Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

Journal of Structural Geology 140 (2020) 104133

Contents lists available at ScienceDirect

Journal of Structural Geology


journal homepage: http://www.elsevier.com/locate/jsg

Deformation history of the Puna plateau, Central Andes of


northwestern Argentina
Susana Henríquez a, *, Peter G. DeCelles a, Bárbara Carrapa a, Amanda N. Hughes a,
George H. Davis a, Patricia Alvarado b
a
Department of Geosciences, University of Arizona, 1040, E. 4th Street, Tucson, AZ, USA
b
CIGEOBIO (CONICET-UNSJ)-Departamento de Geofisica y Astronomía, FCEFN-Universidad Nacional de San Juan, Argentina

A R T I C L E I N F O A B S T R A C T

Keywords: Two tectonic shortening events created the first-order structural characteristics of the Puna plateau: one in the
Puna plateau Paleozoic and a second during the Andean Orogeny in the Cenozoic. To constrain the structural characteristics
Andes and timing of Andean deformation in the Puna plateau, we focus on differentiating these two shortening events
Thrust belt
and provide cooling ages (apatite fission track and apatite (U–Th)/He) to determine the amount of Andean
Structure
exhumation on the hanging-wall of major Cenozoic faults. Two contrasting expressions of shortening are
Thermochronology
Exhumation documented. In Ordovician strata, strain requires four stages: folding, flattening through cleavage formation,
low-angle faulting and local distributed shear. In the Mesozoic and Cenozoic strata, shortening is expressed as
reverse faults, fault-propagation folds and dipping panels. Dissimilarities reflect different P-T conditions and
amount of strain, as structures in the Ordovician rocks require higher temperatures and strain than Andean
structures in the Mesozoic and Cenozoic strata. Furthermore, local Cretaceous AFT cooling ages in pervasively
cleaved Ordovician rocks reveal these rocks did not reach temperatures compatible with cleavage formation
during the Andean event. Lastly, our data show a middle Eocene to Oligocene exhumation history that supports
an overall continuous development of the Andean thrust belt in the Puna plateau since the early Cenozoic.

1. Introduction study of the Andean fold-thrust belt requires consideration of both An­
dean deformation and structures inherited from older shortening events.
Retroarc thrust belts record the dynamics of orogens along conver­ We use the term “Andean deformation” to refer to the Cenozoic-age
gent plate margins (e.g., Willett, 1992; Tavani et al., 2015; Yonkee and deformation, even though the Andean deformation started during the
Weil, 2015). Analyzing the kinematic evolution of structures is funda­ Late Cretaceous (e.g., Cobbold and Rosello, 2003; Mpodozis et al., 2005;
mental to determining the evolution of thrust belt. This includes Arriagada et al., 2006).
studying the strain and mechanisms of deformation operating at Structurally, one of the main issues arises from different observations
different times and scales. In the hinterland of the central Andes, the and interpretations of the nature and timing of structures in the Ordo­
thrust belt involves the Altiplano-Puna plateau, a >3600 m high region vician strata of the Puna. Ordovician strata are ca. 10 km thick north of
(e.g., Isacks, 1988) with crustal thicknesses of up to 70 km (e.g., Beck 21 ◦ S (e.g., Müller et al., 2002) and at least ca. 7 km in the Puna plateau
et al., 2015; Rivadeneyra-Vera et al., 2019). North of 21 ◦ S, several (e.g., Bahlburg, 1991), where they constitute the only exposed Paleozoic
studies in the Altiplano and Eastern Cordillera have documented rocks. Upright folding and cleavage in the Ordovician strata have been
Cenozoic shortening involving Paleozoic to Cenozoic rocks (Fig. 1; e.g., related to a Carboniferous-Permian event (Kley and Reinhardt, 1994;
McQuarrie, 2002; Müller et al., 2002; Ege et al., 2007; Anderson et al., Jacobshagen et al., 2002) but also to high-strained domains spatially
2017). South of 21 ◦ S, shortening across the Puna plateau also involves correlated with Cenozoic-age structures (McQuarrie and Davis, 2002).
Paleozoic to Cenozoic rocks; however, different expressions of short­ This confusion raises questions about strain progression and mecha­
ening have been attributed to Paleozoic and Cenozoic shortening events nisms of cleavage development and overall deformation in the Puna
(e.g., Bahlburg, 1991; Kley et al., 1997; Müller et al., 2002). Thus, the plateau.

* Corresponding author.
E-mail address: susanahg@email.arizona.edu (S. Henríquez).

https://doi.org/10.1016/j.jsg.2020.104133
Received 11 October 2019; Received in revised form 26 June 2020; Accepted 30 June 2020
Available online 6 August 2020
0191-8141/© 2020 Elsevier Ltd. All rights reserved.
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

into four longitudinal tectonomorphic zones (Fig. 1) which, from west to


east are the (1) Western Cordillera, (2) Altiplano-Puna plateau, (3)
Eastern Cordillera, and (4) the Subandean and Santa Bárbara ranges.
The Western Cordillera is composed of the Neogene to Recent vol­
canic arc with high volcanic peaks that reach ca. 6000 m a.s.l. The
Altiplano-Puna plateau is a high elevation (>3600 m a.s.l.) and rela­
tively flat zone in the hinterland. The Altiplano is an internally drained
basin at ca. 3600 m elevation that records sedimentation since the
Paleocene (e.g., Horton, 2005). The surface is dominated by modern
salars and Neogene to Quaternary sediment. Its southern part was
deformed during late Oligocene to Miocene time (e.g., Elger et al.,
2005). South of 21 ◦ S, the Puna plateau has a more rugged topography
formed by a series of ranges that separate small internally drained ba­
sins; base level is at ca. 4200 m (Isacks, 1988; Allmendinger et al., 1997;
Coutand et al., 2001; Strecker et al., 2007). Surface geology in the Puna
plateau is dominated by Ordovician sedimentary and igneous rocks,
with Cenozoic sedimentary rocks and Miocene volcanic rocks along its
western edge (Fig. 2). The Eastern Cordillera is the highest part of the
thrust belt, reaching ca. 6500 m elevation. This part of the thrust belt
also presents differences north and south of 21 ◦ S. North of 21 ◦ S, it is an
externally drained region that accommodated significant shortening in a
doubly-vergent thin-skinned fold-thrust belt developed between late
Eocene and Oligocene time (e.g., McQuarrie, 2002; Müller et al., 2002;
Barnes et al., 2008; Anderson et al., 2018). It exposes mainly Cambrian
to Silurian sedimentary rocks locally covered by Mesozoic and Cenozoic
strata. South of 21 ◦ S, the Eastern Cordillera or Cordillera Oriental
transitions to a bivergent Miocene thrust belt (Kley and Monaldi, 2002;
Fig. 1. Digital Elevation Model of the Central Andes depicting the main tec­ Kley et al., 2005). In this part of the thrust belt, thick-skinned defor­
tonomorphic regions. CC: Coastal Cordillera, CD: Cordillera de Domeyko, WC: mation exposes Proterozoic to Mesozoic rocks that separate Cenozoic
Western Cordillera, EC: Eastern Cordillera, IA: Interandean Ranges, SA: Sub­ depocenters. Finally, the Subandean zone is a thin-skinned fold-thrust
andean Ranges, SB: Santa Bárbara Ranges. belt composed of Silurian through Cenozoic sedimentary rocks, mainly
north of the Argentina-Bolivia border. It was formed during the last 10
The purpose of this contribution is to analyze details of deformation Ma (e.g., McQuarrie, 2002; Echavarria et al., 2003; Barnes et al., 2008;
in the northern Puna plateau in order to examine kinematics, including Hernández and Echavarria, 2009; Anderson et al., 2018; Calle et al.,
strain continuity. In order to understand structural mechanisms oper­ 2018). The Subandean ranges transition southward into the Santa
ating at regional scale, it is necessary to first understand the dominant Bárbara ranges, where upper Cambrian to Devonian, Cretaceous and
mechanism at outcrop scale within individual thrust sheets and struc­ Cenozoic rocks are uplifted and deformed by east-dipping reverse faults
tures. We hypothesize that the strain and exhumation history during (Kley and Monaldi, 2002).
Andean deformation in the northern Puna plateau is not able to fully
account for the strain progression within Ordovician rocks. To test this 2.2. Regional stratigraphic assemblages
hypothesis, we characterize deformation at outcrop scale and document
the distribution of strain across Ordovician lithotectonic assemblages, The stratigraphic record in the northern Puna plateau includes three
investigate the origin of cleavage development, describe Cenozoic main periods of sedimentation during the Ordovician, Cretaceous and
deformation in Andean synorogenic basins, and document cooling and Cenozoic (Fig. 3). These are described in the following sections.
exhumation in the Andean thrust belt. By integrating strain observations
with the exhumation history in basement uplifts, we introduce kine­ 2.2.1. Ordovician
matic constraints for deformation in the region. In this paper, we show The total stratigraphic thickness of Ordovician rocks in the Puna is >
that progressive deformation of Ordovician strata involved at least four 7000 m (Bahlburg, 1991; Moya, 2015). The Ordovician includes three
phases of deformation, two pre-Andean phases, which produced upright unconformity-bound units that are referred to as the Complejo de Pla­
folding and flattening through cleavage formation; and two later phases taforma de la Puna (CPP), a volcaniclastic sequence, and the Complejo
that during the Cenozoic (?) produced low-angle faulting and local do­ Turbíditico de la Puna (CTP) (Figs. 2 and 3). The Complejo de Plata­
mains of distributed shear. Moreover, this study shows that shortening, forma de la Puna (CPP) is mainly composed of the Cobres Group, a 2000
exhumation, and erosion occurred in the northern Puna plateau mainly m thick quartzite and shale sequence in the Del Cobre Range. The vol­
between the middle Eocene (ca. 45-40 Ma) and the Oligocene (ca. 25 caniclastic sequence includes the 3800 m thick Aguada de la Perdiz and
Ma) due to the forward propagation of the thrust belt. Chiquero Formations (Buatois et al., 2009). The third Ordovician unit,
the Complejo Turbidítico de la Puna, is a ca. 3500 m thick succession of
2. Regional geology sandy to conglomeratic turbidites that accumulated in a basin in the
western Puna (Bahlburg, 1991). The Ordovician deposits accumulated
2.1. Regional tectonic setting in a north-south trending foreland basin with increasing accommodation
to the west and a forebulge to the east where shelf to deltaic sediments
The modern Andean retroarc thrust belt in the Central Andes records were deposited (Bahlburg, 1991; Egenhoff, 2007).
the growth of the Andes by shortening and thickening due to the east­ The Ordovician units have been affected by low-grade meta­
ward propagation of the orogenic belt during the Cenozoic (e.g., Isacks, morphism (Jacobshagen et al., 2002). During Ordovician time, two
1988; Allmendinger et al., 1997; McQuarrie, 2002; Kley, 1996; Oncken magmatic arcs developed: the Complejo Igneo y Sedimentario Cordón de
et al., 2006; DeCelles et al., 2011, 2015; Carrapa and DeCelles, 2015). Lila in the forearc (CISL, Furongian-Tremadocian; e.g., Niemeyer,
The central Andean retroarc region between 18◦ and 25 ◦ S is divided 1989), and the Faja Eruptiva de la Puna Oriental in what is now the Puna

2
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 2. Geologic map of the northern Puna plateau.

region. According to Bahlburg et al. (2016), this magmatic event pro­ stromatolitic limestone (Marquillas et al., 2005). The Santa Bárbara
duced a N–S belt that extends for ca. 400 km and includes a middle Subgroup has been interpreted as the results of a second episode of
Ordovician phase (480-460 Ma) followed by a second and main event thermal subsidence in the Salta Rift (Salfity and Marquillas, 1994;
during the late Ordovician (453-444 Ma). Marquillas et al., 2005; Starck, 2011). However, DeCelles et al. (2011)
documented an extremely mature sequence of compound paleosols in
2.2.2. Cretaceous the Maiz Gordo and Lumbrera Formations, which represents a zone of
Up to 6000 m of sedimentary rocks, referred to as the Salta Group, stratigraphic condensation over a period of ca. 10–15 Ma throughout all
accumulated in the intracratonic Salta Rift. The Salta Group consists of of northwestern Argentina and southern Bolivia (DeCelles and Horton,
the Pirgua, Balbuena, and Santa Bárbara Subgroups (Salfity and Mar­ 2003; Horton, 2005). The paleosol disconformity is interpreted to be the
quillas, 1994), which represent formations controlled by normal faulting result of forebulge migration through this part of the Andes, with fore­
and post-rifting thermal subsidence (e.g., Welsink et al., 1995). The deep and wedge-top deposits above the paleosols supporting the east­
Pirgua Subgroup consists of up to 4 km of clastic and volcanic rocks (e.g., ward migration of a foreland basin system during the Cenozoic (DeCelles
Marquillas et al., 2005) of Neocomian to lower Maastrichtian age. The and Horton, 2003; DeCelles et al., 2011).
Pirgua Subgroup was deposited during the syn-rift phase of the Salta Rift Above the Lumbrera Formation, the stratigraphic record can be
(e.g., Grier et al., 1991; Starck, 2011). The Balbuena Subgroup consists divided into the Quebrada de Los Colorados Formation, Río Grande
of 400–500 m of Maastrichtian to lower Paleocene limestone and clastic Formation and the Jujuy Subgroup. The Quebrada de Los Colorados
rocks (e.g., Marquillas et al., 2005; Sempere et al., 1997). The Balbuena Formation is an upward-coarsening succession of sandstone and
Subgroup consists of ca. 250 m of limestone and calcareous sandstone of conglomeratic sandstone interbedded with red siltstone (DeCelles et al.,
the Lecho and Yacoraite Formations. These formations are of Maas­ 2011; Montero-López et al., 2018). This unit crops out from the Susques
trichtian age and provide a structural marker across the region. area to the east, with ca. 860 m measured in the Tres Cruces area (Boll
and Hernández, 1986). A maximum depositional age of 37.6 ± 1.2 Ma
2.2.3. Cenozoic was determined using U–Pb analysis of detrital zircons (Carrapa et al.,
The Lower Cenozoic stratigraphy includes the Mealla, Maiz Gordo, 2012). The Río Grande Formation is composed of 3000–4200 m of eolian
and Lumbrera Formations of the Santa Bárbara Subgroup. The Mealla and fluvial sandstone and conglomerate (Starck and Vergani, 1996; Siks
Formation consists of 100–150 m of sandstone and siltstone; the Maiz and Horton, 2011). Siks and Horton (2011) reported 40Ar/39Ar analyses
Gordo Formation consists of 200–250 m of coarse-to fine-grained of five ashfall tuffs in the Cianzo area that yielded weighted mean ages
sandstone; and the Lumbrera Formation is 400–450 m of sandstone and between 16.34 ± 0.6 Ma and 9.69 ± 1 Ma. Carrapa et al. (2012) reported
mudstone, with intercalations of dark green to grey claystone and a maximum depositional age of 21.4 ± 0.7 Ma for the equivalent

3
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 3. Simplified stratigraphic columns. Modified from Bahlburg (1991), Boll and Hernández (1986), Moya (2015), Kley and Monaldi (2002), Siks and Horton
(2011) and DeCelles et al. (2011).

