You are on page 1of 34

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

net/publication/283344848

Petrological Evolution of the Magmatic Suite Associated with the


Coroccohuayco Cu(–Au–Fe) Porphyry–Skarn Deposit, Peru

Article in Journal of Petrology · November 2015


DOI: 10.1093/petrology/egv056

CITATIONS READS

31 985

4 authors, including:

Cyril Chelle-Michou Massimo Chiaradia


ETH Zurich University of Geneva
78 PUBLICATIONS 1,566 CITATIONS 347 PUBLICATIONS 9,478 CITATIONS

SEE PROFILE SEE PROFILE

Paul Béguelin
University of Münster
17 PUBLICATIONS 200 CITATIONS

SEE PROFILE

All content following this page was uploaded by Cyril Chelle-Michou on 06 December 2015.

The user has requested enhancement of the downloaded file.


JOURNAL OF Journal of Petrology, 2015, Vol. 56, No. 9, 1829–1862
doi: 10.1093/petrology/egv056
PETROLOGY Original Article

Petrological Evolution of the Magmatic Suite


Associated with the Coroccohuayco Cu(–Au–Fe)
Porphyry–Skarn Deposit, Peru
Cyril Chelle-Michou1,*, Massimo Chiaradia1, Paul Béguelin1 and
Alexey Ulianov2

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


1
Department of Earth Sciences, University of Geneva, Rue des Maraı̂chers 13, 1205 Geneva, Switzerland and
2
Institute of Earth Sciences, University of Lausanne, Geopolis, 1015 Lausanne, Switzerland
*Corresponding author. E-mail: cyril.chelle-michou@unige.ch
Received May 8, 2015; Accepted September 4, 2015

ABSTRACT
The petrological evolution of magmatic rocks associated with porphyry-related Cu deposits is
thought to exert a first-order control on ore genesis. It is therefore critical to understand and recog-
nize petrological processes favourable to the genesis of porphyry systems. In this study we present
new petrographic, geochemical (whole-rock and mineral), and isotopic (Pb, Sr, Nd) data for rocks
from the magmatic suite associated with the Eocene Coroccohuayco porphyry–skarn deposit, south-
ern Peru. Previously determined radiometric ages on these rocks provide the temporal framework
for interpretation of the data. Arc-style magmatic activity started at Coroccohuayco with the emplace-
ment of a composite precursor gabbrodiorite complex at c. 404 Ma. After a nearly 5 Myr lull, mag-
matic activity resumed at c. 356 Ma with the rapid emplacement of three dacitic porphyries associ-
ated with mineralization. However, zircon antecrysts in the porphyries show that intra-crustal
magmatic activity started c. 2 Myr before porphyry emplacement and probably built a large intra-
crustal magmatic body with an associated large thermal anomaly. Our data suggest that all magmas
underwent a period of evolution in the deep crust before transfer and further evolution in the upper
crust. The gabbrodiorite complex was sourced from a heterogeneous deep crustal reservoir and was
emplaced at a pressure of 100–250 MPa where it underwent a limited amount of fractionation and
formed a chemically zoned pluton. Its initial water content and oxygen fugacity were estimated to be
around 3 wt % H2O and NNO 6 1 (where NNO is the nickel–nickel oxide buffer), respectively. The
deep crustal source of the porphyries appears to have been more homogeneous. The porphyries are
interpreted to be the product of advanced differentiation of a parental magma similar to the gabbro-
diorite. Most of this evolution occurred at deep crustal levels (around 800 MPa) through fractionation
of amphibole þ pyroxene þ plagioclase 6 garnet, leading to the development of a high Sr/Y signature
characteristic of porphyry-related magmatism worldwide. Subsequent upper crustal evolution
(100–250 MPa) was dominated by crustal assimilation, cannibalism of previously emplaced magma
batches (proto-plutons) and magma recharge. Water content and oxygen fugacity were estimated to
be around 5 wt % H2O and NNO þ 1 to NNO þ 2, respectively, at the end of the period of upper crustal
evolution. This high oxygen fugacity is inferred to have favoured sulphur and metal enrichment in
the melt. The high thermal regime generated through 2 Myr of sustained magmatism in the upper
crust favoured crustal assimilation, proto-pluton cannibalism, and efficient metal extraction upon
fluid exsolution. The Coroccohuayco magmatic suite appears to have acquired its metallogenic
potential (high fO2, high Sr/Y) through several million years of deep crustal evolution.

Key words: deep crustal evolution; high Sr/Y magma; magma fertility; porphyry Cu deposit; proto-
pluton cannibalism

C The Author 2015. Published by Oxford University Press. All rights reserved. For Permissions, please e-mail: journals.permissions@oup.com
V 1829
1830 Journal of Petrology, 2015, Vol. 56, No. 9

INTRODUCTION At Coroccohuayco magmatism spans nearly 5 Myr


Porphyry systems (porphyry Cu–Mo–Au, Cu–Zn–Pb–Au and temporally evolves from basic to silicic, although
skarn, Cu–Zn–Pb–Au–Ag epithermal) result from the only the late silicic stages are directly linked to ore gen-
focused release of saline metal- and sulphur-bearing esis (Chelle-Michou et al., 2014, 2015). Such a long-lived
hydrothermal fluids from a cooling upper crustal system offers the opportunity to provide a time-
magma chamber (e.g. Sillitoe, 2010). These systems are constrained perspective on the petrological evolution of
regarded as the result of the positive interplay between a porphyry system. We use petrographic observations,
tectonic, magmatic and hydrothermal processes major and trace element geochemistry of minerals and
whole-rocks, radiogenic isotopes of whole-rocks (Pb,
(Tosdal & Richards, 2001; Richards, 2013). They com-
Sr, Nd), trace element modelling, and previously pub-
monly occur in Cenozoic to Mesozoic fossil magmatic
lished ages (Chelle-Michou et al., 2014, 2015) to provide
arcs, although some older examples are also found.
a time-resolved reconstruction of the evolution of inten-
Because magmas are recognized as the source of most
sive parameters (P–T–H2O–fO2) and the geochemistry
of the metals and fluids (Hedenquist & Lowenstern,
of the magmatic system. Ultimately, we aim to con-
1994; Dreher et al., 2005; Pettke et al., 2010; Simon &
strain the locus and processes of magma generation
Ripley, 2011; Stern et al., 2011) it is critical to under-

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


and evolution and to understand their impact on
stand how a magmatic system will become productive
the genesis of the Cu(–Au–Fe) porphyry–skarn
and ultimately favour the genesis of economic mineral-
mineralization.
ization. The common association of high Sr/Y magma
with porphyry systems (Sr/Y > 20–40; Thiéblemont
et al., 1997; Sajona & Maury, 1998; Oyarzun et al., 2001;
Rohrlach & Loucks, 2005; Richards & Kerrich, 2007; EOCENE GEODYNAMIC SETTING, ARC
Shafiei et al., 2008; Chiaradia et al., 2009a; Hou et al., MAGMATISM AND PORPHYRY SYSTEMS
2009; Shen et al., 2009; Schütte et al., 2010; Richards IN SOUTHERN PERU
et al., 2012; Loucks, 2014) suggests that specific petro-
The Coroccohuayco porphyry–skarn deposit is part of
logical processes play a key role in the genesis of
the Tintaya mining district (which also includes the de-
economic mineralization.
posits of Tintaya and Antapaccay) located at the south-
Many studies have highlighted the importance of
ern edge of the Eocene Andahuaylas–Yauri batholith in
deep crustal petrological processes in producing hy-
southern Peru (inset Fig. 1; Perelló et al., 2003; Maher,
drous, oxidized, S-rich, metal-rich and high Sr/Y mag-
2010). This batholith has recently emerged as an
mas associated with porphyry systems (e.g. Richards,
important metallogenic belt, owing to major discoveries
2011a, 2011b, and references therein). However, others
of world-class porphyry copper systems (e.g.
have proposed that petrological processes happening
Antapaccay, Las Bambas).
in an upper to mid-crustal magma chamber may be of
During the Eocene period flattening of the subduct-
primary importance. Such processes may involve the
ing Nazca slab in the Central Andean Zone resulted in a
following: (1) mafic magma underplating and degassing northeastward inland migration (up to 200 km) of the
significant amounts of sulphur- and metal-rich mag- magmatic arc compared with its previous positions (in-
matic fluids into an overlying felsic magma (Keith et al., set Fig. 1; Mamani et al., 2010). This timing also corres-
1997; Hattori & Keith, 2001; de Hoog et al., 2004; Blundy ponds to the Incaic compressional event that initiated
et al., 2015); (2) saturation of magmatic sulphides fol- crustal-scale thickening and shortening in the central
lowed by remobilization through their destabilization by Andes (Sandeman et al., 1995; Carlotto, 1998; Mamani
magmatic–hydrothermal fluids (Keith et al., 1997; Halter et al., 2010), oroclinal bending (Roperch et al., 2006,
et al., 2002, 2005; Nadeau et al., 2010; Wilkinson, 2013). 2011; Arriagada et al., 2008) and the initial phase of
Here we present a detailed petrological study of the building of the Altiplano (McQuarrie et al., 2005). As
Eocene magmatic suite associated with the crustal thickening, shortening, bending and uplift con-
Coroccohuayco porphyry–skarn Cu(–Au–Fe) deposit, tinued in the central Andes, the Nazca plate motion vec-
southern Peru. Studies of magmatic systems associated tor changed at c. 28–25 Ma to become normal and the
with porphyry ore deposits aim at reconstructing the convergence rate increased (Somoza & Ghidella, 2012).
depth- and time-integrated evolutionary path (in terms The angle of subduction has steepened and the mag-
of pressure, temperature and chemical composition) of matic arc has slowly migrated trenchward from then on
magma from genesis to emplacement. However, at (inset Fig. 1; Mamani et al., 2010).
most fossil porphyry systems (such as Coroccohuayco) The Andahuaylas–Yauri batholith is the plutonic rem-
the available part of the magmatic system (porphyry nant of this Eocene calc-alkaline magmatic arc (Mamani
stocks and dykes) represents only a very small window et al., 2010), and is continuous to the south with the
to study petrological processes through the entire Chilean Eocene magmatic arc along the Cordillera de
continental crust. Indeed, porphyritic rocks at ore depth Domeyco that hosts some of the largest porphyry sys-
typically represent small apophyses episodically tems in the world (Perelló et al., 2003). It now forms a
extruded from a much larger pluton lying at greater 300 km long and up to 130 km wide batholith emplaced
depth. into Mesozoic to early Cenozoic marine sedimentary
Journal of Petrology, 2015, Vol. 56, No. 9 1831

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 1. Geological map of the Coroccohuayco prospect. Inset shows the location of the Andahuaylas–Yauri Batholith and the south-
ward extension of the Eocene arc in Chile.

sequences. In a regional study, Perelló et al. (2003) iden- the contact between the felsic porphyritic rocks and the
tified an early dominantly mafic phase of magmatism Cretaceous limestone (Ferrobamba formation) through-
followed by a dominantly felsic one, the latest being lo- out the district.
cally associated with porphyry systems.
The Tintaya district represents a tectonically uplifted
window of Mesozoic basement surrounded by GEOLOGY OF COROCCOHUAYCO AND ITS
Cenozoic volcano-sedimentary basins. It was affected MAGMATIC SUITE
by two phases of Eocene magmatic activity that Like porphyry-related magmatism elsewhere (Sillitoe,
intruded Cretaceous sandstones and limestones. Cu– 2010), Coroccohuayco records poly-phasic magmatic
Fe-rich skarn-type mineralization is commonly found at activity (Fig. 1 and Fig. 2). From 404 Ma to 402 Ma a
1832 Journal of Petrology, 2015, Vol. 56, No. 9

heterogeneous, dominantly mafic complex termed the magmatic suite. Instead, it is temporally related to the
gabbrodiorite complex (GDC) was emplaced (Chelle- Tacaza arc that marked the beginning of the trenchward
Michou et al., 2014). It is mainly composed of horn- migration of arc magmatism in southern Peru (see
blende gabbro and diorite with subordinate amounts of Mamani et al., 2010).
their leucocratic equivalents and very minor anortho-
site. The GDC crops out continuously over an area of
10 km2. Drill-hole intersections show that, at least just SAMPLES AND ANALYTICAL METHODS
above the Coroccohuayco skarn ore body, it is Rock samples were collected from both outcrops and
emplaced as an 250 m thick sill-like intrusion capping drillcore. Only the least hydrothermally altered samples
Cretaceous limestones, with local deeper roots of each rock type were selected for geochemical
(Chelle-Michou et al., 2015). The total preserved volume and isotopic analysis on the basis of hand lens and thin
of this complex is estimated to be in the range of 3–10 section observations. However, all samples of the rhyo-
km3. dacite display extensive alteration of mafic minerals
After a nearly 5 Myr lull, magmatism resumed with and moderate to pervasive sericitization of plagioclase
the sequential intrusion of three porphyries of dacitic to and groundmass. To constrain crustal assimilation, we

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


rhyodacitic composition at around 356 Ma (Chelle- also included three Triassic red-bed samples of the Mitu
Michou et al., 2014). This started with the emplacement group from the Mamuera section (60 km NE of
of stocks and NW–SE- and ENE–WSW-trending dykes Coroccohuayco; Reitsma, 2012). On the basis of xeno-
of hornblende-bearing porphyry, closely followed by a crystic zircon age distribution and Hf isotopes, such
hornblende- and biotite-bearing porphyry. On the basis sediments have been proposed as assimilants for the
of high-precision U–Pb geochronology, Chelle-Michou porphyries (Chelle-Michou et al., 2014).
et al. (2014) estimated that the porphyries were Approximate modal mineral abundances were ob-
emplaced in less than 86 kyr. They are temporally and tained from image analysis of scanned thin sections
genetically related to the ore-forming hydrothermal with an estimated uncertainty of <1% (>1000 points).
event (Chelle-Michou et al., 2015). These first two por- Microprobe analysis of amphibole, pyroxene, plagio-
phyries were subsequently crosscut by subvolcanic clase and K-feldspar [Supplementary Data (SD) Tables
NW–SE-trending rhyodacite porphyry dykes marking 1–4; supplementary data are available for downloading
the end of the Eocene magmatism at Coroccohuayco. at http://www.petrology.oxfordjournals.org] were car-
Although the rhyodacite commonly displays pervasive ried out at the University of Lausanne on a JEOL 8200
hydrothermal alteration, it postdates the main ore electron microprobe equipped with five wavelength-dis-
stage. Its emplacement age could not be resolved pre- persive spectrometers. We employed an accelerating
cisely, but it is believed to have intruded in the 356– voltage of 15 kV, a beam current of 15 nA, a spot size of
350 Ma time interval (Chelle-Michou et al., 2014). Zircon 3 mm and measuring time of 10–30 s on peak and half
antecrysts in the three porphyries indicate that deep- this time on the respective backgrounds before and
seated magmatism had actually resumed by 375 Ma, after the peak depending on the element and the min-
and continued until the emplacement of the rhyodacite eral analysed. Both natural and synthetic silicates,
at upper crustal levels (Chelle-Michou et al., 2014). This oxides and sulphate standards were used for external
sustained 2 Myr period of intra-crustal magmatic activ- calibration.
ity led to the building of a stable thermal anomaly that In situ trace element abundances in amphibole,
may have favoured cannibalization of previously plagioclase and pyroxenes (SD Tables 1–4) were deter-
emplaced magma batches (proto-plutons) and crustal mined on polished thin sections by laser ablation in-
assimilation (Chelle-Michou et al., 2014). Using the zir- ductively coupled plasma mass spectrometry (LA-ICP-
con Ce/Nd ratio as a proxy for magma oxidation state MS) using a Thermo ELEMENT XR sector-field ICP-MS
(similar to the zircon Ce anomaly or zircon Ce4þ/Ce3þ system interfaced to a UP-193FX ArF excimer laser ab-
ratio), Chelle-Michou et al. (2014) showed that the por- lation system at the Institute of Earth Sciences of the
phyries (Ce/Nd of 9–46) are significantly more oxidized University of Lausanne. The mass spectrometer opti-
than the GDC (Ce/Nd of 2–12). For Eocene zircons from mization was similar to that described by Ulianov et al.
the porphyries (autocrysts and antecrysts spanning the (2012). Operating conditions of the ablation system
c. 375–356 Ma age interval), they also noted an ab- included a repetition rate of 12 Hz, a pit size of 35–50 mm
sence of correlation between zircon Ce/Nd and age, and an on-sample energy density of 50 J cm–2. The
suggesting that the oxidized character of the magma NIST standard glass SRM 612 was employed for exter-
was already established at 375 Ma and persisted until nal standardization. SiO2 and CaO determined by elec-
the emplacement of the porphyries at 356 Ma. tron microprobe analysis on the same spot as
Magmatism at Coroccohuayco finally terminated at subsequent ablation served as an internal standard for
266 Ma with the intrusion of north–south-trending feldspars (plagioclase, K-feldspar) þ orthopyroxene and
dykes of clinopyroxene-bearing alkali basalt (Chelle- amphibole þ clinopyroxene, respectively. Raw data
Michou et al., 2015). Because of its temporal relation- were reduced off-line using the LAMTRACE software
ship with the other magmatic rock, the alkali basalt is (Jackson, 2008). The reproducibility (1r) of the meas-
not considered to be part of the Coroccohuayco ured trace element abundances in the SRM 612 glass
Journal of Petrology, 2015, Vol. 56, No. 9 1833

