Plantilla de artículo 2013
Andean Geology 41 (3): 447-506, September, 2014
Andean Geology
doi: 10.5027/andgeoV41n3-a01
formerly Revista Geológica de Chile
Timing of the magmatism of the paleo-Pacific border of Gondwana:
U-Pb geochronology of Late Paleozoic to Early Mesozoic igneous
rocks of the north Chilean Andes between 20° and 31°S
Víctor Maksaev1, Francisco Munizaga1, 2, Colombo Tassinari3

1 Departamento de Geología, Universidad de Chile, Plaza Ercilla 803, Santiago, Chile.
vmaksaev@ing.uchile.cl; fmunizag@ing.uchile.cl

2 Geología, Facultad de Ingeniería, Universidad Andrés Bello, Salvador Sanfuentes 2375, Santiago, Chile.

3 Centro de Investigaciones Paleobiológicas (CIPAL), Facultad de Ciencias Exactas, Físicas y Naturales, Universidad Nacional de Córdoba, Avda. Vélez Sársfield 299, X500JJC, Córdoba, Argentina.

4 Instituto de Geociencias, Universidade de Sao Paulo, Rua do Lago, 562-Cidade Universitária CEP 05508-080, Sao Paulo, Brasil.
ccgtassi@usp.br

U-Pb zircon geochronological data provide record of about 130 Ma of igneous activity in the Andes of northern Chile, which extended episodically from the latest Early Carboniferous to Early Jurassic (328-194 Ma). The overall U-Pb data show that volcanism and plutonism were essentially synchronous and major episodes of igneous ac-tivity developed during the Late Carboniferous to Mid-Permian (310 to 260 Ma) and from Late Permian to Late Triassic (255-205 Ma), with less prominent episodes in the mid-Carboniferous (330 to 320 Ma), and Early Jurassic (200-190 Ma). Thus, from the Carboniferous to the Early Triassic dominantly silicic magmatism developed along the Chilean segment of the southwestern border of Gondwana supercontinent. Further magmatism developed during the Mid-Late Triassic (250-194 Ma) was bimodal and synchronous with rift-related, continental and/or marine sedimentary strata related to the early stages of break-up of Gondwana. Most of the silicic volcanic rocks of the Precordillera and Domeyko Cordillera of northern Chile (21°30’ to 25°30’S) are older than the silicic rocks assigned to the Choiyoi succession in Argentina, being instead equivalent in age to Carboniferous to Early Permian marine sedimentary sequences present in the eastern Argentinean foreland. On the other hand, silicic volcanic successions exposed in the easternmost part of northern Chile are equivalent in age to the Choiyoi succession of the San Rafael Block of Argentina. An eastward expansion or migration of the volcanism during the Mid-Permian to Early Triassic is inferred, interpretation that is consistent with expansion of the volcanism at that time in Argentina. The timing of the Late Paleozoic to Early Jurassic magmatism is coincident with that of the Andes of Perú and of western Argentina according to the available U-Pb data, revealing a rather consistent evolution in time of the magmatism along the southwestern, paleo-Pacific border of Gondwana.

Keywords: Geochronology, U-Pb, Andes, Gondwana, Choiyoi, Paleozoic, Carboniferous, Triassic.

Abstract

1. Introduction

The Late Paleozoic-Early Mesozoic magmatism is a significant, but not completely understood geological event, which has been recognized along much of the length of western South America (e.g., Vaughan and Pankhurst, 2008). It is characterized by a high proportion of felsic volcanic rocks and granitic plutons over intermediate and mafic igneous rocks, and represents magmatism that developed along the southwestern paleo-Pacific border of Gondwana supercontinent (Llambías and Sato, 1990; Sato and Llambías, 1993; Mpodozis and Kay, 1992; Strazzere et al., 2006; Munizaga et al., 2008; Vaughan and Pankhurst, 2008; Rocha-Campos et al., 2011). The previous K-Ar dating of this extensive volcano-plutonic complex in northern Chile has yielded an overall range from 332 to 197 Ma for this magmatic activity (e.g., Huete et al., 1977; Nasi et al., 1985; Mpodozis et al., 1993; Lucassen et al., 1999; Tomlinson et al., 2001; Tomlinson and Blanco, 2008). The K-Ar ages have mostly been obtained for intrusive rocks, since volcanic rocks commonly are altered or show low-grade metamorphism, which have prevented the application of the K-Ar dating method or produced inaccurate dates. More recent studies have provided a more accurate geochronological database of U-Pb dates for the Late Paleozoic-Early Mesozoic igneous rocks of northern Chile (north of 31°S latitude), which we have compiled in Tables 1 and 2 (references therein), but also in Argentina (i.e., Pankhurst et al., 2006; Gulbranson et al., 2010; Rocha-Campos et al., 2011). In this contribution we present 41 new SHRIMP U-Pb dates for volcanic and plutonic rocks of northern Chile, which together with the compilation of U-Pb dates from previous studies, provide a more accurate timing for the Late Paleozoic to Early Jurassic magmatism of the north-Chilean segment of the southwestern border of Gondwana. The numerical ages are assigned to the International Stratigraphic Chart of the International Commission on Stratigraphy 2013 (Cohen et al., 2013).

2. Geological background

The Late Paleozoic to Triassic igneous units along the Andes of northern Chile and the Argentinean Frontal Cordillera are commonly known by a variety of names, but Stipanicic et al. (1968) employed the name Choiyoi Group for volcanic rocks attributed to the Permian to Triassic outcropping in the Argentinean Cordillera, and this name has been widely used to refer to ʻPermo-Triassicʼ magmatism in both countries, even though the actual timing of the igneous rocks was quite inaccurate and only limited and inaccurate geochronological data was available. The Choiyoi igneous province consists of volcano-plutonic complexes (Kay et al., 1989; Llambías et al., 1993; Llambías and Sato, 1995) that cover an estimated area of ~500,000 km2, which extends for about 2,500 km, from the Collahuasi area in north-ern Chile (20°30’S) to Neuquén and the northern Patagonian Andes of southern Argentina (44°S; Jordan et al., 1983; Mpodozis and Ramos, 1989; Ramos and Folguera, 2009; Ramos, 2009). Besides, it is also correlated with the Mitu Group (Carlier et al., 1982) and the Yura Group of southern Perú (Sempere et al., 2002; Boekhout et al., 2013), and geological and geochronological data from magmatic and migmatitic rocks of the Northern Andes have shown that the Late Paleozoic to Triassic magmatism was also extensively recorded from northern Perú to the Central Cordillera of the Colombian Andes (Restrepo et al., 1991; Noble et al., 1997; Ordóñez and Pimentel, 2002; Vinasco et al., 2006; Chew et al., 2007; Ibáñez-Mejía et al., 2008; Cardona et al., 2008, 2009).

The main outcrops of Late Paleozoic to Early Jurassic igneous rocks in northern Chile are ex-posed in a semi-continuous belt from about 21°S to 31°S latitudes. These rocks are prominent along the Domeyko Cordillera (24° to 27°S) and its northern extension, the so called Chilean Precordillera (20° to 24°S), but also along the High Andes from 27° to 31°S, which are commonly referred as Chilean Frontal Cordillera. Late Carboniferous and Triassic granitic plutons are also scattered from 21°S latitude southwards along the Coastal Cordillera and then forming an almost continuous intrusive belt between 33° to 38°S (e.g., Deckart et al., 2014).

The isolation of outcrops, fault-bounded blocks, volcanic deposits of similar petrographic compositions, lack of accurate geochronological data, and the absence of guide horizons, have made difficult to know the exact stratigraphic-structural relationships between separate exposures. Thus, a number of lithostratigraphic units were assigned by differ-ent authors to specific time-spans and correlated with other exposures, which at times proved to be incorrect as geochronological data become avail-able. This is reflected by a profuse stratigraphic nomenclature used in previous studies to refer to Late Paleozoic to Early Mesozoic successions of the Andes of northern Chile (e.g., Charrier et al., 2007); for unambiguousness figures 1 and 2 illustrate preliminary stratigraphic schemes for two segments of the Andes of northern Chile, modified according to the new geochronological data.

 

fig.1

 

Fig. 1. Late Paleozoic to Early Jurassic stratigraphic units of northern Chile from 20° to 26°S, ordered according to the new geochronological data. The different units are ordered geographically from western (left) to eastern exposures (right) and face each other according to their approximate latitude of occurrence. The reader is referred to the work of Charrier et al. (2007 and references therein) for the units that are not mentioned in text.

 

 

fig.2

 

Fig. 2. Late Paleozoic to Early Jurassic stratigraphic units of northern Chile from 26° to 31°S, ordered according to the new geochronological data. The different units are ordered geographically from western (left) to eastern exposures (right) and face each other according to their approximate latitude of occurrence. The reader is referred to the work of Charrier et al. (2007 and references therein) for the units that are not mentioned in text.

 

The exposed Late Paleozoic to Mid-Triassic volcanic succession along the Andes in northern Chile has uneven thickness, visible successions typically range between 750 and 2,600 m, but exceeding 3,200 m in the Collahuasi area of the Precordillera (21°S latitude; Vergara and Thomas, 1984). These successions are mostly formed of felsic volcanic rocks that include extensive rhyolitic and dacitic ash flow tuffs (ignimbrites), and banded rhyolite and dacitic domes and lava-flows. A lesser proportion is composed of basaltic and andesitic lavas, and pyroclastic deposits of the same petrographic composition. Sedimentary intercalations of volcanic sandstones, minor conglomerates, and local limestones (i.e., Collahuasi area; Vergara and Thomas, 1984), are also locally present. The volcanic successions unconformably overlie Devonian to Early Carboniferous shallow-marine, sedimentary successions, but its base is only locally exposed (Mercado, 1982; Niemeyer et al., 1985, 1997; Mpodozis and Cornejo, 1988; Nasi et al., 1990; Marinovic et al., 1995). In addition, calcareous and terrigenous, shallow-marine sedimentary successions, containing Early Permian fossils are exposed in the western foothills of the High Andes and in the Central Depression. These sedimentary units represent Early Permian carbonate sedimentation within a shallow-marine platform, but some of them are either overlying or interstratified with rhyolitic rocks (Sepúlveda and Naranjo, 1982; Marinovic et al., 1995; Iriarte et al., 1996; Cortés, 2000; Díaz-Martínez et al., 2000).

Middle to Late Triassic successions are sedimentary-volcanic in nature and are separated from older rocks by unconformities. A bimodal suite composed of basalts and rhyolites, locally with pillow lavas, and associated rhyolitic and dacitic sills, domes and ash-flow deposits occur within the Mid-Late Triassic to Early Jurassic successions (Naranjo and Puig, 1984; Suárez and Bell, 1992; Tomlinson et al., 1999; Cornejo and Mpodozis, 1996, 1997; Cornejo et al., 2009). The Late Triassic to Early Jurassic magmatism is fairly common, but it has been poorly investigated (e.g., Franz et al., 2006), and has not always been recognized as a distinct unit from older Carboniferous-Triassic rocks.

