You are on page 1of 30

TECTONICS, VOL. 30, TC3003, doi:10.

1029/2010TC002707, 2011

Patterns and timing of exhumation and deformation in the Eastern


Cordillera of NW Argentina revealed by (U‐Th)/He
thermochronology
Barbara Carrapa,1,2 John D. Trimble,2 and Daniel F. Stockli3
Received 19 March 2010; revised 14 January 2011; accepted 27 January 2011; published 18 May 2011.

[1] The Eastern Cordillera (EC) and related ranges of Bolivia and Argentina exhibit a
wide variety of structural features, both thick and thin skinned, that make this region
a prime area to study the evolution of these two contrasting styles. Using a combination
of structural, geochronological, and thermochronological techniques, this study
investigates how and in what order the various structures of the Argentinean EC from 25 to
26°S have developed during the Cenozoic. New mapping in the Angastaco area preserves
one of the thickest Cenozoic stratigraphic sections and records a complex structural
evolution during the Neogene, characterized by inversion of Cretaceous Salta Rift
structures. Detrital zircon U‐Pb geochronology combined with stratigraphic and structural
features typical of synsedimentary deformation constrains the age of orgenic growth in the
area to ∼14 Ma. Detrital apatite (U‐Th)/He thermochronology on samples collected
across the width of the southernmost EC at this latitude document an eastward younging of
ages interpreted as the result of sequential eastward propagation of exhumation (and
inferred deformation) from ∼14 to 3 Ma at a rate of ∼8.3 mm/a. Our data, when compared
with existing data, show that the Puna Plateau of NW Argentina was exhuming and
deforming at the same time as the EC and inter‐Andean regions of Bolivia, suggesting that
the deformation front connects along strike despite of the differences in structural style.
Whereas the deformation front reached the sub‐Andes of Bolivia by ∼10 Ma, deformation
localized in the EC of NW Argentina until ∼4 Ma. Rates of propagation through the
whole region seem to be quasi‐uniform regardless of different structural styles.
Citation: Carrapa, B., J. D. Trimble, and D. F. Stockli (2011), Patterns and timing of exhumation and deformation in the Eastern
Cordillera of NW Argentina revealed by (U‐Th)/He thermochronology, Tectonics, 30, TC3003, doi:10.1029/2010TC002707.

1. Introduction geometries, duplex structures and a systematic progression


of deformation that generally propagates toward the foreland
[2] The structural style of mountain belts greatly affects
through a thick package of sedimentary rocks with relatively
foreland basin development, mountain building, exhumation
low mechanical strength [e.g., Bally et al., 1966; Davis et al.,
and sediment recycling. Although each orogenic system is
1983]. Conversely, thick‐skinned belts are characterized by
unique, all mountain belts are subject to the same funda-
steeper faults in crystalline basement, deeper décollements,
mental tectonic processes of continental contraction and
and often experience more erratic, out‐of‐sequence devel-
lateral transport of upper crustal material. On the most
opment [e.g., Jordan, 1995]. Because of the relatively high
fundamental scale, the transported rocks of a fold‐and‐thrust
mechanical strength of basement rocks, faults in thick‐
belt act like a wedge of sand or snow in front of a moving skinned systems often nucleate along preexisting hetero-
bulldozer, incorporating new sediments and rocks at the
geneities, such as fabrics, ancient shear zones, pervasive
frontal tip across the basal thrust, and sometimes deforming
fracture patterns, or inherited fault systems [Allmendinger
internally within the wedge to maintain a critical taper [Bally et al., 1983; Schmidt et al., 1995].
et al., 1966; Dahlstrom, 1970].
[4] The Eastern Cordillera of the central Andes involves
[3] Thin‐skinned systems are characterized by low‐angle
substantial amounts of mechanical ‘basement’, in this case
thrust faults, multiple shallow décollements, ramp flat
previously deformed low‐grade metasedimentary rocks [e.g.,
McQuarrie et al., 2005]. A key question addressed in this
1
Department of Geosciences, University of Arizona, Tucson, Arizona, paper is whether or not the timing of deformation in the
USA. Eastern Cordillera (EC) of NW Argentina is similar to what
2
Department of Geology and Geophysics, University of Wyoming,
Laramie, Wyoming, USA.
observed along strike to the north in Bolivia, where the EC is
3
Department of Geology, Kansas University, Lawrence, Kansas, USA. characterized by a more thin‐skinned style of deformation
[Kley, 1999; McQuarrie, 2002].
Copyright 2011 by the American Geophysical Union.
0278‐7407/11/2010TC002707

TC3003 1 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

samples from exhumed Cretaceous and Cenozoic strata


from the Calchaquí Valley and surrounding ranges in the EC
(Figure 2) to determine how and in what order structures
have evolved. This multidisciplinary study provides infor-
mation on the magnitude and timing of cooling and exhu-
mation related to recent deformational events.
[7] A number of recent studies have documented rates and
styles of deformation in various parts of the central Andes
[Mon and Salfity, 1995; Echavarria et al., 2003; Elger et al.,
2005; Oncken et al., 2006; Ege et al., 2007; Barnes et al.,
2008; McQuarrie et al., 2008; Uba et al., 2009; Barnes
and Ehlers, 2009]. Our new data when combined with
existing data highlight the variations in the timing and rate
of propagation of deformation that exist on a regional scale
between the northern and southern segments of the EC
providing a basis for discussion on possible mechanisms
controlling deformation and erosion in the region.

2. Background
2.1. Regional Setting
[8] Crustal thickening and exhumation in the central
Andes (15°–30°S) of South America is largely the result of
shortening that began in Chile during the Cretaceous and
propagated into Argentina and Bolivia during the Cenozoic
[McQuarrie et al., 2005; Arriagada et al., 2006; Carrapa
and DeCelles, 2008]. The eastward migration of the oro-
genic front has progressively uplifted and exhumed rocks
that were once part of the regional Andean retroarc foreland
basin system, forming the high topography and structures
observed in the Andes today (Figures 1 and 3). The Puna‐
Altiplano plateau forms the high‐altitude, relatively low
internal relief morphology of the central Andes (Figure 1).
With an average elevation of ∼4 km [Isacks, 1988], it is the
Figure 1. Shaded regional topography of the central Andes largest plateau in a retroarc setting. The Puna‐Altiplano
of northern Argentina and southern Bolivia, showing tecto- plateau is bounded to the west by the magmatic arc and
nomorphic zones (labeled), thrust faults (barbed lines), and monoclinal flexure of the Western Cordillera and to the east
volcanoes (circles) [modified after Sobel et al., 2003]. by the varied contractional systems of the Eastern Cordillera
Black box shows area covered in Figure 2. (EC), Santa Barbara System, sub‐Andes, and Sierras Pam-
peanas (Figure 1).
[9] Foreland basin deformation in the central Andes
exhibits a variety of structural styles that can be broadly
[5] The Eastern Cordillera (EC) in Argentina and Bolivia correlated along strike to the angle of subduction of the
is a large retroarc fold‐and‐thrust belt that has developed Nazca plate under the South American plate [Isacks, 1988;
during the Cenozoic along the eastern margin of the central Jordan and Allmendinger, 1986]. Thick‐skinned tectonics
Andean Plateau from ∼13 to 26°S (Figure 1). The basement deformation in the Sierras Pampeanas is associated with the
of the EC in NW Argentina is mainly characterized by high‐ 10° “flat slab” subduction under central Argentina, whereas
angle reverse faults cutting into Precambrian‐Cambrian the thin‐skinned fold‐and‐thrust belts of the EC and sub‐
crystalline and metamorphic rocks, and locally Cretaceous Andes are associated with the 30° “normal” subduction
rift sedimentary rocks, with a décollement at >30 km depth under Bolivia [Cahill and Isacks, 1992]. However, this
[Ramos, 2002]; whereas the basement in Bolivia is charac- relationship breaks down between 25° and 28°S where the
terized by thrust faults cutting into Ordovician and Silurian subducting plate is represented by a broad transitional flex-
metasedimentary rocks [McQuarrie, 2002]. This along‐ ure at an angle between 10° and 30° [Ramos, 2002]. In this
strike variability makes the EC a prime laboratory to region, the angle of the subducting slab appears to have less
examine the factors that control the sequence in which faults influence on the structural style of foreland deformation than
activate within the orogenic system, the rate of exhumation mechanical weaknesses in the crust related to preexisting
associated to deformation on those faults, and the style of basement anisotropy. This study focuses on the southern EC
foreland basin fragmentation. in this transitional interval, where structural style is con-
[6] This study combines detailed mapping of the Cal- trolled by a complex interplay between inherited structures,
chaquí Valley (Figure 1) with apatite (U‐Th)/He thermo- inversion of Cretaceous rift structures and foreland basin
chronology and zircon U‐Pb geochronology of sandstone development [Strecker et al., 1989; Grier et al., 1991; Salfity

2 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Figure 2. General geologic map of the southern Eastern Cordillera in NW Argentina. Location of this
map is outlined as a black box in Figure 1. Apatite (U‐Th)/He sample locations are indicated with yellow
circles, and detrital zircon U‐Pb sample locations are indicated with blue stars. The general east and west
margins of the Brealito, Alemanía, and Metán subbasins of the Salta rift are outlined at the bottom
(modified after Carrera et al. [2006], Hongn and Seggiaro [2001], Marquillas et al. [2005], G. Vergani
and D. Starck (unpublished map. 1988), and new mapping from this study). Black box indicates area
mapped in Figure 5.

and Marquillas, 1994; Kley and Monaldi, 2002; Mon and magnitude of shortening over that distance (Figure 1) with
Salfity, 1995]. average total shortening up to 270 km in Bolivia and 70 km
in NW Argentina [Kley et al., 1999; Mon and Drozdzewski,
2.2. Eastern Cordillera 1999; Elger et al., 2005; McQuarrie et al., 2005, 2008].
[10] The Eastern Cordillera bounds the eastern margin of Both the thick‐skinned and the thin‐skinned portions of the
the plateau for over 1,500 km along strike, and it varies EC are bivergent fold‐and‐thrust belts, with the west half of
greatly in topographic morphology, structural style, and the cordillera verging westward and the east half of the