Angastaco Formation and considered that this formation was deposited Cianzo areas. The Pisungo Formation consists of a succession of poorly
from ca. 21 to ca. 9 Ma. These early to mid-Miocene units were deposited consolidated conglomerate that accumulated in proximal alluvial fans
in the foredeep and wedge-top parts of the foreland basin system on the flanks of growing structures (Siks and Horton, 2011). The Sijes
(DeCelles et al., 2011; Carrapa et al., 2012). The Jujuy Group includes Formations is a 200 m thick pyroclastic unit (Turner, 1960) covered by
the Pisungo and Sijes Formations, which crop out in the Tres Cruces and dacites dated at 12 ± 2 Ma by K–Ar on whole rock (Coira, 1979).

4
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Sedimentation continued until ca. 0.8 Ma in the Tres Cruces area (e.g., Salinas Grandes and Tres Cruces depocenters have been studied
Streit et al., 2017). combining surface observations with seismic reflection profiles and
interpreted to be the results of late Eocene to Quaternary shortening
2.3. Regional deformation events (Coutand, 1999; 2001; Steinmetz and Galli, 2015). Kinematic analyses
of the Cenozoic faults in the northern Puna (Marrett et al., 1994; Cla­
Three main periods of deformation have characterized the evolution douhos et al., 1994; Coutand et al., 2001) have shown shortening di­
of the northern Puna plateau: a series of Paleozoic events, the Early rections trending on average 110◦ (Coutand et al., 2001) and 120 ± 20◦
Cretaceous rifting event, and the Cenozoic Andean shortening event. (Cladouhos et al., 1994) with subvertical stretching. Basin burial and
uplift histories indicate a late Eocene age for the onset of shortening
2.3.1. Pre-Andean deformation between 23 and 25◦ S (Coutand et al., 2001), whereas structural data at
Deformation in the Ordovician section is characterized by N–S to 22 ◦ S indicate that shortening continued during Miocene to Pliocene
NNE-SSW trending open to upright folds, which are marked by slaty times (Cladouhos et al., 1994). In southern Bolivia, Andean deformation
cleavage (e.g., Cladouhos et al., 1994; Müller et al., 2002). Similar styles in the Eastern Cordillera corresponds to development of the
of folding have been documented across the Eastern Cordillera in Bolivia ENE-trending bivergent fold-thrust belt that formed mainly between ca.
and Puna regions (e.g., Bahlburg, 1991; McQuarrie, 2002; Pearson et al., 45 and 25 Ma (e.g., McQuarrie, 2002; Müller et al., 2002; Ege et al.,
2013). This folding took place under low-grade metamorphic condi­ 2007; Anderson et al., 2017). Shortening continued until ca. 9 Ma and
tions. In the northern Puna plateau, the folded and cleaved Ordovician was followed by minor strike-slip faulting (e.g., Müller et al., 2002).
rocks are overlain unconformably by Cretaceous strata.
Information about thermal conditions and timing of folding and 3. Methods
cleaving of the Ordovician rocks is provided by studies of vitrinite
reflectance, infrared spectroscopy of organic matter, illite crystallinity, This work combines field methods and three analytical methods: (1)
and fluid inclusion analyses performed on sedimentary rocks collected Geologic mapping to document structures, contacts and geologic re­
north of the study area in the Eastern Cordillera of southern Bolivia (at lationships as a basis for kinematic analysis; (2) U–Pb geochronology on
ca. 21 ◦ S; Kley and Reinhardt, 1994; Jacobshagen et al., 2002). Results detrital zircons to constrain maximum depositional ages of two Cenozoic
show that: 1) K–Ar ages on synkinematic illite crystals in the Ordovician depocenters and the age of deformation; (3) apatite fission track and
rocks point to a late Carboniferous age for cleavage development apatite (U–Th)/He thermochronology to constrain the cooling history of
(320-290 Ma); 2) fluid inclusions in quartz veins cutting Ordovician Ordovician rocks. Because the depths and styles of the deformation are
strata indicate temperatures of ca. 240–270 ◦ C; and 3) vitrinite reflec­ intimately related to thermal conditions, cooling ages can inform on the
tance data from Ordovician to Carboniferous rocks demonstrates a depth of deformation driven exhumation.
decrease in sediment maturity from the Devonian to Carboniferous rocks
and from west to east in Devonian rocks (Kley and Reinhardt, 1994). 3.1. Field methods
These results suggest that a late Devonian-early Carboniferous heating
event was responsible for the folding and the development of slaty Structural data were collected along geologic transects and during
cleavage (Kley and Reinhardt, 1994). Müller et al. (2002) argued that mapping. The mapping was focused on providing constraint for defor­
the contact between the deformed Ordovician section and the mation along the selected transects; thus, it was concentrated on
un-cleaved Cretaceous units is, in most cases, an angular unconformity regional distribution of lithologic units, cross-cutting relationships and
but without much angular discordance and that the subvertical slaty unconformities. We identified and described dipping panels within the
cleavage is observed in both folded and flat-lying Ordovician rocks. Ordovician strata and collected samples in the hanging-wall of Cenozoic
Therefore, Müller et al. (2002) argued that pre-Cretaceous folding of the thrust faults for low-temperature thermochronologic analysis. For each
Ordovician strata in the Eastern Cordillera of southern Bolivia seems to dipping panel, we described the lithology and bed thickness; searched
be moderate or localized, even though the Ordovician rocks had to have for veins and indicators of bedding orientation; measured bedding and
been buried to 3–7 km depth during peak metamorphism and cleavage cleavage; and characterized fold geometries and fault zones. Work in
formation during late Devonian-late Carboniferous time. McQuarrie and Cenozoic deposits was focused on characterizing deformation by
Davis (2002) observed both highly strained and unstrained domains in measuring bedding, describing geometries and collecting samples for
Paleozoic rocks of Bolivia (at ca. 18 ◦ S). The highly strained domains are U–Pb geochronology. Additionally, we measured and described strati­
mainly in cores of folds and fault zones, whereas the unstrained domains graphic sections and collected samples to provide age constraints in
occur along sections of strata that are structural flats (McQuarrie and depocenters with previously unknown stratigraphy.
Davis, 2002). Additionally, structural conformity between Jurassic and
Devonian rocks in central Bolivia indicates that the folding, faulting and 3.2. U–Pb geochronology
cleavage observed in the Paleozoic strata are results of progressive
deformation in the Andean thrust belt and not Late Paleozoic In order to constrain the age of Cenozoic deposits we used U–Pb
(McQuarrie and Davis, 2002). geochronology on detrital zircons from sedimentary rocks and on zir­
During the Early Cretaceous, the Salta Rift and its three main cons from volcanic rocks. We collected eight sandstone samples for
depocenters accommodated significant regional extension (e.g., Starck, detrital zircon analyses and one volcanic rock for igneous zircon ana­
2011). The N–S trending Tres Cruces depocenter (Fig. 2) in the northern lyses. Descriptions of mineral separation techniques and data reduction
Puna is strongly deformed by Andean deformation, yet preserved methods are in the supplementary material (sections S1 and S2.1). U–Pb
normal faults were described by Monaldi et al. (2008). analytical data can be found in Fig. S1 and Table S1 (Supplementary
material). To determine maximum depositional ages of sedimentary
2.3.2. Andean deformation samples we applied the TuffZirc routine in Isoplot (Ludwig, 2008). This
Exposures of rocks and structures in the Paso de Jama, El Toro, technique yields reliable ages when the sample has either a group of at
Salinas Grandes, and Tres Cruces depocenters and mountain ranges in least five analyses of igneous crystals, or a reasonable fraction of the
between (Fig. 2) provide insight into the nature of Andean deformation. analyses (at least 30–40%) are cogenetic with an eruption age; TuffZirc
These ranges are bounded by high-angle reverse faults that uplift only works well on crystals that are unaffected by Pb loss. For samples
Ordovician strata over Mesozoic and Cenozoic strata. At the surface, the that do not meet this criterion, we used the mean age of the youngest
high-angle reverse faults dip more than 60◦ , strike N–S to NNE-SSW, and cluster of three or more grain ages overlapping at 2σ (Dickinson and
do not show a preferred vergence (e.g., Coutand et al., 2001). The Gehrels, 2009). The clusters or peaks were determined using the DZ Age

5
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Pick Program from the University of Arizona LaserChron Center (www. 3.3.2. Apatite (U–Th)/He and thermal modeling
geo.arizona.edu/alc). Results are shown in Table 1. The apatite (U–Th)/He system is based on α-particles (4He) emitted
during the radioactive decay of the elements 238U, 235U, and 232Th
(Farley et al., 1996; Ehlers and Farley, 2003). The 4He particles are
3.3. Low-temperature thermochronology retained within the crystal at low temperatures, whereas at high tem­
peratures 4He particles diffuse out of crystal. For apatites, the diffusive
The main objective of low-temperature thermochronology in this loss of 4He happens between 40 and 80 ◦ C in the partial retention zone
study was to determine the thermal history of rocks along hanging-walls (PRZ) (e.g., Farley, 2000; Reiners and Brandon, 2006). Procedures for
of structures within the thrust belt (e.g., Ege et al., 2007; Mora et al., grain selection and analysis are described in the supplementary material
2015). We employed apatite fission track (AFT) and apatite (U–Th)/He (sections S1 and S2.3).
(AHe) thermochronology. These two thermochronometers constrain the Our sampling strategy targeted Ordovician and Lower Cretaceous
timing of cooling as rocks transit through the upper few kilometers of intrusive rocks exposed across the Puna plateau. For each sample, we
crust (e.g., Reiners and Brandon, 2006). We were especially interested in analyzed 3 to 5 single aliquots. We looked at the age reproducibility
understanding the cooling history of the Ordovician rocks, which have among grains of a given sample and investigated correlations between
experienced pressures and temperatures necessary for low-grade meta­ eU (ppm) and Rs (μm) and eU (ppm) and corrected age (Ma) to evaluate
morphism. We collected a suite of 11 samples, including three for AHe radiation damage (e.g., Shuster et al., 2006). We used inverse modeling
analysis (SH16-3, SH16-19 and SH16-40), six for AFT analysis (SH16-5, to determine the time-temperature histories of samples using the soft­
SH16-9, SH16-10, SH16-11, SH16-33 and SH16-47), and two for both ware HeFTy (Ketcham, 2005; section S2.4). Results of the apatite
(SH16-1 and SH16-4) (Tables 2 and 3; Fig. 2). We targeted Ordovician (U–Th)/He analyzes are shown in Table 3 and Figs. S3 and S4, and are
sandstones/quartzites and intrusive bodies in the hanging-walls of discussed in the results and interpretation sections from old to young in
major reverse faults (Fig. 2). Apatite grains were separated following the the context of each range. The parameters and data considered for the
description in the supplementary material (section S1). thermal models are presented in Table 4 following protocols proposed
by Flowers et al. (2015).
3.3.1. Apatite fission track thermochronology
AFT thermochronology uses crystallographic damage in apatite 4. Stratigraphy and U–Pb geochronologic constraints on Puna
grains produced by natural fission of 238U. These fission tracks are Cenozoic depocenters
preserved at temperatures below 80–120 ◦ C and rapidly healed at
temperatures higher than 120 ◦ C (e.g., Green et al., 1989). Between 80 The youngest rock units in the northern Puna affected by Andean
and 120 ◦ C, fission tracks begin to anneal or shorten defining a tem­ thrusting and folding are ca. 6.5 Ma (Letcher, 2007). Thus, mapping and
perature range known as the partial annealing zone (PAZ) (e.g., Green dating of Cenozoic rocks provide insight into the character and timing of
et al., 1989). Depending on the time-temperature paths that the sample late Cenozoic deformation (Figs. 4 and 5). Two Cenozoic depocenters in
experienced after deposition or crystallization, the original fission tracks the western part of the Puna were studied: The Paso de Jama and El Toro
can be fully annealed by burial, fully retained by fast and monotonic depocenters (Fig. 6). We also mapped sedimentary rocks in the Susques
exhumation through the PAZ, or partially annealed if the sample resided area and collected samples to establish depositional ages.
within the PAZ. An ~2.5 km-thick, upward-coarsening sequence of Cenozoic strata
For each sample, we counted fission tracks in 20 grains or all the crops out in the Paso de Jama area (Fig. 4). The base of the section is not
available grains in cases of low grain availability or quality. Descriptions exposed but is inferred to rest unconformably upon strongly deformed
of sample preparation and analysis are provided in the supplementary Ordovician turbiditic strata in the Olaroz Range (Fig. 4). The lower
material (sections S1 and S2.2). To assess whether the grain ages belong ~780 m of the section consists of laminated siltstone transitioning up­
to a homogenous population, we used the X2 test where a P(X2) >5% ward into fine-grained sandstone with current ripples, trough cross-
indicates that individual data are consistent with a common ratio of stratification, and lenticular sandstone bodies with erosional bases
spontaneous and induced counts (e.g., Galbraith, 1981; Green, 1981). that are interpreted as fluvial channel deposits (Fig. 6). The sample
The results are shown in Table 2.

Table 1
Compiled U–Pb geochronologic analyses.
Sample Location Type Lithology No. of Youngest Methoda Age Lat. Long. Elevation
crystals population

[Ma ± 1σ] [Ma ± [◦ S] [◦ W] [m a.s.l.]


1σ]

SH16-13 Susques Range Basement Tuff 18 NA Weighted Mean 6.8 ± − 23.001 − 66.358 3995
0.11
SH17-10 Susques Range Detrital Tuffaceous 24 9.12 ± 0.1 Weighted Mean NA − 23.185 − 66.509 4121
sandstone
b
SH17-12 Susques Range Detrital Sandstone 302 61.3 ± 1.7 Weighted NA − 23.197 − 66.481 4242
Meanb
SH17-19 Paso de Jama Detrital Sandstone 178 22.15 ± 0.33 TuffZirc Age NA − 23.230 − 66.926 4249
area
SH17-21 El Toro area Detrital Sandstone 287 425 ± 34 Weighted Mean NA − 23.093 − 66.758 4153
c
SH16-44 El Toro area Detrital Gravel 288 36.1 ± 3.6 Weighted NA − 23.126 − 66.769 4290
Meanc
PJ405DF Paso de Jama Detrital Sandstone 262 270 ± 48 Weighted Mean NA
area
PJ793DZ Paso de Jama Detrital Sandstone 290 35.99 ± 0.87 TuffZirc Age NA − 23.242 − 66.955 4298
area
a
Method used to calculate youngest population.
b
MSWD = 13, Ma ± 2σ, Probability of fit = 0 and n = 4.
c
MSWD = 29, Ma ± 2σ, Probability of fit = 0 and n = 3.