standard generally range from <1 relative % (rel. %) to PETROGRAPHY AND MINERAL CHEMISTRY
<10 rel. %, depending on the element and the analytical Gabbrodiorite complex (GDC)
list (feldspar or amphibole). Rocks from the GDC are mainly composed of plagio-
Whole-rock samples were crushed using a steel jaw clase, calcic amphibole, Fe–Ti oxides and minor
crusher and powdered with an agate mill (<70 mm). amounts of clinopyroxene and/or orthopyroxene with
Fused glass beads (fluxed with Li2B4O7) and pressed variable mineral proportions and grain sizes (Fig. 2).
powder pellets from whole-rock powders were ana- Accessory phases include quartz, K-feldspar, zircon,
lysed for major, minor and some trace elements by X- apatite, and minor titanite and biotite. Plagioclase is eu-
ray fluorescence (XRF) at the University of Lausanne hedral to subhedral and normally zoned, with Ca-richer
(Table 2, SD Table 5). Data quality was controlled with cores and Ca-poorer rims (Fig. 3a and b). Amphibole
the BHVO, NIM-N, NIM-G, SY-2, SDC-1 and QLO stand- crystals are commonly overgrown on a resorbed core
ards. Average reproducibility (1r) of major and trace of either clinopyroxene or orthopyroxene (Fig. 3c and
elements on these standards is <1 rel. % for all major d). Clinopyroxene cores tend to be more resorbed and
elements (except for MgO and K2O, <3 rel. %) and <10 skeletal than orthopyroxene cores and are sometimes
rel. % for trace elements (<03 rel. % for Sr). Rare earth hydrated and replaced by tremolite–actinolite.

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


elements (REE) and additional trace elements were ana- Amphibole is always poikilitic and encloses plagioclase
lysed by LA-ICP-MS on fused glass bead (Table 2, SD and Fe–Ti oxides (Fig. 3b–e). Quartz, sodic plagioclase,
Table 5) fragments previously used for XRF analysis. Fe–Ti oxides (magnetite and ilmenite), K-feldspar, apa-
Analysis were carried out at the University of Lausanne tite, zircon and sparse interstitial titanite and biotite fill
either on a Thermo ELEMENT XR sector-field ICP-MS the inter-mineral spaces (Fig. 3a and b).
system interfaced to an UP-193FX ArF excimer laser The modal mineralogy of each sample (SD Table 5)
ablation system or a Perkin–Elmer ELAN 6100 DRC highlights the diversity of the GDC. Modal amphibole
quadrupole ICP-MS system interfaced to a GeoLas and plagioclase contents are inversely correlated and
200M 193 nm Lambda Physik excimer laser ablation vary from 581 to 137 vol. % and from 780 to 372 vol.
system. Operating conditions of the ablation systems %, respectively. The textural and mineralogical hetero-
included a repetition rate of 15 or 10 Hz, a pit size of 100 geneity of the gabbrodiorite complex is also observed
or 120 lm and an on-sample energy density of 50 or at the hand sample scale. For example, clinopyroxene-
168 J cm–2 depending on the equipment used. The and orthopyroxene-bearing (in the core of amphiboles)
NIST standard glass SRM 612 was employed for exter- zones locally form banded textures (Fig. 3e). Similarly,
nal standardization and the Sr content determined by coarser grained zones and finer grained zones or zones
XRF for each sample served as an internal standard. richer or poorer in mafic minerals are locally observed
Raw data were reduced off-line using the LAMTRACE next to each other, with sharp and sinuous contacts
software (Jackson, 2008). For each sample three points (Fig. 3f).
were measured and results were averaged. The repro- On the basis of modal abundances of plagioclase
ducibility (1r) of the measured trace element abun- and amphibole (Table 1) as well as whole-rock Eu
dances in the SRM 612 glass standard was <3 rel. % anomalies (see below), we have subdivided the GDC
for all elements whereas reproducibilities of repeated into three units: (1) ‘gabbrodiorite cumulate’ (the cumu-
analysis on the same sample are <10% for all elem- late nature of this unit is discussed below) characterized
ents (decreasing uncertainty with increasing element by positive Eu anomalies and an amphibole-rich min-
content). eralogy (>31 vol. %); (2) ‘gabbrodiorite’ sensu stricto,
Selected whole-rock samples were analysed for defined by the absence of Eu anomalies and a modal
their Sr, Nd and Pb isotopic composition (Table 2; amphibole content of 28–33 vol. %; (3) ‘leuco-
SD Table 5) at the University of Geneva. About gabbrodiorite’, with no Eu anomaly and modal
120 mg of powdered rock was dissolved over 7 days amphibole content of 10–20 vol. %. Contacts between
using a mixture of 4 ml conc. HF and 1 ml 15M HNO3 the gabbrodiorite ‘cumulate’ and the gabbrodiorite
in Teflon vials on a hot plate (140 C). The sample were not observed in the field and their respective dis-
was then dried on a hot plate, dissolved again in tribution could not be clearly established. However, the
3 ml of 15M HNO3 in closed Teflon vials at 140 C gabbrodiorite ‘cumulate’ appears to be dominant to-
and dried down again. Sr, Nd and Pb separation ward the southern part of Coroccohuayco. The bodies
was carried out using cascade columns with Sr-spec, of leuco-gabbrodiorite are typically small (10 m2 in out-
TRU-spec and Ln-spec resins following a modified crop area) and their 3D geometry is unclear. They in-
method after Pin et al. (1994). Pb was further puri- trude the gabbrodiorite and gabbrodiorite ‘cumulates’
fied with an AG-MP1-M anion exchange resin in a with all of which they display sharp contacts.
hydrobromic medium. Pb, Sr and Nd isotope ratios
were measured either on a Thermo TRITON mass
spectrometer or on a Thermo NEPTUNE Plus mass Plagioclase
spectrometer at the University of Geneva, following At the hand sample scale plagioclase displays a wide
methods described by Chiaradia et al. (2014) and range of compositions from An90 in some cores to as
Béguelin et al. (2015), respectively. low as An15 in the inter-mineral space (Fig. 4b–g). Most
1834 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 2. Photographs of the magmatic rock types encountered at Coroccohuayco. bt, biotite; cum, cumulate; hbl, hornblende; Kfs,
K-feldspar.

plagioclase crystals are normally zoned with a broad (Na þ K)A  05; VIAl < Fe3þ; Fig. 5a and b]. Slightly fluid-
homogeneous core and relatively thin rims of increas- altered hornblendes have bleached colors and their
ingly albitic composition. The few anorthite-rich cores composition shifts toward the field of actinolite and
(>An85) are rounded and overgrown with sharp con- tremolite (Si > 75) as a result of Fe depletion and Ca
tacts by lower An rims. In each sample, plagioclase dis- and Si addition.
plays one or two dominant compositional modes with For ‘fresh’ amphiboles, Mg# varies from 094 to
higher and lower An contents corresponding to the 064; AlT varies from 08 to 20 p.f.u. and correlates
inner core(s) and the rim(s), respectively (Fig. 4b–g). positively with (Na þ K)A (r ¼ 089), AlVI (r ¼ 070; Fig.
The plagioclase composition in the gabbrodiorite ‘cu- 5a–c) and most compatible and incompatible trace
mulates’ is mostly An75–70 and An60–55, whereas elements including Sr, Ba, Y, Ta and REE (Fig. 6a–e).
plagioclase in the gabbrodiorite is mostly An60–55 In detail, some trace elements (e.g. Zr) define at
(with some deviation towards higher and lower values). least two groups of amphiboles at the sample (thin
Plagioclases in the leuco-gabbrodiorite are the most section) scale (Fig. 6c). Hornblendes from the gab-
evolved with a compositional mode at An35–30. brodiorite exhibit slightly lower Sr/Y values (07)
On trace element profiles, the REE (in particular La) than amphiboles from the gabbrodiorite ‘cumulate’
are inversely correlated with An content (Fig. 4b–g). (15), and both have (La/Yb)N values around unity
(Fig. 6d and e).
Hornblende
Amphiboles of the GDC are Mg-rich [Mg# > 05, where Pyroxene
Mg# ¼ Mg/(Mg þ Fe2þ)] and calcic (CaB  15). In the Clinopyroxene hosted in the core of amphibole plots at
classification scheme of Leake et al. (1997), most amphi- the diopside–augite boundary and has Mg# [Mg# ¼ Mg/
boles are magnesio-hornblendes [65 < Si < 75; (Mg þ Fetot)] of 073–079; AlT varies from 0014 to
(Na þ K)A < 05], with a few outliers lying in the fields 0056 p.f.u. (SD Table 3). Orthopyroxene hosted in
of tschermakite [55 < Si < 65; (Na þ K)A < 05] and the core of amphibole has a composition ranging
less commonly magnesio-hastingsite [65 < Si < 75; from En54 to En63, Mg# of 057–066 and AlT of
Journal of Petrology, 2015, Vol. 56, No. 9 1835

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015

Fig. 3. Photomicrographs of samples from the gabbrodiorite complex. (a) Back-scattered electron (BSE) image of zoned plagioclase
with an anorthite-rich core overgrown by more sodic plagioclase. Hornblende, K-feldspar, quartz and Fe–Ti oxides are interstitial
between the plagioclase crystals. (b) BSE image of plagioclase with slightly more sodic rim compared with the core, enclosed in a
hornblende oikocryst. (c) Amphibole oikocryst with resorbed skeletal clinopyroxene in the core (cross-polarized light).
(d) Amphibole oikocryst containing resorbed orthopyroxene and clinopyroxene (plane-polarized light). (e) Orthopyroxene-rich
and -poor zones displaying banding (thin section scan). (f) Sharp contact between fine- and coarse-grained gabbrodiorite. cpx,
clinopyroxene; hbl, hornblende; Kfs, K-feldspar; opx, orthopyroxene; pl, plagioclase; qtz, quartz.

0027–0088 p.f.u. (SD Table 4). Both pyroxenes display Porphyries


a negative Eu anomaly in chondrite-normalized REE The three types of porphyritic rock are distinguished on
patterns (Fig. 7). Clinopyroxene is characterized by de- the basis of phenocryst content and groundmass–
pletion of light REE (LREE) compared with middle REE phenocryst proportions (Table 1, Fig. 2). The ground-
(MREE) and heavy REE (HREE). mass consists of a microcrystalline patchwork of quartz,
1836 Journal of Petrology, 2015, Vol. 56, No. 9

Table 1: Average modal composition (vol. %) of rocks units from the Coroccohuayco magmatic suite

Gabbrodiorite complex Porphyries

Rock type: Gabbro- Gabbro- Leuco- hbl hbl–bio hbl–bio Rhyodacite


diorite diorite gabbro- porphyry porphyry porphyry (n ¼ 4)
cumulate (n ¼ 5) diorite chilled
(n ¼ 6) (n ¼ 2) (n ¼ 7) (n ¼ 4) margin (n ¼ 2)

Groundmass/ 10 10 10 498 (15) 465 (15) 558 (08) 562 (23)
interstitial pockets
Hornblende 381 (99) 304 (18) 159 (21) 99 (29) 61 (21) 46 (09) 28 (08)
Biotite 06 (08) 04 (04) 12 (04) 10 (02) 17 (05)
Clinopyroxene/ 07 (00) 05 (00)
actinolite
Orthopyroxene 22 (29) 28 (29)
Plagioclase 537 (91) 627 (16) 778 (03) 375 (36) 436 (13) 363 (18) 374 (19)
K-feldspar 02 (01) 14 (08) 09 (01) 09 (09)
Fe–Ti oxides 43 (05) 47 (13) 26 (05) 14 (02) 08 (02) 09 (01) 07 (02)
Titanite 01 (02) 06 (02) 03 (00) 03 (00)

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Apatite 02 (01) 01 (00) 02 (01) 03 (01)

Values in parentheses are 1SD.

albitic plagioclase and K-feldspar. The main mineral observed under the optical microscope (Fig. 8d).
phases are hornblende, plagioclase, magnetite, biotite Plagioclase composition is very restricted within a sin-
and K-feldspar. Zircon and apatite are common acces- gle rock type and becomes increasingly albitic from the
sory minerals in all three porphyries, whereas titanite is hornblende porphyry (median An20) to the horn-
observed only in the hornblende and hornblende–bio- blende–biotite porphyry (median An15) and the rhyo-
tite porphyries. dacite (An12; Fig. 4h–j). In each porphyry fine spongy
In the hornblende porphyry, phenocrysts consist of cellular rings commonly separate the homogeneous
euhedral to subhedral plagioclase (375 vol. %), core of plagioclases from their thin overgrowth rims
amphibole (99 vol. %) and magnetite (14 vol. %, (Fig. 8a–d). In the hornblende porphyry, plagioclase
with common maghemitization at its rim), with occa- rims are compositionally more diverse (An20-40) than
sional ‘book biotite’ and K-feldspar. The hornblende– the cores (An18-26; Fig. 4j).
biotite porphyry has more abundant magmatic ‘book’ Trace element profiles in plagioclase show very scat-
biotite (12 vol. %), K-feldspar (12 vol. %), plagio- tered trends in contrast with the relatively homoge-
clase (436 vol. %) and a lesser amount of amphibole neous anorthite composition (Fig. 4h–j). In some grains,
(61 vol. %). Euhedral, wedge-shaped titanite is very La exhibits inverse zoning, sometimes with a relatively
common (up to 09 vol. %) in both of these porphyries. abrupt (over < 50 mm) contact between a low-La core
Locally, chilled margins of the hornblende–biotite por- and a high-La rim (Fig. 4h and i).
phyry were found to contain as much as 10 vol. % more
groundmass than its more crystalline equivalent and
consequently a lesser amount of phenocrysts (except Hornblende
magnetite). The rhyodacite is pervasively altered and Amphiboles may lack zoning or may display oscillatory
amphibole is replaced by chlorite and clay minerals. or patchy zoning (Fig. 8e–g). It is common to find zon-
Magnetite is not preserved and is replaced by hematite. ing-free cores grading out to faint oscillatory zoned
However, it is still possible to estimate the pre-alteration rims (Fig. 8e and f). In such cases, both the broad core
modal mineralogy of this rock, although with a higher and the oscillatory-zoned rim may exhibit superim-
uncertainty. The rhyodacite contains more groundmass posed patchy zoning.
(562 vol. %) and ‘book biotite’ (17 vol. %), and less Amphiboles in the porphyries have slightly lower
amphibole (28 vol. %) than the other porphyries. Mg# [060–075; Mg# ¼ Mg/(Mg þ Fe2þ)] than those from
Titanite was not observed in this rock. Neither mag- the gabbrodiorite complex. In the classification scheme
matic anhydrite nor sulphides were observed in the un- of Leake et al. (1997), most are magnesio-hornblendes
altered porphyries. and less commonly edenite [65 < Si < 75; (Na þ K)A
 05; Fig. 5d and e]. A few high-AlT (>18 p.f.u.) crystals
plot in the field of magnesio-hastingsite. Slightly altered
Plagioclase amphiboles have bleached colors and exhibit compos-
Plagioclase phenocrysts in the porphyries are mostly ition shifts toward the field of actinolite as a result of Fe
unzoned in backscattered electron (BSE) images and depletion and Ca and Si addition. This bleaching mostly
are albite rich (Figs. 8b, 4h–j). Only few crystals display corresponds to the patchy zoning observed in BSE
dissolution–recrystallization textures with a slightly images (Fig. 8f and g).
more albitic overgrowth compared with the dissolved AlT ranges from 08 to 17 p.f.u. for the main amphi-
core (Fig. 8c). However, oscillatory zoning is sometimes bole population and from 18 to 24 for the high-Al
Journal of Petrology, 2015, Vol. 56, No. 9 1837