3 of 30
TC3003

4 of 30
CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION

Figure 3. Schematic regional cross section illustrating the general geology and major structures of the Eastern Cordillera in
NW Argentina at 25.5°–25.8°S modified after Grier and Dallmeyer [1990]. Location of the cross section is indicated on
Figure 2. Apatite (U‐Th)/He sample locations are indicated with circles, and detrital zircon U‐Pb sample locations are
indicated with stars. Sample LY was collected ∼50 km south of the line of this section and is displayed here in a laterally
equivalent structural location. Existence and depth of regional décollement inferred from Cristallini et al. [1997] and Grier
et al. [1991]. Regional geology interpreted from Carrera et al. [2006], Cristallini et al. [2004], Hongn and Seggiaro [2001],
Marquillas et al. [2005], G. Vergani and D. Starck (unpublished map. 1988), and new mapping in the Calchaquí Valley
from this study (Figure 5).
TC3003
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

cordillera verging eastward. There is also a difference in 2.3. Stratigraphy


timing, with deformation in the Bolivian EC initiating ear- [13] There are three major stratigraphic components in the
lier than the Argentinean EC [McQuarrie et al., 2005]. southern EC (Figure 2), including (1) the metamorphic
[11] From ∼15 to 23°S, the Bolivian EC and sub‐Andes Precambrian‐Cambrian Puncoviscana Formation, (2) rift fill
are connected today as part of a single, retroarc fold‐and‐ sedimentary rocks of the Cretaceous Salta rift, and (3) a thick
thrust belt developed along the eastern margin of the Central Cenozoic sequence of synorogenic Andean foreland basin
Andean Plateau (Figure 1), although the structures observed strata (Figure 4) [Grier et al., 1991]. The Puncoviscana
today are a combination of pre‐Andean and Cenozoic Formation represents a thick passive margin sequence that
deformation. The fold‐and‐thrust belt evolved in places to was deposited during the Neoproterozoic to Cambrian,
form separate east and west verging systems; deformation in shortened and metamorphosed during the Pampean orogeny,
the EC and sub‐Andes has overall propagated eastward into and later intruded by multiple granitic bodies [Jezek and
the foreland during the Cenozoic [Echavarria et al., 2003; Miller, 1985; Willner and Miller, 1985; Grier et al., 1991].
Uba et al., 2009; McQuarrie et al., 2005, 2008]. Defor- The early Paleozoic metamorphism resulted in crustal
mation in the Argentinean EC from 23 to 26°S is markedly imbrication and in the development of east dipping shear
different from Bolivia. It exhibits a complex structural style belts and axial planar cleavages that permeate the formation
dominated by high‐angle reverse faults involving intrusive today [Willner et al., 1987].
and metamorphic basement and Cretaceous sedimentary [14] The Cretaceous‐Eocene Salta group rests uncon-
rocks [Grier et al., 1991; Mon and Salfity, 1995]. formably on the Puncoviscana Formation. The Salta Group
[12] There is some evidence from sedimentology and is over 5 km thick in places and is divided into three sub-
magnetostratigraphy to suggest that the range uplift in this groups: the Pirgua, Balbuena and Santa Barbara subgroups
part of the EC may have swept generally eastward as it (Figure 4). The Pirgua subgroup is composed of over 3 km
did in the Bolivian EC [Carrera and Muñoz, 2008; Reynolds of synrift sandstones, conglomerates, siltstones, and volca-
et al., 2000], but these studies are based on indirect proxies, nics of the La Yesera, La Curtiembres, and Los Blanquitos
and the sequence of deformation and associated exhumation Formations [Marquillas et al., 2005]. The Balbuena sub-
remains unresolved in this part of the EC. This study focuses group is composed of 400–500 m of postrift sandstones,
on the Angastaco area in the southernmost EC (Figures 1 limestones, and pelites. The Santa Barbara subgroup is
and 2), where Cenozoic shortening has inverted older composed of early foreland basin sandstones, siltstones,
north‐south trending normal fault‐bounded rocks of the limestones, and shales [Marquillas et al., 2005]. Early
Cretaceous Salta Rift. In this region basement anisotropies foreland basin strata (Paleogene) were deposited on a mixed
and abrupt changes in sedimentary thickness of Cretaceous substratum within a continuous regional basin that was later
rift strata profoundly affect the style and geometry of recent divided into smaller isolated basins by intrabasin range
contractional deformation (Figure 3) [Mon and Salfity, uplift in the late Miocene [Jordan and Alonso, 1987; Grier
1995; Carrera et al., 2006]. Seismic reflection studies and et al., 1991; Starck and Vergani, 1996; Bosio et al., 2009].
regional balanced cross sections suggest that the ranges of These strata are represented in the southern EC by four
the southwestern EC have been uplifted by the domino formations of the Eocene‐Pliocene Payogastilla group
rotation of semirigid basement blocks along spoon‐shaped [Grier and Dallmeyer, 1990]. In the study area, the Eocene‐
faults, with net shortening of 14–25% [Grier et al., 1991; Oligocene Quebrada de los Colorados Formation is com-
Cristallini et al., 1997; Mon and Salfity, 1995; Cristallini posed of approximately 350 m of coarse sandstone and
et al., 2004; Kley and Monaldi, 2002]. Major folds run conglomerate deposited in a braided fluvial to alluvial plain
parallel to fault trends in the southern EC, and synrift rocks system in a distal foredeep depocenter [Starck and Vergani,
are often thrust over postrift and Neogene foreland strata 1996]. The Miocene Angastaco Formation is composed of
[Grier et al., 1991; Carrera et al., 2006]. These high‐angle ∼4.2 km of proximal fluvial sandstones and conglomerates
reverse faults often have limited displacement, creating that generally coarsen and thicken upward [Starck and
broad to overturned anticlines in their hanging walls and Vergani, 1996]. The Miocene‐Pliocene Palo Pintado For-
synclines in their footwalls (Figure 5) [Grier et al., 1991; mation is composed of ∼2 km of cross‐stratified medium
Carrera et al., 2006]. Kinematic analysis of major and grained sandstones and mudstones, topped by ∼450 m of
minor structures has revealed a Mio‐Pliocene interval of mudrocks, channel sandstones, pebble conglomerates, and
WNW‐ESE shortening that created the major structures intercalated tuffs [Starck and Vergani, 1996]. The Pliocene
exposed at the surface, followed by a minor secondary stage San Felipe Formation is a ∼650 m thick succession of
of Pliocene‐Quaternary WSW‐ESE shortening [Grier et al., pebble to cobble conglomerates, medium‐grained sand-
1991; Marrett et al., 1994]. Kley and Monaldi [2002] note stones, and sparse green mudrocks [Starck and Vergani,
that narrow folds with shallow detachments are rare within 1996]. The reported thickness of the Neogene formations
this part of the EC, but major structures are much more should be considered minimums because unit thicknesses
closely spaced than in the nearby thick‐skinned Sierras vary along strike, likely thicken in the subsurface, and are
Pampeanas to the south [Ramos, 2002]. The Argentinean eroded and crosscut by the Calchaquí Fault to the east
EC represents a type of “fold‐and‐thrust” belt that is tran- (Figure 2).
sitional to the classic thin‐skinned and thick‐skinned models
[e.g., Kley et al., 1999], and therefore likely contains a
2.4. Chronology of Deformation in NW Argentina
hybrid of the structural and kinematic features commonly
found in these two deformation regimes. [15] During the Cretaceous in the area that is now the
southern EC and Santa Barbara System (Figure 2), several

5 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

intracratonic extensional basins of different orientations


radiated from a structural high in NW Argentina as part of
the Salta Rift system [Grier et al., 1991]. The southern arm
of the Salta rift is remarkably coincident with the location
of the modern southern Argentine EC, and can be divided
into several depocenters, including (from west to east) the
Brealito, Alemanía, and Metán subbasins (Figure 2)
[Marquillas et al., 2005]. These subbasins accumulated
more than 3.5 km of sediment from Neocomian to Maas-
trichtian time [Grier et al., 1991], and although Neogene
deformation has overprinted the intricate geometry of the
original rifts [Kley and Monaldi, 2002], extensional struc-
tures striking perpendicular to Andean stresses have been
preferentially reactivated as the major reverse fault bounded
ranges of the modern Argentinean EC during the Cenozoic
[Carrera et al., 2006; Mon and Salfity, 1995]. Deformation
within the Puna region and along its margins developed
since the Paleogene, and significant thick‐skinned defor-
mation has been active along the margins of the Puna
province since the middle Eocene [Bosio et al., 2009;
Coutand et al., 2006; Carrapa et al., 2005; Deeken et al.,
2006; Hongn et al., 2007; Riller et al., 2001]. Paleogene
sedimentation in the southern central Andes occurred in a
foreland basin that was regionally extensive, partially seg-
mented by inherited structural highs and geographically
complex [e.g., Carrapa et al., 2005; Carrapa and
DeCelles, 2008; Hongn et al., 2007]. Original paleo-
drainages within the Puna region were disrupted by local
range uplift starting at ca. 29–24 Ma [Carrapa et al.,
2005]. Apatite fission track thermochronology has docu-
mented that the regional foreland basin became succes-
sively compartmentalized by emergent topography as the
deformation front transferred eastward and into the EC
(Cumbres de Luracatao range) by 21 Ma (Figure 2)
[Deeken et al., 2006], as early as Eocene‐Oligocene time
[Coutand et al., 2001; Hongn et al., 2007].
[16] Hongn et al. [2007] use stratigraphic structural
relationships to argue for syntectonic deposition and growth
structure in the Eocene. If this is correct, deformation
migrated from the Plateau region to the EC almost instan-
taneously, which is a similar pattern to the one proposed for
the Altiplano in Bolivia [McQuarrie et al., 2005]. Alterna-
tively, the EC was the site of regional foredeep deposition in
the Eocene as suggested by sedimentological evidence
described by Starck and Vergani [1996] and DeCelles et al.
[2008]. In this scenario the deformation was within the
Plateau in the Eocene and reached the EC in the Miocene
(∼21 Ma) as recorded by extensive coarse grained sedi-
mentation in the foreland and Miocene exhumation of EC
basement rocks [Starck and Vergani, 1996; Coutand et al.,
2006; Deeken et al., 2006].
[17] Between ∼14 Ma and 4 Ma the deformation front was
in the Angastaco area and migrated into the Santa Barbara
System after ∼4 Ma. The Pucará, Angastaco, and La Viña
areas of the EC were connected in the early Cenozoic as part

Figure 4. General stratigraphy of the southern Eastern


Cordillera in the Calchaquí Valley and adjacent ranges
[Marquillas et al., 2005; Carrapa et al., submitted manu-
script, 2011] with approximate stratigraphic location of the
analyzed samples.