6
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

(PJ405DZ) did not produce a statistically significant Cenozoic age

Note: ρs—spontaneous track density, Ns—number of spontaneous tracks, ρi—induced track density; Ni—number of induced tracks; ρd—dosimeter track density; Nd—number of dosimeter tracks; P(χ)2—χ2 probability; m
Elevation (m a.
(Table S1). A sandstone sample (PJ793DZ) collected at 793 m above the
base of the measured stratigraphic section (Fig. 6) produced a youngest
U–Pb detrital zircon age population (maximum depositional age) with a

3286

3984

4161

4523

4295

4137

3623

4054
mean age of 36.99 ± 0.87 Ma (PJ793DZ, 17 grains). We correlate the
s.l.)

strata in the lower part of the section with the Quebrada de Los Colo­
rados Formation, which has yielded similar ages elsewhere in the Puna
− 65.956

− 66.491

− 67.094

− 66.863

− 66.838

− 66.757

− 66.245

− 66.683
Lat. (◦ S)

plateau and Eastern Cordillera of northwestern Argentina (DeCelles


et al., 2011; Carrapa et al., 2012). Resting conformably above the lower
interval is an ~200 m thick unit of fine-grained sandstone with frosted
− 23.380

− 23.422

− 23.295

− 23.125

− 23.116

− 23.088

− 23.406

− 23.109
grains and large-scale trough cross-stratification (several meter-thick
Long.

cross-sets), which we interpret as an eolian deposit. We tentatively


(◦ W)

correlate this unit with the lower member of the Angastaco Formation
(Carrapa et al., 2012), which includes lower Miocene eolianites (Starck
Age Central (Ma

and Anzótegui, 2001; DeCelles et al., 2011; Kortyna et al., 2019). Above
the eolian interval, a sequence of massive coarse-grained sandstone beds
40.5 ± 3.5

26.3 ± 2.9

69.6 ± 5.8

44.4 ± 5.6
25 ± 2.4

67 ± 6.6

43 ± 4.4

29 ± 3.5
was sampled (SH17-19) for U–Pb detrital geochronology (ca. 50 m in the
± 1σ)

local stratigraphic succession, ca. 1150 m in the composite stratigraphic


section shown in Fig. 6). The maximum depositional age obtained is
22.15 ± 0.33 Ma (SH17-19, 28 grains), which is consistent with previ­
Age Pooled (Ma

a.s.l.: meters above sea level. Error is 1σ, calculated using the zeta calibration method (Hurford and Green., 1983) with zeta of 357.6 ± 7.7 for apatite.
ously reported age of the Angastaco Formation (Carrapa et al., 2012).
40.1 ± 2.4

26.3 ± 2.9

69.6 ± 5.8

40.2 ± 3.8

27.2 ± 2.5

The Cenozoic succession in the Paso de Jama area ends with ~1000 m of
25 ± 2.4

67 ± 6.6

43 ± 5
± 1σ)

very coarse-grained fluvio-alluvial sandstone and conglomerate.


The El Toro depocenter contains an upward-coarsening Cenozoic
sequence (Fig. 6). At the base of the succession, a fine-grained sandstone
4604

4604

4604

4604

4604

4604

4604

4604

intercalated with variegated paleosols is correlated with the Lumbrera


Nd

Formation. A sample (ca. 2 m) above the basal angular unconformity


Dosimeter

1387667

1411361

1140157

1269196

1257349

1277052

1340278

1182279

(SH17-21) on top of Ordovician rocks produced U–Pb zircon ages older


than 250 Ma. Above this unit, a sequence of pink to red, medium-to
ρd

coarse-grained sandstone transitions upward into coarse-grained sand­


stone and conglomerate. A sample from this interval (SH16-44, ~480 m)
P(χ)2

produced youngest detrital zircon U–Pb ages of ca. 36 Ma from three


(%)

0.1
65

90

85

82

10

39

grains (Fig. 6; Table 1); however, the individual ages do not overlap at
5

2σ, such that a maximum depositional age cannot be certified (Dick­


1238

2555

1053

1269
743

492

675

475

inson and Gehrels, 2009). Nevertheless, we tentatively interpret this


Ni

coarse-grained interval as equivalent to the Quebrada de Los Colorados


4253220

4480360

1583270

1421472

1822958

2355152

6057459

1419454

Formation of late Eocene age (Carrapa and DeCelles, 2008). Above this
Induced

unit, a sequence of very coarse-grained sandstone and partially consol­


ρi

idated conglomerate caps the sequence. We tentatively surmise that this


unit is equivalent to either the Rio Grande or Angastaco Formations (e.
125

407

146

210

186

144
96

97
Ns

g., Boll and Hernández, 1986; Carrapa et al., 2012; Siks and Horton,
2011).
429444.663583472

713701.243820431

204567.868100927

421818.826994438

567142.535393444

416009.778271011

687371.239223813

289867.524640791

In the western side of the Susques area, DeCelles et al. (2011)


documented the Lumbrera Formation as an upward-coarsening succes­
Spontaneous

sion that starts with ~100 m of massive red siltstone and paleosols and
records increasing amounts of fluvial sandstone and conglomerate
up-section. The section described by DeCelles et al. (2011) is ~400 m
ρs

thick and was measured ca. 10 km south of the studied area in an


along-strike equivalent position. In the studied part of the Susques area
the sequence of red sandstone intercalated with conglomerate reaches
Crystals
No. of

~600 m thick. This unit is capped by a conglomeratic unit (~300 m


12

18

20

13

20

20

12

13

thick) that we tentatively correlate with the Rio Grande Formation.


Above this unit, a volcanosedimentary unit composed of sandstone and
Sandstone

Sandstone

Sandstone

Sandstone

Sandstone
Granite

Granite

Granite

ignimbrite (~400 m thick) of probable Miocene age marks the top of the
Type

stratigraphic record. A sample from the uppermost ignimbrite (SH16-13,


Fig. 2, Table 1) and a sample from a tuffaceous sandstone from the
middle part of the volcanosedimentary succession (SH17-10, Figs. 5 and
Ordovician/Del Cobre
Cretaceous/Del Cobre

Ordovician/Susques

6, Table 1) yielded weighted mean U–Pb zircon ages of 6.8 ± 0.11 Ma


Ordovician/Lina

Ordovician/Lina

Ordovician/Lina

Ordovician/Lina

Ordovician/Lina

and 9.12 ± 0.1 Ma, respectively (Table 1). This unit is age-equivalent to
Apatite fission track data.
Unit/Locality

the Sijes Formation in the Tres Cruces area and the Pisungo Formation
farther east in the Cianzo area (Siks and Horton, 2011).
Range

Range

Range

Range

Range

Range

Range

Range
SH16-1-

SH16-4

SH16-5

SH16-9
Sample
Table 2

SH16-

SH16-

SH16-

SH16-
10

11

33

47
2

7
S. Henríquez et al.
Table 3
Apatite (U–Th)/He data (Continued).
Sample Unit/Sample type Aliquot name 4He 4He 4He U (ng) U (ng ± Th Th (ng Sm Sm (ng Th/U Raw age Raw age Ft Ft Ft
(pmol) (pmol ± (pmol ± 1σ) (ng) ± 1σ) (ng) ± 1σ) (atomic) (Ma) (Ma ± 238U 235U 232Th
1σ) 1σ) 1σ)

SH16-1 Granite, basement 16B892_SH_SH16_1_1 0.0226 0.0003 1.3109 0.1731 0.0025 0.0492 0.0007 1.8630 0.0283 0.292 22.432 0.414 0.720 0.682 0.682
sample 16B893_SH_SH16_1_2 0.0091 0.0001 1.5579 0.1071 0.0015 0.0318 0.0005 1.1863 0.0177 0.304 14.506 0.293 0.666 0.622 0.622
16B894_SH_SH16_1_3 0.0499 0.0006 1.1374 0.2388 0.0034 0.0743 0.0011 2.6266 0.0389 0.319 35.609 0.611 0.755 0.721 0.721
16B895_SH_SH16_1_4 0.0092 0.0001 1.3652 0.1183 0.0017 0.0376 0.0005 1.1935 0.0178 0.326 13.326 0.249 0.703 0.664 0.664
16B896_SH_SH16_1_5 0.0577 0.0007 1.1885 0.2011 0.0029 0.6321 0.0091 1.7301 0.0259 3.225 30.269 0.470 0.702 0.662 0.662
SH16-3 Granite, basement 16B898_SH_SH16_3_1 0.3289 0.0036 1.1079 2.2462 0.0323 0.0420 0.0018 4.8098 0.0731 0.019 26.937 0.476 0.851 0.829 0.829
sample 16B899_SH_SH16_3_2 0.1004 0.0012 1.1504 0.7904 0.0114 0.0199 0.0003 2.5555 0.0379 0.026 23.316 0.419 0.814 0.787 0.787
16B900_SH_SH16_3_3 0.0220 0.0005 2.0731 0.2108 0.0031 0.0084 0.0001 0.8360 0.0125 0.041 19.107 0.477 0.732 0.696 0.696
16B901_SH_SH16_3_4 0.0734 0.0014 1.9653 0.2710 0.0039 0.0619 0.0009 1.0324 0.0157 0.234 47.278 1.122 0.752 0.718 0.718
16B902_SH_SH16_3_5 0.0353 0.0007 2.0636 0.2604 0.0038 0.0056 0.0001 0.6425 0.0095 0.022 24.880 0.619 0.729 0.692 0.692
SH16-4 Granite, Basement 16B903_SH_SH16_4_1 0.0330 0.0007 2.0307 0.1591 0.0023 0.0300 0.0004 0.4677 0.0071 0.193 36.683 0.891 0.717 0.679 0.679
sample 16B904_SH_SH16_4_2 0.4306 0.0084 1.9539 2.1691 0.0313 0.1384 0.0020 3.4292 0.0536 0.065 36.127 0.864 0.839 0.816 0.816
16B905_SH_SH16_4_3 0.1103 0.0022 1.9511 0.4854 0.0070 0.1344 0.0019 2.5211 0.0387 0.284 39.228 0.922 0.827 0.802 0.802
16B906_SH_SH16_4_4 0.1460 0.0029 1.9620 0.4263 0.0061 0.4527 0.0066 1.8416 0.0278 1.089 50.367 1.148 0.828 0.804 0.804
16B907_SH_SH16_4_5 0.0389 0.0008 2.0103 0.1822 0.0027 0.0441 0.0007 1.0425 0.0158 0.248 37.144 0.896 0.729 0.692 0.692
SH16- Ordovician 16B910_SH_SH16_19_2 0.0613 0.0012 1.9609 0.4565 0.0066 0.3277 0.0047 0.9292 0.0139 0.736 21.212 0.489 0.760 0.727 0.727
19 Diorite, basement 16B911_SH_SH16_19_3 0.0093 0.0002 2.0836 0.0946 0.0014 0.0610 0.0009 0.5902 0.0089 0.662 15.746 0.382 0.698 0.658 0.658
sample 16B912_SH_SH16_19_4 0.0580 0.0011 1.9811 0.3354 0.0048 0.6242 0.0090 0.9589 0.0143 1.909 22.188 0.499 0.717 0.678 0.678
16B918_SH16_19_6 0.0155 0.0007 4.2152 0.0888 0.0013 0.2964 0.0042 0.8697 0.0127 3.424 17.965 0.779 0.664 0.620 0.620
SH16- Granite, Basement 16B914_SH_SH16_40_1 0.0014 0.0001 6.9897 0.0071 0.0001 0.0581 0.0008 1.2106 0.0180 8.447 11.877 0.839 0.762 0.729 0.729
40 sample 16B915_SH_SH16_40_2 0.0036 0.0001 2.7449 0.0088 0.0001 0.0742 0.0011 5.2137 0.0779 8.694 20.185 0.586 0.808 0.781 0.781
16B920_SH16_40_4 0.0073 0.0003 4.2225 0.0682 0.0010 0.0246 0.0004 1.3127 0.0195 0.370 17.932 0.790 0.662 0.618 0.618
8

Ft Ft Rs Corrected 1s err ppm eU ppm U ppm Th ppm Sm nmol mass ap W.A. Std. Long. Lat. (◦ S) Elev. Aliquot name Sample
232Th 147Sm (um) age (Ma) anal. w/Sm (morph) (morph) (morph) 4He/g (μg) age ± σ Dev. (◦ W) (m a.s.
(Ma) (Ca) (morph) [Ma] [Ma] l.)

0.682 0.910 50.689 31.188 0.576 73.241 61.864 17.587 665.741 8.090 2.640 26.5 ± 12.7 − 23.380 − 65.956 3286 16B892_SH_SH16_1_1 SH16-1
0.622 0.891 41.699 21.821 0.442 68.777 65.053 19.300 720.454 5.514 1.747 0.2 16B893_SH_SH16_1_2
0.721 0.921 58.456 47.235 0.812 71.086 57.435 17.881 631.807 12.011 3.777 16B894_SH_SH16_1_3
0.664 0.904 47.575 18.986 0.356 76.703 58.849 18.685 593.502 4.598 1.731 16B895_SH_SH16_1_4
0.662 0.904 47.292 44.135 0.687 174.368 75.691 237.932 651.208 21.720 2.048 16B896_SH_SH16_1_5
0.829 0.952 98.833 31.681 0.561 124.870 92.492 1.731 198.053 13.543 18.248 31.05 ± 15 − 23.423 − 66.195 3547 16B898_SH_SH16_3_1 SH16-3
0.787 0.941 78.417 28.665 0.516 84.301 68.762 1.735 222.323 8.737 9.573 0.3 16B899_SH_SH16_3_2
0.696 0.914 53.225 26.104 0.653 67.705 55.001 2.184 218.111 5.749 3.201 16B900_SH_SH16_3_3
0.718 0.920 57.856 62.953 1.498 64.318 49.365 11.281 188.071 13.364 4.514 16B901_SH_SH16_3_4
0.692 0.913 52.474 34.155 0.851 84.356 90.520 1.941 223.321 12.253 3.139 16B902_SH_SH16_3_5
0.679 0.909 50.159 51.236 1.248 63.810 48.751 9.184 143.328 10.128 2.637 49.6 ± 6.6 − 23.422 − 66.491 3984 16B903_SH_SH16_4_1 SH16-4

