Os ages from nearby related
prospects belonging to the El Abra cluster expand the mineralization period in the district up to ~1.0 m.y., sug-
gesting a multistage, long-lived hydrothermal system. Integration of the molybdenite Re-Os dates and those
from previous works at the El Abra and Toki clusters, accompanied by available tectonic reconstruction along
the Domeyko fault system suggest the development, by late Eocene times, of a single, large (~30 km across)
porphyry copper cluster, which was subsequently separated by an offset of ~35 km along the West fault.

Introduction (34–31 Ma), among others, were formed during this metal-
The El Abra porphyry copper deposit (~8.5 Mt Cu; Camus, logenic cycle (Camus, 2003).
2003) is located in the Central Andes region of northern Chile The El Abra porphyry copper deposit is associated with
(Fig. 1). This ore deposit is part of the middle Eocene to early the El Abra granodiorite complex (Ambrus, 1977, 1979). The
Oligocene porphyry copper belt, which spans over 1,400 km northern extent of this complex is defined by a sharp contact
from 18° to 31°S (Camus, 2003). This mineralized belt was with the older Pajonal diorite (Dilles et al., 1997). Together,
structurally controlled by the transpressional Domeyko fault these form the Pajonal-El Abra igneous suite. This suite crops
system, which shows important along-strike displacements out 2 km eastward of the West fault and ~35 km north of the
and shortening (Coira et al., 1982; Mpodozis et al., 1993). The Chuquicamata porphyry system (Fig. 1). The Pajonal-El Abra
development of this belt is associated with the cessation of a suite has evolved over an ~7.5 m.y. period, from the oldest
compressive tectonomagmatic cycle, when volcanism waned Pajonal diorite (43.2 Ma) to the youngest Lagarto porphyry
and porphyry copper systems formed above large silicic (35.6 Ma) (Valente et al., 2006).
plutonic complexes (Dilles et al., 1997; Ballard et al., 2002; Surface geologic mapping within the El Abra deposit has
Haschke, 2002). World-class deposits such as Chuquica- documented up to seven texturally distinct intrusive units (Bar-
mata (35–31 Ma), La Escondida (38–37 Ma), and Collahuasi rett, 2004). Crosscutting relationships, along with K-Ar (biotite)
and U-Pb (zircon) ages for some of these units, suggest more
than one period of magmatic-hydrothermal activity (Maksaev
† Corresponding author: e-mail, rabbia@udec.cl
et al., 1990; Dilles et al., 1997; Tomlinson et al., 2001; Bal-
*A digital supplement is available at http://economicgeology.org/ and at http:// lard et al., 2002; Campbell et al., 2006). Current geochronol-
econgeol.geoscienceworld.org/. Appendices 1, 2, and 3 follow the text. ogy indicates that magmatic activity within the deposit spans
©2016 by Economic Geology, Vol. 111, pp. 1–28 Submitted: April 22, 2015
0361-0128/16/4367/1-28 1 Accepted: September 6, 2015
2
Iquique Quebrada Blanca
Collahuasi-Ujina
DFS

Cluster Toki El Abra


22° S Radomiro Tomic
Chuquicamata
Calama
Telégrafo Salar de

Pacific Ocean
Atacama basin

Argenna
Antofagasta
EA-1284 24° S
Zaldívar
La Escondida
EA-657
EA-6
Taltal
EA-916 26° S El Salvador

Potrerillos
EA-1116 Copiapo

EA-903 EXPLANATION
EA-1141 Dark diorite

CORREA ET AL.
EA-710 Central diorite
Equis quartz-monzodiorite
EA-561
Rojo Grande granite
Clara granodiorite
Apolo leucogranite and aplite
El Abra porphyry
Intrusive breccias
Hydrothermal breccias
Colluvium and gravels
El Abra pit
Potassic zone
Quartz-sericite zone
Lineaments
U-Pb Ages
Re-Os Ages

Fig. 1. Surface geology map of the El Abra mine, with sample locations. Projection: UTM19; Datum: Provisional South America 1956 (PSAD56). DFS = Domeyko fault
system. Modiied from an unpublished El Abra mine map.
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 3

over an ~3.5 m.y. period, between the older equigranular Cen- et al., 2001). The El Abra granodiorite complex is regionally
tral diorite (40.60 ± 0.30 Ma; U-Pb zircon LA-ICP-MS) and divided into Cerro Panizo de Ojuno quartz-diorite, El Abra
one dacitic variety of the younger El Abra porphyry (37.16 ± granodiorite, and Llareta granodiorite-granite (Tomlinson et
0.38 Ma; U-Pb zircon LA-ICP-MS) (Ballard et al., 2002; Camp- al., 2001). At the El Abra deposit, however, the so-called El
bell et al., 2006). Subsequently, a deep (1 km below surface) Abra granodiorites are known to be represented by the Cen-
drilling program revealed several new texturally distinct por- tral diorite, the Dark diorite and the Equis quartz-monzodi-
phyritic units (Ardila, 2009). Discovery of these new porphyries orite (Barrett, 2004). This nomenclature will be used in this
provides evidence for multiple intrusive events but also raised a work. The El Abra granodiorites along with Clara granodio-
question about the age of these units. Samples from these deep rite are the host rocks of the copper mineralization. In the
drill cores have been dated in this work. northeast part of the open pit mine the geology comprises the
The mineralization age at El Abra has been estimated from Rojo Grande granite, equivalent to the distal Llareta gran-
available K-Ar data (Ambrus, 1977; recalculated by Maksaev, ite (Tomlinson et al., 2001), which in turn has a transitional
1990; Dilles et al., 1997; Tomlinson et al., 2001) obtained on contact with the Llareta granodiorite, both located north of
hydrothermal biotite and sericite associated with potassic and the open pit mine. All above units are intruded by the Clara
quartz-sericite alterations, respectively. These data suggest a granodiorite, the largest intrusion of this complex, which
duration of ~2 m.y. (35.0–36.8 Ma) for the hydrothermal system crops out in the southern sector of the deposit and is the main
and its cooling below the K-Ar closure temperature of micas. host rock of mineralization at depth. The Clara granodiorite
Despite the evidence provided by previous field work and has been previously divided into two units, Clara and South
geochronological studies, the duration of the magmatic system granodiorite (Ambrus, 1979); however, subsequent studies
and the precise timing of mineralization remain poorly con- have found no evidence for a contact between these units
strained, which limits the understanding of ore formation in and instead group them under the name of Cotari granodio-
this porphyry copper deposit and in the metallogenic district. rite (Dilles et al., 1997). More recently, the terms “Clara” or
Here we present and discuss the precise timing of ore deposi- “South” have been used interchangeably to describe the Cotari
tion and duration of the magmatic-hydrothermal system at the granodiorite (e.g., Ballard, 2001; Tomlinson et al., 2001; Bal-
El Abra porphyry copper deposit using new U-Pb ages obtained lard et al., 2002; Campbell et al., 2006). The term Clara grano-
by zircon LA-ICP-MS and CA-TIMS 206Pb/238U analyses (on diorite is adopted here for this unit. All units mentioned so
the same samples) from the El Abra granodiorite complex, far are represented by equigranular rocks, although the Clara
and high precision Re-Os ages on molybdenite. The integra- granodiorite may locally be porphyritic. In contrast, the young-
tion of results from previous workers with those presented here est units are dominantly porphyritic. Three porphyritic intru-
extends the duration of magmatic activity to ~8.6 m.y., conirm- sions occur in the El Abra area: the aplitic granite porphyry,
ing the long-lived nature of the magmatic-hydrothermal system the main Lagarto granodiorite porphyry, and the late Lagarto
associated with this deposit and providing better constraints on maic granodiorite porphyry (Dilles et al., 1997). The oldest of
the timing of ore formation. The mineralization and emplace- these units is represented by a group of dikes and sheets cross-
ment age of porphyritic intrusions can now be correlated with cutting the Clara granodiorite that includes the aplitic granite
those from nearby prospects and other systems along the trace porphyry, the Apolo leucogranite, and the aplite-pegmatite
of the Domeyko fault system. We propose the existence, by late unit. All these porphyritic intrusions are interpreted to rep-
Eocene times, of a large cluster of porphyry copper systems, resent water-saturated, vapor-bubble–bearing, granitic melts
which were subsequently bisected and displaced ~35 km apart that were ilter-pressed from the upper portion of the adja-
along the West fault during post-Eocene times. cent, nearly crystalline Clara granodiorite (Tomlinson et al.,
2001). Leucogranites older than the Clara granodiorite have
The El Abra Granodiorite Complex also been reported (e.g., Ballard et al., 2002). The aplite-peg-
The work of many authors over decades and at different scales matite unit is mostly present toward the southwest, outside the
has led to some complexity in the nomenclature of rocks of open pit mine. The main Lagarto granodiorite porphyry and
the El Abra granodiorite complex. For clarity, this will be the late Lagarto maic granodiorite porphyry (terms used by
briely addressed and the adopted changes in the nomen- Dilles et al., 1997, and Tomlinson et al., 2001) extend from the
clature for this contribution will be indicated. The El Abra El Abra deposit to the Los Volcanes (also known as Conchi or
granodiorite complex can be considered as part of a larger Conchi Viejo) porphyry Cu-Mo prospect, located ~10 km to
unit called the Pajonal-El Abra complex. The late Eocene El the southeast. Campbell et al. (2006) and Ballard et al. (2001)
Abra Granodiorite Complex is a regional unit composed by a used the term El Abra porphyry for a quartz-monzonitic por-
set of calc-alkaline intrusions of intermediate to felsic compo- phyry at the mine site equivalent to the Lagarto porphyries of
sition (Ambrus, 1977). Subsequently Ambrus (1979) applied Dilles et al. (1997). The main Lagarto granodiorite porphyry
the term El Abra granodiorite complex to a group of granodi- is associated with potassic alteration and the late Lagarto
orite intrusions, and this term has been maintained in subse- maic granodiorite porphyry postdates the potassic alteration
quent studies (e.g., Dilles et al., 1997; Tomlinson et al., 2001; (Dilles et al., 1997). Tomlinson et al. (2001) suggest that each
among others). The El Abra granodiorite complex is largely type of the two porphyry varieties actually consists of several
emplaced along its south and east sides into lower Mesozoic porphyries emplaced at slightly different times. Additional
metavolcanic and metasedimentary rocks, along its northeast textural varieties of these porphyritic units were intercepted
side into Upper Cretaceous-lower Tertiary sedimentary rocks by deep drill holes (Ardila, 2009). Table 1 summarizes the
of the Tolar Formation, and along its northern side into the nomenclature and temporal relationships between the dif-
middle Eocene Pajonal monzodioritic complex (Tomlinson ferent rock units described here. The spatial association of
4 CORREA ET AL.