6 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

of a regional retroarc foreland basin that formed in response 2.5. Study Area: Calchaquí Valley
to subsidence and sediment influx from the rising Puna and [21] The Calchaquí Valley is an elongate depression
Cumbres de Luracatao range (EC) directly to the west located in the southwestern EC near the eastern margin of
(Figure 2) [Coutand et al., 2006]. the Puna Plateau (Figure 2). It is subdivided into the
[18] Features typical of growth strata, tilted unconformities, Angastaco and Tin‐Tin areas (or subbasins), and separates
and thermochronological data in the southern EC have sug- the Alemanía and Brealito subbasins of the Cretaceous Salta
gested a west‐to‐east and north‐to‐south migration of defor- rift (Figure 2) [Deeken et al., 2006]. The valley’s linear
mation across the EC and Santa Barbara System starting from shape is controlled in part by the long N‐S trending Cal-
Miocene time [Acocella et al., 2007; Carrera and Muñoz, chaquí fault on its eastern margin (Figure 2). The Cerro
2008; Deeken et al., 2006]. The northern tip of the Sierra Negro, Payogasta, and Tin‐Tin reverse faults in the northern
de Quilmes range (Figure 2) (referred to as Cerro Negro part of the Calchaquí valley expose synrift and postrift rocks
by Coutand et al. [2006]) began exhuming by 13–10 Ma, in their hanging walls, but there is no evidence to suggest
resulting in the removal of Pirgua and early Payogastilla Group that synrift sedimentary rocks exist under the southern
strata from the range and deposition into the Angastaco and part of Calchaquí valley, and in places foreland basin strata
Pucará areas (Figure 4) [Coutand et al., 2006]. From 14.5 Ma lie directly on crystalline basement (Figure 2) [Carrera
to 5.5 Ma, the EC has been the primary source of sediment and Muñoz, 2008; Mon and Salfity, 1995; G. Vergani and
for the Angastaco area, [Coutand et al., 2006]. Multiple apatite D. Starck, unpublished map, 1988]. Because of its excellent
fission track data sets support the exhumation of the Cerro exposures and relative accessibility, the Angastaco area of
Runno and Cerro Durazno range (Figure 2) during the interval the Calchaquí Valley is an ideal location to study the de-
of 13–10 Ma [Coutand et al., 2006; Deeken et al., 2006]. tails of intrabasin deformation that can occur during and
Paleodrainage reorganization in the Angastaco‐La Viña after fragmentation of the larger foreland basin system
areas suggest that the Angastaco basin was fully separated (Figure 2). The Angastaco area is part of a valley that
from the rest of the foreland and was in an intermontane continues ∼100 km north as the Calchaqui Valley and
position by ∼4 Ma (B. Carrapa et al., Cenozoic basin evo- south ∼130 km into the Santa Maria Valley; it is bounded
lution in the Eastern Cordillera of northwestern Argentina to the east by the inverted Calchaquí Fault, to the west by
(25°–26°S): Regional implications for Andean orogenic the Sierras Pampeanas style Sierra de Quilmes uplift,
wedge development, submitted to Basin Research, 2011). (Figure 2). The Angastaco area contains evidence of syn-
[19] As the sedimentary system was fragmented into a tectonic sedimentation, intrabasin faulting and folding, and
series of isolated basins, each of those depocenters became is characterized in the southern part by large drape folds
progressively deformed. Synclinal folding of the Pucará area and thick‐skinned basement uplifts [Carrera and Muñoz,
strata occurred during southward propagation of an adjacent 2008]. Many of these features are common in other parts
anticline (Figure 3), and was determined to be active of the southern EC, such as the Luracatao Valley, the Tin‐
between 13.4 ± 0.4 Ma and 12.1 ± 0.1 Ma by U‐Pb geo- Tin and Amblayo areas, and the Lerma Valley (Figure 2)
chronology on a welded tuff, lahar deposit, and lower [Vergani and Starck, 1988]. Detailed sampling of discrete
Payogastilla strata [Grier and Dallmeyer, 1990; Marrett structures for thermochronology in the Angastaco area can
et al., 1994]. However, the timing of deformation and thus provide cooling ages, and by inference the timing of
exhumation in the Angastaco area and Lerma Valley has so deformation and exhumation, in a variety of Cenozoic
far been based on indirect proxies (i.e., provenance, undated basins within the southern EC.
structures). Provenance and paleocurrent data indicate that
Cretaceous rocks of the adjacent Sierra de los Colorados and
Sierra de Leon Muerto (Figure 2) were uplifted and became
a dominant sediment source for the Angastaco area strata 3. Deformation in the Calchaqui Valley
by ∼4 Ma [Bywater‐Reyes et al., 2010]. Growth strata [22] In order to have a reliable stratigraphic and structural
relationships suggest that the Cumbres de Peñas Blancas frame for our sampling field mapping of geologic structures
range east of the Lerma Valley (Figure 2) was uplifted and within the Angastaco area was conducted using 1:30,000
exhumed during the past 5 Ma [Carrera and Muñoz, 2008]. scale stereoscopic aerial photographs and form the basis of
After deformation swept across the southern EC during the Figure 5. In particular, detailed mapping was conducted to
Miocene and Pliocene, out‐of‐sequence deformation and identify possible growth in the Angastaco Formation, which
minor strike‐slip movement occurred throughout the EC records syndepositional deformation and thus the age of
during the Pleistocene and Holocene [Carrera and Muñoz, strata can be used to determine the timing of deformation
2008; Marrett et al., 1994]. within the basin.
[20] Further south the Sierras Pampeanas, including areas [23] Evidence of synorogenic sedimentation within the
all along the southeastern margin of the Puna Plateau, Angastaco area was observed in one location (Figure 6). To
experienced a transition from regional foreland to com- identify growth, we were looking for truncation, thinning
partmentalized basin at ∼6 Ma [Mortimer et al., 2007; or thickening of beds across strike. A 24° angular uncon-
Carrapa et al., 2008a]. The most recent stratigraphic record formity was found near the town of Angastaco in the lower
in the region is represented by the flat lying Pleistocene lake fluvial Angastaco Formation (Figure 6), with a well‐
deposits in the Calchaquí Valley near the town of Angas- developed paleosol interval along the unconformity (∼1 m
taco, which lie in angular unconformity atop the deformed thick). Detrital U‐Pb geochronology on a sandstone sample
Payogastilla group and document the end of deformation immediately below the unconformity constrains the maxi-
[Salfity et al., 2004]. mum depositional age of these strata to be 14.1 ± 0.5 Ma
(see section 4.3). Continued investigation of the Angastaco

7 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Figure 5. Geologic map of the Angastaco area in the Calchaquí Valley. Location of this map is outlined
as a black box in Figure 2.

8 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

LaserChron center [Gehrels et al., 2008]. For a complete


description of lab methods, consult the appendix to this
paper.
[26] There are several viable methods by which a maxi-
mum depositional age can be calculated from a young
population of detrital zircon ages. Youngest single grain age
has been shown to be compatible with depositional age 90%
of the time, but confident interpretation of young grain age
clusters can only be achieved with a weighted mean of
young ages that overlap within error [Dickinson and
Gehrels, 2009]. We present here mean ages of the youn-
gest two or more grains that overlap within 1s error for
sandstone samples, and for intercalated ash samples the
mean ages were calculated using the Tuffzirc program of
Ludwig [2008].
Figure 6. Photo of the angular unconformity (24° angular 4.2. Detrital U‐Pb Sampling
discordance) in the lower fluvial Angastaco Formation at the
northern edge of the town of Angastaco. A paleosol horizon [27] In order to discern the age of the uppermost San
is well developed along the unconformity. Note person for Felipe Formation, sample 04.14.08‐02 was collected in a
scale. partially reworked ash near the top of the formation in a
small syncline (Figure 2). Because this syncline is crosscut
by the Calchaquí Fault immediately to the east (Figure 3),
the youngest detrital zircon age component constrains the
Formation yielded no other conclusive evidence for timing of recent movement on the fault as well. In order to
growth. discern the age of the angular unconformity in the Angas-
[24] Eastward shallowing of bedding within the lower/ taco Formation near the town of Angastaco previously in
middle Angastaco Formation near the contact with the this paper (Figure 6), two detrital sandstone samples were
basement is suggestive of basin‐scale growth or can alter- collected. Sample 05.07.08‐B.Unc was collected one
natively be attributed to the synclinal flexure of the basin meter below the base of the angular unconformity shown in
(Figure 2). The contacts between the Angastaco/Palo Pin- Figure 6, and sample 05.07.09‐A.Unc was collected one
tado and Palo Pintado/San Felipe Formations were observed meter above the same unconformity. In order to better define
to be conformable at every observed location within the the age of the Angastaco Formation in the Pucará area, a
study area. However, the intense folding and possible boulder of andesite (sample 04.27.08‐AngB) was collected
imbrication of the sedimentary section could easily obscure within an alluvial fan conglomeratic interval of the lower
the shallow fanning of bedding that would provide more Angastaco Formation in the Pucará syncline (Figure 2).
conclusive evidence of large‐scale growth. The angular
unconformity in the lower Angastaco Formation represents 4.3. Detrital U‐Pb Results
the earliest evidence of syndepositional deformation in the [28] Of the four detrital zircon samples analyzed (Table A1),
area and is interpreted here to have formed by movement only samples 05.07.08‐B.Unc and 04.14.08‐02 had a young
along the high‐angle Sierra de Quilmes and Cerro Negro population of ages capable of constraining the maximum
faults to the west (Figure 2) right after ∼14 Ma. depositional age of the Angastaco and San Felipe Forma-
tions. For sample 04.14.08‐02, 40 detrital zircons were
4. Detrital U‐Pb Geochronology analyzed and 34 grains yielded concordant ages (Figure 7a).
Zircon ages are distributed in four clusters, with major
4.1. Detrital U‐Pb Methods peaks around 1–9 Ma, 460–590 Ma, 1.0–1.1 Ga, and one
[25] Detrital zircon U‐Pb analysis has proven a reliable old grain at 1.9 Ga. A young population of 14 grains yielded
technique for determining provenance [Gray and Zeitler, ages younger than 3 Ma. The Tuffzirc routine of Isoplot
1997] and maximum depositional age of sedimentary [Ludwig, 2008] was used to calculate a mean age of 2.3 ±
rocks [DeCelles et al., 2007; Dickinson and Gehrels, 2009]. 0.1 Ma from a coherent group of six detrital zircons, and the
Four detrital sandstone samples from the Cenozoic basin fill youngest single grain age is 1.8 ± 0.1 Ma (Figure 7a). We
preserved in the Angastaco and Pucará areas have been interpret this age to represent the formation and deposition
analyzed in this study to better constrain the maximum of volcanogenic zircons during the deposition of the
depositional age of the lower Angastaco and upper San uppermost San Felipe Formation at ∼2.3 Ma. Because the
Felipe Formations in the region where ashes were not formation here is crosscut by a major fault, this age also
present (Figure 2). In all samples, elongate euhedral zircon demonstrates that the Calchaquí fault has experienced some
crystals were preferentially analyzed to increase the likeli- reverse offset since ∼2.3 Ma. The Paleozoic and Precam-
hood of documenting a young population of ashfall zircons, brian ages are typical of first cycle Paleozoic granites and
and therefore a statistically significant provenance analysis detrital ages in the Puncoviscana Fm.
cannot be made from this data set. All analyses were con- [29] For sample 05.07.08‐A.Unc, 115 detrital zircons
ducted by laser ablation multicollector inductively coupled were analyzed and 102 grains yielded concordant ages
plasma mass spectrometry at the University of Arizona (Figure 7b). Zircon ages show a major Paleozoic age cluster