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015

Fig. 4. Plagioclase composition in (a–g) the GDC and (h–j) the porphyries. Left column: density distribution of plagioclase anorthite
content. Grey shaded bars in samples of the GDC correspond to average plagioclase populations as defined by their anorthite con-
tent (see text for discussion). The red curves corresponds to the density distribution of plagioclase anorthite content theoretically in
equilibrium with hornblende as calculated using the formulation of Holland & Blundy (1994) using the reaction edenite þ alb-
ite ¼ richterite þ anorthite. P and T required for the calculation were obtained by the hornblende-only thermobarometric formula-
tion of Ridolfi & Renzulli (2012). In (i) and (j) the continuous line is without Na-correction, short-dashed line is with an Na – 01 p.f.u.
correction and long-dashed line is with an Na – 02 p.f.u. correction (see text for discussion). Right column: rim (distance ¼ 0 mm) to
core An and La zoning profiles of plagioclases.
1838 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 4. Continued

group. For the main amphibole group, AlT correlates K-feldspar


positively with (Na þ K)A (r ¼ 084), and shows a poor K-feldspars from the hornblende–biotite porphyry and
correlation with AlVI (r ¼ 040; Fig. 5e and f). For a given the rhyodacite are Or60–70. They usually occur as inde-
AlT, amphiboles from the hornblende–biotite porphyry pendent phenocrysts, which incorporate earlier formed
systematically have NaA 01 p.f.u. higher than amphi- plagioclase and hornblende (Figs 2 and 8i), but can also
boles from the hornblende porphyry (Fig. 5e). No other overgrow an earlier plagioclase (Fig. 8j). Both types
systematic difference between amphiboles from the commonly have high Ba contents ranging from 16 to
two rock types was observed. 28 wt % BaO. Barium displays complex concentric zon-
Trace element contents of most amphiboles are ing in BSE images (Fig. 8i) with, from core to rim, a
broadly constant at any given AlT and are typically smooth BaO decrease in each zone and a sharp BaO in-
more homogeneous that in the GDC (Fig. 6). In the horn- crease in the outer rim. K-feldspar overgrown over
blende porphyry, some amphibole cores plot outside plagioclase crystals commonly displays a sieve texture
the main group and define a high-MREE–HREE–Y–Ta similar to the outer rim of K-feldspar phenocrysts (Fig.
group with a more pronounced negative Eu anomaly 8i and j).
(Eu/Eu* ¼ 05–06 versus 06–08 for the main group of
hornblendes) at AlT  14 p.f.u. (circled points in Fig. 6g,
i and j). The high-Al amphiboles are notably enriched in
WHOLE-ROCK GEOCHEMISTRY
Sr (300 ppm) and Ba (150 ppm) and have a high Ba/
La ratio (40) compared with the main amphibole The Coroccohuayco magmatic suite displays a wide
group (Sr ¼ 50 ppm, Ba ¼ 20–120 ppm, Ba/La ¼ 10). range in composition from SiO2 ¼ 489 to 696 wt % with
Amphiboles from the porphyries have Sr/Y and (La/ a compositional gap from SiO2 ¼ 52 to 62 wt % bridged
Yb)N values of 1–4, which is higher than amphiboles by only two volumetrically minor leuco-gabbrodiorite
from the GDC. samples (Table 2, Fig. 9). This chemical evolution is cor-
related with age, with the more basic compositions
being older (i.e. the GDC at c. 404 Ma) and the more sili-
Titanite cic being younger (i.e. the rhyodacite at c. 350 Ma). The
Titanite in the hornblende and hornblende–biotite por- compositional gap also corresponds to the magmatic
phyries is euhedral to subhedral and occurs alone or as lull evidenced by radiometric dating (Chelle-Michou
inclusions in plagioclase, amphibole or biotite. Its in- et al., 2014) from 402 to 356 Ma. Overall, the magmatic
ternal texture can be variable, but it commonly shows suite falls within the calc-alkaline field and extends into
dissolution–overgrowth textures (Fig. 8h). It may con- the high-K calc-alkaline field (Fig. 9g). TiO2, Fe2O3[Tot],
tain ilmenite–hematite inclusions sometimes aligned MgO, MnO, CaO and the high field strength elements
along what could be a growth surface. (HFSE) display negative correlations with SiO2, whereas
Journal of Petrology, 2015, Vol. 56, No. 9 1839

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 5. Amphibole compositions illustrating (a, d) their classification and (b, c, e, f) substitution mechanisms. Colored symbols indi-
cate amphiboles that match the compositional calibration range of the thermobarometric formulations of Ridolfi & Renzulli (2012).
Greyscale symbols correspond to amphiboles that plot outside this compositional range, many of which show clear signs of hydro-
thermal alteration.

K2O and NaO show a positive correlation (Fig. 9a–e, g chondrite-normalized REE patterns and no Eu anoma-
and h). Al2O3 contents are generally high and range lies (Fig. 10a). The P2O5, Zr and REE þ Y contents of the
from 19 wt % for the GDC to 17 wt % for the porphy- gabbrodiorite ‘cumulate’ are typically characterized by
ries (Fig 9f). With increasing SiO2 content P2O5, Ta and lower values (by 30–50%) but similar ratios (e.g. La/Yb,
Y contents increase in the GDC and decrease in the por- Zr/Y) and a positive Eu anomaly (Figs 9i, j, l and 10).
phyries (Fig. 9i, k and l). Zr, Sr and Th are positively cor- Compared with the GDC, the porphyries are more en-
related with SiO2 but are depleted in the rhyodacite riched in LILE and LREE, and depleted in HFSE and
compared with the other porphyries (Fig. 9j, m and n). HREE (Fig. 10b). They show slightly more significant
Over time the magmatic suite exhibits an increase in Sr/ negative Eu and Ti anomalies. REE patterns display a
Y (c. 20–130) and Zr/Y (c. 2–15) from the gabbrodiorite listric shape for all the porphyries, and total REE content
to the rhyodacite (Fig. 10c and d). decreases from the hornblende porphyry to the rhyoda-
Compared with primitive mantle compositions all cite (Fig. 10a).
the magmatic rocks at Coroccohuayco display promin- Isotopic compositions of Sr and Nd are variable for
ent negative Nb and Ta anomalies and an enrichment samples of the GDC but are relatively well correlated for
of large ion lithophile elements (LILE) over HFSE, typical the porphyries (Fig. 11a). The GDC yields higher eNd36
of arc magmas (Fig. 10b); Pb and Sr exhibit positive Ma values (þ26 to þ09) than the porphyries (þ03 to –
anomalies. Samples from the GDC are slightly enriched 13) but similar 87Sr/86Sr36 Ma values (07048–07057). In
in LREE over HREE. Gabbrodiorite and leuco-gabbro- a Sr vs Nd isotope plot, there is a clear temporal trend
diorite samples display a range of overlapping toward increasingly crustal isotopic composition from
1840 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 6. Trace element composition of amphiboles. (a, f) Ba, (b, g) Ta, (c, h) Zr, (d, i) Sr/Y, and (e, j) (La/Yb)N vs AlT. The two groups of
amphiboles best defined in the Zr plots for a given AlT in samples of the gabbrodiorite should be noted; also the high Ta content
and low (La/Yb)N of some amphiboles in the hornblende-porphyry [circled points in (g and j)]. These hornblendes also have high
REE contents (see text for discussion). Legend is the same as in Fig. 5.

the GDC to the rhyodacite (Fig. 11a). The Pb isotopic val- surrounds the accuracy of the calculated intensive vari-
ues of the gabbrodiorite complex and the porphyries, ables (e.g. Walker et al., 2013; Erdmann et al., 2014).
corrected for U and Th decay, overlap (206Pb/204Pb ¼ Therefore, we also indirectly compare the pressure and
1835–1862, 207Pb/204Pb ¼ 1558–1564, 208Pb/204Pb ¼ 38 temperature outputs of this empirical formulation with
37–3861; SD Table 5); the porphyries have more radio- the hornblende–plagioclase thermodynamic formula-
genic present-day values compared with the GDC tion of Holland & Blundy (1994; based on the reaction
(Fig. 11b). All the magmatic rocks at Coroccohuayco edenite þ albite ¼ richterite þ anorthite). Plagioclase–
have transitional values between the Paracas Paleozoic hornblende equilibrium (necessary for the application
basement block and the Arequipa basement block of Holland & Blundy’s formulation) can be difficult to
(Fig. 11b), consistent with the geographical location of evaluate on a thin section scale. Therefore, we have cal-
the Tintaya District at the border between the two culated the composition of plagioclase that would the-
blocks (see Mamani et al., 2010). oretically be in equilibrium with amphibole at the
pressure and temperature values calculated using the
formulation of Ridolfi & Renzulli (2012), leaving the
INTENSIVE PARAMETERS OF MAGMA plagioclase anorthite composition as the only unknown
EVOLUTION (P–T– H2O–fO2) variable. This calculated plagioclase composition is
To determine intensive parameters for the magmas subsequently compared with the analysed one.
(pressure, temperature, melt H2O concentration and
oxygen fugacity) we used the empirical formulation of
Ridolfi & Renzulli (2012). This formulation has the ad- Application of empirical calibrations for
vantage of being based on amphibole-only compos- amphibole
itions, provided that their chemistry matches those of The determination of intensive parameters has been
the amphiboles used for the calibration and that they conducted on amphiboles that match the compositional
grow in equilibrium with their parent melt. However, field used for the calibration of the equations (coloured
despite these obvious advantages, much debate symbols in Figs 5, 6 and 12; Ridolfi & Renzulli, 2012).
Journal of Petrology, 2015, Vol. 56, No. 9 1841

plagioclase than the majority of that observed in these


samples (Fig. 4a–g). In most cases this corresponds to
the actual composition of plagioclase rims. This is con-
sistent with the observation that amphibole crystals are
oikocrysts that grew incorporating previously formed
euhedral plagioclase. Interestingly, for most samples of
the GDC, our calculations show the presence of two
plagioclase populations that are in equilibrium with the
distinct amphibole populations found in these samples
(Fig. 4c–e and g).
Assuming that P and T were correctly estimated for
the porphyries, the plagioclase composition in equilib-
rium with amphibole should be An30–70 (Fig. 4h–j).
Such compositions are rarely found in the porphyries at
Coroccohuayco. However, the albitic compositions

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 7. Clinopyroxene and orthopyroxene chondrite-normalized
observed in plagioclase from the porphyries are not
REE patterns [normalizing values from McDonough & Sun common for water-rich magmas (e.g. Lange et al.,
(1995)]. 2009). Considering a plagioclase composition of 20%
An (i.e. the composition of most plagioclases in the
Amphiboles from both the GDC and the porphyries hornblende porphyry), we also calculated P and T using
yield very similar P, T and H2Omelt values (Fig. 12). the thermobarometric formulation of Holland & Blundy
Profiles within single crystals often result in variable P (1994) in conjunction with the Al-in-hornblende barom-
(and T to a lesser extent), by as much as 200 MPa, with eter of Anderson & Smith (1995) to allow for simultan-
no systematic changes from core to rim. These apparent eous determination of P and T (see Anderson et al.,
magmatic condition changes are unrealistic and may re- 2008). We obtained a range of P and T of 400–100 MPa
sult from the kinetics of amphibole growth where amphi- and 750–650 C, respectively. Although this pressure
bole chemistry is not at equilibrium with the melt (Costa range seems reasonable, such a low temperature range
et al., 2013). A more limited range of pressure is also (around the water-saturated granitic solidus) is unlikely
supported by a lack of significant pressure-sensitive sub- to represent the crystallization temperature of all the
stitutions (e.g. Al-Tschermak; Fig. 5c and f). However, amphiboles in the porphyries. This analysis shows that,
clear evidence for edenite exchange (Fig. 5b and e) sup- in the porphyries, plagioclase and hornblende are no
ports the temperature-controlled compositional variabil- longer in magmatic equilibrium, which prevents the use
ity of the hornblendes. Therefore, we argue that, with the of thermodynamically based thermobarometers for
exception of some anomalous values, our large dataset these rocks.
(400 points) captures the main features of the mag- We conclude that for both the GDC and the porphy-
matic conditions at Coroccohuayco. ries, intensive parameters estimated with the calibrations
The main amphibole population (for both the GDC of Ridolfi & Renzulli (2012) overall provide reasonable es-
and the porphyries) yields pressures between 250 and timates for the Coroccohuayco magmatic suite.
100 MPa (median at 147 MPa), temperatures from 900 to
750 C (median at 833 C) and H2Omelt around 5 6 1 wt %
(Fig. 12). For the high-Al amphiboles in the hornblende Influence of post-crystallization Na addition in
porphyry, calculated pressures vary from 1600 to amphibole on calculated intensive parameters
1200 MPa, temperatures range from 1100 to 960 C and for the porphyries
the melt water content is around 11 wt %. Several lines of evidence suggest that sodium was
The fO2 determinations from amphiboles of the GDC added to plagioclase and amphibole after the porphy-
yield values of DNNO 6 1 (where NNO is the nickel– ries crystallized: (1) plagioclase found in these rocks es-
nickel oxide oxygen buffer, Fig. 12c). In contrast, amphi- sentially lacks anorthite zoning, is significantly more
boles from the porphyries give a higher fO2 from albitic than plagioclase theoretically in equilibrium with
NNO þ 15 to NNO þ 3. At a given pressure, values from amphibole, and displays zoning of La that is not corre-
the hornblende–biotite porphyry are systematically lated with plagioclase composition in contrast to plagio-
05 log unit higher than those from the hornblende clase from the GDC (Fig. 4); (2) the three porphyries
porphyry (Fig. 12f). High-Al amphiboles give high fO2 exhibit high Na2O content (5–6 wt %), above the com-
around NNO þ 35. positional range of magmatic rocks of the Eocene Anta
arc (Fig. 9h); (3) within a single rock type, plagioclase
composition is relatively homogeneous and becomes
Plagioclase composition theoretically in increasingly albitic from the hornblende porphyry to the
equilibrium with hornblende hornblende–biotite porphyry and the rhyodacite (Fig.
For the GDC, our calculations show that amphiboles 4h–j); (4) for a given amphibole AlT, NaA is systematic-
should be in equilibrium with slightly more sodic ally 01 p.f.u. higher in the hornblende–biotite
1842 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015

Fig. 8. Photomicrographs of samples from the porphyries. (a) Amphibole and plagioclase phenocrysts (plane-polarized light). The
irregular spongy cellular ring in the plagioclase should be noted. (b) Plagioclase aggregate with weak zoning (BSE image). (c)
Zoned plagioclase with dissolution texture and overgrowth of a slightly less albitic rim (BSE image). (d) Oscillatory zoning plagio-
clase with spongy cellular ring between core and rim (cross-polarized light). This type of zoning is never observed in BSE images.
(e, f) Amphiboles with broad core and fine oscillatory zoning toward the rim (BSE images). (g) Zoned high-Al amphiboles; darker
zones represent actinolite alteration (BSE image). (h) Titanite (BSE image). ttn-1 has a dissolution–overgrowth texture; ttn-2 has il-
menite–hematite drop-like inclusions. (i) Zoned K-feldspar phenocryst in the hornblende–biotite porphyry (BSE image). Microprobe
analyses of Ba (white circles, spot location; orange circles, BaO content) illustrate this zoning. (j) Plagioclase with K-feldspar over-
growth (BSE image). An, plagioclase anorthite content; ap, apatite; pl, plagioclase; ttn, titanite;.
Journal of Petrology, 2015, Vol. 56, No. 9 1843