Table 1. Summary of Nomenclature and U-Pb Ages for Rock Units in the El Abra and Chuquicamata Districts

El Abra district Chuquicamata district


Pajonal-El Abra Complex Los Picos-Fortuna Complex

District nomenclature Mine nomenclature Los Picos-Fortuna Complex

Unit names Unit names Age (Ma) Unit names Age (Ma)

Pajonal diorite 42.9 ± 0.51 Los Picos monzodiorite 43.1 ± 0.61


43.21 ± 0.212 42.3 ± 0.21
41.84 ± 0.273
Cerro Panizo de Ojuno quartz Dark diorite 40.2 ± 0.51
diorite 40.83 ± 0.342
40.51 ± 0.993
Abra granodiorite Central diorite 40.6 ± 0.31
40.64 ± 0.252
40.62 ± 0.313
Equis quartz-monzodiorite 41.74 ± 0.192
38.5 ± 0.41
Llareta granodiorite and granite Rojo grande granite 39.14 ± 0.242 Antenna granodiorite 38.82 ± 0.643
39.09 ± 0.034 Fortuna Gris 39.1 ± 0.41
Clara granodiorite Clara granodiorite 38.71 ± 0.114 Fiesta granodiorite 39.38 ± 0.483
38.0 ± 0.31 Fortuna Clara 38.5 ± 0.41
37.55 ± 0.202
37.93 ± 0.423
Apolo aplites, leucogranite, Apolo aplites, leucogranite, 37.55 ± 0.262 Tetera porphyries
aplitic porphyries aplitic porphyries 38.5 ± 0.41
San Lorenzo porphyry 38.0 ± 0.41
38.23 ± 0.801
Lagarto granodiorite porphyry Dacite 37.49 ± 0.104 Opache porphyry 37.3 ± 11
porphyry 37.4 ± 0.31 37.9 ± 0.21
El Abra 36.89 ± 0.312 37.67 ± 0.623
poprhyry 37.16 ± 0.383
Lagarto maic granodiorite Rhyolite 36.07 ± 0.154
porphyry porphyry

1 Ballard (2001), LA-ICP-MS


2 Valente (2008) in Santana (2010), LA-ICP-MS
3 Campbell et al. (2006), LA-ICP-MS
4 This study, CA-TIMS

significant copper mineralization with the young porphyritic bornite ± molybdenite, with chalcopyrite > pyrite ± molyb-
units suggests a genetic link between ore formation and the denite dominant in the outer portions of the mineralization
final stages of the magmatic evolution of this large igneous (Ardila, 2009). Albitic alteration is localized in the west and
complex (Ambrus, 1977; Dilles et al., 1997; Tomlinson et al., southwest sectors of the open pit where it occurs in two main
2001; Ballard et al., 2002; Barrett, 2004). styles: (1) small breccias and dikes associated with tourmaline
in thin veinlets with an albitic halo, or (2) in a breccia matrix
Hydrothermal Alteration and Mineralization at in the deepest parts of El Abra deposit. Sulides are rare in
the El Abra Deposit albitic alteration, but when present they include pyrite and
The main alteration zones at El Abra include a potassic core, chalcopyrite. Quartz-sericite-clay alteration overprints the
locally more albitic at depth, surrounded by a quartz-sericite potassic and albitic zones and occurs in hydrothermal brec-
zone with an outer envelope of poorly developed propylitic cias and zones associated with northwest-trending structures
alteration (Fig. 1). Hydrothermal alteration is temporally and to the southeast along the main structural trend in the
and spatially associated with intrusion of the El Abra por- district. These structures mainly correspond to D-veins with
phyries (main Lagarto and maic Lagarto), dikes, and igne- variable thickness (1 mm–1 m) and with sericite, chlorite, and
ous-hydrothermal breccias (Tomlinson et al., 2001; Barrett, quartz halos. Quartz-chalcopyrite ± bornite ± pyrite ± molyb-
2004; Ardila, 2009). Potassic alteration highlights areas with denite, locally along with tourmaline and anhydrite, are the
pervasive replacement of plagioclase by K-feldspar, recrys- typical vein-illing minerals. A proximal propylitic alteration,
tallization of biotite phenocrysts, and presence of “shreddy” characterized by chlorite, epidote, and actinolite, is developed
biotite intergrown with magnetite and relict titanite. Miner- near the periphery of the deposit. K-Ar dating of biotite from
alization associated with potassic alteration is zoned. In the potassic alteration yields ages ranging from 35.0 to 35.3 Ma
central and deeper parts of the mine, bornite + chalcopyrite (Ambrus, 1977; Maksaev, 1990; Tomlinson et al., 2001), with
± magnetite associations prevail, followed by chalcopyrite > K-Ar determinations on sericite and sericitized whole-rock
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 5

providing similar ages of 36.0 ± 1.3, 35.7 ± 3.3, and 34.5 ± background counting time 20/5 s. Trace elements (Sr, Fe, K,
1.4 Ma (Tomlinson et al., 2001). and Mg) were measured on the same spot after major ele-
The NW-SE to WNW-ESE structures control the emplace- ment measurement using peak/background counting time up
ment of the mineralization at El Abra, including porphyry to 400/200 s (depending on the element concentration) and a
dikes and (A, AB) veinlets. D-type veinlets, related to quartz- beam current of 40 nA. Detection limits are 184 ppm for Sr,
sericite alteration show an extensional behavior, while the 450 ppm for Ba, 295 ppm for Fe, and 70 ppm for Mg. Ba and
postmineral structures (trending NW-SE to WNW-ESE, Mg were below detection limits in most analyses, so they were
NNW-SSE and E-W at the mine site), match the axes of prin- no longer considered. Calibration was based on the following
cipal stresses controlling the sinistral movement of the West standards: anorthite USNM 137041 (Al), fayalite USNM 85276
fault (Santana, 2010). Also, postmineral structures are reacti- (Fe), microcline USNM 143966 (Si and K) (Jarosewich et al.,
vations of D-type veinlets. 1980), forsterite (Mg), barite (Ba), and wollastonite (Ca). Data
reduction was performed using the ZAF correction method. A
Sampling and Analytical Methodology summary of results is displayed in Table 2.
Petrography and plagioclase mineral chemistry and U-Pb zircon dating
microtextures High-precision dates, i.e., CA-TIMS 206Pb/238U zircon ages,
This study examined more than 9,000 m of diamond drill core are not available for the El Abra district. Aware about the dis-
from the deep (800–2,000 m) levels of the El Abra deposit. In crepancy among different U-Pb methods (e.g., LA-ICP-MS
addition to petrography we also document the microtextures and CA-TIMS; von Quadt et al., 2011; Chiaradia et al., 2013,
of plagioclase using the electron microprobe (JEOL JXA8600 2014) and taking into account that the bulk of the geochrono-
Superprobe provided with a Bruker SDD EDS spectrometer, logical database compiled for the El Abra district corresponds
model XFLASH 5010, with 123 eV resolution at Mn Ka) at to LA-ICP-MS ages (Table 1), we decided to replicate ages
Instituto de Geología Económica Aplicada, Concepción Uni- using both U-Pb methods (on the same sample).
versity, Chile. Two varieties of porphyritic samples were ana- Two representative samples from the El Abra porphyry
lyzed, including a rhyolite porphyry (EA-1284, also dated by were selected for U-Pb zircon dating, including a dacite (EA-
U-Pb) and a dacite porphyry (EA-874). Selected phenocrysts 561) and a rhyolite (EA-1284). Both of these samples were
were analyzed along a core-rim transect choosing the most selected from deep drill core (500–1,300 m). The dacite
representative domains for each grain using high-contrast, porphyry sample came from a small body; locally, small
backscattered electron (BSE) imaging. The analytical condi- (3–15 cm) inclusions of this porphyry are present as enclaves
tions can be summarized as follows: major elements were mea- in the rhyolite porphyry, a relationship observed at different
sured using 15 kV, 20 nA, 10-µm beam diameter, and peak/ sites in the mine (Fig. 2A). Both porphyries intrude the Clara

A
Dac. P Rhy. P Dac. P
Dac. P Dac. P
Dac. P
Rojo Grande Rhy. P
granite

0 6cm 0 6cm

Clara grd.

Rhy. P

Fig. 2. A. Inclusions of dacite porphyry (Dac. P) in rhyolite porphyry (Rhy. P). B. Multiple phases of B- and D-veinlets cross-
cutting the rhyolite porphyry and Clara granodiorite (Clara grd.).
6
Table 2. Summary of Microtextures in Plagioclase Phenocrysts of Dacitic and Rhyolitic Porphyries (EA-874 and EA-1284)

Texture/unit Interpretation BSE image BSE image Compositional proile

Sieve-texture (ST)/ Partial dissolution due to


dacite porphyry reaction with more An (%) Sr and Fe (ppm)
Ca-rich melt
Core Rim
3000
50
ST 2000
30 1000
10 0
500 µm 125 µm 0 300 600 900 1200
Resorption surfaces (RS)/ Dissolution while reacting
dacite porphyry- with a more primitive
rhyolite porphyry magma
Core Rim
60 2000
RS
1500
40
1000
20 500
0 0

CORREA ET AL.
Rim 0 200 400
400 µm 100 µm
Core dissolution and patchy Dissolution while reacting An (%) Sr and Fe (ppm)
texture (PT) / dacite with a more primitive Core Rim
porphyry magma RS 3000
50
2000
PT 30 1000

10 0
0 50 100 150 200
100 µm 100 µm
approx. distance from the core µm
Fine-scale oscillatory zoning Convection driven small
(FOZ)/dacite porphyry- scale physical-chemical
rhyolite porphyry perturbation at the crystal-
melt interface
FOZ

100 µm 30 µm

Interpretation of microtextures based on Renjith (2013) and Pietranik et al. (2006)


Yellow squares show the extended area; blue, orange, and green solid lines represent An%, Fe (ppm), and Sr (ppm), respectively; green and orange dotted lines are minimum detection limits for Sr
(184 and Fe 295 ppm), respectively.
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 7

granodiorite. In addition, three equigranular samples from than about 50 µm in diameter were picked from the mineral
older units were selected, including the Rojo Grande granite separates and were mounted in an epoxy puck along with sev-
(EA-657), Clara granodiorite (EA-1141), and an amphibole- eral grains of the 337.13 ± 0.13 Ma Plešovice zircon standard
rich variety of the Clara granodiorite (EA-1116). Crosscutting (Sláma et al., 2007), together with a Temora 2 reference zir-
relationships between the Rojo Grande granite and the Clara cons (416.78 ± 0.33 Ma), and brought to a very high polish.
granodiorite are marked by Schlieren and aplitic zones sug- The surface of the mount was washed for 10 min with dilute
gesting that the Rojo Grande granite was not entirely crys- nitric acid and rinsed in ultraclean water prior to analysis. The
talline when the Clara granodiorite intruded (Tomlinson et highest quality portions of each grain, free of alteration, inclu-
al., 2001). Contacts between the Clara granodiorite and the sions, or possible inherited cores, were selected for analysis.
amphibole-rich variety are ambiguous and cannot be easily Line scans rather than spot analyses were employed in order
determined in drill core. Sample locations are shown in Fig- to minimize elemental fractionation during the analyses. A
ure 1 and their coordinates are listed in Table 3. laser power level of 38% and a 30 µm spot size were used.
About 25% of analyzed zircons were screened by cathodo- Backgrounds were measured with the laser shutter closed
luminescence (CL). CL images show textural characteristics for 10 s, followed by data collection with the laser iring for
indicating internal growth zonation (App. 1). Occasionally CL approximately 35 s. The time-integrated signals were analyzed
images reveal the presence of older cores (App. 1) and min- using GLITTER software (van Achterbergh et al., 2001; Grif-
eral inclusions identiied as apatite (App. 2). in et al., 2008), which automatically subtracts background
U-Pb zircon dating on selected samples was performed measurements, propagates all analytical errors, and calculates
using LA-ICP-MS and CA-TIMS methodologies at the Paciic isotopic ratios and ages. Corrections for mass and elemental
Centre for Isotopic and Geochemical Research (PCIGR), fractionation were made by bracketing analyses of unknown
University of British Columbia, Canada. The full U-Pb data- grains with replicate analyses of the Plešovice zircon standard.
set is provided in an Excel table online as a digital supplement A typical analytical session at the PCIGR consists of four anal-
(digital supp.). yses of the Plešovice zircon standard, followed by two analyses
of the Temora 2 zircon standards, ive analyses of unknown
LA-ICP-MS zircons, two standard analyses, ive unknown analyses, etc.,
The LA-ICP-MS zircon analyses were carried out follow- and inally two Temora 2 zircon standards and four Plešovice
ing the methodology described by Tafti et al. (2009). Instru- standard analyses. Temora 2 reference zircon was analyzed as
mentation employed for LA-ICP-MS dating of zircons at the an unknown in order to monitor the reproducibility of the age
PCIGR comprises a New Wave UP-213 laser ablation sys- determinations on a run-to-run basis (see digital supp.). Final
tem and a Thermo Finnigan™ Element 2™ single collector, interpretation and plotting of the analytical results employed
double-focusing, magnetic sector ICP-MS. All zircons greater the ISOPLOT software of Ludwig (2003).