9 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

at 460–540 Ma, with other zircons as old as 1.1 Ga.


Unfortunately, no zircons yielded Cenozoic ages capable of
constraining maximum depositional age of the Angastaco
Formation.
[30] For sample 05.07.08‐B.Unc, 100 detrital zircons
were analyzed and 86 grains yielded concordant ages
(Figure 7c). Zircon ages are distributed in four clusters,
with peaks around 14 Ma, 468–690 Ma, 0.9–1.1 Ga, and
1.8–2.1 Ga. Two detrital grains yielded Cenozoic ages that
overlap within 1s error, and give a maximum depositional
age (weighted mean) of 14.1 ± 0.5 Ma, with a youngest
single grain age of 13.7 ± 1.1 Ma. These young zircons
constrain growth in the lowermost Angastaco Formation to
have begun after ∼14.1 Ma. For sample 04.27.08‐AngB,
mineral separation yielded only a small zircon fraction. For
this reason only twenty zircons were analyzed, 14 of which
yielded concordant ages, and all zircons gave Precambrian
to Silurian ages (∼425–568 Ma) (Figure 7d).
[31] Although our decision to preferentially analyze
euhedral zircons prevents an unbias statistical analysis of
zircon populations to be conducted for provenance pur-
poses, it is interesting to note that zircon ages from the lower
Angastaco and upper San Felipe Formations cluster around
three similar intervals of time in the early Paleozoic at
∼450–600 Ma, in the middle Proterozoic at ∼950–1150 Ma,
and in the early Proterozoic at ∼1.9 Ga (Figure 7). The 450–
600 Ma cluster is likely due to the dominance of zircon
sourcing from adjacent outcrops of the Puncoviscana For-
mation, and the presence of similar clustering in the lower
Angastaco and upper San Felipe Formations suggests that
there was no significant change in basement provenance
during this interval.

5. Apatite U‐Th/He Thermochronology


5.1. AHe System
[32] Apatite (U‐Th)/He thermochronometry (AHe) con-
strains the timing and magnitude of cooling through the
40°–80°C temperature window, a range that defines the
AHe partial retention zone (HePRZ) [Stockli et al., 2000;
Wolf et al., 1998]. This technique provides information on
shallow crustal processes, responds quickly to abrupt
changes in erosion rate, and is therefore especially useful
when investigating the recent history of range exhumation
and basin incision in tectonically active regions [e.g.,
Reiners and Brandon, 2006]. The AHe dating system is
based upon on the principle that radiogenic production of a
particles by 238U, 235U, 232Th, and 147Sm introduces 4He
into the crystal lattice of apatite at a predictable rate over
geologic time [Farley, 2002]. Because 4He is retained in
apatite at earth surface temperatures, but systematically

Figure 7. (a‐d) Probability density functions of zircon U‐


Pb analyses from four detrital samples. Blue boxes represent
the number of zircon ages falling into 50 My bins, with
probability density function overlain in red. Insets in boxes
Figures 7a and 7c display the youngest population of ages
from each sample. Inset of Figure 7a was calculated with
the Isoplot routine of Tuffzirc [Ludwig, 2008]. Inset of
Figure 7c displays the two youngest ages that overlap within
1s error. Sample locations are shown in Figure 2.

10 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

diffuses out of apatite at shallow crustal temperatures, the conglomerate, approximately three meters east of the Cal-
AHe system can be used as a precise low‐temperature chaquí Fault in the overturned west limb of the same anti-
thermochronometer sensitive to as little as 2–3 km of cline. Because sample 04.29.08‐01 was collected in the core
exhumation [Ehlers and Farley, 2003; Farley, 2002; Stockli of the anticline, it lies stratigraphically ∼400 m below
et al., 2000; Zeitler et al., 1987] and >4 km in case of low sample 04.14.08‐01.
geothermal gradients (<20°C). The AHe closure temperature 5.2.4. Rio de Las Conchas Sample Locations
is well constrained to be 75 ± 5°C in Durango apatite [Wolf [37] Two samples were collected along the Rio de las
et al., 1996; Ehlers and Farley, 2003], but can vary sig- Conchas, and represent two different fault‐bounded seg-
nificantly depending on grain size, crystal morphology, ments of the EC. Sample LY is from the lowermost part of
cooling rate, and presence of radiation damage [Shuster the Pirgua Subgroup in the La Yesera Formation. It was
et al., 2006; Ehlers and Farley, 2003; Wolf et al., 1996]. collected in the hanging wall of the southernmost tip of the
Calchaquí Fault, also known as the Zorrito Fault [Mon and
5.2. AHe Sampling Salfity, 1995], at the south tip of the Sierra de Leon Muerto
[33] Ten detrital sandstone and conglomerate rock sam- where Cretaceous rift rocks are thrust over Cenozoic sedi-
ples were collected in the deepest exposed strata of five mentary rocks (Figure 2). Sample ALP is from the upper
fault‐bounded segments of the southern EC (Figure 2 and Pirgua Subgroup in the Los Blanquitos Formation, and was
Table 1). The fault‐bounded segments were sampled along a collected in the hanging wall of the La Viña Fault near the
roughly east‐west oriented transect in the Pucará Valley, the town of Alemanía along the southwestern margin of the
Angastaco area of the Calchaquí Valley, the Sierra de los Lerma Valley (Figure 2).
Colorados anticline, the Rio de las Conchas, and off transect 5.2.5. Tin‐Tin Sample Locations
in the Tin‐Tin area of the Calchaquí Valley (Figure 2). Two [38] Two samples were collected in the northern Calcha-
samples were collected from fault‐bounded segments of the quí Valley southeast of the town of Cachi. Sample
EC, in multiple formations at each location where possible, QCTin250, was collected in the Quebrada de los Colorados
and are therefore described in the following as sample pairs. Formation in the northern Calchaquí Valley in the hanging
All analyses were conducted using the ultra high vacuum wall of the Payogasta Fault (Figure 2). Sample 1TT18FT
noble gas extraction and purification line and ICP‐MS at the was collected in the Santa Barbara Subgroup at the northeast
University of Kansas (U‐Th)/He laboratory, following corner tip of Cerro Tin‐Tin in the hanging wall of the Tin‐
analytical procedures of Stockli et al. [2000] and Biswas Tin fault (Figure 2).
et al. [2007]. For complete AHe laboratory methods, con-
sult the appendix to this paper. 5.3. AHe Results
5.2.1. Pucará Valley Sample Locations [39] Fifty individual grain analyses from ten individual
[34] The Pucará valley consists of a large N‐S trending rock samples show a scatter of raw AHe dates that sys-
syncline that contains Cretaceous through Neogene sedi- tematically vary with the effective uranium concentration of
mentary rocks. Two samples were collected on the west the grain (Table 1) (eU = U + 0.235Th) [Shuster et al.,
limb of the syncline, in the deepest parts of the Cenozoic 2006]. For detrital grains, the presence of a strong age‐eU
strata. Sample QCPucara1 was collected near the base of (Figure 8) relationship implies an episode of partial He loss
the Quebrada de los Colorados Formation, and Sample [Flowers et al., 2007]. Although these detrital samples all
AngPuc06 was collected at the base of the Angastaco For- come from different parts of the EC, geologic constraints
mation just above the contact that separates these two for- suggest that they were all buried to depths greater than 3 km
mations (Figure 3). and subsequently exhumed as indicated by >6 km of
5.2.2. Angastaco Area Sample Locations Cenozoic section that sits stratigraphically above the col-
[35] The Angastaco area is a structurally complex segment lected samples [Uliana et al., 1989; Coutand et al., 2006].
of the Calchaquí Valley that contains Oligocene‐Pliocene Assuming a standard geothermal gradient and standard
synorogenic sedimentary rocks. Two samples were collected Durango apatite diffusion kinetics, all these grains would
on the westernmost part of the valley, just south of the town have been thermally reset during burial, and therefore AHe
of Angastaco, in the footwall of a large reverse fault that ages should represent cooling following basin incision or
bounds the northeastern tip of the Sierra de Quilmes range. exhumation, presumably related to deformation.
Sample QC1EB080 was collected in the upper Quebrada de [40] However, most samples have two groups of grain
los Colorados Formation, and sample Ang2EB was col- ages, including a scatter of dates older than depositional
lected just above in the lower eolian facies of the Angastaco age and a cluster of dates younger than depositional age
Formation (Figure 2) (B. Carrapa et al., submitted manu- (Figure 8 and Table 1). AHe dates that are older than
script, 2011). depositional age (DA) of the formation in which they were
5.2.3. Sierra de los Colorados Sample Locations collected routinely correlate with high [eU], suggesting that
[36] The Sierra de los Colorados is a broad, west vergent higher radiation damage increased He retentivity and raised
anticline developed in the hanging wall of the Calchaquí the effective closure temperature (Tec) in these grains,
Fault (Figure 3). The ridge is composed of Pirgua Subgroup allowing only partial resetting, and resulting in older AHe
rift sediments, and is dissected in several places by ante- dates. In contrast, the young clusters of dates in many
cedent drainage canyons. Sample 04.29.08‐01 was collected samples correspond to low [eU], indicating that these
in the core of the anticline where it is cut by a narrow grains have less radiation damage. The clustering of these
canyon, in a conglomeratic interval of the Los Blanquitos young populations suggests that they were more thermally
Formation within the Pirgua Subgroup (Figure 2). Sample reset during burial. Thus, the older, partially annealed
04.14.08‐01 was also collected in upper Pirgua Subgroup populations can be momentarily disregarded, and the

