You are on page 1of 19

Gondwana Research 76 (2019) 303e321

Contents lists available at ScienceDirect

Gondwana Research
journal homepage: www.elsevier.com/locate/gr

Tectonic evolution of the southwestern margin of Pangea and its


global implications: Evidence from the mid PermianeTriassic
magmatism along the Chilean-Argentine border

Alvaro del Rey a, b, *, Katja Deckart a, c, Noah Planavsky d, Ce
sar Arriagada a,
a, e
Fernando Martínez
a
Departamento de Geología, Universidad de Chile, Santiago, Chile
b
Globe Institute, University of Copenhagen, Copenhagen, Denmark
c
Advanced Mining Technology Centre (AMTC), Universidad de Chile, Santiago, Chile
d
Department of Geology and Geophysics, Yale University, New Haven, CT, USA
e
Departamento de Ciencias Geolo gicas, Universidad Cato
lica del Norte, Antofagasta, Chile

a r t i c l e i n f o a b s t r a c t

Article history: The tectonic evolution of the southwestern margin of Pangea supercontinent is represented by the
Received 21 March 2018 extensive late PaleozoiceTriassic magmatism along the southwestern margin of South America, including
Received in revised form the Chilean Frontal Andes batholiths as part of the Choiyoi province. Several models have proposed
16 May 2019
cessation of subduction as the reason behind the vast amounts of felsic magmatism and apparent lack of
Accepted 17 May 2019
Available online 2 July 2019
typical arc magmas. Here, new U-Pb in zircon ages, and geochemical and isotope analyses (Rb-Sr, Sm-Nd,
Re-Os) indicate that mid PermianeTriassic granitic magmatism originated in a subduction-related exten-
Handling Editor: I. Safonova sional setting (slab rollback). Subduction and anatexis of lower continental crust were the main magma-
generation mechanisms, the latter caused by asthenospheric upwelling, decompression and subsequent
Keywords: accumulation of underplated basalts. A comparison with coeval igneous units along the Chilean-Argentine
PermianeTriassic magmatism border allows extension of this model from at least 21 to 40 S. The key elements triggering slab rollback are
Pangea low subduction plate velocities and convergence rates, which can be attributed to the assembly of Pangea
Andean subduction supercontinent (mid PermianeTriassic). Therefore, subduction of the oceanic plate beneath South America
Slab rollback
has been a continuous process from early Paleozoic times onwardsdrather than having a period without
Mass extinctions
subduction before the onset of the Andean cycle as previous models have invoked. New geochronological
constraints indicate that the peak of the voluminous crustal-derived magmatism and related explosive
volcanism (Choiyoi province) was contemporaneous with the emplacement of the Emeishan and Siberian
Traps LIPs, potentially conditioning the Earth system for the environmental collapse and biotic crises related
to those LIPs. The observed tectonic changes, magmatism and related environmental implications could
potentially be linked to the assembly of Pangea supercontinent.
© 2019 Published by Elsevier B.V. on behalf of International Association for Gondwana Research.

1. Introduction activity can be observed in the Chilean Precordillera (20 e24 S,


Munizaga et al. (2008)), the Cordillera de Domeyko (24 e27 S;
Key aspects of the late Paleozoic to Triassic magmatism along the Mpodozis et al. (1993)), the Chilean Frontal Andes (28 e31 S, also
southwestern margin of South America and how it relates to the known as Chilean Frontal Cordillera; e.g., Mpodozis and Kay, 1992;
Pangea assembly (completed in the earliest Permian, ca. 280 Ma, Li Pankhurst et al., 1996) and the Coastal Cordillera (33 e38 S, Deckart
and Powell, 2001), the Pangea supercontinent (from ca. 280 Ma until squez et al., 2011) in Chile. In Argentina, the
et al., 2014; Parada, 1990; Va
its break up ca. 200 Ma, Stampfli et al., 2013) and MesozoiceCenozoic active region occurs in the Argentine Frontal Cordillera (33 S, Gregori
‘Andean’ magmatism (late Lower Jurassicepresent, Charrier et al., and Benedini, 2013), the San Rafael Block (34 e36 S; Kleiman and
2014) are still poorly constrained. This protracted interval of igneous Japas, 2009; Rocha-Campos et al., 2011), and La Pampa province (38 S,
Kay et al., 1989). In addition, there is PermianeTriassic magmatism in
the North Patagonian Massif, Tierra del Fuego and Antarctic Peninsula
* Corresponding author at: Globe Institute, University of Copenhagen, Copen-
(approx. south 40 S, e.g., Pankhurst et al., 2006). Traditionally, this
hagen, Denmark magmatism has been divided into two groups: Carboniferouseearly/
 del Rey).
E-mail address: alvarodelrey@ign.ku.dk (A. late Permian subduction-related (e.g., Charrier et al., 2014; Mpodozis

https://doi.org/10.1016/j.gr.2019.05.007
1342-937X/© 2019 Published by Elsevier B.V. on behalf of International Association for Gondwana Research.
304  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

and Kay, 1992), and mid/late PermianeTriassic widely distributed and Carboniferous metabasites and metasediments overlaid by upper
voluminous felsic magmatism emplaced during important extensional 
CarboniferousePermian carbonates and clastic sediments (e.g., Alvarez
conditions (e.g., Kleiman and Japas, 2009; Llambias and Sato, 1990; et al., 2011).
Mpodozis and Kay, 1992). It has also been grouped into the Choiyoi Traditionally, the Elqui complex has been interpreted as subduction-
province (Kay et al., 1989), which at the same time can be divided into related magmatism emplaced into a normal to thickened continental
the older, Lower Choiyoi and younger, Upper Choiyoi, respectively (Kay crust, whereas the Ingagua s complex has been described as crustal-
et al., 1989; Kleiman and Japas, 2009; Sato et al., 2015). Nevertheless, derived generated during a period without subduction after a
despite the similarities that allow grouping this magmatism and the compressional episode (the San Rafael orogenic event, ca. 284 to
agreement on the vast extensional conditions particularly throughout 276 Ma) that resulted in extensional conditions during Permian and
Triassic times until earliest Jurassic (e.g., Charrier et al., 2014; Willner Triassic times (Llambias and Sato, 1990; Mpodozis and Kay, 1992).
et al., 2005), several different explanations have been given for the
extensional regime, including: rifting related to the first stage of Pangea 3. Sampling and analytical methods
breakup (Charrier, 1979), dextral strike-slip movements (Rapela and
Pankhurst, 1992), orogenic collapse and slab break-off after the accre- Thirty-three samples from the Montosa-El Potro batholith were
tion of a hypothetic terrane (north 33 S, Mpodozis and Kay, 1992), analysed. They correspond to several intrusive units named La Estan-
cessation of subduction together with dextral strike-slip movements n and El Colorado (Fig. 1). Thin-section
cilla, Montosa, Chollay, El Leo
parallel to the continental margin (Franzese and Spalletti, 2001), con- petrographic descriptions were made for each sample. Whole-rock
tinental dextral rotations with subsequent slab break-off and orogenic compositions were determined by Fusion-ICP (major elements plus Sc,
collapse (south 31 S, Kleiman and Japas, 2009), and slab steepening Be, V, Sr, Zr, Ba) and Fusion-ICPMS (trace elements) at Actlabs com-
and delamination of the lower continental crust related to a flat-slab panies, Canada. The results are accurate at a precision level of <9% (2s).
segment (Ramos and Folguera, 2009). Notably, various of the afore- Zircon (LA-ICPMS U-Pb) and magnetite (TIMS Re-Os) grains
mentioned models indicate cessation of subduction between latest were separated from 2 kg whole-rock samples (11 and 5 samples
Permian and earliest Jurassic time as the reason behind the extensional for zircon and magnetite, respectively) at the Sample Preparation
regime and apparent lack of typical arc magmas thus incompatible with Laboratory (University of Chile), through grinding, Gemini table,
work showing evidence of subduction during Triassic times (e.g., heavy liquids and Frantz separation, followed by a final hand
Brown, 1991; V asquez et al., 2011; Willner et al., 2005). picking under binocular microscope.
In addition, the timing of this voluminous felsic mid LA-ICPMS U-Pb zircon geochronology was undertaken in 11
PermianeTriassic magmatism coinciding with the biotic crisis representative samples of the different rock units in addition to 5
observed during the late Permian (e.g., Bond and Wignall, 2014), previously dated by SHRIMP II (del Rey et al., 2016). Analyses were
could potentially mean that this magmatic event (the Choiyoi carried out using a Thermo X-series Q-ICPMS coupled to a Reso-
province) contribute to the environmental collapse. netics Resolution M050 Excimer laser following the analytical
Here we present new results that bolster the case that subduction protocol after Solari et al. (2010). The laser ablation spots were
along the South American margin has been a continuous process since 23 mm. U-Pb reference grain and uncertainties were obtained as in
late Paleozoic. We provide new evidence that latest early Carboniferouse Paton et al. (2010). 207Pb/206Pb ratios, ages and errors were calcu-
mid Permian (ca. 325e270 Ma) magmatism occurred as subduction- lated and corrected according to Andersen (2002). Weighted mean
206
related during the Gondwanide Orogeny while Pangea was being Pb/238U ages have been calculated and the uncertainties are
assembled, and it was followed by mid PermianeTriassic (ca. 270e reported at 95% confidence limits. Analytical work was performed
210 Ma) subduction-related extensional conditions due to slab rollback at the Laboratory of Isotopic Studies (LEI) of the Geoscience Centre
and steepening as a result of important tectonic changes brought by the of the National Autonomous University of Mexico (UNAM).
assembly of Pangea supercontinent, as proposed by del Rey et al. (2016). Re-Os analyses were performed using Negative Thermal Ion Mass
Using new Rb-Sr, Sm-Nd and Re-Os isotopic data; LA-ICPMS U- Spectrometer (NTIMS-VG54 MS) following the standard procedure
Pb in zircon ages; and major and trace element whole-rock described in Creaser et al. (1991). Magnetite grains were loaded in
geochemical analysis of rock samples from the Montosa-El Potro Carius tubes (Shirey and Walker, 1995). Re and Os spikes and 16 ml of
batholith (Chilean Frontal Andes, 28 e28 300 S), together with a aqua regia were added and kept under cold conditions to avoid evap-
comparison with coeval late PaleozoiceTriassic plutonic rocks from oration and early chemical reaction, then sealed. Afterwards, the tubes
Chile and Argentina, we assess the general tectonic framework for were placed into an oven at 240  C overnight. Os was separated from
the contemporaneous magmatism, evaluate the causes of the vast the solution in a two-stage distillation process (N€agler and Frei, 1997),
magmatic event and indicate its environmental implications. then purified and loaded with Ba(OH)2 on a Pt filament. Re was sepa-
rated after Os by using the remaining acid solution. After drying, the
2. Geological setting solution was re-dissolved in 0.1 HNO3. Re was extracted and purified by
using AG1-X8 (100e200 mesh) resins. Subsequently, Re was loaded
The Chilean Frontal Andes (28 e31 S) possesses a series of batho- with Ba(SO)4 on Pt filaments. The analytical work was performed at the
liths that record most of the mid PermianeTriassic evolution of the Arizona Geochronology Centre of the University of Arizona.
southwestern margin of Pangea. From north to south, these batholiths Sr and Nd isotopic analyses were performed on 17 dated samples
are named Montosa-El Potro, Chollay and Elqui-Limarí; and according (11 in this work and 6 previously dated) using Thermo Neptune Plus
to their geochronological and geochemical characteristics, they have MC-ICPMS following the previous protocol of Go  mez-Tuena et al.
been classified as belonging to the Elqui and Ingagu as complexes (e.g., (2011). 200 mg of sample was used for separating the elements. Nd
Mpodozis and Kay, 1992) (Fig. 1). The Elqui complex (Carboniferouse isotopic compositions of samples and standards were measured in ca.
early Permian) is constituted by the Elqui and Montosa batholiths. 200 ppb solutions, and Sr isotopes were determined using solutions
Compositionally, they correspond to metaluminous to peraluminous, diluted to ca. 300 ppb concentrations. Analyses of samples and stan-
calc-alkaline and I- to S-type diorites, tonalites, granodiorites and dards consist of 70e100 static measurement cycles, with each cycle
granites (Mpodozis and Kay, 1992). The Ingagua s complex (Permiane measured in ca. 4 s integrations. Beam intensities of 5 V/ppm and
Triassic), which is included in the Choiyoi province (Kay et al., 1989), is 100 V/ppm were routinely achieved for 88Sr and 142Nd, respectively,
constituted by the Limarí, El Potro and Chollay batholiths, and they using wet plasma and a 100 ml/mm free aspirating nebulizer. The
correspond to hyper-siliceous, peraluminous, calc-alkaline, S- to A-type interference-corrected 87Sr/86Sr ratios were exponentially normalized
granitoids (Mpodozis and Kay, 1992). Host rocks include lower for mass fractionation to 86Sr/88Sr ¼ 0.1194 and corrected to a NIST
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 305

Fig. 1. Simplified geological map of the Chilean Frontal Andes batholiths. Late PaleozoiceTriassic intrusives, units belonging to the Elqui and Ingagua s complexes (after Mpodozis
and Kay, 1992), the location of the analysed samples, and Paleozoic, Mesozoic and Cenozoic sedimentary deposits are shown (after Martínez et al., 2015). The Montosa-El Potro
 n, El Colorado and Chollay units. In the last, official geological map (Martínez et al., 2015) units Montosa, El Leo
batholith is constituted by La Estancilla, Montosa, El Leo  n, El Colorado
and Chollay have been grouped into the plutonic complex ‘Complejo Pluto  nico Montosa’ (the same is applied for units studied in del Rey et al., 2016). Based on the new
geochronological results, the Montosa unit can be included into the Ingaguas complex and no longer into the Elqui complex. All samples have geochemical analyses. Only dated
samples have their identification shown in the map. Samples IC-17, -22, -58, -91 and -93 were SHRIMP U-Pb dated (del Rey et al., 2016), the rest was dated using LA-ICPMS. 1, Herve 
et al. (2014); 2, Maksaev et al. (2014); 3, Martínez et al. (2016).