Table 3. Summary of LA-ICP-MS (weighted mean age), CA-TIMS (youngest grain age), and Coordinates for Each Dated Sample

UTM coordinates1 LA-ICP-MS CA-TIMS

Unit Sample North East Elevation 206Pb/238U Ma 206Pb/238U Ma n

Rojo Grande granite EA-657 7576599.78 516599.79 3965.6 38.75 ± 0.61 39.09 ± 0.03 5
MSWD = 0.15 MSWD = 1.4
n = 20, 4 of 20 rej. n=4
102.07 ± 0.151
Clara granodiorite EA-1141 7576199.97 516900.15 3804.93 37.82 ± 0.57 38.71 ± 0.11 4
MSWD = 0.30
n = 19, 1 of 19 rej.
Mesocratic Clara granodiorite EA-1116 7576300.22 517100.22 3805.22 38.83 ± 0.47 37.45 ± 0.09 5
MSWD = 0.39
n = 19, 1 of 19 rej.
El Abra dacite porphyry EA-561 7576000.15 517399.44 3804.64 38.94 ± 0.65 37.49 ± 0.10 5
MSWD = 0.59
n = 18, 3 of 18 rej.
177.6 ± 3.72
El Abra rhyolite porphyry EA-1284 7576748.52 517848.51 3774.39 38.48 ± 0.62 36.07 ± 0.15 5
MSWD = 1.7
n = 19, 2 of 19 rej.
236.91 ± 2.12 202.35 ± 0.301
126.95 ± 0.271

Data for old xenocrystic core ages are included for each method; coordinates: projection-UTM 19K; Provisional South American Datum 1956; rej = rejected
1 Collar coordinates
2 Old zircon xenocrysts
8 CORREA ET AL.

Chemical abrasion-thermal ion mass spectrometry the Excel-based program of Schmitz and Schoene (2007).
(CA-TIMS) Standard concordia diagrams were constructed and regres-
CA-TIMS procedures described here are modiied from sion intercepts, as well as weighted averages were calculated
Mundil et al. (2004); Mattinson (2005), and Scoates and with Isoplot (Ludwig, 2003). Unless otherwise noted all errors
Friedman (2008). After rock samples have undergone stan- are quoted at the 2σ or 95% level of conidence. Isotopic dates
dard mineral separation procedures, zircons are handpicked are calculated with the decay constants λ238U = 1.55125E-10
in alcohol. The clearest, crack- and inclusion-free grains are and λ235U = 9.8485E–10 (Jaffey et al., 1971). Three EARTH-
selected, photographed and then annealed in quartz glass TIME U-Pb (ET100 Ma) synthetic solutions were analyzed
crucibles at 900°C for 60 h. Annealed grains are transferred to monitor the accuracy of CA-TIMS results. The averages of
into 3.5 mL PFA screwtop beakers, ultrapure HF (up to those three runs are as follows:
50% strength, 500 µL) and HNO3 (up to 14 N, 50 µL) are 206Pb/238U± random (+tracer) [+λ]
added and caps are hand tightened. The beakers are placed = 100.144 ± 0.07 (0.12) [0.16] Ma; MSWD = 0.06
in 125 mL PTFE liners (up to four per liner) and about 2 mL
207Pb/235U ± random (+tracer) [+λ]
HF and 0.2 mL HNO3 of the same strength as acid within
beakers containing samples are added to the liners. The liners = 100.19 ± 0.16 (0.19) [0.23] Ma; MSWD = 0.24
are then slid into stainless steel Parr™ high pressure dissolu- 207Pb/206Pb ± random [+λ]
tion devices, which are sealed and brought up to a maximum = 100.4 ± 3.0 [3.99] Ma; MSWD = 0.23
of 200°C for 8 to 16 h (typically, 175°C for 12 h). Beakers
are removed from liners and zircon is separated from leach-
ate. Zircons are rinsed with >18 MΩ.cm water and subboiled Re-Os in molybdenite
acetone. Then 2 mL of subboiled 6N HCl is added and bea- Samples for Re-Os dating in molybdenite were chosen tak-
kers are set on a hotplate at 80° to 130°C for 30 min and again ing into account the type of veinlet and their distribution
rinsed with water and acetone. Masses are estimated from within the deposit (following a SW-NE transect, see Fig. 1;
the dimensions (volumes) of grains. Single grains are trans- for coordinates, see Table 4). Analyzed molybdenites come
ferred into clean 300 µL PFA microcapsules (crucibles), and from B- and D-type veinlets hosted by El Abra porphyry
50 µL 50% HF and 5 µL 14 N HNO3 are added. Each zir- units. Sample EA-916 (dacite porphyry) corresponds to a 3- to
con is spiked with a 233-235U-205Pb tracer solution (EARTH- 7-mm-thick D-veinlet, mainly illed by quartz with molybde-
TIME ET535), capped and again placed in a Parr liner (8–15 nite bands toward the edges. It presents an up to 6-mm-wide
microcapsules per liner). HF and nitric acids in a 10:1 ratio, quartz-sericite alteration halos, which locally follows previous
respectively, are added to the liner, which is then placed in thin AB-veinlets with chalcopyrite and quartz-potassium feld-
the Parr high pressure device and dissolution is achieved at spar ine halos. Grain size of molybdenite varies from <1 to 4
240°C after 40 h. The resulting solutions are dried on a hot- mm. Samples EA-903 (dacite porphyry) and EA-710 (rhyolite
plate at 130°C, 50 µL 6N HCl is added to microcapsules and porphyry) correspond to a 5- to 10-mm-thick B-type vein-
luorides are dissolved in high pressure Parr devices for 12 h lets, consisting of quartz and molybdenite (grain size <1–3
at 210°C. HCl solutions are transferred into clean 7 µL PFA mm) present as aggregates and/or bands preferentially at
beakers and dried with 2 µL of 0.5 N H3PO4. Samples are the edge of the veinlet, with disseminated chalcopyrite (Fig.
loaded onto degassed, zone-reined Re ilaments in 2 µL of 2B). Analyses were performed in the Laboratory for Sulide
silicic acid emitter (Gerstenberger and Haase, 1997). Isotopic and Source Rock Geochronology and Geochemistry in the
ratios are measured in a modiied single collector VG-54R or Durham Geochemistry Center at the University of Durham
354S (with Sector 54 electronics) TIMS equipped with ana- (UK). The complete analytical procedure for Re-Os determi-
logue Daly photomultipliers. Analytical blanks are 0.2 pg for nations is described in Selby and Creaser (2004) and Selby et
U and 1.0 pg for Pb. Uranium fractionation was determined al. (2007), and it is only briely summarized here. Molybde-
directly on individual runs using the EARTHTIME ET535 nite samples were dissolved and equilibrated with a known
mixed 233-235U-205Pb isotope tracer and Pb isotope ratios were amount of 185Re and isotopically normal Os in inverse aqua
corrected by fractionation by 0.25%/amu, based on replicate regia (2:1 16 N HNO3 and 12 N HCl, 3 mL) at 240°C for 24 h
analyses of the NBS-982 reference material and the values in a Carius tube. Rhenium and Os were isolated and puriied
recommended by Thirlwall (2000). Data reduction employed by solvent extraction, microdistillation, and anion exchange

Table 4. Sample Coordinates and Results of Re-Os Molybdenite Dating (2σ level)

UTM Coordinates

Sample Vein/ unit North East wt (g) Re (ppm) ± 187Re (ppm) ± 187Os (ppb) ± Age (Ma) ± ±1

EA-916 D/EADPP 7576460 517959 0.010 508.2 2.5 319.4 1.6 193.4 0.9 36.34 0.15 0.18
EA-903 B/EADP 7576200 517400 0.012 242.8 1.1 152.6 0.7 92.0 0.4 36.18 0.15 0.18
EA-710 B/EARP 7576100 516700 0.012 300.7 1.4 189.0 0.9 114.1 0.5 36.22 0.15 0.18

Coordinates: Projection-UTM 19K; Provisional South American Datum 1956


EADP = El Abra dacite porphyry; EARP = El Abra rhyolite porphyry
1 Includes decay constant
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 9

chromatography, and analyzed by negative TIMS on a Fisher constant (Smoliar et al., 1996; Selby et al., 2007) is also con-
Scientiic TRITON mass spectrometer using Faraday col- sidered (Table 4).
lectors. Total procedural blanks for Re and Os, are 2 pg and
0.5 pg, respectively, with an 187Os/188Os blank composition Results and Discussion: Magmatism
of 0.24 ± 0.02 (n = 6). Rhenium and Os concentrations and The discrepancy among the CA-TIMS and LA-ICP-MS
Re-Os molybdenite date uncertainties are presented at the results (Fig. 3A, B) is discussed below case by case. We can
2σ level, which includes the uncertainties in Re and Os mass anticipate, however, that in three cases (samples EA-1116,
spectrometer measurement, spike and standard Re and Os EA-561, and EA-1284), there is clearly a demonstration of
isotope compositions, and calibration uncertainties of 185Re the fact that LA-ICP-MS ages are “artiicially” systematically
and 187Os. Because a mixed 185Re and Os tracer solution is older than the youngest CA-TIMS age as shown already by
used, uncertainties in weights of sample and tracer solution von Quadt et al. (2011) and Chiaradia et al. (2013, 2014).
do not affect the calculated age, and are not considered. How-
ever, sample and tracer solution weight uncertainties are con- Rojo Grande granite (EA-657)
sidered in determining the uncertainty in the Re and 187Os The oldest analyzed rock corresponds to the Rojo Grande equi-
concentrations. Uncertainty with and without the 187Re decay granular granite. This is a leucocratic, phaneritic rock (Fig. 4A,

A. CA-TIMS B. LA-ICP-MS

Fig. 3. Ages of intrusive rocks from the El Abra deposit. A. CA-TIMS 206Pb/238U dates. Best estimated ages (youngest grain),
or mean weighted age for equivalent zircon population, are indicated for each unit. B. LA-ICP-MS 206Pb/238U dates with
probability plots and density functions. Light gray histograms are from the density plot and dark gray columns are 206Pb/238U
mean weighted ages. Rej. = rejected data points (black bars).
10 CORREA ET AL.

