You are on page 1of 78

Journal Pre-proof

Arc magmatic evolution and porphyry copper deposit formation


under compressional regime: A geochemical perspective from the
Toquepala arc in Southern Peru

Nian Chen, Jingwen Mao, Zhaochong Zhang, Zheng Duan, Alan


Santos, Hongying Li

PII: S0012-8252(23)00072-7
DOI: https://doi.org/10.1016/j.earscirev.2023.104383
Reference: EARTH 104383

To appear in: Earth-Science Reviews

Received date: 1 December 2022


Revised date: 21 February 2023
Accepted date: 6 March 2023

Please cite this article as: N. Chen, J. Mao, Z. Zhang, et al., Arc magmatic evolution
and porphyry copper deposit formation under compressional regime: A geochemical
perspective from the Toquepala arc in Southern Peru, Earth-Science Reviews (2023),
https://doi.org/10.1016/j.earscirev.2023.104383

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2023 Published by Elsevier B.V.


Journal Pre-proof

Arc Magmatic Evolution and Porphyry Copper Deposit

Formation under Compressional Regime: A Geochemical

Perspective from the Toquepala Arc in Southern Peru

Nian Chena, Jingwen Maoa,b,*, Zhaochong Zhangc, Zheng Duand,e, Alan Santosf,

Hongying Lib

of
a
MNR Key Laboratory for Exploration Theory & Technology of Critical Mineral

ro
Resources, China University of Geosciences, Beijing 100083, China
-p
b
MNR Key Laboratory of Metallogeny and Mineral Assessment, Institute of Mineral
re

Resources, Chinese Academy of Geological Sciences (CAGS), Beijing 100037, China


lP

c
na

State Key Laboratory of Geological Processes and Mineral Resources, China

University of Geosciences, Beijing 100083, China


ur
Jo

d
Nanjing Center of China Geological Survey, Nanjing 210016, China

e
Chinese Academy of Geological Sciences (CAGS), Beijing 100037, China

f
Faculty of Earth Resources, China University of Geosciences, Wuhan 430074, China

*
Corresponding author: Jingwen Mao, email, jingwenmao@263.net
Journal Pre-proof

Abstract

The Paleocene–Eocene southern Peru metallogenic belt contains numerous very large

to supergiant porphyry Cu‒Mo deposits. The deposits were dated previously as coeval

with the Incaic I orogeny at ca. 60 Ma. However, tectono-magmatic processes during

the formation of the deposits are poorly constrained. Here, we integrate published

geochronological, geochemical, and mineralogical data from barren and ore-related

of
granitoids in the region. The syn-orogenic unit (60–53 Ma) has significantly higher

ro
average ratios of Sr/Y (103.5), La/Yb (29.5), Dy/Yb (1.9), and EuN/Eu* (1.07) than
-p
the pre-orogenic unit (69–60 Ma; Sr/Y = 18.6, La/Yb = 11.5, Dy/Yb = 1.7, and
re

EuN/Eu* = 0.66). The syn-orogenic unit exhibits high Dy/Yb ratios in relatively
lP

primitive rocks and has a lower average Dy/Yb(zircon) ratio (0.23) and higher
na

EuN/Eu*(zircon) ratio (0.43) than the pre-orogenic unit (Dy/Yb(zircon) = 0.29 and
ur

EuN/Eu*(zircon) = 0.21). These integrated data suggest that the syn-orogenic magmas

underwent high-pressure, garnet-dominated fractionation prior to


Jo

amphibole-dominated fractionation, distinct from the pre-orogenic magmas which

underwent low-pressure, plagioclase-dominated fractionation. Comprehensive

investigation of the geological background suggests that the geochemical variations

between pre- and syn-orogenic units were caused by the long-term evolution of

mantle-derived magmas at the crust‒mantle boundary during arc compression, instead

of crustal thickening. Fractionation of ferric iron-depleted minerals (e.g., garnet and


Journal Pre-proof

amphibole) in the mantle-derived basaltic magmas probably has positive effects on

magmatic fertility. The dataset presented highlights the critical role of high-pressure

differentiation, thus providing an insight into the relationship between the arc

compressional regime and giant porphyry Cu deposit formation in the Andes. Zircon

trace elements can be useful indicators for porphyry deposit exploration in the

Toquepala arc in southern Peru. However, the regional tectonic settings of the Andean

of
arcs should be studied thoroughly before applying these indicators.

ro
Keywords: arc compression; magmatic evolution; oxygen fugacity; magmatic fertility;
-p
porphyry deposit
re
lP

1. Introduction
na

The Andes host the most important porphyry Cu‒Mo‒Au ore belts in the world (Mao
ur

et al., 2018; Sillitoe and Perelló, 2005). The deposits in these belts were formed

mainly during the Cenozoic, although some smaller Jurassic‒Cretaceous porphyry


Jo

deposits have been discovered (Clark et al., 1990; Drobe et al., 2013; Maksaev et al.,

2010; Richards et al., 2017; Santos et al., 2019; Sillitoe and Perelló, 2005). Porphyry

Cu deposits tend to occur in clusters along belts in close alignment with regional

faults (Noble et al., 2004; Sillitoe and Perelló, 2005), and they are generally

associated with shallow and multiphase porphyry intrusions (Chen and Wu, 2020;

Sillitoe, 2010). With respect to geochemistry, most mineralized porphyries exhibit


Journal Pre-proof

strong adakitic affinities (Richards and Kerrich, 2007) with relatively high SiO2 (≥57

wt.%) and Al2O3 (≥15 wt.%), high ratios of Sr/Y (>50) and La/Yb (>20), slightly to

no negative Eu anomalies, and steep to listric-shaped patterns of rare earth elements

(REE) (Bourdon et al., 2002; Hollings et al., 2005; Reich et al., 2003). Some studies

have suggested that oxidized magmas (fayalite-magnetite-quartz buffer, ΔFMQ >+0.5)

with high volatiles contents (e.g., S6+, Cl, and H2O) favor porphyry Cu deposit

of
formation (Audétat and Simon, 2012; Chambefort et al., 2008; Chambefort et al.,

ro
2013; Chelle-Michou and Chiaradia, 2017; Meng et al., 2021a; Munoz et al., 2012;
-p
Rezeau and Jagoutz, 2020; Sun et al., 2015; Wang et al., 2014a,b; Wu et al., 2021;
re
Zhu et al., 2018). An auto-oxidation model that builds upon garnet fractionation has
lP

been proposed (Lee and Tang, 2020). Since garnet preferentially incorporates Fe2+, its

fractionation elevates progressively the oxidation state of residual melt. A similar


na

model based on amphibole fractionation has also been proposed (Zhang et al., 2022).
ur

However, it remains unclear how these fractionation processes are connected to


Jo

specific tectono-magmatic setting during the generation of ore-forming arc magmas

(Green, 1970; Loucks, 2021; Park et al., 2021; Wang et al., 2014a).

Although many arc magmas are oxidized and have high volatile contents

(Audétat and Simon, 2012; Cao et al., 2021; Chambefort et al., 2008; Chen et al.,

2022a; Chen et al., 2022b; Meng et al., 2021a; Richards, 2011; Richards et al., 2012;

Richards, 2015; Wang et al., 2014a; Wang et al., 2014b), porphyry Cu deposit

formation in magmatic arcs is quite rare and occurs in discrete events (Cooke et al.,
Journal Pre-proof

2005; Richards, 2013; Simmons et al., 2013). In the Andes, various metallogenic

mechanisms have been proposed to explain the formation of the Cenozoic porphyry

Cu deposits, including buoyant slab subduction of topographic anomalies (oceanic

plateau, oceanic ridge, and island chain): (1) buoyant slab subduction increases the

coupling between subducting and overriding plates, thus increasing the horizontal

compression and prolonging the evolution and maturation of mantle-derived magmas

of
in the mid‒lower crust (Chen et al., 2022b; Chiaradia, 2009; Chiaradia et al., 2009b;

ro
Loucks, 2021); (2) higher horizontal compression leads to horizontal shortening and
-p
crustal thickening, which are responsible for the transition of the field in the magma
re
source region from the plagioclase-stable to the amphibole ± garnet-stable (Haschke
lP

et al., 2002; Hollings et al., 2005); (3) stronger coupling increases subduction erosion

and sediment input into the mantle wedge, thus promoting the generation of fertile
na

magmas with high oxidation states and volatile contents (Bissig et al., 2003; Hollings
ur

et al., 2005; Kay et al., 2005); (4) tectonic disturbance of deeper magma chambers
Jo

causes fertile magma recharge (Blundy et al., 2015; Richards, 2018); and (5) buoyant

flat-slab subduction reduces the amount of heat transferred from the mantle wedge to

the magma source, and the reduced magma supply causes magma cooling and fluid

exsolution (Skewes and Stern, 1994; Stern and Skewes, 2005). All the above

processes may contribute to the formation of giant porphyry Cu deposits, and in

similar or different arc segments in the Andes, they are not necessarily mutually

exclusive.
Journal Pre-proof

Here, we tested these possible mechanisms in the Paleocene‒Eocene southern

Peru metallogenic belt (Fig. 1), which contains numerous very large to supergiant

porphyry Cu‒Mo deposits (Table 1, based on the criterion of Clark, 1993). These

deposits include Toquepala, Cerro Verde‒Santa Rosa, Quellaveco, Cuajone, Los

Calatos, and Don Javier (Martínez et al., 2017; Shatwell, 2021; Simmons, 2013;

Webster et al., 2013; Ye et al., 2022; Zweng et al., 1995), which contain a total of

of
about 60 million metric tons of copper metal resources (Acosta et al., 2020; Shatwell,

ro
2021). Based on previous geochronological studies, these deposits were formed
-p
coevally with the Incaic I orogeny (Noble et al., 1979; Noble et al., 1985; Roperch et
re
al., 2011), which was speculatively caused by the subduction of the San Felix Ridge at
lP

ca. 60 Ma (Fig. 1A) (Noury et al., 2017). To reveal the tectono-magmatic processes

that caused the formation of these deposits, we compiled published geochronological,


na

geochemical, and mineralogical data, including whole-rock elements, whole-rock Sr‒


ur

Nd isotopes, and zircon trace elements. This work helps us to understand the
Jo

magmatic response to a tectonic event, and its relationship with the formation of giant

porphyry Cu deposits. Moreover, this study will be beneficial for mineral exploration

in the Andes.

2. Tectono-magmatic setting

The oldest strata exposed in the study area belong to the metamorphic Arequipa

Massif (Fig. 1C) (Coira et al., 1982), which is considered to be an exotic terrane that
Journal Pre-proof

accreted to the western Gondwana margin during the Mesoproterozoic (Ramos, 2008,

2010). Steep subduction may have caused significant intra-arc and back-arc rifts in

the western Gondwana margin (Simmons, 2013), including the Arequipa back-arc

basin in southern Peru during the Jurassic (Boekhout et al., 2012; Sempere et al.,

2002). This basin was filled by a succession of >4.5-km-thick volcanic/volcaniclastic,

deep-/shallow-marine, deltaic, and locally continental sediments (Boekhout et al.,

of
2012; Manrique, 2011; Sempere et al., 2002), which were cut by the progressively

ro
eastward migrating Coastal Batholith plutons (190–95 Ma) (Manrique, 2011; Mukasa,
-p
1986; Simmons, 2013). The Mesozoic sediments are overlain unconformably by
re
continental and volcaniclastic sediments of the Toquepala Group, which
lP

are >1.5-km-thick (75–55 Ma) (Fig. 2). Within this group, unconformities also

developed locally, which were formed likely by the Incaic I orogeny at ca. 60 Ma in
na

southern Peru (Noble et al., 1979; Noble et al., 1985; Roperch et al., 2011).
ur

Magmatism during the period was represented mainly by the composite Coastal
Jo

Batholith, and formed the Toquepala arc (91–45 Ma) (Fig. 1B) (Demouy et al., 2012;

Mamani et al., 2010; Manrique, 2011). Many of the very large to supergiant porphyry

deposits in this arc are associated with Late Paleocene‒Early Eocene porphyry

intrusions (60–53 Ma) (Fig. 1C) (Clark et al., 1990; Simmons, 2013). These deposits

occur in belts along the Cincha‒Lluta‒Incapuquio fault system (SFCLLI) and its

secondary faults (Fig. 1C) (Mamani et al., 2011). In the Eocene‒Oligocene (52–30

Ma), southern Peru was dominated by a flat-slab subduction regime (James and Sacks,
Journal Pre-proof

1999), which caused the northwestward magmatic migration and widening of the

Toquepala arc (Fig. 2C) (Mamani et al., 2010). The Incaic II orogenic event (ca. 45

Ma) caused regional compression and uplift (Noble and Wise, 2016), and the

associated magmatism formed the Andahuaylas‒Anta arc in southern Peru (45–30 Ma)

(Fig. 1B) (Chelle-Michou et al., 2015; Perelló et al., 2003). Subsequently, back-arc

crustal shortening and uplift associated with several pulses of orogenic events

of
occurred, and these are recorded in several regional unconformities (Fig. 2A)

ro
(Chapman et al., 2015; MéGard, 1987; Noble et al., 1985; Roperch et al., 2011;

Sandeman et al., 1995).


-p
re

3. Data compilation
lP

To study the tectono‒magmatic processes of the Toquepala arc in southern Peru at


na

around 60 Ma, geochronology, whole-rock elements, and zircon trace elements from
ur

regionally distributed igneous rocks were compiled. The Jurassic‒Quaternary Sr‒Nd


Jo

isotopes of igneous rocks from the study area were also compiled as reference.

The whole-rock K‒Ar/Ar‒Ar isotopic systems have low closure temperatures and

are susceptible to thermal disturbance and hydrothermal activity (Chiaradia et al.,

2013; Duan et al., 2022); thus, we selected only zircon U‒Pb ages to better constrains

the timelines of plutonic and porphyry magmatism and the magmatic lulls between

them. Muscovite K‒Ar/Ar‒Ar or molybdenite Re‒Os isotopic dating was used to


Journal Pre-proof

determine the mineralization ages. After 17 studies were screened, a total of 118

geochronological data were compiled (Fig. 3) (Bellon and Lefevre, 1976; Chen et al.,

2022a; Clark et al., 1990; Demouy et al., 2012; Martínez and Cervantes, 2003;

Mukasa and Tilton, 1985; Mukasa, 1986; Nathwani et al., 2021; Noury et al., 2017;

Quang et al., 2003; Sillitoe and Mortensen, 2010; Simmons, 2013; Simmons et al.,

2013; Stegen et al., 2018; Stewart et al., 1974; Ye et al., 2022; Zweng et al., 1995).

of
To track the magmatic evolution in the time interval 69–53 Ma, whole-rock

ro
elemental data were filtered using the following criteria: (1) loss on ignition (LOI)
-p
<3.5 wt.% to minimize the influence of alteration and (2) EuN/Eu* <1.3 or Al2O3 <20
re
wt.% to exclude cumulates (Wu et al., 2021, 2022). A dataset of major and trace
lP

elements on 130 whole-rock samples was obtained from five studies (Chen et al.,
na

2022a; Demouy et al., 2019; Nathwani et al., 2021; Simmons 2013; Valencia et al.,

2020).
ur
Jo

The whole-rock Sr‒Nd isotopic systems may be susceptible to modification by

hydrothermal fluids (Davies, 2002); thus, we screened the data using more strict

criteria for analytical totals (volatile-free basis) of >98.5 wt.% and LOI values of <1.5

wt.%. A total of 224 isotopic data from various studies passed the criteria (Boekhout

et al., 2012; Chen et al., 2022a; Delacour et al., 2007; Demouy et al., 2019; Gerbe and

Thouret, 2004; Mamani et al., 2008; Rivera et al., 2017; Ruprecht and Wörner, 2007;

Sørensen and Holm, 2008; Thouret et al., 2005).


Journal Pre-proof

Regarding zircon trace elements, spot analyses that may indicate contamination

by mineral inclusions (e.g., apatite and titanite; i.e., La >1 ppm and Ti >50 ppm) were

excluded (Lu et al., 2016; Meng et al., 2021b; Zhu et al., 2018). A total of 30 zircon

trace-element samples from three studies were selected (Simmons, 2013; Nathwani et

al., 2021; Chen et al., 2022a).