Carboniferous to Triassic granitic, granodioritic, and tonalitic batholiths and lesser dioritic or gabbroic plutons were emplaced in different segments along the Precordillera (21°-24°S), Domeyko Cordillera (24°-27°S) and an extensive sector of the Chilean Frontal Cordillera between 28° and 31°S (Figs. 3 and 4) is formed of granitic rocks of Late Paleozoic to Triassic age (Parada et al., 1981, 2007; Nasi et al., 1985; Mpodozis and Kay, 1990, 1992; Mpodozis et al., 1993; Cornejo et al., 1998; Tomlinson et al., 1999; Tomlinson et al., 2001; Tomlinson and Blanco, 2008; Hervé et al., 2014). Scattered Late Carboniferous and Triassic granitic plutons also extend from 21°S latitude southwards along the Coastal Cordillera, and then forming an almost continuous intrusive belt between 33° to 38°S (Berg et al., 1983; Shibata et al., 1984; Maksaev and Marinovic, 1980; Skarmeta and Marinovic, 1981; Berg and Baumann, 1985; Mpodozis and Kay, 1992; Parada et al., 2007; Deckart et al., 2014). Some of the Carboniferous and Early Permian plutons show foliation with preferred orientation of their tabular minerals, and local bands of mylonites with high-temperature, ductile, shearing deformation (Nasi et al., 1985; Ribba et al., 1988; Marinovic et al., 1995; Murillo et al., 2012; Hervé et al., 2014). Close associations of granitic plutons and ignimbrites in the Andes have been interpreted as large relict calderas (Davidson et al., 1985; Breitkreuz, 1995; Tomlinson et al., 2001). In addition, a belt of porphyry copper prospects and hydrothermal alteration zones are associated with the Permian and Triassic intrusions in Chile and Argentina (Sillitoe, 1977; Camus, 2003; Cornejo et al., 2006; Tomlinson and Blanco, 2008).

 

fig.3

 

Fig. 3. Map showing the distribution of the Late Paleozoic to Late Triassic rocks of northern Chile between latitudes 20° and 26°S and location of U-Pb zircon ages (Ma±2σ). New SHRIMP U-Pb ages in bold; previous U-Pb dates compiled from previous works in italics (references in Tables 1 and 2).

 

fig.4

 

Fig. 4. Map showing the distribution of the Late Paleozoic to Late Triassic rocks of northern Chile between latitudes 26° and 31°S and location of U-Pb zircon ages (Ma±2σ). New SHRIMP U-Pb ages in bold; previous U-Pb dates compiled from previous works in italics (references in Tables 1 and 2).

 

The Carboniferous to Mid-Triassic igneous rocks constitute a calc-alkaline to potassium-rich calc-alkaline suite, are metaluminous to peraluminous in composition, and have overall geochemical characteristics consistent with subduction-related, magmatic arc origin, but with evolution into extensional magmatism and significant involvement of crustal melts, evidencing significant crustal recycling (Nasi et al., 1985; Kay et al., 1989; Breitkreuz et al., 1989; Mpodozis and Kay, 1990, 1992; Mpodozis et al., 1993; Breitkreuz and Zeil, 1994; Marinovic et al., 1995; Lucassen et al., 1999; Llambías et al., 2003; Parada et al., 2007; Urzúa, 2009; Vásquez et al., 2011; Parada, 2013). Brown (1990) compared the major and trace elements geochemistry of the Permian to Middle Triassic Coastal granitic plutons at latitude 26°S with those exposed some 120 km to the east in the High Andes, showing their similar compositions with subduction-related affinity, and concluded that they were part of the same magmatic arc split during Mesozoic extension.

The Middle to Late Triassic igneous rocks show a distinct bimodal composition with a within plate tendency, consistent with an origin in an extensional, rift-related setting (Vergara et al., 1991; Parada et al., 1991; Morata et al., 2000; Vásquez et al., 2011; Parada, 2013).

 

3. Previous U-Pb geochronological data

U-Pb geochronological data for Late Paleozoic to Early Mesozoic rocks compiled from previous studies are presented in Tables 1 and 2 and figures 3 and 4.

The oldest zircon U-Pb ages compiled for volcanic rocks range from 309.9±2.2 to 273.5±11.0 Ma (Table 1; Fig. 3); these rocks are exposed from 20°30’ to 25°30’S in a ~50 km wide, longitudinal belt along ~69°W longitude along the Precordillera and the Domeyko Cordillera (Fig. 3), and according to the U-Pb data represent a Late Carboniferous to Early Permian, felsic volcanism and subordinate intermediate and basic volcanic rocks, and minor sedimentary intercalations (Collahuasi, Agua Dulce and La Tabla Formations, and the Sierra del Jardín Rhyolites; Vergara and Thomas, 1984; Mpodozis et al., 1993; Marinovic and García, 1999; Masterman, 2003; Cornejo et al., 2006; Marinovic, 2007; Munizaga et al., 2008; Tomlinson and Blanco, 2008; Urzúa, 2009; Jara et al., 2009; Hervé et al., 2012). The Late Carboniferous to Early Permian volcanism is less known southward of 25°30’S, but a U-Pb zircon age of 300.8±4.6 Ma has been reported by Salazar et al. (2013) for rhyolites outcropping in the El Tránsito river valley in the High Andes, 65 km west of the divide (28°46’S; Fig. 4).

Rhyolitic volcanic rocks exposed along the Domeyko Cordillera and Frontal Cordillera south of 25°30’S latitude have yielded U-Pb ages of 262.9±2.0 and 265.0±5.6 Ma (Table 1; Fig. 4), thus essentially representing a Mid-Permian felsic volcanism, which has been previously ascribed to the to the Permian-Triassic (Martin et al., 1999) or to the Carboniferous-Permian (Cornejo et al., 2009).

A U-Pb zircon U-Pb age of 248±3 Ma, resulting from combining analyses of four separate samples, was reported by Breitkreuz and van Schmus (1996) for felsic volcanic rocks of the Peine Group, exposed in on the eastern border of the Salar de Atacama salt flat at about 23°30’S (Fig. 3). Thus, these easternmost outcrops of northern Chile represent an Early to Middle Triassic volcanism, which originally was assigned to the Late Permian by Breitkreuz and van Schmus (1996), as befitted to the geologic time scale in that year.

Middle Triassic to Early Jurassic volcanism has yielded U-Pb dates from 249.7±3.8 to 194.0±9.2 Ma (Table 1). During this time span andesitic to basaltic volcanism coexisted with rhyolitic volcanism (bimodal volcanism), and was synchronous with rift-related, continental and/or marine sedimentary deposits widely distributed in northern Chile (Naranjo and Puig, 1984; Charrier, 1979; Suárez and Bell, 1992; Martin et al., 1999; Tomlinson et al., 1999; Cornejo and Mpodozis, 1996, 1997; Cornejo et al., 2009).

The zircon U-Pb ages for intrusive rocks exposed along the High Andes and Domeyko Cordillera range from 328.1±2.8 to 205.9±0.5 Ma (Table 2), overlapping with the corresponding interval of U-Pb dates of volcanic rocks (ca. 310 to 194 Ma; Table 1). The intrusions include coarse-grained granitoids and porphyritic felsic plutons; the late are closely related to the volcanic rocks. Foliated plutons (ductile deformation) show U-Pb ages in the range 328 to 288 Ma (Table 2).

There are scarce and inaccurate previous U-Pb dates for granitic plutons scattered along the Coastal Cordillera of northern Chile from 21° to 31°S, which range from 291.8±14.3 to 217±12 Ma (Berg et al., 1983; Berg and Baumann, 1985). Yet, a number of whole rock Rb-Sr dates that range from 296±5.4 to 222±3 Ma have been published (Berg et al., 1983; Shibata et al., 1984; Berg and Baumann, 1985), and biotite K-Ar dates of 318±6 and 322±5 Ma (Maksaev and Marinovic, 1980; Skarmeta and Marinovic, 1981). Besides, eight U-Pb ages ranging from 319.6±2.4 to 308.9±2.4 Ma have been recently obtained for the granitoids that are exposed from 33° to 38°S along the Coastal Cordillera of central-southern Chile (Deckart et al., 2014).

 

4. Analytical methods

SHRIMP II U-Pb analyses of zircon grains were carried out at the Geochronological Research Center (Centro de Pesquisas Geocronológicas) of the Sao Paulo University, Brazil. Zircon grains were obtained from rock samples following normal mineral separation procedures in Universidad de Chile, which included crushing, washing, and heavy liquid and magnetic separation. Hand-picked zircon grains were mounted in epoxy, together with chips of the Temora reference zircons, sectioned approximately in one half, and polished in Brazil. Cathodoluminescence (CL) Scanning Electron Microscope (SEM) images were prepared for all zircons. The CL images were used to decipher the internal structures of the sectioned grains and to ensure that the ~20 μm SHRIMP spot was wholly within a single age component within the sectioned grains. Uranium-Th-Pb analyses were made using a sensitive high resolution ion microprobe (SHRIMP II) for determining the respective zircon Concordia U-Pb ages following procedures given in Williams (1998, and references therein). Each analysis consisted of six scans through the mass range, with a Temora U-Pb reference grain analyzed for every three unknown analyses. The data have been reduced using the SQUID Excel Macro of Ludwig (2001); U-Pb ratios have been normalized relative to a value of 0.0668 for the Temora reference zircon, equivalent to an age of 417 Ma (Black et al., 2003). Uncertainties given for individual analyses (ratios and ages) are at the one-sigma level (Appendix 2). Tera and Wasserburg (1972) concordia plots, and mean U-Pb age calculations were carried out using ISOPLOT/EX (Ludwig, 2003). The data tabulations for all 43 samples are presented in Appendix 1 and 2.

5. Results

The new SHRIMP U-Pb zircon concordia ages are summarized in Tables 3 and 4 for volcanic and intrusive rocks, respectively and sample locations are shown in figures 3 and 4; Tera-Wasserburg concordia plots and SHRIMP zircon U-Pb analytical data are included in the Appendix 1 and 2, respectively.

5.1. Volcanic rocks

The new zircon U-Pb ages for volcanic rocks range from 324.7±4.0 to 221.0±2.6 Ma (Table 3), which is consistent with the ca. 310 to 194 Ma interval showed by the compiled U-Pb dates (Table 1). Nevertheless, we obtained some of the oldest age determinations for Late Paleozoic volcanic rocks in Chile to date, particularly for volcanic deposits exposed in the current Intermediate Depression and the western foothills of the High Andes. These include U-Pb ages of 324.7±4.0 and 323.1±5.8 for rhyolites outcropping in the El Tránsito river valley (at 28°46’S; Fig. 4), which were previously mapped as the Permian to Triassic Pastos Blancos Formation (Moscoso et al., 2010), but were recognized as an older separate unit named Cerro Bayo Formation by Salazar et al. (2013). Besides, a U-Pb age of 320.1±9.0 Ma was obtained for an ignimbrite outcropping in the Intermediate Depression of the Antofagasta Region, which is part of a rhyolitic volcanic succession that underlies and is partly interfingered with carbonate strata that contains Early Permian marine fossils in its upper section (Cerro del Árbol Formation; Marinovic et al., 1995; Marinovic, 2007). In addition, U-Pb ages of 296.6±4.2 and 292.2±4.2 Ma were obtained for rhyolites that are part of a succession exposed in the Paipote valley which was assigned to the Pantanoso Formation (Sepúlveda and Naranjo, 1982; Iriarte et al., 1996) and unconformably underlies Late Triassic conglomerates of La Ternera Formation in the Atacama Region (Fig. 4). The same rhyolitic unit underlies limestone beds with Early Permian marine fossils in the region, though the contact between them has been interpreted as a low-angle, normal fault (Quebrada de Las Represas Beds; Mpodozis and Allmendinger, 1993; Iriarte et al., 1996).

In the High Andes a U-Pb age of 291.5±4.0 Ma was obtained for an andesitic tuff intercalated in the upper section of El Mono Beds (sensu Mercado, 1982; Mpodozis et al., 2012), which is a mostly alluvial and lacustrine sedimentary succession, ca. 1,000 m thick, formed of sandstone and conglomerate strata that unconformably overlie Devonian to Early Carboniferous sandstones of the Chinches Formation in the High Andes at ~27°S (Fig. 4) and underlies in apparent conformity Jurassic marine strata (Mercado, 1982; Mpodozis et al., 2012). This sedimentary-volcanic succession was previously attributed to the Late Triassic based on fossil flora and conchostracans (Davidson et al., 1978; Mercado, 1982; Suárez and Bell, 1991; Blanco, 1994; Cornejo et al., 1998; Gallego and Covacevich, 1998; Mpodozis et al., 2012).