A B

1cm
0 3cm

C D

1cm
0 3cm

E F

1cm
0 3cm

G H

1cm
0 4cm

I J

1cm
0 3cm

Quartz Amphibole(Hornblende)
Feldspars Magnetite
K-Feldspars Titanite
Biotite/Phlogopite Ilmenite
Fig. 4. False color QEMSCAN and hand-sample photos of dated samples. A, B. EA-657, Rojo grande granite. C, D. EA-1141,
Clara granodiorite. E, F. EA-561, dacite porphyry. G, H. EA-1116, mesocratic Clara granodiorite. I, J. EA-1284, rhyolite
porphyry.
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 11

B). The major constituents are subhedral plagioclase (34.56 data points rejected (black bars, Fig. 3B). The small MSWD
vol %), commonly zoned and slightly altered to sericite, anhe- value may suggest the analytical errors are likely to be over-
dral K-feldspar (28.65 vol %) with perthitic texture, and anhe- estimated. On the other hand, if rejected data are included
dral quartz crystals (27.50 vol %). Additionally, minor biotite (n = 20), the weighted mean becomes 39.40 ± 0.63 Ma with
(6 vol %), mostly secondary, is present in association with a still acceptable MSWD of 1.2. However, in the cumulative
chlorite and magnetite. Accessory phases include rhombic probability plot, the four removed zircon data form a distinc-
titanite, magnetite (locally replaced by hematite with ilmen- tive, slightly older (~40.5–42.5 Ma) subpopulation (see digital
ite exsolutions), apatite, and zircon. Zircons are euhedral, supp.) and therefore we prefer to reject these data, despite
typically smaller than 50 µm, and are common inclusions in this decreases the MSWD value below 1. The CA-TIMS
plagioclase. This sample has thin D-veinlets of quartz-chlo- methods (Fig. 5A, B shows concordia diagrams, respectively)
rite-chalcopyrite with quartz-sericite halos up to 3 mm thick. yields a weighted mean 206Pb/238U date of 39.088 ± 0.026 Ma
The cumulative probability plot method (e.g., Campbell et (MSWD = 1.4; Fig. 6). Therefore, the LA-ICP-MS age of
al., 2006; see also Chiaradia et al., 2014) was used for removal Rojo Grande granite is overlapping within the large uncer-
of spot analyses representing subtle inheritance and Pb-loss, a tainty of the LA-ICP-MS method (38.75 ± 0.61 or 39.40 ±
common type of open-system behavior in zircons, from each 0.63 Ma) to the CA-TIMS 206Pb/238U ages. These dates are
LA-IPC-MS data population (digital supp.). The LA-ICP-MS in perfect agreement with previously reported ages for this
data have a nearly unimodal distribution (Fig. 3B), yielding unit (Table 1). The Rojo Grande granite unit shows evidence
a 206Pb/238U weighted mean age of 38.75 ± 0.61 Ma (mean of a subtle inheritance (slightly older subpopulation in the
square of weighted deviates, MSWD = 0.15) with 4 out of 20 cumulative probability plot, digital supp.) from proto-plutons

A B

C D

Fig. 5. Concordia diagrams, left side: LA-ICP-MS, right side: CA-TIMS.


12 CORREA ET AL.

E F

G H

I J

Fig. 5. (Cont.)
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 13

Fig. 6. Weighted mean 206Pb/238U CA-TIMS dates of Rojo Grande granite (EA-657).

or slightly early magmatic products. An old inherited zircon discordant 206Pb/238U CA-TIMS date (45.54 ± 0.06 Ma) inter-
core from this sample yields an Early Cretaceous age of 102 ± preted to represent an inherited zircon grain. The remain-
0.15 Ma (CA-TIMS). ing three CA-TIMS dates are not statistically equivalent and
range from 38.91 ± 0.06 to 38.71 ± 0.11 Ma (Figs. 3A, 5D)
Clara granodiorite (EA-1141) (weighted mean CA-TIMS age 38.87 ± 0.24  Ma, MSWD =
The Clara granodiorite is a phaneritic, leucocratic, equigran- 4.8). The difference between the youngest CA-TIMS date
ular rock that is macroscopically characterized by irregular (38.71 ± 0.11 Ma, Fig. 3A), which is the closest approximation
aggregates of biotite (2–3 mm), biotite-replaced hornblende to the intrusion emplacement age (Chiaradia et al., 2013) and
(euhedral, up to 6 mm) and prismatic plagioclase surrounded the LA-ICP-MS weighted mean age (37.82 ± 0.57 Ma) for the
by K-feldspar (Fig. 4C, D). Plagioclase (54.25 vol %) shows Clara granodiorite may result either from undetected Pb loss
moderate sericitization, patchy texture, oscillatory zoning, (see the complex cumulative probability plot, digital supp.) in
and locally myrmekitic texture with anhedral quartz inclu- LA-ICP-MS ages or from undersampling in CA-TIMS analy-
sions. Anhedral quartz (18.02 vol %) crystallized late, illing ses, or a combination of both. Our best estimated CA-TIMS
interstices. Potassium feldspar (14.76 vol %) is anhedral with age for the Clara granodiorite intrusion (38.71 ± 0.11 Ma) is
perthitic texture. Biotite (5%) and amphibole (3%) are minor older than previously reported ages by conventional U-Pb and
constituents partially or totally replaced by secondary biotite LA-ICP-MS dating, which range from 37.5 ± 0.5 to 38.0 ±
and chlorite. Titanite, magnetite, apatite, and zircon are com- 0.3 Ma (Table 1; Fig. 7; Dilles et al., 1997; Ballard et al., 2001;
mon accessory minerals. Titanite occurs as subhedral crystals Campbell et al., 2006). However, considering that sampling
partially altered to rutile. Magnetite partially oxidized to hema- sites (and depths) were different in each case, and taking into
tite is associated with biotite and titanite. Apatite appears as account the batholithic size of the Clara granodiorite (Dilles
tiny prismatic crystals usually included in plagioclase. Zircon et al., 1997), it is possible that the observed age range in this
tends to be euhedral and appears included in plagioclase and unit relects a long-lived magma system, rather than only dif-
biotite, as well as associated with magnetite aggregates. Crys- ferences in dating methods.
tal sizes vary between 10 and 50 µm. This sample contains
early quartz veinlets (1–8 mm) with a discontinuous suture of Mesocratic Clara granodiorite (EA-1116)
bornite and K-feldspar and biotite toward the vein margins. The amphibole-rich, mesocratic Clara granodiorite (Fig. 4E,
Despite the fact that LA-ICP-MS data from this sample have F) represents a variety of coarse-grained equigranular rock
a unimodal distribution (Fig. 3B), the cumulative probability intercepted at depths of over 1 km from the present pit bot-
plot shows a complex distribution of data (digital supp.). One tom level. It has a particularly distinctive texture visually domi-
out of 19 zircons analyzed was rejected. It was interpreted to nated by large (up to 20 mm) euhedral, subequant hornblende
be a slightly older inherited grain (~40.5 Ma) using the cumu- crystals. It appears as up to 0.5 m wide meso- to melanocratic
lative probability plot (digital supp.). The remaining 18 zircons bands in the Clara granodiorite without a clearly deined
yield a weighted mean 206Pb/238U LA-ICP-MS age of 37.82 ± intrusive contact. Large amphibole crystals (up to 20  mm)
0.57 Ma (MSWD = 0.30). The same applies to the oldest and have previously been described in the Clara granodiorite
14 CORREA ET AL.

Age references: Geochemical data


references:
This work U-Pb (Zircon/CA-TIMS youngest grain) Ballard et al., 2001, U-Pb (Zircon/LA-ICP-MS) This work
This work Re-Os (Molybdenite) Ballard et al., 2001, U-Pb (Zircon/SHRIMP) Ballard et al., 2001
This work U-Pb (Zircon/LA-ICP-MS) Campbell et al., 2006, U-Pb (Zircon/LA-ICP-MS) Cabrera, 2011
This work, subtle inherited grains (LA-ICP-MS) Dilles et al., 1997, U-Pb (Zircon/Conventional) Rojas, 2015
Tomlinson et al., 2001, K-Ar (Biotite) Barra et al., 2013, U-Pb (Zircon/LA-ICP-MS)
Rivera and Pardo, 2004, U-Pb (Zircon/SHRIMP) Barra et al., 2013, Re-Os (Molybdenite)
Proffet, 2008 in Cabrera, 2011, U-Pb (Zircon/SHRIMP) Valente, 2008 in Santana, 2010, U-Pb (Zircon/LA-ICP-MS)
Perelló et al., 2012, U-Pb (Zircon/LA-ICP-MS) Perelló et al., 2012, Re-Os (Molybdenite)
Rojas, 2015, U-Pb (Zircon/LA-ICP-MS)

Fig. 7. Compilation of dates in a timeline for the El Abra and Chuquicamata districts, by different methods and authors.
Errors in U-Pb ages are reported at the 2σ level. Uncertainties in Re-Os ages are reported with a 2σ uncertainty. D = diorite,
Grd = granodiorite, mzD = monzodiorite, P = porphyry, Gr = granite, Mesoc. = mesocratic, RT = Radomiro Tomic, PFF =
ine porphyry (minor facies) (Campbell et al., 2006), PFE = Este porphyry (major facies) (Campbell et al., 2006).
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 15