4. Results

of
ro
4.1. Regional geochronology
-p
The intrusions of the Toquepala arc are conventionally grouped into superunits or
re

batholiths (Linga, Tiabaya, Toquepala, and Yarabamba) (Demouy et al., 2019;


lP

Manrique, 2011; Mukasa, 1986). The youngest of these intrusions are found in the
na

Yarabamba Superunit, which hosts the porphyry complexes of the Cerro Verde‒Santa

Rosa, Don Javier, Cuajone, Quellaveco, and Toquepala deposits (Manrique, 2011;
ur

Sillitoe and Mortensen, 2010; Simmons, 2013). Based on the compiled data, the
Jo

plutons generally predate 60 Ma, but some were emplaced at 59.0 ± 0.2 Ma, near the

porphyry deposits (Fig. 3) (Nathwani et al., 2021).

In Cerro Verde‒Santa Rosa (Fig. 3), the youngest intrusions of the Yarabamba

Superunit were dated at ca. 60.7 Ma (Mukasa, 1986). The mineralization is associated

with granodiorite porphyry stocks (61 ± 1 to 60.9 ± 0.6 Ma) (Mukasa, 1986; Simmons,

40
2013; Stegen et al., 2018). Hydrothermal sericite was dated by the Ar/39Ar method
Journal Pre-proof

from 62.0 ± 1.1 to 61.0 ± 1.2 Ma (Quang et al., 2003). The late-mineralization

granodiorite porphyries have zircon U‒Pb ages ranging from 59.8 ± 0.6 to 58.5 ± 0.5

Ma (Stegen et al., 2018).

In Don Javier, the youngest intrusions of the Yarabamba Superunit were dated

from 65.4 ± 0.7 to 63.5 ± 0.8 Ma (Chen et al., 2022a; Ye et al., 2022). The ore-related

dacite porphyry was dated from 59.2 ± 1.1 to 59.9 ± 0.4 Ma (Chen et al., 2022a; Ye et

of
al., 2022). The latter overlaps within uncertainty with the mineralization age (60.4 ±

ro
1.4 Ma) determined by molybdenite Re‒Os dating, which identified also a younger
-p
mineralization event at ca. 45 Ma (Ye et al., 2022).
re
lP

In Cuajone, the youngest plutonic intrusion was dated at 64.2 ± 0.9 Ma (Simmons,

2013; Simmons et al., 2013). The mineralization is primarily associated with latite
na

porphyry 1 (55.6 ± 0.6 Ma), which hosts the bulk of the Cu‒Mo resources. The
ur

younger latite porphyry 2 (56.2 ± 0.7 Ma) hosts minor Cu‒Mo resources, and it was
Jo

intruded by andesite dikes at 56.2 ± 0.5 Ma. The youngest latite porphyry 3 (53.5 ±

0.6 Ma) is generally altered and not mineralized (Simmons et al., 2013).

Hydrothermal sericite is slightly younger and was dated from 52.4 ± 1.7 to 51.0 ± 1.6

Ma by the 40Ar/39Ar method (Clark et al., 1990; Estrada, 1978).

In Quellaveco, zircon U‒Pb dating showed that the youngest plutonic rocks of the

Yarabamba Superunit emplaced at 59.0 ± 0.2 Ma (Nathwani et al., 2021). Regarding

the porphyry complexes, at least seven intrusive porphyry phases have been identified,
Journal Pre-proof

i.e., from the oldest to the youngest: granodiorite porphyry (57.8 ± 0.8 to 58.2 ± 0.8

Ma), early porphyry (54.4 ± 0.21 to 58.41 ± 0.53 Ma), intermineralization porphyry 1

(54.88 ± 0.2 to 57.2 ± 0.9 Ma), intermineralization porphyry 2 (54.19 ± 0.19 Ma), late

intermineralization monzodiorite (55.26 ± 0.28 to 57.2 ± 0.7 Ma), late porphyry (53.7

± 0.9 to 56 ± 0.8 Ma), and post-mineralization dacite dikes (55.1 ± 0.6 to 57.4 ± 0.9

Ma) (Nathwani et al., 2021; Sillitoe and Mortensen, 2010; Simmons, 2013). The

of
oldest granodiorite porphyry was revealed also as xenoliths in the later intrusions. The

ro
early porphyry, intermineralization porphyries, and late intermineralization
-p
monzodiorite are spatially associated with the bulk of the Cu‒Mo resources.
re
40
Hydrothermal sericite Ar/39Ar dating of mineralized porphyries showed that the
lP

mineralization occurred from 54.5 to 54.7 Ma (Clark, 2003; Simmons, 2013). The late

porphyry and post-mineralization dacite dikes do not contain significant economic


na

Cu‒Mo resources.
ur

In Toquepala, zircon U‒Pb dating showed that the youngest part of the
Jo

Yarabamba Superunit was emplaced at 59.8 ± 0.9 Ma (Simmons, 2013). The main

mineralization at this deposit is associated with the 56.8 ± 0.6 Ma dacite porphyry

(Simmons, 2013), which is consistent with hydrothermal muscovite K‒Ar ages (56.0

± 1.0 to 52.3 ± 1.6 Ma) (Clark et al., 1990; Zweng et al., 1995). Late-stage barren

dacite agglomerates and latite porphyry dikes have zircon U‒Pb ages from 56.2 ± 0.6

to 54.3 ± 0.6 Ma (Simmons, 2013).


Journal Pre-proof

Overall, the geochronological data suggest that the plutonic magmatism of the

Yarabamba Superunit waned at 60‒59 Ma, and was followed by the emplacement of

the porphyry complexes. The magmatic lull between plutons and porphyries in the

districts of Don Javier and Cuajone lasted for 3‒8 Myr; however, only a very small

gap can be observed in the districts of Cerro Verde‒Santa Rosa, Quellaveco, and

Toquepala districts (Fig. 3). The mineralization in the Paleocene‒Eocene southern

of
Peru metallogenic belt occurred likely from northwest to southeast in diachronous

ro
intervals of 4‒6 Myr (e.g., in Cerro Verde‒Santa Rosa and Toquepala at ca. 61 and ca.
-p
56 Ma, respectively) (Fig. 1C). The youngest magmatic-hydrothermal activity
re
persisted probably to 45‒46 Ma, as suggested by the molybdenite Re‒Os dates of Don
lP

Javier (44.5 ± 1.4 to 44.4 ± 1.6 Ma) (Ye et al., 2022) and the whole-rock K‒Ar dates

of Quellaveco (47.0 ± 1.3 to 45.9 ± 0.3 Ma) (Zimmermann and Collado, 1983). The
na

persistence of magmatism is also supported by the data obtained from the


ur

southernmost area of Peru (Clark et al., 1990; Martínez and Cervantes, 2003), where
Jo

the Toquepala and Andahuaylas‒Anta arcs overlap (Fig. 1B).

4.2. Whole-rock geochemistry

The compiled samples range from low-K calc-alkaline to shoshonitic (Fig. 4). They

are enriched in large-ion lithophile elements (LILEs, e.g., Rb, Ba, Th, and K) and

depleted in high-field-strength elements (HFSEs, e.g., Nb and Ta), respectively. They

also have negative Ti anomalies (Fig. 5), which are typical characteristics of
Journal Pre-proof

subduction-related magmas. The igneous rock samples could be divided into two units

that are distinct in age and geochemical composition: (1) Late Cretaceous‒Early

Paleocene (69 to 60 Ma) pre-orogenic unit emplaced before the Incaic I orogeny and

(2) Late Paleocene‒Early Eocene (60 to 53 Ma) syn-orogenic unit emplaced during

the orogeny and normal-angle subduction (Fig. 2). However, some of the igneous

rocks were probably emplaced during the subsequent flat-slab subduction. They are

of
relatively young, and most are distantly related to the main mineralization, e.g., late-

ro
to post-mineralization porphyries in Cuajone, Quellaveco, and Toquepala. These
-p
young igneous rocks were grouped into the syn-orogenic unit. The geochemical data
re
compiled from the pre-orogenic unit relate mainly to diorite, (quartz-)monzonite, and
lP

granodiorite, whereas the syn-orogenic unit has mainly granodiorite to granite

compositions.
na

In chondrite-normalized REE patterns (Fig. 5), the pre-orogenic unit exhibits a


ur

flatter pattern from the middle to heavy REEs compared to the syn-orogenic unit. In
Jo

these regional igneous rocks, the EuN/Eu* (EuN/(SmN × GdN)˄0.5) ratios increase

generally with decreasing age (Fig. 6). The pre-orogenic unit exhibits significant

negative Eu anomalies (average EuN/Eu* = 0.66, n = 48), different to the syn-orogenic

unit (average EuN/Eu* = 1.07, n = 82).

To evaluate the REE patterns, we plotted the Sr/Y, La/Yb, and Dy/Yb ratios

against intrusive ages (Fig. 6). Although the ratios generally show gradual increases

with decreasing age, their rates of increase rise abruptly at ca. 60 Ma. For the
Journal Pre-proof

pre-orogenic unit, the average Sr/Y, La/Yb, and Dy/Yb ratios are 18.6 (n = 48), 11.5

(n = 48), and 1.7 (n = 48), respectively. In contrast, the syn-orogenic unit has average

Sr/Y, La/Yb, and Dy/Yb ratios of 103.5 (n = 82), 29.5 (n = 79), and 1.9 (n = 79),

respectively.

Igneous rocks from both units also show fractionation trends in their Dy/Yb vs.

SiO2 diagrams (Fig. 7). The pre-orogenic unit shows constant Dy/Yb ratios with

of
increasing SiO2 contents, whereas the syn-orogenic unit shows higher Dy/Yb ratios in

ro
relatively primitive rocks (Dy/Yb ≈ 2.5 at 65 wt.% SiO2); then, the ratios decrease
-p
gradually with increasing SiO2 contents (Dy/Yb ≈ 1.5 at 70 wt.% SiO2).
re
4.3. Whole-rock Sr‒Nd isotopes
lP

Given the limited amount of whole-rock Sr‒Nd isotopic data (Fig. 8), we combined
na

the pre- and syn-orogenic units. To determine the magma sources and crustal
ur

contamination, we followed previous methods (Chen et al., 2022a) for choosing the
Jo

rear-arc sample 126175 (37.3°S, Chile), as representative of the trace-element

composition of a mantle-derived magma (MDM) with both minimal/no subducted

slab nor crustal contamination (Søager et al., 2013; Søager and Holm, 2013). The

isotopic composition of the MDM and baseline ambient mantle (BAM), representing

no contribution from the subducted slab, was taken equal to sample 126175 (Søager

and Holm, 2013; Søager et al., 2013). The trace-element composition of BAM was

taken from the average depleted mid-ocean ridge basalt mantle (Workman and Hart,
Journal Pre-proof

2005). Most of the Jurassic to Paleogene samples were found to fall on the mixing

array between MDM and metamorphic basement rocks, and their total crustal

contaminations are in the range of 5%–20%. The Neogene‒Quaternary igneous rocks

show more crustal contamination (15%–35%), while some Jurassic‒Cretaceous

igneous rocks were found to fall on the mixing array between the BAM and

subducting sediment.

of
4.4. Zircon trace elements

ro
-p
The zircons of pre- and syn-orogenic units show distinct trace-element characteristics

in their EuN/Eu*(zircon) and Dy/Yb(zircon) ratios (Fig. 9). The pre-orogenic unit shows
re

average EuN/Eu*(zircon) (0.21, n = 202) ratios, lower than those of the syn-orogenic unit
lP

(0.43, n = 888). However, the pre-orogenic unit shows a higher average Dy/Yb(zircon)
na

ratio (0.29, n = 201) than the syn-orogenic unit (0.23, n = 888) (Fig. 7). In the
ur

Dy/Yb(zircon) vs. Hf diagram (Fig. 9), both pre- and syn-orogenic units show
Jo

decreasing Dy/Yb(zircon) ratios with increasing Hf contents. In the EuN/Eu*(zircon) vs. Hf

diagram, both pre- and syn-orogenic units show decreasing EuN/Eu*(zircon) ratios with

increasing Hf contents. In the Hf vs. Ti diagrams, a positive relationship was found

between Hf and Ti contents in both pre- and syn-orogenic units (Fig. 9E).

The temperature-susceptible CeN/Ce*(zircon) ratios are unreliable to evaluate the

magmatic oxidation state (Fig. 9) (Loucks et al., 2018; Loader et al., (2022); and Zou

et al., 2019). Different tectonic settings can have similar impacts on the Ce/U ratios of
Journal Pre-proof

melts during magmatic differentiation, but the zircon‒melt partition coefficients of Ce

and U are strongly affected by the magmatic oxidation state (Loucks et al., 2020). The

Ti content in zircon can be used as an index to differentiate parental magmas (Hayden

and Watson, 2007); thus, we employed the Ce/U(initial) and U(initial)/Ti ratios to evaluate

the magmatic oxidation state according to the method of Loucks et al. (2020). Initial

U content of zircon was calibrated according to the zircon U‒Pb age. The results

of
show that oxygen fugacity (fO2) values increase with decreasing emplacement age, i.e.

ro
increasing from ΔFMQ = −0.3 at 67 Ma to ΔFMQ = +1.9 at 54 Ma (Fig. 10A).
-p
5. Discussion
re
lP

5.1. Igneous geochemistry reveals regional tectonic transition


na

While both the pre- and syn-orogenic units exhibit typical subduction-related

geochemical features, the syn-orogenic unit is more evolved and has higher average
ur

SiO2 content than the pre-orogenic unit (Fig. 4). The pre-orogenic unit shows
Jo

relatively low Sr/Y ratios (average 18.6) and negative Eu anomalies (average EuN/Eu*

= 0.66) (Fig. 6). In contrast, the syn-orogenic unit shows strong adakitic affinity

(Richards, 2011), with average Sr/Y ratios of 103.5 and weak/no negative Eu

anomalies (average EuN/Eu* = 1.07; Fig. 6). Moreover, similar to some mature

magmatic arcs, the syn-orogenic unit exhibits a trend of decreasing Dy/Yb ratios with

increasing SiO2 content (Fig.7), implying amphibole-dominated fractionation

(Davidson et al., 2007). However, the Dy/Yb ratios cannot be projected back to the
Journal Pre-proof

field of mantle-derived basaltic arc magma, implying participation of slab-derived

components or presence of garnet in cumulates during the early stage of

differentiation (Davidson et al., 2007; Defant and Drummond, 1990; Errázuriz-Henao

et al., 2019; Gómez-Tuena et al., 2014; Lee and Tang, 2020; Richards, 2011; Zellmer

et al., 2012).

5.1.1. The role of slab-derived components

of
ro
Owing to the uniformity of the zircon Hf–O isotopic data between pre- and
-p
syn-orogenic units (Appendix A), the slab-derived melt model is not able to explain

the origin of high Sr/Y and Dy/Yb signatures during the orogeny (Chen et al., 2022a;
re

Chen et al., 2022b; Munoz et al., 2012). This is because participation of additional
lP

slab-derived melt would have modified the isotopic characteristics (Munoz et al.,
na

2012). In addition, the whole-rock Sr‒Nd isotope data indicate that no additional
ur

subducting sediment was integrated into the magmatic source during the Paleogene in
Jo

southern Peru (Fig. 8).