Journal of Structural Geology 140 (2020) 104133


0.816 0.949 91.277 43.099 1.031 115.546 112.991 7.211 178.626 22.430 19.193 0.5 16B904_SH_SH16_4_2
0.802 0.945 84.664 47.512 1.119 54.067 44.896 12.430 233.201 10.202 9.778 16B905_SH_SH16_4_3
0.804 0.945 85.381 61.148 1.396 62.848 43.513 46.211 187.990 14.899 8.606 16B906_SH_SH16_4_4
0.692 0.913 52.463 51.051 1.234 48.173 41.765 10.103 238.964 8.920 4.098 16B907_SH_SH16_4_5
0.727 0.923 59.880 28.100 0.649 265.707 120.111 86.222 244.490 16.121 2.023 26.76 ± 12 − 22.590 − 65.859 3800 16B910_SH_SH16_19_2 SH16-
0.658 0.902 46.685 22.709 0.552 61.242 53.898 34.765 336.452 5.312 1.822 0.35 16B911_SH_SH16_19_3 19
0.678 0.909 49.999 31.481 0.710 224.491 146.213 272.115 418.011 25.293 2.165 16B912_SH_SH16_19_4
0.620 0.890 41.399 27.836 1.208 136.115 53.876 179.829 527.624 9.417 1.192 16B918_SH16_19_6
0.729 0.924 60.375 15.796 1.116 6.933 1.981 16.308 339.768 0.401 3.800 23.1 ± 4.8 − 23.274 − 65.787 3536 16B914_SH_SH16_40_1 SH16-
0.781 0.939 75.946 24.661 0.715 7.113 1.217 10.311 724.607 0.494 7.116 0.5 16B915_SH_SH16_40_2 40
0.618 0.890 41.211 27.056 1.194 50.774 33.372 12.032 642.745 3.579 1.577 16B920_SH16_40_4

Note: FT—alpha–ejection correction (see methods section and references therein); 1s err anal.—one–sigma analytical precision; eU—equivalent uranium concentration (U + 0.235 x Th); m a.s.l. - meters above sea level.
W.A.: Weighted Average. Aliquot name in cursive was not considered for age.
S. Henríquez et al.
Table 4
Constraints for HeFTy models.
Sample location Thermochronologic data Additional geological information Model parameters

Sample Long. Lat. (◦ S) Elev. AHe Aliquots Modeled AFT Data Modeled Crystallization intrusive Total Paths Acceptable Good
(◦ W) (m a.s. Paths Paths
Grain AHe Error Error N of Age Model Age T ( C)

Reference
l.)
Age (Ma) (%) grains (Ma) Age (Ma)
(Ma) (Ma)

SH16-1 − 23.380 − 65.956 3286 16B892_SH_SH16_1_1 22.4 5 22 11 25 25 145–90 350–400 147 ± 10 to 96 ± 5 Ma, 10,97,543 25,613 1,000
16B893_SH_SH16_1_2 14.5 3 21 Halpern and Latorre
16B894_SH_SH16_1_3 35.6 NC NC (1973)
16B895_SH_SH16_1_4 13.3 3 23
16B896_SH_SH16_1_5 30.3 NC NC
SH16-3 − 23.423 − 66.195 3547 16B898_SH_SH16_3_1 26.9 5 19 – – – 500–460 350–400 478.40 ± 3.46 Ma, UPb, – 18,933 1,227 1,000
16B899_SH_SH16_3_2 23.3 5 21 Insel et al. (2012) and 483
16B900_SH_SH16_3_3 19.1 4 21 ± 3 Ma, UPb, –Hauser et al.
16B901_SH_SH16_3_4 47.3 NC NC (2011)
16B902_SH_SH16_3_5 24.9 5 20
9

SH16-4 − 23.422 − 66.491 3984 16B903_SH_SH16_4_1 36.7 9 25 18 40.1 40.1 500–450 350–400 476 ± 1 Ma, Lork and 2,73,78,220 30,835 5
16B904_SH_SH16_4_2 36.1 NC NC Bahlburg, 1993), 453.3 ±
16B905_SH_SH16_4_3 39.2 10 25 9.3 Ma, UPb, –Bahlburg
16B906_SH_SH16_4_4 50.4 15 30 et al. (2016)
16B907_SH_SH16_4_5 37.1 10 27
SH16- − 22.590 − 65.859 3800 16B910_SH_SH16_19_2 21.2 4 19 – – – 500–450 350–400 447.1 ± 6 Ma, Bahlburg 20,241 1,577 1,000
19 16B911_SH_SH16_19_3 15.7 3 19 et al. (2016)
16B912_SH_SH16_19_4 22.2 5 23
16B918_SH16_19_6 18.0 4 22
SH16- − 23.274 − 65.787 3536 16B914_SH_SH16_40_1 11.9 4 34 – – – 145–90 350–400 150.4 ± 0.9 Ma, UPb, – 22,118 2,966 1,000
40 16B915_SH_SH16_40_2 20.2 4 20 Insel et al. (2012)
16B920_SH16_40_4 17.9 4 22

Model specific parameters: Search Method used was Monte Carlo. The present condition (0 Ma) for all the samples was considered 20 ◦ C. We used the Nodal, GOF Product for the weighted mean path funstion. The merit
value for "good" and "accepatable" fit are 0.5 and 0.05 respectively.

Journal of Structural Geology 140 (2020) 104133


Model prameters for (U–Th)/He system: AHe ages are the uncorrected ages (Table 3).The calibration model used for apatite (U–Th)/He data is RDAAM from Flowers et al. (2009). The alpha ejection corection model used
was from Ketcham et al., (2011). The error used for finding paths is reported as a percentage of the uncorrected age. Radious, U, Th and Sm concentrations required for the model correspond to ppm U (morph), ppm Th
(morph) and ppm (Sm (morph) values from Table 3.
Model prameters for AFT system: The Zetha value is 357.6 ± 7.7. The annealing model is from Ketcham et al. (2007). No lenght data was obtained from the samples.
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 4. Map and composite cross-section of the Olaroz and Lina ranges. Stereographic projections of the bedding and cleavage across the cross-section.

5. Structures in the northern Puna plateau perspective, we are able to document systematic variations in how these
structures are distributed spatially across the northern Puna plateau and
5.1. Overview of contrasting structural styles thereby identify the critical observations that permit differentiation of
these deformational events. In the following section we lay out some key
The main structures in the Ordovician assemblage are mesoscale examples that document these two contrasting expressions of shortening
upright folds, generally close to tight, with most limbs dipping from 45◦ in maps, cross-sections, and photographs.
to 80◦ , and flattened by widely distributed cleavage (Figs. 7 and 8).
Distinct areas characterized by symmetric folding and systematic
asymmetric folding were documented, with local subregions exhibiting 5.2. Nature of deformation
east and west-vergent folding. These structures contrast profoundly with
structures in the Mesozoic and Cenozoic rocks that are marked by the 5.2.1. Structures within the Ordovician rocks
presence of thrust faults, steeply dipping reverse faults, local asymmetric The Ordovician assemblage is generally characterized by interca­
macroscopic folds and extensive homoclinal dip panels indicative of lated weak layers (shales to very fined-grained sandstones) and strong
folding at larger spatial scales (Figs. 4 and 5). Folding and cleavage in layers (medium to coarse-grained quartz-rich sandstones) that vary in
the Ordovician assemblage of the northern Puna plateau have been thickness from a few centimeters up to a few tens of meters. The con­
generally described in previous studies (e.g., Coira and Zappettini, 2008; trasting strength of the two main compositional layers correlates with
Bahlburg, 1991; Mon and Salfity, 1995). However, by taking a regional variations in structural style. Folds in the Ordovician assemblage trend
15 ± 10◦ (stereographic projections, Figs. 4 and 5) and have concentric

10
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 5. Map and composite cross-section of the Susques and Del Cobre Ranges.

(Fig. 7c and f and 8d) and chevron geometries (Figs. 7b and 8f). Bed located in weak layers to penetrative across folds; the significance of
thicknesses of strong layers are generally preserved across folds. Some these variations will be further explored in section 5.2.1. Within rocks
folds display thickening in their cores accommodated by smaller-scale that are both folded and cleaved, highly strained domains are charac­
folds and layer-parallel slip in weak lithologies (Fig. 7c). Fold sizes terized by tight to isoclinal folds, steeply dipping fold limbs, small scale
and geometries vary systematically based on the mechanical layering folds (8e), penetrative cleavage and cleavage orientations subparallel to
and the amount of strain. Fold wavelengths range from about 150 m to bedding (Fig. 8c). The cleavage and scale of the folds is at least partially
2.5 km. In areas such as the Del Cobre Range where weak lithologies are controlled by the amount of strain and rheologic strength of the strata.
separated by thick stiff intervals (quartzitic sandstone several meters Other associated features are locally abundant quartz veins (Fig. 8b),
thick), deformation in the Ordovician section includes second-order kink bands (Fig. 7a) and low-angle thrust faults (Fig. 7d).
folds with wavelengths of a few tens of meters due to slip on internal The structure of the Ordovician rocks exposed in the Olaroz Range
detachments in areas in which weak lithologies predominate (Fig. 8a exhibits notable systematic local variations. On the western side of the
and d). However, in other portion of the Del Cobre Range, the Ordovi­ range, axial surfaces and cleavage dip preferentially eastward (Fig. 4). In
cian section is minimally folded or cleaved (due to its mechanical the middle of the range, folds are more symmetrical; in this area,
competence), resulting in the formation of large, open, kilometer-scale conglomerate pebbles in the center of a syncline exhibit sub-horizontal
folds (Fig. 5). open cracks that indicate vertical extension (Fig. 7e and g). In the syn­
Cleavage distribution and spacing varies from being preferentially cline along the eastern side of the range, folds are tighter to isoclinal

11
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 6. Schematic stratigraphic sections of the Paso de Jama and El Toro basins and relative probability plots showing the results of the detrital U–Pb geochronology.

with cleavage subparallel to the bedding. Axial planes of folds are west two east-vergent reverse faults produce regional lineaments in the area.
dipping but display an eastward variation from high-angle to low-angle Reverse faults in the Del Cobre Range cannot be observed directly cut­
in a fanning arrangement (Fig. 7h). ting Cenozoic strata; however, Cretaceous rocks are preserved in the
footwall of the west-vergent reverse fault (Coira and Zappettini, 2008),
5.2.2. Mesozoic and Cenozoic rocks suggesting that these faults are post-Cretaceous.
Steeply dipping reverse faults that bound the modern ranges are the Folds involving the Lumbrera Formation were observed in the El
main structures that characterize the Mesozoic to Cenozoic assemblage Toro and Susques areas. Two small-scale east-vergent kink anticlines at
(Fig. 2). Reverse faults trend NS to NNE, are subparallel to the Ordovi­ the base of the tilted domain in the El Toro area and an asymmetrical,
cian strata, and dip >50◦ . Folds and faults in the Mesozoic and Cenozoic partially covered, east-vergent anticline with a short hinge zone in the
strata trend 10◦ –25◦ (Figs. 4 and 5) and produce dip domains and large Paso de Jama area indicate east-vergent deformation. In the Susques
asymmetric folds. Backlimbs map out as broad homoclines where ca. area, two tight and overturned east-vergent anticlines within a west-
2.5 km of Eocene to Miocene strata dip between 30◦ and 80◦ . Schematic dipping panel indicate that the east vergent deformation likely pre­
cross-sections in Figs. 4 and 5 represent the scale difference in geometry ceded west-vergent deformation, although it is possible that these events
and deformation styles observed within the Mesozoic to Cenozoic as were contemporaneous.
compared to Ordovician assemblage. Folds are asymmetrical and show Differences in fold geometries provide insight into the deformation
both east- and west-vergence in accordance with the vergence of adja­ age. Folds involving Eocene to early Miocene strata are open to over­
cent reverse faults. Westward dipping domains in the Paso de Jama and turned, whereas folds involving late Miocene and younger deposits are
El Toro areas are clear expressions of Cenozoic faulting (Fig. 4). gentle to open, suggesting that most shortening was accommodated
In the Paso de Jama area, the contact between Ordovician and prior to the late Miocene. Examples of these fold relationships are
Cenozoic rocks is covered but is inferred to be a west-dipping reverse observed in a gentle syncline in the eastern side of Paso de Jama and in a
fault that uplifts the Ordovician rocks over the Cenozoic strata. A west- north-plunging fold in the western part of the Susques area. There, the
dipping reverse fault and an onlapping unconformity over Ordovician Ordovician basement (exposed to the south) is in direct contact with late
strata form the western and eastern boundaries, respectively, of Ceno­ Miocene ignimbrites, indicating either paleotopography or the erosion
zoic strata in the El Toro area. In the Susques area, a west-vergent of Eocene to early-mid Miocene strata prior to the later folding event.
reverse fault uplifts Ordovician rocks and imposes a west-vergent fold-
thrust system to the west. In the Del Cobre Range, a west-vergent and

12
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 7. Photos documenting strain development in Olaroz and Lina ranges. a) Kink bands b) Chevron folds c) Flexure slip fold: Anticline showing thickening in the
hinge region. Quarzitic sandstone beds preserve original thickness, whereas intercalated finer-grained strata show variable thickness. This can be seen in the eastern
side of the picture where a syncline includes steep dipping panels at the base that flatten to sub-horizontal beds toward the top of the outcrop. d) Low angle thrust
faults e) Conglomerate with quarzite clast affected by vertical extension (see g in this figure) f) Folds with east-dipping axial surfaces. g) Clast with open sub-
horizontal fractures due to vertical extension h) High strain domain: folds with west axial surfaces that steepen from east to west. This shear zone aligns with a
west-dipping Cenozoic fault to the south.

13
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 8. Photos documenting strain development in the Del Cobre Range. a) Anticline in the core of a syncline (Photo d). Concentric fold detachment within the
section. Other structures in the picture are tension gashes, small-scale folds and thrust faults that accommodate shortening in the core of the folds and axial-planar
cleavage which provide evidence of fold flattening. b) Quartz veins. c) Cleavage subparallel to the bedding. d) Relatively large-scale syncline. Parasitic folds in the
core are described in a). The quarzitic beds control these bigger folds, while smaller rheological changes control smaller scale structures. e) Small-scale folds are
indicating east-verging shear parallel to the bedding. f) Chevron folds in the core of a syncline (ca. 700 m wide). The wavelength is ca. 200 m, suggesting a shallow
detachment. Unconformity at the top with Cenozoic strata (undifferentiated).

14
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 9. Strain Progression. Proposed strain stages for deformation in the Ordovician basement.