(Dilles et al., 1997), but with a lower modal abundance than agglomerates of secondary biotite and epidote. Subrounded
the melanocratic bands described in this study, which have quartz eye phenocrysts (<1 vol %) are small (0.5–2 mm) and
variable amphibole contents, commonly greater than 10 wt some show embayments. Magnetite is a minor constituent
%. Biotite-altered diorite xenoliths of uncertain origin (Dark (2 vol %) along with apatite, titanite, and zircon. Zircons are
or Central) are locally hosted by these melanocratic bands in euhedral to subhedral and frequently included in plagioclase
the Clara granodiorite. Sample EA-1116 consists of euhedral and hornblende phenocrysts, but are also locally found as part
prismatic crystals of amphibole (11.24 vol %), 2 to 16  mm of the groundmass. Their average size is about 36 µm. Chal-
in size, partially replaced by secondary biotite and chlorite, cocite, bornite, and digenite are disseminated throughout the
subhedral plagioclase (34.65 vol %), often surrounded by groundmass; locally, a bornite-chalcocite “myrmekite”-type
anhedral K-feldspar (13.51 vol %), and quartz (17.93 vol %). texture is observed. Quartz-rich A- and AB-type veinlets with
Secondary biotite aggregates (11.45 vol %) predominate over biotite-bornite-chalcopyrite-K-feldspar halos are present in
primary biotite. Magnetite (2 vol %), relict titanite (± rutile), this sample.
apatite, and zircon are present as accessory phases. Large The LA-ICP-MS method yields a weighted mean 206Pb/238U
(20–100 µm) euhedral zircons occur as inclusions in horn- age of 38.94 ± 0.65 Ma (MSWD = 0.59), from 15 analyzed zir-
blende, and less frequently in plagioclase microphenocrysts, cons. Three data were rejected using the cumulative probabil-
associated with quartz-magnetite aggregates. Some zircon ity plot (digital supp.), two of these being slightly older (44.70
crystals display thin Th-, U-, and Hf-enriched growth bands ± 0.53 and 42.70 ± 0.53 Ma) zircons interpreted as inherited
(if viewed in cross section) as part of the compositional zon- grains. The older rejected zircon age is concordant while the
ing and small (5–20 µm) Cl-F-OH apatite inclusions are often youngest one is slightly discordant, possibly due to Pb loss
present (see App. 2). Also present are quartz-rich A-type and (Fig. 5G). If both data are considered the weighted mean age
AB-type veinlets with K-feldspar, abundant anhydrite, born- becomes 41.1±1.3 with MSWD = 6.2, a value that indicates
ite, and chalcopyrite. scatter well above the expected analytical uncertainty. Addi-
The 206Pb/238U LA-ICP-MS weighted mean age is 38.83 ± tionally, both data plot outside the weighted mean age range
0.47 Ma (MSWD = 0.39). The LA-ICP-MS data have an almost (black bar, Fig. 3B). The third rejected result is from an old
unimodal distribution, with only one rejected discordant zir- zircon core that yielded a Jurassic age of 177.6 ± 3.1 Ma (digi-
con (45.70 ± 1.71 Ma) out of 19 (Fig. 5E). The removed zircon tal supp.). Sample EA-561 yields CA-TIMS 206Pb/238U ages
is interpreted as an inherited grain (digital supp.). Concordia ranging from 38.39 ± 0.07 to 37.49 ± 0.10 Ma. Despite the
diagrams are displayed in Figure 5E and F. The ive 206Pb/238U relatively small age difference (0.9 m.y.) between the oldest
CA-TIMS dates are not statistically equivalent and range from and youngest zircon grain, the ive ages are not statistically
38.47 ± 0.08 to 37.45 ± 0.09 Ma (weighted mean age is 38.01 equivalent (weighted mean age is 38.19 ± 0.32 Ma, MSWD =
± 0.49, MSWD = 101). The LA-ICP-MS weighted mean is out 57) (Figs. 3A, 5H). This is a typical case showing that LA-ICP-
of the CA-TIMS ages range, even though most CA-TIMS data MS weighted mean dates are often older than the youngest
are in the range of the LA-ICP-MS data population (Fig. 3A, CA-TIMS age (von Quadt et al., 2011; Chiaradia et al., 2013,
B). The youngest CA-TIMS date (37.45 ± 0.09 Ma) is the best 2014). Multiple populations in LA-ICP-MS dates are difi-
approximation of the age of intrusion. This CA-TIMS age is cult to interpret, and a consequence is that the LA-ICP-MS
among the youngest reported ages for the Clara granodiorite weighted mean ages, in some cases, may not be geologically
(Table 1, Fig. 7) and the full range of CA-TIMS ages (~1 Ma) meaningful (von Quadt et al., 2011; Chiaradia et al., 2013,
covers almost the entire range of reported ages for this unit 2014; Chelle-Michou et al., 2014). Assuming that chemical
(Dilles et al., 1997; Ballard et al., 2001; Campbell et al., 2006; abrasion performed in the high-precision CA-TIMS method
Valente, 2008, in Santana, 2010). allows complete removal of the effect of postcrystallization Pb
loss (damaged zircon domains) (Mattinson, 2005), the young-
Dacite porphyry (EA-561) est zircon grains might best approximate the emplacement
The selected dacite porphyry (sample EA-561) comes from age (maximum crystallization age) resolving the 206Pb/238U
a dike that intrudes the Clara granodiorite. It is a mesocratic, LA-ICP-MS data ambiguities. Therefore, the best estimated
porphyritic rock (Fig. 4G, H) with 50 vol % phenocrysts age for the dacite porphyry is 37.49 ± 0.10 Ma (youngest
(1–4  mm) in a ine-grained groundmass (0.08–0.15  mm) grain, CA-TIMS). This age is identical (within uncertainty)
composed of plagioclase, quartz, and K-feldspar. Plagioclase to 206Pb/238U CA-TIMS age of the equigranular, mesocratic
phenocrysts (48.54 vol %) are subhedral to euhedral with Clara granodiorite (37.45 ± 0.09 Ma, sample 1116), but older
K-feldspar rims. Plagioclase phenocryst cores have com- than the CA-TIMS age for the rhyolite porphyry (Table 3,
positions that range from An36 to An29, and also contain Sr Fig. 3A). This proves that at least the early generations of por-
(1,140–1,800 ppm) and Fe (1,100–1,700 ppm). Patchy tex- phyry dikes, such as the dacite units (sample EA-561), were
ture is observed, with and without dissolution zones and with essentially coeval with some of the coarse-grained, equigranu-
ine-scale oscillatory zoning. Sieve resorption surfaces are fol- lar intrusive rocks seated directly below (sample EA-1116),
lowed by zones with variable, but commonly high increments which may be part of a larger magmatic reservoir at depth.
in anorthite content (An53-An31) and accompanied by incre-
ments of ~100 ppm of Sr and up to ~500 ppm of Fe. Toward Rhyolite porphyry (EA-1284)
the rims, anorthite content decreases down to An21-An7, with This sample comes from a nearly vertical dike intruding
Sr below 1,200 ppm and Fe below 900 ppm (Table 2). Primary the Clara granodiorite. It is composed of abundant quartz
biotite phenocrysts (1.3 vol %) are small (0.5 mm) and appear eyes, plagioclase, and minor biotite phenocrysts in a micro-
as subhedral to euhedral single crystals, locally replaced by felsic groundmass. It is a leucocratic, porphyritic to locally
16 CORREA ET AL.

glomerophyric rock (Fig. 4I, J) dominated by subhedral, Subtle inheritance


concentrically zoned plagioclase phenocrysts (46.86 vol %). The presence of (concordant) zircons slightly older than the
Microtexture analysis (Table 2) shows large (1,400 µm) phe- crystallization age of the intrusion being dated has been called
nocryst cores with anorthite compositions ranging from An30 subtle inheritance (Campbell et al., 2006). Recycling of these
to An21, Sr contents from 1,100 to 1,700 ppm, and Fe con- zircon antecrysts during successive magmatic injections is the
tents from 700 to 1,380 ppm. Some grains exhibit multiple primary cause of the modest age dispersion of concordant
resorption surfaces, with anorthite contents ranging from zircon ages, and is compatible with progressive growth of a
An42 to An35, Sr contents from 1,800–2,000 ppm, and Fe large, long-lived, crystal mush body (Miller et al., 2007). The
contents from 900–1,500  ppm. These surfaces are followed inclusion of these inherited grains in the calculation of the
by ine-scale oscillatory zoning with compositions similar to
LA-ICP-MS weighted mean age may produce high values of
the core. Medium-sized (900–1,000  µm) phenocrysts pres-
MSWD (>2), suggesting that the spread of ages is not within
ent rounded cores and wavy, ine-scale oscillation zoning with
analytical uncertainty and thus it could have geologic mean-
anorthite content ranging between An35 and An23, Sr between
ing (Campbell et al., 2006; Chiaradia et al., 2014). Despite
1,699 and 1,193 ppm, and Fe between 1,381 and 880 ppm.
the statistical processing to remove inherited grains and
Toward the rims, anorthite contents range from An12 to An19,
grains with Pb loss, the ages determined by lower precision
Sr <800 ppm, and Fe <700 ppm. Other phenocrysts include
(1–3%) LA-ICP-MS tend to be older than those obtainable by
subhedral K-feldspar (11.04 vol %) with perthitic texture,
measuring single zircon grains by higher precision (≤0.1%)
anhedral subrounded quartz crystals (8.36 vol %) with embay-
ments, and tabular biotite (1 vol %). Phenocrysts sizes range CA-TIMS, implying that inferring an intrusion emplacement
between 300 and 800  µm, while groundmass grain size is age by in situ methods may be erroneous (von Quadt et al.,
restricted to 30 to 120 µm. The groundmass is composed of 2011; Chiaradia et al., 2014). In this context, the CA-TIMS
quartz (12.96 vol %) and K-feldspar (11.35 vol %), minor pla- age of the youngest single zircon grain is considered the one
gioclase (4.58 vol %) and scarce, ine-grained biotite altered that most closely approximates the real timing of intrusion
to chlorite (2 vol %). Magnetite, titanite, apatite, and zircon emplacement (Chiaradia et al., 2013, 2014). Additionally, a
occur as accessory phases. Zircons are less abundant than in number of recent studies carried out on intermediate and fel-
other unit samples, are subhedral to euhedral, and are found sic magmatic intrusions show that CA-TIMS age of individual
both in the groundmass and as inclusions in plagioclase phe- zircon crystals from the same hand-sized sample are not over-
nocrysts. The average size is 35 µm, similar to those in the lapping in age, which cannot be attributed to Pb loss (all zir-
dacite porphyry, but smaller than the zircons of equigranular con treated with chemical abrasion). It is rather interpreted
rocks. Biotite, chlorite, and epidote are present as alteration to be the consequence of zircon grains recording a protracted
minerals. The mineralization in this sample is associated with evolution of magmas within the crust (von Quadt et al., 2011;
thin A- and AB-type veinlets with chalcopyrite and K-feldspar Chiaradia et al., 2013, 2014 and references therein).
halos. In this work, an important spread in 206Pb/238U LA-ICP-MS
The LA-ICP-MS weighted mean 206Pb/238U age for this and CA-TIMS dates (beyond analytical uncertainty) is docu-
sample is 38.48 ± 0.62 Ma (MSWD = 1.7) (Fig. 5I). For the mented (Fig. 3A, B). The systematic mixing of old inherited
weighted mean 206Pb/238U age calculation (Fig. 3B), two con- zircon cores with younger Eocene overgrowths cannot account
cordant zircons (Fig. 5 I, J) with slightly older ages (42.90 ± for the range of 206Pb/238U dates (see App. 1, 3). The time
0.62 and 51.50 ± 1.18 Ma) were rejected (black bars, Fig. 3B, elapsed since emplacement until the magmatic body is cooled
digital supp.). The weighted mean age considering these two to the solidus can neither reproduce the observed zircon age
zircons becomes 39.1 ± 1.3 Ma (MSWD = 8.7). The removed variability. For example, if the amphibole-rich variety of the
zircons are interpreted as inherited grains. Additionally, a Clara granodiorite (sample EA-1116) is considered, the 1.02 ±
Middle Triassic (236.9 ± 2.1 Ma) inherited zircon core was 0.12 Ma (2σ) CA-TIMS age spread is incompatible with the geo-
also removed (digital supp.). Five grains were dated by CA- logical evidence. This sample represents a thin band (~0.5 m)
TIMS, two of which are older inherited zircons (126.95 ± with gradual contacts toward a more typical Clara granodiorite.
0.27  Ma, Early Cretaceous, and 202.35 ± 0.30 Ma, Juras- Such a range (e.g., 1.02 ± 0.12 Ma in the amphibole-rich variety
sic). The remaining three 206Pb/238U CA-TIMS dates are not of the Clara granodiorite, sample EA-1116) most likely derives
statistically equivalent (weighted mean age is 37.1 ± 1.9 Ma, from zircon recycling from older units—i.e., already consoli-
MSWD = 359), and show a wide age dispersion (~2.5 m.y.) dated parts of the large Clara granodiorite, and perhaps also
ranging from 38.58 ± 0.13 to 36.07 ± 0.15 Ma (Figs. 3A, 5J). partial assimilation of the Equis quartz-monzodiorite (Fig. 7).
Considering the removal of zircon damage zones by the CA- We speculate that this “cannibalization” of early-formed
TIMS method in the same way as for the dacite porphyry, rocks by later magmatic activity (subtle inheritance of zircons)
the youngest zircon (36.07 ± 0.15 Ma, CA-TIMS) is the best was probably associated with building of a magmatic reservoir
estimated age for the rhyolite porphyry intrusion. Again this from which porphyritic dike intrusions associated with miner-
is a demonstration of the fact that weighted mean LA-ICPMS alization were fed (e.g., Chiaradia et al., 2013, 2014). This may
ages are often systematically older than the youngest CA- have resulted from the intrusion of a hotter, slightly less evolved
TIMS ages. The 206Pb/238U CA-TIMS best estimated intru- magma, as evidenced by the coeval dacite porphyry event, with
sion ages of both porphyry samples (37.49 ± 0.10 and 36.07 SiO2 contents lower than the typical Clara granodiorite (Fig.
± 0.15 Ma) are younger than the best estimated 206Pb/238U 7). Additional evidence of a thermal input during this canni-
CA-TIMS age for the Clara granodiorite (38.71 ± 0.11 Ma), a balization stage can be found in microtextures preserved in
result consistent with ield observations. plagioclase phenocrysts from porphyry dike samples, including
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 17