SRM 987 (formerly NBS 987) standard ratio of 87Sr/86Sr ¼ 0.710230. potentially critical isobaric interference of Sm. The calculated
143
External reproducibility for the standard was 0.000016 (2s, n ¼ 10). Nd/144Nd ratios were exponentially normalized for mass bias to
146
Nd isotopic ratios were measured statically using the entire collector Nd/144Nd ¼ 0.72190, and then corrected to the JNdi standard value
array in order to monitor all relevant isotopic species and the of 143Nd/144Nd ¼ 0.512115. Analyses on the JNdi standard yielded an

Table 1
87
Summarized geochemical classification, U-Pb in zircon ages and initial eNdi, Sr/86Sri and gOsi values of the dated samples from the Montosa-El Potro batholith.

Sample Geographical coordinates Lithologya S-I-A type ASI (Eu/Eu*) (La/Yb)N 206
Pb/238U age (Ma) eNdi (87Sr/86Sr)i gOsi
La Estancilla
IC-17 28 000 13.8200 S 69 530 05.2200 W Hbl-bt tonalite I-type 0.93 0.89 7.6 286 ± 2b þ2.09 0.705279 e

Montosa
IC-22 28 050 03.4500 S 69 500 57.0100 W Hbl-bt tonalite I-type 0.97 0.66 9.0 253 ± 2b 2.13 0.703985 125e140
IC-23 28 050 03.0800 S 69 500 48.2100 W Bt syenogranite I-type, slightly S 1.03 0.83 28.0 264 ± 2 1.36 e 1056e1290
IC-99 28 050 17.0600 S 69 490 05.8700 W Bt-Hbl tonalite I-type 1.00 1.21 16.5 253 ± 2 0.51 0.704599 e
IC-102 28 050 49.9000 S 69 470 40.7000 W Bt porphyritic rhyolite I-type 1.02 0.61 9.5 252 ± 2 þ6.17 0.707463 e
IC-106 28 070 01.2300 S 69 450 44.4900 W Bt-Hbl granodiorite I-type, slightly S 1.06 0.53 14.9 244 ± 2 þ10.04 e e
RdM-09 28 150 11.2600 S 69 480 07.8600 W Bt syenogranite I-type 1.02 0.72 22.6 255 ± 1 2.62 0.707936 e

n
El Leo
IC-58 28 010 05.7400 S 69 320 18.9200 W Bt-Hbl monzogranite I-type 1.02 0.45 8.3 257 ± 2b 2.87 0.707902 447
IC-94 28 230 41.5900 S 69 470 40.9000 W Syenogranite I-type 1.04 1.14 5.5 243.9 ± 2.1c 2.60 0.708099 e

El Colorado
IC-14 28 130 53.0000 S 69 390 47.3200 W Bt monzogranite I-type 1.04 0.68 11.2 250 ± 2 2.09 0.705693 e
IC-44 28 030 20.8300 S 69 160 53.2300 W Hbl-bt porph. rhyolite A-type 1.03 0.11 3.9 261 ± 2 2.64 e e
IC-93 28 240 37.5300 S 69 470 10.9900 W Bt syenogranite I-type 1.02 0.58 9.2 248 ± 2b 0.92 0.706140 409e423

Chollay
IC-46 28 030 30.1200 S 69 180 54.1800 W Bt syenogranite I-type 1.03 0.19 2.2 248 ± 1 þ0.29 e e
IC-47 28 020 49.9900 S 69 180 37.1100 W Afs granite I-type, slightly S 1.03 0.40 7.9 245 ± 1 0.73 e e
IC-81 28 040 29.0500 S 69 210 39.1600 W Bt monzogranite I-type 0.98 0.43 12.0 242 ± 1 1.64 0.705923 e
IC-91 28 280 16.6900 S 69 430 25.1300 W Hbl-bt syenogranite I-type 0.98 0.62 8.5 259 ± 2b 12.17 0.705090 23e37
IC-92 28 270 33.0400 S 69 450 54.6000 W Porphyritic quartz latite I-type 1.00 0.45 7.8 216 ± 1 1.18 0.703301 e

Notes: Afs: alkali feldspar; Bt: biotite; Hbl: hornblende. Aluminium Saturation Index, ASI: A/CNK ¼ Al2O3/(CaO þ Na2O þ K2O) (molar). Detailed isotopic calculations for eNdi-
(87Sr/86Sr)i and gOsi are in Tables 3 and 4 respectively. i ¼ isotopic initial ratios calculated using the obtained U-Pb ages. (La/Yb)N, N: normalized to chondrite, McDonough and
Sun (1995).
a
Lithologies were obtained after petrographic descriptions (Supplementary material).
b
Age from del Rey et al. (2016).
c
Age from Martínez et al. (2016).
306  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

Fig. 2.1. Geochronological analyses. Tera Wasserburg Concordia and Weighted mean age plots of the Montosa unit samples: IC-23, IC-99, IC-102, IC-106, RdM-09. Location of
analysed samples is shown in Fig. 1.
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 307

Fig. 2.2. Geochronological analyses. Tera Wasserburg Concordia and Weighted mean age plots of El Colorado: IC-14, IC-44; and Chollay: IC-46, IC-47, IC-81, IC-92 units. Location of
analysed samples is shown in Fig. 1.
308
A.
 del Rey et al. / Gondwana Research 76 (2019) 303e321
Table 2
Geochemical data of the U-Pb dated samples from the Montosa-El Potro batholith.

Unit La Estancilla El Colorado n


El Leo Chollay Montosa

Sample IC-17 IC-14 IC-44 IC-93 IC-58 IC-94 IC-46 IC-47 IC-81 IC-91 IC-92 IC-22 IC-23 IC-99 IC-102 IC-106 RdM-9

wt.% d.l.
SiO2 0.01 59.87 68.19 76.78 73.54 71.23 76.19 78.45 76.19 72.07 68.80 75.38 68.82 78.31 68.87 70.88 67.25 78.46
Al2O3 0.01 16.46 15.69 12.18 13.40 14.10 12.63 12.03 12.11 13.73 14.71 12.70 14.88 11.85 15.06 13.07 14.64 11.96
Fe2O3(T) 0.01 6.68 3.35 1.68 1.80 3.18 0.69 1.00 1.57 2.91 3.76 2.44 4.20 0.97 4.31 2.40 4.25 1.05
MnO 0.001 0.10 0.07 0.04 0.03 0.06 0.01 0.02 0.04 0.06 0.06 0.02 0.06 0.01 0.08 0.06 0.09 0.03
MgO 0.01 3.12 0.95 0.16 0.38 0.88 0.07 0.03 0.19 0.55 0.98 0.13 1.27 0.14 1.33 0.63 1.33 0.13
CaO 0.01 6.14 2.86 0.33 1.16 2.26 0.67 0.30 0.87 1.46 2.58 0.71 3.44 0.68 3.81 1.68 2.44 0.83
Na2O 0.01 2.73 4.27 4.05 3.98 3.31 3.70 3.57 2.97 3.88 3.82 4.41 3.28 2.69 3.94 3.78 3.77 3.72
K2O 0.01 1.85 2.66 4.17 4.12 3.98 4.48 4.89 4.89 4.57 3.69 3.85 3.37 5.40 1.49 3.26 2.91 3.78
TiO2 0.001 0.90 0.48 0.13 0.22 0.36 0.11 0.10 0.19 0.37 0.48 0.25 0.49 0.12 0.50 0.30 0.55 0.11
P2O5 0.01 0.14 0.16 0.04 0.06 0.10 0.04 0.03 0.05 0.11 0.15 0.06 0.13 0.03 0.18 0.09 0.15 0.03
LOI 1.89 0.61 0.78 1.42 0.84 1.02 0.51 1.55 0.54 1.78 0.78 0.81 0.62 0.69 2.25 1.54 0.26
Total 99.87 99.28 100.30 100.10 100.30 99.62 100.90 100.60 100.20 100.80 100.70 100.70 100.80 100.30 98.39 98.93 100.40

ppm
Sc 1 17 6 2 3 7 3 2 3 5 8 4 10 3 8 5 12 1
Be 1 2 2 3 3 2 1 6 4 4 3 3 2 <1 2 2 2 1
V 5 156 51 11 26 52 11 9 13 29 66 20 76 14 74 32 84 13
Cr 20 30 < 20 < 20 < 20 20 < 20 < 20 < 20 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20
Co 1 16 4 <1 2 5 <1 <1 1 3 5 2 6 <1 6 3 7 <1
Ni 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20 < 20
Cu 10 20 < 10 < 10 10 < 10 < 10 < 10 < 10 < 10 < 10 10 < 10 < 10 < 10 < 10 < 10 < 10
Zn 30 60 < 30 60 < 30 40 < 30 < 30 30 30 40 < 30 < 30 < 30 60 < 30 40 < 30
Ga 1 17 17 17 14 15 12 15 14 18 16 17 16 11 17 14 17 12
Ge 0.5 1.7 1.7 1.1 2 1.9 1.8 2.2 1.8 2 1.8 1.6 1.7 1.9 1.5 1.4 1.9 1.7
As 5 11 <5 <5 <5 <5 <5 <5 <5 <5 7 <5 6 <5 <5 <5 <5 <5
Rb 1 85 89 194 126 168 132 347 239 185 133 81 157 106 57 117 85 69
Sr 2 297 247 30 115 199 64 27 40 140 235 55 245 86 303 65 338 37
Y 0.5 14.8 24.5 47.9 16.1 24.9 18 37.1 23.5 35.7 19.5 22.9 21.7 5.9 9.9 21 19.1 5.8
Zr 1 139 210 193 103 137 54 84 115 225 184 242 200 76 162 144 160 72
Nb 0.2 5.1 7.4 14.3 7.8 6.8 5.1 25.7 12.3 16 7.2 5.8 7.8 0.8 0.9 6 5.2 < 0.2
Mo 2 <2 <2 <2 <2 <2 <2 <2 <2 2 <2 <2 <2 <2 <2 <2 <2 <2
Ag 0.5 0.7 1.3 0.9 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 0.6 < 0.5 1 1.1 < 0.5 < 0.5 < 0.5 < 0.5 1.4
In 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1
Sn 1 1 1 2 1 2 <1 1 178 2 2 <1 2 <1 <1 <1 <1 <1
Sb 0.2 0.3 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 0.4 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2
Cs 0.1 3.8 1.7 5.3 1.9 4.1 1.9 5 3.8 4.1 3.8 1.1 8.3 1.5 1.9 2.4 2.3 0.7
Ba 3 323 818 201 657 793 846 101 365 559 727 683 846 1101 728 552 1176 218
La 0.05 17.9 42.1 30.6 27.2 34.7 16.6 18.2 34.7 69 29.7 33 31 33.7 28.1 32.7 45.1 19.9
Ce 0.05 35.6 77.8 62.2 52.1 68.3 32.7 49.1 65.9 133 63.2 64.7 62.6 60.6 49.9 63.6 84.8 38.5
Pr 0.01 4.1 8.76 8.06 5.8 7.64 3.72 4.95 7.47 14.3 7.51 7.3 7.1 6.22 5.21 7.11 8.87 4.33
Nd 0.05 15.7 31.8 30.5 19.4 26.6 12.9 17.6 25.7 47.4 27.2 25.9 25.6 19.3 17.4 25 30 14.5
Sm 0.01 3.28 6.22 7.29 3.57 5.36 2.73 4.31 4.76 7.96 4.71 4.89 4.87 2.68 2.7 4.44 4.99 2.38
Eu 0.005 0.937 1.28 0.271 0.628 0.883 0.388 0.259 0.586 1 0.886 0.674 0.982 0.617 0.964 0.826 1.09 0.363
Gd 0.01 3.02 4.96 7.26 2.86 4.45 2.48 4.09 3.87 5.93 3.72 4.03 4.04 1.68 2 3.63 3.99 1.64
Tb 0.01 0.48 0.83 1.41 0.5 0.75 0.47 0.91 0.69 1 0.62 0.7 0.68 0.22 0.31 0.64 0.61 0.21
Dy 0.01 2.72 4.63 8.38 2.79 4.2 2.96 5.85 4.05 5.9 3.5 4.02 3.83 1.15 1.74 3.66 3.44 1.05
Ho 0.01 0.55 0.89 1.82 0.56 0.9 0.62 1.35 0.86 1.2 0.72 0.86 0.76 0.22 0.35 0.73 0.69 0.21
Er 0.01 1.5 2.59 5.02 1.67 2.6 1.82 4.25 2.53 3.52 2.09 2.62 2.2 0.63 1.02 2.11 1.98 0.59
Tm 0.005 0.254 0.411 0.871 0.303 0.45 0.319 0.796 0.46 0.628 0.368 0.434 0.367 0.104 0.186 0.361 0.304 0.085
Yb 0.01 1.61 2.56 5.37 2.01 2.86 2.05 5.75 3.01 3.91 2.39 2.9 2.36 0.82 1.16 2.36 2.07 0.6
Lu 0.002 0.226 0.338 0.736 0.308 0.423 0.291 0.809 0.429 0.572 0.349 0.431 0.329 0.128 0.177 0.338 0.291 0.092
Hf 0.1 3.2 4.7 6.1 2.9 3.9 1.8 3.3 3.5 5.6 4.5 6 4.6 2.7 3.8 4.1 4.1 1.8