5.3. Interpretation of kinematics and strain however, cleavage varies from weak (spaced cleavage) to slaty (pene­
trative cleavage). The tectonic flattening that produces cleavage is
5.3.1. Folding and cleavage formation in Ordovician rocks accommodated by pressure dissolution (e.g., Wood, 1974; McClay,
Strain within the Ordovician section is heterogeneous and parti­ 1977; Davis et al., 2011). The presence of quartz, which could have been
tioned, with different structural elements distributed within the two dissolved and locally reprecipitated in veins, further supports pressure
main compositional and rheologic rock types. Folding is the main dissolution. Flattening via cleavage formation is likely the result of the
mechanism accommodating strain within the Puna ranges, and flexural natural progression of flexural-slip folds after they hit a tightening limit,
slip is the dominant mechanism by which the folding was accomplished. though formation during a slightly later but distinct event cannot be
The intercalation of rheologically contrasting rocks in Ordovician strata ruled out.
is such that within the folded domains, strong layers largely maintain Although two (Paleozoic and Cenozoic) shortening events are
their thickness, while bedding-plane slip between stiff layers and documented in the region, the absence of intersecting cleavage planes
distributed shear and flexural flow within incompetent layers leads to indicates one event of cleavage formation. If cleavage in the Ordovician
folds that are roughly parallel. Incompetent shales therefore absorbed rocks was formed in the same tectonic event across the Altiplano and
more of the strain than the more competent sandstone/quartzite Puna plateau, then the synkinematic illites dated in southern Bolivia
lithologies. (Jacobshagen et al., 2002) suggest that recrystallization and pressure
Fold geometries (especially symmetric and tight to isoclinal folds), dissolution in the northern Puna plateau is Paleozoic in age.
variable scales of folding, and observations of intraformational de­
tachments at outcrop scale (Fig. 8 a, d) suggest that detachment folding 5.3.2. Interpretation of the major reverse faults
is operating at the mesoscale. We speculate that weak layers within the Reverse faults in the study area are large-scale structures that extend
Ordovician strata acted as intraformational detachments at depth due to for over 100 km along strike. The lateral continuity and area uplifted
the short wavelengths of the structures relative to the overall known above regional stratigraphic elevation in the hanging-walls of these
thickness of the stratigraphic section. Although fault-propagation faults suggest that they are deep-seated structures. Fault throw is at least
folding might be operating locally, the lack of systematic overturned 5–6 km based on fault dip, the thicknesses of local Cenozoic strati­
limbs and faulted domains in the deeper stratigraphic levels make the graphic sections that were overridden, and cooling ages presented in the
role of fault-propagation folding difficult to assess. following section. Fault-related folding associated with these structures
Folds were subsequently flattened during cleavage formation; was also accommodated by flexural slip but within more spatially

15
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Fig. 10. Summary of the AFT and AHe ages for the northern Puna at 23–24 ◦ S.

extensive uniform dip domains than in the Ordovician assemblage, in one way in which Andean shortening might have increased strain in the
this case involving unmetamorphosed and otherwise internally un­ Ordovician rocks. Thus, we propose that strain progression evolved in
strained stratigraphic packages up to ca. 4 km thick. The long, gently the Ordovician assemblage in four main stages depicted in Fig. 9: 1)
dipping backlimbs and steep, narrow front limbs of these structures main stage of upright folding; 2) flattening stage accompanied by
point toward fault-propagation folding as the dominant structural style. cleavage formation, vertical extension and possibly rotations; 3) low-
Dip domains are an expression of fault slip along footwall ramps that angle thrusting, and 4) local domains of distributed shear.
formed during fault-bend or fault-propagation folding.
6. Low-temperature thermochronology

5.4. History of deformation 6.1. Apatite fission track thermochronology results

Structural relationships allow for determination of the relative Although AFT was attempted on all samples, low apatite yield
timing of events. Structures that only affect the Ordovician rocks are the limited the results to eight cooling ages. We sampled Ordovician sedi­
first expressions of deformation in the northern Puna plateau. Because mentary rocks, Ordovician intrusives, and a Cretaceous intrusive from
cleavage in Ordovician rocks are not systematically rotated around fold six ranges across the Puna plateau (Fig. 2). The results are shown in
hinges, we propose that cleavage formation either post-dated an initial Table 2 and Figs. S5 and S6. Seven of the eight samples passed the χ2 test,
folding event or, more likely, accompanied the development of the folds. which indicates a single population. All samples of the Ordovician
We believe that initial buckling transitioned into flexural-slip folding sedimentary unit show ages younger than Ordovician. The results are
that in latter stages was accompanied by flattening and the development described below from older to younger in the context of each mountain
of axial planar cleavage. Layer-parallel strain due to further shortening range.
continued to form and rotate cleavage planes, flattening the folds. Low- In the Olaroz Range (Fig. 4), two samples from the Ordovician
angle thrust faults that crosscut fold limbs affected by cleavage imply a sedimentary succession yielded Cretaceous cooling ages. SH16-9 and
later stage of horizontal compression (Fig. 7d). SH16-10 yielded pooled ages of 67 ± 6.6 Ma and 69.5 ± 5.8 Ma
Deformation in the northern Puna plateau ended with Cenozoic respectively. To the east, two samples from the Lina Range show Eocene
reverse and thrust faulting and folding clearly observed in the Mesozoic cooling ages; pooled ages from samples SH16-11 and SH16-47 are 40.2
and Cenozoic rocks; however, the expression of Cenozoic deformation ± 3.8 Ma and 43 ± 5 Ma, respectively. One sample from an Ordovician
within the Ordovician assemblage is conspicuous. To better constrain intrusive exposed in the Susques Range (SH16-4) yielded a pooled age of
the deformation history and explore possible effects of Andean defor­ 40.1 ± 2.4 Ma. This sample was multi-dated using (U–Th)/He thermo­
mation in the Ordovician section, we combined cleavage orientations chronometry (see below). To the east, two samples from Ordovician
with observations of Cenozoic deformation. In the Olaroz Range, two sedimentary units exposed along the Del Cobre Range produced Oligo­
cleavage domains are observed: a west-dipping domain on the eastern cene cooling ages. Sample SH16-33 collected from an Ordovician granite
side and an east-dipping domain on the western side. Cleavage orien­ within the central thrust sheet of the Del Cobre Range produced a pooled
tation in the hanging-wall mirrors the Cenozoic faults below. One age of 27.2 ± 2.5 Ma, whereas a granitic sample from the Tusaquillas
possible explanation for this is that because the Ordovician rocks were Range (Tusaquillas Intrusive, Fig. 2) yielded an age of 25 ± 2.4 Ma.
already folded, cleaved and unable to accommodate more flexural slip, Sample SH16-5 from the Ordovician sedimentary succession exposed
slip along two bounding Cenozoic reverse faults further strained the west of the Paso de Jama area cooled through the AFT closure temper­
Ordovician rocks through distributed shear producing the systematic ature at 26.3 ± 2.9 Ma.
rotation of cleavage (Fig. 9 part 4). These structures are not the first-
order expression of Cenozoic deformation; however, they demonstrate

16
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

6.2. Apatite (U–Th)/He thermochronology results thermal model for this sample (HeFTy, Table 4, Fig. S3) suggests cooling
from temperatures of at least 80 ◦ C and possibly higher than 120 ◦ C at
We performed (U–Th)/He analyses on five samples (22 single grains; ca. 45-35 Ma. Only five ‘good’ paths were found. Although these paths
SH16-1, SH16-3, SH16-4, SH16-19 and SH16-40). AHe dates from the do not represent a unique solution, all of them suggest that rapid cooling
Susques Range (SH16-4) are between ca. 43 and ca. 61 Ma. Single-grain through the AFT PAZ and AHe PRZ is a plausible solution. We interpret
dates from the Del Cobre Range (SH16-3) are between ca. 26 and ca. 62 this cooling event as the result of exhumation due to thrusting and
Ma. Single-grain AHe ages from the Del Cobre Range E (SH16-1) are erosion of the hanging-wall of the fault at ca. 45-35 Ma.
between ca. 18 and ca. 47 Ma. Additionally, sample SH16-19, which we A previous study by Letcher (2007) reported three AFT ages for the
consider to be an along-strike equivalent of sample SH16-1, produced Susques Range (ST-1, ST-2, and ST-3, Fig. 2, Table S2). Additionally,
dates between ca. 22 and 31 Ma. Sample SH16-40 from the western flank these authors reported Eocene apatite (U–Th)/He ages for the same
of the Aguilar Range yielded dates between ca. 15 and ca. 27 Ma. This samples (ST-1, weighted mean of 5 grains, 46.7 ± 3.1Ma; ST-2, weighted
sample did not contain many apatites. Two of the analyzed grains mean of 3 grains, 46.3 ± 1.5 Ma; ST-3, weighted mean of 3 grains, 37.9
proved not to be apatites, and two others were broken and partially ± 1.2 Ma; ST-4, weighted mean of 3 grains, 29.4 ± 2.9 Ma). Letcher
complete (SH16_40_1 and SH16_40_2) although larger than the only (2007) attributed the systematically older AHe ages compared to the
complete aliquot analyzed (SH16_40_4). The two incomplete grains AFT ages to possible contamination of the AHe ages by inclusions or He
produced younger dates and low eU concentrations (SH16_40_1, 15.7 ± implantation. She used AFT data from two samples (ST-2 and ST-3) to
1.1 Ma, 6.9 ppm eU and SH16_40_2, 24.6 ± 0.7 Ma, 7.1 ppm eU) model plausible T-t thermal histories using HeFTy (Ketcham, 2005)
compared to the third grain (SH16_40_4, 27 1.1 Ma, 50. 7 ppm eU). We based on which they proposed cooling at ca. 26-20 Ma.
used the three grains to constrain the cooling history through a thermal Previous and new AHe ages are similar (single ages between ca. 35
model discussed in the next section. and 50 Ma for 5 samples) which would require that any external process
responsible for producing older AHe ages would have had to be sys­
6.3. Interpretation: Cooling history and exhumation of Ordovician ranges tematic. Additionally, older AFT ages to the west (Lina Range) provide a
reasonable context for middle Eocene AHe ages in the Susques Range.
The new thermochronology data from samples across the northern Thus, although it is possible that the AHe ages are in general older than
Puna plateau record Eocene to Oligocene (Fig. 10) cooling of rocks AFT ages due to inclusions or He implantation, long residence time in
passing through the uppermost 3.3 km of crust (assuming 30 ◦ C/km and the PAZ could have produced radiation damage which can increase 4He
surface temperature of 20 ◦ C). We interpret the cooling ages as the retentivity and thus, produce AHe ages older than AFT ages (inverted
timing of exhumation and erosion in the hanging-walls of active struc­ ages; Shuster et al., 2006; Fox and Shuster, 2014). In any case, partially
tures (e.g., McQuarrie and Ehlers, 2017) and thus to represent the timing reset AFT ages and slightly older AHe ages suggest limited exhumation
of shortening in the Andean thrust belt. and erosion (ca. 2.0–3.3 km assuming a paleogeothermal gradient of 30

C/km) in the range, likely starting at ca. 45-35 Ma.
6.3.1. Olaroz and Lina Ranges
Two new AFT ages from the Olaroz Range indicate cooling at 67 ± 6.3.3. Del Cobre and Tusaquillas Ranges
6.6 Ma (SH16-9) and 69.5 ± 5.8 Ma (SH16-10). The Olaroz Range is Two new cooling ages from the Del Cobre Range (SH16-33 and
bounded by an east-vergent fault to the east and by an inferred west- SH16-3) and one multi-dated sample in the Tusaquillas Range (SH16-1)
vergent fault on the western flank (Fig. 4). These faults uplifted the help constrain exhumation in three main thrust sheets (Figs. 2 and 5). An
range during the Miocene (after ca. 22 Ma) disconnecting the Paso de AFT sample (SH16-33) from the central part of the Del Cobre Range
Jama depocenter from the El Toro depocenter (Fig. 4). The stratigraphic (Fig. 5) shows cooling at 27.2 ± 2.5 Ma. Five AHe ages from sample
record suggests that these units were deposited within the same depo­ SH16-3 contrain cooling in the thrust sheet to the east. The oldest aliquot
center before being deformed; this implies erosion of at least 2 km of (SH16_3_4, 62.9 ± 1.5 Ma) has the lowest eU (ca. 64.3 ppm). The
Cenozoic strata from above the Olaroz Range. These two samples were combination of a relatively older age with low eU suggests He implan­
collected in the uppermost part of the Ordovician succession, suggesting tation in the grain. He implantation also explains an older AHe age
that only a small part of the Ordovician section is missing. Moreover, the compared to the AFT ages reported for samples from the same range (SC-
age distribution observed in the AFT radial plots (Fig. S5) shows a clear 1, SC-3, and SC-4, Fig. 2). The remaining grains produced ages between
trend toward younger ages that suggests that these samples (SH16-9 and ca. 26 and ca. 34 Ma, with eU between ca. 67 and 124 ppm and a size
sH16-10) resided in the lower part of the PAZ (O’Sullivan and Parrish, (Rs) between ca. 52 and 98 μm. These results depict a strong correlation
1995). We interpret these two Cretaceous ages to indicate <3.3 km of between age and eU, and between Rs and eU (Fig. S2) suggesting radi­
erosion (assuming 30 ◦ C/km and surface temperature of 20 ◦ C) due to ation damage in the sample. A thermal model based on four AHe grains
the Cenozoic shortening and exhumation, ca. 2 km of which would be from sample SH16-3 (Fig. S4) shows monotonic cooling through the AHe
required to remove the Cenozoic foreland deposits. PRZ between ca. 40 and 10 Ma in the next thrust sheet to the east.
To the east, two samples from the Lina Range record the onset of For the Tusaquillas Range, five AHe ages and one AFT age allow
exhumation during Andean deformation. The two AFT ages evince determination of the cooling history of the range. Grain SH16_1_3 yiel­
cooling at 40.2 ± 3.8 Ma (SH16-11) and 43 ± 5 Ma (SH16-47). These ded the oldest age (47.2 ± 0.8 Ma) and lowest eU (76.7 ppm), suggesting
new ages suggest that the thrust front was located in the western Puna He implantation. Grain SH16_1_5 yielded an age of 44.1 ± 0.7 Ma with a
plateau producing shortening and inducing exhumation and erosion by high eU (174.3 ppm) and U/Th (3.225 atomic) ten times larger than the
mid-Eocene time (ca. 45-40 Ma), consistent with previous results re­ ratios of the other grains. Moreover, this grain is older than the AFT age
ported by Carrapa and DeCelles (2008) and Carrapa et al. (2012). for the same sample (SH16-1, 25 ± 2.4 Ma) thus inverting the expected
relation between these two techniques. The other grains produced ages
6.3.2. Susques Range between ca. 18 and ca. 31 Ma, with similar eU (ca. 68–76 ppm) and U/
In the Susques Range, a sample (SH16-4) collected in the hanging- Th (ca. 0.29–0.31 atomic). These three aliquots show no correlation
wall of a west-vergent thrust fault was multi-dated using AFT and AHe between age and eU, or Rs and eU (Fig. S2); however, these ages are,
techniques. The youngest aliquot (SH16_3_4) has a very high eU (ca. within error, equivalent to the AFT age. Thus we combined results from
115 ppm) suggesting possible He implantation or zonation (Farley et al., these two techniques in a thermal model to determine the thermal his­
2011; Gautheron et al., 2012; Murray et al., 2014). The other single ages tory of the sample. The thermal model (Fig. S3) indicates rapid cooling
are between ca. 47 and 61 Ma and show a weak correlation between eU through the PRZ and PAZ between ca. 30 and ca. 25 Ma in the east­
and the corrected age (Fig. S2), suggesting some radiation damage. A ernmost thrust sheet. Finally, four AHe ages from an along-strike