the presence of plagioclase core dissolution textures, patchy age of 45.5 Ma (whole-rock K/Ar; Tomlinson et al., 2001).
textures, and sieve textures as summarized in Table 2. These Similarly, ages of about 45 Ma have also been reported for
microtexture features and compositional variation of plagio- the Los Picos quartz monzodiorite, located at present ~35 km
clase (Table 2) are interpreted as evidence of interaction with a southward from El Abra and westward from the West fault,
hot melt (e.g., Pietranik et al., 2006; Renjith, 2014). The large in the Chuquicamata district (Fig. 8). The Los Picos unit is
CA-TIMS age spread in the El Abra dacite (892 ± 122 ka, 2σ) broadly correlative with the Pajonal diorites and was spatially
and rhyolite (2.50 ± 0.20 Ma, 2σ) porphyry dikes, along with associated with the El Abra deposit in pre-Oligocene times
their small sizes (typically narrower than ~80 m) supports the (e.g., Dilles et al., 1997; see below).
idea of zircon recycling as a result of fresh magma input provid- Finally, magmatism at the El Abra deposit has evolved
ing additional heat for assimilation of earlier intrusives. within (and assimilated) Mesozoic crustal rocks at depth,
All rejected 206Pb/238U LA-ICP-MS and CA-TIMS con- as indicated by xenocrystic zircons with ages ranging from
cordant data in the range of ~42 to 45 Ma may represent a 102.07 ± 0.15 to 236.9 ± 2.1 Ma (Table 3).
mixing of autocrystic and antecrystic (slightly older) zircons
(e.g., Miller et al., 2007) (subtle inheritance) evidencing a Results and Discussion: Mineralization
protracted evolution of the El Abra magmas.
Age of ore formation: Re-Os molybdenite geochronology
Chronological history of the magmatic activity The concentrations of Re and Os in molybdenite and calcu-
We have documented protracted magmatic activity at the El lated ages are summarized in Table 4. The total Re and 187Os
Abra deposit. It extends back until ~45 Ma, as evidenced by concentrations range from 242.8 to 508.2 ppm and from 92 to
a concordant zircon grain of this age (44.70 ± 0.53 Ma; LA- 193.4 ppb, respectively. The highest concentrations of Re and
ICP-MS, Fig. 7) from the dacite porphyry (sample EA-561), Os are reported in sample EA-916, located in the northeast
despite the fact that there is no surface evidence for such part of the mine (Fig. 1), corresponding to a D-type veinlet
early magmatic activity. After a ~1.8 m.y. magmatic lull, igne- within the dacite variety of the El Abra porphyry. The age is
ous activity resumed at about 43 Ma with the emplacement of 36.34 ± 0.18 Ma (2σ uncertainty). Samples EA-903 (36.18 ±
Pajonal diorite (42.9 ± 0.5 to 43.21 ± 0.21 Ma; Table 1). The 0.18 Ma) and EA-710 (36.22 ± 0.18 Ma) (belonging to B-type
zircon record of magmatic activity for the next 2 Ma, from veinlets), hosted by the dacite and rhyolite porphyry respec-
about 43 and 41 Ma, is scarce for surface outcrops (Fig. 7), tively, have lower Re abundances and are within the 2σ uncer-
but is documented by slightly older (antecrystic) zircons in tainty of the age of sample EA-916.
Rojo Grande granite, dacite, and rhyolite porphyries (Fig. Three Re-Os dates were obtained from molybdenite sam-
7). The next phase of igneous activity is indicated by ~41 to ples that spread over a proile of ~1.3 km (Fig. 1), and from
36 Ma zircon U-Pb dates (Fig. 7). During this new magmatic different veinlet (B and D) and rock types (rhyolite and dacite
cycle the El Abra granodiorite complex was emplaced, start- porphyries). Nevertheless, the Re-Os data deine a tight clus-
ing at 40.83 ± 0.34 Ma with the earliest phases of Central and ter, between 36.34 ± 0.18 and 36.18 ± 0.18 Ma. These high-
Dark diorites, and ending with the rhyolite dikes at 36.07 ± precision Re-Os dates are in agreement with previous works,
0.15 Ma (Fig. 7). With few exceptions, the early magmatic which (based on K-Ar data of potassic alteration) suggest a
period, from ~45 until ~37.5 Ma, was dominated by equi- cooling window for the hydrothermal luids from about 36.8
granular intrusive rocks with relatively low La/Yb and Sr/Y to 35.0 Ma (Ambrus, 1977; recalculated by Maksaev, 1990;
ratios (Ballard et al., 2001; Fig. 7). In contrast, the late and Dilles et al., 1997; Tomlinson et al., 2001). According to these
more focused igneous activity, from ~37.5 until ~36.1 Ma new data, the ore formation age (mean: 36.25 ± 0.25 Ma, 2σ)
(~1.4 m.y.), was exclusively porphyritic at present exposure is identical within uncertainty, to the age of the latest mag-
levels, with a geochemical signature characterized by high La/ matic activity occurred in the deposit site represented by the
Yb and Sr/Y ratios (Ballard et al., 2001; Fig. 7). At the open rhyolite porphyry (36.07 ± 0.15 Ma).
pit mine, an 8.63 ± 0.55-m.y., long-lived magmatic system is The geologic evidence at the El Abra deposit suggests
thus recognized. multiple events of porphyry intrusion lasting for ~1.4 m.y.
At the district scale (Fig. 8), porphyry-style magmatism is and multiple pulses of veining (Fig. 2B), but the statistically
almost coeval with the porphyritic magmatism at El Abra. It equivalent Re-Os dates constrain the mineralizing pulses to
was dated by the LA-ICP-MS method in at least three pros- a focused episodes over a 0.16 ± 0.25 m.y. period, consider-
pects deining a linear array, which includes Brujulina (37.5 ing the upper and lower bounds of uncertainty of the Re-Os
± 0.4 and 38.3 ± 0.9 Ma), Cerro Las Papas (also known as ages. Even if these Re-Os results could be the outcome of
Pinorca) (37.2 ± 0.5 to 34.3 ± 0.4 Ma), and Los Volcanes (36.8 undersampling, it does not downplay the signiicance of the
± 0.5 to 35.2 ± 0.5 Ma) (Perelló et al., 2012; Rojas, 2015). widespread ore-forming event at 36.25 ± 0.25 Ma (2σ).
Therefore, at the district scale, igneous activity extended until At the Los Volcanes porphyry Cu-Mo prospect, in the El
34.3 ± 0.5 Ma, making the magmatic history even longer, up Abra district (Fig. 9), Re-Os geochronology indicates that
to 10.40 ± 0.73 m.y. mineralization was coeval with that of the El Abra deposit,
The early magmatic period, between ~45 and 43.2 Ma is despite being located ~10 km away from it. In this pros-
about 2 m.y. older than previously recognized at the El Abra pect, the Re-Os dates deine a series of mineralization pulses
deposit, and may represent the earliest phase of the Pajonal spanning 1.02 ± 0.21 m.y. (2σ), from 36.47 ± 0.15 to 35.45 ±
complex. Another alternative is a group of rocks without a for- 0.14 Ma (Perelló et al., 2012; Rojas, 2015).
mational name (quartz monzodiorite to diorite), which crop Together, ield observations and the U-Pb and Re-Os geo-
out ~5 km northeast of El Abra and are reported to have an chronology data reveal a cluster of magmatic-hydrothermal
18 CORREA ET AL.

E 535.000
E 525.000
E 505.000

E 515.000
N 7.580.000

El Abra Cluster

Brujulina El Abra Los Volcanes


(Conchi)

Cerro Las Papas


(Pinorca) N 7.570.000

Eocene location of
Toki Cluster
N 7.560.000 EOCENE COMPLEXES
Pajonal-El Abra Complex

Pajonal diorite
(42.9±0.5 - 43.21±0.21 Ma)

El Abra granodiorites:
-Central diorite (40.62±0.31 Ma)
-Equis qtz-monzonite
N 7.550.000 (41.74±0.19 - 38.5±0.4 Ma)

Llareta granodiorite
(40.10± 0.48 Ma)

Clara granodiorite
Radomiro Tomic (38.74±0.10 - 37.55±0.20 Ma)

Los Picos-Fortuna Complex


N 7.540.000 Los Picos monzodiorite complex
(41.5-43.7 Ma)

Fortuna granodioritic complex


Chuquicamata -Tetera aplitic porphyries
-San Lorenzo granodiorite and
Tonalite porphyries (38.2 ± 0.8 Ma)
-Fiesta granodiorite (36.9 - 39.9 Ma)
-Antena granodiorite (39.3±0.4 Ma)
Mina Sur N 7.530.000
Chuquicamata Porphyry Complex
East porphyry (35.2±0.4 Ma)
West porphyry (34±0.3 Ma)
Ministro Hales Banco porphyry (34.1±0.3 Ma)
Toki Cluster
Other units
Q T Opache prospect
O
G
M N 7.520.000
G Genoveva prospect
Q Quetena prospect
O
Calama M Miranda prospect
T Toki prospect

Fig. 8. Location map for the principal deposits and prospects of the El Abra and Chuquicamata districts, showing the Toki
cluster position at 37 to 39 Ma following the 35-km left lateral offset of the West fault (after Dilles et al., 1997). Projection:
UTM19; Datum: Provisional South America 1956 (PSAD56). Modiied from Barra et al. (2013) and Barrett (2004).
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 19

Fig. 9. Location and ages of prospects in the El Abra cluster. Ages in black = U-Pb ages of porphyry units, Ages in blue =
Re-Os in Mo ages. Modiied from Tomlinson et al. (2001).