A.
 del Rey et al. / Gondwana Research 76 (2019) 303e321
Ta 0.01 0.6 0.93 1.81 1.2 1.12 0.84 5.26 1.77 2.01 0.95 0.85 1.21 0.34 0.25 0.89 0.52 0.3
W 0.5 < 0.5 < 0.5 < 0.5 < 0.5 7.3 < 0.5 0.9 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5 < 0.5
Tl 0.05 0.55 0.57 2.62 0.72 1.16 0.84 2.48 2.73 1.24 0.93 0.88 0.86 0.61 0.34 0.61 0.45 0.57
Pb 5 12 11 16 10 12 14 19 13 17 17 <5 11 12 7 8 13 16
Bi 0.1 0.2 < 0.1 < 0.1 0.2 0.6 0.3 0.3 0.9 < 0.1 < 0.1 < 0.1 0.1 < 0.1 < 0.1 < 0.1 < 0.1 < 0.1
Th 0.05 8.93 10 25.2 11.9 16.9 8.17 38.7 19.7 24.9 16.6 10.3 18.8 10.6 4.49 12.6 9.94 16.1
U 0.01 2.09 1.42 4.65 2.31 3.8 1.51 6.49 3.01 3.9 2.03 2.65 3.86 1.49 1.3 2.32 1.85 0.98

Notes: Major element oxides and trace elements were analysed by ACTLabs, Canada
Mayor elements plus Sc, Be, V, Sr, Zr and Ba were analysed using FUS-ICP and trace elements using FUS-ICPMS
Total iron as Fe2O3
d.l. ¼ detection limit

309
310  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

Fig. 3. Classification diagrams. a. A/CNK vs. A/NK diagram indicating that samples are divided between metaluminous and peraluminous. El Leo n and El Colorado units are strictly
peraluminous, and La Estancilla, metaluminous. b. Peccerillo and Taylor (1976) diagram displaying a calc-alkaline to high-K calc-alkaline trend typical of continental magmatic arcs.
Units Montosa and La Estancilla can be defined as calc-alkaline, whereas El Leo n, Chollay and El Colorado as high-K calc-alkaline. c. 10,000  Ga/Al vs. Ce þ Nb þ Zr þ Y granite
discrimination diagram (Whalen et al., 1987). Only few samples from El Leo n, Chollay and El Colorado plot in the A-type granites area.

Fig. 4. Trace element diagrams. Primitive mantle-normalized multielement variation diagrams of trace elements (a, b, c), and chondrite-normalized REE plots (d, e, f) (after
 n (257e244 Ma), El Colorado (261e248 Ma), and
McDonough and Sun, 1995) of units from the Montosa-El Potro batholith: La Estancilla (ca. 286 Ma), Montosa (264e244 Ma), El Leo
Chollay (259e216 Ma).
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 311

Table 3
Sm-Nd and Rb-Sr isotopic data of the U-Pb dated samples from the Montosa-El Potro batholith.
147
Sample SiO2 Age Sm Nd Sm/144Nd (143Nd/144Nd)today (143Nd/144Nd)i CHURi eNdi Rb Sr 87
Rb/86Sr (87Sr/86Sr)today (87Sr/86Sr)i
(wt%) (Ma) (ppm) (ppm) (ppm) (ppm)

La Estancilla
IC-17 59.87 286a 3.28 15.7 0.126 0.512612 0.512377 0.512270 þ2.09 85 297 0.8281 0.708645 0.705279

Montosa
IC-22 68.82 253a 4.87 25.6 0.115 0.512393 0.512203 0.512312 2.13 157 245 1.8543 0.710667 0.703985
IC-23 78.31 264 2.68 19.3 0.084 0.512372 0.512227 0.512296 1.36 106 86 3.5665 e e
IC-99 68.87 253 2.7 17.4 0.093 0.512440 0.512284 0.512310 0.51 57 303 0.5443 0.706574 0.704599
IC-102 70.88 252 4.44 25.0 0.107 0.512806 0.512630 0.512314 þ6.17 117 65 5.2085 0.726127 0.707463
IC-106 67.25 244 4.99 30.0 0.100 0.512993 0.512827 0.512313 þ10.04 85 338 0.7277 e e
RdM-09 78.46 255 2.38 14.5 0.099 0.512340 0.512176 0.512310 2.62 69 37 5.3962 0.727496 0.707936

n
El Leo
IC-58 71.23 257a 5.36 26.6 0.121 0.512364 0.512160 0.512307 2.87 168 199 2.4428 0.716826 0.707902
IC-94 76.19 244b 2.73 12.9 0.127 0.512394 0.512189 0.512322 2.60 132 64 5.9681 0.728907 0.708099

El Colorado
IC-14 68.19 250 6.22 31.8 0.118 0.512403 0.512213 0.512319 2.09 89 247 1.0426 0.709363 0.705693
IC-44 76.78 261 7.29 30.5 0.144 0.512415 0.512176 0.512311 2.64 194 30 18.7119 e e
IC-93 73.54 248a 3.57 19.4 0.111 0.512451 0.512271 0.512319 0.92 126 115 3.1704 0.717329 0.706140

Chollay
IC-46 78.45 248 4.31 27.2 0.104 0.512573 0.512334 0.512319 þ0.29 347 27 37.1881 e e
IC-47 76.19 245 4.76 17.6 0.147 0.512464 0.512285 0.512323 0.73 239 40 17.2893 e e
IC-81 72.07 242 7.96 47.4 0.101 0.512404 0.512246 0.512330 1.64 185 140 3.8237 0.718922 0.705923
IC-91 68.80 259a 4.71 25.7 0.112 0.511858 0.511682 0.512305 12.17 133 235 1.6377 0.711113 0.705090
IC-92 75.38 216 4.89 25.9 0.114 0.512461 0.512300 0.512360 1.18 81 55 4.2615 0.716368 0.703301

Notes: The decay constant used in the calculation is l147Sm ¼ 6.54  1012 year1 and l87Rb ¼ 1.42  1011 year1 (Steiger and Ja €ger, 1977). eNd values were calculated
relative to a chondrite isotopic initial ratios using present day values (143Nd/144Nd)todayCHUR ¼ 0.512638; (143Sm/144Nd)todayCHUR ¼ 0.1967 (Jacobsen and Wasserburg, 1980).
i ¼ isotopic initial ratios calculated using the obtained U-Pb ages. For samples IC-44, -46, -47, -23, and -106 (87Sr/86Sr)i values were unable to obtain due to extremely low Sr
values relative to Rb.
a
Age from del Rey et al. (2016).
b
Age from Martínez et al. (2016).

external reproducibility of 0.000010 (2s, n ¼ 8). Sr and Nd analytical 4.1. U-Pb in zircon geochronology
work was performed at the Laboratory of Isotopic Studies (LEI) of the
Geoscience Centre of the National Autonomous University of Mexico New U-Pb dates ranging from 264 ± 2 to 244 ± 2 Ma for Mon-
(UNAM). tosa; from 248 ± 1 to 216 ± 1 Ma for Chollay; and from 261 ± 2 to
245 ± 2 Ma for El Colorado were obtained (Table 1, Figs. 2.1, 2.2). The
analysed samples represent crystallization ages from late Permian
4. Results
to late Triassic. Overall, the Montosa-El Potro batholith (Fig. 1) was
emplaced between early Permian and late Triassic.
Isotopic and geochronological results, including previously pub-
lished ages together with sample locations, are summarized in Table 1.
Tera Wasserburg Concordia and Weighted mean age plots are given 4.2. Whole-rock geochemistry
for each new dated sample (Figs. 2.1 and 2.2). Full geochronological,
major and trace element data and petrographic descriptions are given Geochemical data of the dated samples is summarized in Table 2.
in the Supplementary material. Absolute ages were interpreted using Geochemical analyses of all samples are given in Supplementary
the IUGS International Chronostratigraphic Chart v2018/08. material.

Table 4
Re-Os in magnetite isotopic data of some of the U-Pb dated samples from the Montosa-El Potro batholith.
187 187 187
Magnetite grain Age (Ma) Re (ppb) Re (ppb) Os (ppb) Os (ppb) Re/188Os 187
Os/188Os (187Os/188Os)i (gOs)i

Montosa
IC-22-mt-2 253a 0.187 0.117 0.043 0.002 21.60 0.376 0.285 125
IC-22-mt-3 253a 0.117 0.073 0.050 0.002 11.64 0.352 0.302 140
IC-23-mt-1 264 0.308 0.193 0.037 0.007 48.64 1.674 1.459 1056
IC-23-mt-2 264 0.170 0.107 0.032 0.006 31.41 1.893 1.755 1290

n
El Leo
IC-58-mt-2 257a 0.915 0.573 0.046 0.006 108.92 1.158 0.691 447

El Colorado
IC-93-mt-2 248a 2.361 1.478 0.520 0.046 23.71 0.741 0.642 409
IC-93-mt-3 248a 1.684 1.054 0.521 0.046 16.85 0.730 0.660 423

Chollay
IC-91-mt-2 259a 0.556 0.348 0.224 0.006 12.12 0.207 0.155 23
IC-91-mt-3 259a 0.472 0.296 0.168 0.005 13.76 0.232 0.173 37

Notes: The decay constant used in the calculation is l187Re ¼ 1.666  1011 year1 (Smoliar et al., 1996). gOs values were calculated relative to primitive mantle peridotites
using present day values (187Re/188Os)todayPM ¼ 0.3924; (187Os/188Os)todayPM ¼ 0.1262 (Walker et al., 2002). i ¼ isotopic initial ratios calculated using the obtained U-Pb ages.
a
Age from del Rey et al. (2016).
312  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

Fig. 5. Isotopic composition diagrams. a. Sr-Nd initial isotope composition diagram of the units from the Montosa-El Potro batholith. A comparison is made with coeval units along
the Chilean border (ca. 28 e34 S). Late Carboniferousemid Permian units (La Estancilla, Elqui-Limarí Batholith and Santo Domingo complex) apparently show a trend from higher
initial 87Sr/86Sr ratios (Santo Domingo complex)dindicating higher upper crustal involvementdtowards less evolved signatures (La Estancilla), which overlap with part of the mid
PermianeTriassic units. In general terms, a trend from higher upper crustal involvement towards a more mantle-like can be seen from late Carboniferous to Jurassic. b. Re/Os ratio
versus Os initial isotope composition for the units from the Montosa-El Potro batholith, plus a variety of terrestrial rocks data (Blusztajn et al., 2000) for comparison. 1, Parada et al.
(1999); 2, Gonz alez et al. (2017); 3, Mpodozis and Kay (1992); 4, Deckart et al. (2010).