5.1.2. The role of crustal thickening

Crustal thickness has been suggested as a critical factor that controls magmatic

differentiation (Farner and Lee, 2017; Chiaradia, 2015; Lee and Tang, 2020; Park et

al., 2021; Tang et al., 2020). The high average ratios of La/Yb (29.6) and Dy/Yb (1.9)

in the syn-orogenic rocks are comparable with those of many igneous rocks from
Journal Pre-proof

thickened crust (Fig. 1), e.g., the Middle‒Late Miocene igneous rocks in northern

Peru (5–10°S, average La/Yb ≈ 28 and Dy/Yb ≈ 2.1) (Chen et al., 2022b), Quaternary

igneous rocks in northernmost Chile (SP-L, ~22°S, average La/Yb ≈ 22 and Dy/Yb ≈

2.1) (Godoy et al., 2014), Pliocene igneous rocks in central Chile (PCC, 32–34°S,

average La/Yb ≈ 24 and Dy/Yb ≈ 1.9) (Hollings et al., 2005), and the nearby Misti

volcanic rocks (average La/Yb ≈ 30 and Dy/Yb ≈ 2.2) (Rivera et al., 2017).

of
Crustal thickening is assumed to contribute to the increase in the ratios of La/Yb

ro
and Dy/Yb in syn-orogenic igneous rocks; thus, we used the La/Yb ratios to estimate
-p
roughly the crustal thickness (Haschke et al., 2002; Haschke and Gunther, 2003). The
re
estimated thickness values are 30–35 and 40‒55 km for the pre- and syn-orogenic
lP

crusts, respectively (Fig. 6C). If the high ratios of La/Yb and Dy/Yb in the
na

syn-orogenic unit originated fully from crustal thickening, the estimated values

suggest the presence of an abrupt 10- to 20-km vertical crustal growth during this
ur

rapid tectonic transition period.


Jo

Although the subduction of topographic anomalies may have led to arc

compression that caused crustal shortening and thickening (Espurt et al., 2008), as

suggested by the local unconformity found within the Toquepala Group (Fig. 2)

(Manrique, 2011), crustal shortening (deformation) was not found in most places in

this region before Miocene (Demouy et al., 2012; Mamani et al., 2010; Noble et al.,

1985; Noury et al., 2017). The low degree of crustal contamination (5%–20%) in the
Journal Pre-proof

Paleogene rocks indicates also the occurrence of relatively little crustal shortening

during this period (Fig. 8), consistent with the conclusion of previous studies

(Mamani et al., 2010; Demouy et al., 2012). Thus, the crustal thickening cannot be

explained by shortening alone. If a 10- to 20-km vertical crustal growth occurred, the

Incaic I orogeny likely influenced the thickening by increasing the rate of

mantle-derived basalt underplating. However, the underplating rate would be reduced

of
because the subduction of the topographic anomalies would have lowered the degree

ro
of partial melting in the mantle wedge through the following mechanisms: (1)
-p
shortening the length of the melting column (Plank and Langmuir, 1988; Turner and
re
Langmuir, 2015), (2) constraining the corner flow (Farner and Lee, 2017), (3)
lP

increasing the pressure, and (4) lowering the temperature. The latter is related to the

reduced subduction rate caused by the enhanced coupling between oceanic and
na

continent plates (England and Katz, 2010; Hall, 2012; Perrin et al., 2018), although
ur

some works have argued that the subduction rate in southern Peru at 60‒50 Ma was
Jo

slightly higher than that at 70‒60 Ma (Pardo-Casas and Molnar, 1987).

Consequently, although arc compression could have caused crustal thickening, it

is likely not the reason for the abrupt increase in the ratios of La/Yb and Dy/Yb in the

syn-orogenic unit at ca. 60 Ma.

5.1.3. The role of lower-crustal differentiation

Some studies have argued that arc compression will force the long-term residence of
Journal Pre-proof

mantle-derived magma at the crust‒mantle boundary (Chen et al., 2022b; Chiaradia et

al., 2009b; Loucks, 2021), contributing potentially to magmatic evolution and

increasing the possibility for porphyry Cu deposit formation. Using the fractional

crystallization experiments on mantle-derived magma at a pressure of 1.0 GPa (Ulmer

et al., 2018), we were able to reveal the evolutionary paths of both pre- and

syn-orogenic magmas (Fig.7). We selected the Jurassic gabbro sample 09SD230

of
(Demouy et al., 2019) to represent mantle-derived magma. The sample has a Dy/Yb

ro
ratio of 1.63, and it is assumed to have initial magmatic and amphibole H 2O contents
-p
of 2.9 wt.% and of 2.1 wt.%, respectively (Ulmer et al., 2018). The pressure and
re
initial magmatic H2O content are characterized by
lP

olivine-clinopyroxene-(amphibole-orthopyroxene) dominated fractionation and

absence of garnet (Ulmer et al., 2018). A total of 68% fractionation of the initial melt
na

would increase the H2O content to 8.8 wt.% in the residual melt (stage A; Fig. 7), and
ur

would result in garnet saturation as the temperature decreases (Alonso-Perez et al.,


Jo

2009; Ulmer et al., 2018). The wide range of the stability field of garnet in andesitic‒

dacitic melt has been proven at pressures down to 0.8 GPa and H2O contents up to

12.2 wt.% (Alonso-Perez et al., 2009; Ulmer et al., 2018). Further, 40%

garnet-(amphibole-clinopyroxene-plagioclase) dominated fractionation (stage B; Fig.

7), which is approximately 81% relative to the initial melt, would produce a high

Dy/Yb ratio of 2.8 similar to that of relatively primitive rocks in the syn-orogenic unit.

It would also result in a H2O content of 14.1 wt.%, which is below the solubility limit
Journal Pre-proof

of pure H2O in a basaltic‒rhyolitic melt at 1.0 GPa (Newman and Lowenstern, 2002;

Ulmer et al., 2018; Zhang, 1999). This is consistent with a recent observation of

extensively fractionated arc magmas in the lower crust (Urann et al., 2022). The

evolved magmas will ascend ultimately and differentiate in mid- to lower-crustal

magma reservoirs or escape directly to shallow-crustal magma chambers by buoyancy

or crustal relaxation (Chen et al., 2022b; Loucks, 2021). This stage of H2O-saturated

of
fractionation is dominated by amphibole (Fig. 7) (Nandedkar et al., 2014), and

ro
fractionation of the mineral assemblage would produce an evolutionary trend toward
-p
Dy/Yb ratios similar to those of the syn-orogenic unit. Exsolution of fluid at this stage
re
would occur because of decompression during ascent (Chiaradia et al., 2012; Loucks,
lP

2021; Rezeau and Jagoutz, 2020; Wilkinson, 2013).


na

In contrast, the computation assumed that mantle-derived magma that underwent

low-pressure, H2O-unsaturated fractionation (stage D; Fig 7) with gabbro


ur

(plagioclase-dominated) cumulates would produce an evolutionary trend in Dy/Yb


Jo

ratios similar to the pre-orogenicl unit.

Although the model for residual melt compositions is sensitive to sources and

partition coefficients, it provides insights into the fractional crystallization processes

of both pre- and syn-orogenic units and a plausible explanation of the geochemical

divergences between the units. Here we propose the critical role of early garnet

fractionation in the evolution of the syn-orogenic unit; however, garnet would not
Journal Pre-proof

persist in surface samples because of their complete removal during deep

fractionation during arc compression. The Colombian Middle Cauca Au‒Cu belt

provides an example of garnet-bearing pre-mineralization intrusions followed by a

garnet-absent syn-mineralization intrusion emplaced during arc compression,

supporting an evident fractionation of garnet (Bissig et al., 2017).

5.2. Zircon trace element reveals magmatic evolution

of
ro
The pre- and syn-orogenic units define two distinct groups in the Hf vs. EuN/Eu*(zircon)
-p
diagram (Fig. 9). Hf in zircon is co-controlled by magmatic differentiation and

temperature (Blundy and Wood, 1994; Claiborne et al., 2010; Nathwani et al., 2021).
re

Progressive magmatic evolution increases Hf concentrations in residual melt


lP

(Nathwani et al., 2021), and decrease in temperature increases the values of KdHf
na

(zircon/melt) (Blundy and Wood, 1994). Together, these would result in more Hf in
ur

zircon, which is reflected in the observed coupling between increasing Hf and


Jo

decreasing Ti in zircon (Fig. 9E). The syn-orogenic unit exhibits significantly higher

EuN/Eu* ratios than the pre-orogenic unit. The Eu3+ (1.066 Å, where Å is unit of ionic

radius) preferentially substitutes Zr4+ (0.84 Å) relative to Eu2+ (1.25 Å) (Ballard et al.,

2002; Blundy and Wood, 1994; Shannon, 1976). High magmatic oxidation states

increase the Eu3+/Eu2+ ratio in melt, which leads to high EuN/Eu* ratios in zircon

(Ballard et al., 2002; Dilles et al., 2015). However, recent studies have suggested that

the EuN/Eu* ratio in zircon is mainly controlled by the melt EuN/Eu* ratio, which is
Journal Pre-proof

affected by co-crystallization of REE-bearing phases (e.g., plagioclase and titanite)

(Loader et al., 2017; Loader et al., 2022; Nathwani et al., 2021). Thus, higher

EuN/Eu*(zircon) ratios in the syn-orogenic unit may reflect higher EuN/Eu*(melt) ratios in

syn-orogenic magmas compared to pre-orogenic magmas at the time of zircon

crystallization. This is consistent with the whole-rock elemental observations (Fig.

6A), which show distinct evolutionary paths for the two units. Both igneous units

of
show broadly decreasing EuN/Eu*(zircon) ratios with increasing Hf contents (Fig. 9D),

ro
and this may reflect the reduced EuN/Eu*(melt) ratios caused by fractionation of
-p
Eu-bearing minerals (e.g., plagioclase) during magmatic evolution (Drake and Weill,
re
1975; Nathwani et al., 2021).
lP

Both pre- and syn-orogenic units show decreases in Dy/Yb(zircon) ratios with
na

increasing Hf and Ti contents (Fig. 9B), which may reflect the continuous amphibole

fractionation and decreasing temperatures during magmatic evolution. These trends


ur

are consistent with decreasing Dy/Yb(melt) with increasing SiO2 in the syn-orogenic
Jo

unit. However, the pre-orogenic unit shows a higher average Dy/Yb(zircon) ratio (0.29

vs. 0.23), which is significantly different from the whole-rock Dy/Yb ratios (1.7 vs.

1.9) (Fig. 7). We employed a thermodynamically-based model using equations 1 and 2

to further explore the observed divergence (Blundy and Wood, 1994).

Kdi = Kd0e −ΔG/RT (1)

𝑟0 𝟏
ΔG = 4πENA[ ( 𝑟𝑖 – 𝑟0 )2 + (r𝑖 – 𝑟0 ) 3
]
2 3
Journal Pre-proof

(2)

where, Kdi is the zircon‒melt partition coefficient of REE i; Kd0 is the hypothetical

strain-free distribution coefficient; ΔG is the lattice-strain energy; R is the gas

constant; T is the absolute temperature; ri is the ionic radius of the cation of REE i; r0

is the optimal site radius through the Young’s modulus (E) of the zircon; and NA is

Avogadro’s number.

of
ro
Equations 3 and 4 are obtained from equations 1 and 2:
-p
𝑟𝐷𝑦 𝑟 𝑟 𝑟
−4𝜋𝐸𝑁𝐴 [( + 0 )(𝑟𝐷𝑦 − 𝑟0 )2 − ( 𝑌𝑏 + 0 )(𝑟𝑌𝑏 − 𝑟0 )2 ]
3 6 3 6
Dy/Yb(zircon) =𝑒 𝑅𝑇 × Dy/Yb(melt)
re
(3)
lP

𝑟𝐷𝑦 𝑟 𝑟 𝑟
4𝜋𝑁𝐴 [( + 0 )(𝑟𝐷𝑦 − 𝑟0 )2 − ( 𝑌𝑏 + 0 )(𝑟𝑌𝑏 − 𝑟0 )2 ]
Definition: 3 6 3 6
= Ø
na

(4)
ur

Since Ø is a positive constant because of trivalent rDy >rYb (Shannon, 1976), equation
Jo

3 can be transformed to equation 5:

𝐸
Dy/Yb(zircon) = 𝑒 −Ø 𝑇 Dy/Yb(melt) (5)

Where θ is given by equation 6:

𝐸
θ = 𝑒 −Ø 𝑇 (6)

Thus, Dy/Yb(zircon) is expressed finally by equation 7:


Journal Pre-proof

Dy/Yb(zircon) = θ × Dy/Yb(melt) (7)

E is a function of pressure and temperature parameters (Dutta and Mandal, 2012;

Luo and Ayers, 2009; Timms et al., 2018), thus the co-control of θ values by

crystallization pressure and temperature has been shown. Equation 7 suggests that the

syn-orogenic unit has lower θ values than the pre-orogenic unit, and therefore reflects

distinct pressure and temperature conditions in the two units at the time of zircon

of
crystallization. This is consistent with the whole-rock geochemical observations and

ro
proposed fractional crystallization models (Fig. 7).
-p
re
5.3. Potential effects of high-pressure differentiation on magmatic fertility
lP

Mantle-derived basaltic arc magmas generally have Fe3+/ƩFe ratios in the range of
na

0.18–0.32 (Grocke et al., 2016; Kelley and Cottrell, 2009, 2012), which are higher

than the ratios in orthopyroxene (e.g., ~0.04) (Nimis et al., 2015), olivine (e.g., ~0.03)
ur

(Dyar et al., 1998), and garnet (e.g., ~0.077) (Nimis et al., 2015; Yaxley et al., 2012).
Jo

However, the ratios are equivalent or lower than the ratios in clinopyroxene (e.g., ~

0.44) (Rötzler and Wilke, 2005), amphibole (e.g., ~0.29) (Zhang et al., 2022), and

plagioclase (e.g., ~0.85) (Lac, 2009). Thus, the mineral assemblages in the cumulates

can significantly influence the Fe3+/ƩFe ratio of residual melt during fractional

crystallization and control magmatic oxidation state (Cottrell and Kelley, 2011; Kelley

and Cottrell, 2009, 2012; Ulmer, 1989). It is difficult to achieve accurate estimates of

magmatic Fe3+/ƩFe ratios. In this study, we evaluated semi-quantitatively the


Journal Pre-proof

influence of differentiation on the Fe3+/ƩFe ratio in residual melt by employing two

fractional crystallization models (Fig. 11). The first model assumes constant partition

coefficients of Fe3+ and Fe2+ between mineral and melt during fractional

crystallization (Fig. 11 A‒C), whereas the second model assumes constant Fe3+/ƩFe

ratios in the fractionated minerals (Fig. 11 D‒F). Both models predict high Fe3+/ƩFe

ratios in the residual melt during high-pressure fractionation of basaltic melt (stages A

of
and B; Fig. 11A and D). Subsequent differentiation (stage C; Fig. 11A and D) at mid-

ro
to lower-crustal levels (amphibole-dominated) can further elevate the Fe3+/ƩFe ratios
-p
in the residual melt. However, model 2 overestimated significantly the Fe3+/ƩFe ratios
re
because this ratio does not reach 1.0 in a terrestrial melt. If differentiation occurs only
lP

in an upper-crustal magma reservoir (plagioclase-dominated), both models generally

predict low Fe3+/ƩFe ratios in the residual melt (Fig. B and E). This scenario with
na

substantially different Fe3+/ƩFe ratios in residual melts depends on the bulk partition
ur

coefficient ratio of Fe3+/Fe2+ (model 1) and the Fe3+/ƩFe ratio of cumulates (model 2).
Jo

If the bulk partition coefficient ratio of Fe3+/Fe2+ <1 or the Fe3+/ƩFe ratio of cumulates

is lower than that in the initial melt, differentiation will produce a continuously

increasing Fe3+/ƩFe ratio in the residual melt. In contrast, high bulk partition

coefficient ratios (Fe3+/Fe2+ >1) and high Fe3+/ƩFe ratios of cumulates will produce a

continuously decreasing Fe3+/ƩFe ratio in residual melt by differentiation. It is

worthwhile to note that the ore-forming magmas in island arcs are only prone to

amphibole-dominated differentiation (e.g. Philippines) (Hollings et al., 2011), and


Journal Pre-proof

their oxygen fugacity equals continental arcs. This may reflect the contribution of

slab-derived melt to an elevated oxidation state and higher Fe3+/ƩFe ratio in island arc

basaltic magma than those in continental arcs (Cooke et al., 2014; Debret et al., 2020;