11 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Table 1. Detrital Apatite (U‐Th)/He Data


4
Mass U Th eU Sm He Length Width rb Raw Date Est 1s Corr Date Est 1s Latitude
(mg) (ppm) (ppm) (ppm) (ppm) (nmol/g) (mm) (mm) Fta (mm) (Ma) (Ma) (Ma) (Ma) Longitude
Pucará Valley Samples
AngPuc06, base of Angastaco
Fm., eolian sandstone,
deposited since 21.4 ±
0.8 Mac
a1 2.0 2.7 11.6 5.5 16.4 0.65 136 86 0.65 44 8.5 0.5 13.2 0.8 25.794°S,
66.282°W
a2 1.5 13.8 25.2 19.7 109.5 0.61 145 72 0.61 39 47.1 2.8 77.0 4.6
a3 1.5 62.9 25.8 68.9 17.6 0.64 121 79 0.64 40 36.2 2.2 56.4 3.4
a4 2.7 6.6 5.4 7.8 18.2 0.70 136 100 0.70 49 24.6 1.5 35.3 2.1
a5 2.0 2.4 9.7 4.7 23.4 0.65 125 89 0.65 44 8.6 0.5 13.3 0.8
a6 1.2 29.3 9.1 31.4 94.1 0.61 127 69 0.61 36 38.3 2.3 62.7 3.8
QCPucara1, base of Quebrada
de Los Colorados Fm, sandstone,
deposited ∼37.6 ± 1.2 Mac
a1 1.4 30.6 9.7 32.9 37.0 0.64 114 79 0.64 40 31.7 1.9 49.7 3.0 25.784°S,
66.291°W
a2 1.4 5.5 11.6 8.3 59.7 0.60 156 67 0.60 37 22.1 1.3 37.2 2.2
a3 1.5 16.2 20.6 21.0 54.8 0.62 132 76 0.62 39 20.1 1.2 32.3 1.9
a4 1.1 19.4 25.7 25.4 55.1 0.56 150 60 0.56 33 26.8 1.6 47.4 2.8
a5 1.1 8.3 17.1 12.3 19.5 0.59 109 72 0.59 36 15.8 0.9 26.7 1.6

Angastaco Area Samples


Ang2EB, base of Angastaco
Fm., eolian sandstone,
deposited since
19.2 ± 0.7 Mac
a2 2.0 26.9 7.4 28.6 60.3 7.77 146 83 0.67 43 49.4 3.0 73.9 4.4 25.714°S,
66.175°W
a3 2.0 15.2 18.8 19.6 53.9 5.89 145 83 0.65 43 54.3 3.3 83.0 5.0
a4 1.1 2.9 21.8 8.1 66.7 0.27 102 71 0.57 36 5.7 0.3 9.9 0.6
a5 1.0 2.3 17.3 6.3 44.2 0.22 99 72 0.57 36 6.1 0.4 10.6 0.6
QC1EB080, top of Quebrada
de los Colorados Fm.,
sandstone, deposited
since 37.6 ± 1.2 Mac
a1 11.5 6.1 40.1 15.6 110.0 0.59 232 157 0.79 79 6.6 0.4 8.4 0.5 25.715°S,
66.178°W
a2 10.4 4.2 32.0 11.7 56.9 0.76 222 152 0.79 76 11.5 0.7 14.6 0.9
a3 10.1 25.7 6.2 27.1 83.7 7.85 231 147 0.80 75 52.1 3.1 65.1 3.9
a4 13.2 57.9 6.9 59.6 77.8 23.68 247 163 0.82 82 72.6 4.4 88.6 5.3
a5 9.6 6.0 4.0 7.0 41.8 1.51 172 167 0.80 76 38.2 2.3 47.9 2.9
a6 6.4 26.0 3.5 26.9 61.8 7.82 160 141 0.78 66 52.8 3.2 67.8 4.1
a7 8.7 4.9 36.5 13.5 80.9 0.71 186 152 0.78 73 9.3 0.6 11.9 0.7

Sierra de los Colorados Samples (Alemania Subbasin)


04.29.08‐01, Pirgua Subgroup,
conglomerate, core of anticline,
deposited 128–70 Mad
a1 0.7 9.4 20.4 14.3 18.6 0.24 83 65 0.54 32 3.1 0.2 5.7 0.3 25.624°S,
65.991°W
a2 0.6 11.8 44.6 22.3 56.5 0.55 70 65 0.51 30 4.5 0.3 8.7 0.5
a3 0.9 7.9 26.2 14.1 97.4 0.46 80 76 0.57 35 5.7 0.3 9.9 0.6
a4 0.8 20.5 63.1 35.4 63.6 0.98 92 64 0.54 32 5.1 0.3 9.4 0.6
04.14.08‐01, Pirgua Subgroup,
conglomerate, east of Calchaqui
Fault, deposited 128–70 Mad
a1 2.2 9.6 42.8 19.6 127.2 1.98 113 99 0.66 46 17.7 1.1 26.6 1.6 25.725°S,
65.967°W
a2 1.3 20.6 3.7 21.5 41.2 5.35 176 60 0.59 34 45.3 2.7 76.2 4.6
a3 0.4 17.7 61.1 32.0 163.3 0.21 69 56 0.47 27 1.1 0.1 2.4 0.1
a4 1.6 8.8 23.5 14.3 18.7 0.30 134 76 0.62 40 3.8 0.2 6.1 0.4
a5 2.3 9.9 3.8 10.8 14.2 0.99 144 88 0.68 45 16.8 1.0 24.8 1.5
a6 0.9 2.8 12.9 5.9 33.3 0.38 98 66 0.55 33 11.5 0.7 21.0 1.3

12 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Table 1. (continued)
4
Mass U Th eU Sm He Length Width rb Raw Date Est 1s Corr Date Est 1s Latitude
(mg) (ppm) (ppm) (ppm) (ppm) (nmol/g) (mm) (mm) Fta (mm) (Ma) (Ma) (Ma) (Ma) Longitude
Rio de las Conchas Samples
LY, La Yesera Fm, sandstone,
hangingwall of Zorrito Fault,
deposited 128–90 Mad
a3 1.0 30.7 175.0 71.8 480.1 1.4 89 74 0.57 35 3.4 0.2 6.0 0.4 25.962°S,
65.751°W
a4 1.4 5.5 15.0 9.1 95.6 0.1 122 75 0.61 38 2.6 0.2 4.3 0.3
a5 1.4 33.8 183.5 76.9 321.1 1.2 147 69 0.59 37 2.8 0.2 4.7 0.3
a6 0.7 7.0 35.9 15.5 166.7 0.2 86 62 0.52 31 1.7 0.1 3.4 0.2
ALP, Los Blanquitos Fm., sandstone,
hangingwall of La Viña Fault,
deposited 76–70 Mad
a1 2.5 60.4 7.9 62.3 69.0 1.1 179 83 0.68 45 3.1 0.6 4.5 0.8 25.632°S,
65.629°W
a2 0.9 37.3 14.5 40.8 60.9 0.5 112 63 0.57 33 2.4 0.1 4.2 0.3
a3 3.0 15.4 5.7 16.8 70.0 0.3 154 99 0.71 50 2.8 0.2 4.0 0.2
a4 10.0 0.7 1.9 1.1 1.2 0.0 220 150 0.79 75 2.3 0.1 3.0 0.2
a5 1.7 60.2 4.3 61.2 54.7 1.1 121 84 0.67 42 3.2 0.2 4.8 0.3
a6 4.4 55.1 4.0 56.0 88.1 1.3 150 120 0.75 58 4.3 0.3 5.7 0.3

Tin‐Tin Samples
QCTin250, Quebrada de los
Colorados Fm, sandstone,
deposited since ∼37.6 ± 1.2 Mad
a1 8.8 16.0 7.6 17.8 29.0 4.7 223 140 0.79 72 48.8 2.9 61.9 3.7 25.254°S,
66.120°W
a2 4.7 15.2 12.8 18.2 22.7 2.0 206 107 0.73 57 20.2 1.2 27.7 1.7
a3 4.0 30.2 5.4 31.5 61.5 9.0 160 112 0.74 56 51.9 3.1 70.4 4.2
a4 3.8 5.4 8.3 7.3 7.6 0.8 169 106 0.71 54 19.5 1.2 27.3 1.6
1TT18FT, Pirgua Subgroup,
sandstone, deposited 128–70 Mad
a1 1.0 57.0 116.0 84.2 106.5 3.1 114 65 0.56 34 6.7 0.4 11.9 0.7 25.113°S,
66.005°W
a2 1.0 63.0 236.0 118.5 82.7 2.1 85 75 0.57 35 3.3 0.2 5.7 0.3
a3 0.9 65.3 250.9 124.2 97.7 36.5 92 71 0.56 34 53.8 3.2 95.4 5.7
a4 1.2 71.6 158.9 109.0 100.9 2.6 106 76 0.60 38 4.4 0.3 7.4 0.4
a
Ft is alpha ejection correction.
b
r is spherical radius.
c
Max depositional age based on detrital U‐Pb geochronology (B. Carrapa et al., submitted manuscript, 2011).
d
Salfity and Marquillas [1994].