U-Pb dates ranging from 282.0±11.4 to 270.4±4.6 Ma were obtained for 4 samples of banded rhyolites and rhyolitic tuff from the type locality of La Tabla Formation, (sensu García, 1967), in the eastern part of the Domeyko Cordillera and southwest of the Pedernales Salar in the Atacama Region (26°30’S; Fig. 4). One sample (CHY-14; Table 3) returned a bimodal age distribution with an older population at 311.3±8.2 Ma and a younger one at 282.0±11.4; the analytic spots on the cathode luminescence images of the zircons indicate that the older age population corresponds to inherited zircon cores, while the younger ages to zircon rims, thus the younger population corresponding to the actual crystallization age of the dated rock. The La Tabla Formation is an 800 m thick succession formed of rhyolites, rhyolitic tuffs, breccias, and basaltic intercalations (without exposed base) which originally was thought to be Early Triassic in age (García, 1967), but later assigned to the Late Paleozoic by Tomlinson et al. (1999), based on K-Ar ages of intrusions that crosscut the unit. The volcanic rocks of La Tabla Formation are crosscut by granitoids of the Pedernales Batholith and andesitic dikes that yielded U-Pb ages from 287.0±4.2 to 264.6±7.0 Ma (Table 4). According to the new U-Pb dates it represent Permian volcanism, which is a bit younger but overlapping with the U-Pb ages from 309.9±2.2 to 282±2 Ma published for rocks assigned to the La Tabla Formation along the Domeyko Cordillera north of 25°30’S latitude (Table 1). In addition, a U-Pb age of 262.0±2.8 was obtained for a sample of fluidal rhyolite farther south (26°49’S; Table 3), mapped as La Tabla Formation by Cornejo et al. (1998) that is part of a succession crosscut by granite of the Caballo Muerto Pluton, which yielded a U-Pb age of 263.5±3.4 Ma (Table 4).

A U-Pb age of 270.3±4.2 Ma was obtained for a basal dacite horizon and additional U-Pb ages of 269.9±4.0 and 266.1±4.4 Ma were obtained for dacitic tuff and rhyolitic ignimbrite of the Pantanoso Formation (Mercado, 1982; Mpodozis et al., 2012), at its type locality (Table 3; Fig. 4). The Pantanoso Formation is crosscut by granite of the El Hielo batholith that yielded a U-Pb age of 267.9±3.8 Ma (Table 4). The Pantanoso Formation is a 1,000 thick succession of rhyolitic and dacitic volcanic rocks and minor conglomerate and volcanic sandstone intercalations exposed in the High Andes at 27°28’S latitude, and according to the new U-Pb data it rep-resent Mid-Permian volcanism. It unconformably overlies mainly Devonian to Early Carboniferous sandstones of the Chinches Formation (Mercado, 1982; Bell, 1985), but also a local lens of folded sandstone and conglomerate beds up to 200 m thick of the Elba Beds, which were ascribed to the Carboniferous-Permian by Mercado (1982).

U-Pb zircon ages of 255.9±6.2 and 242.7±5.8 Ma (Table 3) were obtained for samples of ignimbrite and rhyolite from the Cas Formation (sensu Ramírez and Gardeweg, 1982), which is a succession of about 300 m thick (without exposed base), outcropping immediately east of the Peine town, in the eastern part of the Antofagasta Region (Fig. 3). The new U-Pb ages are consistent with the mean weighted U-Pb age of 248±3 based on 4 separate determinations published by Breitkreuz and van Schmus (1996) and indicate that the rhyolitic rocks of the Cas Formation represent a Late Permian to Early Triassic felsic volcanism.

A sample of basaltic tuff from the basal meters of the mostly basaltic-andesitic La Totora Formation in the homonymous creek in the High Andes of the Vallenar region (28º42’; Fig. 4) yielded a U-Pb age of 249.7±3.8 Ma. The dated sample (CHY-56) actually returned a bimodal age distribution with an older population at 323.6±5.8 Ma; the analytic spots on the cathode luminescence images of the zircons indicate that the older age population corresponds to inherited zircon cores, while the younger ages to zircon rims, thus the younger population corresponding to the actual crystallization age of the dated rock. Another sample of basalt of the same succession in the Totora creek yielded an U-Pb age of 221.0±2.6 Ma; these ages are coherent with the U-Pb ages from 217.9±1.4 to 210.4±2.9 Ma published by Salazar et al. (2013) for the same formation. This volcanic unit unconformably overlies a mylonitized tonalite of the Quebrada Pintado intrusive unit (sensu Salazar et al., 2013) for which we obtained a U-Pb age of 258.9±5.0 Ma (Table 4), and conformably underlies fossiliferous, Mid-Late Jurassic (Sinemurian) marine limestones (Reutter, 1974; Nasi et al., 1990; Salazar et al., 2013); stratigraphic position that is consistent with the new U-Pb data. In addition, strata of the western belt of La Totora Formation are in part conformably overlying the 4,000 m thick, mostly clastic-sedimentary Mid-Late Triassic San Felix Formation to the southwest (Salazar et al., 2013).

U-Pb ages of 232.1±4.6, 224.0±3.4, and 221.6±3.4 were obtained for samples of banded rhyolites and rhyolitic tuffs from the Los Tilos sequence of the Pastos Blancos Group (sensu Martin et al., 1999), exposed in the Potrerillos river valley in the High Andes, 20 km west of the divide (29°30’S; Fig. 4); this new U-Pb data confirms the Late Triassic age of the Los Tilos volcanic sequence of the Pastos Blancos Group, which unconformably overlies the Permian silicic volcanics of the Guanaco Sonso sequence as defined by Martin et al. (1999).

A U-Pb age of 154.3±3.8 Ma was obtained for andesite that is part of a succession outcropping in the west flank of the El Carmen river valley (29°17’S; Fig. 4). This intermediate volcanic unit was originally mapped as part of the Late Paleozoic-Triassic Pastos Blancos Formation (with a question mark) by Nasi et al. (1990). Actually, the new U-Pb age shows that these andesitic rocks are Late Jurassic in age, being equivalent in age and lithology to the Picudo Formation (Reutter, 1974; Moscoso et al., 2010) exposed in the El Tránsito river valley and to the Lagunillas Formation, which has been recognized farther north between 27°30’ and 28°30´S in the High Andes of the Atacama Region and these represent Late Jurassic back-arc volcanism (Iriarte et al., 1996; Oliveros et al., 2011).

5.2. Intrusive rocks

The new zircon U-Pb ages for Late Paleozoic intrusive rocks range from 315.7±4.6 to 229.6±5.2 Ma (Table 4), well within the 328 to 206 Ma interval of the compiled U-Pb dates (Table 2).

Granitoids exposed along the Domeyko Cordillera and its southern continuation in the Frontal Cordillera yielded U-Pb ages from 287 to 263 Ma (Table 4; Fig. 4), representing Permian granitic plutonism, these include: a U-Pb age of 285.3±6.4 Ma for a foliated tonalite that is part of the Sierra Castillo batholith (sensu Tomlinson et al., 1999; Fig. 4); a U-Pb age of 287.0±4.2 Ma for a hornblende-biotite monzodiorite and U-Pb ages of 280.6±4.2, and 264.9±7.0 Ma for biotite-hornblende, monzogranite plutons of the Pedernales batholiths (sensu Tomlinson et al., 1999; Fig. 4); a U-Pb age of 263.5±3.4 Ma for monzogranite of the Caballo Muerto pluton (sensu Cornejo et al., 1998); and a U-Pb age of 267.9±3.8 Ma for leucocratic, monzogranite of the El Hielo batholith (sensu Mercado, 1982; Bell, 1985; Mpodozis et al., 2012) in the High Andes of the Atacama Region (Fig. 4).

A U-Pb age of 244.0±2.8 Ma was obtained for a coarse-grained monzogranite of the Chollay Batholith (Fig. 4), which is consistent with U-Pb ages from 247.7±3.4 to 240.3±1.7 Ma that have recently been reported for the same batholith (Coloma et al., 2012; Salazar et al., 2013; Álvarez et al., 2013; Hervé et al., 2014). These U-Pb ages confirm the Mid-Triassic age ascribed to the granitic Chollay batholith by Nasi et al. (1985) and Martin et al. (1999).

The dating of plutons of the northern part of the composite Elqui-Limarí Batholith in the Frontal Cordillera of the Atacama Region (Fig. 4) have yield U-Pb ages from 315.7±4.6 to 229.6±5.2 Ma (Table 4). These are consistent with the 330 to 215 U-Pb interval of recently published U-Pb ages for the same batholith (e.g., Pineda and Calderón, 2008; Salazar et al., 2009, 2013; Coloma et al., 2012, Álvarez et al., 2013; Hervé et al., 2014). We have obtained a U-Pb age of 315.7±4.6 Ma for a coarse-grained, biotite monzogranite from the Guachicay pluton (sensu Nasi et al., 1985) exposed in the Del Carmen river valley and 311.9±6.0 Ma for a coarse-grained, biotite syenogranite exposed in the Potrerillos river valley. The latter was included in the extensive tonalitic Guanta pluton by Nasi et al. (1985, 1990). In addition, we obtained a zircon U-Pb age of 307.1±4.8 Ma for a foliated, coarse-grained, biotite-hornblende tonalite of the Guanta pluton (sensu Nasi et al., 1985, 1990) exposed in the Del Carmen river valley, and a U-Pb zircon age of 286.3±6.2 Ma was obtained for a coarse-grained, muscovite-biotite granite of the Burro Muerto pluton (sensu Nasi et al., 1985), which is emplaced within the foliated tonalite of the Guanta pluton in the El Carmen river valley, and corresponds to the Cochiguaz intrusive unit of Nasi et al. (1985, 1990); and a U-Pb zircon age of 229.6±5.2 Ma was obtained for a brick-red, fine-grained, syenogranite of the La Coneja pluton (sensu Nasi et al., 1985) that corresponds to the Colorado intrusive unit of Nasi et al. (1985, 1990).

A U-Pb zircon age of 258.9±5.0 Ma (Late Permian) was obtained for mylonitized tonalite of the Quebrada Pintado tonalitic unit of Salazar et al. (2013) in the La Totora creek (28°40’S; Fig 4). The mylonites at La Totora creek show a N35°E/75°E foliation and centimetric to decimetric white, felsic bands alternating with dark gray, biotitic bands. This mylonitized intrusion unconformably underlies the Mid-Late Triassic volcanic rocks of La Totora Formation.

Along the Pacific coast north of Chañaral, about 120 km west of the High Andes, U-Pb ages of 276.6±3.6 and 264.6±7.0 Ma were obtained for biotite-muscovite, monzogranite of the Pan de Azúcar pluton (sensu Mercado, 1980; Godoy and Lara, 1998), while two U-Pb ages of 246.5±4.0 and 237.0±4.0 were obtained for coarse-grained, muscovite-biotite monzogranite of the Cerros del Vetado pluton of the Coastal Cordillera immediately east of Chañaral (Fig. 4).

6. Discussion

The U-Pb data provide evidence for the onset of volcanism at ~325 Ma (by the end of the Missisippian), which is approximately coeval with the onset of granitic plutonism at ~328 Ma (Tables 2 and 3), confirming a previous conclusion of Hervé et al. (2014). This marks reestablishment of magmatism during the Carboniferous after a gap in igneous activity during the Devonian in northern Chile and Perú (e.g., Charrier et al., 2007; Boekhout et al., 2013).