Hollings et al., 2011; Lee and Tang, 2020). Correspondingly, both models suggest the

generation of high Fe3+/ƩFe ratios in evolved melts due to amphibole-dominated

differentiation of basaltic melt with an initial Fe3+/ƩFe ratio >0.32 (Fig. C and F),

of
which is consistent with the conclusion of a recent study (Zhang et al., 2022).

ro
Some studies have argued that rising dissolved H2O and degassing could elevate
-p
oxygen fugacity (Gaillard et al., 2001; Gaillard et al., 2003; Humphreys et al., 2015).
re
However, other studies have suggested that neither dissolved H2O (Botcharnikov et al.,
lP

2005; Moore et al., 1995; Schuessler et al., 2008; Sisson and Grove, 1993; Wilke et al.,
na

2002) nor degassing (Crabtree and Lange, 2012; Waters and Lange, 2016) showed

obviously positive effects on Fe3+/ƩFe ratios. The effects of both dissolved H2O and
ur

degassing on oxygen fugacity are not yet well understood (Borisov et al., 2018).
Jo

In summary, we conclude that the Incaic I orogeny at ca. 60 Ma led to

quasi-instantaneous arc compression, permitting long-term evolution of

mantle-derived magmas at the crust‒mantle boundary. High-pressure differentiation

may have positive effects on the magmatic oxidation state, and this is consistent with

the results obtained from the zircon oxybarometer since the fO2 values of the

syn-orogenic unit (ΔFMQ = +0.7 to +2.1) are significantly higher than those of the
Journal Pre-proof

pre-orogenic unit (ΔFMQ = −1.0 to +0.4) (Fig.7). It has been shown that high

oxidation states oxidize sulfur and inhibit sulfide saturation during early

differentiation as well as maintain high sulfur contents in residual melts (Jugo et al.,

2005a, b; Jugo, 2009; Li and Ripley, 2009). A recent experiment has shown that at 1.0

GPa, the S solubility in a dacitic melt in the initial range of 200‒400 ppm at ΔFMQ =

−1 increased to 2000‒8000 ppm at ΔFMQ = +2 (Xu et al., 2022). This is consistent

of
with the observations found in the Don Javier deposit in the study area (Fig. 10)

ro
(Chen et al., 2022a). A similar scenario was also suggested for the Quechua II
-p
orogeny (ca. 12 Ma) in northern Peru (Chambefort et al., 2013; Chen et al., 2022b).
re
Furthermore, the quantitative model of Chiaradia and Caricchi (2017) also shows that
lP

huge amounts of H2O and Cu could be accumulated in the mid- to lower-crustal

reservoirs during long-term evolution, which is an essential step in the formation of a


na

giant porphyry Cu deposit.


ur

5.4. Implications for mineral exploration


Jo

It has been widely accepted that porphyry Cu deposits are generally formed from

magmas with high fO2, high oxidized S, and high H2O contents (Audétat et al., 2004;

Audétat and Pettke, 2006; Chelle-Michou and Chiaradia, 2017; Meng et al., 2021a;

Wang et al., 2014b; Xu et al., 2021; Zhu et al., 2018), although there have been

examples of relatively reduced magmas and low melt S during porphyry Cu deposit

formation (Meng et al., 2021b). Therefore, targeting fertile magmas is more


Journal Pre-proof

productive for the exploration of magmatic-hydrothermal deposits. Zircon trace

elements have been suggested as indicators for porphyry Cu deposit exploration (Lu

et al., 2016; Wade et al., 2022). In this case study, the syn-orogenic unit underwent a

distinct evolutionary path compared to the pre-orogenic unit, and this led to distinct

divergences in magmatic compositions recorded in their zircons. Therefore, zircon

trace elements can be used effectively as fingerprint to identify the two distinct

of
tectonic regimes, and they can act as robust indicators for porphyry deposit

ro
exploration in the Toquepala arc. Based on the regional dataset, we updated the
-p
recently proposed whole-rock and zircon indicators (Chen et al., 2022a) to the
re
following: no or weak whole-rock Eu anomalies, Sr/Y >103, fO2 >ΔFMQ(zircon) >+0.7
lP

(Loucks et al., 2020), and EuN/Eu*(zircon) >0.43 favor porphyry Cu mineralization in

the Toquepala arc.


na

Some studies argued that the compositional variations of zircon are not specific
ur

indicators of causative intrusions (Chiaradia and Caricchi, 2022; Viala and Hattori,
Jo

2021). For example, in the Rio Blanco mining district in northern Peru, the

pre-mineralization intrusions (Portachuela batholith) have been found to have high

fO2 and high igneous apatite S contents (Chen et al., 2022b), similar to those of the

Don Javier dacite porphyry (Chen et al., 2022a); however, the intrusions did not result

in mineralization (Fig. 13). This is because the Portachuela batholith formed under

moderate crustal thickness and normal subduction (Chen et al., 2022b). Crustal

thickness is an alternative critical factor that controls magmatic oxidation states and S
Journal Pre-proof

contents (Chiaradia, 2015; Lee and Tang, 2020; Park et al., 2021; Tang et al., 2020).

Although oxidized and S-rich magmas are considered the precondition for porphyry

mineralization (Audétat et al., 2004; Audétat and Pettke, 2006; Cao et al., 2021;

Chelle-Michou and Chiaradia, 2017; Chen et al., 2022a; Grondahl and Zajacz, 2022;

Richards, 2015; Richards et al., 2017; Streck and Dilles, 1998; Zhu et al., 2018),

compressional regimes more likely caused the formation of giant porphyry Cu

of
deposits (Chiaradia et al., 2009a; Cooke et al., 2005; Richards et al., 2017), even in

ro
thick crustal arcs in the Andes (Bissig et al., 2017; Chen et al., 2022b). These
-p
examples imply spatial-temporal limitations of zircon-based indicators. Thus, before
re
applying indicators during mineral exploration, it is vital to understand thoroughly the
lP

regional tectonic settings in specific Andean arc segments.


na

6. Conclusions
ur

We investigated the tectono-magmatic processes underlying the formation of the very


Jo

large to supergiant porphyry Cu deposits in the Paleocene‒Eocene southern Peru

metallogenic belt. The syn-orogenic igneous rocks exhibit geochemical characteristics

significantly different from those of the pre-orogenic igneous rocks, as indicated by

their higher ratios of Sr/Y, La/Yb, Dy/Yb, and EuN/Eu*. The geochemical differences

are associated with the Incaic I orogeny at ca. 60 Ma. The quasi-instantaneous arc

compression promoted long-term differentiation of mantle-derived basaltic magmas at

the crust‒mantle boundary, resulting in the stabilization of garnet and amphibole as


Journal Pre-proof

near-liquidus phases. Our survey suggests that deep crustal differentiation plays a

critical role in the formation of the oxidized and S-rich magmas with large volume

under a compressional regime, which is favorable for the formation of giant porphyry

Cu deposits. The zircon trace elements are potentially robust indicators for

distinguishing magmas generated in the two distinct tectonic regimes. However, the

spatial-temporal limitations of zircon-based indicators require a thorough

of
understanding of the regional tectonic setting before they can be applied to other

ro
Andean arc segments.
-p
Declaration of Competing Interest
re
lP

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.
na
ur

Acknowledgement
Jo

The authors have benefitted from discussions with Meng Xuyang and Bernd Lehmann.

We thank Editor-in-Chief Arturo Gómez-Tuena for editorial handling, and Massimo

Chiaradia and two anonymous reviewers for their constructive comments that greatly

improved the manuscript. This research was supported by the National Natural

Science Foundation of China (No. 41820104010).


Journal Pre-proof

References

Acosta, J.G., Huanacuni, D., Manrique, M., 2020. Base de datos de recursos y
reservas de operaciones y proyectos mineros del Perú. Instituto Geológico,
Minero y Metalúrgico. Lima, Peru.
https://www.gob.pe/institucion/ingemmet/informes-publicaciones/1423541-ba
se-de-datos-de-metalogenia (accessed 03 January 2022)

Alonso-Perez, R., Müntener, O., Ulmer, P., 2009. Igneous garnet and amphibole
fractionation in the roots of island arcs: experimental constraints on andesitic
liquids. Contrib. Mineral. Petrol. 157 (4), 541–558.

of
https://doi.org/10.1007/s00410-008-0351-8

ro
Audétat, A., Pettke, T., Dolejš, D., 2004. Magmatic anhydrite and calcite in the
ore-forming quartz-monzodiorite magma at Santa Rita, New Mexico (USA):
-p
genetic constraints on porphyry-Cu mineralization. Lithos 72 (3), 147–161.
https://doi.org/10.1016/j.lithos.2003.10.003
re
Audétat, A., Pettke, T., 2006. Evolution of a Porphyry-Cu Mineralized Magma
System at Santa Rita, New Mexico (USA). J. Petrol. 47 (10), 2021–2046.
lP

https://doi.org/10.1093/petrology/egl035
na

Audétat, A., Simon, A.C., 2012. Magmatic controls on porphyry copper genesis. In:
Hedenquist, J.W., Harris, M., Camus, F. (Eds.), Geology and genesis of major
copper deposits and districts of the world: A tribute to Richard H. Sillitoe,
ur

Society of Economic Geologists, Special Publications, p. 553–572.


https://doi.org/10.5382/SP.16.21
Jo

Ballard, J.R., Palin, M.J., Campbell, I.H., 2002. Relative oxidation states of magmas
inferred from Ce(IV)/Ce(III) in zircon: application to porphyry copper
deposits of northern Chile. Contrib. Mineral. Petrol. 144 (3), 347–364.
https://doi.org/10.1007/s00410-002-0402-5

Bellon, H., Lefevre, C., 1976. Données géochronométriques sur le volcanisme ̀ andin
dans le sud du Pérou: implications volcano-tectoniques. C.R. Acad. Sci., Ser.
D 283, 1–4.

Bissig, T., Clark, A.H., Lee, J.K., von Quadt, A., 2003. Petrogenetic and
metallogenetic responses to Miocene slab flattening: new constraints from the
El Indio-Pascua Au–Ag–Cu belt, Chile/Argentina. Miner. Deposita 38 (7),
844–862. https://doi.org/10.1007/s00126-003-0375-y
Journal Pre-proof

Bissig, T., Leal-Mejía, H., Stevens, R.B., Hart, C.J., 2017. High Sr/Y magma
petrogenesis and the link to porphyry mineralization as revealed by
Garnet-Bearing I-type granodiorite porphyries of the Middle Cauca Au-Cu
Belt, Colombia. Econ. Geol. 112 (3), 551–568.
https://doi.org/10.2113/econgeo.112.3.551

Blundy, J., Wood, B., 1994. Prediction of crystal–melt partition coefficients from
elastic moduli. Nature 372, 452–454. https://doi.org/10.1038/372452a0

Blundy, J., Mavrogenes, J., Tattitch, B., Sparks, S., Gilmer, A., 2015. Generation of
porphyry copper deposits by gas–brine reaction in volcanic arcs. Nat. Geosci.
8 (3), 235–240. https://doi.org/10.1038/ngeo2351

of
Boekhout, F., Spikings, R., Sempere, T., Chiaradia, M., Ulianov, A., Schaltegger, U.,
2012. Mesozoic arc magmatism along the southern Peruvian margin during

ro
Gondwana breakup and dispersal. Lithos 146, 48–64.
https://doi.org/10.1016/j.lithos.2012.04.015
-p
Borisov, A., Behrens, H., Holtz, F., 2018. Ferric/ferrous ratio in silicate melts: a new
re
model for 1 atm data with special emphasis on the effects of melt composition.
Contrib. Mineral. Petrol. 173 (12), 98.
lP

https://doi.org/10.1007/s00410-018-1524-8

Botcharnikov, R., Koepke, J., Holtz, F., McCammon, C., Wilke, M., 2005. The effect
na

of water activity on the oxidation and structural state of Fe in a ferro-basaltic


melt. Geochim. Cosmochim. Acta 69 (21), 5071–5085.
ur

https://doi.org/10.1016/j.gca.2005.04.023

Bottazzi, P., Tiepolo, M., Vannucci, R., Zanetti, A., Brumm, R., Foley, S.F., Oberti, R.,
Jo

1999. Distinct site preferences for heavy and light REE in amphibole and the
prediction of Amph/LDREE. Contrib. Mineral. Petrol. 137 (1), 36–45.
https://doi.org/10.1007/s004100050580

Bourdon, E., Eissen, J.P., Gutscher, M.A., Monzier, M., Samaniego, P., Robin, C.,
Bollinger, C., Cotten, J., 2002. Slab melting and slab melt metasomatism in
the Northern Andean Volcanic Zone: adakites and high-Mg andesites from
Pichincha volcano (Ecuador). Bull. Soc. Geol. Fr. 173 (3), 195–206.
https://doi.org/10.2113/173.3.195

Cao, K., Yang, Z.M., White, N.C., Hou, Z.Q., 2021. Generation of the giant porphyry
Cu-Au deposit by repeated recharge of mafic magmas at Pulang in eastern
Tibet. Econ. Geol. 114 (1), 57–90. https://doi.org/10.5382/econgeo.4860
Journal Pre-proof

Chambefort, I., Dilles, J.H., Kent, A.J.R., 2008. Anhydrite-bearing andesite and dacite
as a source for sulfur in magmatic-hydrothermal mineral deposits. Geology 36
(9), 719–722. https://doi.org/10.1130/G24920A.1

Chambefort, I., Dilles, J.H., Longo, A.A., 2013. Amphibole geochemistry of the
Yanacocha Volcanics, Peru: Evidence for diverse sources of magmatic
volatiles related to gold ores. J. Petrol. 54 (5), 1017–1046.
https://doi.org/10.1093/petrology/egt004

Chapman, A.D., Ducea, M.N., McQuarrie, N., Coble, M., Petrescu, L., Hoffman, D.,
2015. Constraints on plateau architecture and assembly from deep crustal
xenoliths, northern Altiplano (SE Peru). Geol. Soc. Am. Bull. 127 (11-12),
1777–1797. https://doi.org/10.1130/B31206.1

of
Chelle-Michou, C., Chiaradia, M., Béguelin, P., Ulianov, A., 2015. Petrological

ro
evolution of the magmatic suite associated with the coroccohuayco Cu(–Au–
Fe) porphyry–skarn deposit, Peru. J. Petrol. 56 (9), 1829–1862.
-p
https://doi.org/10.1093/petrology/egv056
re
Chelle-Michou, C., Chiaradia, M., 2017. Amphibole and apatite insights into the
evolution and mass balance of Cl and S in magmas associated with porphyry
lP

copper deposits. Contrib. Mineral. Petrol. 172 (11), 105.


https://doi.org/10.1007/s00410-017-1417-2
na

Chen, H.Y., Wu, C., 2020. Metallogenesis and major challenges of porphyry copper
systems above subduction zones. Sci. China Earth Sci. 63 (7), 899‒918.
ur

https://doi.org/10.1007/s11430-019-9595-8

Chen, N., Mao, J.W., Ye, Z.C., Duan, Z., Li, H.Y., 2022a. Rapid transition to fertile
Jo

magma and promotion of porphyry mineralization: A case study from the Don
Javier deposit. Ore Geol. Rev., 104964.
https://doi.org/10.1016/j.oregeorev.2022.104964

Chen, N., Meng, X.Y., Mao, J.W., Xie, G.Q., 2022b. Genetic relationship between
subduction of slab topographic anomalies and porphyry deposit formation:
Insight from the source and evolution of Rio Blanco magmas. J. Petrol. 63 (6),
egac045. https://doi.org/10.1093/petrology/egac045

Chiaradia, M., 2009. Adakite-like magmas from fractional crystallization and


melting-assimilation of mafic lower crust (Eocene Macuchi arc, Western
Cordillera, Ecuador). Chem. Geol. 265 (3–4), 468–487.
https://doi.org/10.1016/j.chemgeo.2009.05.014
Journal Pre-proof

Chiaradia, M., Caricchi, L., 2017. Stochastic modelling of deep magmatic controls on
porphyry copper deposit endowment. Sci. Rep. 7(1): 1–11.
https://doi.org/10.1038/srep44523

Chiaradia, M.,Caricchi, L., 2022. Supergiant porphyry copper deposits are failed large
eruptions. Commun. Earth Environ. , 3(1): 107.
https://doi.org/10.1038/s43247-022-00440-7