young clustered populations of dates can be interpreted to AngPuc06. We interpret the two youngest AngPuc06 grain
represent the most recent signal of exhumation (Figure 9). dates to represent exhumation of the Pucará area because
Overall, there is a remarkable eastward younging trend they have comparable ages, show the lowest eU of any
among the young populations of AHe ages. Raw AHe apatites from the Pucará area, and are consistent with other
dates systematically young eastward from ∼13 Ma in the constraints on basin folding at ∼13.4–12.1 Ma [Marrett
Pucará area to ∼4 Ma in the La Viña area, suggesting an et al., 1994]. Although we recognize that model results for
eastward younging wave of exhumation that propagated sample QCPucara1 are equivocal, the fact that this age
across the southern EC during the Mio‐Pliocene at a rate match well with evidence of growth in the adjacent
of ∼8.3 mm/yr (Figure 9). Angastaco area at ∼14 Ma supports our interpretation that
5.3.1. Pucará Valley Sample Results this part of the basin was exhuming at this time possibly as a
[41] Six apatites were analyzed from sample AngPuc06. result of tilting and uplift of the east verging limb of the
Two of the six grains were younger than the DA of ∼21.4 Ma Sierra de Quilmes hanging wall (Figure 3).
(B. Carrapa et al., submitted manuscript, 2011), with com- 5.3.2. Angastaco Area Sample Results
parable cooling ages of 13.2 ± 0.8 and 13.3 ± 0.8 Ma (Table 1) [43] Four apatites were analyzed from sample Ang2EB
and a weighted mean age of 13.3 ± 1.1 Ma. Five apatites were (Table 1). Two of the four grain ages are younger than
analyzed from sample QCPucara1. Three of the five grains DA, and have similar grain ages of 9.9 ± 0.6 Ma and 10.6 ±
were younger than DA, and have disparate grain ages of ca. 0.6 Ma. Seven apatites were analyzed from sample
26.7 ± 1.6 Ma, 32.3 ± 1.9 Ma, and 37.2 ± 2.2 Ma. QC1EB080. Three of the seven grain ages are younger than
[42] QCPucara1 is stratigraphically lower than AngPuc06 DA, with dates of 8.4 ± 0.5 Ma, 11.9 ± 0.7 Ma, and 14.6 ±
(Figure 4), and therefore should have been heated to a 0.9 Ma. Young AHe grain ages from these two samples
greater temperature than AngPuc06 during burial, but correlate strongly with eU, suggesting a basin history where
QCPucara1 shows AHe dates considerably older than the lower Angastaco and upper Quebrada de los Colorados

13 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Figure 8. (U‐Th)/He ages versus effective U concentration for the analyzed samples. Green line repre-
sents depositional age.

Formations spent a significant amount of time within the yield significantly older ages, which cluster around 21.0 ±
HePRZ. Additionally, because these two samples are stra- 1.3 to 26.6 ± 1.6 Ma. The grain yielding an age of 2.4 ±
tigraphically close to one another, it can be assumed that 0.1 Ma is the smallest crystal of the entire study, with crystal
they underwent similar thermal histories. Whereas the oldest dimensions of 56 mm × 69 mm; this is likely the cause for its
group of ages (Table 1) can be explained by partial an- anomalously young AHe date. Both of these samples were
nealing of different closure temperatures depending on dif- collected within the same structural, and topographic feature
ferent compositions, the weighted mean of the youngest five (the Sierra de los Colorados anticline), however sample
grain ages is used to approximate the timing of exhumation 04.29.08‐01 despite being farther from the fault, with
at 10.4 ± 2.5 Ma. respect to sample 04.14.08‐01, it is from a deeper strati-
5.3.3. Sierra de los Colorados Sample Results graphic level (Figure 2). The emplacement of a relatively
[44] Four apatites were analyzed from sample 04.29.08‐ warm hanging wall over a cool footwall causes thermal
01, all of which are significantly younger than DA, with perturbations that depress isotherms and reduce lateral
ages between 5.7 ± 0.3 Ma and 9.4 ± 0.6 Ma. Six grains thermal homogeneity, and there is some evidence that fric-
were analyzed from sample 04.14.08‐01. Five of the six tional heating along the fault plane can locally affect ther-
grains give ages younger than DA, and cluster into two mochronological systems [Ruppel and Hodges, 1994;
distinct populations. The two youngest ages of 2.4 ± 0.1 Ma Ehlers and Farley, 2003]. This could explain the wide
and 6.1 ± 0.4 Ma generally correlate with the young exhu- spread of AHe dates in sample 04.14.08‐01. Because sam-
mation signal of sample 04.29.08‐01, but three other grains ple 04.29.08‐01 was strategically collected within the dee-

14 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Figure 9. Young populations of detrital apatite (U‐Th)/He grain ages plotted against sample longitude.
Errors are 1s. Dashed line represents linear best fit (r2 = 0.7). Gray halo represents the time range of sam-
ple good cooling paths through the HePRZ as modeled in HeFTy.

pest stratigraphic interval of the anticline, it was likely 2009]. Sandstone samples in this study show a relationship
subjected to greater temperatures, and it is not surprising between grain ages and [eU], making them prime candidates
that it represents a simpler, less altered, and more thermally for inverse modeling using the radiation damage trapping
reset signal of cooling and exhumation. model of Shuster et al. [2006]. Radiation damage impedes the
5.3.4. Rio de Las Conchas Sample Results diffusion of 4He from the apatite crystal lattice, and therefore
[45] Four apatites were analyzed from sample LY, all diffusion coefficient (D) of an apatite varies between grains
of which are younger than DA, with dates between 3.4 ± and evolves within grains as a function of time. In a series of
0.2 Ma and 6.0 ± 0.4 Ma. Six apatites were analyzed from laboratory diffusion experiments, Shuster et al. [2006]
sample ALP, all of which are younger than DA, with ages empirically calculated the relationship between the diffu-
between 3.0 ± 0.2 Ma and 5.7 ± 0.3 Ma. The fact that grain sion coefficient (D) and the volume of radiation damage (Vrd)
ages are significantly younger than DA suggest that both within the grain, and the radiation damage trapping model
samples were completely thermally reset and that all these exploits this laboratory constrained age‐Vrd relationship to
ages represent a signal of recent cooling. greatly restrict the number of possible time‐temperature (t‐T)
5.3.5. Tin‐Tin Sample Results paths that could reproduce the AHe dates observed in nature.
[46] Four apatites were analyzed from sample QCTin250. Inverse modeling of AHe grain ages predicts t‐T paths that
Two of the four grain ages are younger than DA, with would produce the observed relationship between Tc and
similar ages of 27.3 ± 1.6 Ma and 27.7 ± 1.7 Ma. Four radiation damage. Numerical inverse modeling can thus help
apatites were analyzed from sample 1TT18FT. Three of determine the detailed thermal history of a sample.
the four grain ages are younger than DA, with ages of
5.7 ± 0.3 Ma, 7.4 ± 0.4 Ma, and 11.9 ± 0.7 Ma. The 5.5. HeFTy Modeling Constraints
young 6–12 Ma signal of sample 1TT18FT can thus be [48] Thermal models were run for multiple grain ages from
interpreted to represent recent main exhumation of the six AHe samples (Table 2). In all cases, He models were run
northern Calchaquí Valley. using the radiation damage trapping model calibration of
Shuster et al. [2006] (Do/a2) and the static alpha ejection
5.4. Modeling of AHe Data correction of Farley et al. [1996]. Gradual node randomi-
[47] Sample thermal history information is extracted employ- zation was chosen to reflect steady burial, and episodic node
ing an inverse modeling approach using a constrained Monte randomization was used to represent exhumation accom-
Carlo simulations that takes into account the effects of grain modated by potentially erratic fault movements. For each
size, radiation damage, and cooling rate on the thermal his- sample, wide model constraints based on geological plausi-
tory of each sample (HeFTy) [Ketcham, 2005; Flowers et al., ble scenarios were used as input for deposition, burial, and

15 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

exhumation so to allow maximum modeling freedom (see


Table 2). The model was allowed to run until 100 good fits
were found. If no good fits could be found, then the model
was allowed to run until 100 acceptable fits were found. All
samples were started at an age at least double the deposi-
tional age of the given sample; only the most recent, relevant
thermal history is shown in Figure 10. Known constraints on
deposition of the sample were input as a 20° temperature
boundary condition at the age of deposition. Constraints on
burial were input from the earliest known depositional age to
present, with minimum burial temperatures beginning at
30°–40°C. Depending on a sample’s grain age distributions,
we used one of two different values for maximum burial
temperature. For samples that had a group of grain ages older
than DA and a group younger than DA, we assumed that the
sample was never heated significantly above the HePRZ, and
imposed a maximum burial temperature of 100°C. For
samples that contained a single, well defined, group of grain
ages younger than DA, we assumed that the samples had
been heated well above the HePRZ, were thermally reset,
and therefore allowed a maximum model temperature of
200°C. Because the apatite system does not retain any He at
high temperatures, we did not find it necessary to impose
maximum temperatures above 200°C for any sample. A final
constraint on present‐day temperature was placed at 15°–
22°C for 0 Ma, which is consistent with average surface
temperatures. A complete list of modeling parameters can be
found in Table 2.
[49] HeFTy can accept simultaneous input from up to five
apatite aliquots, and because the apatite has a different Tec,
the model simulates thermal histories that are allowed by all
the grains, thereby constraining more time‐temperature
paths much more than would be possible with any single
grain age. Grain ages older than DA represent a partially
reset and complex T‐t signal and therefore they constrain
neither source exhumation nor basin exhumation, but
instead a combination of the two. Attempts were made to
model as many grains per sample as possible, but HeFTy is
often unable to find any good model fits when constrained
by three or more grains (especially when grains with ages
older than DA were added), even when allowed to simulate
>200,000 paths. For these reasons only grains younger than
DA were used for modeling purposes and for each of the six
successful models presented here, two to four representative
grains were used. We refer to Table 2 and Figure A1 for a
complete list and results of the grains modeled.