6.1. Volcanic rocks

The oldest rhyolitic rocks (325 to 292 Ma; Table 1 and 3) are present within the current Intermediate Depression and the western foothills of the High Andes of northern Chile (Figs. 3 and 4). Part of this initial volcanism developed within a shallow marine platform represented by carbonate strata with Late Permian marine fossils that are overlying and partly interfingered with the dated rhyolitic rocks (Juan de Morales Formation in the Tarapacá Region; Galli, 1956, 1968; Díaz-Martínez et al., 2000; Cerro del Árbol Formation in Antofagasta Region; Marinovic et al., 1995; Marinovic, 2007; and Quebrada de Las Represas Beds in the Atacama Region; Iriarte et al., 1996). While, the rhyolitic succession exposed in the El Tránsito river valley includes conglomerates, sandstones and andesites (Cerro Bayo Formation; Salazar, 2012; Salazar et al., 2013).

The mainly felsic volcanic rocks widely exposed along the Domeyko Cordillera and Precordillera of northern Chile north of 25°30’S latitude have yielded U-Pb ages from 310 to 282 Ma Tables 1 and 3), thus represent a Late Carboniferous to Early Permian volcanism that is older than the Choiyoi succession of the San Rafael Block in western Argentina for which U-Pb ages from 281 to 251 Ma have been published (Rocha-Campos et al., 2011). Actually, the Chilean Carboniferous to Early Permian volcanic rocks are chronologically equivalent to sedimentary successions with submarine volcanic deposits that are present farther east in Argentina (Azcuy et al., 1999; López-Gamundí, 2006; Koukharsky et al., 2009; Gulbranson et al., 2010). Therefore, when the Carboniferous to Early Permian magmatism was active along the current Domeyko Cordillera and Precordillera northwards, in Argentina farther east developed contemporary marine basins; this situation has been interpreted as a magmatic arc and back-arc basins by Gulbranson et al. (2010).

The U-Pb age of 291.5±4.0 Ma that we obtained for an andesitic tuff intercalated in the upper part of El Mono Beds at its type locality (Mpodozis et al., 2012), in the High Andes at ~27°S, indicates an Early Permian age for the upper part of the succession. Even though it underlies in apparent conformity Jurassic marine strata, implying a hiatus in what was considered a continuous sedimentary sequence (e.g., Blanco, 1994). The mostly clastic, alluvial and lacustrine sedimentary El Mono Beds have previously been assigned to the Late Triassic, based mainly on fossil flora attributed to Dicroidium sp. (Mercado, 1982; Suárez and Bell, 1991; Blanco, 1994; Cornejo et al., 1998; Gallego and Covacevich, 1998; Mpodozis et al., 2012). We cannot exclude the possibility that Triassic sedimentary rocks exists in the same region, but the new U-Pb age of 291.5±4.0 Ma is consistent with the stratigraphic position of the El Mono Beds unconformably overlying Devonian to Early Carboniferous sandstones of the Chinches Formation (Bell, 1985; Mpodozis et al., 2012), and evidences that this unit represents a Late Carboniferous to Early Permian sedimentation and coetaneous pyroclastic volcanism, at least where sampled for dating. Therefore, El Mono Beds could be equivalent to a sedimentary unit with Late Carboniferous flora and marine fossils exposed near the watershed of the Andes at ~29°S latitude (Las Placetas Formation; Reutter, 1974; Nasi et al., 1990); these sedimentary deposits probably represent the westernmost part of sedimentary basins developed mainly farther east in Argentina during the Carboniferous to Early Permian (Azcuy et al., 1999; López-Gamundí, 2006; Koukharsky et al., 2009; Gulbranson et al., 2010).

Rhyolitic volcanism with andesitic and basaltic intercalations persisted during the Permian along the current Domeyko Cordillera and the Chilean Frontal Cordillera (High Andes) as the volcanic rocks exposed south of 25°30’S latitude have yielded U-Pb ages from 282 to 262 Ma (Figs. 3 and 4). These are represented by the La Tabla, and Pantanoso Formations, and the Guanaco Sonso Sequence of the Pastos Blancos Group (Mercado, 1982; Cornejo et al., 1993, 1998, 2009; Martin et al., 1999; Tomlinson et al., 1999; Mpodozis et al., 2012) and their U-Pb ages fall well within the 281 to 251 Ma time span of U-Pb ages published by Rocha-Campos et al. (2011) for the Choiyoi Succession of the San Rafael Block in western Argentina.

The U-Pb age of 270.3±4.6 Ma for the basal unit of the Pantanoso Formation (Table 3) indicates the onset of volcanic accumulation in the current High Andes at ~27°S latitude unconformably on top of deformed Devonian to Early Carboniferous sedimentary rocks (Chinches Formation; Mercado, 1982; Bell, 1985), but also over folded sandstone and conglomerate beds ascribed to the Carboniferous-Permian (Elba Beds; Mercado, 1982). Thus, according to the new U-Pb data the accumulation of the volcanic rocks of the Pantanoso Formation at its type locality occurred at least 20 My later than the rhyolitic succession that is exposed 47 km to the west in the Paipote valley, which in previous works has also been assigned to the Pantanoso Formation (Iriarte et al., 1996).

Late Permian to Early Triassic rhyolitic volcanism is represented by the Cas Formation that yielded U-Pb ages from 256 to 243 Ma (Fig. 3; Tables 1 and 3) and is exposed in the easternmost part of Chile (68°W), along the eastern side of the Salar de Atacama salt flat in the Antofagasta Region (Ramírez and Gardeweg, 1982; Breitkreuz and van Schmus, 1996). The rhyolitic rocks of the Cas Formation at its type locality are significantly younger than the Early Permian rhyolitic rocks exposed some 90 km to the west along the Domeyko Cordillera (e.g., Sierra del Jardín Rhyolites; Marinovic, 2007), even though the latter were mapped as the same formation by Mpodozis et al. (1993).

The occurrence of Mid-Permian to Early Triassic volcanic rocks eastward from the Carboniferous to Early Permian volcanic rocks in Chile is interpreted as a Mid-Permian to Early Triassic expansion and/or eastward migration of the volcanic front, which is consistent with evidences of expansion of the volcanic activity farther east in Argentina at that time (e.g., Ramos and Folguera, 2009; Ramos, 2009).

Middle Triassic to Early Jurassic volcanic rocks yielded zircon U-Pb ages from 249.7±3.8 to 194.0±9.2 Ma, their outcrops are more discontinuous (patchy) and dispersed compared to the older volcanic rocks, and are separated from them by erosional surfaces. During this period andesitic to basaltic volcanism coexisted with felsic volcanism (bimodal volcanism), and was synchronous with rift-related, continental and/or marine sedimentary deposits in northern Chile (e.g., Suárez and Bell, 1992; Bell and Suarez, 1994). Middle to Upper Triassic strata show strong facies changes along the longitudinal axis of the High Andes, for example according to the U-Pb data the mostly basaltic-andesitic La Totora Formation (250-210 Ma) overlaps in age with the upper part of the mainly clastic, sedimentary San Félix Formation (223-212 Ma), but also with rhyolites of Los Tilos Sequence (232-221 Ma), within less than 100 km in the Atacama Region. These striking north-south facies changes probably reflect the main northwest orientation of extensional basins that developed in the Mid-Late Triassic in Chile and Argentina (Charrier, 1979; Chong and Von Hillebrandt, 1985; Uliana et al., 1989; Ramos and Kay, 1991; Jenchen and Rosenfeld, 2002; Charrier et al., 2007; Barredo et al., 2012).

A probability density plot and stacked histogram of all U-Pb ages for volcanic rocks exposed along the segment of the Chilean Andes from 20° to 31°S (Fig. 5) evidences a protracted volcanic activity that extended for about 130 million years, from about 325 to 194 Ma (Late Mississippian to Early Jurassic). The histogram (Fig. 5) shows two largest concentrations of U-Pb dates from 310 to 250 Ma and from 240 to 210 Ma, and other less prominent concentrations of U-Pb ages from 330 to 320 Ma and at about 190 Ma. Even though the histogram probably is biased to zircon-bearing rocks and easily accessible outcrops for sampling, the statistical analysis of the U-Pb data may signify that major volcanic episodes extended from the Late Carboniferous to Late Permian and then from Mid- to Late Triassic, but were preceded by a mid-Carboniferous volcanic episode and followed by an Early Jurassic one.

 

fig.5

 

Fig. 5. Probability density plot and stacked histogram of zircon U-Pb ages for volcanic rocks.

 

6.2 Intrusive rocks

U-Pb data for plutonic rocks show that earliest intrusions crystallized from 328.1±2.8 to 311.9±6.0 Ma, including coarse-grained, biotite-muscovite granite to granodiorite (S-type) and biotite monzogranite and biotite-hornblende granodiorite plutons (I-type). From 308.5±2.2 to 290±2 Ma rhyolitic and granitic porphyries and equigranular monzogranite plutons crystallized and biotite-muscovite granitic and granodioritic plutons, and hornblende-diorite tonalitic to dioritic plutons were emplaced. Furthermore, 33 compiled and 9 new U-Pb zircon ages, which altogether range from 298.9±2.6 to 252.5±1.8 Ma attest that plutonism was active during the whole Permian in northern Chile, when there was a strong predominance of coarse-grained monzogranite plutons, both equigranular and porphyritic, followed by coarse-grained biotite-hornblende tonalites, and only minor monzonite, diorite and gabbro intrusions; two-mica (S-type) granites, granodiorites, and tonalites also intruded during this time span. Most of the Permian plutons are exposed along the High Andes and Domeyko Cordillera, but Permian granites with U-Pb ages of 277-270 Ma are also present some 120 km west in the Pacific coast (Pan de Azúcar pluton, Fig 4).

Triassic plutonism is represented by granitic porphyries, biotite- and two-mica granites, granodiorites, and brick-red syenogranite intrusions with U-Pb ages that range from 247.0±3.1 to 215.6±1.9 Ma. Some of these intrusions are of large volume, as most of the Chollay and Montosa granitic batholiths are formed of Triassic medium to coarse-grained monzogranites for which U-Pb ages of 245.0±3.1 and 240.3±1.7 Ma have been reported (Coloma et al., 2012; Hervé et al., 2014) and we obtained a U-Pb age of 244.0±2.8 Ma. Triassic granitic plutons are also present in the current Coastal Cordillera as Cerros del Vetado pluton yielded U-Pb ages of 247-237 Ma (Fig. 4).

The U-Pb data now available evidence that muscovite-bearing intrusive, with S-type granite characteristics, were not confined to a single discrete Permian magmatic episode as previously thought, but crystallized within the entire time span from Late Carboniferous to Late Triassic and synchronically with biotite- and hornblende-bearing intrusions with petrographic characteristics of I-type granitoids; this is in agreement with a similar conclusion based on U-Pb data for the Elqui-Limarí Batholith (Hervé et al., 2014). Actually, the synchronic emplacement of I- and S-Type granitoids was already noted and sustained by chemical data by Marinovic et al. (1995) in the Domeyko Cordillera, though it was attributed to a local phenomenon related to a hypothetical longitudinal segmentation of the Late Paleozoic to Triassic plutonism, rather than to a regional characteristic.