The rocks analysed here represent compositions ranging from Saturation Index, ASI ¼ 0.93) and I-type (Fig. 3). In primitive mantle-
diorite to granite. Samples are calc-alkaline to high-K calc-alkaline, normalized multi-element diagrams it shows no Ti anomaly, negative
metaluminous to weakly peraluminous, and mostly present clear Ba and P, Nb-Ta troughs, and a positive Pb anomaly. It has relatively
I-type characteristics with some samples showing transition to flat REE patterns (ca. La/YbN ¼ 7.6) and a slight negative Eu anomaly
A-type and others with slight S-type tendencies (e.g., Chappell and (Eu/Eu* ¼ 0.89) (Fig. 4).
White, 2001) (Fig. 3). Multi-element variation diagrams of trace el- The Montosa unit (mid Permianemid Triassic) shows the widest
ements and REE, normalized to primitive mantle and chondrite compositional range ranging from diorite through granodiorite to
(McDonough and Sun, 1995) respectively, display relatively similar granite (SiO2 ¼ 56e78%). All diorites and some granodiorites are
patterns for all the early Permian-late Triassic intrusive rocks (Fig. 4). metaluminous (ASI between 0.87 and 1.00), whereas all granites and
The rocks are enriched in incompatible elements, showing negative most granodiorites are weakly peraluminous (ASI between 1.02 and
Sr, P, Ti, Ba, Nb and Ta anomalies, and positive Pb. Most samples have 1.06). Samples are calc-alkaline and I-type with some showing slight S-
negative Eu anomalies and relatively flat REE patterns. type characteristics (Fig. 3). According to the new ages obtained,
The oldest sample, corresponding to the La Estancilla unit geochemical variability occurred mostly during late Permian and early
(286 ± 2 Ma, early Permian, del Rey et al., 2016), is of dioritic Triassic. Multi-element diagrams indicate negative Sr, P, Ti and Nb-Ta
composition (SiO2 ¼ 59.9%), calc-alkaline, metaluminous (Aluminun troughs, and positive Pb. The unit presents two general groups of REE

Fig. 6. Tectonic discrimination diagrams. a. Y-Nb-3Ga diagram (Eby, 1992) showing that all the A-type samples from the Montosa-El Potro batholith plot in the volcanic arc and
post-orogenic tectonic setting (A2-type, crustal derived). b. Tectonic discrimination Y þ Nb vs. Rb and Y vs. Nb diagrams (Pearce et al., 1984). Most samples plot in the volcanic arc
granite field (La Estancilla, Montosa and El Leo  n units) and some show transitional features towards within-plate granites (Chollay and El Colorado units). VAG: volcanic arc
granites; WPG: within-plate granites; syn-COLG: syn-collision granites; ORG: ocean ridge granites.
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 313

patterns: relatively flat ((La/Yb)N ¼ 6.0e14.9) and steep ((La/ mass extinction in Earth’s history (Burgess and Bowring, 2015).
Yb)N ¼ 16.5e52.1) (Fig. 4). Typically, they do not have very pronounced Nevertheless, protracted magmatism from latest Mississippian to
negative Eu anomalies (Eu/Eu* ¼ 0.53e0.89, with an average of 0.82) Upper Triassic (328.1 ± 2.8e215.6 ± 1.9 Ma) has been described for
with some samples showing neutral or slightly positive values (Eu/ the Chilean Frontal Andes (Maksaev et al., 2014).
Eu* ¼ 0.92e1.21). Geochemically and isotopically, all the studied samples present
The El Leo n unit (late Permianemid Triassic) is granitic clear subduction-related arc magma characteristics, i.e., negative
(SiO2 ¼ 71e76%), high-K calc-alkaline and weakly (ASI between 1.02 Nb-Ta and positive Pb anomalies (normalized to primitive mantle)
and 1.10) to strongly (ASI ¼ 1.14) peraluminous. Most samples are (e.g., Pearce and Peate, 1995); and occurred mainly from a lower
I-type with one showing slight A-type characteristics and another continental crust source with sometimes more direct mantle in-
S-type affinity (Fig. 3). Multi-element diagrams show negative Sr, fluence and lesser upper continent contributions (Fig. 5).
P and Ti anomalies, somewhat negative Ba, positive Pb and Nb-Ta
troughs. It has relatively flat REE patterns ((La/Yb)N ¼ 5.5e15.1) and 5.1.1. La Estancilla
fairly negative Eu anomalies (Eu/Eu* ¼ 0.44e0.63) (Fig. 4). This unit corresponds to the oldest analysed sample (286 ± 2 Ma,
The Chollay unit (late Permianemid Triassic) presents granitic del Rey et al., 2016) It shows subduction-related arc magma signatures
composition (SiO2 ¼ 69e78%), and is high-K calc-alkaline and metal- together with no major fractionation of plagioclase, thus indicating
uminous to weakly peraluminous (ASI between 0.94 and 1.03). Samples possibly greater depths of the source (emplaced into normal thickness
are mostly I-type with lesser A-type, and one with slight S-type char- crust). The latter evidenced in a distinct trace element composition
acteristics (Fig. 3). Regarding the multi-element diagrams, the unit with absence of Eu, Sr and Ti anomalies. This sample is also one of the
presents negative Sr, P, Ti and Ba anomalies, Nb-Ta troughs and positive least evolved with dioritic composition and eNdi ¼ þ2.1 and
87
Pb. It has relatively flat REE patterns ((La/Yb)N ¼ 2.2e12.0) and pro- Sr/86Sri ¼ 0.705 (Fig. 5a).
nounced negative Eu anomalies (Eu/Eu* ¼ 0.12e0.62) (Fig. 4).
The El Colorado unit (mid Permianeearly Triassic) presents a 5.1.2. Montosa, El Leo n, Chollay and El Colorado units
granitic to granodioritic composition and is high-K calc-alkaline All four units to the east of the La Estancilla carry subduction-
and weakly peraluminous (ASI between 1.02 and 1.08). Samples are related arc magma features, however they display a wider composi-
I- and A-type with one showing slight S-type tendencies (Fig. 3). tional range: diorite to granite and I-type with minor S-, and A-type
Multi-element diagrams show negative Ba, Sr, P and Ti anomalies, affinities. The similar age of the analysed samples means that it is not
Nb-Ta troughs and positive Pb. It has relatively flat REE patterns possible to differentiate or separate successive events with regard to
((La/Yb)N ¼ 3.9e11.2) and pronounced negative Eu anomalies (Eu/ major geochemical characteristics, but they represent the peak of the
Eu* ¼ 0.11e0.68) (Fig. 4). magmatic event between late Permian and earliest Triassic and pre-
cisely correlate with the Upper Choiyoi (e.g., Kleiman and Japas, 2009).
4.3. Whole-rock Rb-Sr and Sm-Nd isotopic data The Montosa unit displays a less prominent role of plagioclase
fractionation (less pronounced negative Sr and Ti anomalies) and
All samples analysed for 87Sr/86Sr and 143Nd/144Nd (12 and 17, higher pressures, meaning melts generated under a thicker crust (high
respectively) have been dated (Table 1). Initial eNdi values and LREE, (La/Yb)N ¼ 16.5 to 52.1, and no Eu anomaly). In some samples,
Sr-isotope ratios were calculated at the given U-Pb zircon age of each slightly concave REE patterns support amphibole fractionation. Mon-
sample. The obtained ratios range from þ9.9 to 12.2 for eNdi, and tosa also presents the strongest negative Nb-Ta anomalies, indicating
from 0.703 to 0.708 for 87Sr/86Sri (Table 3). The La Estancilla unit subduction zone magma affinities probably with a source from the
presents eNdi ¼ þ2.1 and 87Sr/86Sri ¼ 0.705, indicating a less evolved shallower parts of the subduction zone (i.e., closer to the trench, Baier
magma source compared to the rest of the units. El Colorado shows et al., 2008). Lower peraluminosity and less enrichment in incompat-
eNdi values between 2.6 and 0.9 and 87Sr/86Sr ¼ 0.706; El Leon ible elements (U, Th and Pb) indicate less magma differentiation
eNdi between 2.9 and 2.6 and 87Sr/86Sri ¼ 0.708; Chollay eNdi compared to the other units. More positive eNdi values and low-to-
between 12.2 and þ0.3 and 86Sr/87Sri values between 0.703 and moderate 87Sr/86Sri (0.704e0.708) (Fig. 5a) imply minor continental
0.706; and Montosa eNdi between 2.6 to þ10, and 87Sr/86SrI be- contamination. This unit is also I-type, which together with its weak
tween 0.7004 and 0.708. A lower continental crust signature pre- peraluminous character, indicates that some of its magmas derived
dominates during early Permianelate Triassic. from a metaluminous source during partial melting of the crust at
pressures below the garnet stability field (Chappell et al., 2012).
4.4. Re-Os in magnetite isotopic data  n, El Colorado and Chollay units exhibit fractionation of
El Leo
plagioclase or its retention in the source, suggesting low pressure
The calculated initial isotope compositions expressed as gOsi
melting and differentiation in a garnet-free source, possibly under
values are: Montosa, þ125 to þ1290; El Leo  n, þ447; Chollay, þ23
thinned crust. This is evidenced by relatively flat REE patterns ((La/
to þ37; and El Colorado, þ409 to þ423 (Table 4). There is an
Yb)N ¼ 2.2e15.1), and strong negative Eu anomalies together with
apparent trend of increasingly lower gOsi as the rocks get younger.
negative Sr, Ba (especially El Colorado and Chollay) and Ti anomalies.
5. Discussion These units are mostly weakly peraluminous (apart from some
metaluminous rocks from Chollay) and present enrichment in
5.1. Timing and source of the Frontal Andes magmatism incompatible elements (mainly U, Th and Pb) displaying crustal
contributions and/or high amounts of magma differentiation by
The new geochronological results show that the rocks were fractional crystallization. Compared with the La Estancilla and Mon-
emplaced in magmatic pulses represented by different recognisable tosa, these units show higher K content (they are high-K calc-alkaline
episodes, with a peak between Guadalupian (late Permian) and rather than calc-alkaline plutons) and higher enrichment in trace
earliest Triassic, and minor activity during the Cisuralian (early elements. Chollay and El Colorado are the most enriched, followed by
Permian) and Upper Triassic (Table 1). The most pronounced El Leo  n. Negative eNdi and moderate 87Sr/86Sri values for all these
magmatic pulse is coeval with the upper section of the voluminous samples evidence a lower continental crust signature (Fig. 5a), and
felsic Choiyoi province (ca. 265e248 Ma, Upper Choiyoi, Kleiman eNdi close to zero (Table 3) indicates minimal contribution from
and Japas, 2009), and indicates that there was extensive magma- evolved crustal material (e.g., Alasino et al., 2012).
tism and related explosive volcanism before, during and in the El Colorado and Chollay units show stronger transitional to A-type
recovery stage of the end Permian biotic crisis, the most severe characteristics suggesting emplacement in an extensional regime, in
314  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

Fig. 7. Paleogeographic model. Schematic cross-sections through Chilean-Argentine Frontal Andes (at. ca. current 28 S) during mid PermianeTriassic subduction-related slab
rollback extensional setting with increasing thinning of continental crust and slab steepening towards the late Triassic. West-to-east geochemical zoning indicates decreasing
crustal thickness eastwards, in the same manner with increasing magma differentiation and A-type granitoids in the backarc region. Towards the west, the closer to the shallower
part of the subducting slab is evidenced by stronger negative Nb-Ta anomalies, I-type magmas and trace element patterns indicating thicker crust thus defining the corresponding
arc. Mid Permian-Triassic magmas are due to anatexis of mafic to intermediate lower continental crust and subduction-derived. I-, S-, and A-type magmas are all from mid
PermianeTriassic (lilac coloured). Early Permian La Estancilla unit magmatism is represented by the latest early Carboniferousemid Permian group (orange coloured), according to
the model proposed by del Rey et al. (2016). The red line in the globe (modified after del Rey et al., 2016) indicates the approximate paleo-location of the cross-section. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

either continental backarc, post-collisional extension or within-plate Montosa with the least negative. Regarding trace element enrich-
scenarios (e.g., Eby, 1992; Whalen et al., 1996). They mostly corre- ment, a similar pattern can be seen: the further inland, the higher
spond to magmas derived from melting of continental crust with or overall enrichment in trace elements (Fig. 4).
without mantle input (A2-type, Eby, 1992) (Fig. 6a). gOsi values indicate that between late Permian and earliest
A geographical distinction of the contemporaneous samples can Triassic a mixed magma source can be described, with some
be made. According to Fig. 1, the Montosa unit corresponds to the samples showing higher crustal contributions (e.g., Montosa, El
western part of the batholith whereas El Leo n, Chollay and El Col- n and El Colorado) and others more mantle-like affinities (e.g.,
Leo
orado units are to the east. In fact, only the Montosa unit presents Chollay) (Table 4). Lower gOsi values, such as in Chollay
very steep REE patterns, which together with the clear absence of (gOsi ¼ 23e37, 259 ± 2 Ma) imply a more direct mantle influence
Eu anomalies indicates a westward increase in crustal thickness or source (Johnson et al., 1996), whereas higher values, as
(i.e. towards the trench) whereas the rest of the units, an eastward observed in the Montosa unit (gOsi ¼ 1056e1290, 264 ± 2 Ma),
thinning of the crust (flat REE patterns and negative Eu anomalies). indicate crustal contributions and/or fractional crystallization in
However, a lower pressure magma source is most probably also absence of mantle influence (e.g., Hart et al., 2002; Johnson et al.,
represented in the Montosa unit (evidenced by the Montosa sam- 1996).
ples with flatter REE patterns), possibly indicating magma genesis
in different levels of the continental crust. The west-to-east 5.2. Tectonic setting and evolution
geochemical zoning can also explain the broader geochemical
variations and the strongest negative Nb-Ta anomalies seen in this  n units
The tectonic setting of the La Estancilla, Montosa and El Leo
unit, i.e., Montosa was emplaced in the shallower part of the sub- can be described as volcanic arc granites whereas Chollay and El Col-
ducting slab. El Colorado and Chollay showing A-type characteris- orado units as a combination between volcanic arc and within plate
tics indicate that this signature increases eastwards whereas I-type granites according to Pearce et al. (1984) (Fig. 6b). This means that
westwards. Also, these units present the most negative Eu anom- volcanic arc magmatism occurred from early Permian to at least late
alies, followed by El Leo n with intermediate values and then Triassic, when the youngest dated sample (Chollay unit, 216 ± 1 Ma)
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 315