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 8. Continued

porphyry relative to the hornblende porphyry (Fig. 5e). than those without correction. Na-corrections induce
This Na addition could result from high-temperature the most significant changes on the determination of
sodic alteration that is well documented in porphyry the oxygen fugacity. Reduction of Na of 01 p.f.u and
systems (e.g. Seedorff et al., 2005, 2008) and has been 02 p.f.u yields oxygen fugacities lower by 06 log unit
shown experimentally to occur rapidly on short time- and 12 log units, respectively (Fig. 13d). Importantly, it
scales (Hövelmann et al., 2010). However, such alter- shows that even applying Na-corrections to the data,
ation usually results in dusty and porous plagioclase the oxygen fugacity of the porphyry is NNO þ 1, which
(Engvik et al., 2008; Plümper & Putnis, 2009; Hövelmann is higher than the average oxygen fugacity of the gab-
et al., 2010) and in albite rims around magmatic K-feld- brodiorite complex (Figs 12c, f and 13d). Finally, the ef-
spar (Norberg et al., 2011), which are not observed at fect of these corrections on the calculation of the
Coroccohuayco. Instead, the optical zoning of plagio- plagioclase composition theoretically in equilibrium
clase appears to be preserved (Fig. 8d). with amphibole is negligible (Fig. 4h–j).
Regardless of the actual process responsible for this
Na addition, the impact of post-crystallization incorpor-
ation of Na in the A site of the amphibole on the deter- Influence of the presence of S in the magma on
mination of the intensive parameters need to be calculated intensive parameters for the porphyries
evaluated. We have recalculated the intensive param- The equations of Ridolfi & Renzulli (2012) were cali-
eters by artificially decreasing NaA by 01 p.f.u (035 wt brated with data from a set of S-free experiments.
% Na2O) and 02 p.f.u. (070 wt % Na2O; Fig. 13) for all However, the genetic link between the porphyries and
amphiboles from the porphyries. Subtracting 02 p.f.u the mineralization suggests that the magmatic system
Na corresponds to removing all NaA from the low-Al may have been S-rich at the time of porphyry emplace-
amphiboles, and is therefore considered to be a max- ment. The presence of S in a water-saturated dacitic
imum correction. For the main amphibole population, magma has been experimentally shown to increase the
pressure is found to change relatively by –6 6 14% and Mg# [calculated as Mg/(Mg þ Fe)] of amphibole by up to
–4 6 15% (1SD) when removing 01 and 02 p.f.u. Na, 01 for oxygen fugacities in the field of sulfide stability
respectively. For high-Al amphiboles, the effect of the (<NNO þ 1; Scaillet & Evans, 1999). In contrast, no sig-
Na-correction is much larger with –47% and –64% for nificant difference in amphibole chemistry for S-free
Na – 01 and Na – 02, respectively. For the main group and S-bearing experiments was detected above
of low-Al amphiboles, the Na-correction tends to reduce NNO þ 15 (Scaillet & Evans 1999; Costa et al., 2004).
the scatter in the calculated pressure (Fig. 13a). For tem- Compared with the gabbrodiorite complex, the pres-
perature, Na-corrections yield values 27 6 04% and ence of titanite relative to ilmenite in the porphyries
51 6 05% lower than without correction. At 800 C, this (Wones, 1989; Nakada, 1991; Xirouchakis & Lindsley,
corresponds to a temperature decrease of 22 C and 1998; Frost et al., 2001; McLeod et al., 2011) and the
41 C, respectively (Fig. 13b). Melt water content calcu- higher Ce/Nd in zircon from the porphyries (used as a
lated with Na-correction is mostly within the model un- proxy for the zircon Ce anomaly and magma oxidation
certainty of 6078 wt % H2O (Fig. 13c; Ridolfi & Renzulli, state; Chelle-Michou et al., 2014) suggest that the oxy-
2012). Only for the high-Al amphiboles do Na-correc- gen fugacity of the porphyries is higher than that of the
tions yield water contents 21 wt % and 30 wt % lower GDC; that is, above NNO þ 1. Consequently, the likely
1844 Journal of Petrology, 2015, Vol. 56, No. 9

Table 2: Representative geochemical and isotopic data for the Coroccohuayco magmatic suite
Rock type: Gabbrodiorite Gabbrodiorite Leuco- Hornblende Hornblende– Rhyodacite
‘cumulate’ gabbrodiorite porphyry biotiteporphyry
Sample no.: 10CC61 10CC09 10CC55 10CC16 10CC22 10CC94 10CC40
Sample type: Outcrop Outcrop Outcrop Outcrop Outcrop Drillcore Outcrop
Coord E/DDH#: 256574 256706 256812 256546 257328 A400-199 257497
Coord N/depth: 8344662 8346226 8345420 8345406 8344904 317 m 8345019

Major elements (wt %)


SiO2 4726 5209 4915 5483 6318 6556 6960
TiO2 096 093 100 068 054 036 021
Al2O3 1958 1776 1909 1926 1694 1717 1717
Fe2O3 (tot) 1217 1034 1143 606 450 287 169
MnO 021 016 016 008 004 003 003
MgO 525 449 418 355 160 085 036
CaO 1149 896 963 908 352 283 100
Na2O 175 274 312 446 539 595 568
K2O 028 077 070 063 273 263 288
P2O5 012 011 020 018 021 014 008
LOI 043 070 054 045 053 137 109

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Total 9950 9905 9919 9926 9917 9975 9980
Mg# 046 046 042 054 041 037 030
Trace elements (ppm)
Sr 429 345 428 416 794 877 612
Ga 19 19 21 19 20 21 21
Cu 72 43 304 21 361 5 6
Cr 14 11 12 11 16 8 6
Ba 77 226 168 236 1029 1172 2339
Sc 3143 3991 3370 2610 1126 649 300
V 335 351 283 209 130 86 50
Co 373 299 270 125 113 53 24
Ni 151 77 163 111 93 n.d. 55
Zn 732 524 413 218 164 240 1931
As 090 086 098 096 069 131 120
Rb 68 211 150 86 411 566 754
Y 1181 1981 1784 1771 1546 927 613
Zr 267 759 449 335 1213 1343 1016
Nb 098 245 158 172 493 418 304
Mo 091 032 031 021 063 026 072
Sb n.d. 022 009 023 023 021 024
Cs 026 080 017 036 015 067 180
La 409 755 627 742 1691 1428 620
Ce 980 1793 1508 1553 3423 2910 1018
Pr 135 239 208 230 425 346 165
Nd 664 1119 1040 1072 1763 1388 679
Sm 186 296 274 300 351 276 137
Eu 074 102 096 108 085 060 031
Gd 215 343 338 320 307 207 131
Tb 0327 0520 0502 0548 0436 0296 0189
Dy 233 381 334 356 289 154 113
Ho 0431 0726 0693 0725 0546 0310 0223
Er 130 194 176 189 160 091 063
Tm 0177 0324 0257 0275 0247 0147 0101
Yb 123 209 180 180 158 095 073
Lu 0169 0317 0269 0245 0264 0177 0106
Hf 094 210 134 106 330 403 312
Ta 0072 0149 0096 0108 0309 0260 0191
W 012 2733 046 016 067 183 071
Pb 27 33 20 22 25 43 124
Th 056 152 079 064 460 474 202
U 0200 0550 0265 0170 1549 1697 1204
Measured isotopic ratios
87
Sr/86Sr 07048528 07051751 07052239 07053211 07054340 07055095 07056774
143
Nd/144Nd 05127635 05127206 05127070 05126710 05126206 05126017 05125519
206
Pb/204Pb 18533 18551 18569 18536 18817 18725 18647
207
Pb/204Pb 15616 15616 15625 15626 15651 15639 15635
208
Pb/204Pb 38478 38506 38522 38509 38825 38693 38599
Initial isotopic ratios
87
Sr/86Sri 07048265 07050736 07051728 07052909 07053583 07054151 07054973
143
Nd/144Ndi 05127182 05126779 05126701 05126318 05125924 05125735 05125235
eNdi 2594 1807 1655 0908 0139 0228 1340
206
Pb/204Pbi 18504 18484 18430 18509 18597 18585 18613
207
Pb/204Pbi 15614 15613 15618 15624 15640 15632 15634
208
Pb/204Pbi 38451 38446 38507 38476 38610 38564 38580

DDH#, drillcore number. LOI, loss on ignition; n.d., not detected.


Journal of Petrology, 2015, Vol. 56, No. 9 1845

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015

Fig. 9. Harker variation diagrams for whole-rock samples of the Coroccohuayco magmatic suite. Grey field delineates the compos-
itional range of magmatic rocks of the Eocene Anta arc (mostly from northern Chile) as compiled by Mamani et al. (2010).

presence of S in the melt would not have significantly PETROLOGICAL EVOLUTION


affected the Mg# of amphiboles from the porphyries,
and no influence on the calculation of intensive param- Our main focus is to constrain the deep and upper crus-
eters can be expected. tal processes responsible for the 5 Myr petrological
1846 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 9. Continued

evolution of the Coroccohuayco magmatic system lead- porphyries also suggest that they have a similar
ing to mineralization. Below, we discuss all the source (Fig. 11). The most primitive samples of the
Coroccohuayco magmatic suite evolutionary stages magmatic suite (the gabbrodiorite samples,
from the magma source to deep crustal and upper crus- SiO2  50 wt %) have high Ba/La (20–30) and Pb/Ce
tal evolution from the inception of the magmatic activity (01–02), which is diagnostic of the contribution of flu-
(the GDC) to its end (the rhyodacite). ids that have promoted melting of the mantle wedge.
In line with established models about subduction-
related magmatism, these fluids are likely to have ori-
Magma source ginated by dehydration of the subducting slab (e.g.
All the magmatic rocks at Coroccohuayco display the Tatsumi et al., 1986; Tatsumi, 1989) or of a mélange
strong negative Nb, Ta and Ti anomalies and LILE en- diapir (e.g. Marschall & Schumacher, 2012). The low
richment that are typical of subduction-related mag- Th/La (<02) and Th/Nb (<1) ratios of the GDC suggest
mas (Fig. 10b). The overlapping Pb isotopic values minimal involvement of subducted sediment recycling
corrected for U, Th and Rb decay for the GDC and the (e.g. Plank, 2005).
Journal of Petrology, 2015, Vol. 56, No. 9 1847

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 10. Whole-rock trace element variations. (a) Chondrite-normalized REE patterns [normalizing values from McDonough & Sun
(1995)]. (b) Primitive mantle-normalized trace element patterns [normalizing values from McDonough & Sun (1995)]. (c) Y vs Sr/Y.
(d) Ta vs Zr/Y. Grey lines in (c) and (d) indicate the AFC paths modeled using a Monte Carlo approach that simultaneously repro-
duce the trace element composition (REE, Y, Sr, Zr and Ta) of sample 10CC22 (path 1) and sample 10CC94 (path 2). Large symbols
highlight end-member compositions used for modeling (see text for details).

Deep crustal evolution crystallization of primitive mantle melts and that some
The gabbrodiorite complex (GDC) fractionation of olivine, pyroxene and possibly amphi-
Owing to the mafic composition of the GDC bole has occurred, most probably in the DCHZ.
(SiO2 < 52 wt %), this unit is the one that most closely Experimental studies on hydrous (>3 wt % H2O) primi-
reflects the signature of its source (i.e. least crustal con- tive arc basalts at pressure conditions appropriate for
tamination). In the GDC, the initial zircon Hf isotopic the deep crust (07–12 GPa) show plagioclase to be
composition (eHfi of þ106 to þ52; Chelle-Michou et al., minimally stable, and that fractionation of olivine, pyr-
2014) and the whole-rock Nd isotope composition oxene and amphibole, as well as possible hybridization
(eNd36 Ma of þ26 to þ09; Fig. 11a) are not compatible with amphibole-bearing lithologies (from previous
with a 36 Ma depleted mantle source [eNd of þ6 to þ11; magma batches) result in the formation of low-Mg,
eHf of þ25 to þ11; calculated using present-day values high-Al basalt (MgO < 8 wt % and Al2O3 > 18 wt %) simi-
from Workman & Hart (2005)]. However, their Nd and lar to the gabbrodiorite at Coroccohuayco (Müntener
Sr isotopic composition is similar to the most primitive et al., 2001; Nandedkar et al., 2014).
composition of deep crustal xenoliths exhumed near Several lines of evidence indicate that the GDC does
Cusco (including diorite, gabbro and clinopyroxenite; not represent a single homogeneous magma batch.
inset Fig. 11a; Chapman et al., 2015). This suggests These include: (1) evidence of micro-scale mingling be-
tha the parental magmas of the GDC were sourced tween clinopyroxene-rich and orthopyroxene-rich mag-
from either sub-arc mantle intensely metasomatized by mas that suggests contrasted initial melt compositions
slab-derived fluids or that the mantle-derived melts in- (Fig. 3e and f); (2) variable Sr and Nd isotopic compos-
tensively interacted with less radiogenic material, prob- itions of whole-rock samples (Fig. 11a); (3) heteroge-
ably in a deep crustal hot zone (DCHZ; Annen et al., neous Hf isotopic compositions of zircons at the hand
2006). sample scale (up to 5 epsilon units; Chelle-Michou
However, the very low Ni (10 ppm) and Cr et al., 2014); (4) multiple groups of amphiboles at the
(15 ppm) contents and Mg-numbers (42–48) of the hand sample scale defined by contrasting concentra-
GDC indicate that it does not represent the products of tions of trace elements (e.g. Zr; Fig. 6c). This suggests
1848 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015

Fig. 11. Radiogenic isotope compositions of the studied samples. (a) 87Sr/86Sr vs eNd calculated at 36 Ma. Continuous lines repre-
sent a simple mixing model with three Triassic red sandstones (from the Mamuera section close to Sicuani) with 10% increments
(see text for details). Inset shows the isotopic composition of the Coroccohuayco magmatic rocks compared with depleted MORB
source mantle (DMM; data from Workman & Hart, 2005), and metasedimentary and igneous deep crustal xenoliths from the Cusco
area (Chapman et al., 2015) all calculated at 36 Ma. (b) Pb isotope plots for samples from Coroccohuayco compared with isotopic
reference fields delineated with regional data from Mamani et al. (2010). Legend is the same as in Fig. 9.
Journal of Petrology, 2015, Vol. 56, No. 9 1849

that the GDC was sourced from one compositionally would imply that the DCHZ was located within the lower
heterogeneous magma reservoir or from several dis- crust.
tinct magma reservoirs within a deep crustal hot zone. With respect to the gabbrodiorite complex, the more
Because the GDC marks the beginning of arc magma- homogeneous whole-rock Sr and zircon Hf isotopic
tism in this region, it would be expected that the DCHZ compositions of the porphyries (Fig. 11a; Chelle-Michou
below the Tintaya district at this time had a lower ther- et al., 2014) suggest that they were sourced from a
mal maturity compared with regions where arc magma- more homogeneous magma reservoir in the DCHZ.
tism had persisted at more or less the same position for Prolonged intracrustal magmatic activity (5 Myr) prior
several millions of years. Under these conditions, it is to the emplacement of the porphyries (Chelle-Michou
likely that a newly formed DCHZ hosts several discon- et al., 2014) may have favoured thermal maturation of
nected magma reservoirs, each having a slightly differ- the DCHZ in which large volumes of evolved magma
ent isotopic composition. could be generated and homogenized on thousands to
millions of years timescales (Sisson et al., 2005; Solano
et al., 2012; Melekhova et al., 2013).
The porphyries