The new U-Pb data from 277 to 237 Ma (Table 4) confirm that Permian and Mid-Triassic granitic magmatism is also exposed in the current coastal area, which combined with previously published geochronological data suggests a protracted plutonism extending from 322 to 217 Ma along the Coastal Cordillera of northern Chile from 21° to 33°S (Maksaev and Marinovic, 1980; Skarmeta and Marinovic, 1981; Berg et al., 1983; Shibata et al., 1984; Berg and Baumann, 1985). However, there appears to be a difference with the Coastal Batholith exposed to the south between 33° to 38°S, which has yielded eight 320 to 300 Ma (Pennsylvanian) U-Pb zircon ages (Deckart et al., 2014), but to date there are no Permian and Early Triassic granitoids intruded in the main body of this Coastal Batholith of central-southern Chile, and only some small Late Triassic-Early Jurassic stocks of calc-alkaline to transitional A-type affinities, with U-Pb zircon ages from 225 to 197 Ma that are emplaced into its metamorphic envelope (Vásquez et al., 2011; Hervé et al., 2014).

A probability density plot and stacked histogram of all U-Pb ages for intrusive rocks (Fig. 6) shows that plutonic activity extended for at least 120 My from about 328 to 206 Ma (Late Missisipian to Late Triassic) with a largest concentration of U-Pb dates between 310 and 280 Ma, followed by another significant concentration of U-Pb dates from 255 to 205 Ma, and other less prominent U-Pb age concentrations from 330 to 310 Ma and 280 to 260 Ma. This statistical analysis of the U-Pb data may signify that major intrusive episodes took place from the Late Carboniferous to Early Permian and from the Late Permian to Late Triassic, but also there were mid-Carboniferous and mid-Permian intrusive episodes. The inferred plutonic episodes are fundamentally coincident with the major episodes of volcanism inferred from the statistical analysis of U-Pb data (Figs. 5 and 6). A sequential plot of the U-Pb ages (Fig. 7) shows that there are no apparent periods when the volcanism or plutonism prevailed, but both developed simultaneously from the Carboniferous to the Early Jurassic. A probability density plot and stacked histogram of all U-Pb ages for igneous rocks (Fig. 8) show the largest concentration of dates from 310 to 260 Ma and from 255 to 205 Ma, which are interpreted as the major magmatic episodes developed during from the Late Carboniferous to Mid-Permian and from Late Permian to Late Triassic. Other less prominent U-Pb age concentrations reveal a mid-Carboniferous episode (330 to 320 Ma) and an Early Jurassic episode (200-190 Ma).

 

fig.6

 

Fig. 6. Probability density plot and stacked histogram of zircon U-Pb ages for intrusive rocks.

 

fig.7

 

Fig. 7. Plot of all zircon U-Pb ages for Late Paleozoic to Early Jurassic intrusive and volcanic rocks ordered from older to younger with 2σ error bars; black bars=intrusive rocks and gray bars=volcanic rocks.

 

 

fig.8

 

Fig. 8. Probability density plot and stacked histogram of all zircon U-Pb ages for igneous rocks.

 

The magmatic episodes of the Chilean Andes fall well within the 336 to 230 Ma interval of published U-Pb geochronological data for igneous rocks of Argentina (Pankhurst, 2006; Gulbranson et al., 2010; Rocha-Campos et al., 2011; Barredo et al., 2012; Suárez et al., 2012; Gregori and Benedini, 2013), but also with the interval from 342 to 190 of U-Pb data from the Andes of Perú, Ecuador and Colombia (Restrepo et al., 1991; Noble et al., 1997; Ordoñez and Pimentel, 2002; Vinasco et al., 2006; Chew et al., 2007; Ibáñez-Mejía et al., 2008; Cardona et al., 2008, 2009; Mišković et al., 2009, Mišković and Shaltegger, 2009; Witt et al., 2013; Boekhout et al., 2013). Therefore, despite local particularities, there was a rather consistent evolution in time of the magmatism along the southwestern, paleo-Pacific border of Gondwana.

7. Conclusions

U-Pb zircon geochronological data provide record of about 130 Ma of igneous activity in the Andes of northern Chile that extended episodically from the latest Early Carboniferous to Early Jurassic (328-194 Ma). The overall U-Pb data show that volcanism and plutonism were essentially synchronous and inferred major episodes of igneous activity developed during the Late Carboniferous to Permian (310 to 260 Ma) and from Late Permian to Late Triassic (255-205 Ma), with less prominent episodes in the mid-Carboniferous (330 to 320 Ma), and Early Jurassic (200-190 Ma). Thus, from the Carboniferous to the Early Triassic dominantly silicic magmatism developed along the Chilean segment of the southwestern border of Gondwana supercontinent. Further magmatism developed during the Mid-Late Triassic (250-194 Ma) was bimodal and synchronous with rift-related, continental and/or marine sedimentary strata related to the early stages of break-up of Gondwana.

Most of the siliceous volcanic rocks of the Precordillera and Domeyko Cordillera of northern Chile (21°30’ to 25°30’S) are older than the silicic rocks assigned to the Choiyoi succession in Argentina, being instead equivalent in age to Carboniferous to Early Permian marine sedimentary sequences present in the eastern Argentinean foreland. On the other hand, silicic volcanic successions exposed in the easternmost part of northern Chile are equivalent in age to the Choiyoi succession of the San Rafael Block of Argentina. An eastward expansion or migration during the Mid-Permian to Early Triassic is inferred by the age difference between older western rhyolitic deposits and younger eastern rhyolitic volcanic that originally were supposed to be of the same age; interpretation that is consistent with expansion of the volcanism at that time in Argentina.

The timing of the magmatism is coincident with that of the Andes of Perú and of Argentina according to the available U-Pb data, revealing a rather consistent evolution in time of the magmatism along the southwestern, paleo-Pacific border of Gondwana.

The correct allocation of lithostratigraphic units to the geological time results in changes in the stratigraphy modifying the tectonic interpretation and evolution of the studied volcano-plutonic event. The current U-Pb data allow a better understanding of the temporal evolution of the Late Paleozoic to Early Jurassic volcano-plutonic event in northern Chile, but certainly further work is required to solve particularly stratigraphic problems still remaining and some outcrops that have not been sampled by logistic reasons.

 

Acknowledgements
This study was supported by CONICYT, Chile, through Fondecyt 1110093 grant to V. Maksaev and F. Munizaga (Universidad de Chile). We are obliged to F. Hervé and his Fondecyt project 1095099 for sharing data from the Elqui-Limarí batholith and provided some samples for dating during this study. We thank K. Sato in charge of the SHRIMP and the staff of the Geochronological Research Center (Centro de Pesquisas Geocronológicas) of the University of Sao Paulo, Brazil, for their cooperation during the U-Pb dating on the SHRIMP. Reviews by V. Oliveros and C. Mpodozis contributed to improve this paper.

 

References

Álvarez, J.; Mpodozis, C.; Blanco-Quintero, I.; García-Casco, A.; Arriagada, C.; Morata, D. 2013. U-Pb ages and metamorphic evolution of the La Pampa Gneisses: Implications for the evolution of the Chilenia Terrane and Permo-Triassic tectonics of north Central Chile. Journal of South American Earth Sciences 47: 100-115. doi: 10.1016/j.jsames.2013.07.001.

 

Azcuy, C.L.; Carrizo, H.A.; Caminos, R. 1999. Carbonífero y Pérmico de las Sierras Pampeanas, Famatina, Precordillera, Cordillera Frontal y Bloque de San Rafael. Geología Argentina, Anales del Instituto de Geología y Recursos Minerales 29: 261-318.

 

Barredo, S.; Chemale, F.; Marsicano, C.; Ávila, J.N.; Ottone, E.G.; Ramos, V. 2012. Tectono-sequence stratigraphy and U-Pb zircon ages of the Rincón Blanco Depocenter, northern Cuyo Rift, Argentina. Gondwana Research 21: 624-536.

 

Basso, M. 2004. Carta Baquedano, Región de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 82: 26 p. 1 mapa escala 1:100.000. Santiago.

 

Basso, M.; Marinovic, N. 2003. Antecedentes geocronológicos de volcanismo Triásico en la zona de los Estratos El Bordo, Antofagasta, Chile. In Congreso Geológico Chileno, No. 10, Sesión Temática 5: CD-ROM. Concepción.

 

Bell, C.M. 1985. The Chinches Formation: an early Carboniferous lacustrine succession in the Andes of northern Chile. Revista Geológica de Chile 24: 29-48. doi: 10.5027/andgeoV12n1-a03.

 

Bell, C.M.; Suárez, M. 1994. The sedimentation and tectonics of a marine fan-delta developed on anactive continental margin: the Triassic San Felix Formation in the Andes of northern Chile. Journal of South American Earth Sciences 7 (3-4): 403-413.

 

Berg, K.; Baumann, A. 1985. Plutonic and metasedimentary rocks from the Coastal Range of northern Chile: Rb-Sr and U-Pb systematic. Earth and Planetary Science Letters 75: 101-115.

 

Berg, K.; Breitkreuz, C.; Damm, K-W.; Pichowiak, S.; Zeil, W. 1983. The North-Chilean Coast Range-an Example for the Development of an Active Continental Margin. Geologische Rundschau 72 (2): 715-731.

 

Black, L.P.; Kamo, S.L.; Allen, C.M.; Aleinikoff, J.N.; Davis, D.W.; Korsch, R.J.; Foudoulis, C. 2003. TEMORA1: a new zircon standard for phanerozoic U-Pb geochronology. Chemical Geology 200: 155-170.

 

Blanco, N. 1994. Sedimentología y ambientes sedimentarios de los estratos El Mono, Triásico de la Región de Atacama, Chile. In Congreso Geológico Chileno, No. 7, Actas 1: 409-413. Concepción.

 

Boekhout, F.; Sempere, T.; Spikings, R.; Schaltegger, U. 2013. Late Paleozoic to Jurassic chronostratigraphy of coastal southern Perú: Temporal evolution of sedimentation along an active margin. Journal of South American Earth Sciences 47: 179-200.

 

Breitkreuz, C. 1995. The Late Permian Peine and Cas Formations at the Eastern margin of the Salar de Atacama, Northern Chile: stratigraphy, volcanic facies, and tectonics. Revista Geológica de Chile 22 (1): 3-23. doi: 10.5027/andgeoV22n1-a01.

 

Breitkreuz, C.; Zeil, W. 1994. The Late Carboniferous to Triassic Volcanic Belt in Northern Chile. In Tectonics of the Southern Andes (Reutter, K.J.; Scheuber, E.; Wigger, P.J.; editors). Springer: 277-292.

 

Breitkreuz, C.; Van Schmus, W.R. 1996. U-Pb geochronology and significance of Late Permian ignimbrites in Northern Chile. Journal of South American Earth Sciences 9: 281-293.

 

Breitkreuz, C.; Bahlburg, H.; Delakowitz, B.; Pichowiak, S. 1989. Paleozoic volcanic events in the Central Andes. Journal of South American Earth Sciences 2: 171-189.

 

Brown, M. 1990. Comparative geochemical interpretation of Permian-Triassic plutonic complexes of the Coastal Range and Altiplano (25º30’ to 26º30’S), northern Chile. In Andean Magmatism and its Tectonic Setting (Harmon, R.S.; Rapela, C.E.; editors). Geological Society of America, Special Paper 265: 157-177.

 

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

 

Cardona, A.; Cordani, U.; Nutman, A. 2008. U-Pb SHRIMP zircon, 40Ar/39Ar geochronology and Nd isotopes from gneissic and granitoid rocks of the Illescas massif, Perú: a southern extension of a fragmented Late Paleozoic orogen? In South American Symposium on Isotope Geology, No. 6. Proceedings, Abstract: p. 78. San Carlos de Bariloche.