Fig. 8. Geographical distribution of the late Paleozoic-Triassic igneous rocks from the Chilean-Argentine border and tectonic discrimination diagrams. Areas are the following
studies: 1, Brown (1991); 2, this work; 3, Mpodozis and Kay (1992); 4, Parada et al. (1999); 5, Vasquez et al. (2011); 6, Gregori and Benedini (2013); and 7, Kleiman and Japas (2009).
8, Charrier et al. (2014). Units are separated according to the 270 Ma limit defined by del Rey et al. (2016) (latest early Carboniferousemid Permian and mid PermianeTriassic) and
then plotted into the Y þ Nb vs. Rb discrimination diagram by Pearce et al. (1984) (a and b plots). Mid PermianeTriassic units from this work are plotted in Fig. 6b. Mpodozis and Kay
(1992) samples are not in the diagram because they lacked Y and Nb values, but they are considered the southern prolongation of Montosa-El Potro batholith. a. Pre-mid Permian
(325e270 Ma) units plot almost exclusively in the VAG field, indicating subduction related magmatism in a typical arc setting. b. Post-mid Permian (270e210 Ma) units show results
from VAG towards WPG, correspondingly with the mid PermianeTriassic units herewith studied. In the same manner, some units are predominantly VAG (1 and 4 in the map),
whereas others show intermediate features between VAG and WPG (5, 6 and 7), which may be due to its relative paleogeographical position in the subduction-related slab rollback
extensional setting as seen in Fig. 7. South of 31 S, westward displacement of the Coastal Batholith relative to the Frontal Andes (seen in map) has been attributed to Mesozoic and
Cenozoic extension (e.g. Brown, 1991; Pankhurst et al., 1988; Parada et al., 2007).

was emplaced. The latter being a typical metaluminous, calc-alkaline, REE patterns, indicating that the thickness of the continental crust
I-type, subduction-related intrusive. varied laterally, i.e., progressive lithospheric thinning eastwards,
The overlapping emplacement episodes of the intrusive units, allowing incorporation of more juvenile material in the backarc
volcanic arc and within plate mixed scenario and the presence of A- region. Higher trace element and K enrichment eastwards could also
type intrusives during the main peak of magmatism allow inferring mean greater magma differentiation in this direction.
a tectonic setting where rocks were emplaced in high temperature All analysed A-type granites are peraluminous, meaning that they
conditions (presence of A-type granites, Whalen et al. (1987)) into a can be interpreted as being generated by partial melting fractionation
thin continental crust (overall relatively flat REE patterns and from source compositions such as lower crustal (compositionally
negative Eu anomalies) with volcanic arc activity. These observa- equivalent to I-type sources) but with limited availability of H2O and
tions are compatible with a tectonic scenario of a convergent margin relatively low oxygen fugacity at high temperatures (King et al., 1997).
with subduction and predominant extensional conditions, sug- The high temperatures required may have been produced by
gesting thinned continental crust and high temperatures. In addi- asthenospheric upwelling or mafic magma influx in a localized area,
tion, Montosa presents exclusively volcanic arc granitoids whereas perhaps separated vertically from the other granite source regions in
the eastern units increasing within-plate tendencies, in the same the crust. These tectonic implications are coherent with extensional
manner as the presence of A-type intrusives and the changes in the conditions and partial melting of lower continental crust with variable
316  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

degrees of mantle influence. In addition, Andean I-type rocks have Chica complexes (ca. 292e270 Ma, Brown, 1991). In Argentina they
been explained by multiple episodes of partial melting of a basaltic include the Cerro Punta Blanca stock (ca. 276 Ma, Gregori and Bene-
underplate accompanied by crystal and/or restite fractionation from dini, 2013) and the early stage of the Lower Choiyoi volcanic province
the resulting magmas (Chappell and Stephens, 1988). Thus, the tec- (ca. 281 Ma, Kleiman and Japas, 2009).
tonic scenario for the Montosa-El Potro batholith can be envisaged as These units are predominantly metaluminous, calc-alkaline to
magmatism generated by partial melting of a basaltic underplate and a high-K calc-alkaline and mostly I- and S-types. According to the tec-
metaluminous lower crust in a subduction-related extensional setting. tonic discrimination diagrams (Pearce et al., 1984) (Fig. 8a), all units fall
The proposed geodynamic conditions are in agreement with the within the VAG field (in the same way as the only sample older than
model proposed by del Rey et al. (2016), where extension occurred 270 Ma from this work, the La Estancilla pluton, 286 ± 2 Ma, Fig. 6b). In
inboard from a subduction zone in the backarc region as a result of particular, the Guanta unit is interpreted as subduction-related gran-
slab rollback. The foundering slab creates a gravitational potential itoids (Mpodozis and Kay, 1992), the Santo Domingo complex as
well into which the orogen collapses (Lister and Forster, 2009), orogenic magmatism (Parada et al., 1999), and Cifuncho, Pan de Azúcar
creating the extensional conditions. As the hinge retreats, strong and Don ~ a Ine
s Chica complexes as subduction-related magmatism
crustal thinning causes upwelling of the asthenosphere followed by (Brown, 1991). The Cerro Punta Blanca stock is also subduction-related
underplating basalts and subsequent anatexis of the lower conti- (Gregori and Benedini, 2013) and Lower Choiyoi volcanics are inter-
nental crust to form granitic melts (Collins, 2002), with simulta- preted as arc-related magmatism (Kleiman and Japas, 2009). The
neous subduction-related magmatism (Fig. 7). granitoid units are predominantly granodiorite with minor tonalite
This scenario explains why stronger subduction-related signatures and granite.
are seen in the west at the same time as more mantle-like transitional The similarities among all these units are also in agreement with
magmatism inland. A similar setting has been described for early the model proposed by del Rey et al. (2016), reinforcing the idea
Carboniferous magmatism in the Eastern Sierras Pampeanas, that early/late Carboniferousemid Permian magmatism is sub-
Argentina, where A-type granitoid magmatism inland evolved towards duction-related and formed a continuous belt between at least
I-type to the west through time (27 300 e30 S, Alasino et al., 2012). 21 e40 S (considering also contemporaneous units between 21
del Rey et al. (2016) presented zircon Hf- and O-isotope data that and 26 S, e.g., Maksaev et al. (2014) and the southernmost limit of
showed rocks younger than ca. 270 Ma have mantle-like signatures, but the Coastal Batholith, Deckart et al., 2014).
because of their granitic nature, it was suggested that the obtained
signature was inherited from mafic igneous rocks without the 5.3.2. Post-mid Permian
involvement of new mantle material (Pietranik et al., 2013, and refer- In Chile, mid PermianeTriassic units include the following: El Col-
ences therein). This work geochemical data indicates highly evolved orado, Chollay, El Leo n, Montosa units (this work); the Ingagu as
magmatism (i.e., mostly granites), further supporting the idea that most complex (Mpodozis and Kay, 1992), which equivalent in Argentina
mid PermianeTriassic rocks were formed due to melting of mafic corresponds to the Colangüil batholith (leucocratic granodiorites and
continental crust; thus, inheriting d18O mantle-like signatures. Because granites, Llambias and Sato, 1990); the Limarí complex (203 Ma, Parada
of this, anatexis of a mafic crust and fractional crystallization are the et al., 1999), Triassic granitoids of the Coastal Cordillera between 34
main processes producing highly evolved granitic suits rather than and 37 S (ca. 224e197 Ma, Va squez et al., 2011); and the Chan ~ aral,
melting or contamination/assimilation of felsic, upper continental crust, Cerros del Vetado, Corral del Alambre, Pedernales and La Ola com-
with only minor participation of it. Considering that lower continental plexes (ca. 257e212 Ma, Brown, 1991). Equivalent units in Argentina
crust can be described as of mafic to intermediate composition (e.g., are Cerro Bayo and Cerro Punta Negra stocks (ca. 262 and 234 Ma
Hacker et al., 2015; Rudnick and Fountain, 1995), most peraluminous respectively, Gregori and Benedini, 2013), and the later (Upper) stage of
granitoids originated by crustal anatexis of the lower continental the Choiyoi province (intrusive units 265 Ma, Kleiman and Japas, 2009).
crustdin agreement with their Sr-Nd signaturesdreaching the needed These suits are predominantly weakly peraluminous, calc-
high temperatures by underplating. Furthermore, highly evolved calc- alkaline to high-K calc-alkaline and I- and S-type with clear tran-
alkaline granites (such as most of the late PermianeTriassic units) are sitional signatures to A-type. The discrimination diagrams (Figs. 6b,
interpreted as products of melting of biotite ± hornblende bearing in- 8b) shows mixed VAG and WPG affinities, with some units being
termediate composition rocks in the mid to lower crust (e.g. Whitney, mostly WPG (e.g., El Colorado, Cerro Punta Negra). One in partic-
1988). If such rocks have a metaluminous character (e.g., Chappell et al., ular, the Cerro Bayo stock (262 Ma, Gregori and Benedini, 2013),
2012), they could be the required protolith for most of the weakly shows clear transitional signatures between VAG and WPG.
peraluminous I-type granites observed in the Frontal Andes batholiths. As with the rocks studied here, mid PermianeTriassic magmatism
presents higher amounts of variability, reflected in contemporaneous
5.3. Contemporaneous magmatism along the southwestern margin I-, S- and A-type granotoids emplaced in both VAG and WPG settings.
of South America The Ingagua s complex is interpreted due to anatexis of a thinner crust
(Mpodozis and Kay, 1992), the Limarí complex as crust-derived (Parada
Considering previous published data from Chile and Argentina et al., 1999), the Triassic granitoids of the Coastal Cordillera as
(Fig. 8), a comparison has been made with that and the model pro- emplaced in a extensional subduction-related setting with mantle and
posed here. Thus, the ca. 270 Ma boundary (del Rey et al., 2016) has crustal sources (Va squez et al., 2011), the Chan
~ aral, Cerros del Vetado,
been used to separate units according to the hypothesis. Permian Corral del Alambre, Pedernales and La Ola complexes as subduction-
granites south 40 S are not used for comparison because of their related magmatism (Brown, 1991), the Cerro Bayo and Cerro Punta
overall different tectonic setting (e.g., mid Carboniferous collisional Negra stocks as the result of a transition between subduction-related
history, Pankhurst et al., 2006). and intraplate extension-related magmatism (Gregori and Benedini,
2013), and the Upper Choiyoi as transitional to intraplate post-orogenic
5.3.1. Pre-mid Permian extension-related magmatism (Kleiman and Japas, 2009). In addition,
Igneous units older than ca. 270 Ma (and younger than early all the available data of A-type rocks from this period (i.e., in the upper
Pennsylvanian) correspond to the following: in Chile, the La Estancilla Choiyoi (Kleiman and Japas, 2009) and Chilean Triassic granitoids
(ca. 286 Ma, del Rey et al., 2016); the Guanta unit of the Elqui-Limarí (Vasquez and Franz, 2008; V asquez et al., 2011)) can be described as
batholith (ca. 310e297 Ma, Mpodozis and Kay, 1992; 286 ± 2 Ma, aluminous A-type, suggesting the same tectonic implications for such
Pankhurst et al., 1996), the Santo Domingo complex (308 Ma, Parada rocks in the Chilean Frontal Andes (28 e28 300 S), the Coastal Cordil-
et al., 1999) and rocks from Cifuncho, Pan de Azúcar and Don ~ a Ine
s lera (34 e37 S) and the Choiyoi province (34 e36 S, Argentina).
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 317

Fig. 9. Summary of the tectonic evolution of the southwestern margin of Pangea, supercontinent cycles, large igneous provinces (LIPs) and biotic crises. Geodynamic changes
brought by the assembly of Pangea (and its breakup) coincide temporarily with the emplacement of important magmatic events across the globe and their related mass extinctions:
End-Guadalupian Mass Extinction (EGME, ca. 260 Ma, Bond and Wignall, 2014), End-Permian Mass Extinction (EPME, ca. 252 Ma, Burgess et al., 2014) and End-Triassic Mass
Extinction (ETME, ca. 202 Ma, Blackburn et al., 2013). CAMP: Central Atlantic Magmatic Province. Purple dotted line within the Choiyoi event represents the period of important
crustal anatexis. Plutons of each schematic section are coloured representing igneous unit as in Fig. 8. Apart from the LIPs, protracted arc magmatism occurred from Carboniferous
and throughout Triassic times (Maksaev et al., 2014). The San Rafael event is considered part of the Gondwanide orogeny and its beginning is denoted with the dotted black line. (For
interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)

As before, the described characteristics are coherent with the PermianeTriassic magmatism but also part of the previous subduc-
proposed model in which mid/late PermianeTriassic magmatism tion-related magmatism further inland (del Rey et al., 2016), e.g., the
occurred under extensional conditions with important contributions Cerro Bayo stock and the lower Choiyoi volcanics.
of melting of lower continental crust in a subduction setting. Observed
bimodal magmatism in this period (e.g., Kleiman and Japas, 2009; 5.4. Pangea supercontinent: how global tectonics played a key role
Maksaev et al., 2014; V asquez et al., 2011) is also coherent with strong in the changes of magmatism
slab rollback extensional conditions, and explains the presence of
mafic mantle-derived units such as Triassic gabbroic stocks at ca. 36 S The most important tectonic change discussed above in the late
(Vasquez et al., 2011). Unlike the pre-mid Permian period, this mag- Paleozoic regional magmatism is the shift from normal to thickened
matism occurred in an easterly position, possibly linked to the crust during latest early Carboniferousemid Permian to thinned
development of several NWeSE to NNWeSSE basins (e.g., Charrier continental crust with increasing and widespread extensional condi-
et al., 2014), including the Domeyko Basin (between 21 and 27 S in tions during Triassic times. Throughout this entire period, subduction-
northern Chile), which is interpreted as a subduction-related rift basin related signatures can be seen in almost all samples (e.g., Nb-Ta
in a backarc position during the late Triassic (Espinoza et al., 2018). In troughs and positive Pb anomaly as in the upper Choiyoi (Kleiman and
addition, the post ca. 270 Ma boundary proposed is in agreement with Japas, 2009) and Nb-Ta troughs in the Cerro Bayo stock (Gregori and
the PermianeTriassic thermal anomaly (between 265 and 215 Ma) Benedini, 2013)), implying that subduction was a continuous process
defined by Riel et al. (2018), a period of important crustal magmatic from at least latest early Carboniferous and throughout Triassic times.
activity based on the compilation of available data all along the Pangea In addition, Willner et al. (2005) demonstrated continuous accretion
margin facing the Panthalassa ocean (>30.000 km). as a result of long-term basal underplating in the accretionary prism
In a similar way to the latest early Carboniferousemid Permian of central Chile during the interval 320e224 Ma. After this time, the
magmatism, these units formed a continuous marginal belt that was late Paleozoic accretional system was inactive despite continuing
progressively displaced eastwards south of 31 S, possibly due to slab subduction at the convergent margin (Willner et al., 2005). This model
shallowing and final flat-slab subduction at ca. 36 S (del Rey et al., would explain why VAG signatures are dominant during the entire
2016; Kleiman and Japas, 2009). This process not only displaced late period of time.
318  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