17
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

equivalent sample (SH16-19) show a strong correlation between age and sedimentation in the Tres Cruces area at that time. Since the sedimen­
eU, and between Rs and eU (Fig. S2), a process that produces older ages tary units that overlie the Casa Grande Formation are thick (up to 4 km)
due to a higher He retentivity in the apatite crystals. A thermal model and coarse (the Río Grande and Pisungo Formations), we speculate that
based on these samples shows monotonic cooling through the AHe PRZ significant exhumation and erosion happened in the Aguilar Range after
between ca. 30 and ca. 10 Ma. The Del Cobre Range has been extensively the Oligocene. We interpret that cooling due to faulting, exhumation,
dated by previous thermochronometric studies. Published cooling ages and erosion in the Aguilar Range could have started as early as ca. 27
in this range consist of seven AFT ages and four apatite (U–Th)/He ages Ma, but that at least 2 km of erosion happened between ca. 25 and 15
(Deeken, 2005; Letcher, 2007). Deeken (2005) reported four AFT ages Ma.
(SUS-1, SUS-2, SUS-3 and SUS-4, Fig. 10, Table S2) for the central range.
According to their results, the central part of the Del Cobre Range 7. Discussion: Regional implications
(Figs. 5 and 10) experienced cooling through the PAZ between ca. 31
and ca. 37 Ma (Figs. 5 and 10). For the next thrust to the east, Letcher 7.1. Timing of Andean exhumation and shortening in northern Puna
(2007) reported three AFT ages (SC-1, SC-3 and SC-4, Figs. 5 and 10, plateau
Table S2) and four apatite (U–Th)/He ages (SC-1, weighted mean of 2
grains, 30.7 ± 2.4 Ma; SC-2, weighted mean of 3 grains, 27.6 ± 0.9 Ma; The combination of bedrock cooling ages and timing of Cenozoic
SC-3, weighted mean of 2 grains, 26.4 ± 1.0 Ma; SC-4, weighted mean of basin formation shows that Andean shortening was accommodated by
3 grains, 32.4 ± 2.3 Ma). These authors considered that these AHe ages an eastward propagating orogenic system with local, west-vergent, out-
were affected by either inclusions or He implantation and used two AFT of-sequence faulting. This shortening contributed to different stages of
ages (SC-1 and SC-4) to model thermal histories using HeFTy (Ketcham, crustal thickening in the frontal and hinterland parts of the thrust belt.
2005). Based on the modeled T-t paths they proposed punctuated Henriquez et al. (2019) documented late Cretaceous-Eocene exhu­
cooling between ca. 28 and ca. 15 Ma or continuous cooling since ca. 28 mation coeval with shortening in the Cordillera de Domeyko of northern
Ma until the present. Chile, which was probably the location of the earliest part of the central
Our new ages mainly complement the previous ages in the Del Cobre Andean thrust belt (Arriagada et al., 2006). Cooling ages in the Lina
Range and point to the onset of exhumation and erosion between ca. 37- Range record the relocation of the thrust front by the middle Eocene (ca.
29 Ma in the central part of the range, between ca. 28 and 10 Ma for the 40–45 Ma). During this time, significant exhumation and deposition of
next thrust sheet to the east, and rapid exhumation and erosion between growth strata have been reported in the Cordillera de Domeyko and
ca. 30-25 Ma for the easternmost thrust sheet. The cooling ages decrease Salar de Atacama Basin (Maksaev and Zentilli, 1999; Mpodozis et al.,
eastward and require ca. 3.3 km of erosion (assuming a paleogeothermal 2005; Arriagada et al., 2006; Henriquez et al., 2019). To the north at 21
gradient of 30 ◦ C/km) induced by upper Eocene to Oligocene east- ◦
S, reset zircon (U–Th)/He ages indicate cooling at ca. 43-40 Ma in the
vergent shortening in the middle of the northern Puna plateau. During central Eastern Cordillera (Anderson et al., 2018). To the south at ca.
the Oligocene, cooling due to exhumation and erosion in the hanging- 24.5 ◦ S, the Geste Formation accumulated in a wedge-top depozone no
wall of an inferred east-vergent thrust sheet is also recorded to the later than ca. 38 Ma (DeCelles et al., 2007; Carrapa and DeCelles, 2008)
west (west of the Paso de Jama depocenter). A new AFT age of 26.3 ± indicating Eocene shortening in the western Puna plateau. Additional
2.9 Ma (SH16-5) suggests out-of-sequence shortening in the Puna information regarding the evolution of the thrust belt comes from the
plateau. exhumation history in the Olaroz Range, where Cretaceous cooling ages
indicate that rocks now at the surface did not experience temperatures
6.3.4. Aguilar Range higher than 120 ◦ C since that time; therefore, they were too shallow to
In the Aguilar Range, one sample (SH16-40) from the western side of have developed cleavage during the Andean shortening event. Thus,
the range yielded few apatites to analyze. A thermal model constrained most of the penetrative deformation within the Paleozoic strata in this
by three AHe ages shows slow cooling through the AHe PRZ between ca. area is pre-Andean, and probably late Paleozoic. This limited exhuma­
30 and ca. 15 Ma. However, two of the three apatites analyzed have very tion also indicates that the westernmost Puna plateau was never buried
low eU, increasing the uncertainty of the results. by a thick foreland basin, possibly because of the presence of
Two previous studies recorded cooling in the Aguilar Range. Deeken pre-Cenozoic paleotopography (Carrapa and DeCelles, 2015).
(2005) reported cooling between ca. 17 and ca. 23 Ma based on two AFT By late Eocene to early Oligocene time (ca. 37-31 Ma), the thrust
ages in the eastern part of the range (GA6 and GA7, Fig. 10). Insel et al. front propagated into the Del Cobre Range. During this time, exhuma­
(2012) used K-feldspar multi–diffusion domain modeling to propose a tion continued to the west in the Cordillera de Domeyko (Maksaev and
rapid cooling event (ca. 150-100 ◦ C) driven by exhumation and erosion Zentilli, 1999; Henriquez et al., 2019) after magmatism ceased (by ca.
in the Aguilar Range at ca. 35-25 Ma. 37 Ma; e.g., Coira et al., 1993; Haschke et al., 2002). Middle Eocene
Our AHe ages overlap and are slightly older than the AFT ages re­ exhumation driven by shortening in the western Puna contemporaneous
ported by Deeken (2005). Slow cooling suggested by the thermal model with exhumation to the west in the Cordillera de Domeyko suggests that
(Fig. S4) suggests that the sample (SH16-40) was close to the AHe PRZ the orogen was growing by shortening both in the hinterland and along
and, therefore, the eastern part of the Aguilar Range likely sat at a higher the eastern orogenic front. At 21 ◦ S, the thrust belt propagated westward
structural level than the western side. On the other hand, the rapid from the central Eastern Cordillera to the Altiplano (AFT ages; Ege et al.,
cooling event proposed by Insel et al. (2012) could also be reconciled 2007; Anderson et al., 2018). At ca. 26 ◦ S, an AFT age in the
with our AHe ages; however, the three thermochronometric contraints hanging-wall of a thrust fault west of Arizaro basin (37.3 ± 4.8 Ma;
(AHe, AFT and K-feldspar multi-diffusion domain modeling) cannot be Carrapa and DeCelles, 2015) and detrital AFT and AHe ages reveal late
reconciled in a single cooling history. Low apatite yields in the samples Eocene exhumation in the western Puna plateau (Carrapa et al., 2009).
could have affected the AFT ages, as discussed by Insel et al. (2012). If Moreover, stable isotope paleoelevation data suggest that the western
that is the case, the rapid cooling between ca. 35-25 Ma would have side of the Puna plateau had high elevations (>4 km) and dry conditions
required rapid erosion and sedimentation at the time. The stratigraphic no later than ca. 38-36 Ma (Canavan et al., 2014; Quade et al., 2015).
record to the east in the Tres Cruces area includes ~800 m of sedi­ The samples that record high paleoelevations are located ~330 km to
mentary rocks (Casa Grande Formation according to Boll and Hernán­ the south in a proximal foreland basin position during Eocene time. Our
dez, 1986, and Geste Formation according to DeCelles et al., 2011) data support the hypothesis that shortening could have been responsible
between late Eocene and Oligocene times. It is possible that a later for Eocene crustal thickening and surface uplift in the Puna plateau as
erosion event could have removed part of this stratigraphic record; suggested by Quade et al. (2015).
however, uplift of the ranges farther to the west could easily explain During the Oligocene (ca. 31-25 Ma), exhumation due to east-

18
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

vergent shortening is recorded in the Del Cobre Range and possibly in synkinematic illite growth. Since most of the folding and flattening in
the Aguilar Range. Exhumation due to shortening is also recorded in the the Ordovician strata happened prior to and during cleavage formation,
westernmost ridge of the northern Puna plateau at 23 ◦ S. These results most of the shortening in the Ordovician rocks is likely to be Carbonif­
document eastward propagation of the thrust front and out-of-sequence erous in age (K–Ar ages on synkinematic illites; Jacobshagen et al.,
shortening in the hinterland. At 21 ◦ S, AFT cooling ages in the Eastern 2002).
Cordillera and southern Altiplano (Elger et al., 2005; Ege et al., 2007) New apatite fission track and apatite (U–Th)/He ages document
record a similar exhumation history. There, shortening in the hinterland eastward propagation of the Andean thrust belt during middle Eocene
was accommodated by east-vergent deformation in the southern Alti­ through Oligocene times. Exhumation in the western part of the north­
plano at ca. 30-27 Ma (Elger et al., 2005; Ege et al., 2007), equivalent to ern Puna plateau started at ca. 45-40 Ma before propagating eastward
deformation in the westernmost ridge of the Puna plateau. However, the until at least ca. 25 Ma. Cretaceous AFT cooling ages in Ordovician strata
significant west-vergent shortening in the Eastern Cordillera which involved in Cenozoic deformation demonstrate that locally Ordovician
reached the Altiplano by ca. 27 Ma (e.g., Ege et al., 2007) is apparently rocks did not experience temperatures higher than ca. 120 ◦ C and thus,
not recorded in the northern Puna plateau. Andean deformation did not occur at temperatures required for pressure
After ca. 25 Ma, exhumation is recorded mainly in the Aguilar Range dissolution. Additionally, new thermochronologic data reveal two pe­
before propagating eastward into the Eastern Cordillera. Shortening riods of out-of-sequence exhumation: during the Oligocene as recorded
continued in the plateau, deforming Cenozoic deposits (Paso de Jama, El by an AFT age (ca. 26 Ma), and during the Miocene, as recorded by
Toro, Susques and Tres Cruces areas). Deformation ceased at ca. 6.5 Ma deformation of dated Cenozoic strata.
in the Susques area (Letcher, 2007). A similar deformation history is This study shows that Andean shortening in the northern Puna
observed at 21 ◦ S, where the thrust front migrated into the easternmost plateau took place at the same time as shortening in the Eastern
Eastern Cordillera at ca. 25 Ma (e.g., Anderson et al., 2018). Cordillera at Altiplano latitudes. However, a significant difference is that
Out-of-sequence deformation is recorded at ca. 19-10 Ma by AFT ages in strain within the Paleozoic strata in the northern Puna plateau cannot be
the southern Altiplano (Elger et al., 2005), by Miocene synorogenic explained entirely by the Cenozoic Andean event as documented in
deposits in the western Eastern Cordillera (Horton, 1998) and at ca. central Bolivia (McQuarrie and Davis, 2002). This along-strike differ­
15-10 Ma by AHe ages in the western Eastern Cordillera (Anderson et al., ence might reflect the distribution of Paleozoic shortening and/or the
2017, 2018). Out-of-sequence shortening at this latitude ended by ca. 10 style of Cenozoic deformation and mechanism for cleavage formation
Ma (Gubbels et al., 1993; Horton, 2005). A similar dynamic in the within the Paleozoic strata.
hinterland at these two latitudes suggests continuous evolution of the
thrust belt along the transition from the Altiplano to the Puna plateau CRediT authorship contribution statement
since late Oligocene.
Susana Henriquez: Conceptualization, Investigation, Methodology,
7.2. Accommodation of shortening Data curation, Funding acquisition, Project administration, Writing -
original draft, Writing - review & editing. Bárbara Carrapa: Concep­
Shortening is mainly accommodated by folding and cleavage tualization, Funding acquisition, Supervision, Writing - review & edit­
development within Ordovician strata and thrust faulting and associated ing. Amanda Hughes: Formal analysis, Writing - review & editing.
folding of Cenozoic rocks. Shortening of Ordovician basement has not George Davis: Conceptualization, Writing - review & editing. Patricia
been precisely estimated in this study. Strain varies across the ranges, Alvarado: Resources.
with general estimates based on line-length balancing from outcrops
suggesting 25–50% shortening. Additionally, strong spaced cleavage Declaration of competing interest
would add as much as 25–35% (Alvarez et al., 1978). If cleavage was
formed during the late Paleozoic, significant shortening and flattening The authors declare that they have no known competing financial
probably took place at that time. Andean strain likely increased the interests or personal relationships that could have appeared to influence
strain and at least locally rotated folds and cleavage. Microstructural the work reported in this paper.
studies of Ordovician rocks, in tandem with petrochronologic dating,
would be required to estimate the Andean contribution to shortening Acknowledgements
within Ordovician rocks. We speculate that it will be significantly less
than the Paleozoic shortening. Moreover, if a significant part of the Funding for this research was provided by GSA research grant,
shortening in the Ordovician section is of Paleozoic age, then it is likely Tinker Field Research Grant Program, AAPG grant, Lewis and Clark
that a Paleozoic-age detachment-controlled shortening, exhumation, Fund for Exploration and Field Research and by the National Science
and erosion during both Paleozoic and Andean shortening. Foundation CSEDI grant EAR-1565475. The manuscript was signifi­
cantly improved in response to comments by an anonymous reviewer.
8. Conclusions We are also grateful for the support of the editor.

New structural, geochronologic and thermochronologic data help Appendix A. Supplementary data
constrain the thermal conditions and time of deformation in the Puna
plateau. Cenozoic strata were deformed by horizontal shortening mainly Supplementary data to this article can be found online at https://doi.
through reverse and thrust faults and folding. The amounts of strain and org/10.1016/j.jsg.2020.104133.
the thermal conditions for deforming Ordovician versus Cenozoic strata
are significantly different and could not have taken place during a single References
shortening event.
The strain progression in the Ordovician basement requires at least 4 Allmendinger, R.W., Jordan, T.E., Kay, S.M., Isacks, B.L., 1997. The evolution of the
Altiplano-Puna plateau of the central Andes. Annu. Rev. Earth Planet Sci. 25,
stages: a first stage of upright folding, a second stage of flattening
139–174.
through cleavage formation, a third stage of low-angle thrusting, and a Alvarez, W., Englender, T., Geiser, P.E., 1978. Classification of solution cleavage in
fourth stage of local distributed shear. Cenozoic Andean deformation pelagic limestones. Geology 6, 263–266.
happened under near-surface conditions locally no deeper than ca. 3.3 Anderson, R.B., Long, S.P., Horton, B.K., Calle, A.Z., Ramirez, V., 2017. Shortening and
structural architecture of the Andean fold-thrust belt of southern Bolivia (21◦ S):
km, whereas the strain history in Ordovician strata requires conditions implications for kinematic development and crustal thickening of the central Andes.
for pressure dissolution of quartz and formation of cleavage planes with Geosphere 13. https://doi.org/10.1130/GES01433.1.