 

Cardona, A.; Cordani, U.G.; Ruiz, J.; Valencia, V.A.; Armstrong, R.; Chew, D.; Nutman, A.; Sanchez, W. 2009. U-Pb Zircon Geochronology and Nd Isotopic Signatures of the Pre-Mesozoic Metamorphic Basement of the Eastern Peruvian Andes: Growth and Provenance of a Late Neoproterozoic to Carboniferous Accretionary Orogen on the Northwest Margin of Gondwana. The Journal of Geology 117 (3): 285-305.

 

Carlier, G.; Grandin, G.; Laubacher, G.; Marocco, R.; Megard, F. 1982. Present knowledge of the magmatic evolution of the eastern Cordillera of Perú. Earth-Science Reviews 18: 253-283.

 

Casquet, C.; Hervé, F.; Pankhurst, R.J.; Baldo, E.; Calderón, M.; Fanning, C.M.; Rapela, C.W.; Dahlquist, J. 2014. The Mejillonia suspect terrane (Northern Chile): Late Triassic fast burial and metamorphism of sediments in a magmatic arc environment extending into the Early Jurassic. Gondwana Research 25 (3): 1272-1286. doi: 10.1016/j.gr.2013.05.016.

 

Charrier, R. 1979. El Triásico en Chile y regiones adyacentes de Argentina: Una reconstrucción paleogeográfica y paleoclimática. Universidad de Chile, Departamento de Geología. Comunicaciones 26: 1-47. Santiago.

 

Charrier, R.; Pinto, L.; Rodríguez, M.P. 2007. Tectono-stratigraphic evolution of the Andean Orogen in Chile. In The Geology of Chile (Moreno, T.; Gibbons, W.; editors). The Geological Society: 21-114. London.

 

Chew, D.M.; Schaltegger, U.; Kosler, J.; Whitehouse, M.J.; Gutjahr, M.; Spikings, R.A.; Miskovíc, A. 2007. U-Pb geochronologic evidence for the evolution of the Gondwanan margin of the north-central Andes. Geological Society of America Bulletin 119 (5-6): 697-711.

 

Chong, G.; Von Hillebrandt, A. 1985. El Triásico preandino de Chile entre los 23º30’ y 26º00’ de lat. Sur. In Congreso Geológico Chileno, No. 4, Actas 1: 138-154. Antofagasta.

 

Cohen, K.M.; Finney, S.; Gibbard, P.L. 2013. International Chronostratigraphic Chart, v2013/01. International Commission on Stratigraphy. http//www.stratigraphy.org/ICSchart/ ChronostratChart2013-01.pdf: 1 p.

 

Coloma, F.; Salazar, E.; Creixell, C. 2012. Nuevos ante-cedentes acerca de la construcción de los plutones Pérmicos y Permo-Triásicos en el valle del río Tránsito, Región de Atacama, Chile. In Congreso Geológico Chileno, No. 13, Actas: 330-332. Antofagasta.

 

Cornejo, P.; Mpodozis, A.C. 1996. Geología de la región de Sierra Exploradora (25°-26ºS). Servicio Nacional de Geología y Minería, Informe Registrado IR-96-09: 330 p. Santiago.

 

Cornejo, P.; Mpodozis, C. 1997. El rift Triásico-Sinemuriano de Sierra Exploradora, Cordillera de Domeyko (25°-26°): asociaciones de facies y reconstrucción tectónica. In Congreso Geológico Chileno, No. 8, Actas 1: 550-554. Antofagasta.

 

Cornejo, P.; Mpodozis, C.; Ramírez, C.F.; Tomlinson, A. 1993. Estudio geológico de la región de Potrerillos y El Salvador (26°-27° Lat. S). Servicio Nacional de Geología y Minería, Informe Registrado IR-93-01: 258 p. Santiago.

 

Cornejo, P.; Mpodozis, A.C.; Tomlinson, A.J. 1998. Hoja Salar de Maricunga, Región de Atacama. Servicio Nacional de Geología y Minería, Mapas Geológicos 7, 1 mapa escala 1:100.000. Santiago.

 

Cornejo, P.; Matthews, S.; Marinovic, N.; Pérez de Arce, C.; Basso, M.; Alfaro, J.; Navarro, M. 2006. Alteración hidrotermal y mineralización recurrente de Cu y Cu-Mo durante el Pérmico y el Triásico en la Cordillera de Domeyko (Zona de Zaldívar-Salar de los Morros): antecedentes geocronológicos U-Pb, 40Ar/39Ar y Re-Os. In Congreso Geológico Chileno, No. 11, Actas 2: 219-222. Antofagasta.

 

Cornejo, P.; Mpodozis, C.; Rivera, O.; Matthews, S.J. 2009. Carta Exploradora, Región de Antofagasta y Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 119: 100 p. Santiago.

 

Cortés, J. 2000. Hoja Palestina, Región de Antofagasta. Servicio Nacional de Geología y Minería, Mapas Geológicos 19, 1 mapa escala 1:100.000. Santiago.

 

Damm, W.; Pichowiak, S.; Harmon, R.S.; Todt, W.; Kelley, S.; Omarini, R.; Niemeyer, H. 1990. Pre-Mesozoic evolution of the Central Andes, the basement revisited. In Plutonism from Antarctica to Alaska (Kay, M.S.; Rapela, C.W.; editors). Geological Society of America, Special Papers 241: 101-126.

 

Davidson, J.; Mpodozis, C.; Bunger, C.; Cornejo, P.; Faunes, A.; Foerster, H.; Fuster, N.; Gardeweg, M.; Godoy, S.; Jannas, R.; Muzzio, G.; Reyes, J.C.; Salinas, P.; Sepúlveda, J.; Solano, A.; Velasco, L. 1978. Geología de la Precordillera de Copiapó: las nacientes de la quebrada Paipote al oeste del Salar de Maricunga (resultados del curso de Geología de Campo II, 1977). Universidad de Chile, Departamento de Geología, Comunicaciones 23: 1-34.

 

Davidson, J.; Ramírez, C.F.; Gardeweg, M.; Brook, M.; Pankhurst, M. 1985. Calderas del Palaeozoico superior-Triásico Inferior y mineralización asociada. Universidad de Chile, Departamento de Geología, Comunicaciones 35: 53-57. Santiago.

 

Deckart, K.; Hervé, F.; Fanning, M.; Ramírez, V.; Calderón, M.; Godoy, E. 2014. Geocronología U-Pb e isótopos de Hf-O en circones del batolito de la costa Pensilvaniana, Chile. Andean Geology 41 (1): 49-82. doi: 10.5027/andgeoV41n1-a03.

 

Díaz-Martínez, E.; Mamet, B.; Isaacson, P.E.; Grader, G.W. 2000. Permian marine sedimentation in northern Chile: new paleontological evidence from Juan de Morales Formation, and regional paleogeographic implications. Journal of South American Earth Sci-ences 13 (6): 511-525.

 

Franz, G.; Lucassen, F.; Kramer,W.; Trumbull, R.B.; Romer, R.L.; Wilke, H.-G.; Viramonte, J.G.; Becchio, R.; Siebel, W. 2006. Crustal evolution at the Central Andean continental margin: a geochemical record of crustal growth, recycling and destruction. In The Andes-Active Subduction Orogeny (Oncken, O.; Chong, G.; Franz, G.; Giese, P.; Götze, H.-J.; Ramos, V.; Strecker, M.R.; Wigger, P.; editors). Springer: 45-64.

 

Gallego, O.F.; Covacevic, V. 1998. Conchostracos triásicos de las regiones de Antofagasta, Atacama y Coquimbo, Chile. Revista Geológica de Chile 25 (2): 115-131. doi: 10.5027/andgeoV25n2-a01.

 

Galli, C. 1956. Nota sobre el hallazgo del Palaeozoico superior en la Provincia de Tarapacá. Revista Minerales 53-54: 30-31.

 

Galli, C. 1968. Cuadrángulo Juan de Morales, Provincia de Tarapacá. Instituto de Investigaciones Geológicas, Carta Geológica de Chile 18: 53 p. Santiago.

 

García, F. 1967. Geología del Norte Grande de Chile. In Simposio sobre el Geosinclinal Andino. Sociedad Geológica de Chile, Publicación 3: 138 p. Santiago.

 

Godoy, E.; Lara, L. 1998. Hojas Chañaral y Diego de Almagro, Región de Atacama. Servicio Nacional de Geología y Minería, Mapa Geológico 5-6, 1 mapa escala 1:100:000. Santiago.

 

Gregori, D.; Benedini, L. 2013. The Cordón del Portillo Permian magmatism, Mendoza, Argentina, plutonic and volcanic sequences at the western margin of Gondwana. Journal of South American Earth Sciences 42: 61-73.

 

Gulbranson, E.L.; Montañez, I.P.; Schmitz, M.D.; Limarino, C.O.; Isbell, J.L.; Marenssi, S.A.; Crowley, J.L. 2010. High-precision U-Pb calibration of Carboniferous glaciation and climate history, Paganzo Group, NW Argentina. Geological Society of America, Bulletin 122 (9-10): 1480-1498.

 

Hervé, F.; Fanning, C.M.; Calderón, M.; Mpodozis, C. 2014. Early Permian to Late Triassic batholiths of the Chilean Frontal Cordillera (28°-31°S): SHRIMP U-Pb zircon ages and Lu-Hf and O isotope systematic. Lithos 184-187: 436-446. doi: 10.1016/j.lithos.2013.10.018.

 

Hervé, M.; Sillitoe, R.H.; Wong, C.; Fernández, P.; Crignola, F.; Ipinza, M.; Urzúa, F. 2012. Geologic Overview of the Escondida Porphyry Copper District, Northern Chile. In Geology and Genesis of Major Copper Deposits and Districts of the World: A Tribute to Richard H. Sillitoe (Hedenquist, J.W.; Harris, M.; Camus, F.; editors). Society of Economic Geologists, Special Publication 16: 55-78.

 

Huete, C.; Maksaev, V.; Moscoso, R.; Ulriksen, C.; Vergara, H. 1977. Antecedentes geocronológicos de rocas intrusivas y volcánicas de la Cordillera de los Andes comprendida entre la Sierra de Moreno y el río Loa y los 21º y 22º de latitud Sur, II Región, Chile. Revista Geológica de Chile 4: 35-41. doi: 10.5027/andgeoV4n1-a03.

 

Ibáñez-Mejía, M.; Jaramillo-Mejía, J.M.; Valencia, V. 2008. U-Th/Pb zircon geochronology by multicollector LA-ICP-MS of the Samaná Gneiss: a Middle Triassic syn-tectonic body in the Central Andes of Colombia, related to the latter stages of Pangea assembly. In South American Symposium on Isotope Geology, No. 6, Extended Abstracts. San Carlos de Bariloche.

 

Iriarte, S.; Arévalo, C.; Mpodozis, C.; Rivera, O. 1996. Hoja Carrera Pinto, Región de Atacama. Servicio Nacional de Geología y Minería, Mapas Geológicos 3, 1 mapa escala 1:100.000.

 

Santiago. Iriarte, S.; Arévalo, C.; Mpodozis, C. 1999. Hoja La Guardia, Región de Atacama. Servicio Nacional de Geología y Minería, Mapas Geológicos 13, 1 mapa escala 1:100.000. Santiago.

 

Jara, C.; Rabbia, O.; Valencia, V. 2009. Petrología y dataciones U-Pb del depósito tipo pórfido Cu Zaldívar, II Región de Antofagasta, Chile. In Congreso Geológico Chileno, No. 12, Actas 4: 4 p. Santiago.

 

Jenchen, U.; Rosenfeld, U. 2002. Continental Triassic in Argentina: response to tectonic activity. Journal of South American Earth Sciences 15: 461-479.