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Compared with other magmatic rocks of the Eocene
Anta arc, the dacitic porphyries at Coroccohuayco have Magma ascent and upper crustal evolution
higher Al2O3 (17 wt %) and Sr (600–1000 ppm) and The gabbrodiorite complex (GDC)
lower Fe2O3, MgO, MnO and TiO2 contents (Fig. 9). This Petrographic observations (Fig. 3) indicate that the ear-
suggests that the porphyries result from magmatic dif- liest crystallizing phases in the GDC were An90 plagio-
ferentiation at mid- to lower crustal levels during which clase, clinopyroxene and orthopyroxene. Natural
ferromagnesian minerals fractionate from an evolving observations and experimental studies suggest that
basalt precursor or are residual during melting of a such anorthite-rich plagioclase must crystallize near the
lower crust amphibolite outside the plagioclase stability liquidus at high temperature (1050–1200 C), under hy-
field. The generally high Sr/Y values (increasing drous conditions (2–6 wt % H2O) from arc basalts with
through time from 50 for the hornblende porphyry to high Al and/or high CaO/Na2O values (>8), at various
100 for the rhyodacite; Fig. 10c), the moderate La/Yb pressures (from 200 to 800 MPa; Sisson & Grove, 1993a,
(12) and the listric shape of the chondrite-normalized 1993b; Panjasawatwong et al., 1995; Takagi et al., 2005;
REE patterns (Fig. 10a) point towards the predominance Pichavant & Macdonald, 2007; Alonso-Perez et al., 2009;
of amphibole 6 titanite (6garnet 6 clinopyroxene) in the Zellmer et al., 2012; Nandedkar et al., 2014). The low La
fractionating mineral assemblage or in the restite. The contents of the An90 plagioclase (La < 1 ppm; Fig. 4c
decreasing Ta and P2O5 contents with increasing SiO2 and f) are also in agreement with their crystallization
(Fig. 9i and k) indicate the predominance of titanite and from basaltic melts that have undergone limited frac-
apatite fractionation, respectively. The small negative tionation. The negative Eu anomalies of clinopyroxene
Eu anomalies of the porphyries (Fig. 10a) additionally and orthopyroxene cores (Fig. 7) indicate that they crys-
suggest that some plagioclase fractionation may have tallized just after plagioclase, in agreement with hy-
occurred. drous basalt phase diagrams at 200 MPa (e.g. Sisson &
Experimental studies of fractional crystallization of Grove, 1993a, 1993b; Blatter et al., 2013). Their low Al,
hydrous andesitic and low-Si dacitic melts at 800– Ni and Cr contents also confirm that they crystallized at
450 MPa, 950–850 C and fO2 around the quartz–fayalite– low pressure from a derivative basalt. Clinopyroxene
magnetite (QFM) buffer yield dacite and high-Si dacite crystallization probably caused a rapid decrease in CaO/
melts, respectively, through crystallization of amphibole Na2O that triggered a rapid decrease of the plagioclase
with a lesser amount of plagioclase, garnet (only above molar anorthite content to An75. Subsequent cooling
800 MPa), Fe–Ti oxides and pyroxenes (Sisson et al., and crystallization of clinopyroxene and later horn-
2005; Müntener & Ulmer, 2006; McCanta et al., 2007; blende explains the normal zoning of plagioclase
Alonso-Perez et al., 2009; Nandedkar et al., 2014). (Fig. 4b–g).
Pressures above 800 MPa stabilize garnet at the ex- Towards the south, where the GDC is thicker, an
pense of amphibole (Alonso-Perez et al., 2009), which early crystallizing assemblage from now eroded higher
would result in a higher La/Yb ratio (typically > 20) of levels of the magma chamber probably fractionated
the extracted melt. Assuming that these experimental from the melt and formed an immature ‘cumulate’ (the
studies apply to the Coroccohuayco magmatic system, gabbrodiorite cumulate) with anorthite-rich plagioclase
they would suggest that the DCHZ was located at pres- and positive Eu anomalies. The lower Zr, P2O5 and
sures around (with a small amount of fractionating gar- REE þ Y concentrations of the gabbrodiorite ‘cumulate’
net) or below (no fractionating garnet) 800 MPa (i.e. at a compared with the gabbrodiorite, but with similar ratios
maximal depth of 30 km). The crustal thickness of (e.g. La/Yb, Zr/Y; Figs 9i, j, l and 10; Table 2), cannot be
southern Peru in the Eocene is estimated to be 35– explained by simple crystal accumulation. The higher
40 km, prior to a major orogenic shortening and crustal modal content of poikilitic hornblende (although with
thickening event in the Miocene (Hindle et al., 2005; slightly lower incompatible element concentrations)
Gotberg et al., 2010; Eichelberger et al., 2015), which relative to plagioclase in the ‘cumulate’ (Table 1) cannot
1850 Journal of Petrology, 2015, Vol. 56, No. 9

explain the lower concentrations in these whole-rock minerals originated from this pluton. The emplacement
samples. The low Zr and P2O5 concentrations in the ‘cu- pressure of this magmatic (granitic, sensu lato) body as
mulate’ suggest that these samples host on average constrained by hornblende barometry is around 250–
30–50% less zircon and apatite than the gabbrodiorite. 100 MPa (Fig. 12d–f), which is typical for plutons under-
Thus, because zircon and apatite are large repositories lying porphyry systems (e.g. Dilles, 1987; Cloos, 2001;
of REE þ Y (Bea, 1996), the lower whole-rock REE þ Y Stavast et al., 2008; Sillitoe, 2010; Steinberger et al.,
contents of the gabbrodiorite ‘cumulate’ are most prob- 2013). At such pressures a silicic melt is water saturated
ably due to lower amounts of zircon and apatite. This at 5 6 1 wt % (Fig. 12d and e; Papale et al., 2006).
implies that the interstitial melt from which these min- Although Erdmann et al. (2014) showed that the calibra-
erals crystallized (as well as possibly some hornblende) tion of Ridolfi & Renzulli (2012) does not reliably predict
was compositionally zoned at the scale of the GDC. In the melt H2O content at the Mount Merapi volcano,
this scenario, the more evolved melt (incompatible water content calculated by hornblende hygrometry in
element-rich) formed the gabbrodiorite, whereas the the porphyries at Coroccohuayco (Fig. 12e) is perfectly
less evolved melt (incompatible-poor) formed the gab- consistent with water saturation of the melt at 250–
brodiorite ‘cumulate’. 100 MPa.

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


An90 plagioclase–liquid hygrometry (Lange et al., A previous study of the Hf isotopic compositions and
2009) at a near-liquidus temperature of 1200 C yields a ages of zircon xenocrysts and antecrysts from the por-
melt water content of around 3 wt % (for a melt compos- phyries at Coroccohuayco (Chelle-Michou et al., 2014)
ition similar to the whole-rock gabbrodiorite samples). has highlighted that the upper crustal magmatic
In turn, hornblende hygrometry (Ridolfi & Renzulli, evolution of the felsic magmas is characterized by
2012) yields melt water contents of around 5 wt %, open-system behaviour and dominated by crustal as-
which is the water solubility of an andesitic melt at similation, proto-pluton cannibalism and magma re-
around 200 MPa (e.g. Newman & Lowenstern, 2002; charge. U–Pb age spectra of zircon xenocrysts and
Papale et al., 2006). These contrasted hygrometry re- xenocrystic cores from the porphyries show that during
sults, together with the early and late crystallization of upper crustal evolution, the magmas assimilated
An90 plagioclase and hornblende, respectively, suggest Triassic siliciclastic sediments (Mitu group; Chelle-
that volatile saturation was promoted by isobaric Michou et al., 2014). Furthermore, zircons (autocrysts
plagioclase crystallization from a melt containing ini- and antecrysts) from the porphyries display a temporal
tially 3 wt % dissolved H2O. Mass-balance consider- decrease in initial eHf from 7 at 375 Ma to 4 at
ations require that about 40 wt % of anhydrous 356 Ma, consistent with progressive assimilation of
minerals (plagioclase, Fe–Ti oxides, pyroxenes) crystal- these sediments (with zircon eHf36 Ma of 0 to –6). Such
lize prior to amphibole saturation. This is consistent Triassic sediments (red sandstones, conglomerates and
with both textural observations (Fig. 3) and the modal pelites) crop out 60 km NE of Coroccohuayco at the bor-
mineralogy (Table 1) of the gabbrodiorite. der of the Mesozoic and Cenozoic basins within which
The volumetrically minor leuco-gabbrodiorite is the Andahuaylas–Yauri Batholith was emplaced.
more evolved than the gabbrodiorite and has similar Crustal assimilation is also highlighted by initial iso-
trace (including REE and Eu/Eu*) and minor element topic compositions of Sr and Nd that become increas-
chemistry (Figs 9 and 10; Table 2). Plagioclases in the ingly crustal-like with time (Fig. 11a). A simple mixing
leuco-gabbrodiorite are much more albitic than those model between a magma with the isotopic composition
from the gabbrodiorite and are in equilibrium with of gabbrodiorite sample 10CC55 and the Sr and Nd con-
hornblendes in the same samples (Fig. 5b). This sug- centrations of the porphyries and the Mitu sediments
gests that the leuco-gabbrodiorite magma does not re- suggests that the isotopic composition of the porphy-
sult from advanced fractionation of the gabbrodiorite ries is best explained by 20–30% bulk assimilation of
magma in the upper crust, which would be dominated the sediments (Fig. 11a). Thermodynamic models of as-
by plagioclase and amphibole and would significantly similation–fractional crystallization (AFC) indicate that
change the trace element signature of the melt. Instead, this is an unrealistic amount of assimilation under the
we propose that it acquired its composition through estimated P–T conditions of the upper crust (e.g.
fractional crystallization of ferromagnesian minerals, Glazner, 2007). One possibility to explain this inconsist-
mostly Fe–Ti oxides and pyroxenes, probably in the ency would be that the upper crustal magmatic system
deep crust. underwent periodic recharge that allowed the mainten-
ance of high temperatures and promoted crustal as-
similation (e.g. Spera & Bohrson, 2004). Some amount
The porphyries of reactive bulk assimilation in which the assimilant
It is important to stress that, in line with established partly remains in solid form may also provide a mech-
models for porphyry deposits (Dilles & Proffett, 1995; anism for lower energy assimilation (e.g. Beard et al.,
Shinohara & Hedenquist, 1997; Sillitoe, 2010), the por- 2005; Erdmann et al., 2010). Alternatively, the isotopic
phyries from which we obtained samples probably rep- composition of the parental magma at 36 Ma (before as-
resent small apophyses of a larger pluton lying at similation) may have been more evolved than that of the
greater depth and that the majority of the rock-forming gabbrodiorite, possibly owing to the involvement of a
Journal of Petrology, 2015, Vol. 56, No. 9 1851

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 12. Plot of P vs (a, d) T, (b, e) H2Omelt, and (c, f) fO2 (expressed as DNNO) calculated from the amphibole-only formulations of
Ridolfi & Renzulli (2012) for amphiboles that match the compositional calibration range. Error bars are model uncertainties. Curves
represent H2O solubility calculated for 10 ppm CO2 (continuous line) and 100 ppm CO2 (dashed line) using the model of Papale et al.
(2006) with the average composition of the hornblende porphyry. Dashed arrows indicate the effect of the maximal Na-correction
on the calculated parameters for the high-Al, intermediate-Al and low-Al amphiboles in the porphyritic rocks (see text and Fig. 13).
Legend is the same as in Fig. 5.

greater amount of deep crustal material than in the GDC of years of magma injection (Thompson et al., 2002;
and/or of a higher amount of slab-derived material in the Annen et al., 2006).
fluids metasomatizing the sub-arc mantle. Another pos- Porphyry samples have a concordant zircon age re-
sibility could be that the Mitu samples that we analysed cord spanning nearly 2 Myr, from 375 to 356 Ma; this
are not representative of the composition of the rocks has been interpreted to reflect cannibalism of previ-
assimilated at Coroccohuayco. We should also keep ously emplaced magma batches (or proto-plutons;
open the possibility that the assimilated Mitu sediments Chelle-Michou et al., 2014). The high-Al amphiboles
were not located in the upper crust but in the deep crust, found in the hornblende porphyry are also interpreted
as recently suspected near Cusco on the basis of felsic in terms of proto-pluton cannibalism. Indeed, these
granulite xenoliths sampled by Pliocene to Quaternary amphiboles record Na-corrected pressures of 800 to
volcanism (Chapman et al., 2015). Under this hypothesis, 500 MPa and melt water content around 8 wt %
such an amount of crustal assimilation in a DCHZ can (Figs 12e and 13a, c), which is close to the solubility of
readily happen after a few hundred thousand to millions water in a dacitic melt at this pressure. The high melt
1852 Journal of Petrology, 2015, Vol. 56, No. 9

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Fig. 13. Effect of correcting for amphibole excess NaA on the results of (a) pressure, (b) temperature, (c) melt water content and (d)
oxygen fugacity calculations for the hornblende and hornblende–biotite porphyries. Open symbols correspond to calculations per-
formed with an Na decrease of 01 p.f.u. Filled symbols correspond to calculations performed with an Na decrease of 02 p.f.u.
Shaded areas refer to model uncertainties, which are P 6 115%, T 6 235 C, H2Omelt 6 078 wt % and NNO 6 037 log units (Ridolfi &
Renzulli, 2012).

water content is consistent with the high Ba/La and Sr of titanite inclusions in hornblendes and biotites.
content of these amphiboles. This suggests that the Titanites that display dissolution–recrystallization tex-
magma from which these amphiboles crystallized tures (Fig. 8h) indicate that their early cores were desta-
reached volatile saturation in the deep to middle crust, bilized, probably upon hot magma recharge. Titanite
which induced rapid and abundant crystallization and hosting ilmenite–hematite droplets along growth sur-
inhibited magma ascent to shallower levels. These faces (Fig. 8h) may indicate fluctuating oxygen fugacity
amphiboles were probably sampled by subsequent (e.g. McLeod et al., 2011), possibly as a result of
magma batches at 356 Ma and transferred to upper episodic fluid exsolution (Burgisser & Scaillet, 2007) or
crustal levels. recharge by more reduced magma. Petrographic obser-
Textures of titanite and K-feldspar also highlight vations show that K-feldspar was a late crystallizing
open-system evolution in the porphyries. Although phenocryst in the porphyries (Fig. 8i and j). The Ba zon-
some titanite may have fractionated in the deep crust, ing of K-feldspar phenocrysts from the hornblende–
some hornblendes crystallized in the upper crust retain biotite porphyry and the rhyodacite, which displays a
a record of titanite co-crystallization. Indeed, some smooth decrease and an abrupt increase at dissolution–
hornblendes from the hornblende porphyry with recrystallization surfaces (Fig. 8i), also suggests the oc-
high-MREE–HREE–Y–Ta cores (circled points in Fig. 6g, i currence of magma recharge during K-feldspar
and j; with Na-corrected pressure around 150 MPa) crystallization.
evolving to lower MREE–HREE–Y–Ta rims (part of the Finally, the zircon record of the rhyodacite porphyry
main amphibole group in Fig. 6g, i and j) show that at proves that remelting of proto-plutons was an import-
least part of the titanite crystallization took place in the ant process at Coroccohuayco. Indeed, most zircons
early stage of upper crustal evolution upon intrusion of from the rhyodacite crystallized c. 05 Myr before the
the dacitic magma. This is consistent with the presence supposed emplacement of this porphyry as constrained
Journal of Petrology, 2015, Vol. 56, No. 9 1853

by cross-cutting relationship with the hornblende and primitive gabbrodiorite composition (10CC55), we mod-
hornblende–biotite porphyries (Chelle-Michou et al., elled the evolution to the hornblende (10CC22) and
2014). The Nd and Sr isotopic compositions of this rock hornblende–biotite (10CC94) porphyries as an average
also show that it records the maximal amount of crustal of upper and deep crustal processes. To overcome the
assimilation (Fig. 11a). We propose that magma re- large uncertainties associated with numerous input par-
charge may have triggered partial melting of previous ameters we used a Monte Carlo approach; this consists
magma batches (proto-plutons) and crustal material of generating a large number of outputs by randomly
(Mitu sediments), possibly hybridized with these partial selecting the input parameters within known possible
melts, finally yielding modified magma compositions. ranges. To obtain internally consistent sets of partition
Overall, the magnitude of crustal assimilation, proto- coefficients for the REE3þ þ Y3þ we generated them
pluton cannibalism and remelting appears to have using lattice strain fitting (Blundy & Wood, 2003) from
increased with time in the felsic magmatic system at available sets of partition coefficients. The Sr, Zr and Ta
Coroccohuayco, from the inception of intracrustal mag- partition coefficients were selected within a range of
matism at 375 Ma to the emplacement of the porphy- published values [see Béguelin et al. (2015) for more de-
ries at 356 Ma. This requires an increasing amount of tails on the method]. The fractionating assemblages