19
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

Anderson, R.B., Long, S.P., Horton, B.K., Thomson, S.N., Calle, A.Z., Stockli, D.F., 2018. Cenozoic central Andes. In: DeCelles, P.G., Ducea, M.N., Carrapa, B., Kapp, P.A.
Orogenic wedge evolution of the central Andes, Bolivia (21◦ S): implications for (Eds.), Geodynamics of a Cordilleran Orogenic System: the Central Andes of
Cordilleran cyclicity. Tectonics 37, 3577–3609. https://doi.org/10.1029/ Argentina and Northern Chile, 212. Geological Society of America Memoir. https://
2018TC005132. doi.org/10.1130/2015.1212(22.
Arriagada, C., Cobbold, P.R., Roperch, P., 2006. Salar de Atacama Basin: a record of Deeken, A., 2005. Construction of the Southern Eastern Cordillera, NW Argentina: from
compressional tectonics in the central Andes since the mid-Cretaceous. Tectonics 25, Early Cretaceous Extension to Middle Miocene Shortening, Constrained by AFT-
TC1008. https://doi.org/10.1029/2004TC001770. Thermochronometry. Ph.D. thesis, Berlin, Freien Universität Berlin.
Bahlburg, H., 1991. The Ordovician back-arc to foreland successor basin in the Dickinson, W.R., Gehrels, G.E., 2009. Use of U-Pb ages of derital zircons to infer
Argentinean–Chilean Puna: tectonosedimentary trends and sea-level changes. In: maximum depositional ages of strata: a test against a Colorado Plateau Mesozoic
Macdonald, D.I.W. (Ed.), Sedimentation, Tectonics and Eustasy, 12. International database. Earth Planet Sci. Lett. 288, 115–125.
Association of Sedimentologists Special Publication, pp. 465–484. Echavarria, L., Hernández, R., Allmendinger, R., Reynolds, J., 2003. Subandean thrust
Bahlburg, H., Berndt, J., Gerdes, A., 2016. The ages and tectonic setting of the Faja and fold belt of northwestern Argentina: geometry and timing of the Andean
Eruptiva de la Puna oriental, ordovician, NW Argentina. Lithos 256–257, 41–54. evolution. AAPG (Am. Assoc. Pet. Geol.) Bull. 87, 965–985. https://doi.org/
Barnes, J.B., Ehlers, T.A., McQuarrie, N., O’Sullivan, P.B., Tawackoli, S., 2008. 10.1306/01200300196.
Thermochronometer record of central Andean Plateau growth, Bolivia (19.5◦ S). Ege, H., Sobel, E.R., Scheuber, E., Jacobshagen, V., 2007. Exhumation history of the
Tectonics 27, TC3003. https://doi.org/10.1029/2007TC002174. southern Altiplano plateau (southern Bolivia) constrained by apatite fission track
Beck, S.L., Zandt, G., Ward, K.M., Scire, A., 2015. Multiple styles and scales of thermochronology. Tectonics 26, TC1004. https://doi.org/10.1029/2005TC001869.
lithospheric foundering beneath the Puna plateau, central Andes. In: DeCelles, P.G., Egenhoff, S.O., 2007. Life and death of a Cambrian–Ordovician basin: an Andean three-
Ducea, M.N., Carrapa, B., Kapp, P.A. (Eds.), Geodynamics of a Cordilleran Orogenic act play featuring Gondwana and the Arequipa-Antofalla terrane. In: Linnemann, U.,
System: the Central Andes of Argentina and Northern Chile, 212. Geological Society et al. (Eds.), The Evolution of the Rheic Ocean: from Avalonian-Cadomian Active
of America Memoir. https://doi.org/10.1130/2015.1212(03. Margin to Alleghenian- Variscan Collision, 423. Geological Society of America
Boll, A., Hernández, R., 1986. Interpretación estructural del área de Tres Cruces. Bol. Inf. Special Paper, pp. 511–524. https://doi.org/10.1130/2007.2423(26).
Pet. (1924) 7, 2–14. Ehlers, T.A., Farley, K.A., 2003. Apatite (U-Th)/He thermochronometry: methods and
Buatois, L., Mangano, M., Brussa, E., Benedetto, J., Pompei, J., 2009. The changing face applications to problems in tectonic and surface processes. Earth Planet Sci. Lett.
of the deep: colonization of the early Ordovician deep-seafloor, Puna, northwest 206, 1–14.
Argentina. Palaeogeogr. Palaeoclimatol. Palaeoecol. 280, 291–299. Elger, K., Oncken, O., Glodny, J., 2005. Plateau style accumulation of deformation:
Calle, A.Z., Horton, B.K., Limachi, R., Stockli, D.F., Uzeda-Orellana, G.V., Anderson, R.B., southern Altiplano. Tectonics 24, TC4020. https://doi.org/10.1029/2004TC001675.
Long, S.P., 2018. Cenozoic provenance and depositional record of the Subandean Farley, K.A., 2000. Helium diffusion from apatite: general behavior as illustrated by
foreland basin during growth and advance of the central Andean fold-thrust belt, Durango fluorapatite. J. Geophys. Res. 105 (B2), 2903–2914. https://doi.org/
southern Bolivia. In: Zamora, G., McClay, K., V, R. (Eds.), Petroleum Basins and 10.1029/1999JB900348.
Hydrocarbon Potential of the Andes of Peru and Bolivia, 117. American Association Farley, K.A., Wolf, R.A., Silver, L.T., Wolf, L.T., 1996. The effects of long alpha-stopping
of Petroleum Geologists Memoir, pp. 459–506. https://doi.org/10.1306/ distances on (U–Th)/He ages. Geochem. Cosmochim. Acta 60, 4223–4229.
13622132M1173777. Farley, K.A., Shuster, D.L., Ketcham, R.A., 2011. U and Th zonation in apatite observed
Canavan, R., Carrapa, B., Clementz, M.T., Quade, J., DeCelles, P.G., Schoenbohm, L.M., by laser ablation ICPMS, and implications for the (U–Th)/He system. Geochem.
2014. Early Cenozoic uplift of the Puna Plateau, central Andes, based on stable Cosmochim. Acta 75, 4515–4530. https://doi.org/10.1016/j.gca.2011.05.020.
isotope paleoaltimetry of hydrated volcanic glass. Geology 42, 447–450. https://doi. Flowers, R.M., Ketcham, R.A., Shuster, D.L., Farley, K.A., 2009. Apatite (U–Th)/He
org/10.1130/G35239.1. thermochronometry using a radiation damage accumulation and annealing model.
Carrapa, B., DeCelles, P.G., 2008. Eocene exhumation and basin development in the Puna Geochem. Cosmochim. Acta 73 (8), 2347–2365.
of northwestern Argentina. Tectonics 27, TC1015. https://doi.org/10.1029/ Flowers, R.M., Farley, K.A., Ketcham, R.A., 2015. A reporting protocol for
2007TC002127. thermochronologic modeling illustrated with data from the Grand Canyon. Earth
Carrapa, B., DeCelles, P.G., 2015. Regional exhumation and kinematic history of the Planet Sci. Lett. 432, 425–435. https://doi.org/10.1016/j.epsl.2015.09.053.
central Andes in response to cyclical orogenic processes. In: DeCelles, P.G., Fox, M., Shuster, D.L., 2014. The influence of burial heating on the (U–Th)/He system in
Ducea, M.N., Carrapa, B., Kapp, P.A. (Eds.), Geodynamics of a Cordilleran Orogenic apatite: grand Canyon case study. Earth Planet Sci. Lett. 397, 174–183. https://doi.
System: the Central Andes of Argentina and Northern Chile, 212. Geological Society org/10.1016/j.epsl.2014.04.041.
of America Memoir. https://doi.org/10.1130/2015.1212(11. Galbraith, R.F., 1981. On statistical models for fission-track counts. J. Mathemati. Geosci.
Carrapa, B., DeCelles, P.G., Reiners, P.W., Gehrels, G.E., Sudo, M., 2009. Apatite triple 13, 471–478.
dating and white mica 40Ar/39Ar thermochronology of syntectonic detritus in the Gautheron, C., Tassan-Got, L., Ketcham, R.A., Dobson, K.J., 2012. Accounting for long
central Andes: a multiphase tectonothermal history. Geology 37, 407–410. https:// alpha- particle stopping distances in (U–Th–Sm)/He geochronology: 3D modeling of
doi.org/10.1130/G25698A.1. diffusion, zoning, implantation, and abrasion. Geochem. Cosmochim. Acta 96,
Carrapa, B., Bywater-Reyes, S., DeCelles, P.G., Mortimer, E., Gehrels, G., 2012. 44–56. https://doi.org/10.1016/j.gca.2012.08.016.
Eocene–Miocene synorogenic basin evolution in the Eastern Cordillera of Green, P.F., 1981. A new look at statistics in fission track dating. Nucl. Tracks Radiat.
northwestern Argentina (25◦ –26◦ S): regional implications for Andean orogenic Meas. 5, 77–86.
wedge development. Basin Res. 23, 1–20. Green, P.F., Duddy, I.R., Laslett, G.M., Hegarty, K.A., Gleadow, A.J.W., Lovering, J.F.,
Cladouhos, T.T., Allmendinger, R.W., Coira, B., Farrar, E., 1994. Late Cenozoic 1989. Thermal annealing of fission tracks in apatite 4: quantitative modeling
deformation in the central Andes: fault kinematics from the northern Puna, techniques and extension to geologic timescales. Chem. Geol. 79, 155–182.
northwestern Argentina and southwestern Bolivia. J. S. Am. Earth Sci. 7, 209–228. Grier, M.E., Salfity, J.A., Allmendinger, R.W., 1991. Andean reactivation of the
Cobbold, P.R., Rosello, E.A., 2003. Aptian to recent compressional deformation, foothills Cretaceous Salta rift, northwestern Argentina. J. S. Am. Earth Sci. 4, 351–372.
of the Neuquen Basin, Argentina. Mar. Petrol. Geol. 20, 429–433. Gubbels, T.L., Isacks, B.L., Farrar, E., 1993. High-level surfaces, plateau uplift, and
Coira, B.L., 1979. Descripción geológica de la Hoja 3c abra pampa, provincia de Jujuy, foreland development, Bolivian central Andes. Geology 21, 695–698.
Argentina. Servicio Geológico Nacional 170, 1–90. Halpern, M., Latorre, C.O., 1973. Estudio geocronológico inicial de rocas del Noroeste de
Coira, B., Zappettini, E.O., 2008. Mapa Geológico de la provincia de Jujuy. Servicio la República Argentina. Rev. Asoc. Geol. Argent. 28, 195–205.
Geológico Minero Argentino, 1. Instituto de Geología y Recursos Minerales, Buenos Haschke, M., Scheuber, E., Gunthner, A., Reutter, K.J., 2002. Evolutionary cycles during
Aires, 500,000. the Andean orogeny: repeated slab breakoff and fl at subduction? Terra. Nova 14,
Coira, B., Kay, S.M., Viramonte, J., 1993. Upper Cenozoic magmatic evolution of the 49–55. https://doi.org/10.1046/j.1365-3121.2002.00387.x.
Argentine Puna A model for changing subduction geometry. Int. Geol. Rev. 35 (8), Hauser, N., Matteini, M., Omarini, R., Pimentel, M.M., 2011. Combined U–Pb and Lu–Hf
677–720. https://doi.org/10.1080/00206819309465552. isotope data on turbidites of the Paleozoic basement of NW Argentina and petrology
Coutand, I., 1999. Ph.D. theses. Tectonique cenozoique du haut plateau de la Puna, Nord- of associated igneous rocks: implications for the tectonic evolution of western
Ouest argentin, Andes Centrales, 1. Universite de Rennes. Gondwana between 560 and 460 Ma. Gondwana Res. 19, 100–127. https://doi.org/
Coutand, I., Cobbold, P.R., de Urreiztieta, M., Gautier, P., Chauvin, A., Gapais, D., 10.1016/j.gr.2010.04.002.
Rossello, E.A., López-Gamundí, O., 2001. Style and history of andean deformation, Henriquez, S., DeCelles, P.G., Carrapa, B., 2019. Cretaceous to middle Cenozoic
Puna plateau, northwestern Argentina. Tectonics 20, 210–234. https://doi.org/ exhumation history of the Cordillera de Domeyko and Salar de Atacama basin,
10.1029/2000TC900031. northern Chile. Tectonics 38, 395–416. https://doi.org/10.1029/2018TC005203.
Davis, G.H., Reynolds, S.J., Kluth, C.F., 2011. Structural Geology of Rocks and Regions. Hernández, R.M., Echavarria, L., 2009. Faja plegada y corrida Subandina del noroeste
John Wiley and Sons, United States. Argentina: estratigrafía, geometría y cronología de la deformación. Rev. Asoc. Geol.
DeCelles, P.G., Horton, B.K., 2003. Early to Middle Tertiary foreland basin development Argent. 65, 68–80.
and the history of Andean crustal shortening in Bolivia. Geol. Soc. Am. Bull. 115, Horton, B.K., 1998. Sediment accumulation on top of the Andean orogenic wedge:
58–77. https://doi.org/10.1130/0016-7606(2003)115<0058:ETMTFB>2.0.CO;2. Oligocene to late Miocene basins of the Eastern Cordillera, southern Bolivia. Geol.
DeCelles, P.G., Carrapa, B., Gehrels, G.E., 2007. Detrital zircon U-Pb ages provide new Soc. Am. Bull. 110, 1174–1192.
provenance and chronostratigraphic information from Eocene synorogenic deposits Horton, B.K., 2005. Revised deformation history of the central Andes: inferences from
in northwestern Argentina. Geology 35, 323–326. Cenozoic foredeep and intermontane basins of the Eastern Cordillera, Bolivia.
DeCelles, P.G., Carrapa, B., Horton, B.K., Gehrels, G.E., 2011. Cenozoic foreland basin Tectonics 24, TC3011. https://doi.org/10.1029/2003TC001619.
system in the central Andes of northwestern Argentina: implications for Andean Hurford, A.J., Green, P.F., 1983. The Zeta age calibration of fission track dating.
geodynamics and modes of deformation. Tectonics 30, TC6013. https://doi.org/ Chemical Geology (Isotope Geoscience Section) 1, 285–317.
10.1029/2011TC002948. Insel, N., Grove, M., Haschke, M., Barnes, J.B., Schmitt, A.K., Strecker, M.R., 2012.
DeCelles, P.G., Zandt, G., Beck, S.L., Currie, C.A., Ducea, M.N., Kapp, P., Gehrels, G.E., Paleozoic to early Cenozoic cooling and exhumation of the basement underlying the
Carrapa, B., Quade, J., Schoenbohm, L.M., 2015. Cyclical orogenic processes in the