 

Jordan, T.E.; Isacks, B.L.; Allmendinger, R.W.; Brewer, J.A.; Ramos, V.A.; Ando, C.J. 1983. Andean tectonics related to geometry of subducted Nazca plate: Geological Society of America Bulletin 94: 341-361.

 

Kay, S.M.; Ramos, V.A.; Mpodozis, C.; Sruoga, P. 1989. Late Palaeozoic to Jurassic silicic magmatism at the Gondwana margin: Analogy to the middle Proterozoic in North America? Geology 17: 324-328.

 

Koukharsky, M.; Kleiman, L.; Etcheverría, M.; Quenardelle, S.; Bercowski, F. 2009. Upper Car-boniferous retroarc volcanism with submarine and subaerial facies at the western Gondwana margin of Argentina. Journal of South American Earth Sciences 27: 299-308.

 

Llambías, E.J.; Sato, A.M. 1990. El Batolito de Colangüil, cordillera frontal de Argentina: Estructura y marco tectónico. Revista Geológica de Chile 17 (1): 89-108. doi: 10.5027/andgeoV17n1-a04.

 

Llambías, E.J.; Sato, A.M. 1995. El Batolito de Colangüil: transición entre orogénesis y anorogénesis. Revista de la Asociación Geológica Argentina 50 (1-4): 111-131. Buenos Aires.

 

Llambías, E.J.; Kleiman, L.E.; Salvarredi, J.A. 1993. El Magmatismo Gondwánico. In Geología y Recursos Naturales de Mendoza (Ramos, V.; editor). Congreso Geológico Argentino, No. 12 y Congreso de Exploración de Hidrocarburos, No. 2. Relatorio: 53-64. Mendoza.

 

Llambías, E.J.; Quenardelle, S.; Montenegro, T. 2003. The Choiyoi Group from central Argentina: a subalkaline transitional to alkaline association in the craton adjacent to the active margin of the Gondwana continent. Journal of South American Earth Sciences 16: 243-257.

 

López-Gamundí, O. 2006. Permian plate margin volcanism and tuffs in adjacent basins of west Gondwana: Age constraints and common characteristics. Journal of South American Earth Sciences 22: 227-238.

 

Lucassen, F.; Franz, G.; Thirlwall, M.F.; Mezger, K. 1999. Crustal recycling of metamorphic basement, Late Paleozoic granitoids of northern Chile (~22°S): implications for the composition of the Andean crust. Journal of Petrology 40: 1527-1551.

 

Ludwig, K.R. 2001. SQUID 1.02, A User’s Manual. Berkeley Geochronology Center, Special Publication 2, 2455 Ridge Road, Berkeley, CA 94709, USA.

 

Ludwig, K.R. 2003. User’s Manual for Isoplot/Ex, Version 3.0, A Geochronological Toolkit for Microsoft Excel. Berkeley Geochronology Center, Special Publication 4, 2455 Ridge Road, Berkeley CA 94709, USA.

 

Maksaev, V.; Marinovic, N. 1980. Cuadrángulos Cerro de la Mica, Quillagua, Cerro Posada y Oficina Prosperidad, Región de Antofagasta. Instituto de Investigaciones Geológicas, Carta Geológica de Chile 45-48: 63 p.

 

Marinovic, N. 2007. Carta Oficina Domeyko, Región de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 105: 41 p., 1 mapa escala 1:100.000. Santiago.

 

Marinovic, N.; Smoje, I.; Maksaev, V.; Hervé, M.; Mpodozis, C. 1995. Hoja Aguas Blancas, Región de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 70: 142 p. 1 mapa escala 1:250.000. Santiago.

 

Marinovic, N.; García, M. 1999. Hoja Pampa Unión, Región de Antofagasta. Servicio Nacional de Geología y Minería, Mapas Geológicos 9, 1 mapa escala 1:100.000. Santiago.

 

Martin, M.W.; Clavero, J.; Mpodozis, C. 1999. Late Palaeozoic to Early Jurassic tectonic development of the high Andean Principal Cordillera, El Indio region, Chile (29º-30ºS). Journal of South American Earth Sciences 12: 33-49.

 

Masterman, G.L. 2003. Structural and geochemical evolution of the Rosario Copper-molybdenum porphyry deposit and related copper veins, Collahuasi District, Northern Chile. Doctorate Thesis (Unpublished), University of Tasmania: 253 p. Australia.

 

Mercado, M. 1980. Geología del área Pan de Azúcar, Región de Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 37: 30 p., 1 mapa escala 1:100.000. Santiago.

 

Mercado, M. 1982. Geología de la Hoja Laguna del Negro Francisco, Región de Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 56: 73 p., 1 mapa escala 1:100.000. Santiago.

 

Miskovic, A.; Schaltegger, U. 2009. Crustal growth along a non-collisional cratonic margin: A Lu-Hf isotopic survey of the Eastern Cordilleran granitoids of Perú. Earth and Planetary Science Letters 279 (3-4): 303-315.

 

Miskovic, A.; Spikings, R.A.; Chew, D.M.; Kosler, J.; Ulianov, A.; Schaltegger, U. 2009. Tectonomagmatic evolution of Western Amazonia: Geochemical characterization and zircon U-Pb geochronologic constraints from the Peruvian Eastern Cordilleran granitoids. Geological Society of America Bulletin 121 (9-10): 1298-1324.

 

Morata, D.; Aguirre, L.; Oyarzún, M.; Vergara, M. 2000. Crustal contribution in the genesis of the bimodal Triassic volcanism from the Coastal Range, central Chile. Revista Geológica de Chile 27 (1): 83-98. doi: 10.5027/andgeoV27n1-a06.

 

Moscoso, R.; Mpodozis, C.; Nasi, C.; Ribba, L.; Arévalo, C. 2010. Geología de la Hoja El Tránsito, Región de Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Preliminar 7: 17 p., 1 mapa escala 1:250.000. Santiago.

 

Mpodozis, C.; Cornejo, P. 1988. Hoja Pisco Elqui, IV Región de Coquimbo. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 68: 164 p., 1 mapa escala 1:250.000. Santiago.

 

Mpodozis, C.; Ramos, V. 1989. The Andes of Argentina and Chile. In Geology of the Andes and its relation to hydrocarbon and mineral resources (Ericksen, G.E.; Cañas Pinochet, M.T.; Reinemund, J.A.; editors). Circumpacific Council for Energy and Mineral Resources, Earth Science Series 11: 59-90.

 

Mpodozis, C.; Kay, S. 1990. Provincias magmáticas ácidas y evolución tectónica de Gondwana: Andes chilenos (28-31°S). Revista Geológica de Chile 17 (2): 153-180. doi: 10.5027/andgeoV17n2-a03.

 

Mpodozis, C.; Kay, S. 1992. Late Paleozoic to Triassic evolution of the pacific Gondwana margin: evidence from Chilean frontal Cordilleran batholiths. Geological Society of America Bulletin 104: 999-1014.

 

Mpodozis, C.; Allmendinger, R.W. 1993. Extensional tectonics, Cretaceous Andes, northern Chile (27°S). Geological Society of America Bulletin 105: 1462-1477.

 

Mpodozis, C.; Marinovic, N.; Smoje, I.; Cuitiño, L. 1993. Estudio Geológico-Estructural de la Cordillera de Domeyko entre Sierra Limón Verde y Sierra Mariposas, Región de Antofagasta. Servicio Nacional de Geología y Minería, Informe Registrado IR-93-04: 282 p. Santiago.

 

Mpodozis, C.; Iriarte, S.; Gardeweg, M.; Valenzuela, M. 2012. Carta Laguna del Negro Francisco, Región de Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 145: 30 p. 1 mapa escala 1:100.000. Santiago.

 

Munizaga, F.; Maksaev, V.; Fanning, C.M.; Giglio, S.; Yaxley, G.; Tassinari, C.C.G. 2008 Late Paleozoic-Early Triassic magmatism on the western margin of Gondwana: Collahuasi area, Northern Chile. Gondwana Research 13: 407-427.

 

Murillo, I.; Álvarez, J.; Arriagada, C.; Calderón, M.; Charrier, R. 2012. Geología estructural del Valle de El Tránsito con énfasis en las Milonitas El Portillo, III Región, Chile. In Congreso Geológico Chileno, No. 13, T2: 314-316. Antofagasta.

 

Naranjo, J.A.; Puig, A. 1984. Hojas Taltal y Chañaral, Regiones de Antofagasta y Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 62-63: 140 p., 1 mapa escala 1:250.000. Santiago.

 

Nasi, C.; Mpodozis, C.; Cornejo, P.; Moscoso, R.; Maksaev, V. 1985. El Batolito Elqui-Limarí (Paleozoico Superior-Triásico): características petrográficas, geoquímicas y significado tectónico. Revista Geológica de Chile 25-26: 77-11. doi: 10.5027/andgeoV12n2-3-a06.

 

Nasi, C.; Moscoso, R.; Maksaev, V. 1990. Hoja Guanta, Regiones de Atacama y Coquimbo. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 67: 140 p., 1 mapa escala 1:250.000. Santiago.

 

Niemeyer, H. 2013. Geología del Cerro Lila-Peine, Región de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 147.

 

Niemeyer, H.; Venegas, R.; Baeza, L.; Soto, H. 1985. Reconocimiento geológico del sector sur-occidental del Cuadrángulo Cerro Yocas, ubicado en la zona de falla Quebrada Blanca-Chuquicamata, Región de Antofagasta. In Congreso Geológico Chileno, No. 4, Actas 1: 629-653. Antofagasta.

 

Niemeyer, H.; Urzúa, F.; Rubinstein, C. 1997. Nuevos antecedentes estratigráficos y sedimentológicos de la Formación Zorritas, Devónico-Carbonífero de Sierra de Almeida, Región de Antofagasta, Chile. Revista Geológica de Chile 24 (1): 25-43. doi: 10.5027/andgeoV24n1-a02.

 

Noble, S.R.; Aspden, J.A.; Jemelita, R. 1997. Northern Andean crustal evolution: New U-Pb geochronological constraints from Ecuador. Geological Society of America Bulletin, 109: 789-798.

 

Oliveros, V.; Labbé, M.; Rossel, P.; Charrier, R.; Encinas, A. 2011. Late Jurassic paleogeographic evolution of the Andean back-arc basin: New constrains from the Lagunillas Formation, northern Chile (27°30’-28°30’S). Journal of South American Earth Sciences 37: 25-40.

 

Ordóñez, O.; Pimentel, M. 2002. Rb-Sr and Sm-Nd isotopic study of the Puqui complex, Colombian Andes. Journal of South American Earth Sciences 15: 173-182.

 

Padel, M.; Salazar, E.; Coloma, F. 2012. Arquitectura y evolución tectonoestratigráfica del Depocentro de San Félix, Triásico Medio a Superior: Resultados preliminares. In Congreso Geológico Chileno, No. 13, T5: 715-717. Antofagasta.

 

Pankhurst, R.; Millar, I.; Hervé, F. 1996. A Permo-Carboniferous U-Pb age for part of the Guanta unit of the Elqui-Limarí Batholith at Río del Tránsito, Northern Chile. Revista Geológica de Chile 23 (1): 35-42. doi: 10.5027/andgeoV23n1-a03.

 

Pankhurst, R.J.; Rapela, C.W.; Fanning, C.M.; Márquez, M. 2006. Gondwanide continental collision and the origin of Patagonia. Earth-Science Reviews 76: 235-257.

 

Parada, F. 2013. Geoquímica de las rocas ígneas del Carbonífero-Triásico de la Alta Cordillera, Región de Atacama, Chile. Memoria de Título (Inédito), Universidad de Chile, Departamento de Geología: 93 p. Santiago.