Chiaradia, M., Merino, D., Spikings, R., 2009a. Rapid transition to long-lived deep
crustal magmatic maturation and the formation of giant porphyry-related
mineralization (Yanacocha, Peru). Earth Planet. Sci. Lett. 288 (3–4), 505–515.
https://doi.org/10.1016/j.epsl.2009.10.012

of
Chiaradia, M., Müntener, O., Beate, B., Fontignie, D., 2009b. Adakite-like volcanism
of Ecuador: lower crust magmatic evolution and recycling. Contrib. Mineral.

ro
Petrol. 158 (5), 563–588. https://doi.org/10.1007/s00410-009-0397-2
-p
Chiaradia, M., Ulianov, A., Kouzmanov, K., Beate, B., 2012. Why large porphyry Cu
deposits like high Sr/Y magmas? Sci. Rep. 2 (1), 1–7.
re
https://doi.org/10.1038/srep00685
lP

Chiaradia, M., Schaltegger, U., Spikings, R., Wotzlaw, J.F., Ovtcharova, M., 2013.
How accurately can we date the duration of magmatic-hydrothermal events in
porphyry systems?—an invited paper. Econ. Geol. 108 (4), 565–584.
na

https://doi.org/10.2113/econgeo.108.4.565
ur

Chiaradia, M., 2015. Crustal thickness control on Sr/Y signatures of recent arc
magmas: an Earth scale perspective. Sci. Rep. 5 (1), 1–5.
https://doi.org/10.1038/srep08115
Jo

Claiborne, L.L., Miller, C.F., Flanagan, D.M., Clynne, M.A., Wooden, J.L., 2010.
Zircon reveals protracted magma storage and recycling beneath Mount St.
Helens. Geology 38 (11), 1011–1014. https://doi.org/10.1130/G31285.1

Clark, A.H., Farrar, E., Kontak, D.J., Langridge, R.J., Arenas F, M.J., France, L.J.,
McBride, S.L., Woodman, P.L., Wasteneys, H.A., Sandeman, H.A., 1990.
Geologic and geochronologic constraints on the metallogenic evolution of the
Andes of southeastern Peru. Econ. Geol. 85 (7), 1520–1583.
https://doi.org/10.2113/gsecongeo.85.7.1520

Clark, A.H., 1993. Are Outsize Porphyry Copper Deposits Either Anatomically or
Environmentally Distinctive? In: Whiting, B.H., Hodgson, C.J., Mason, R.
Journal Pre-proof

(Eds.), Giant Ore Deposits, Society of Economic Geologists, Special


Publications, p. 213–282. https://doi.org/10.5382/SP.02.06

Clark, A.H., 2003. The Paleocene-middle Eocene porphyry copper-molybdenum


province of southern Perú: Hypogene and supergene ore-genesis and
metallogenesis. ProExplo Meeting, Instituto Geológico, Minero y
Metalúrgico, Lima, Perú, April, 2003, Field Guidebook; pp. 115.

Coira, B., Davidson, J., Mpodozis, C., Ramos, V., 1982. Tectonic and magmatic
evolution of the Andes of northern Argentina and Chile. Earth Sci. Rev. 18 (3–
4), 303–332. https://doi.org/10.1016/0012-8252(82)90042-3

Cooke, D.R., Hollings, P., Walshe, J.L., 2005. Giant porphyry deposits: characteristics,

of
distribution, and tectonic controls. Econ. Geol. 100 (5), 801–818.
https://doi.org/10.2113/gsecongeo.100.5.801

ro
Cooke, D.R., Hollings, P., Wilkinson, J.J., Tosdal, R.M., 2014. Geochemistry of
-p
Porphyry Deposits. In: Holland, H.D., Turekian, K.K. (Eds.), Treatise on
Geochemistry (Second Edition), Elsevier, Oxford, p. 357‒381.
re
https://doi.org/10.1016/B978-0-08-095975-7.01116-5
lP

Cottrell, E., Kelley, K.A., 2011. The oxidation state of Fe in MORB glasses and the
oxygen fugacity of the upper mantle. Earth Planet. Sci. Lett. 305 (3–4), 270–
282. https://doi.org/10.1016/j.epsl.2011.03.014
na

Crabtree, S.M., Lange, R.A., 2012. An evaluation of the effect of degassing on the
ur

oxidation state of hydrous andesite and dacite magmas: a comparison of pre-


and post-eruptive Fe2+ concentrations. Contrib. Mineral. Petrol. 163 (2), 209‒
224. https://doi.org/10.1007/s00410-011-0667-7
Jo

Davidson, J., Turner, S., Handley, H., Macpherson, C., Dosseto, A., 2007. Amphibole
“sponge” in arc crust? Geology 35 (9), 787–790.
https://doi.org/10.1130/G23637A.1

Davies, R.C., 2002. Tectonic, magmatic and metallogenic evolution of the Cajamarca
mining district, Northern Peru (Ph.D thesis). James Cook University, Australia,
P. 323.

Debret, B., Reekie, C., Mattielli, N., Savov, I., Beunon, H., Ménez, B., Williams, H.,
2020. Redox transfer at subduction zones: insights from Fe isotopes in the
Mariana forearc. Geochem. Perspect. Lett. 12, 46–51.
https://doi.org/10.7185/geochemlet.2003
Journal Pre-proof

Defant, M.J., Drummond, M.S., 1990. Derivation of some modern arc magmas by
melting of young subducted lithosphere. Nature 347, 662–665.
https://doi.org/10.1038/347662a0

Delacour, A., Gerbe, M.C., Thouret, J.C., Wörner, G., Paquereau-Lebti, P., 2007.
Magma evolution of Quaternary minor volcanic centres in southern Peru,
Central Andes. Bull. Volcanol. 69 (6), 581–608.
https://doi.org/10.1007/s00445-006-0096-z

Demouy, S., Paquette, J.L., de Saint Blanquat, M., Benoit, M., Belousova, E.A.,
O'Reilly, S.Y., García, F., Tejada, L.C., Gallegos, R., Sempere, T., 2012.
Spatial and temporal evolution of Liassic to Paleocene arc activity in southern
Peru unraveled by zircon U–Pb and Hf in-situ data on plutonic rocks. Lithos

of
155, 183–200. https://doi.org/10.1016/j.lithos.2012.09.001

ro
Demouy, S., Benoit, M., de Saint-Blanquat, M., Ganne, J., 2019. Evolution of a
long-lived continental arc: a geochemical approach (Arequipa Batholith,
-p
Southern Peru). Solid Earth Discuss., 1–42. https://doi.org/10.5194/se-2019-43
re
Dilles, J.H., Kent, A.J., Wooden, J.L., Tosdal, R.M., Koleszar, A., Lee, R.G., Farmer,
L.P., 2015. Zircon compositional evidence for sulfur-degassing from
lP

ore-forming arc magmas. Econ. Geol. 110 (1), 241–251.


https://doi.org/10.2113/econgeo.110.1.241
na

Drake, M.J., Weill, D.F., 1975. Partition of Sr, Ba, Ca, Y, Eu2+, Eu3+, and other REE
between plagioclase feldspar and magmatic liquid: an experimental study.
ur

Geochim. Cosmochim. Acta 39 (5), 689–712.


https://doi.org/10.1016/0016-7037(75)90011-3
Jo

Drobe, J., Lindsay, D., Stein, H., Gabites, J., 2013. Geology, Mineralization, and
Geochronological Constraints of the Mirador Cu–Au Porphyry District,
Southeast Ecuador. Econ. Geol. 108 (1), 11–35.
https://doi.org/10.2113/econgeo.108.1.11

Duan, Z., Guo, W.M., Xiang, H.L., Liu, J.A., Jaimes, F., Astete, I., 2022.
Geochronology and geochemistry of Early Cretaceous volcanic sequences in
Northwestern Peru: implications for Farallon Plate subduction. Int. Geol. Rev.,
1–24. https://doi.org/10.1080/00206814.2021.2024769

Dutta, R., Mandal, N., 2012. Effects of pressure on the elasticity and stability of
zircon (ZrSiO4): First-principle investigations. Comput. Mater. Sci 54, 157–
164. https://doi.org/10.1016/j.commatsci.2011.09.035
Journal Pre-proof

Dyar, M.D., Delaney, J.S., Sutton, S.R., Schaefer, M.W., 1998. Fe3+ distribution in
oxidized olivine: A synchrotron micro-XANES study. Am. Miner. 83 (12),
1361–1365. https://doi.org/10.2138/am-1998-1227

England, P.C., Katz, R.F., 2010. Melting above the anhydrous solidus controls the
location of volcanic arcs. Nature 467 (7316), 700–703.
https://doi.org/10.1038/nature09417

Errázuriz-Henao, C., Gómez-Tuena, A., Duque-Trujillo, J., Weber, M., 2019. The role
of subducted sediments in the formation of intermediate mantle-derived
magmas from the Northern Colombian Andes. Lithos 336-337, 151–168.
https://doi.org/10.1016/j.lithos.2019.04.007

of
Espurt, N., Funiciello, F., Martinod, J., Guillaume, B., Regard, V., Faccenna, C.,
Brusset, S., 2008. Flat subduction dynamics and deformation of the South

ro
American plate: Insights from analog modeling. Tectonics 27 (3), TC3011.
https://doi.org/10.1029/2007TC002175
-p
Estrada, F., 1978. Estudio de correlacion de los porfidos de cobre del sur del Peru:
re
edades radiometricas de los principales eventos geologicos de los porfidos del
sur del Peru. INGEMMET. Lima, Peru,, P. 18.
lP

Farner, M.J., Lee, C.T.A., 2017. Effects of crustal thickness on magmatic


differentiation in subduction zone volcanism: A global study. Earth Planet. Sci.
na

Lett. 470, 96–107. https://doi.org/10.1016/j.epsl.2017.04.025


ur

Fujimaki, H., Tatsumoto, M., Aoki, K.i., 1984. Partition coefficients of Hf, Zr, and
REE between phenocrysts and groundmasses. J. Geophys. Res.: Solid Earth 89,
B662–B672. https://doi.org/10.1029/JB089iS02p0B662
Jo

Gaillard, F., Scaillet, B., Pichavant, M., Bény, J.-M., 2001. The effect of water and fO2
on the ferric–ferrous ratio of silicic melts. Chem. Geol. 174 (1), 255–273.
https://doi.org/10.1016/S0009-2541(00)00319-3

Gaillard, F., Pichavant, M., Mackwell, S., Champallier, R.m., Scaillet, B.,
McCammon, C., 2003. Chemical transfer during redox exchanges between H2
and Fe-bearing silicate melts. Am. Miner. 88 (2-3), 308‒315.
https://doi.org/10.2138/am-2003-2-308

Gerbe, M.C., Thouret, J.C., 2004. Role of magma mixing in the petrogenesis of tephra
erupted during the 1990–98 explosive activity of Nevado Sabancaya, southern
Peru. Bull. Volcanol. 66 (6), 541–561.
Journal Pre-proof

https://doi.org/10.1007/s00445-004-0340-3

Godoy, B., Wörner, G., Kojima, S., Aguilera, F., Simon, K., Hartmann, G., 2014.
Low-pressure evolution of arc magmas in thickened crust: The San Pedro–
Linzor volcanic chain, Central Andes, northern Chile. J. South Am. Earth Sci.
52, 24–42. https://doi.org/10.1016/j.jsames.2014.02.004

Gómez-Tuena, A., Straub, S.M., Zellmer, G.F., 2014. An introduction to orogenic


andesites and crustal growth. Geol. Soc., London, Spec. Publ. 385 (1), 1‒13.
https://doi.org/10.1144/SP385.16

Green, T.H., 1970. High pressure experimental studies on the mineralogical


constitution of the lower crust. Phys. Earth Planet. Inter. 3, 441–450.

of
https://doi.org/10.1016/0031-9201(70)90086-5

ro
Grocke, S.B., Cottrell, E., de Silva, S., Kelley, K.A., 2016. The role of crustal and
eruptive processes versus source variations in controlling the oxidation state of
-p
iron in Central Andean magmas. Earth Planet. Sci. Lett. 440, 92–104.
https://doi.org/10.1016/j.epsl.2016.01.026
re

Grondahl, C., Zajacz, Z., 2022. Sulfur and chlorine budgets control the ore fertility of
lP

arc magmas. Nat. Commun. 13 (1), 1–11.


https://doi.org/10.1038/s41467-022-31894-0
na

Hall, P.S., 2012. On the thermal evolution of the mantle wedge at subduction zones.
Phys. Earth Planet. Inter. 198, 9–27.
ur

https://doi.org/10.1016/j.pepi.2012.03.004

Haschke, M., Siebel, W., Günther, A., Scheuber, E., 2002. Repeated crustal thickening
Jo

and recycling during the Andean orogeny in north Chile (21–26 S). J. Geophys.
Res.: Solid Earth 107 (B1), ECV 6-1-ECV 6-18.
https://doi.org/10.1029/2001JB000328

Haschke, M., Gunther, A., 2003. Balancing crustal thickening in arcs by tectonic vs.
magmatic means. Geology 31 (11), 933–936.
https://doi.org/10.1130/G19945.1

Hayden, L.A., Watson, E.B., 2007. Rutile saturation in hydrous siliceous melts and its
bearing on Ti-thermometry of quartz and zircon. Earth Planet. Sci. Lett. 258
(3–4), 561–568. https://doi.org/10.1016/j.epsl.2007.04.020

Hollings, P., Cooke, D.R., Clark, A., 2005. Regional geochemistry of Tertiary igneous
Journal Pre-proof

rocks in central Chile: Implications for the geodynamic environment of giant


porphyry copper and epithermal gold mineralization. Econ. Geol. 100 (5),
887–904. https://doi.org/10.2113/gsecongeo.100.5.887

Hollings, P., Cooke, D.R., Waters, P.J., Cousens, B., 2011. Igneous Geochemistry of
Mineralized Rocks of the Baguio District, Philippines: Implications for
Tectonic Evolution and the Genesis of Porphyry-Style Mineralization. Econ.
Geol. 106 (8), 1317–1333. https://doi.org/10.2113/econgeo.106.8.1317

Humphreys, M.C.S., Brooker, R.A., Fraser, D.G., Burgisser, A., Mangan, M.T.,
McCammon, C., 2015. Coupled interactions between volatile activity and Fe
oxidation state during arc crustal processes. J. Petrol. 56 (4), 795–814.
https://doi.org/10.1093/petrology/egv017

of
IGEMMENT, 2015. Mapa geologico del Perú, Escala 1: 3,000,000. Lima, Peru:

ro
INGEMMET. -p
James, D.E., Sacks, I.S., 1999. Cenozoic formation of the Central Andes: A
geophysical perspective. In: Skinner, B.J. (Eds.), Geology and ore deposits of
re
the Central Andes, Society of Economic Geologists, Special Publications, p.
1–25. https://doi.org/10.5382/SP.07.01
lP

Johnson, K., 1994. Experimental cpx/and garnet/melt partitioning of REE and other
trace elements at high pressures: petrogenetic implications. Mineral. Mag. 58
na

(1), 454–455. https://doi.org/10.1180/minmag.1994.58A.1.236


ur

Jugo, P.J., Luth, R.W., Richards, J.P., 2005a. Experimental data on the speciation of
sulfur as a function of oxygen fugacity in basaltic melts. Geochim.
Cosmochim. Acta 69 (2), 497–503. https://doi.org/10.1016/j.gca.2004.07.011
Jo

Jugo, P.J., Luth, R.W., Richards, J.P., 2005b. An Experimental Study of the Sulfur
Content in Basaltic Melts Saturated with Immiscible Sulfide or Sulfate Liquids
at 1300°C and 1.0 GPa. J. Petrol. 46 (4), 783–798.
https://doi.org/10.1093/petrology/egh097

Jugo, P.J., 2009. Sulfur content at sulfide saturation in oxidized magmas. Geology 37
(5), 415–418. https://doi.org/10.1130/G25527A.1