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


heat to be added to the system. We propose that sus- were chosen according to the mineralogy and petrog-
tained and possibly increasing magmatic injections in raphy of the samples. However, for the porphyries, be-
the upper crust for 2 Myr before the emplacement of cause part of the fractionation probably happened in
the porphyries (as evidenced by zircon antecrysts) the deep crust and at least part of the corresponding
could have favoured the thermal maturation of the sys- fractionating assemblage may not be accessible (e.g.
tem and increased the local geothermal gradient that garnet, clinopyroxene?), the fractionating assemblage
peaked at the time of porphyry emplacement. Under must be inferred from experimental data (e.g. Müntener
these conditions, convective melting of roof-blocks & Ulmer, 2006). For the porphyries, the assimilant was
(proto-plutons and/or older crust) and/or remobilization modelled with various proportions of Mitu sediments
and mingling or mixing of crystal mush by hot magma and proto-pluton partial melts that were assumed to be
during high flux recharge events may be facilitated. similar to the rhyodacite. Simulations that reproduced
This could provide a mechanism for efficient crustal as- simultaneously all the trace elements of the targeted
similation and proto-pluton cannibalism and re-melting samples within 10–20% (depending on the element)
(Reubi & Blundy, 2008; Schmitt et al., 2010; Chiaradia were considered valid. Results are presented in Fig. 10c
et al., 2011, 2014; Reubi et al., 2011; Kennedy et al., and d, Table 3, and SD Tables 6 and 7.
2012; Simakin & Bindeman, 2012; Cashman & Blundy, The modelling results show that the chemistry of the
2013; Walker et al., 2013). In addition, a large part of the porphyries is consistent with fractionation dominated
crustal assimilation might have occurred in the DCHZ by amphibole and pyroxene with lesser amounts of
through 2 Myr of mantle-derived magma injection. plagioclase. Fractionation of apatite and titanite is
The nature of the recharge magma in the upper crust poorly constrained by the model and is better sug-
is unclear. Crustal assimilation and proto-pluton canni- gested by the whole-rock data (i.e. decreasing P2O5 and
balism would readily be facilitated by mafic magma re- Ta with increasing SiO2; Fig. 9i and k). Interestingly, the
charges. However, we could not find any compelling successful models require the presence of 12–58 wt %
evidence for the involvement of such mafic magma at of garnet in the bulk fractionating assemblage (Table 3).
Coroccohuayco. Alternatively, we could envision inter- This amount of fractionating garnet indicates that deep
mediate to felsic recharge magma that would in any crustal fractionation processes occurred at around
case be hotter than the cooling magmatic system into 800 MPa (c. 30 km depth), where the assemblage amphi-
which it intrudes. The efficiency of cannibalism and as- bole þ pyroxene þ plagioclase 6 garnet is stable (e.g.
similation would then critically depend on flux and vol- Alonso-Perez et al., 2009). As expected, trace element
ume of the recharge events. Combining zircon age data modelling did not help in constraining the amount of
and thermal modelling, Caricchi et al. (2014) estimated assimilation (e.g. Powell, 1984) and the relative propor-
that the volume of such a felsic magma body associated tion of assimilants. The evolution from the hornblende
with the porphyries could be of the order of 500–2000 porphyry to the hornblende–biotite porphyry was
km3, consistent with such potential high-volume re- caused by more extensive hornblende fractionation,
charge events. mostly at the expense of plagioclase. This most prob-
ably reflects limited evolution at upper crustal levels
and more extensive deep crustal evolution for the horn-
Quantitative constraints on the evolutionary blende–biotite porphyry with respect to the hornblende
path of the Coroccohuayco magmatic suite from porphyry.
trace element modelling
To strengthen the petrological model depicted above
we have modelled the end-member whole-rock trace Evidence for coeval uplift and erosion
element compositions (REE, Y, Sr, Ta, Zr) in terms of Amphibole crystallization pressure for the GDC (250–
AFC processes (DePaolo, 1981). Starting with the most 100 MPa) and the porphyries (250–100 MPa) are
1854 Journal of Petrology, 2015, Vol. 56, No. 9

Table 3: Summary of results from the AFC Monte Carlo simulations

F r Assimilant Amph Garnet Plag Apatite Titanite Cpx Opx


(Mm/M0) (Ma/Mc) (%) (%) (%) (%) (%) (%) (%)

Input
Minimum 01 0 1 0 0 0 0 0 0 0
Maximum 06 1 1 80 5 50 1 1 30 5
10CC55 to 10CC22
Minimum 034 005 013 347 12 50 000 001 255 02
Average 036 010 057 454 25 115 015 026 370 32
Maximum 040 015 099 579 50 210 058 099 494 63
10CC55 to 10CC94
Minimum 030 002 001 541 14 00 000 000 164 00
Average 033 005 047 633 28 27 049 025 273 32
Maximum 034 008 099 774 58 58 095 062 367 62

Assimilant: 0 ¼ 100% of average rhyodacite; 1 ¼ 100% of average Mitu sediments. F is final mass of the melt (Mm) over the initial
mass (M0). r is assimilated mass (Ma) over the crystallized mass (Mc).

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


equivalent (Fig. 12). However, whereas the amphiboles plagioclase (Rohrlach & Loucks, 2005; Richards, 2011a;
from the GDC crystallized in situ at c. 402 Ma, amphi- Chiaradia et al., 2012; this study). The common associ-
boles from the porphyries were sampled at c. 356 Ma ation of high Sr/Y magma and porphyry Cu deposits
from a pluton lying at some depth below the GDC and suggests that the fertility of the magmatic system was
transferred into small stocks and dykes. The porphyritic mostly acquired in the DCHZ (e.g. Richards, 2011a;
intrusions and the associated copper mineralization are Chiaradia et al., 2012; Chiaradia, 2014). Magmatic evolu-
now juxtaposed to the GDC. These observations sug- tion in the DCHZ was probably responsible for increas-
gest that erosion was taking place in the Tintaya district ing magma water content from 3 wt % in the basaltic
between 402 and 356 Ma. magma to 5 wt % in the felsic magma (Fig. 14).
Based on fluid inclusion data and textural observa- Concurrent with the increase in the water content,
tion of the skarn and breccia, Maher (2010) proposed the magma oxidation state increased from NNO to
that the ore deposit formed at pressures around around NNO þ 1 to NNO þ 2 (Fig. 14). During such a
25 MPa. This indicates that 75 (100–25) to 225 (250–25) transition in fO2 the dominant sulphur species rapidly
MPa of unloading occurred between 402 and 356 Ma. changes from S2– to S6þ (Jugo et al., 2010), the latter
This corresponds to an exhumation rate of 06–18 mm being one order of magnitude more soluble in silicate
a–1. Such erosion rates are higher than those measured melts (Jugo, 2009), allowing chalcophile elements (such
today in the Andes (e.g. Gregory-Wodzicki, 2000; Safran as Cu) to behave incompatibly during magmatic evolu-
et al., 2005; Bookhagen & Strecker, 2012) but are in the tion. This is probably a key factor for the genesis of a
range of those observed in other porphyry systems porphyry Cu deposit (e.g. Ishihara, 1981; Candela, 1992;
(e.g. Houston & Dilles, 2013). They are also consistent Lynton et al., 1993; Richards, 2003) such as
with the rapid subsidence and infill of the coeval Coroccohuayco. The reasons for this higher oxygen fu-
Eocene sedimentary basins in the future Altiplano re- gacity remain unclear. On a global scale water and
gion associated with the Incaic compressional event element tracers of slab-derived fluids have been found
(Carlotto, 2013; Chelle-Michou et al., 2015; Horton et al., to correlate with fO2 in primitive melts of the mantle
2015). These observations confirm that the wedge (Kelley & Cottrell, 2009). However, the high
Coroccohuayco deposit formed during a period of in- water and LILE contents (e.g. Ba) observed in the por-
tense exhumation and erosion similar to porphyry sys- phyries (Figs 10b and 12e) can be attained by recharge–
tems worldwide (e.g. Skewes & Stern, 1994; Tosdal & assimilation–fractional crystallization processes and do
Richards, 2001; Richards, 2003; Cooke et al., 2005; not require additional input of oxidized slab-derived
Chiaradia et al., 2009b; Sillitoe, 2010; Schütte et al., fluid (e.g. Chiaradia et al., 2014). Some researchers
2011. have proposed that the sub-arc mantle may become in-
creasingly oxidized after several millions years of slab
fluid metasomatism (Rowe et al., 2009; Evans &
THE ROLE OF MAGMATIC PROCESSES IN THE Tomkins, 2011; Evans et al., 2012). On the other hand,
GENESIS OF THE COROCCOHUAYCO ORE trace element systematics suggests that the sub-arc
DEPOSIT mantle may not be more oxidized than mid-ocean ridge
Similar to porphyry-related magmatism elsewhere (e.g. basalt (MORB)-source mantle (Lee et al., 2005, 2010,
Richards & Kerrich, 2007; Loucks, 2014), the magmatic 2012; Mallmann & O’Neill, 2009). These contradictory
rocks temporally associated with the mineralization (the views highlight the need for continued efforts to under-
porphyries) display a high Sr/Y signal as opposed to the stand the process leading to high oxygen fugacity in arc
precursor magmas (Fig. 10c). This is interpreted to re- magmas that favours the genesis of porphyry deposits
flect hydrous magma evolution in the mid- to deep crust (e.g. Richards, 2015). The increase in fO2 recorded at
where amphibole 6 garnet are more stable than Coroccohuayco may be explained by fractional
Journal of Petrology, 2015, Vol. 56, No. 9

Fig. 14. Schematic petrogenetic model for the Coroccohuayco magmatic suite. (a) Emplacement of the GDC from a heterogeneous source in the deep crust. Some amount of differen-
tiation in the upper crust generated the gabbrodiorite and the gabbrodiorite ‘cumulate’ rocks. The leuco-gabbrodiorite probably represents a more evolved magma from the deep
crust. (b) Emplacement of a deep-seated proto-pluton. (c) Growth of proto-pluton by injection of new magma batches sourced from a large and homogeneous deep crustal hot zone.
Transitory periods of crustal and proto-pluton assimilation triggered by new magma batches. (d) Sustained recharge of pristine silicic magma from the deep crust (i.e. uncontamin-
ated magma) where it acquired most of its chemical characteristic through mostly amphibole þ clinopyroxene þ plagioclase 6 garnet fractionation (i.e. high Sr/Y). Increasing amount
of proto-pluton cannibalism and crustal assimilation dominantly in the upper crust. Porphyry emplacement occurred at peak thermal conditions followed by ore-forming fluid exsolu-
tion. Remelting of the early laccolith probably started at this time and may have been caused by heating from the underlying magma reservoir. (e) Remelting of the laccolith continued
until it could be remobilized and emplaced as a rhyodacite (f) Situation today with the outcropping Coroccohuayco magmatic suite, near-surface ore body and its deep-seated plutonic
root. From (a) to (e) erosion of 28–66 km occurred in response to uplift associated with the Incaic orogenic event.
1855

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


1856 Journal of Petrology, 2015, Vol. 56, No. 9

crystallization, degassing, assimilation, varying remelting of a proto-pluton, as suggested by zircon


amounts of slab-fluid metasomatism of the mantle high-precision U–Pb geochronology and cross-cutting
source, or a combination of these [see review by relationships. Partial melt similar to the rhyodacite is
Richards (2015)]. However, we note that the hypothesis likely to have hybridized with pristine (uncontaminated),
that the Triassic Mitu red beds or other oxidized sedi- deep crustal silicic melt. This magmatic history is coeval
ments could be located in the DCHZ (see Chapman with active erosion probably associated with uplift
et al., 2015) instead of in the upper crust (Chelle-Michou caused by the Incaic compressional event. The fertility
et al., 2014) means that their efficient assimilation could of the felsic magma with respect to porphyry genesis
readily drive the observed increase in fO2. was probably acquired in the deep crust through the in-
Sustained upper crustal magmatism from 375 to crease of water content and oxygen fugacity in the melt
356 Ma appears to have increasingly favoured crustal and associated S and Cu enrichment. Proto-pluton can-
assimilation and proto-pluton cannibalism. This neces- nibalism and building of a stable thermal anomaly in
sarily implies that a large volume of magma (of the the upper crust may have favoured Cu and S recycling
order of 1000 km3; Caricchi et al., 2014) was emplaced as well as their efficient extraction from the magma to
in the crust, which constituted a large reservoir of cop- form the ore deposit.

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


per and sulphur able to source the Coroccohuayco min-
eral deposit. The resulting thermal anomaly was
greatest just before the mineralization event and prob- ACKNOWLEDGEMENTS
ably favoured slow cooling of the underlying silicic We thank Anders McCarty for assistance during micro-
magma body and more efficient metal diffusion and probe analysis, and Jean-Marie Boccard, Fabio Capponi
partitioning into exsolving fluids (Stavast, 2006; and Michèle Senn for their help with sample preparation.
Vigneresse, 2007). If previous magma batches were Early discussions with John Dilles, Bruno Scaillet, Luca
able to saturate and sequester some amount of sulfides, Caricchi and Hervé Rezeau, and constructive reviews by
subsequent cannibalism of these sulfide-bearing do- the Editor Gerhard Wörner, Jeremy Richards, Charles
mains may contribute to the final metal budget of the Stern and an anonymous reviewer helped to improve
deposit (Wilkinson, 2013). A similar process could also the quality of this paper. Logistic support by Xstrata
occur in the deep crust where several million years of Tintaya SA during fieldwork was greatly appreciated.
magmatism could build up a sulfide-enriched reservoir
(Lee et al., 2012; Chiaradia, 2014). Subsequent remobili-
zation of these sulphides and transfer to the upper crust FUNDING
could also contribute to the magma metal endowment. This study was funded by the Swiss National Science
Foundation (projects N. 200020-126896 and 200020-
CONCLUSIONS 137663).

Figure 14 presents a general model for the evolution of


the Coroccohuayco magmatic system based on age SUPPLEMENTARY DATA
constraints from Chelle-Michou et al. (2014). At c.
404 Ma the GDC magmas were sourced from a hetero- Supplementary data for this paper are available at
geneous deep crustal hot zone and subsequently under- Journal of Petrology online.
went minor degrees of differentiation in the upper crust
to yield the gabbrodiorite and the gabbrodiorite ‘cumu- REFERENCES
late’. The leuco-gabbrodiorite was probably sourced
Alonso-Perez, R., Müntener, O. & Ulmer, P. (2009). Igneous gar-
from the same deep crustal magma chamber that net and amphibole fractionation in the roots of island arcs:
underwent a slightly higher degree of differentiation. experimental constraints on andesitic liquids. Contributions
After a magmatic lull, magmatism resumed at c. to Mineralogy and Petrology 157, 541–558.
375 Ma with the injection of several magma batches Anderson, J. L. & Smith, D. R. (1995). The effects of temperature
within the upper crust, probably sourced from a more and fO2 on the Al-in-hornblende barometer. American
mature and homogeneous deep crustal hot zone. The Mineralogist 80, 549–559.
Anderson, J. L., Barth, A. P., Wooden, J. L. & Mazdab, F. (2008).
compositions of these magmas were mostly acquired
Thermometers and thermobarometers in granitic systems.
in the deep crust but some amount of fractionation and In: Putirka, K. D. & Tepley, F. J., III (eds) Minerals, Inclusions
assimilation also occurred in the upper crust. Sustained and Volcanic Processes. Mineralogical Society of America
magma injection (recharge) until c. 356 Ma favoured and Geochemical Society, Reviews in Mineralogy and
upper (and to some extent mid-) crustal assimilation of Geochemistry 69, 121–142.
previous magma batches (proto-plutons) and of older Annen, C., Blundy, J. D. & Sparks, R. (2006). The genesis of
intermediate and silicic magmas in deep crustal hot zones.
crust. Because the porphyries record increasing
Journal of Petrology 47, 505–539.
amounts of assimilation with time they are thought to Arriagada, C., Roperch, P., Mpodozis, C. & Cobbold, P. R.
have been extruded from an upper crustal silicic (2008). Paleogene building of the Bolivian Orocline: Tectonic
magma reservoir undergoing steadily increasing ther- restoration of the central Andes in 2-D map view. Tectonics
mal maturation. The rhyodacite results from partial 27, TC6014.
Journal of Petrology, 2015, Vol. 56, No. 9 1857