5.6. HeFTy Modeling Results


[50] All model results allow relatively wide boundaries on
timing of burial, but show strong, well‐constrained paths of

Figure 10. Time‐temperature paths of six AHe samples


modeled using the HeFTy computer program. The most
recent 20 Myr are illustrated here. Purple areas represent
halos around all “good” paths modeled, and green areas rep-
resent halos around all “acceptable” paths modeled. All
modeling was conducted using the radiation trapping model
of Shuster et al. [2006] (Do/a2) and the static alpha ejection
correction of Farley et al. [1996]. Complete model para-
meters can be found in Table 2. Sample locations are indi-
cated on Figures 2, 3, and 4.

16 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Table 2. HeFTy Modeling Constraints


Deposition Burial Present Grains Total Acceptable Good
Sample Age (Ma) T (°C) Age (Ma) T (°C) Age (Ma) T (°C) Modeled Paths Paths Paths
AngPuc06 18–21.4 20 0–21.4 30–100 0 10–25 a1, a5 2,003 168 100
Ang2EB 14–19.2 20 0–19.2 40–100 0 10–25 a4, a5 36,892 2,871 100
QC1EB080 19.2–37.6 20 0–19.2 40–100 0 10–25 a1, a7 308,236 1,500 0
04.29.08‐01 70–128 20 0–128 40–200 0 10–25 a2, a3, a4 120,547 1,381 100
LY 90–128 20 0–128 40–200 0 10–25 a4, a5 34,514 357 100
ALP 70–76 20 0–76 40–200 0 10–25 a1, a2, a3, a5 27,180 237 100

most recent cooling and exhumation (Figure 10). Although Angastaco, and Tin‐Tin samples), it can be confidently
thermal models cannot determine the exact path of exhu- interpreted that the samples were not completely buried
mation, they can greatly constrain a window of time and beneath the HePRZ, and therefore young populations of AHe
temperature when the sample began cooling through the dates closely represent the onset of exhumation. In the
HePRZ (40°–80°C). For all models, this window of exhu- Angastaco area, where corresponding apatite fission track
mation has been interpreted, and it superimposed upon ages have not been annealed during burial [Coutand et al.,
Figure 9 as a cloud of time during which exhumation 2006], there is further confidence in interpreting young
through the HePRZ was occurring for each location. Model AHe ages as reflecting the onset of exhumation of the basin
AngPuc06 was conducted using grains a1 and a5, and strata.
suggests that early cooling likely initiated by ∼14 Ma, and [55] For the two Cretaceous sandstone samples collected
that the sample had cooled through the HePRZ by ∼12 Ma in the Sierra de los Colorados anticline, just east of
(Figure 10a). These results and presence of an age‐eU Angastaco area, the sample collected in the deepest strati-
relationship suggests that sample AngPuc06 cooled slowly, graphic portion of anticline has been fully reset during burial
and likely spent an extended period of time in the HePRZ. (sample 04.29.08‐01), whereas the shallower sample con-
[51] Model Ang2EB was conducted using grains a4 and tains a mixture of nonreset, partially reset, and fully reset
a5, and suggests that early cooling initiated ∼11 Ma, with ages (sample 04.14.08‐01) (see Table 1). Thus, the shallower
cooling through the HePRZ by ∼8 Ma (Figure 9). These sample was likely buried to a maximum depth within the
results and the presence of an age‐eU relationship suggests HePRZ, while the deeper sample experienced temperatures
that sample Ang2EB spent significant time in the HePRZ. below the HePRZ; differences in composition have con-
[52] Model QC1EB080 was conducted using grains a1 tributed to the observed scatter of ages. Because these sam-
and a7. Although HeFTy was never able to find a “good” ples are stratigraphically ∼400 m apart, sample 04.29.08‐01
path for these grains, 1500 “acceptable” paths were found was buried to a maximum depth just below the HePRZ, and
that suggest cooling began by ∼13 Ma, with cooling through therefore should have entered the HePRZ shortly to imme-
70°C by 12 Ma (Figure 10c). Modeling of samples diately after the onset of Neogene exhumation. HeFTy
04.29.08‐01, LY, and ALP all suggest rapid cooling. Model modeling suggests that this sample cooled through the
04.29.08‐01 was conducted using grains a2, a3, and a4, and HePRZ at rate ≥0.5 mm/yr, and therefore the ∼400 m
suggests that early cooling began ∼10 Ma, with cooling stratigraphic offset between the two samples can be used to
through the HePRZ at ca. 6–7 Ma (Figure 10d). Model LY infer that the maximum time between onset of exhumation
was conducted using grains a4 and a5, and suggests that and entering the HePRZ is 0.8 Myr. This suggests that the
early cooling began by ∼6 Ma, with final cooling through AHe ages of sample 04.29.08‐01 do represent the onset of
the HePRZ at ∼2–4 Ma (Figure 10e). Model ALP was exhumation for this range, and that any variability is within
conducted using grains a1, a2, a3, and a5, and suggests error of the AHe signal.
that early cooling through the HePRZ had initiated at ∼35– [56] The Cretaceous samples LY and ALP have been
22 Ma, with final cooling through the 80°–40° T‐window at completely thermally reset, and all grain ages are signifi-
ca. 0–4 Ma (Figures A1 and 10f). cantly younger than DA. While the AHe data for these
samples do not necessarily constrain the onset of exhuma-
5.7. Interpretation of AHe Ages and HeFTy Modeling tion for these samples, HeFTy modeling shows with confi-
Results dence that they have experienced rapid cooling during the
[53] The detrital AHe system provides information on the last ∼6 Myr, which we interpret to reflect major range
latest stage of exhumation in the uppermost crust, and by exhumation caused by the onset of major active deformation
exploiting the effects of radiation damage on multiple on adjacent reverse faults.
crystals using HeFTy modeling, we can expand our confi- [57] The majority of the analyzed grains give ages that can
dence in the interpretation of recent cooling. While AHe be interpreted to roughly represent the onset of basin
thermochronology cannot tell us about events that occurred exhumation in the southern EC. The AHe grain ages and
while the sample was buried beneath the HePRZ, incorpo- HeFTy model results indicate that the sequence of exhu-
rating the data with reasonable assumptions and known mation in the southern EC has swept consistently eastward
geological constraints allows us to make interpretations that during the Miocene and Pliocene (Figure 10). We interpret
greatly elucidate the recent patterns of exhumation in the the younging in cooling ages eastward as the result of ero-
region. sion related to progressive eastward younging of fault acti-
[54] For the six Cenozoic samples that contain an old vation. A linear regression placed through the combined
population of partially to nonreset grain ages (Pucará, young grain age populations suggests that this wave of