Considering all gathered geochemical and isotope data, the possibly subduction plate velocities triggering important rollback
general tectonic evolution can be described as follows: latest early extension in the backarc region, as predicted by the models of Schellart
Carboniferousemid Permian subduction-related magmatism was (2008, 2005). Moreover, the work of Gorczyk et al. (2007), a petro-
emplaced into a normal to thickened continental crust and ac- logical-thermomechanical model, also indicates that under low
quired the input of continental material. Mid PermianeTriassic convergence rates an extensional regime can be expected, comprising
subduction-related magmatism was emplaced in a progressively trench retreat, backarc spreading and high melt productivity. On the
thinned crust as consequence of persistent extensional conditions contrary, before the mid PermianeTriassic period, higher convergence
due to slab rollback and steepening. This generated magmatism rates led to the development of an accretionary prism (Gorczyk et al.,
with within-plate affinities in a possible backarc region when 2007), into which latest early Carboniferousemid Permian magmatism
normal arc magmas were being emplaced farther west (Fig. 7). was emplaced (Willner et al., 2005). Accretion began shortly before ca.
South of 31 S Permian magmatism is unproven in Chile (but pre- 320 Ma (Willner et al., 2005), approximately at the onset of the Pangea
sent in Argentina, Fig. 8). This could be explained by an early and assembly (Li and Powell, 2001). Towards the end Triassic/early Jurassic,
progressive shallowing of the slab that displaced magmatism and a relative increase in the convergent rates together with subduction
extension eastwards (del Rey et al., 2016). The same magmatic shift velocities along the western margin of Pangeadand possibly in rela-
(from volcanic arc to within plate) can be seen in the Argentine Frontal tion to the supercontinent breakupdstopped slab rollback extension
Andes (Cordo  n del Portillo, ca. 33 e34 S, Gregori and Benedini (2013)) and related magmatism, allowing the transition to the well-known
and the San Rafael Massif (34 e36 S, Kleiman and Japas, 2009). Flat- Andean cycle from earliest Jurassic onwards (e.g., Charrier et al., 2014)
slab subduction was developed (at least at 31 e32 S) during most of (Fig. 9). The major geodynamic process controlling magmatism along
the early Permian (ca. 300e290 Ma, García-Sansegundo et al., 2014), the southwestern margin of South America is and has been the sub-
reinforcing the idea that a flat or very shallow segment from 31 to duction of the oceanic plate (Panthalassa/Pacific ocean) beneath Pan-
40 S hindered early Permian magmatism in Chile. Furthermore, the gea/Gondwana/South America since early Carboniferous times (Fig. 9),
possible subduction of an ocean ridge (or plume-influenced ridge) with no evidence for arrested subduction during the Triassic Period as
between 34 e39 S at ca. 300 Ma (Hyppolito et al., 2014) could explain suggested by some of the aforementioned models. Therefore, we
the final complete flat slab at ca. 36 e37 S (del Rey et al., 2016; Klei- propose that the tectonic differences between the defined Pre-Andean
man and Japas, 2009). Less voluminous Chilean late Triassic magma- (Lopingian (late Permian)elate Lower Jurassic) and Andean cycle (late
tism between 34 and 37 S (compared to mid PermianeTriassic Lower Jurassicepresent, Charrier et al., 2014) are convergence rates
magmatism in the Chilean Frontal Andes) has also been interpreted as and subduction plate velocities linked to the assembly of Pangea su-
subduction-related magmatism emplaced in a extensional setting percontinent, which resulted in important changes of the subduction
(Vasquez et al., 2011). Mid Permianeearly Triassic extension occurred plate geometry (Fig. 9); thus, the type of magmatism. These constitute
mostly inland (in Argentine territory, Fig. 8) and increased until it the key factors controlling the structural and magmatic changes.
reached Chilean territory towards the end of the Triassic, likely in as-
sociation with a progressively steeper slab, which allowed the 5.5. Tectonics, volcanism and mass extinctions
emplacement of the late Triassic extension-related units in southern
Chile (e.g., Vasquez et al., 2011). This series of events depicts the The prolonged mid PermianeTriassic magmatism and associated
evolving geometry of the subducting slab as the one described for explosive volcanism along the Chilean-Argentine border, included into
delamination processes (e.g., Kay and Coira, 2009), suggesting it could the Choiyoi event (ca. 286e247 Ma, Sato et al., 2015) and analogous
be involved in the origin of the mid PermianeTriassic magmas. Rapid magmatism in central Antarctica as part of the magmatic belt along
steepening of the shallower or flat slab segment would lead to the southwestern margin of Pangea (e.g, Nelson and Cottle, 2019)
delamination and mantle and crustal melting, producing widespread contributed to the greenhouse conditions during Permian and Triassic
volcanism and large local ignimbrites (Kay et al., 2013; Kay and Coira, times (e.g., Nelson and Cottle, 2019; Spalletti and Limarino, 2017), in
2009), in this case represented by the Upper Choiyoi in Argentina. agreement with arc volcanism as the main driver of atmospheric CO2
Upper Choiyoi rhyolites with extremely high La/Yb ratios (Kleiman and (McKenzie et al., 2016). Moreover, given the potential of significant
Japas, 2009) could point to a gravitationally unstable thickened lower carbonate rocks, organic-rich shales and peat as older sedimentary
crust. This idea is analogous to the model previously proposed by formations being heated and intruded by this geographically extensive
Ramos and Folguera (2009), where slab steepening after a period of magmatism thus increasing the release of CO2 (Spalletti and Limarino,
flat slab allowed the emplacement of the Choiyoi province. Progressive 2017) in addition to the volcanic outgassing; the emplacement of this
and significant steepening of the slab towards the end-of the Triassic magmatic event dramatically increased CO2 fluxes to the atmosphere
along the entire margin (21 e40 S) would help to explain the west- conditioning the global climate during mid Permian and Triassic times.
ward jump of early-to-late Jurassic magmatism to the Chilean Coastal The contemporaneous but abrupt emplacement of the large
Range (Charrier et al., 2014) (Figs. 8, 9). igneous provinces (LIPs) Emeishan (ca. 260e257 Ma Bond and
The geochemical, isotopic and geochronological characteristics of Wignall, 2014) and the Siberian Traps (ca. 252e250 Ma, Burgess et al.,
all the units allow identification of a continuous late PaleozoiceTriassic 2017) (Fig. 9) contribute to brief but dramatic spikes of CO2 release in
magmatic belt from 21 S to 40 S, which at the same time can be an already warmer, CO2-loaded world (e.g., Nelson and Cottle, 2019),
encompassed to the belt formed by the subduction of the Panthalassa thus potentially achieving a tipping point of climate stress that ended
oceanic plate under Pangea (Riel et al., 2018). The hypothesis presented up in environmental crisis (global warming, ocean anoxia, ocean
here provides a unifying tectonic framework for the magmatism. acidification) and mass extinctions (Bond and Wignall, 2014). Simi-
Extension in a subduction setting due to slab rollback and steepening larly, the emplacement of the Central Atlantic Magmatic Province
explains the observed continuity of subduction during the Triassic and (CAMP) and its related abrupt and important CO2 release at the end
the backarc location of transitional to within-plate magmatism. Also, Triassic (ca. 202 Ma)dunder prevailing greenhouse conditions due to
in highly extended regions widespread occurrence of high tempera- the previous events plus the prolonged Choiyoi magmatismdhas also
ture metamorphism, migmatisation and anatexis is expected (Lister been related to environmental crisis and mass extinction (Blackburn
and Forster, 2009) thus explaining the origin of the aluminous A-type et al., 2013). Therefore, we suggest that the Choiyoi event promoted
granites. On a more global scale, the Pangea supercontinent was fully the anomalously prolonged and long-lived extreme greenhouse gas
assembled during mid PermianeTriassic times (ca. 280e200 Ma, state observed from mid Permian and Triassic times, with a peak
Fig. 9) and in a possible static reference mode (Vilas and Valencio, between the end Permian and earliest Triassic. This, however, does not
1978). This potentially greatly decreased convergence rates and mean a causal relationship between the Choiyoi event and the mass
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 319

extinctions; we are simply suggesting that this magmatism contrib- iii. Subduction of the oceanic plate beneath South America has
uted to the observed environmental crises. In addition, this peak of been a continuous process from early Carboniferous and
magmatism corresponds to the main period of crustal anatexis throughout TriassiceJurassic times until present day. There is
observed along the southwestern margin of Pangea (ca. 265e245 Ma, no period without subduction between the end of Pangea as-
Kleiman and Japas (2009), this work), and overlaps with the sembly and the beginning of the Andean subduction cycle. On
emplacement of the Emeishan and Siberian Traps LIPs (Bond and the contrary, there is a fundamental change in the subducting
Wignall, 2014) (Fig. 9), thus implying that the tectonic drivers were plate geometry. As a result, the evolution of the subduction in
likely global. In this regard, it has been suggested that asthenospheric the convergent margin, from a compressional episode during
flow linked to the geodynamics of the paleo-Tethys domain (including late Paleozoic to progressively shallower southwards (with
asthenosphere upwelling under a subduction setting) during the as- development of a flat slab segment), followed by important
sembly of Pangea can explain the emplacement of several LIPs, extension, slab steepening (probably in association with
including the Emeishan and Siberian Traps LIPs (eastern Pangea, Wang delamination) and retreat of the arc towards late Triassic/
et al., 2015), comparably to the herewith described extensional con- earliest Jurassic, is similar to the idealized model for the Andean
ditions (slab rollback) and crustal anatexis promoted by the assembly cycle (Ramos, 2009), yet the mid PermianeTriassic magmatism
of the supercontinent (Panthalassa domain, western Pangea). In both is unexpectedly different compared to the typical Andean one.
cases, magma genesis is attributed to lithospheric extension and Such discrepancy can be attributed to the exceptional tectonic
asthenospheric flows triggered by subduction and convergent systems setting into which the magmas originated, i.e., unusually low
in response to the geodynamic changes brought by the supercontinent subduction plate velocities and convergence rates triggered by
formation. Likewise, the emplacement of the CAMP LIP can also be the assembly of Pangea supercontinent. Consequently, the
related to global tectonic changes, in this case, Pangea breakup (e.g., Andean subduction has its roots in the Gondwanide orogeny as
Blackburn et al., 2013; Bond and Wignall, 2014), which constitutes the part of the Terra Australis orogen (Cawood, 2005).
main late Triassic extensional event (Golonka et al., 2018). Further- iv. The contemporaneous magmatic activity of the Choiyoi
more, it has been suggested that this event is also driven by slab province, and the Emeishan and Siberian Traps LIPs and
pulling forces related to the subduction regime of eastern Pangea, in related environmental stresses, together with their rela-
addition to ridge-pushing forces associated with mantle upwelling tionship with the assembly of Pangea (Fig. 9), allow sug-
(Golonka et al., 2018). In all these cases, the geodynamic changes can gesting that global plate tectonic reconfigurations may have
be attributed to variations on the speed and/or direction of plate a key role controlling important magmatic events that ulti-
movements in relation to the supercontinent. mately are related to biotic crises.
In consequence, we propose that global tectonic plate reorga-
nization, the assembly of Pangea, may have had a key role con- Acknowledgements
trolling the emplacement of major magmatic events and therefore
their related CO2 release that ultimately led to the environmental We thank J. Vargas and R. Valles (Universidad de Chile) for the
crises that culminated in mass extinctions. zircon and magnetite separation, O. Pe rez and A. Go
mez-Tuena for the
preparation and measurement of Rb-Sr and Sm-Nd samples (LEI,
6. Conclusions UNAM) and J. Kirk for Os isotope analyses of magnetite (University of
Arizona). We also thank Robert Pankhurst and Arne Willner for their
Whole-rock geochemical composition together with geochro- constructive comments on the manuscript. This research was funded
nological data show that the Chilean Frontal Andes batholiths by the Servicio Nacional de Geología y Minería (SERNAGEOMIN) project
(28 e28 300 S) crystallized from calc-alkaline, I-, S-, and transi- Plan Nacional de Geología (Geological Map ‘Iglesia Colorada-Cerro del
tional to aluminous A-type magmas during most of mid Potro y Cerro Mondaquita’), personal research resources supplied by
PermianeTriassic time. Isotopic data suggests a lower continental Katja Deckart and the Masters fellowship of the Comisio n Nacional de
crust source with minor degrees of upper crust contributions and n Científica y Tecnolo
Investigacio  gica, CONICYT (grant no. 221320626).
important crustal anatexis. A comparison with other coeval units The first author thanks Tais Dahl for his valuable comments and to the
allows the following general conclusions (Fig. 9): Carlsberg Foundation for his current financial support.

i. Latest early Carboniferousemid Permian magmatism origi- Appendix A. Supplementary data


nated in a compressive convergent margin setting with
normal to thickened continental crust during the Gondwa- Supplementary data to this article can be found online at
nide orogeny and it was emplaced within an early accre- https://doi.org/10.1016/j.gr.2019.05.007.
tionary prism. I- and S-type calc-alkaline dioritic, tonalitic
and granodioritic magma intrusions constitute a continuous References
north to south belt from at least 21 S to 40 S. Isotopic data
(d18O; Deckart et al., 2014; del Rey et al., 2016; Herve et al., Alasino, P.H., Dahlquist, J.A., Pankhurst, R.J., Galindo, C., Casquet, C., Rapela, C.W.,
2014) show upper continental crust contributions. Larrovere, M.A., Fanning, C.M., 2012. Early Carboniferous sub- to mid-alkaline
magmatism in the Eastern Sierras Pampeanas, NW Argentina: a record of crustal
ii. Mid PermianeTriassic magmatism occurred in an exten- growth by the incorporation of mantle-derived material in an extensional setting.
sional convergent margin setting caused by slab rollback Gondwana Res. 22, 992e1008. https://doi.org/10.1016/j.gr.2011.12.011.