Jugo, P.J., Wilke, M., Botcharnikov, R.E., 2010. Sulfur K-edge XANES analysis of
natural and synthetic basaltic glasses: Implications for S speciation and S
content as function of oxygen fugacity. Geochim. Cosmochim. Acta 74 (20),
5926–5938. https://doi.org/10.1016/j.gca.2010.07.022
Journal Pre-proof

Kay, S.M., Godoy, E., Kurtz, A., 2005. Episodic arc migration, crustal thickening,
subduction erosion, and magmatism in the south-central Andes. Geol. Soc. Am.
Bull. 117 (1–2), 67–88. https://doi.org/10.1130/B25431.1

Kelley, K.A., Cottrell, E., 2009. Water and the Oxidation State of Subduction Zone
Magmas. Science 325 (5940), 605–607.
https://doi.org/10.1126/science.1174156

Kelley, K.A., Cottrell, E., 2012. The influence of magmatic differentiation on the
oxidation state of Fe in a basaltic arc magma. Earth Planet. Sci. Lett. 329–330,
109–121. https://doi.org/10.1016/j.epsl.2012.02.010

Lac, D., 2009. Using the oxidation state of iron plagioclase to evaluate magma

of
oxygen fugacity: a micro-XANES study (M. Sc. thesis). University of
Massachusetts Amherst, Massachusetts, USA, P. 91.

ro
Le Bas, M.J., Le Maitre, R.W., Streckeisen, A., Zanettin, B., 1986. A chemical
-p
classification of volcanic rocks based on the total alkali-silica diagram. J.
Petrol. 27 (3), 745–750. https://doi.org/10.1093/petrology/27.3.745
re

Lee, C.-T.A., Tang, M., 2020. How to make porphyry copper deposits. Earth Planet.
lP

Sci. Lett. 529, 115868. https://doi.org/10.1016/j.epsl.2019.115868

Li, C., Ripley, E.M., 2009. Sulfur contents at sulfide-liquid or anhydrite saturation in
na

silicate melts: empirical equations and example applications. Econ. Geol. 104
(3), 405–412. https://doi.org/10.2113/gsecongeo.104.3.405
ur

Li, W., Yang, Z., Cao, K., Lu, Y., Sun, M., 2019. Redox-controlled generation of the
giant porphyry Cu–Au deposit at Pulang, southwest China. Contrib. Mineral.
Jo

Petrol. 174 (2), 12. https://doi.org/10.1007/s00410-019-1546-x

Loader, M.A., Wilkinson, J.J., Armstrong, R.N., 2017. The effect of titanite
crystallisation on Eu and Ce anomalies in zircon and its implications for the
assessment of porphyry Cu deposit fertility. Earth Planet. Sci. Lett. 472, 107–
119. https://doi.org/10.1016/j.epsl.2017.05.010

Loader, M.A., Nathwani, C.L., Wilkinson, J.J., Armstrong, R.N., 2022. Controls on
the magnitude of Ce anomalies in zircon. Geochim. Cosmochim. Acta, 242–
257. https://doi.org/10.1016/j.gca.2022.03.024

Loucks, R.R., Fiorentini, M.L., Rohrlach, B.D., 2018. Divergent T–ƒO2 paths during
crystallisation of H2O-rich and H2O-poor magmas as recorded by Ce and U in
Journal Pre-proof

zircon, with implications for TitaniQ and TitaniZ geothermometry. Contrib.


Mineral. Petrol. 173 (12), 104. https://doi.org/10.1007/s00410-018-1529-3

Loucks, R.R., Fiorentini, M.L., Henríquez, G.J., 2020. New magmatic oxybarometer
using trace elements in zircon. J. Petrol. 61, 1–30.
https://doi.org/10.1093/petrology/egaa034

Loucks, R.R., 2021. Deep entrapment of buoyant magmas by orogenic tectonic stress:
Its role in producing continental crust, adakites, and porphyry copper deposits.
Earth Sci. Rev. 220, 103744. https://doi.org/10.1016/j.earscirev.2021.103744

Luo, Y., Ayers, J.C., 2009. Experimental measurements of zircon/melt trace-element


partition coefficients. Geochim. Cosmochim. Acta 73 (12), 3656–3679.

of
https://doi.org/10.1016/j.gca.2009.03.027

ro
Lu, Y.J., Loucks, R.R., Fiorentini, M., McCuaig, T.C., Evans, N.J., Yang, Z.M., Hou,
Z.Q., Kirkland, C.L., Parra-Avila, L.A., Kobussen, A., Richards, J.P., 2016.
-p
Zircon compositions as a pathfinder for porphyry Cu ± Mo ± Au deposits. In:
Richards, J.P. (Eds.), Tectonics and Metallogeny of the Tethyan Orogenic Belt,
re
Society of Economic Geologists, p. 329–347. https://doi.org/10.5382/SP.19.13
lP

Maksaev, V., Almonacid, T.A., Munizaga, F., Valencia, V., McWilliams, M., Barra, F.,
2010. Determinaciones geocronologicas y termocronologicas para la
mineralizacion de cobre porfidico en la zona de alteracion de Domeyko, norte
na

de Chile. Andean Geol. 37 (1), 144–176.


http://dx.doi.org/10.5027/andgeoV37n1-a07
ur

Mamani, M., Tassara, A., Wörner, G., 2008. Composition and structural control of
crustal domains in the central Andes. Geochem. Geophys. Geosyst. 9 (3).
Jo

https://doi.org/10.1029/2007GC001925

Mamani, M., Wörner, G., Sempere, T., 2010. Geochemical variations in igneous rocks
of the Central Andean orocline (13 °S to 18 °S): Tracing crustal thickening and
magma generation through time and space. Geol. Soc. Am. Bull. 122 (1–2),
162–182. https://doi.org/10.1130/B26538.1

Mamani, M., Acosta, H., Rodriguez, J.P., Cutipa, M., Cacya, L., 2011. Geotectonic
domains: based on the tectonics, geology, magmatism, geochemistry, mineral
deposits (Fe-Cu-Au-Mo) and geophysics. 30th Convención Minera. Lima,
Peru: PERUMIN; pp. 20.

Mamani, M., Carlotto, V., Choquehuanca, S., Santos, A., Rodriguez, J., Cueva, E.,
Journal Pre-proof

Chavez, L., Cereceda, C., Rodríguez, R., Cacya, L., 2016. Base de datos
litogeoquímica del perú. INGEMMET. Lima, Peru

Manrique, J.P.R., 2011. Magmatismo meso-cenozoico de la cordillera de la costa y


borde oeste de la cordillera occidental del sur de Perú (Tesis para optar el
Título Profesional de ingeniero geólogo). Universudad Nacional San Agustín,
Arequipa, Peru, P. 138.

Mao, J., Xie, G., Yuan, S., Liu, P., Meng, X., Zhou, Z., Zheng, W., 2018. Current
research progress and future trends of porphyry-skarn copper and
granite-related tin polymetallic deposits in the Circum Pacific metallogenic
belts. Acta Petr. Sinica 34 (9), 2501–2517.

of
Martínez, W., Cervantes, J., 2003. Rocas ígneas en el sur del Perú: nuevos datos
geocronométricos, geoquímicos y estructurales entre los paralelos 16° y 18°

ro
30' Latitud Sur. INGEMMET, Lima, Peru, p. 140.
-p
Martínez, W., Marchena, A.A., Otero, J.F., Cervantes, J., León, W.R., 2017. Geología
y controles tectonomagmáticos de los sistemas porfiríticos en el arco
re
magmático occidental sur del Perú. INGEMMET, Lima, Peru, p. 113.
lP

MéGard, F., 1987. Cordilleran Andes and Marginal Andes: a Review of Andean
Geology North of the Arica Elbow (18 °S). In: Monger, J.W.H., Francheteau,
J. (Eds.), Circum-Pacific Orogenic Belts and Evolution of the Pacific Ocean
na

Basin, p. 71–95. https://doi.org/10.1029/GD018p0071


ur

Meng, X.Y., Kleinsasser, J.M., Richards, J.P., Tapster, S.R., Jugo, P.J., Simon, A.C.,
Kontak, D.J., Robb, L., Bybee, G.M., Marsh, J.H., 2021a. Oxidized sulfur-rich
arc magmas formed porphyry Cu deposits by 1.88 Ga. Nat. Commun. 12 (1),
Jo

1–9. https://doi.org/10.1038/s41467-021-22349-z

Meng, X.Y., Richards, J.P., Kontak, D.J., Simon, A.C., Kleinsasser, J.M., Marsh, J.H.,
Stern, R.A., Jugo, P.J., 2021b. Variable modes of formation for
tonalite-trondhjemite-granodiorite-diorite (TTG)-related porphyry-type Cu ±
Au deposits in the Neoarchean southern Abitibi subprovince (Canada):
Evidence from petrochronology and oxybarometry. J. Petrol. 62 (11), egab079.
https://doi.org/10.1093/petrology/egab079

Moore, G., Righter, K., Carmichael, I.S.E., 1995. The effect of dissolved water on the
oxidation state of iron in natural silicate liquids. Contrib. Mineral. Petrol. 120
(2), 170–179. https://doi.org/10.1007/BF00287114
Journal Pre-proof

Mukasa, S.B., Tilton, G.R., 1985. Zircon U-Pb ages of super-units in the Coastal
batholith, Peru. In: Pitcher, W.S., Atherton, M.P., Cobbing, E.J., Beckingsale,
R.D. (Eds.), Magmatism at the plate edge: The Peruvian Andes:,
Glasgow, UK,, p. 203–207. https://doi.org/10.1007/978-1-4899-5820-4_17

Mukasa, S.B., 1986. Zircon U–Pb ages of super-units in the Coastal batholith, Peru:
Implications for magmatic and tectonic processes. Geol. Soc. Am. Bull. 97 (2),
241–254.
https://doi.org/10.1130/0016-7606(1986)97<241:ZUAOSI>2.0.CO;2

Munoz, M., Charrier, R., Fanning, C., Maksaev, V., Deckart, K., 2012. Zircon trace
element and O–Hf isotope analyses of mineralized intrusions from El Teniente
ore deposit, Chilean Andes: constraints on the source and magmatic evolution

of
of porphyry Cu–Mo related magmas. J. Petrol. 53 (6), 1091–1122.
https://doi.org/10.1093/petrology/egs010

ro
Nandedkar, R.H., Ulmer, P., Müntener, O., 2014. Fractional crystallization of
-p
primitive, hydrous arc magmas: an experimental study at 0.7 GPa. Contrib.
Mineral. Petrol. 167 (6), 1015. https://doi.org/10.1007/s00410-014-1015-5
re

Nandedkar, R.H., Hürlimann, N., Ulmer, P., Müntener, O., 2016. Amphibole–melt
lP

trace element partitioning of fractionating calc-alkaline magmas in the lower


crust: an experimental study. Contrib. Mineral. Petrol. 171 (8), 71.
https://doi.org/10.1007/s00410-016-1278-0
na

Naney, M.T., 1983. Phase equilibriua of rock-forming ferromagnesian silicates in


ur

granitic systems. Amer. Jour. Sci. 283, 993–1033.


https://doi.org/10.2475/ajs.283.10.993
Jo

Nathwani, C.L., Simmons, A.T., Large, S.J., Wilkinson, J.J., Buret, Y., Ihlenfeld, C.,
2021. From long-lived batholith construction to giant porphyry copper deposit
formation: petrological and zircon chemical evolution of the Quellaveco
District, Southern Peru. Contrib. Mineral. Petrol. 176 (2), 1–21.
https://doi.org/10.1007/s00410-020-01766-1

Newman, S., Lowenstern, J.B., 2002. VolatileCalc: a silicate melt–H2O–CO2 solution


model written in Visual Basic for excel. Comput. Geosci. 28 (5), 597–604.
https://doi.org/10.1016/S0098-3004(01)00081-4

Nimis, P., Goncharov, A., Ionov, D.A., McCammon, C., 2015. Fe3+ partitioning
systematics between orthopyroxene and garnet in mantle peridotite xenoliths
and implications for thermobarometry of oxidized and reduced mantle rocks.
Journal Pre-proof

Contrib. Mineral. Petrol. 169 (1), 6.


https://doi.org/10.1007/s00410-014-1101-8

Noble, D., Vidal, C., Perelló, J., Rodríguez, O., 2004. Space-time relationships of
some porphyry Cu-Au, epithermal Au, and other magmatic-related mineral
deposits in northern Perú. In: Sillitoe, R.H., Perelló, J., Vidal, C.E. (Eds.),
Andean Metallogeny: New Discoveries, Concepts, and Updates, Society of
Economic Geologists, Special Publications, p. 313–318.
https://doi.org/10.5382/SP.11.17

Noble, D.C., McKEE, E.H., Mégard, F., 1979. Early Tertiary “Incaic” tectonism,
uplift, and volcanic activity, Andes of central Peru. Geol. Soc. Am. Bull. 90
(10), 903–907.

of
https://doi.org/10.1130/0016-7606(1979)90<903:ETITUA>2.0.CO;2

ro
Noble, D.C., Sébrier, M., Megard, F., McKee, E.H., 1985. Demonstration of two
pulses of Paleogene deformation in the Andes of Peru. Earth Planet. Sci. Lett.
-p
73 (2-4), 345–349. https://doi.org/10.1016/0012-821X(85)90082-2
re
Noble, D.C., Wise, J.M., 2016. Timing of Incaic deformation and subsequent erosion,
middle Eocene volcanism, and plutonism in central Southern Peru with
lP

implications for mineralization. XVIII Congreso Nacional de Geológia,


Resumes Extendidos.
na

Noury, M., Philippon, M., Bernet, M., Paquette, J.-L., Sempere, T., 2017. Geological
record of flat slab–induced extension in the southern Peruvian forearc.
ur

Geology 45 (8), 723–726. https://doi.org/10.1130/G38990.1

Pardo-Casas, F., Molnar, P., 1987. Relative motion of the Nazca (Farallon) and South
Jo

American plates since Late Cretaceous time. Tectonics 6 (3), 233–248.


https://doi.org/10.1029/TC006i003p00233

Park, J.-W., Campbell, I.H., Chiaradia, M., Hao, H., Lee, C.-T., 2021. Crustal
magmatic controls on the formation of porphyry copper deposits. Nat. Rev.
Earth Environ. 2 (8), 542–557. https://doi.org/10.1038/s43017-021-00182-8

Perelló, J., Carlotto, V.c., Zárate, A., Ramos, P., Posso, H.c., Neyra, C., Caballero, A.,
Fuster, N.s., Muhr, R., 2003. Porphyry-style alteration and mineralization of
the middle Eocene to early Oligocene Andahuaylas-Yauri belt, Cuzco region,
Peru. Econ. Geol. 98 (8), 1575–1605.
https://doi.org/10.2113/gsecongeo.98.8.1575
Journal Pre-proof

Perrin, A., Goes, S., Prytulak, J., Rondenay, S., Davies, D.R., 2018. Mantle wedge
temperatures and their potential relation to volcanic arc location. Earth Planet.
Sci. Lett. 501, 67–77. https://doi.org/10.1016/j.epsl.2018.08.011

Pietranik, A., Holtz, F., Koepke, J., Puziewicz, J., 2009. Crystallization of quartz
dioritic magmas at 2 and 1 kbar: experimental results. Mineral. Petrol. 97 (1–
2), 1–21. https://doi.org/10.1007/s00710-009-0070-5

Plank, T., Langmuir, C.H., 1988. An evaluation of the global variations in the major
element chemistry of arc basalts. Earth Planet. Sci. Lett. 90 (4), 349–370.
https://doi.org/10.1016/0012-821X(88)90135-5

Plank, T., 2014. The chemical composition of subducting sediments. In: Turekian,

of
K.K. (Eds.), Treatise on Geochemistry (Second Edition), Elsevier, Amsterdam,
p. 607–629. https://doi.org/10.1016/B978-0-08-095975-7.00319-3

ro
Quang, C.X., Clark, A.H., Lee, J.K., Guillén, J., 2003. 40Ar-39Ar ages of hypogene
-p
and supergene mineralization in the Cerro Verde-Santa Rosa porphyry Cu-Mo
cluster, Arequipa, Peru. Econ. Geol. 98 (8), 1683–1696.
re
https://doi.org/10.2113/gsecongeo.98.8.1683
lP