Bea, F. (1996). Residence of REE, Y, Th and U in granites and formation of giant porphyry-related mineralization
crustal protoliths; implications for the chemistry of crustal (Yanacocha, Peru). Earth and Planetary Science Letters 288,
melts. Journal of Petrology 37, 521–552. 505–515.
Beard, J. S., Ragland, P. C. & Crawford, M. L. (2005). Reactive Chiaradia, M., Vallance, J., Fontboté, L., Stein, H., Schaltegger,
bulk assimilation: A model for crust–mantle mixing in silicic U., Coder, J., Richards, J. P., Villeneuve, M. & Gendall, I.
magmas. Geology 33, 681–684. (2009b). U–Pb, Re–Os, and 40Ar/39Ar geochronology of the
Béguelin, P., Chiaradia, M., Beate, B. & Spikings, R. (2015). The Nambija Au-skarn and Pangui porphyry Cu deposits,
Yanaurcu volcano (Western Cordillera, Ecuador): A field, Ecuador: implications for the Jurassic metallogenic belt of
petrographic, geochemical, isotopic and geochronological the Northern Andes. Mineralium Deposita 44, 371–387.
study. Lithos 218–219, 37–53. Chiaradia, M., Müntener, O. & Beate, B. (2011). Enriched bas-
Blatter, D. L., Sisson, T. W. & Hankins, W. B. (2013). altic andesites from mid-crustal fractional crystallization, re-
Crystallization of oxidized, moderately hydrous arc basalt at charge, and assimilation (Pilavo volcano, western Cordillera
mid- to lower-crustal pressures: implications for andesite of Ecuador). Journal of Petrology 52, 1107–1141.
genesis. Contributions to Mineralogy and Petrology 166, Chiaradia, M., Ulianov, A., Kouzmanov, K. & Beate, B. (2012).
861–886. Why large porphyry Cu deposits like high Sr/Y magmas?
Blundy, J. D. & Wood, B. (2003). Partitioning of trace elements Scientific Reports 2, 685.
between crystals and melts. Earth and Planetary Science Chiaradia, M., Müntener, O. & Beate, B. (2014). Quaternary

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Letters 210, 383–397. sanukitoid-like andesites generated by intracrustal proc-
Blundy, J., Mavrogenes, J., Tattitch, B., Sparks, S. & Gilmer, A. esses (Chacana caldera complex, Ecuador): implications for
(2015). Generation of porphyry copper deposits by gas– Archean sanukitoids. Journal of Petrology 55, 769–802.
brine reaction in volcanic arcs. Nature Geoscience 8, 235– Cloos, M. (2001). Bubbling magma chambers, cupolas, and por-
240. phyry copper deposits. International Geology Review 43,
Bookhagen, B. & Strecker, M. R. (2012). Spatiotemporal trends 285–311.
in erosion rates across a pronounced rainfall gradient: Cooke, D. R., Hollings, P. & Walshe, J. L. (2005). Giant porphyry
Examples from the southern Central Andes. Earth and deposits: characteristics, distribution, and tectonic controls.
Planetary Science Letters 327–328, 97–110. Economic Geology 100, 801–818.
Burgisser, A. & Scaillet, B. (2007). Redox evolution of a degass- Costa, F., Scaillet, B. & Pichavant, M. (2004). Petrological and
ing magma rising to the surface. Nature 445, 194–197. experimental constraints on the pre-eruption conditions of
Candela, P. A. (1992). Controls on ore metal ratios in granite- Holocene dacite from Volcan San Pedro (36 S, Chilean
related ore systems: an experimental and computational ap- Andes) and the importance of sulphur in silicic subduction-
proach. Transactions of the Royal Society of Edinburgh: related magmas. Journal of Petrology 45, 855–881.
Earth Sciences 83, 317–326. Costa, F., Andreastuti, S., Bouvet de Maisonneuve, C. &
Caricchi, L., Simpson, G. & Schaltegger, U. (2014). Zircons Pallister, J. S. (2013). Petrological insights into the storage
reveal magma fluxes in the Earth’s crust. Nature 511, conditions, and magmatic processes that yielded the cen-
457–461. tennial 2010 Merapi explosive eruption. Journal of
Carlotto, V. (1998). Evolution andine et raccourcissement au Volcanology and Geothermal Research 261, 209–235.
niveau de Cusco (13 –16 S), Pérou: Enregistrement sédi- de Hoog, J. C. M., Hattori, K. H. & Hoblitt, R. P. (2004).
mentaire, chronologie, contrôles paléogéographiques, Oxidized sulfur-rich mafic magma at Mount Pinatubo,
évolution cinématique. PhD thesis, Université Joseph Philippines. Contributions to Mineralogy and Petrology 146,
Fourier, Grenoble, 199 pp. 750–761.
Carlotto, V. (2013). Paleogeographic and tectonic controls on DePaolo, D. J. (1981). Trace element and isotopic effects of
the evolution of Cenozoic basins in the Altiplano and combined wallrock assimilation and fractional crystalliza-
Western Cordillera of southern Peru. Tectonophysics 589, tion. Earth and Planetary Science Letters 53, 189–202.
195–219. Dilles, J. H. (1987). Petrology of the Yerington Batholith,
Cashman, K. & Blundy, J. D. (2013). Petrological cannibalism: Nevada; evidence for evolution of porphyry copper ore flu-
the chemical and textural consequences of incremental ids. Economic Geology 82, 1750–1789.
magma body growth. Contributions to Mineralogy and Dilles, J. H. & Proffett, J. M. (1995). Metallogenesis of the
Petrology 166, 703–729. Yerington batholith, Nevada. In: Pierse, F. W. & Bolm, J. G.
Chapman, A. D., Ducea, M. N., McQuarrie, N., Coble, M., (eds) Porphyry copper deposits of the American Cordillera.
Petrescu, L. & Hoffman, D. (2015). Constraints on plateau Arizona Geological Society Digest 20, 306–315.
architecture and assembly from deep crustal xenoliths, Dreher, S. T., Macpherson, C. G., Pearson, D. G. &
northern Altiplano (SE Peru). Geological Society of America Davidson, J. P. (2005). Re–Os isotope studies of
Bulletin, doi: 0.1130/B31206.1. Mindanao adakites: Implications for sources of metals
Chelle-Michou, C., Chiaradia, M., Ovtcharova, M., Ulianov, A. & and melts. Geology 33, 957–960.
Wotzlaw, J.-F. (2014). Zircon petrochronology reveals the Eichelberger, N., McQuarrie, N., Ryan, J., Karimi, B., Beck, S. &
temporal link between porphyry systems and the magmatic Zandt, G. (2015). Evolution of crustal thickening in the cen-
evolution of their hidden plutonic roots (the Eocene tral Andes, Bolivia. Earth and Planetary Science Letters 426,
Coroccohuayco deposit, Peru). Lithos 198–199, 129–140. 191–203.
Chelle-Michou, C., Chiaradia, M., Selby, D., Ovtcharova, M. & Engvik, A. K., Putnis, A., Fitz Gerald, J. D. & Austrheim, H.
Spikings, R. A. (2015). High-resolution geochronology of the (2008). Albitization of granitic rocks: the mechanism of re-
Coroccohuayco porphyry–skarn deposit, Peru: a rapid prod- placement of oligoclase by albite. Canadian Mineralogist 46,
uct of the Incaic orogeny. Economic Geology 110, 423–443. 1401–1415.
Chiaradia, M. (2014). Copper enrichment in arc magmas con- Erdmann, S., Scaillet, B. & Kellett, D. A. (2010).
trolled by overriding plate thickness. Nature Geoscience 7, Xenocryst assimilation and formation of peritectic
43–46. crystals during magma contamination: An experimen-
Chiaradia, M., Merino, D. & Spikings, R. (2009a). Rapid transi- tal study. Journal of Volcanology and Geothermal
tion to long-lived deep crustal magmatic maturation and the Research 198, 355–367.
1858 Journal of Petrology, 2015, Vol. 56, No. 9

Erdmann, S., Martel, C., Pichavant, M. & Kushnir, A. (2014). Mineralogical Association of Canada, Short Course Series
Amphibole as an archivist of magmatic crystallization condi- 40, 305–307.
tions: problems, potential, and implications for inferring Jugo, P. J. (2009). Sulfur content at sulfide saturation in oxi-
magma storage prior to the paroxysmal 2010 eruption of dized magmas. Geology 37, 415–418.
Mount Merapi, Indonesia. Contributions to Mineralogy and Jugo, P. J., Wilke, M. & Botcharnikov, R. E. (2010). Sulfur K-
Petrology 167, 1016. edge XANES analysis of natural and synthetic basaltic
Evans, K. A. & Tomkins, A. G. (2011). The relationship between glasses: Implications for S speciation and S content as func-
subduction zone redox budget and arc magma fertility. tion of oxygen fugacity. Geochimica et Cosmochimica Acta
Earth and Planetary Science Letters 308, 401–409. 74, 5926–5938.
Evans, K. A., Elburg, M. A. & Kamenetsky, V. S. (2012). Keith, J. D., Whitney, J. A., Hattori, K., Ballantyne, G. H.,
Oxidation state of subarc mantle. Geology 40, 783–786. Christiansen, E. H., Barr, D. L., Cannan, T. M. & Hook, C. J.
Frost, B. R., Chamberlain, K. R. & Schumacher, J. C. (2001). (1997). The role of magmatic sulfides and mafic alkaline
Sphene (titanite): phase relations and role as a geochronom- magmas in the Bingham and Tintic mining districts, Utah.
eter. Chemical Geology 172, 131–148. Journal of Petrology 38, 1679–1690.
Glazner, A. F. (2007). Thermal limitations on incorporation of Kelley, K. A. & Cottrell, E. (2009). Water and the oxidation state
wall rock into magma. Geology 35, 319–322. of subduction zone magmas. Science 325, 605–607.
Gotberg, N., McQuarrie, N. & Carlotto Caillaux, V. (2010). Kennedy, B., Wilcock, J. & Stix, J. (2012). Caldera resurgence

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Comparison of crustal thickening budget and shortening es- during magma replenishment and rejuvenation at
timates in southern Peru (12–14 S): Implications for mass Valles and Lake City calderas. Bulletin of Volcanology 74,
balance and rotations in the ‘Bolivian orocline.’. Geological 1833–1847.
Society of America Bulletin 122, 727–742. Lange, R. A., Frey, H. M. & Hector, J. (2009). A thermodynamic
Gregory-Wodzicki, K. M. (2000). Uplift history of the Central and model for the plagioclase–liquid hygrometer/thermometer.
Northern Andes: A review. Geological Society of America American Mineralogist 94, 494–506.
Bulletin 112, 1091–1105. Leake, B. E., Woolley, A. R., Arps, C. E. S., et al. (1997).
Halter, W. E., Pettke, T. & Heinrich, C. A. (2002). The origin of Nomenclature of amphiboles; report of the subcommittee
Cu/Au ratios in porphyry-type ore deposits. Science 296, on amphiboles of the International Mineralogical
1844–1846. Association, Commission on New Minerals and Mineral
Halter, W. E., Heinrich, C. A. & Pettke, T. (2005). Magma evolu- Names. Canadian Mineralogist 35, 219–246.
tion and the formation of porphyry Cu–Au ore fluids: Lee, C.-T. A., Leeman, W. P., Canil, D. & A Li Z.-X., (2005).
evidence from silicate and sulfide melt inclusions. Similar V/Sc systematics in MORB and arc basalts:
Mineralium Deposita 39, 845–863. Implications for the oxygen fugacities of their mantle source
Hattori, K. H. & Keith, J. (2001). Contribution of mafic melt to regions. Journal of Petrology 46, 2313–2336.
porphyry copper mineralization: evidence from Mount Lee, C.-T. A., Luffi, P., Le Roux, V., Dasgupta, R., Albarède, F. &
Pinatubo, Philippines, and Bingham Canyon, Utah, USA. Leeman, W. P. (2010). The redox state of arc mantle using
Mineralium Deposita 36, 799–806. Zn/Fe systematics. Nature 468, 681–685.
Hedenquist, J. W. & Lowenstern, J. B. (1994). The role of mag- Lee, C.-T. A., Luffi, P., Chin, E. J., Bouchet, R., Dasgupta, R.,
mas in the formation of hydrothermal ore deposits. Nature Morton, D. M., Le Roux, V., Yin, Q. Z. & Jin, D. (2012). Copper
370, 519–527. systematics in arc magmas and implications for crust–
Hindle, D., Kley, J., Oncken, O. & Sobolev, S. (2005). Crustal bal- mantle differentiation. Science 336, 64–68.
ance and crustal flux from shortening estimates in the Loucks, R. R. (2014). Distinctive composition of copper-ore-
Central Andes. Earth and Planetary Science Letters 230, forming arc magmas. Australian Journal of Earth Sciences
113–124. 61, 5–16.
Holland, T. & Blundy, J. D. (1994). Non-ideal interactions in cal- Lynton, S. J., Candela, P. A. & Piccoli, P. M. (1993). An experi-
cic amphiboles and their bearing on amphibole–plagioclase mental study of the partitioning of copper between pyrrho-
thermometry. Contributions to Mineralogy and Petrology tite and a high silica rhyolitic melt. Economic Geology 88,
116, 433–447. 901–915.
Horton, B. K., Perez, N. D., Fitch, J. D. & Saylor, J. E. (2015). Maher, K. C. (2010). Skarn alteration and mineralization at
Punctuated shortening and subsidence in the Altiplano Coroccohuayco, Tintaya district, Peru. Economic Geology
Plateau of southern Peru: Implications for early Andean 105, 263–283.
mountain building. Lithosphere 7, 117–137. Mallmann, G. & O’Neill, H. St. C. (2009). The crystal/melt parti-
Hou, Z., Yang, Z., Qu, X., Meng, X., Li, Z., Beaudoin, G., Rui, Z., tioning of V during mantle melting as a function of oxygen
Gao, Y. & Zaw, K. (2009). The Miocene Gangdese porphyry fugacity compared with some other elements (Al, P, Ca, Sc,
copper belt generated during post-collisional extension in Ti, Cr, Fe, Ga, Y, Zr and Nb). Journal of Petrology 50, 1765–
the Tibetan Orogen. Ore Geology Reviews 36, 25–51. 1794.
Houston, R. A. & Dilles, J. H. (2013). Structural geologic evolu- Mamani, M., Wörner, G. & Sempere, T. (2010). Geochemical
tion of the Butte district, Montana. Economic Geology 108, variations in igneous rocks of the Central Andean orocline
1397–1424. (13 S to 18 S): Tracing crustal thickening and magma gener-
Hövelmann, J., Putnis, A., Geisler, T., Schmidt, B. C. & ation through time and space. Geological Society of
Golla-Schindler, U. (2010). The replacement of plagioclase America Bulletin 122, 162–182.
feldspars by albite: observations from hydrothermal Marschall, H. R. & Schumacher, J. C. (2012). Arc magmas
experiments. Contributions to Mineralogy and Petrology sourced from mélange diapirs in subduction zones. Nature
159, 43–59. Geoscience 5, 862–867.
Ishihara, S. (1981). The granitoid series and mineralization. McCanta, M. C., Rutherford, M. J. & Hammer, J. E. (2007). Pre-
Economic Geology, 75th Anniversary Volume, 458–484. eruptive and syn-eruptive conditions in the Black Butte,
Jackson, S. (2008). LAMTRACE data reduction software for LA- California dacite: Insight into crystallization kinetics in a sili-
ICP-MS. In: Sylvester, P. (ed.) Laser Ablation ICP-MS in the cic magma system. Journal of Volcanology and Geothermal
Earth Sciences: Current Practices and Outstanding Issues. Research 160, 263–284.
Journal of Petrology, 2015, Vol. 56, No. 9 1859

McDonough, W. F. & Sun, S. S. (1995). The composition of the Pin, C., Briot, D., Bassin, C. & Poitrasson, F. (1994). Concomitant
Earth. Chemical Geology 120, 223–253. separation of strontium and samarium–neodymium for
McLeod, G. W., Dempster, T. J. & Faithfull, J. W. (2011). isotopic analysis in silicate samples, based on specific
Deciphering magma-mixing processes using zoned titanite extraction chromatography. Analytica Chimica Acta 298,
from the Ross of Mull Granite, Scotland. Journal of 209–217.
Petrology 52, 55–82. Plank, T. (2005). Constraints from thorium/lanthanum on sedi-
McQuarrie, N., Horton, B. K., Zandt, G., Beck, S. & DeCelles, P. ment recycling at subduction zones and the evolution of the
G. (2005). Lithospheric evolution of the Andean fold–thrust continents. Journal of Petrology 46, 921–944.
belt, Bolivia, and the origin of the central Andean plateau. Plümper, O. & Putnis, A. (2009). The complex hydrothermal his-
Tectonophysics 399, 15–37. tory of granitic rocks: multiple feldspar replacement reac-
Melekhova, E., Annen, C. & Blundy, J. D. (2013). tions under subsolidus conditions. Journal of Petrology 50,
Compositional gaps in igneous rock suites controlled by 967–987.
magma system heat and water content. Nature Geoscience Powell, R. (1984). Inversion of the assimilation and fractional
6, 385–390. crystallization (AFC) equations; characterization of contam-
Müntener, O. & Ulmer, P. (2006). Experimentally derived high- inants from isotope and trace element relationships in vol-
pressure cumulates from hydrous arc magmas and conse- canic suites. Journal of the Geological Society, London 141,
quences for the seismic velocity structure of lower arc crust. 447–452.