Alvarez, J., Mpodozis, C., Arriagada, C., Astini, R., Morata, D., Salazar, E.,
extension and subsequent progressive thinning of conti-
Valencia, V.A., Vervoort, J.D., 2011. Detrital zircons from late Paleozoic accre-
nental crust. Most magmas originated due to melting of
tionary complexes in north-central Chile (28 e32 S): possible fingerprints of
lower continental crust with increasingly less upper crust the Chilenia terrane. J. S. Am. Earth Sci. 32, 460e476. https://doi.org/10.1016/
contributions and variable degrees of mantle influence. j.jsames.2011.06.002.
Andersen, T., 2002. Correction of common lead in UePb analyses that do not report
Crustal anatexis and extension occurred in the backarc re- 204
Pb. Chem. Geol. 192, 59e79.
gion and was simultaneous with subduction-related arc tat, A., Keppler, H., 2008. The origin of the negative niobium-tantalum
Baier, J., Aude
magmas. Widespread extensional conditions were triggered anomaly in subduction zone magmas. Earth Planet. Sci. Lett. 267, 290e300.
by slab rollback in association with steepening under low Blackburn, T.J., Olsen, P.E., Bowring, S.A., McLean, N.M., Kent, D.V., Puffer, J.,
McHone, G., Rasbury, E.T., Et-Touhami, M., 2013. Zircon U-Pb geochronology
subduction plate velocities and convergence rates as a result links the end-Triassic extinction with the Central Atlantic Magmatic Province.
of the assembly of Pangea supercontinent. Science 340, 941e945. https://doi.org/10.1126/science.1234204.
320  del Rey et al. / Gondwana Research 76 (2019) 303e321
A.

Blusztajn, J., Hart, S.R., Ravizza, G., Dick, H.J.B., 2000. Platinum-group elements and Hart, G.L., Johnson, C.M., Shirey, S.B., Clynne, M.A., 2002. Osmium isotope con-
Os isotopic characteristics of the lower oceanic crust. Chem. Geol. 168, 113e122. straints on lower crustal recycling and pluton preservation at Lassen Volcanic
Bond, D.P.G., Wignall, P.B., 2014. Large igneous provinces and mass extinctions: an up- Center, CA. Earth Planet. Sci. Lett. 199, 269e285. https://doi.org/10.1016/S0012-
date. Geol. Soc. Am. Spec. Pap. 505, 29e55. https://doi.org/10.1130/2014.2505(02). 821X(02)00564-2.
Brown, M., 1991. Comparative geochemical interpretation of Permian-Triassic Herve , F., Fanning, C.M., Caldero n, M., Mpodozis, C., 2014. Early Permian to Late
plutonic complexes of the Coastal Range and Altiplano (25 300 to 26 300 S), Triassic batholiths of the Chilean Frontal Cordillera (28 e31 S): SHRIMP U-Pb
northern Chile. In: Geological Society of America, vol. 265. zircon ages and Lu-Hf and O isotope systematics. Lithos 184e187, 436e446.
Burgess, S.D., Bowring, S.A., 2015. High-precision geochronology confirms volumi- https://doi.org/10.1016/j.lithos.2013.10.018.
nous magmatism before, during, and after Earth’s most severe extinction. Sci. Hyppolito, T., Juliani, C., García-Casco, A., Meira, V.T., Bustamante, A., Herve , F., 2014.
Adv. 1 https://doi.org/10.1126/sciadv.1500470. The nature of the Palaeozoic oceanic basin at the southwestern margin of
Burgess, S.D., Bowring, S.A., Shen, S., 2014. High-precision timeline for Earth’s most Gondwana and implications for the origin of the Chilenia terrane (Pichilemu
severe extinction. Proc. Natl. Acad. Sci. 111, 3316e3321. https://doi.org/10.1073/ region, central Chile). Int. Geol. Rev. 56, 1097e1121.
pnas.1317692111. Jacobsen, S.B., Wasserburg, G.J., 1980. Sm-Nd isotopic evolution of chondrites. Earth
Burgess, S.D., Muirhead, J.D., Bowring, S.A., 2017. Initial pulse of Siberian Traps sills Planet. Sci. Lett. 50, 139e155.
as the trigger of the end-Permian mass extinction. Nat. Commun. 8, 1e4. Johnson, C.M., Shirey, S.B., Barovich, K.M., 1996. New approaches to crustal evolu-
https://doi.org/10.1038/s41467-017-00083-9. tion studies and the origin of granitic rocks: what can the Lu-Hf and Re-Os
Cawood, P.A., 2005. Terra Australis Orogen: Rodinia breakup and development of isotope systems tell us? Trans. R. Soc. Edinb. Earth Sci. 87, 339e352.
the Pacific and Iapetus margins of Gondwana during the Neoproterozoic and Kay, S.M., Coira, B.L., 2009. Shallowing and Steepening Subduction Zones, Conti-
Paleozoic. Earth Sci. Rev. 69, 249e279. https://doi.org/10.1016/j.earscirev. nental Lithospheric Loss, Magmatism, and Crustal Flow Under the Central An-
2004.09.001. dean Altiplano-Puna Plateau 1204. https://doi.org/10.1130/2009.1204(11).
Chappell, B.W., Stephens, W.E., 1988. Origin of infracrustal (I-type) granite magmas. Kay, S.M., Ramos, V.A., Mpodozis, C., Sruoga, P., 1989. Late Paleozoic to Jurassic silicic
Trans. R. Soc. Edinb. Earth Sci. 79, 71e86. https://doi.org/10.1017/ magmatism at the Gondwana margin: analogy to the Middle Proterozoic in
S0263593300014139. North America?. In: Geological Society of America, vol. 17, pp. 324e328. https://
Chappell, B.W., White, A.J.R., 2001. Two contrasting granite types: 25 years later. doi.org/10.1130/0091-7613(1989)017<0324.
Aust. J. Earth Sci. 48, 489e499. Kay, S.M., Mpodozis, C., Gardeweg, M., 2013. Magma sources and tectonic setting
Chappell, B.W., Bryant, C.J., Wyborn, D., 2012. Peraluminous I-type granites. Lithos of Central Andean forearc subduction erosion and delamination magma
153, 142e153. https://doi.org/10.1016/j.lithos.2012.07.008. sources and tectonic setting of Central Andean andesites (25.5e28 S) related
Charrier, R., 1979. El Tria sico en Chile y regiones adyacentes de Argentina: Una to crustal thickening, forearc subduction erosion and delamination. In:
reconstruccio n paleogeogr afica y paleoclim atica. Comunicaciones 26, 1e37. Geological Society, London, Special Publications. https://doi.org/10.1144/
Charrier, R., Ramos, V.A., Tapia, F., Sagripanti, L., 2014. Tectono-stratigraphic evo- SP385.11.
lution of the Andean Orogen between 31 and 37 S (Chile and Western King, P.L., White, A.J.R., Chappell, B.W., Allen, C.M., 1997. Characterization and Origin
Argentina). Geol. Soc. Lond., Spec. Publ. 399, 13e61. https://doi.org/10.1144/ of Aluminous A-type Granites From the Lachlan Fold Belt, Southeastern
SP399.20. Australia, vol. 38, pp. 371e391.
Collins, W.J., 2002. Hot orogens, tectonic switching, and creation of continental Kleiman, L.E., Japas, M.S., 2009. The Choiyoi volcanic province at 34 Se36 S (San
crust. Geology 30, 535e538. Rafael, Mendoza, Argentina): implications for the Late Palaeozoic evolution of
Creaser, R.A., Papanastassiou, D.A., Wasserburg, G.J., 1991. Negative thermal ion the southwestern margin of Gondwana. Tectonophysics 473, 283e299. https://
mass-spectrometry of osmium, rhenium, and iridium. Geochim. Cosmochim. doi.org/10.1016/j.tecto.2009.02.046.
Acta 55, 397e401. https://doi.org/10.1016/0016-7037(91)90427-7. Li, Z.X., Powell, C.M., 2001. An outline of the palaeogeographic evolution of the
Deckart, K., Godoy, E., Bertens, A., Jerez, D., Saeed, A., 2010. Barren Miocene gran- Australasian region since the beginning of the Neoproterozoic. Earth-Sci. Rev.
itoids in the Central Andean metallogenic belt, Chile: geochemistry and Nd-Hf 53, 237e277. https://doi.org/10.1016/S0012-8252(00)00021-0.
and U-Pb isotope systematics. Andean Geol. 37, 1e31. https://doi.org/10.4067/ Lister, G., Forster, M., 2009. Tectonic mode switches and the nature of orogenesis.
s0718-71062010000100001. Lithos 113, 274e291. https://doi.org/10.1016/j.lithos.2008.10.024.
Deckart, K., Herve , F., Fanning, M., Ramírez, V., Caldero n, M., Godoy, 2014. U-Pb Llambias, E.J., Sato, A.M., 1990. El Batolito de Colangüil (29-31 S) Cordillera Frontal
geochronology and Hf-O isotopes of zircons from the Pennsylvanian Coastal de Argentina: estructura y marco tecto nico. Rev. Geol. Chile 17, 89e108.
Batholith, South-Central Chile. Andean Geol. 41, 49e82. https://doi.org/ Maksaev, V., Munizaga, F., Tassinari, C., 2014. Timing of the magmatism of the
10.5027/andgeoV41n1-a03. paleo-Pacific border of Gondwana: U-Pb geochronology of Late Paleozoic to
 Deckart, K., Arriagada, C., Martínez, F., 2016. Resolving the paradigm of
del Rey, A., Early Mesozoic igneous rocks of the north Chilean Andes between 20 and 31 S.
the late PaleozoiceTriassic Chilean magmatism: isotopic approach. Gondwana Andean Geol. 41, 447e506. https://doi.org/10.5027/andgeoV41n3-a01.
Res. 37, 172e181. https://doi.org/10.1016/j.gr.2016.06.008. Martínez, F., Pen ~ a, M., Arriagada, C., 2015. Geología de las  areas Iglesia Colorada-
Eby, G.N., 1992. Chemical subdivision of the A-type granitoids: petrogenetic and Cerro del Potro y Cerro Mondaquita, Regio  n de Atacama. In: Carta Geolo  gica de
tectonic implications. Geology 20, 641e644. Chile, pp. 179e180.
Espinoza, M., Montecino, D., Oliveros, V., Astudillo, N., Va squez, P., Reyes, R., Celis, C., Martínez, F., Arriagada, C., Pen ~ a, M., Deckart, K., Charrier, R., 2016. Tectonic styles
Gonza lez, R., Contreras, J., Creixell, C., Martínez, A., 2018. The synrift phase of and crustal shortening of the Central Andes “Pampean” flat-slab segment in
the early Domeyko Basin (Triassic, northern Chile): sedimentary, volcanic and 
northern Chile (27e29 S). Tectonophysics 667, 144e162. https://doi.org/
tectonic interplay in the evolution of an ancient subduction-related rift basin. 10.5027/andgeoV41n3-a01.
Basin Res. https://doi.org/10.1111/bre.12305. McDonough, W.F., Sun, S.S., 1995. Composition of the Earth. Chem. Geol. 120,
Franzese, J.R., Spalletti, L.a., 2001. Late triassic-early jurassic continental extension 223e253.
in southwestern Gondwana: tectonic segmentation and pre-break-up rifting. McKenzie, N.R., Horton, B.K., Loomis, S.E., Stockli, D.F., Planavsky, N.J., Lee, C.-T.A.,
J. S. Am. Earth Sci. 14, 257e270. https://doi.org/10.1016/S0895-9811(01)00029-3. 2016. Continental arc volcanism as the principal driver of icehouse-greenhouse
García-Sansegundo, J., Farias, P., Heredia, N., Gallastegui, G., Charrier, R., Rubio- variability. Science 352, 444e448.
Ordo n
~ ez, A., Cuesta, A., 2014. Structure of the Andean Palaeozoic basement in Mpodozis, C., Kay, S.M., 1992. Late Paleozoic to Triassic evolution of the Gondwana
the Chilean coast at 31300 S: geodynamic evolution of a subduction margin. margin: evidence from Chilean Frontal cordilleran batholiths (28 S to 31 S).
J. Iber. Geol. 40, 293e308. https://doi.org/10.5209/rev_JIGE.2014.v40.n2.45300. Geol. Soc. Am. Bull. 104, 999e1014.
Golonka, J., Embry, A., Krobicki, M., 2018. Late Triassic global plate tectonics. In: Mpodozis, C., Marinovic, N., Smoje, I., Cuitin ~ o, L., 1993. Estudio Geologico-Estruc-
Tanner, L.H. (Ed.), The Late Triassic World: Earth in a Time of Transition. tural de la Cordillera de Domeyko entre Sierra Limo n Verde y Sierra Mariposas,
Springer International Publishing, Cham, pp. 27e57. https://doi.org/10.1007/ Region de Antofagasta. Servicio Nacional de Geología y Minería, Informe Reg-
978-3-319-68009-5_2. istrado IR-93-04 282.
Gomez-Tuena, A., Mori, L., Goldstein, S.L., Pe rez-Arvizu, O., 2011. Magmatic diversity Munizaga, F., Maksaev, V., Fanning, C.M., Giglio, S., Yaxley, G., Tassinari, C.C.G., 2008.
of western Mexico as a function of metamorphic transformations in the sub- Late Paleozoic-early Triassic magmatism on the western margin of Gondwana:
ducted oceanic plate. Geochim. Cosmochim. Acta 75, 213e241. Collahuasi area, Northern Chile. Gondwana Res. 13, 407e427. https://doi.org/
Gonza lez, J., Oliveros, V., Creixell, C., Vel
asquez, R., V asquez, P., Lucassen, F., 2017. 10.1016/j.gr.2007.12.005.
The Triassic magmatism and its relation with the Pre-Andean tectonic evolu- N€agler, T.F., Frei, R., 1997. Plug in osmium distillation. In: Schweizer Mineralogische
tion: geochemical and petrographic constrains from the High Andes of north und Petrographische Mitteilungen, vol. 77, pp. 123e127.
Central Chile (29 300 e30 S). J. S. Am. Earth Sci. https://doi.org/10.1016/ Nelson, D.A., Cottle, J.M., 2019. Tracking voluminous Permian volcanism of the
j.jsames.2017.12.009. Choiyoi Province into central Antarctica. In: Geological Society of America,
Gorczyk, W., Willner, A.P., Gerya, T.V., Connolly, J.A.D., Burg, J.P., 2007. Physical pp. 1e13. https://doi.org/10.1130/L1015.1.
controls of magmatic productivity at Pacific-type convergent margins: nu- Pankhurst, R., Hole, M., Brook, M., 1988. Isotope evidence for the origin of Andean
merical modelling. Phys. Earth Planet. Inter. 163, 209e232. https://doi.org/ granites. Trans. R. Soc. Edinb. Earth Sci. 79, 123e133.
10.1016/j.pepi.2007.05.010. Pankhurst, R., Millar, I., Herve , F., 1996. A Permo-Carboniferous U-Pb age for part of
Gregori, D., Benedini, L., 2013. The Cordon del Portillo Permian magmatism, Men- the Guanta Unit of the Elqui-Limari Batholith at Rio del Transito, Northern
doza, Argentina, plutonic and volcanic sequences at the western margin of Chile. Rev. Geol. Chile 23, 35e42.
Gondwana. J. S. Am. Earth Sci. 42, 61e73. https://doi.org/10.1016/ Pankhurst, R.J., Rapela, C.W., Fanning, C.M., Ma rquez, M., 2006. Gondwanide con-
j.jsames.2012.07.010. tinental collision and the origin of Patagonia. Earth Sci. Rev. 76, 235e257.
Hacker, B.R., Kelemen, P.B., Behn, M.D., 2015. Continental lower crust. Annu. Rev. https://doi.org/10.1016/j.earscirev.2006.02.001.
Earth Planet. Sci. 43, 167e205. https://doi.org/10.1146/annurev-earth-050212- Parada, M.A., 1990. Granitoid plutonism in central Chile and its geodynamic im-
124117. plications; a review. Geol. Soc. Am. Spec. Pap. 241, 51e66.
 del Rey et al. / Gondwana Research 76 (2019) 303e321
A. 321