20
S. Henríquez et al. Journal of Structural Geology 140 (2020) 104133

eastern Puna plateau margin prior to plateau growth. Tectonics 31, TC6006. https:// Mpodozis, C., Arriagiada, C., Basso, M., Roperch, P., Cobbold, P., Reich, M., 2005. Late
doi.org/10.1029/2012TC003168. mesozoic to paleogene stratigraphy of the salar de Atacama basin, antofagasta,
Isacks, B.L., 1988. Uplift of the central Andean plateau and bending of the Bolivian northern Chile: implications for the tectonic evolution of the central Andes.
orocline. J. Geophys. Res. 93, 3211–3231. Tectonophysics 399, 125–154. https://doi.org/10.1016/j.tecto.2004.12.019.
Jacobshagen, V., Müller, J., Wemmer, K., Ahrendt, H., Manutsoglu, E., 2002. Hercynian Müller, J.P., Kley, J., Jacobshagen, V., 2002. Structure and cenozoic kinematics of the
deformation and metamorphism in the Cordillera Oriental of southern Bolivia, eastern cor- dillera, southern Bolivia (21◦ S). Tectonics 21, 1037. https://doi.org/
central Andes. Tectonophysics 345, 119–130. 10.1029/2001TC001340.
Ketcham, R.A., 2005. Forward and inverse modeling of low-temperature Murray, K.E., Orme, D.A., Reiners, P.W., 2014. Effects of U–Th-rich grain boundary
thermochronometry data. In: Reiners, P.W., Ehlers, T.A. (Eds.), Low-temperature phases on apatite helium ages. Chem. Geol. 390, 135–151. https://doi.org/10.1016/
Thermochronology, 58. Mineralogical Society of America, Chantilly, Virginia, j.chemgeo.2014.09.023.
pp. 275–314. Niemeyer, H., 1989. El Complejo ígneo-sedimentario del Cordón de Lila, Región de
Ketcham, R.A., Carter, A.C., Donelick, R.A., Barbarand, J., Hurford, A.J., 2007. Improved Antofa- gasta: significado tectónico. Rev. Geol. Chile 16, 163–181.
modeling of fission-track annealing in apatite. Am. Mineral. 92 (5–6), 799–810. Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P., Schemmann, K., 2006. Deformation
https://doi.org/10.2138/am.2007.2281. of the central Andean upper plate system-Facts, fiction, and constraints for plateau
Ketcham, R.A., Gautheron, C., Tassan-Got, L., 2011. Accounting for long alpha-particle models; central and southern Andean tectonic evolution inferred from arc
stopping distances in (U–Th–Sm)/He geochronology: refinement of the baseline magmatism. In: Oncken, O., Chong, G., Franz, G., Giese, P., Götze, H.-J., Ramos, V.,
case. Geochimica et Cosmochimica Acta. 75, 7779–7791. Strecker, M., Wigger, P. (Eds.), The Andes-Active Subduction Orogeny. Frontiers in
Kley, J., 1996. Transition from basement-involved to thin- skinned thrusting in the Earth Sciences. Springer, Berlin, pp. 3–27.
Cordillera Oriental of southern Bolivia. Tectonics 15, 763–775. O’Sullivan, P.B., Parrish, R.R., 1995. The importance of apatite composition and single-
Kley, J., Monaldi, C.R., 2002. Tectonic inversion in the Santa Bárbara system of the grain ages when interpreting fission track data from plutonic rocks: a case study from
central Andean foreland thrust belt, northwestern Argentina. Tectonics 21 (6), 1061. the Coast Ranges, British Columbia. Earth Planet Sci. Lett. 132, 213–224.
https://doi.org/10.1029/2002TC902003. Pearson, D.M., Kapp, P., DeCelles, P.G., Reiners, P.W., Gehrels, G.E., Ducea, M.N.,
Kley, J., Reinhardt, M., 1994. Geothermal and tectonic evolution of the eastern Pullen, A., 2013. Influence of pre-Andean crustal structure on Cenozoic thrust belt
Cordillera and the subandean ranges of southern Bolivia. In: Reutter, K.J., kinematics and shortening magnitude: northwestern Argentina. Geosphere 9 (6),
Scheuber, E., Wigger, P.J. (Eds.), Tecronics of the Southern Cenrrul Andes. Springer 1766–1782. https://doi.org/10.1130/GES00923.1.
Verlag, Berlin, pp. 155–170. Quade, J., Dettinger, M.P., Carrapa, B., DeCelles, P., Murray, K.E., Huntington, K.A.,
Kley, J., Müller, J., Tawackoli, S., Jacobshagen, V., Manutsoglu, E., 1997. Pre-andean Cartwright, A., Canavan, R.R., Gehrels, G., Clementz, M., 2015. The growth of the
and andean-age deformation in the eastern Cordillera of southern Bolivia. J. S. Am. central Andes 22–26◦ S. In: DeCelles, P.G., Ducea, M.N., Carrapa, B., Kapp, P.A.
Earth Sci. 10, 1–19. (Eds.), Geodynamics of a Cordilleran Orogenic System: the Central Andes of
Kley, J., Rossello, E.A., Monaldi, C.R., Habighorst, B., 2005. Seismic and field evidence Argentina and Northern Chile, 212. Geological Society of America Memoir. https://
for selective inversion of Cretaceous normal faults, Salta rift, northwestArgentina. doi.org/10.1130/2015.1212(15).
Tectonophysics 399, 155–172. https://doi.org/10.1016/j.tecto.2004.12.020. Reiners, P.W., Brandon, M.T., 2006. Using thermochronology to understand orogenic
Kortyna, C., DeCelles, P.G., Carrapa, B., 2019. Structural and thermochronologic erosion. Annu. Rev. Earth Planet Sci. 34, 419–466.
constraints on kinematics and timing of inversion of the Salta rift in the Tonco- Rivadeneyra-Vera, C., Bianchi, M., Assumpção, M., Cedraz, V., Julià, J., Rodríguez, M.,
Amblayo sector of the Andean retroarc fold-thrust belt, northwestern Argentina. In: et al., 2019. An updated crustal thickness map of central South America based on
Horton, B.K., Folguera, A. (Eds.), Andean Tectonics. Elsevier, pp. 429–464. receiver function measurements in the region of the Chaco, Pantanal, and Paraná
Letcher, A.J., 2007. Deformation History of the Susques Basin (~23◦ S, 66◦ W), Puna Basins, southwestern Brazil. J. Geophys. Res. Solid. Earth. 124 https://doi.org/
Plateau, NW Argentina: New Constraints by Apatite (U-Th)/He Thermochronology 10.1029/2018JB016811.
and 40Ar/39Ar Geochronology, and Implications for Plateau Formation in the Salfity, J.A., Marquillas, R.A., 1994. Tectonic and sedimentary evolution of the
Central Andes. MSc Thesis. Stanford University, USA. cretaceous-eocene Salta group basin, Argentina. In: Salfity, J.A. (Ed.), Cretaceous
Lork, A., Bahlburg, H., 1993. Precise U-Pb ages of monazites from the Faja Eruptiva de la Tectonics of the Andes. Vieweg Verlag, Wiesbaden, Germany, pp. 266–315.
Puna oriental and the Cordillera oriental, NW Argentina. In: XII Congreso Geológico Sempere, T., Butler, R.F., Richards, D.R., Marshall, L.G., Sharp, W., Swisher, C.C., 1997.
Argentino y II Congreso de Exploración de Hidrocarburos. Actas IV, Argentina, Stratigraphy and chronology of late cretaceous-early paleogene strata in Bolivia and
pp. 1–6. northwest Argentina. Geol. Soc. Am. Bull. 109, 709–727.
Ludwig, K.R., 2008. Isoplot 3.6, 4. Berkeley Geochronology Center Special Publication. Shuster, D.L., Flowers, R.M., Farley, K.A., 2006. The influence of natural radiation
Maksaev, V., Zentilli, M., 1999. Fission track thermochronology of the Domeyko damage on helium diffusion kinetics in apatite. Earth Planet Sci. Lett. 249, 148–161.
Cordillera, northern Chile: implications for Andean tectonics and porphyry copper https://doi.org/10.1016/j.epsl.2006.07.028.
metallogenesis. Explor. Min. Geol. 8, 65–89. Siks, B.C., Horton, B.K., 2011. Growth and fragmentation of the Andean foreland basin
Marquillas, R.A., Del Papa, C., Sabino, I.F., 2005. Sedimentary aspects and during eastward advance of fold-thrust deformation, Puna plateau and Eastern
paleoenvironmental evolution of a rift basin: Salta Group (Cretaceous-Paleogene), Cordillera, northern Argentina. Tectonics 30, TC6017. https://doi.org/10.1029/
northwestern Argentina. Int. J. Earth Sci. 94, 94–113. https://doi.org/10.1007/ 2011TC002944.
s00531-004-0443-2. Starck, D., 2011. Cuenca Cretácica-Paleógena del noroeste Argentino: VIII Congreso de
Marrett, R.A., Allmendinger, R.W., Alonso, R.N., Drake, R.E., 1994. Late Cenozoic Exploración y Desarrollo de Hidrocarburos Simposio Cuencas Argentinas: vision
tectonic evolution of the Puna plateau and adjacent foreland, northwestern actual. IAPG, Instituto Argentino del Petróleo y el Gas, pp. 407–453.
Argentine Andes. J. S. Am. Earth Sci. 7, 179–207. Starck, D., Anzótegui, L.M., 2001. The late Miocene climate change-persistence of
McClay, K.R., 1977. Pressure solution and Coble creep in rocks. J. Geolo. Stud. London. climate signal through the orogenic stratigraphic record in northwestern Argentina.
134, 57–70. J. S. Am. Earth Sci. 14, 763–774.
McQuarrie, N., 2002. The kinematic history ofthe central Andean fold-thrust belt, Starck, D., Vergani, G., 1996. Desarrollo tecto-sedimentario del Cenozoico en el sur de la
Bolivia: implications for building a high plateau. Geol. Soc. Am. Bull. 114, 950–963. Provincia de Salta-Argentina. In: XIII Congreso Geológico Argentino. Actas I, Buenos
https://doi.org/10.1130/0016-7606(2002)114<0950:TKHOTC>2.0.CO;2. Aires, pp. 433–452.
McQuarrie, N., Davis, G.H., 2002. Crossing the several scales of strain-accomplishing Steinmetz, R.L., Galli, C.I., 2015. Basin development at the eastern border of the
mechanisms: the central Andean fold-thrust belt. J. Struct. Geol. 24, 1587–1602. Northern Puna and its relationship with the plateau evolution. J. S. Am. Earth Sci.
https://doi.org/10.1016/S0191-8141(01)00158-4. 63, 244–259. https://doi.org/10.1016/j.jsames.2015.07.017.
McQuarrie, N., Ehlers, T.A., 2017. Techniques for understanding fold-and-thrust belt Strecker, M.R., Alonso, R.N., Bookhagen, B., Carrapa, B., Hilley, G.E., Sobel, E.R.,
kinematics and thermal evolution. In: Law, R.D., Thigpen, J.R., Merschat, A.J., Trauth, M.H., 2007. Tectonics and climate of the southern central Andes. Annu. Rev.
Stowell, H.H. (Eds.), Linkages and Feedbacks in Orogenic Systems, 213. Geological Earth Planet Sci. 35, 747–787.
Society of America Memoir, pp. 25–54. Streit, R.L., Burbank, D.W., Strecker, M.R., Alonso, R.N., Cottle, J.M., Kylander-Clark, A.
Mon, R., Salfity, J.A., 1995. Tectonic evolution of the Andes of northern Argentina. In: R.C., 2017. Controls on intermontane basin filling, isolation and incision on the
Tankard, A.J., Welsink, H.J. (Eds.), Petroleum Basins of South America. AAPG margin of the Puna plateau, NW Argentina (~23◦ S). Basin Res. 29, 131–155.
Memoir, Tulsa, pp. 269–283. Tavani, S., Storti, F., Lacombe, O., Corradetti, A., Munoz, J.A., Mazzoli, S., 2015.
Monaldi, C.R., Salfity, J.A., Kley, J., 2008. Preserved extensional structures in an A review of deformation pattern templates in foreland basin systems and fold-and-
inverted Cretaceous rift basin, north- western Argentina: outcrop examples and thrust belts: implications for the state of stress in the frontal regions of thrust wedges.
implications for fault reactivation. Tectonics 27, TC1011. https://doi.org/10.1029/ Earth Sci. Rev. 141, 82–104.
2006TC001993. Turner, J.C.M., 1960. Estratigrafía de la Sierra de Santa Victoria y adjacencias, 41.
Montero-López, C., del Papa, C., Hongn, F., Strecker, M.R., Aramayo, A., 2018. Academia Nacional de Ciencias de Córdoba, pp. 163–196.
Synsedimentary broken-foreland tectonics during the Paleogene in the Andes of NW Welsink, H.J., Martinez, E., Aranibar, O., Jarandilla, J., 1995. Structural inversion of a
Argentine: new evidence from regional to centimetre-scale deformation features. Cretaceous rift basin, southern Altiplano, Bolivia. In: Tankard, A.J., Suarez
Basin Res. 30, 142–159. https://doi.org/10.1111/bre.12212. Soruco, R., Welsink, H.J. (Eds.), Petroleum Basins of South America. AAPG Memoir,
Mora, A., Casallas, W., Ketcham, R.A., Gomez, D., Parra, M., Namson, J., Stockli, D., Tulsa, pp. 305–324.
Almendral, A., Robles, W., Ghorbal, B., 2015. Kinematic restoration of con- Willett, S.D., 1992. Dynamic and kinematic growth and change of a Coulomb wedge. In:
tractional basement structures using thermokinematic models: a key tool for McClay, K.R. (Ed.), Thrust Tectonics. Chapman and Hall, New York, pp. 19–31.
petroleum system modeling. AAPG (Am. Assoc. Pet. Geol.) Bull. 99, 1575–1598. Wood, D.S., 1974. Current views of the development of slaty cleavage. Annu. Rev. Earth
https://doi.org/10.1306/04281411108. Planet Sci. 2, 369–401.
Moya, C., 2015. La “Fase Oclóyica” (Ordovícico Superior) en el noroeste argentino. Yonkee, W.A., Weil, A.B., 2015. Tectonic transect of the sevier and laramide belts, north
Interpretación histórica y evidencias en contrario. Contribuciones a la Geología American Cordillera: evolution of a complex orogenic system. Earth Sci. Rev. 150,
Argentina. Ser. Correlación Geol. 31 (1), 73–110. 531–593. https://doi.org/10.1016/j.earscirev.2015.08.001.

21

You might also like