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Geophysical Research Letters 33, L21308. Reitsma, M. J. (2012). Reconstructing the Late Paleozoic-
Müntener, O., Kelemen, P. B. & Grove, T. L. (2001). The role of Early Mesozoic plutonic and sedimentary record of
H2O during crystallization of primitive arc magmas under south-east Peru: Orphaned back-arcs along the western
uppermost mantle conditions and genesis of igneous pyrox- margin of Gondwana. PhD thesis, Université de Genève, 246
enites: An experimental study. Contributions to Mineralogy pp.
and Petrology 141, 643–658. Reubi, O. & Blundy, J. (2008). Assimilation of plutonic roots, for-
Nadeau, O., Williams-Jones, A. E. & Stix, J. (2010). Sulphide mation of high-K ‘exotic’ melt inclusions and genesis of an-
magma as a source of metals in arc-related magmatic desitic magmas at Volcán de Colima, Mexico. Journal of
hydrothermal ore fluids. Nature Geoscience 3, 501–505. Petrology 49, 2221–2243.
Nakada, S. (1991). Magmatic processes in titanite-bearing da- Reubi, O., Bourdon, B., Dungan, M. A., Koornneef, J. M., Sellés,
cites, central Andes of Chile and Bolivia. American D., Langmuir, C. H. & Aciego, S. (2011). Assimilation of the
Mineralogist 76, 548–560. plutonic roots of the Andean arc controls variations in U-ser-
Nandedkar, R. H., Ulmer, P. & Muentener, O. (2014). Fractional ies disequilibria at Volcan Llaima, Chile. Earth and Planetary
crystallization of primitive, hydrous arc magmas: an experi- Science Letters 303, 37–47.
mental study at 07 GPa. Contributions to Mineralogy and Richards, J. P. (2003). Tectono-magmatic precursors for por-
Petrology 167, 1015. phyry Cu–(Mo–Au) deposit formation. Economic Geology
Newman, S. & Lowenstern, J. B. (2002). VolatileCalc: a silicate 98, 1515–1533.
melt–H2O–CO2 solution model written in Visual Basic for Richards, J. P. (2011a). High Sr/Y arc magmas and porphyry
Excel. Computers and Geosciences 28, 597–604. Cu 6 Mo 6 Au deposits: just add water. Economic Geology
Norberg, N., Neusser, G., Wirth, R. & Harlov, D. (2011). 106, 1075–1081.
Microstructural evolution during experimental albitization Richards, J. P. (2011b). Magmatic to hydrothermal metal fluxes
of K-rich alkali feldspar. Contributions to Mineralogy and in convergent and collided margins. Ore Geology Reviews
Petrology 162, 531–546. 40, 1–26.
Oyarzun, R., Márquez, A., Lillo, J., López, I. & Rivera, S. (2001). Richards, J. P. (2013). Giant ore deposits formed by optimal
Giant versus small porphyry copper deposits of Cenozoic alignments and combinations of geological processes.
age in northern Chile: adakitic versus normal calc-alkaline Nature Geoscience 6, 911–916.
magmatism. Mineralium Deposita 36, 794–798. Richards, J. P. (2015). The oxidation state, and sulfur and Cu
Panjasawatwong, Y., Danyushevsky, L. V., Crawford, A. J. & contents of arc magmas: implications for metallogeny.
Harris, K. L. (1995). An experimental study of the effects of Lithos 233, 27–45.
melt composition on plagioclase–melt equilibria at 5 and 10 Richards, J. P. & Kerrich, R. (2007). Special Paper: Adakite-like
kbar: implications for the origin of magmatic high-An rocks: their diverse origins and questionable role in metallo-
plagioclase. Contributions to Mineralogy and Petrology 118, genesis. Economic Geology 102, 537–576.
420–432. Richards, J. P., Spell, T., Rameh, E., Razique, A. & Fletcher, T.
Papale, P., Moretti, R. & Barbato, D. (2006). The compositional (2012). High Sr/Y magmas reflect arc maturity, high mag-
dependence of the saturation surface of H2O þ CO2 fluids in matic water content, and porphyry Cu 6 Mo 6 Au potential:
silicate melts. Chemical Geology 229, 78–95. examples from the Tethyan arcs of central and eastern Iran
Perelló, J., Carlotto, V., Zarate, A., Ramos, P., Posso, H., Neyra, and western Pakistan. Economic Geology 107, 295–332.
C., Caballero, A., Fuster, N. & Muhr, R. (2003). Porphyry-style Ridolfi, F. & Renzulli, A. (2012). Calcic amphiboles in calc-alka-
alteration and mineralization of the Middle Eocene to Early line and alkaline magmas: thermobarometric and chemo-
Oligocene Andahuaylas–Yauri belt, Cuzco region, Peru. metric empirical equations valid up to 1,130 C and 22 GPa.
Economic Geology 98, 1575–1605. Contributions to Mineralogy and Petrology 163, 877–895.
Pettke, T., Oberli, F. & Heinrich, C. A. (2010). The magma and Rohrlach, B. D. & Loucks, R. R. (2005). Multi-million-year cyclic
metal source of giant porphyry-type ore deposits, based on ramp-up of volatiles in a lower crustal magma reservoir
lead isotope microanalysis of individual fluid inclusions. trapped below the Tampakan copper–gold deposit by Mio-
Earth and Planetary Science Letters 296, 267–277. Pliocene crustal compression in the southern Philippines. In:
Pichavant, M. & Macdonald, R. (2007). Crystallization of primi- Porter, T. M. (ed.) Super Porphyry Copper and Gold
tive basaltic magmas at crustal pressures and genesis of the Deposits: a Global Perspective. PGC Publishing, pp.
calc-alkaline igneous suite: experimental evidence from St 369–407.
Vincent, Lesser Antilles arc. Contributions to Mineralogy Roperch, P., Sempere, T., Macedo, O., Arriagada, C., Fornari,
and Petrology 154, 535–558. M., Tapia, C., Garcı́a, M. & Laj, C. (2006). Counterclockwise
1860 Journal of Petrology, 2015, Vol. 56, No. 9

rotation of late Eocene–Oligocene fore-arc deposits in south- Simon, A. C. & Ripley, E. M. (2011). The role of magmatic sulfur
ern Peru and its significance for oroclinal bending in the cen- in the formation of ore deposits. In: Behrens, H. & Webster,
tral Andes. Tectonics 25, TC3010. J. D. (eds) Sulfur in Magmas and Melts: its Importance for
Roperch, P., Carlotto, V., Ruffet, G. & Fornari, M. (2011). Natural and Technical Processes. Mineralogical Society of
Tectonic rotations and transcurrent deformation south of America and Geochemical Society, Reviews in Mineralogy
the Abancay deflection in the Andes of southern Peru. and Geochemistry 73, 513–578.
Tectonics 30, TC2010. Sisson, T. W. & Grove, T. L. (1993a). Temperatures and H2O
Rowe, M. C., Kent, A. J. R. & Nielsen, R. L. (2009). Subduction in- contents of low-MgO high-alumina basalts. Contributions to
fluence on oxygen fugacity and trace and volatile elements Mineralogy and Petrology 113, 167–184.
in basalts across the Cascade volcanic arc. Journal of Sisson, T. W. & Grove, T. L. (1993b). Experimental investiga-
Petrology 50, 61–91. tions of the role of H2O in calc-alkaline differentiation and
Safran, E. B., Bierman, P. R., Aalto, R., Dunne, T., Whipple, K. X. subduction zone magmatism. Contributions to Mineralogy
& Caffee, M. (2005). Erosion rates driven by channel network and Petrology 113, 143–166.
incision in the Bolivian Andes. Earth Surface Processes and Sisson, T. W., Ratajeski, K., Hankins, W. B. & Glazner, A. F.
Landforms 30, 1007–1024. (2005). Voluminous granitic magmas from common basaltic
Sajona, F. G. & Maury, R. C. (1998). Association of adakites sources. Contributions to Mineralogy and Petrology 148,
with gold and copper mineralization in the Philippines. 635–661.

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Comptes Rendus de l’Académie des Sciences, Série IIA Skewes, M. A. & Stern, C. R. (1994). Tectonic trigger for the for-
326, 27–34. mation of late Miocene Cu-rich breccias in the Andes of cen-
Sandeman, H. A., Clark, A. H. & Farrar, E. (1995). An integrated tral Chile. Geology 22, 551–554.
tectono-magmatic model for the evolution of the southern Solano, J. M. S., Jackson, M. D., Sparks, R. S. J., Blundy, J. D. &
Peruvian Andes (13–20 S) since 55 Ma. International Annen, C. (2012). Melt segregation in deep crustal hot zones:
Geology Review 37, 1039–1073. A mechanism for chemical differentiation, crustal assimila-
Scaillet, B. & Evans, B. W. (1999). The 15 June 1991 eruption of tion and the formation of evolved magmas. Journal of
Mount Pinatubo. I. Phase equilibria and pre-eruption P–T– Petrology 53, 1999–2026.
fO2–fH2O conditions of the dacite magma. Journal of Somoza, R. & Ghidella, M. E. (2012). Late Cretaceous to recent
Petrology 40, 381–411. plate motions in western South America revisited. Earth and
Schmitt, A. K., Stockli, D. F., Lindsay, J. M., Robertson, R., Planetary Science Letters 331–332, 152–163.
Lovera, O. M. & Kislitsyn, R. (2010). Episodic growth and Spera, F. J. & Bohrson, W. A. (2004). Open-system magma
homogenization of plutonic roots in arc volcanoes from chamber evolution: an energy-constrained geochemical
combined U–Th and (U–Th)/He zircon dating. Earth and model incorporating the effects of concurrent eruption, re-
Planetary Science Letters 295, 91–103. charge, variable assimilation and fractional crystallization
Schütte, P., Chiaradia, M., Barra, F., Villagómez, D. & Beate, B. (EC-E’RAvFC). Journal of Petrology 45, 2459–2480.
(2011). Metallogenic features of Miocene porphyry Cu and Stavast, W. J. A. (2006). The fate of magmatic sulfides during in-
porphyry-related mineral deposits in Ecuador revealed by trusion or eruption, Bingham and Tintic districts, Utah.
Re–Os, 40Ar/39Ar, and U–Pb geochronology. Mineralium Economic Geology 101, 329–345.
Deposita 47, 383–410. Stavast, W. J. A., Butler, R. F., Seedorff, E., Barton, M. D. &
Schütte, P., Chiaradia, M. & Beate, B. (2010). Petrogenetic evo- Ferguson, C. A. (2008). Tertiary tilting and dismemberment
lution of arc magmatism associated with late Oligocene to of the Laramide arc and related hydrothermal
late Miocene porphyry-related ore deposits in Ecuador. systems, Sierrita Mountains, Arizona. Economic Geology
Economic Geology 105, 1243–1270. 103, 629–636.
Seedorff, E., Dilles, J. H., Proffett, J. M., Einaudi, M. T., Zurcher, Steinberger, I., Hinks, D., Driesner, T. & Heinrich, C. A. (2013).
L., Stavast, W. J. A., Johnson, D. A. & Barton, M. D. (2005). Source plutons driving porphyry copper ore formation:
Porphyry deposits: characteristics and origin of hypogene combining geomagnetic data, thermal constraints, and
features. Economic Geology, 100th Anniversary Volume, chemical mass balance to quantify the magma chamber be-
251–298. neath the Bingham Canyon deposit. Economic Geology 108,
Seedorff, E., Barton, M. D., Stavast, W. J. A. & Maher, D. J. 605–624.
(2008). Root zones of porphyry systems: extending the por- Stern, C. R., Skewes, M. A. & Arevalo, A. (2011). Magmatic
phyry model to depth. Economic Geology 103, 939–956. evolution of the giant El Teniente Cu–Mo deposit, central
Shafiei, B., Haschke, M. & Shahabpour, J. (2008). Recycling of Chile. Journal of Petrology 52, 1591–1617.
orogenic arc crust triggers porphyry Cu mineralization in Takagi, D., Sato, H. & Nakagawa, M. (2005). Experimental study
Kerman Cenozoic arc rocks, southeastern Iran. Mineralium of a low-alkali tholeiite at 1–5 kbar: optimal condition for the
Deposita 44, 265–283. crystallization of high-An plagioclase in hydrous arc
Shen, P., Shen, Y., Liu, T., Meng, L., Dai, H. & Yang, Y. (2009). tholeiite. Contributions to Mineralogy and Petrology 149,
Geochemical signature of porphyries in the Baogutu por- 527–540.
phyry copper belt, western Junggar, NW China. Gondwana Tatsumi, Y. (1989). Migration of fluid phases and genesis of
Research 16, 227–242. basalt magmas in subduction zones. Journal of Geophysical
Shinohara, H. & Hedenquist, J. W. (1997). Constraints Research 94, 4697–4707.
on magma degassing beneath the Far Southeast Tatsumi, Y., Hamilton, D. L. & Nesbitt, R. W. (1986). Chemical
porphyry Cu–Au deposit, Philippines. Journal of Petrology characteristics of fluid phase released from a subducted
38, 1741–1752. lithosphere and origin of arc magmas: Evidence from high-
Sillitoe, R. H. (2010). Porphyry copper systems. Economic pressure experiments and natural rocks. Journal of
Geology 105, 3–41. Volcanology and Geothermal Research 29, 293–309.
Simakin, A. G. & Bindeman, I. N. (2012). Remelting in caldera Thiéblemont, D., Stein, G. & Lescuyer, J.-L. (1997). Gisements
and rift environments and the genesis of hot, ‘recycled’ épithermaux et porphyriques: la connexion adakite.
rhyolites. Earth and Planetary Science Letters 337–338, Comptes Rendus de l’Académie des Sciences, Série IIA 325,
224–235. 103–109.
Journal of Petrology, 2015, Vol. 56, No. 9 1861

Thompson, A. B., Matile, L. & Ulmer, P. (2002). Some thermal crustal magma cycle: thermobarometry of the Aucanquilcha
constraints on crustal assimilation during fractionation of Volcanic Cluster. Contributions to Mineralogy and Petrology
hydrous, mantle-derived magmas with examples from cen- 165, 663–682.
tral Alpine batholiths. Journal of Petrology 43, 403–422. Wilkinson, J. J. (2013). Triggers for the formation of porphyry
Tosdal, R. M. & Richards, J. P. (2001). Magmatic and structural ore deposits in magmatic arcs. Nature Geoscience 6,
controls on the development of porphyry Cu 6 Mo 6 Au de- 917–925.
posits. In: Richards, J. P. & Tosdal, R. M. (eds) Structural Wones, D. R. (1989). Significance of the assemblage titan-
Controls on Ore Genesis. Reviews in Economic Geology, ite þ magnetite þ quartz in granitic rocks. American
157–181. Mineralogist 74, 744–749.
Ulianov, A., Müntener, O., Schaltegger, U. & Bussy, F. (2012). Workman, R. K. & Hart, S. R. (2005). Major and trace element
The data treatment dependent variability of U–Pb zircon composition of the depleted MORB mantle (DMM). Earth
ages obtained using mono-collector, sector field, laser abla- and Planetary Science Letters 231, 53–72.
tion ICPMS. Journal of Analytical Atomic Spectrometry 27, Xirouchakis, D. & Lindsley, D. H. (1998). Equilibria among
663–676. titanite, hedenbergite, fayalite, quartz, ilmenite, and mag-
Vigneresse, J.-L. (2007). The role of discontinuous magma in- netite: Experiments and internally consistent thermo-
puts in felsic magma and ore generation. Ore Geology dynamic data for titanite. American Mineralogist 83,
Reviews 30, 181–216. 712–725.

Downloaded from http://petrology.oxfordjournals.org/ at Université de Genève on October 31, 2015


Walker, B. A., Klemetti, E. W., Grunder, A. L., Dilles, J. H., Zellmer, G. F., Iizuka, Y., Miyoshi, M., Tamura, Y. & Tatsumi, Y.
Tepley, F. J. & Giles, D. (2013). Crystal reaming during the (2012). Lower crustal H2O controls on the formation of ada-
assembly, maturation, and waning of an eleven-million-year kitic melts. Geology 40, 487–490.

View publication stats

You might also like