Parada, M.A., Nystro €m, J.O., Levi, B., 1999. Multiple sources for the Coastal Batholith Schellart, W.P., 2005. Influence of the subducting plate velocity on the geometry of
of central Chile (31e34 S): geochemical and Sr-Nd isotopic evidence and tec- the slab and migration of the subduction hinge. Earth Planet. Sci. Lett. 231,
tonic implications. Lithos 46, 505e521. https://doi.org/10.1016/S0024-4937(98) 197e219. https://doi.org/10.1016/j.epsl.2004.12.019.
00080-2. Schellart, W.P., 2008. Overriding plate shortening and extension above subduction
Parada, M.A., Lo  pez-Escobar, L., Oliveros, V., Fuentes, F., Morata, D., Caldero  n, M., zones: a parametric study to explain formation of the Andes Mountains. Bull.
raud, G., Espinoza, F., Moreno, H., Figueroa, O., Mun
Aguirre, L., Fe ~ oz Ravo, J., Tron- Geol. Soc. Am. 120, 1441e1454. https://doi.org/10.1130/B26360.1.
coso V asquez, R., Stern, C.R., 2007. Andean magmatism. In: Moreno, T., Gibson, W. Shirey, S.B., Walker, R.J., 1995. Carius tube digestion for low-blank rhenium-osmium
(Eds.), The Geology of Chile. The Geological Society, London, pp. 115e146. analysis. Anal. Chem. 67, 2136e2141. https://doi.org/10.1021/ac00109a036.
Paton, C., Woodhead, J.D., Hellstrom, J.C., Hergt, J.M., Greig, A., Maas, R., 2010. Smoliar, M.I., Walker, R.J., Morgan, J.W., 1996. ReeOs ages of group IIA, IIIA, IVA, and
Improved laser ablation UePb zircon geochronology through robust downhole IVB iron meteorites. Science 271, 1099e1102.
fractionation correction. Geochem. Geophys. Geosyst. 11. Solari, L.A., Go  mez-Tuena, A., Bernal, J.P., Pe
rez-Arvizu, O., Tanner, M., 2010. U-Pb
Pearce, J.A., Peate, D.W., 1995. Tectonic implications of Volcanic Arc Magmas. Annu. zircon geochronology by an integrated LAICPMS microanalytical workstation:
Rev. Earth Planet. Sci. 23, 251e285. https://doi.org/10.1146/annurev.ea. achievements in precision and accuracy. Geostand. Geoanal. Res. 34, 5e18.
23.050195.001343. Spalletti, L.A., Limarino, C.O., 2017. The Choiyoi magmatism in south western
Pearce, J.A., Harris, N.B.W., Tindle, A.G., 1984. Trace Element Discrimination Dia- Gondwana: implications for the end-permian mass extinction - a review. An-
grams for the Tectonic Interpretation of Granitic Rocks, vol. 25, pp. 956e983. dean Geol. 44, 328. https://doi.org/10.5027/andgeov44n3-a05.
Peccerillo, A., Taylor, S.R., 1976. Geochemistry of eocene calc-alkaline volcanic rocks Stampfli, G.M., Hochard, C., Ve rard, C., Wilhem, C., vonRaumer, J., 2013. The formation of
from the Kastamonu area, Northern Turkey. Contrib. Mineral. Petrol. 58, 63e81. Pangea. Tectonophysics 593, 1e19. https://doi.org/10.1016/j.tecto.2013.02.037.
https://doi.org/10.1007/BF00384745. Steiger, R.H., J€ ager, E., 1977. Subcommission of geochronology: convention on the
Pietranik, A., Słodczyk, E., Hawkesworth, C.J., Breitkreuz, C., Storey, C.D., use of decay constants in geo- and cosmochronology. Earth Planet. Sci. Lett. 1.
Whitehouse, M., Milke, R., 2013. Heterogeneous zircon cargo in voluminous Vasquez, P., Franz, G., 2008. The Triassic Cobquecura Pluton (Central Chile): an
Late Paleozoic rhyolites: Hf, O isotope and Zr/Hf records of plutonic to volcanic example of a fayalite-bearing A-type intrusive massif at a continental margin.
magma evolution. J. Petrol. 54, 1483e1501. https://doi.org/10.1093/petrology/ Tectonophysics 459, 66e84. https://doi.org/10.1016/j.tecto.2007.11.067.
egt019. Vasquez, P., Glodny, J., Franz, G., Frei, D., Romer, R., 2011. Early Mesozoic Plutonism
Ramos, V.A., 2009. Anatomy and global context of the Andes: main geologic fea- of the Cordillera de la Costa (34 e37 S), Chile: constraints on the Onset of the
tures and the Andean orogenic cycle. In: Geological Society of America Memoir, Andean Orogeny. J. Geol. 119, 159e184.
vol. 204, pp. 31e65. https://doi.org/10.1130/2009.1204(02). Vilas, J.F., Valencio, D.A., 1978. Paleomagnetism of South American and African rocks
Ramos, V.A., Folguera, A., 2009. Andean flat-slab subduction through time. Geol. and the age of the South Atlantic. In: Revista Brasileira de Geocie ^ncias, vol. 8,
Soc. Lond., Spec. Publ. 327, 31e54. https://doi.org/10.1144/SP327.3. pp. 3e10.
Rapela, C.W., Pankhurst, R.J., 1992. The granites of northern Patagonia and the Walker, R.J., Horan, M.F., Morgan, J.W., Becker, H., Grossman, J.N., Rubin, A.E., 2002.
Gastre Fault System in relation to the break-up of Gondwana. Geol. Soc. Lond., Comparative 187Ree187Os systematics of chondrites: implications regarding
Spec. Publ. 68, 209e220. early solar system processes. Geochim. Cosmochim. Acta 66, 4187e4201.
Riel, N., Jaillard, E., Martelat, J.-E., Guillot, S., Braun, J., 2018. Permian-Triassic Wang, Y., Santosh, M., Luo, Z., Hao, J., 2015. Large igneous provinces linked to super-
Tethyan realm reorganization: implications for the outward Pangea margin. J. S. continent assembly. J. Geodyn. 85, 1e10. https://doi.org/10.1016/j.jog.2014.12.001.
Am. Earth Sci. 81, 78e86. https://doi.org/10.1016/j.jsames.2017.11.007. Whalen, J.B., Currie, K.L., Chappell, B.W., 1987. A-type granites: geochemical char-
Rocha-Campos, A.C., Basei, M.A., Nutman, A.P., Kleiman, L.E., Varela, R., Llambias, E., acteristics, discrimination and petrogenesis. Contrib. Mineral. Petrol. 95,
Canile, F.M., da Rosa, O. de C.R., 2011. 30 million years of Permian volcanism 407e419. https://doi.org/10.1007/BF00402202.
recorded in the Choiyoi igneous province (W Argentina) and their source for Whalen, J.B., Jenner, G.A., Longstafe, Frederick J., Robert, F., Gariepy, C., 1996.
younger ash fall deposits in the Parana  Basin: SHRIMP UePb zircon geochro- Geochemical and isotopic (O, Nd, Pb and Sr) constraints on A-type granite
nology evidence. Gondwana Res. 19, 509e523. https://doi.org/10.1016/ petrogenesis based on the Topsails igneous suite, Newfoundland Appalachians.
j.gr.2010.07.003. J. Petrol. 37, 1463e1489. https://doi.org/10.1093/petrology/37.6.1463.
Rudnick, R.L., Fountain, D.M., 1995. Nature and Composition of the Continental Whitney, J.A., 1988. The origin of granite: the role and source of water in the evo-
Crust: A Lower Crustal Perspective, pp. 267e309. https://doi.org/10.1029/ lution of granitic magmas. Bull. Geol. Soc. Am. 100, 1886e1897. https://doi.org/
95RG01302. 10.1130/0016-7606(1988)100<1886:TOOGTR>2.3.CO;2.
Sato, A.M., Llambías, E.J., Basei, M.A.S., Castro, C.E., 2015. Three stages in the Late Willner, A.P., Thomson, S.N., Kro €ner, A., Wartho, J.A., Wijbrans, J.R., Herve , F., 2005.
Paleozoic to Triassic magmatism of southwestern Gondwana, and the re- Time markers for the evolution and exhumation history of a Late Palaeozoic
lationships with the volcanogenic events in coeval basins. J. S. Am. Earth Sci. 63, paired metamorphic belt in North-Central Chile (34 e35 300 S). J. Petrol. 46,
48e69. https://doi.org/10.1016/j.jsames.2015.07.005. 1835e1858. https://doi.org/10.1093/petrology/egi036.

You might also like