centers in the El Abra district (Fig. 9), pointing to the pres- Opache (37.6 ± 0.2 to 36.4 ± 0.2 Ma), the youngest porphyry
ence of a large parental reservoir to which these shallow-level copper deposits from the Toki cluster (Rivera and Pardo,
porphyry systems are linked, which makes the El Abra district 2004), which is located about 50 km southward of the El Abra
attractive for further ore exploration at the district scale. deposit, in the Chuquicamata area (Fig. 8). In this cluster, the
nominally older deposits of Quetena, Toki, and Genoveva have
Regional correlation along the Domeyko fault system Re-Os ages at around 38 Ma (Barra et al., 2013). Together,
The hypothesis of post-Oligocene sinistral displacement the ore formation ages of the ive porphyry Cu-Mo deposits
along the West fault, a structure related to the Domeyko display a ~2.4 m.y. spread, from 38.4 ± 0.2 to 36.0 ± 0.2 Ma
fault system, has been widely studied and conirmed by the (Barra et al., 2013). Inferred resources for the Toki cluster are
petrographic and geochronological correlation between the estimated at about 20 Mt of Cu (Camus, 2003; Rivera et al.,
Los Picos-Fortuna and the Pajonal-El Abra intrusive com- 2006).
plexes (Reutter et al., 1996; Dilles et al., 1997; Tomlinson et When tectonic reconstruction of the ~35-km offset along
al., 2001; Ballard, 2001; Ballard et al., 2002; Campbell et al., the Domeyko fault system is considered, then the location
2006; among others). of the Toki cluster at 37 to 39 Ma is ~15 km south from the
The Re-Os geochronological data obtained in this study and present position of the El Abra region (Fig. 8). This scenario
those reported from Los Volcanes (Perelló et al., 2012) are suggests that during the late Eocene both clusters might have
similar to the Re-Os mineralization ages reported by Barra formed a single, large (~30 km across) porphyry copper cen-
et al. (2013) for Miranda (36.5 ± 0.2 to 36.0 ± 0.2 Ma) and ter associated with the youngest porphyritic facies from the
20 CORREA ET AL.

Pajonal-El Abra-Los Picos-Fortuna batholith. Subsequently, Linares E., ed., Magmatic evolution of the Andes, Earth Science Reviews,
the “El Abra-Toki” porphyry copper cluster (and batholith) Special Issue 18, p. 303–332.
Dilles, J.H., Tomlinson, A., Marti, M.W., and Blanco, N., 1997, El Abra
containing resources estimated at ~28.5 Mt of Cu (Camus, and Fortuna Complexes: A porphyry copper batholith sinistrally dis-
2003; Rivera et al., 2006) was tectonically bisected and the placed by the Falla Oeste: Actas VIII, Congreso Geológico Chileno, v. 3,
individual parts were displaced to their present positions (Fig. p. 1883–1887.
8) by post-Eocene left-lateral movements along the same Gerstenberger, H., and Haase, G., 1997, A highly effective emitter substance
for mass spectrometric Pb isotopic ratio determinations: Chemical Geology,
strike-slip structural system that controlled their formation, as v. 136, p. 309–312.
depicted by earlier tectonic models (e.g., Dilles et al., 1997, Grifin, W.L., Powell, W.J., Pearson, N.J., and O’Reilly, S.Y., 2008, Glitter:
and references therein). Data reduction software for laser ablation ICP-MS, in Sylvester, P.J., ed.,
Laser ablation ICP-MS in the earth sciences: Current practices and out-
Acknowledgments standing issues: Vancouver, B.C., Mineralogical Association of Canada
Short Course Series, Short Course 40, p. 308–311.
This work is part of a PhD study of the irst author, at the Uni- Haschke, M.R., 2002, Evolutionary geochemical patterns of Late Cretaceous
versidad de Concepción. We thank the Sociedad Contractual to Eocene arc magmatic rocks in north Chile: Implications for Archean
Minera El Abra (SCMEA) and Freeport-McMoRan Inc., for crustal growth: European Geosciences Union, EGU Stephan Mueller Spe-
funding this research, for the logistical support during ield cial Publication Series, v. 2, p. 207–218.
work of the irst author, and inally for the permission to pub- Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bentley, W.C., and Essling, A.M.,
1971, Precision measurement of half-lives and speciic activities of 235U and
lish the results. We also thank Ricardo Ardila for his coop- 238U: Physical Review Letters, v. 4, p. 1889–1906.
eration during ield work activities and Richard Friedman for Jarosewich, E., Nelen, J.A., and Norbery, J.A., 1980, Reference samples for
the LA-ICP-MS and CA-TIMS zircon analyses. Stephanie electron microprobe analysis: Geostandards Newsletter, no. 4, p. 43–47.
Mrozek and Massimo Chiaradia are thanked for thorough Ludwig, K.R., 2003, Isoplot/Ex. Version 3: A geochronological toolkit for
Microsoft Excel, Berkeley Geochronology Center, California.
reviews and constructive comments that helped to improve Maksaev, V., 1990, Metallogeny, geological evolution and thermocronology of
this paper. the Chilean Andes between latitudes 21° and 26° south, and the origin of
major porphyry copper deposits: Unpublished PhD thesis, Halifax, Canada,
REFERENCES Dalhousie University, 544 p.
Ambrus, J., 1977, Geology of the El Abra porphyry copper deposit, Chile: Mattinson, J.M., 2005, Zircon U-Pb chemical abrasion (“CA-TIMS”) method:
Economic Geology, v. 72, p. 1062–1085. Combined annealing and multi-step partial dissolution analysis for improved
——1979, Emplazamiento y mineralización de los póridos cupríferos de precision and accuracy of zircon ages: Chemical Geology, v. 220, p. 47–66.
Chile: Unpublished Ph.D. thesis, España, Universidad de Salamanca, Miller, J.S., Matzel, J.E.P., Miller, C.F., Burgess, S.D., and Miller, R.B.,
314 p. 2007, Zircon growth and recycling during the assembly of large, compos-
Ardila, R., 2009, Eventos de mineralización de sulfuros en la mina El Abra: ite arc plutons: Journal of Volcanology and Geothermal Research, v. 167,
Internal report from Sociedad Contractual Minera El Abra (SCMEA), 54 p. p. 282−299.
Ballard, J.R., 2001, A comparative study between the geochemistry of ore- Mpodozis, C., Marinovic, N., Smoje, I., and Coutiño, L., 1993, Estudio
bearing and barren calc-alkaline intrusions: Unpublished PhD thesis, Can- geológico estructural de la cordillera de Domeyko entre Sierra Limón
berra, Australia, The Australian National University, 254 p. Verde y Sierra Mariposas, región de Antofagasta: Servicio Nacional de
Ballard, J.R., Palin, J.M., Williams, I.S., Campbell, I.H., and Faunes, A., Geología y Minería-Corporación Nacional del Cobre (Chile), Informe reg-
2001, Two ages of porphyry intrusion resolved for the super-giant Chuqui- istrado IR-93-04, 282 p.
camata copper deposit of northern Chile by ELA-ICP-MS and SHRIMP: Mundil, R., Ludwig, K.R., Metcalfe, I., and Renne, P.R., 2004, Age and tim-
Geology, v. 29, p. 383–386. ing of the Permian mass extinctions: U/Pb dating of closed-system zircons:
Ballard, J., Palin, M., and Campbell, I., 2002, Relative oxidation states of Science, v. 305, p. 1760–1763.
magmas inferred from Ce(IV)/Ce(III) in zircon: application to porphyry Perelló, J., Saldías, G., Gonzales, M., Hernríquez, R., and Giglio, S., 2012,
copper deposits of northern Chile: Contributions to Mineralogy and Petrol- Hypogene alteration and mineralization at the Conchi porphyry Cu-Mo
ogy, v. 144, p. 347–364. deposit, northern Chile: assemblages, zoning, and age: XVI Peruvian Geo-
Barra, F., Alcota, H., Rivera, S., Valencia, V., Munizaga, F., and Maksaev, logic Congress, Lima, 2 p.
V., 2013, Timing and formation of porphyry Cu-Mo mineralization in the Pietranik, A., Koepke, J., and Puziewicz, J., 2006, Crystallization and resorp-
Chuquicamata district, northern Chile: New constraints from the Toki clus- tion in plutonic plagioclase: Implications on the evolution of granodiorite
ter: Mineralium Deposita, v. 48, p. 629–651. magma (Gȩsiniec granodiorite, Strzelin Crystalline Massif, SW Poland):
Barrett, L.F., 2004, Geology and status of exploration projects, El Abra dis- Lithos, v. 86, p. 260–280.
trict, II Region, Chile, El Abra: Sociedad Contractual Minera El Abra Renjith, M.L., 2014, Micro-textures in plagioclase from 1994–1995 eruption,
(SCMEA), Unpublished report, p. 6–37. Barren Island Volcano: Evidence of dynamic magma plumbing system in
Campbell, I.H., Ballard, J.R., Palin, J.M., Allen, C., and Faunes, A., 2006, the Andaman subduction zone: Geoscience Frontiers, v. 5, p. 113–126.
U-Pb zircon geochronology of granitic rocks from the Chuquicamata-El Reutter, K.J., Scheuber, E., and Chong, G., 1996, The Precordilleran fault
Abra porphyry copper belt of northern Chile: Excimer laser ablation ICP- system of Chuquicamata, northern Chile: evidence for reversals along arc-
MS analysis: Economic Geology, v. 101, p. 1327–1344. parallel strike-slip faults: Tectonophysics, v. 259, p. 213–228.
Camus, F., 2003, Geología de los sistemas porfíricos en los Andes de Chile: Rivera, S.L., and Pardo, R., 2004, Discovery and geology of the Toki por-
Corporación Nacional del Cobre de Chile-Servicio de Geología y Minería phyry copper deposit, Chuquicamata district, northern Chile: Society of
de Chile, 267 p. Economic Geologists, Special Publication 11, p. 15–54.
Chelle-Michou, C., Chiaradia, M., Ovtcharova, M., Ulianov, A., and Wotzlaw, Rivera, S.l., Pardo, R., Alcota, H., Fontecilla, C., Kovacic, P., Pizarro, J., and
J.F., 2014, Zircon petrochronology reveals the temporal link between por- Rojo, J., 2006, Reseña de la exploración y geología del cluster de póridos de
phyry systems and the magmatic evolution of their hidden plutonic roots cobre Toki, distrito Chuquicamata: Actas XI Congreso Geológico Chileno,
(the Eocene Coroccohuayco deposit, Peru): Lithos, v. 198–199, p. 129–140. Antofagasta (electronic version), p. 343–346.
Chiaradia, M., Schaltegger, U., Spikings, R., Wotzlaw, J.F., and Ovtcharova, Rojas, G., 2015, Petrografía, geoquímica y geocronología del distrito Los Vol-
M., 2013, How accurately can we date the duration of magmatic-hydrother- canes (Conchi), Región de Antofagasta, Chile: Unpublished thesis, Chile,
mal events in porphyry systems?: Economic Geology, v. 108, p. 565–584. Universidad de Concepción, 139 p.
Chiaradia, M., Schaltegger, U., and Spikings, R.A., 2014, Time scales of min- Santana, C.A., 2010, Informe estudio geológico-estructural y su aplicación
eral systems—advances in understanding over the past decade: Society of a la exploración de cuerpos mineralizados en el distrito El Abra: Internal
Economic Geologists, Special Publication 18, p. 37–58. report for El Abra Company (SCMEA), 25 p.
Coira, B., Davidson, J., Mpodozis, C., and Ramos, V., 1982, Tectonic and Schmitz, M.D., and Schoene, B., 2007, Derivation of isotope ratios, errors, and
magmatic evolution of the Andes of northern Argentina and Chile, in error correlations for U-Pb geochronology using 205Pb-235U-(233U)-spiked
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 21