 

Parada, M.A.; Munizaga, F.; Kawashita, K. 1981. Edades Rb-Sr roca total del Batolito compuesto de los ríos Elqui-Limarí a la latitud 30°S. Revista Geológica de Chile 13-14: 87-93. doi: 10.5027/andgeoV8n2-3-a07.

 

Parada, M.A.; Levi, B.; Nystrom, J.O. 1991. Geochemistry of the Triassic to Jurassic plutonism of Central Chile (30 to 33°S); petrogenetic implications and a tectonic discussion. In Andean Magmatism and its Tectonic Setting (Harmon, R.S.; Rapela, C.W.; editors). Geological Society of America, Special Papers 265: 99-112. Boulder.

 

Parada, M.A.; López-Escobar, L.; Oliveros, V.; Fuentes, F.; Morata, D.; Calderón, M.; Aguirre, L.; Féraud, G.; Espinoza, F.; Moreno, H.; Figueroa, O.; Muñoz, J.; Troncoso, R.; Stern, C.R. 2007. Andean Magmatism. In The Geology of Chile (Moreno, T.; Gibbons, W.; editors). The Geological Society, Special Publication 4: 115-146. London.

 

Pineda, G.; Calderón, M. 2008. Geología del área Monte Patria-El Maqui, Región de Coquimbo. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 116: 44 p., 1 mapa escala 1:100.000. Santiago.

 

Ramírez, C.F.; Gardeweg, M. 1982. Geología de la Hoja Toconao, Región de Antofagasta. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 54: 117 p., 1 mapa escala 1:250.000. Santiago.

 

Ramos, V. 2009. Anatomy and global context of the Andes: Main geologic features and the Andean orogenic cycle. The Geological Society of America, Memoir 204: 31-65.

 

Ramos, V.A.; Kay, S.M. 1991. Triassic rifting and associated basalts in the Cuyo basin, central Argentina. In Andean magmatism and its tectonic setting (Harmon, R.S.; Rapela, C.W.; editors). Geological Society of America, Special Papers 265: 79-91. Boulder.

 

Ramos, V.; Folguera, A. 2009. Andean flat-slab subduction through time. Geological Society Special Publication 327: 31-54. London.

 

Restrepo, J.J.; Toussaint, J.F.; González, H.; Cordani, U.; Kawashita, K.; Linares, E.; Patricia, C. 1991. Precisiones geocronológicas sobre el occidente Colombiano. In Simposio sobre magmatismo andino y su marco tectónico, Manizales 1: 1-21.

 

Reutter, K.J. 1974. Entwicklung und Bauplan der chilenischen Hochkordillere im Bereich 29º südlicher Breite. Neues Jahrbuch für Geologie und Paläontologie 146: 153-178.

 

Ribba, L.; Mpodozis, C.; Hervé, F.; Nasi, C.; Moscoso, R. 1988. El basamento del valle del Tránsito, Cordillera de Vallenar: Eventos magmáticos y metamórficos y su relación con la evolución paleozoica de los Andes Chileno-Argentinos. Revista Geológica de Chile 15 (2): 129-149. doi: 10.5027/andgeoV15n2-a03.

 

Richards, J.P.; Noble, S.R.; Pringle, M.S. 1999. A revised late Eocene age for porphyry Cu magmatism in the Escondida area, northern Chile. Economic Geology 94: 1231-1247.

 

Rocha-Campos, A.C.; Basei, M.A.; Nutman, A.P.; Kleiman, L.; Varela, R.; Llambías, E.; Canile, F.M.; da Rosa, O. de C.R. 2011. 30 million years of Permian volcanism recorded in the Choiyoi igneous province (W Argentina) and their source for younger ash fall deposits in the Paraná Basin: SHRIMP U-Pb zircon geochronology evidence. Gondwana Research 19 (2): 509-523.

 

Salazar, E. 2012. Evolución tectonoestratigráfica post-Paleozoica de la Cordillera de Vallenar. Tesis de Magister en Ciencias, Mención Geología (Inédito), Universidad de Chile, Departamento de Geología: 126 p. Santiago.

 

Salazar, E.; Arriagada, C.; Mpodozis, M.; Martínez, F.; Peña, M.; Álvarez, J. 2009. Análisis estructural del Oroclino de Vallenar. Resultados preliminares. In Congreso Geológico Chileno, No. 12, Abstract: S9-026. Santiago.

 

Salazar, E.; Mpodozis, C.; Arriagada, C.; Coloma, F. 2012. Evolución tectonoestratigráfica post-paleozoica de la Cordillera de Vallenar. In Congreso Geológico Chileno, No. 13, T2: 259-261. Antofagasta.

 

Salazar, E.; Coloma, F.; Creixell, C. 2013. Geología del área El Tránsito-Lagunillas, Región de Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 149.

 

Sato, A.M.; Llambías, E.J. 1993. El Grupo Choiyoi, provincia de San Juan: equivalente del Batolito de Colangüil. In Congreso Geológico Argentino, No. 12, Actas 4: 156-165. Mendoza.

 

Sempere, T.; Carlier, G.; Soler, P.; Fornari, M.; Carlotto, V.; Jacay, J. 2002. Late Permian-Middle Jurassic lithospheric thinning in Perú and Bolivia, and its bearing on Andean-age tectonics. Tectonophysics 345 (1-4): 153-181.

 

Sepúlveda, P.; Naranjo, J.A. 1982. Geología de la Hoja Carrera Pinto, Región de Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile 53: 60 p. 1 mapa escala 1:100.000. Santiago.

 

Shibata, K.; Ishihara, S.; Ulriksen, C. 1984. Rb-Sr Ages and Initial 87Sr/86Sr ratios of Late Paleozoic Granitic Rocks from Northern Chile. Bulletin of the Geological Survey of Japan 35 (11): 537-545.

 

Sillitoe, R.H. 1977. Permo-Carboniferous, Upper Cretaceous, and Miocene porphyry copper-type mineralization in the Argentinean Andes. Economic Geology 72: 99-109.

 

Skarmeta, J.; Marinovic, N. 1981. Geología de la Hoja Quillagua, Región de Antofagasta. Instituto de Investigaciones Geológicas, Carta Geológica de Chile 51: 63 p. 1 mapa escala 1:250.000. Santiago.

 

Stipanicic, P.; Rodrigo, F.; Baulier, O.; Martínez, C. 1968. Las Formaciones Presenonianas en el denominado Macizo Norpatagónico y regiones adyacentes. Revista de la Asociación Geológica Argentina 23: 67-98.

 

Strazzere, D.A.; Gregori, D.A.; Dristas, J.A. 2006. Genetic evolution of Permo-Triassic volcaniclastic sequences at Uspallata, Mendoza Precordillera, Argentina. Gondwana Research 9: 485-499.

 

Suárez, M.; Bell, M. 1991. Sedimentos lacustres de probable edad triásica en el área de La Coipa, Región de Atacama, Chile. In Congreso Geológico Chileno, No. 6, Actas 1: 660-663. Viña del Mar.

 

Suárez, M.; Bell, M. 1992. Triassic rift-related sedimen-tary basins in northern Chile (24º-29ºS). Journal of South American Earth Sciences 6: 109-121.

 

Suárez, M.; Fanning, C.M.; Etchart, H.; De La Cruz, R. 2012. New Carboniferous to Toarcian U-Pb SHRIMP ages from Cordillera del Viento, Neuquén, Argentina. In Congreso Geologico Chileno, No. 13, Proceedings, CD-ROM. Antofagasta.

 

Tera, F.; Wasserburg, G. 1972. U-Th-Pb systematics in three Apollo 14 basalts and the problem of initial Pb in lunar rocks. Earth and Planetary Science Letters 14: 281-304.

 

Tomlinson, A.J.; Blanco, N. 2008. Geología de la franja El Abra-Chuquicamata, II Región (21°45ʼ-22°30ʼS). Servicio Nacional de Geología y Minería, Informe Registrado IR-08-35: 196 p. Santiago.

 

Tomlinson, A.; Cornejo, P.; Mpodozis, C. 1999. Hoja Potrerillos, Región de Atacama. Servicio Nacional de Geología y Minería, Mapas Geológicos 14, 1 mapa escala 1:100.000. Santiago.

 

Tomlinson, A.J.; Blanco, N.; Maksaev, V.; Dilles, J.H.; Grunder, A.L.; Ladino, M. 2001. Geología de la Precordillera Andina de Quebrada Blanca-Chuquicamata, Regiones I y II (20°30’-22°30’S). Servicio Nacional de Geología y Minería, Informe Registrado IR-01-20: 444 p. Santiago.

 

Uliana, M.A.; Biddle, K.T.; Cerdan, J. 1989. Mesozoic extension and the formation of Argentine sedimentary basins. In Extensional Tectonics and Stratigraphy of the North Atlantic Margins (Tankard, A.J.; Balkwill, H.R.; editors). American Association of Petroleum Geologists, Memoir 46: 599-614.

 

Urzúa, F. 2009. Geology, geochronology and structural evolution of La Escondida copper district, northern Chile. Ph.D. Thesis (Unpublished), University of Tasmania: 486 p. Hobart, Australia.

 

Vásquez, P.; Glodny, J.; Franz, G.; Frei, D.; Romer, R.L. 2011. Early Mesozoic Plutonism of the Cordillera de la Costa (34°-37°S), Chile: Constraints on the Onset of the Andean Orogeny. The Journal of Geology 119 (2): 159-184.

 

Vaughan, A.P.M.; Pankhurst, R.J. 2008. Tectonic overview of the West Gondwana margin. Gondwana Research 13: 150-162.

 

Venegas, C.; Cervetto, M.; Astudillo, N.; Espinoza, F. 2013. Carta Sierra Vaquillas Altas, Regiones de Anto-fagasta y Atacama. Servicio Nacional de Geología y Minería, Carta Geológica de Chile, Serie Geología Básica 159.

 

Vergara, H.; Thomas, A. 1984. Hoja Collacagua, Región de Tarapacá, Servicio Nacional de Geología y Minería, Carta Geológica de Chile 59: 79 p., 1 mapa escala 1:250.000. Santiago.

 

Vergara, M.; López-Escobar, L.; Cancino, A.; Levi, B. 1991. The Pichidangui Formation; some geochemical characteristics and tectonic implications of the Triassic marine volcanism in central Chile (31º55’ to 32º20’S). In Andean Magmatism and its Tectonic Setting (Harmon, R.S.; Rapela, C.W.; editors). Geological Society of America, Special Paper 265: 93-98.

 

Vinasco, C.J.; Cordani, U.; González, G.; Weber, M.; Peláez, C. 2006. Geochronological, isotopic, and geochemical data from Permo-Triassic granitic gneisses and granitoids of the Colombian Central Andes. Journal of South American Earth Sciences 21: 355-371.

 

Williams, I.S. 1998. U-Th-Pb geochronology by ion microprobe. In Applications of Microanalytical Techniques to Understanding Mineralizing Processes (McKibben, M.A.; Shanks III, W.C.; Ridley, W.I.; editors). Reviews in Economic Geology 7: 1-35.

 

Witt, W.K.; Hagemann, S.G.; Villanes, C.; Zeng, Q. 2013. New geochronological results and structural evolution of the Pataz gold mining district: Implications for the timing and origin of the batholith-hosted veins. Ore Geology Reviews 50: 143-170.

 

 

Appendix 1

Tera-Wasserburg concordia plots of SHRIMP U-Pb zircon ages for samples dated during this study (ordered according Tables 3 and 4). Ellipses and error in ages in the inset are at 1σ; decay constant errors included.

Appendix 2

Shimp zircon U-Pb analytical data.