Ramos, V.A., 2008. The basement of the Central Andes: the Arequipa and related
terranes. Annu. Rev. Earth Planet. Sci. 36, 289–324.
https://doi.org/10.1146/annurev.earth.36.031207.124304
na

Ramos, V.A., 2010. The Grenville-age basement of the Andes. J. South Am. Earth Sci.
ur

29 (1), 77–91. https://doi.org/10.1016/j.jsames.2009.09.004

Reich, M., Parada, M.A., Palacios, C., Dietrich, A., Schultz, F., Lehmann, B., 2003.
Jo

Adakite-like signature of Late Miocene intrusions at the Los Pelambres giant


porphyry copper deposit in the Andes of central Chile: metallogenic
implications. Miner. Deposita 38 (7), 876–885.
https://doi.org/10.1007/s00126-003-0369-9

Rezeau, H., Jagoutz, O., 2020. The importance of H2O in arc magmas for the
formation of porphyry Cu deposits. Ore Geol. Rev. 126, 103744.
https://doi.org/10.1016/j.oregeorev.2020.103744

Richards, J.P., Kerrich, R., 2007. Special paper: adakite-like rocks: their diverse
origins and questionable role in metallogenesis. Econ. Geol. 102 (4), 537–576.
https://doi.org/10.2113/gsecongeo.102.4.537
Journal Pre-proof

Richards, J.P., 2011. High Sr/Y arc magmas and porphyry Cu±Mo±Au deposits: just
add water. Econ. Geol. 106 (7), 1075–1081.
https://doi.org/10.2113/econgeo.106.7.1075

Richards, J.P., Spell, T., Rameh, E., Razique, A., Fletcher, T., 2012. High Sr/Y
magmas reflect arc maturity, high magmatic water content, and porphyry
Cu±Mo±Au potential: examples from the Tethyan arcs of central and eastern
Iran and western Pakistan. Econ. Geol. 107 (2), 295–332.
https://doi.org/10.2113/econgeo.107.2.295

Richards, J.P., 2013. Giant ore deposits formed by optimal alignments and
combinations of geological processes. Nat. Geosci. 6 (11), 911–916.
https://doi.org/10.1038/ngeo1920

of
Richards, J.P., 2015. The oxidation state, and sulfur and Cu contents of arc magmas:

ro
implications for metallogeny. Lithos 233, 27–45.
https://doi.org/10.1016/j.lithos.2014.12.011
-p
Richards, J.P., López, G.P., Zhu, J.J., Creaser, R.A., Locock, A.J., Mumin, A.H., 2017.
re
Contrasting tectonic settings and sulfur contents of magmas associated with
Cretaceous porphyry Cu±Mo±Au and intrusion-related iron oxide Cu-Au
lP

deposits in northern Chile. Econ. Geol. 112 (2), 295–318.


https://doi.org/10.2113/econgeo.112.2.295
na

Richards, J.P., 2018. A Shake-Up in the Porphyry World? Econ. Geol. 113 (6), 1225–
1233. https://doi.org/10.5382/econgeo.2018.4589
ur

Rivera, M., Martin, H., Le Pennec, J.L., Thouret, J.C., Gourgaud, A., Gerbe, M.C.,
2017. Petro-geochemical constraints on the source and evolution of magmas at
Jo

El Misti volcano (Peru). Lithos 268, 240–259.


https://doi.org/10.1016/j.lithos.2016.11.009

Rollinson, H.R., 2014. Using geochemical data: evaluation, presentation,


interpretation. Routledge, Abingdon, London, p.

Roperch, P., Carlotto, V., Ruffet, G., Fornari, M., 2011. Tectonic rotations and
transcurrent deformation south of the Abancay deflection in the Andes of
southern Peru. Tectonics 30 (2). https://doi.org/10.1029/2010TC002725

Rötzler, J., Wilke, M., 2005. Micro-XANES on Fe-Mg minerals from granulites.
University of Potsdam. Potsdam, Germany, P. 1–2.
Journal Pre-proof

Ruprecht, P., Wörner, G., 2007. Variable regimes in magma systems documented in
plagioclase zoning patterns: El Misti stratovolcano and Andahua monogenetic
cones. J. Volcanol. Geotherm. Res. 165 (3), 142–162.
https://doi.org/10.1016/j.jvolgeores.2007.06.002

Sandeman, H.A., Clark, A.H., Farrar, E., 1995. An integrated tectono-magmatic


model for the evolution of the southern Peruvian Andes (13-20 °S) since 55
Ma. Int. Geol. Rev. 37 (12), 1039–1073.
https://doi.org/10.1080/00206819509465439

Santos, A., Guo, W., Rivera, F., Tassinari, C., Cerpa, L., Kojima, S., 2019. Early
Jurassic arc related magmatism associated with porphyry copper
mineralization at Zafranal, Southern Peru unraveled by zircon U-Pb ages.

of
Andean Geol., 445–470. https://doi.org/10.5027/andgeoV46n3-3041

ro
Schuessler, J.A., Botcharnikov, R.E., Behrens, H., Misiti, V., Freda, C., 2008.
Amorphous Materials: Properties, structure, and Durability†: Oxidation state
-p
of iron in hydrous phono-tephritic melts. Am. Miner. 93 (10), 1493–1504.
https://doi.org/10.2138/am.2008.2795
re

Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J., Arispe, O.,
lP

Néraudeau, D., Cárdenas, J., Rosas, S., 2002. Late Permian–Middle Jurassic
lithospheric thinning in Peru and Bolivia, and its bearing on Andean-age
tectonics. Tectonophysics 345 (1–4), 153–181.
na

https://doi.org/10.1016/S0040-1951(01)00211-6
ur

Shannon, R.D., 1976. Revised effective ionic radii and systematic studies of
interatomic distances in halides and chalcogenides. Acta Crystallogr., Sect. A
32 (5), 751–767. https://doi.org/10.1107/S0567739476001551
Jo

Shatwell, D., 2021. Mesozoic Metallogenesis of Peru: A Reality Check on


Geodynamic Models. SEG Discovery (124), 15–24.
https://doi.org/10.5382/SEGnews.2021-124.fea-01

Sillitoe, R.H., Perelló, J., 2005. Andean copper province: Tectonomagmatic settings,
deposit types, metallogeny, exploration, and discovery. Econ. Geol. 100th
Anniv. Volume, 845–890. https://doi.org/10.5382/AV100.26

Sillitoe, R.H., 2010. Porphyry copper systems. Econ. Geol. 105 (1), 3–41.
https://doi.org/10.2113/gsecongeo.105.1.3

Sillitoe, R.H., Mortensen, J.K., 2010. Longevity of porphyry copper formation at


Journal Pre-proof

Quellaveco, Peru. Econ. Geol. 105 (6), 1157–1162.


https://doi.org/10.2113/econgeo.105.6.1157

Simmons, A.T., 2013. Magmatic and hydrothermal stratigraphy of Paleocene and


Eocene porphyry Cu-Mo deposits in southern Peru University of British
Columbia, P.

Simmons, A.T., Tosdal, R.M., Wooden, J.L., Mattos, R., Concha, O., McCracken, S.,
Beale, T., 2013. Punctuated magmatism associated with porphyry Cu-Mo
formation in the Paleocene to Eocene of southern Peru. Econ. Geol. 108 (4),
625–639. https://doi.org/10.2113/econgeo.108.4.625

Sisson, T.W., Grove, T.L., 1993. Experimental investigations of the role of H2O in

of
calc-alkaline differentiation and subduction zone magmatism. Contrib.
Mineral. Petrol. 113 (2), 143–166. https://doi.org/10.1007/BF00283225

ro
Skewes, M.A., Stern, C.R., 1994. Tectonic trigger for the formation of late Miocene
-p
Cu-rich breccia pipes in the Andes of central Chile. Geology 22 (6), 551–554.
https://doi.org/10.1130/0091-7613(1994)022<0551:TTFTFO>2.3.CO;2
re

Søager, N., Holm, P.M., 2013. Melt–peridotite reactions in upwelling eclogite bodies:
lP

Constraints from EM1-type alkaline basalts in Payenia, Argentina. Chem.


Geol. 360, 204–219. https://doi.org/10.1016/j.chemgeo.2013.10.024
na

Søager, N., Holm, P.M., Llambías, E.J., 2013. Payenia volcanic province, southern
Mendoza, Argentina: OIB mantle upwelling in a backarc environment. Chem.
ur

Geol. 349, 36–53. https://doi.org/10.1016/j.chemgeo.2013.04.007

Sørensen, E.V., Holm, P.M., 2008. Petrological inferences on the evolution of magmas
Jo

erupted in the Andagua Valley, Peru (Central Volcanic Zone). J. Volcanol.


Geotherm. Res. 177 (2), 378–396.
https://doi.org/10.1016/j.jvolgeores.2008.05.021

Stegen, R.J., Barton, M.D., Waegli, J.A., 2018. Cerro Verde-Santa Rosa
copper-molybdenum deposits, Peru: magmatic, hydrothermal, and supergene
characteristics of two adjacent porphyry systems. In: A.M. Arribas R and J.L.
Mauk (Editors), Metals, minerals, and society. Society of Economic
Geologists, Special Publications, p293–319.
https://doi.org/10.5382/SP.21.13Stern, C.R., Skewes, M.A., 2005. Origin of
giant Miocene and Pliocene Cu-Mo deposits in central Chile: Role of ridge
subduction, decreased subduction angle, subduction erosion, crustal thickening
and long-lived, batholith sized, open-system magma chambers. In: Porter, T.M.
Journal Pre-proof

(Eds.), Super Porphyry Copper & Gold Deposits: A Global Perpective, PGC
Publishing, Adelaide, Australia, p. 65–82.

Stewart, J., Evernden, J., Snelling, N., 1974. Age determinations from Andean Peru: a
reconnaissance survey. Geol. Soc. Am. Bull. 85 (7), 1107–1116.
https://doi.org/10.1130/0016-7606(1974)85<1107:ADFAPA>2.0.CO;2

Streck, M.J., Dilles, J.H., 1998. Sulfur evolution of oxidized arc magmas as recorded
in apatite from a porphyry copper batholith. Geology 26 (6), 523–526.
https://doi.org/10.1130/0091-7613(1998)026<0523:SEOOAM>2.3.CO;2

Sun, S.S., McDonough, W.F., 1989. Chemical and isotopic systematics ofoceanic
basalts: Implications for mantle composition and processes. In: Saunders, A.D.,

of
Norry, M.J. (Eds.), Magmatism in the Ocean Basins, Geological Society
Special Publication, p. 313–345.

ro
https://doi.org/10.1144/GSL.SP.1989.042.01.1
-p
Sun, W.D., Huang, R.f., Li, H., Hu, Y.B., Zhang, C.C., Sun, S.J., Zhang, L.P., Ding, X.,
Li, C.Y., Zartman, R.E., 2015. Porphyry deposits and oxidized magmas. Ore
re
Geol. Rev. 65, 97–131. https://doi.org/10.1016/j.oregeorev.2014.09.004
lP

Takagi, D., Sato, H., Nakagawa, M., 2005. Experimental study of a low-alkali
tholeiite at 1–5 kbar: optimal condition for the crystallization of high-An
plagioclase in hydrous arc tholeiite. Contrib. Mineral. Petrol. 149 (5), 527–540.
na

https://doi.org/10.1007/s00410-005-0666-7
ur

Tang, M., Lee, C.-T.A., Ji, W.Q., Wang, R., Costin, G., 2020. Crustal thickening and
endogenic oxidation of magmatic sulfur. Sci. Adv. 6 (31), eaba6342.
https://doi.org/10.1126/sciadv.aba634
Jo

Thouret, J.C., Rivera, M., Wörner, G., Gerbe, M.C., Finizola, A., Fornari, M.,
Gonzales, K., 2005. Ubinas: the evolution of the historically most active
volcano in southern Peru. Bull. Volcanol. 67 (6), 557–589.
https://doi.org/10.1007/s00445-004-0396-0

Timms, N.E., Healy, D., Erickson, T.M., Nemchin, A.A., Pearce, M.A., Cavosie, A.J.,
2018. Role of elastic anisotropy in the development of deformation
microstructures in zircon. In: Moser, D.E., Corfu, F., Darling, J.R., Reddy,
S.M., Tait, K. (Eds.), Microstructural geochronology: Planetary records down
to atom scale, p. 183-202. https://doi.org/10.1002/9781119227250.ch8

Turner, S.J., Langmuir, C.H., 2015. What processes control the chemical
Journal Pre-proof

compositions of arc front stratovolcanoes? Geochem. Geophys. Geosyst. 16


(6), 1865–1893. https://doi.org/10.1002/2014GC005633

Ulmer, P., 1989. The dependence of the Fe2+-Mg cation-partitioning between olivine
and basaltic liquid on pressure, temperature and composition. Contrib. Mineral.
Petrol. 101 (3), 261–273. https://doi.org/10.1007/BF00375311

Ulmer, P., Kaegi, R., Müntener, O., 2018. Experimentally Derived Intermediate to
Silica-rich Arc Magmas by Fractional and Equilibrium Crystallization at
1·0 GPa: an Evaluation of Phase Relationships, Compositions, Liquid Lines of
Descent and Oxygen Fugacity. J. Petrol. 59 (1), 11–58.
https://doi.org/10.1093/petrology/egy017

of
Urann, B., Le Roux, V., Jagoutz, O., Müntener, O., Behn, M., Chin, E., 2022. High
water content of arc magmas recorded in cumulates from subduction zone

ro
lower crust. Nat. Geosci. 15, 1–8.
https://doi.org/10.1038/s41561-022-00947-w
-p
Valencia, M.M., Santisteban, A., Marchena, A.A., León, W.R., 2020. GE33A-5:
re
Características geológicas y geoquimicas de los depósitos minerales de la Faja
magmática Cretáceo-Paleógena entre Huancavelica y Tacna-Bloque Norte.
lP

https://repositorio.ingemmet.gob.pe/handle/20.500.12544/3519 (accessed 04
August 2022)
na

Viala, M., Hattori, K., 2021. Hualgayoc mining district, northern Peru: Testing the use
of zircon composition in exploration for porphyry-type deposits. J. Geochem.
ur

Explor. 223, 106725. https://doi.org/10.1016/j.gexplo.2021.106725

Wade, C.E., Payne, J.L., Barovich, K., Gilbert, S., Wade, B.P., Crowley, J.L., Reid, A.,
Jo

Jagodzinski, E.A., 2022. Zircon trace element geochemistry as an indicator of


magma fertility in iron oxide copper-gold provinces. Econ. Geol. 117 (3),
703–718. https://doi.org/10.5382/econgeo.4886

Wang, R., Richards, J.P., Hou, Z., Yang, Z., DuFrane, S.A., 2014a. Increased
magmatic water content—the key to Oligo-Miocene porphyry Cu-Mo ± Au
formation in the eastern Gangdese belt, Tibet. Econ. Geol. 109 (5), 1315–1339.
https://doi.org/10.2113/econgeo.109.5.1315

Wang, R., Richards, J.P., Hou, Z.Q., Yang, Z.M., Gou, Z.B., DuFrane, S.A., 2014b.
Increasing magmatic oxidation state from paleocene to miocene in the eastern
Gangdese Belt, Tibet: implication for collision-related porphyry Cu-Mo±Au
mineralization. Econ. Geol. 109 (7), 1943–1965.
Journal Pre-proof

https://doi.org/10.2113/econgeo.109.7.1943

Waters, L.E., Lange, R.A., 2016. No effect of H2O degassing on the oxidation state of
magmatic liquids. Earth Planet. Sci. Lett. 447, 48–59.
https://doi.org/10.1016/j.epsl.2016.04.030

Webster, R., Pitman, C., Zamora, C., 2013. Don Javier Cu-Mo project technical report.
prepared by AMC Mining Consultants Ltd. Canada, P. 140.