isotope dilution thermal ionization mass spectrometric data: Geochemistry, Tafti, R., Mortensen, J.K., Lang, J.R., Rebagliati, M., and Oliver, J.L., 2009,
Geophysics, Geosystems, v. 8, p. 1–20. Jurassic U-Pb and Re-Os ages for newly discovered Xietongmen Cu-Au
Scoates, J.S., and Friedman, R.M., 2008, Precise age of the platiniferous porphyry district, Tibet: Implications for metallogenic epochs in the south-
Merensky reef, Bushveld Complex, South Africa, by the U-Pb ID-TIMS ern Gangdese belt: Economic Geology, v. 104, p. 127–136.
chemical abrasion ID-TIMS technique: Economic Geology, v. 103, Thirlwall, M.F., 2000, Inter-laboratory and other errors in Pb isotope analyses
p. 465–471. investigated using a 207Pb-204Pb double spike: Chemical Geology, v. 163,
Selby, D., and Creaser, R.A., 2004, Macroscale NTIMS and microscale p. 299–322.
LA-MC-ICP-MS Re-Os isotopic analysis of molybdenite: Testing spatial Tomlinson, A.J., Blanco, N., Maksaev, V., Dilles, J.H., Grunder, A.L., and
restrictions for reliable Re-Os age determinations, and implications for the Ladino, M., 2001, Geología de la precordillera Andina de Quebrada
decoupling of Re and Os within molybdenite: Geochimica et Cosmochi- Blanca- Chuquicamata, Regiones I y II (20°30'-22°30' S): Servicio Nacional
mica Acta, v. 68, p. 3897–3908. de Geología y Minería (Chile), Informe Registrado IR-01-20, 444 p.
Selby, D., Creaser, R.A., Stein, H. J., Markey, R. J., and Hannah, J. L., 2007. Valente, D.L., Campbell, I.H., and Allen, C.M., 2006, The geology, geochem-
Assessment of the 187Re decay constant accuracy and precision: Cross istry and geochronology of the El Abra mine, Chile and the adjacent El
calibration of the 187Re-187Os molybdenite and U-Pb zircon chronometers: Abra-Pajonal suite [abs.]: Geochimica et Cosmochimica Acta, v. 70, p. A664.
Geochimica et Cosmochimica Acta, v. 71, 1999–2013. van Achterbergh, E., Ryan, C.G., Jackson, S.E., and Grifin, W.L., 2001,
Sláma, J., Košler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M., Data reduction software for LA-ICP-MS: Appendix; in Sylvester, P.J., ed.,
Horstwood, M.S.A., Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, Laser ablation–ICP-mass spectrometry in the earth sciences: Principles
U., Xchoene, B., Tubrett, M.N., and Whitehouse, M.J., 2007, Plešovice and applications: Mineralogical Association of Canada Short Course Series,
zircon—a new natural reference material for U–Pb and Hf isotopic micro- Ottawa, Ontario, Canada, v. 29, p. 239–243.
analysis: Chemical Geology, v. 249, p. 1–35. von Quadt, A., Emi, M., Martinek, K., Moll, M., Peycheva, and Heinrich,
Smoliar, M.I., Walker, R.J., and Morgan, J.W., 1996, Re-Os isotope con- C.A., 2011, Zircons crystallization and the lifetimes of ore-forming mag-
straints on the age of Group IIA, IIIA, IVA, and IVB iron meteorites: Sci- matic-hydrothermal systems: Geology, v. 39, p. 731–734.
ence, v. 271, p. 1099–1102.
22 CORREA ET AL.

APPENDIX 1
CL Images of Dated Zircons from Different Rock Types

1 2 3 6
8 9

4 5
15 16 17
11 12

13 18 19
14

20 21 22

23
24

App. 1, Fig. 1. Examples of CL images from dated zircons: 1-5. Rojo Grande granite (EA-657); 6-9. Dacite porphyry
(EA-561); 11-14. Mesocratic Clara granodiorite (EA-1116); 15-19. Rhyolite porphyry (EA-1284); 20-24. Clara granodiorite
(EA-1141).
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 23

APPENDIX 2
Apatite Inclusions in Zircon

A B

C D

A B
D
App. 2, Fig. 1. Images of apatite inclusions in zircons. A. BSE. B. SEI. C. BSE. D. Ca Kα X-ray map.

App. 2, Table 1. Apatite Composition (EDS analysis spot 74


App. 2, Fig. 1C, above)

Spectrum: 74
unn. C norm. C Atom. C Error (1σ)
El AN Series [wt %] [wt %] [at.%] [w %]

O 8 K-series 17.76 23.78 41.62 2.57


F 9 K-series 1.29 1.72 2.54 0.33
P 15 K-series 13.34 17.86 16.14 0.54
Cl 17 K-series 0.99 1.32 1.05 0.06
Ca 20 K-series 41.33 55.32 38.65 1.26

Total: 74.71 100.00 100.00


24 CORREA ET AL.

A B

App. 2, Fig. 2. Images of apatite inclusions and trace element-rich zones in zircons. A. BSE. B. SEI. C. BSE.

App. 2, Table 2. Apatite Composition (EDS analysis spots


App. 2, Fig. 2B, above)

Norm. mass percent (%)

Spectrum O F P Cl Ca

50 23.88 1.05 22.21 2.06 50.79


51 24.28 2.13 19.10 0.69 53.80
52 24.08 0.85 19.09 1.92 54.06

Mean value: 24.08 1.34 20.14 1.56 52.88


Sigma: 0.20 0.69 1.80 0.75 1.82
Sigma mean: 0.11 0.40 1.04 0.43 1.05
MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 25

cps/eV
2.2 Hf Th O Hf Si Zr
2.0
U
1.8

1.6

1.4

1.2
U
1.0 Th
0.8

0.6

0.4
Hf
0.2
Th U
0.0
2 4 6 8 10 12 14
keV

App. 2, Fig. 3. EDS spectra for the trace element-enriched zone in zircon (spot 80 in image App. 2, Fig. 2C.)

App. 2, Table 3. Compositional Analysis Th, U, Hf


(enriched zone spot 80 in image App. 2, Fig. 2C)

unn. C norm. C Atom. C Error (1σ)


El AN Series [wt %] [wt %] [at.%] [w %]

O 8 K-series 16.22 15.99 41.95 2.00


Si 14 K-series 20.24 19.96 29.83 0.87
Zr 40 L-series 60.08 59.23 27.26 2.26
Hf 72 L-series 1.68 1.65 0.39 0.10
Th 90 M-series 1.60 1.58 0.29 0.08
U 92 M-series 1.61 1.59 0.28 0.08

Total: 101.43 100.00 100.00


26 CORREA ET AL.

A B

App. 2, Fig. 4. Images of apatite inclusions in zircons. A. SEI. B. SEI.

App. 2, Table 4. Apatite Composition (EDS analysis spot 63 in


App. 2, Fig. 4B, above)

Spectrum: 63

unn. C norm. C Atom. C Error (1σ)


El AN Series [wt %] [wt %] [at.%] [w %]

O 8 K-series 19.14 24.21 42.31 2.76


F 9 K-series 1.20 1.52 2.23 0.32
P 15 K-series 13.57 17.17 15.50 0.54
Cl 17 K-series 1.04 1.32 1.04 0.07
Ca 20 K-series 44.09 55.79 38.92 1.35

Totals: 79.04 100.00 100.00


MAGMATISM AND ORE FORMATION IN THE EL ABRA PORPHYRY Cu DEPOSIT, N. CHILE 27

APPENDIX 3
Th/U Ratio Versus 206Pb/238U CA-TIMS Age Diagrams

To check if the spread in 206Pb/238U CA-TIMS dates observed the mixing line (App. 3, Fig. 1). If the older early Jurassic
in most samples (EA-561, EA-1141, EA-1116, and EA-1284) age of 177.6 ± 3.7 Ma (206Pb/238U LA-ICP-MS) is considered
may result from mixing different proportions of old inherited instead, the outcome does not change (App. 3, Fig. 2). Simi-
cores and Eocene overgrowths, an evaluation using Th and U lar conclusions can be drawn for samples EA-1141, EA-1116,
contents in zircons and age information is done. If the presence and EA-1284 (black dots in App. 3, Figs. 1, 2). Therefore, the
of small xenocrystic cores were responsible for the spread in presence of small inherited zircon cores cannot adequately
206Pb/238U CA-TIMS dates, data in a Th/U ratio vs. 206Pb/238U explain the disparity of 206Pb/238U CA-TIMS dates observed
CA-TIMS age diagram should be plotted along a mixing line to a sample scale (Fig. 4A).
between the old core and the younger overgrowth. Using
the middle Cretaceous age of 102.07 ± 0.15 Ma (206Pb/238U REFERENCE
CA-TIMS) from the Rojo Grande (sample EA-567) as the Belousova, E.A., Grifin W.L., O’Reilly S.Y., and Fisher, N.I., 2002, Igneous
zircon: Trace element composition as an indicator of source rock type: Con-
old core, the result shows that the 206Pb/238U CA-TIMS dates tributions to Mineralogy and Petrology, v. 143, p. 602–622.
from the dacite porphyry (EA-561) zircons do not plot along

App. 3, Fig. 1. Th/U vs. 206Pb/238U CA-TIMS diagram showing the theoretical mixing line between the youngest zircon from
the dacite porphyry and the inherited Cretaceous zircon from the Rojo Grande granite. Black dots represent data from other
samples (Clara granodiorite, melanocratic band in Clara granodiorite, and rhyolite porphyry).
28 CORREA ET AL.

0,9
Youngest Dacite porphyry
0,8 zircon (sample EA-561)

0,7 Crustal zircon: Median ~300 ppm U


Green areas >90% crustal Zrn
0,6
50

0,5 100
Th/U

200
0,4
400
0,3 800
1600
3200 Old
0,2
6400
zircon
0,1 core

0,0
30 50 70 90 110 130 150 170 190
206Pb/ 238U age (Ma)
App. 3, Fig. 2. Th/U vs. 206Pb/238U CA-TIMS diagram showing different mixing lines, for different U concentrations, between
the youngest zircon from the dacite porphyry and the inherited Jurassic zircon from the same dacite porphyry and several
mixing lines. Crustal zircon median value estimated after Belousova et al. (2002). Note that the only candidates for old-core
end members (those that have a low enough Th/U ratio to account for the mixing) are the middle Cretaceous (102.07 ±
0.15 Ma; 206Pb/238U CA-TIMS age) zircons from the Rojo Grande granite (sample EA-567) and the early Jurassic (177.6 ±
3.7 Ma; 206Pb/238U LA-ICP-MS age) zircons from the dacite porphyry (sample EA-561). To assess the mixing hypothesis using
the Jurassic zircon data, several mixing lines for different U concentrations were drawn since the U content of this zircon was
not available. Here, the dacite porphyry ages plot along a mixing line, but only at very high U contents (≥5,000 ppm). These
U contents, however, are higher than expected for crustal zircons from granodiorites and tonalites, which are lower than
~3,000 ppm and typically less than ~1,000 ppm (Belousova et al., 2002), making that mixing highly improbable.

You might also like