Wilke, M., Behrens, H., Burkhard, D.J.M., Rossano, S., 2002. The oxidation state of
iron in silicic melt at 500 MPa water pressure. Chem. Geol. 189 (1), 55–67.
https://doi.org/10.1016/S0009-2541(02)00042-6

of
Wilkinson, J.J., 2013. Triggers for the formation of porphyry ore deposits in magmatic
arcs. Nat. Geosci. 6 (11), 917–925. https://doi.org/10.1038/ngeo1940

ro
Workman, R.K., Hart, S.R., 2005. Major and trace element composition of the
-p
depleted MORB mantle (DMM). Earth Planet. Sci. Lett. 231 (1–2), 53–72.
https://doi.org/10.1016/j.epsl.2004.12.005
re

Wu, C., Chen, H., Lu, Y., 2021. Magmatic water content and crustal evolution control
lP

on porphyry systems: insights from the Central Asian Orogenic Belt. J. Petrol.
62 (2), egab021. https://doi.org/10.1093/petrology/egab021
na

Wu, C., Chen, H., Lu, Y., 2022. Crustal structure control on porphyry copper systems
in accretionary orogens: insights from Nd isotopic mapping in the Central
ur

Asian Orogenic Belt. Miner. Deposita 57 (4), 631–641.


https://doi.org/10.1007/s00126-021-01074-z
Jo

Xu, B., Hou, Z.Q., Griffin, W.L., Lu, Y.J., Belousova, E., Xu, J.F., O'Reilly, S.Y., 2021.
Recycled volatiles determine fertility of porphyry deposits in collisional
settings. Am. Miner. 106 (4), 656–661. https://doi.org/10.2138/am-2021-7714

Xu, T., Liu, X., Xiong, X., Wang, J., 2022. Sulfur dissolution capacity of highly
hydrated and fluid-saturated dacitic magmas at the lower crust and
implications for porphyry deposit formation. Geochim. Cosmochim. Acta.
https://doi.org/10.1016/j.gca.2022.07.004

Yaxley, G.M., Berry, A.J., Kamenetsky, V.S., Woodland, A.B., Golovin, A.V., 2012.
An oxygen fugacity profile through the Siberian Craton — Fe K-edge XANES
determinations of Fe3+/ ∑ Fe in garnets in peridotite xenoliths from the
Udachnaya East kimberlite. Lithos 140–141, 142–151.
Journal Pre-proof

https://doi.org/10.1016/j.lithos.2012.01.016

Ye, Z.C., Mao, J.W., Lu, M.J., Zhu, X.S., Chen, N., Wei, H.T., Jin, W.Q., Meng, X.Y.,
2022. Geology and geochronology of the Don Javier Cu-Mo porphyry deposit,
southern Peru. Ore Geol. Rev. 143, 104777.
https://doi.org/10.1016/j.oregeorev.2022.104777

Zellmer, G.F., Iizuka, Y., Miyoshi, M., Tamura, Y., Tatsumi, Y., 2012. Lower crustal
H2O controls on the formation of adakitic melts. Geology 40 (6), 487–490.
https://doi.org/10.1130/G32912.1

Zhang, J., Wang, R., Hong, J., 2022. Amphibole fractionation and its potential redox
effect on arc crust: Evidence from the Kohistan arc cumulates. Am. Miner.

of
Zhang, Y., 1999. H2O in rhyolitic glasses and melts: measurement, speciation,

ro
solubility, and diffusion. Rev. Geophys. 37 (4), 493–516.
https://doi.org/10.1029/1999RG900012
-p
Zhu, J.J., Richards, J.P., Rees, C., Creaser, R., DuFrane, S.A., Locock, A., Petrus, J.A.,
re
Lang, J., 2018. Elevated magmatic sulfur and chlorine contents in ore-forming
magmas at the Red Chris porphyry Cu-Au deposit, northern British Columbia,
lP

Canada. Econ. Geol. 113 (5), 1047–1075.


https://doi.org/10.5382/econgeo.2018.4581
na

Zimmermann, J.L., Collado, A.K., 1983. Détermination par la Méthode K/Ar de l'âge
des Intrusions et des Minéralisations Associées dans le Porphyre Cuprifère de
ur

Quellaveco (Sud Ouest du Pérou). Miner. Deposita 18 (2), 207–213.


https://doi.org/10.1007/BF00206209
Jo

Zou, X., Qin, K., Han, X., Li, G., Evans, N.J., Li, Z., Yang, W., 2019. Insight into
zircon REE oxy-barometers: A lattice strain model perspective. Earth Planet.
Sci. Lett. 506, 87–96. https://doi.org/10.1016/j.epsl.2018.10.031

Zweng, P.L., Clark, A.H., Pierce, F.W., Bolm, J.B., 1995. Hypogene evolution of the
Toquepala porphyry copper-molybdenum deposit, Moquegua, southeastern
Peru. In: Pierce, F.W., Bolm, J.G. (Eds.), Porphyry copper deposits of the
American Cordillera, Tucson, Arizona Geological Society, Digest 20, p. 566–
612.
Journal Pre-proof

Figure captions

Fig.1. Maps showing the (A) locations of San Felix Ridge at ca. 60 Ma and Inca

Oceanic Plateau at 12‒10 Ma, (B) locations of the Toquepala and Andahuaylas‒Anta

arcs, and (C) geology of the Paleocene‒Eocene southern Peru metallogenic belt. The

≤60 Ma Yarabamba Superunit granodiorite and porphyries are too small in area to be

shown in the map. Maps are modified from IGEMMENT (2015) and mineralization

of
ages are from Quang et al. (2003), Simmons (2013), Martínez et al. (2017), and Ye et

ro
al. (2022). Abbreviations: Hua = Huagayoc, IC = Inca Oceanic Plateau, PCC =
-p
Pliocene igneous rocks in central Chile, RB = Rio Blanco, SFR = San Felix Ridge
re

(inferred), SP-L = San Pedro‒Linzor volcanic chain, and YC = Yanacocha.


lP
na
ur

Fig. 2. (A) Schematic representation of the geological framework of southern Peru

(modified from Manrique, 2011). (B) Two stage geodynamic evolution of southern
Jo

Peru, which shows the inferred San Felix Ridge collision at ca. 60 Ma and the

subsequent flat-slab subduction at 52‒30 Ma (modified from Noury et al., 2017).

Fig. 3. Geochronological summary of the Toquepala arc in southern Peru.

Abbreviations: CJ = Cuajone, CV = Cerro Verde‒Santa Rosa, DJ = Don Javier, and


Journal Pre-proof

TP = Toquepala.

Fig. 4. Major element diagram for igneous rocks from the Toquepala arc (A) (Na2O +

K2O) vs. SiO2 (Le Bas et al., 1986) and (B) K2O vs. SiO2 (Rollinson, 2014).

of
ro
Fig. 5. Primitive mantle-normalized trace-element patterns for the pre- (A) and
-p
syn-orogenic (B) units, respectively. (C) and (D) are the chondrite-normalized REE
re
patterns for the pre- and syn-orogenic units, respectively. Primitive mantle and
lP

chondrite-normalized values are from Sun and McDonough (1989).


na
ur

Fig. 6. Zircon U‒Pb age vs. (A) Eu anomaly, (B) Sr/Y, (C) La/Yb, and (D) Dy/Yb
Jo

ratios observed in southern Peru igneous rocks. Crustal thickness that was correlated

with La/Yb ratio is from Haschke et al. (2002).

Fig. 7. Binary plots of SiO2 contents vs. Dy/Yb ratios with the modeled trends of

fractional crystallization. Cumulate assemblages of stages A and B are from Ulmer et

al. (2018), stage C from Nandedkar et al. (2014), and stage D from Takagi et al. (2005)
Journal Pre-proof

and Pietranik et al. (2009). Dy and Yb mineral‒melt partition coefficients of

plagioclase, orthopyroxene, clinopyroxene, and olivine are from Fujimaki et al.

(1984), garnet from Johnson (1994), high-pressure (crust-mantle boundary)

amphibole from Bottazzi et al. (1999) and middle-pressure (mid-lower crust)

amphibole from Nandedkar et al. (2016). The fO2 values were calculated using the

zircon oxybarometer of Loucks et al. (2020). Abbreviation: D = degree of fractional

of
crystallization.

ro
-p
re
Fig. 8. 143Nd/144Ndt vs. 87Sr/88Srt diagram. Regional isotopic data were compiled from
lP

143
Mamani et al. (2016). The Nd/144Ndt and 87
Sr/88Srt isotope values calculated at 60

Ma for rocks from >60 Ma, 14 Ma for Miocene, 5.3 Ma for Miocene‒Pliocene, 4 Ma
na

for Pliocene, 2 Ma for Pleistocene, and 1–0 Ma for Quaternary. GLOSS: global
ur

subducting sediment (Plank, 2014).


Jo

Fig. 9. Zircon trace-element ratios. (A) CeN/Ce* vs. EuN/Eu*, (B) Dy/Yb vs. Hf, (C)

Dy/Yb vs. Ti, (D) EuN/Eu* vs. Hf, and (E) Hf vs. Ti. Blue dashed lines show the

trends of the pre-orogenic unit and the yellow dashed lines show the trends of the

syn-orogenic unit.
Journal Pre-proof

Fig. 10. Zircon-based fO2 values vs. zircon U‒Pb age of the pre- and syn-orogenic

units, and the average apatite S contents vs. zircon U‒Pb age from Yarabamba

Superunit and Don Javier dacite porphyry (Chen et al., 2022a). The fO2 values

corresponding to melt S6+/ƩS ratios in basaltic melt are taken from Jugo et al. (2010).

of
Fig. 11. Semi-quantitative evaluation of Fe3+/ƩFe ratios in residual melt as a

ro
consequence of differentiation. Model 1: (A‒C) assumes constant Fe3+ and Fe2+
-p
mineral‒melt partition coefficients. Model 2: (D‒F) assumes constant Fe3+/ƩFe ratios
re
in fractionated minerals. (A) and (D) show the modeled results of initial magmas that
lP

underwent Fe3+/ƩFe ratio pre-elevation by high-pressure fractionation at the crust‒

mantle boundary (stages A and B in Fig. 7) and subsequent amphibole-dominated


na

fractionation at the mid‒lower crust (stage C in Fig. 7). (B) and (E) show the modeled
ur

results of initial magmas that underwent low-pressure (plagioclase-dominated)


Jo

fractionation. (C) and (F) show the modeled results of initial magmas that underwent

only amphibole-dominated fractionation. Fe3+/ƩFe ratios in garnet and amphibole are

from Zhang et al. (2022), in plagioclase from Lac (2009), in olivine from Dyar et al.

(1998), and in orthopyroxene and clinopyroxene from Nimis et al. (2015). The

Fe3+/ƩFe ratios and Fe contents in minerals were used to calculate the Fe 3+ and Fe2+

mineral‒melt partition coefficients at each stage.


Journal Pre-proof

Fig. 12. Schematic of the evolution of pre- and syn-orogenic magmas. (A)

Pre-orogenic magmas generated under normal subduction, producing magmas with

low oxidation state and low melt S content by plagioclase-dominated fractional

crystallization. (B) Syn-orogenic magmas generated under an arc compressional

regime, producing magmas with high oxidation state and high melt S content by

garnet-dominated fractionation at crust‒mantle boundary and subsequent

of
amphibole-dominated fractionation at mid-lower crust. Crustal thickening might

ro
occur but is insignificant during this period. Based on Chen et al. (2022a) and Zhang
-p
et al. (2022). Abbreviation: Amp = amphibole, Cpx = clinopyroxene, Gar = garnet,
re
Opx = orthopyroxene, Ol = olivine, and Pl = plagioclase.
lP
na

Fig. 13. Zircon-based fO2 values vs. apatite-based melt S content for representative
ur

porphyry Cu deposits. The calculation of fO2 values is based on the method of Loucks
Jo

et al. (2020) and that of melt S content is based on the method of Meng et al. (2021a).

Data sources: Haib from Meng et al. (2021a); Pulang from Li et al. (2019) and Cao et

al. (2021); Rio Blanco from Chen et al. (2022b); and Red Chris from Zhu et al.

(2018).
Journal Pre-proof

Tables

Table1. Porphyry deposits in the Toquepala arc in southern Peru

References: 1 = Acosta et al. (2020); 2 = Quang et al. (2003), 3 = Webster et al.

(2013), 4 = Ye et al. (2022), 5 = Simmons et al. (2013), 6 = Zweng et al. (1995), 7 =

Clark et al. (1990), 8 = Martínez et al. (2017), 9 = Shatwell (2021).

of
ro
-p
re
lP
na
ur
Jo
Journal Pre-proof

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.

of
ro
-p
re
lP
na
ur
Jo
Journal Pre-proof

Table 1. Porphyry deposits in the Toquepala arc

Resources
Deposit Age Referen
Ore Cu grade Mo grade Ag grade
(Ma) ces
(Mt) (wt.%) (wt.%) (g/t)
Cerro Verde‒Santa
Rosa ~61 4,048 0.37 0.014 1.5 1, 2
Don Javier ~60 303 0.43 0.017 2.6 3, 4
Quellaveco 58–54 2,904 0.46 0.015 / 1, 5
Toquepala 57–52 4,424 0.37 0.01–0.03 / 1, 6
Cuajone 56–52 2,252 0.48 0–0.02 / 1, 5, 7

of
Los Calatos ~55 352 0.76 0.03 / 1, 8, 9

References: 1 = Acosta et al. (2020); 2 = Quang et al. (2003), 3 = Webster et al.

ro
(2013), 4 = Ye et al. (2022), 5 = Simmons et al. (2013), 6 = Zweng et al. (1995), 7 =
-p
Clark et al. (1990), 8 = Martínez et al. (2017), 9 = Shatwell (2021). / = not given.
re
lP
na
ur
Jo
Journal Pre-proof

Abstract

The Paleocene–Eocene southern Peru metallogenic belt contains numerous very large

to supergiant porphyry Cu‒Mo deposits. The deposits were dated previously as coeval

with the Incaic I orogeny at ca. 60 Ma. However, tectono-magmatic processes during

the formation of the deposits are poorly constrained. Here, we integrate published

geochronological, geochemical, and mineralogical data from barren and ore-related

of
granitoids in the region. The syn-orogenic unit (60–53 Ma) has significantly higher

ro
average ratios of Sr/Y (103.5), La/Yb (29.5), Dy/Yb (1.9), and EuN/Eu* (1.07) than
-p
the pre-orogenic unit (69–60 Ma; Sr/Y = 18.6, La/Yb = 11.5, Dy/Yb = 1.7, and
re

EuN/Eu* = 0.66). The syn-orogenic unit exhibits high Dy/Yb ratios in relatively
lP

primitive rocks and has a lower average Dy/Yb(zircon) ratio (0.23) and higher
na

EuN/Eu*(zircon) ratio (0.43) than the pre-orogenic unit (Dy/Yb(zircon) = 0.29 and
ur

EuN/Eu*(zircon) = 0.21). These integrated data suggest that the syn-orogenic magmas

underwent high-pressure, garnet-dominated fractionation prior to


Jo

amphibole-dominated fractionation, distinct from the pre-orogenic magmas which

underwent low-pressure, plagioclase-dominated fractionation. Comprehensive

investigation of the geological background suggests that the geochemical variations

between pre- and syn-orogenic units were caused by the long-term evolution of

mantle-derived magmas at the crust‒mantle boundary during arc compression, instead

of crustal thickening. Fractionation of ferric iron-depleted minerals (e.g., garnet and


Journal Pre-proof

amphibole) in the mantle-derived basaltic magmas probably has positive effects on

magmatic fertility. The dataset presented highlights the critical role of high-pressure

differentiation, thus providing an insight into the relationship between the arc

compressional regime and giant porphyry Cu deposit formation in the Andes. Zircon

trace elements can be useful indicators for porphyry deposit exploration in the

Toquepala arc in southern Peru. However, the regional tectonic settings of the Andean

of
arcs should be studied thoroughly before applying these indicators.

ro
-p
re
lP
na
ur
Jo
Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8
Figure 9
Figure 10
Figure 11
Figure 12
Figure 13

You might also like