17 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

exhumation (and inferred deformation) propagated eastward ∼10 Ma, at which point deformation and exhumation
at a rate of ∼8.3 mm/yr (r2 = 0.7) (Figure 10). There is an migrated eastward as the Sierra de los Colorados anticline
interesting recurrence of a second population of young grain experienced rapid exhumation ∼9 Ma. The Pucará, Angas-
ages between 21 and 28 Ma in multiple parts of the study taco, and Sierra de los Colorados areas continued to exhume
area. We interpret this to represent a possible early, minor as the Sierra de Leon Muerto and western Lerma Valley
exhumation signal between deposition of the Quebrada de experienced significant exhumation ∼6–3 Ma (Figure 10).
los Colorados and Angastaco Formations. This timeline corresponds to the timing of shortening pro-
posed by Marrett et al. (∼13–1 Ma) [1994], as well as the
6. Discussion eastward propagation of growth structures proposed in the
Mio‐Pliocene by Carrera and Muñoz [2008]. This pattern
6.1. Angastaco Area of eastward younging deformation involved mainly west-
[58] This study confirms that deposition of the Miocene‐ ward verging thrust faults, and suggests that, regardless of
Pliocene sedimentary package in the Angastaco area was local structural geometry, overall orogenic strain propagated
contemporaneous with local deformation. U‐Pb dating in eastward as expected during the development of a tapered
this study and B. Carrapa et al. (submitted manuscript, orogenic wedge.
2011) limits the age range of Cenozoic strata in the An-
gastaco area to ∼37–2.3 Ma. The dominance of flexural slip 6.3. Evolution of Orogenic Propagation Across the
folding implies that the exposed strata experienced brittle Southern Central Andes
deformation at low temperature and confining pressure. [60] Our new data set, when compared to existing data,
Based on detrital U‐Pb geochronology from strata below the documents a wave of exhumation that propagated across the
angular unconformity shown in Figure 6, synsedimentary southern EC of NW Argentina at 8.3 mm/yr between ∼14
deformation initiated in the area by ∼14 Ma. AHe thermo- and ∼3 Ma (Figure 11), which is roughly in agreement with
chronology documents significant rapid cooling along the the average rate of strain propagation across the entire
western margin of the Angastaco area by no later than central Andes (∼7.8 mm/yr) at the same latitude (Figure 11).
10 Ma, demonstrating that the western part of the basin was Major development of the southern part of the central Andes
exhuming coevally with active deposition in the central and began during the mid to late Cretaceous in the Salar de
eastern parts of the basin. Thus, much of the sediment in the Atacama basin [Arriagada et al., 2006; Mpodozis et al.,
Palo Pintado and San Felipe Formations may have been 2005]. Later, during the middle to late Eocene, the defor-
recycled from the Quebrada de los Colorados and Angastaco mation front jumped eastward across the width of the Puna
Formations just a few kilometers to the west. It is interesting region, causing a roughly synchronous onset of deformation
to note that samples 04.29.08‐01 and 04.29.08‐01 from the from ∼40 to 38 Ma in the Salar de Antofalla, Salar de
Sierra de los Colorados yield cooling ages between 6 and Arizaro and Salar de Pastos Grandes (and possibly La
10 Ma. This indicates that cooling and exhumation in, at least Poma) basins (Figure 11) [Kraemer et al., 1999; Voss, 2002;
a portion of, this range initiated before the change in prove- Carrapa et al., 2005; Mpodozis et al., 2005; Arriagada et al.,
nance and basin reorganization documented by paleocurrent 2006; Jordan and Mpodozis, 2006; Hongn et al., 2007;
and conglomerate deposition at 4 Ma [Bywater‐Reyes et al., Carrapa and DeCelles, 2008]. Deformation was localized
2010]. In turn, this suggests that despite being deformed within the eastern Puna/western EC of NW Argentina
and eroded as early as ∼10 Ma, the Sierra de los Colorados between ∼39 and ∼22 Ma and swept eastward across the
did not have enough topographic relief to shed coarse eastern EC during the Miocene‐Pliocene (Figure 11). Evi-
material to the surrounding areas and that it was likely not dence of deformation and exhumation in the Puna Plateau
an orographic barrier. More likely this earlier exhumation is at ca. 25 Ma [Carrapa et al., 2005] suggests out‐of‐sequence
related to the formation of local, low relief. This is supported deformation of the hinterland at this time.
by fine‐grained Miocene deposits of the Anta and Jesus [61] Despite similarities of deformation during the early
Maria formations in the La Viña area, which constitute the Cenozoic, the Oligocene and Neogene kinematic histories
distal equivalents of the Angastaco Formation [Starck and in Bolivia and NW Argentina are significantly different.
Anzótegui, 2001; B. Carrapa et al., submitted manuscript, Shortening in Bolivia continued to propagate eastward
2011]. Overall the data indicate that the main river system through the EC and inter‐Andean zone throughout the
was still draining to the east and was not disrupted by the Oligocene [McQuarrie et al., 2005; Ege et al., 2007],
regional uplift of ranges to the east of the Angastaco area, whereas the Puna experienced internal deformation and
which did not occur until ∼4 Ma when conglomerates exhumation [Carrapa et al., 2005] with very limited east-
derived from the Sierra de los Colorados range were ward propagation during this time interval (Figure 11).
deposited in the Angastaco and La Viña areas. During Miocene‐Pliocene time (∼22–3 Ma) deformation
was concentrated within the EC/Santa Barbara system of
6.2. Recent Exhumation in the Southernmost EC NW Argentina (∼26°S, 66°W) and it was migrating through
[59] AHe data show an eastward younging of cooling ages the study area in sequence between ∼14 and 3 Ma. In the
that suggests an eastward migrating wave of exhumation sub‐Andes of NW Argentina‐S Bolivia [Echavarria et al.,
and related deformation during the Miocene‐Pliocene. 2003] deformation was migrating through the region
These data document middle Miocene exhumation of the between ∼8 and 3 Ma and it was in the sub‐Andes of
Pucará area and initiation of structural growth within the Bolivia by ∼10 Ma [Gubbels et al., 1993].
Angastaco area at ∼14 Ma (Figure 10). The western [62] Interesting questions that arise from this comparison
Angastaco area experienced significant exhumation by are (1) why did different segments of the central Andes

18 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

behave differently during the Neogene and (2) what are the migration of the deformation front into the Santa Barbara
possible controlling mechanisms? Mass removal underneath system after 4 Ma.
the plateau region [Kay et al., 1999] and subsequent
adjustment of the orogenic system has been proposed as one
of the main controlling factors on the evolution of Cordil- 7. Conclusions
leran type orogenic systems such as the Andes [DeCelles [63] Although the central Andes express great variability
et al., 2009]. In particular the abrupt eastward shifts of the in timing and structural style along strike, remarkable sim-
deformation front observed in Bolivia at ca. 45–40 Ma and ilarities exist between the timing and rates of propagation of
again at 10 Ma have been explained as a response to iso- deformation and exhumation in the northern (Bolivian) and
static elevation gain following the removal of dense arc southern (Argentinean) portions of the central Andes, par-
roots beneath the orogenic hinterland [DeCelles et al., 2009; ticularly when major structural features are followed along
Kay et al., 1999]. However, significant along‐strike varia- strike regardless of the established tectonomorphic nomen-
tions in the Neogene kinematic history between the study clature. The development of the Bolivian EC was roughly
area and regions farther north suggest differences in oro- coeval with the Eocene‐Oligocene internal deformation of
genic processes. For example the Altiplano plateau to the the Puna region, and the development of the Argentinean
north has been explained in terms of orogenic‐scale litho- EC was coeval with the development of the sub‐Andean
spheric removal [Garzione et al., 2006]. Alternatively the system (Figure 11) [Carrera and Muñoz, 2008; Echavarria
whole region has been explained by continuous shortening et al., 2003; McQuarrie et al., 2005, 2008; Uba et al.,
and gradual uplift [Ehlers and Poulsen, 2009]. The Puna 2009]. Many of the geomorphic and structural features of
Plateau to the south has been associated with smaller‐scale, the Puna can be traced northward along strike into the
and perhaps diachronous, lithospheric removal [Carrapa Bolivian EC (Figure 11), and because both provinces contain
et al., 2009]. Thus, orogenic cyclicity has likely not oper- copious evidence of Oligocene deformation, it appears that
ated uniformly over along‐strike distances of >500 km, the eastern part of the Puna correlates both geologically and
perhaps because of differences in shortening rates, arc melt temporally with the Bolivian EC and inter‐Andean zone.
and dense residue production rates resulting in different [64] If we temporarily overlook the geographic constric-
upper crustal to surficial structural responses. A younger tions implied by predefined tectonomorphic provinces, and
lithospheric removal event in the Puna Plateau at ∼3 Ma as instead focus on regions of the central Andes that correlate
proposed by DeCelles et al. [2009] is consistent with in time, the propagation of deformation rates (and exhu-

Figure 11. Regional map of the central Andes illustrating the timing and rates of propagation of deformation across the
area. White circles represent timing (in Ma) of the earliest evidence for Cenozoic deformation, wedge top deposition, or
synorogenic deformation. Values along white lines represent rates at which the orogenic wedge propagated eastward. Base
map is a NASA SRTM digital elevation model. Tectonomorphic domains follow description in Figure 1 (tectonomorphic
provinces modified after Schoenbohm and Strecker [2009]). Timing of deformation and propagation rates are based on the
following: Balanced cross‐section restoration, apatite fission track, and zircon fission track thermochronology of McQuarrie
et al. [2008] (letter a). Modeling of 40Ar/39Ar and AFT ages after Gillis et al. [2006] (letter a*). Balanced cross‐section
restoration of McQuarrie [2002] and McQuarrie et al. [2005] (letter b). Loose estimates based on flexural modeling of
stratigraphic data combined with GPS convergence rates. Flexural modeling of stratigraphic data by DeCelles and Horton
[2003] suggests that the foreland flexural wave migrated eastward at a rate of 12–20 mm/yr during the Oligocene, which
combined with shortening rates of ∼3–6 mm/yr suggests an orogenic propagation rate of 6–17 mm/yr [after DeCelles and
DeCelles, 2001; DeCelles and Horton, 2003; Norabuena et al., 1998] (letter c). Seismic interpretation and balanced cross‐
section restoration of Uba et al. [2009] (letter d). Cross‐cutting relationships, growth strata, unconformities, provenance,
migration of facies, and accumulation rate history of Echavarria et al. [2003] (letter e). Middle‐to‐late Cretaceous onset of
deformation inferred from seismic interpretation, structural mapping, and sedimentology of Arriagada et al. [2006] and
Mpodozis et al. [2005] (letter f). Sedimentology, provenance analysis, paleocurrent interpretation, and apatite fission track
thermochronology of Carrapa et al. [2005], Kraemer et al. [1999], and Voss [2002] (letter g). Carrapa et al. [2009];
stratigraphy of Jordan and Mpodozis [2006] (letter h). Apatite fission track data and sedimentology stratigraphy; [after
Carrapa and DeCelles, 2008] (letter i). Apatite fission track thermochronology of Deeken et al. [2006] (letter j). Growth
strata deted in this study (letter k). Detrital (U‐Th)/He thermochronology in this study (letter l). Seismic interpretation,
apatite fission track thermochronology, and geochronology of Mortimer et al. [2007] (letter m). Sedimentology, stratig-
raphy, structural geology and U‐Pb geochronology of Carrapa et al. [2008a] (letter n). Apatite fission track data from
Carrapa et al. [2006] (letter o). Shaded gray lines show the correlation along strike of the deformation front independently
from predefined tectonomorphic domains. In this figure only studies that have a significant W‐E area coverage and/or use a
similar approach to ours (structural relationships/thermochronology) and are interpreted here as representing unequivocal
timing of in‐sequence deformation are included. Extra data are discussed in the text. For a review on available multidis-
ciplinary data in the Andean Plateau we refer to Barnes and Ehlers [2009]. Deformation was localized within the eastern
Puna/western EC of NW Argentina between ca. 39 and ca. 22 Ma resulting in very slow (1.8 mm/yr) propagation rates.
Propagation rates in the sub‐Andean fold‐and‐thrust belt were similar, although slightly more erratic in places, migrating
eastward at ∼9.9 mm/yr from ca. 9 to 1.2 Ma at 22–23°S [Echavarria et al., 2003], ∼15 mm/yr from ca. 12 to 0 Ma at 21°S
[Uba et al., 2009], and 6–8 mm/yr from ca. 20 to 0 Ma at 19°–20°S [McQuarrie et al., 2005].

19 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Figure 11

20 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

mation) of the NW Argentinean EC is comparable to simultaneously in static mode. Measurement of 238U, 232U,
propagation of deformation rates in the thin‐skinned sub‐ 208
Pb, and 206Pb was conducted with eight 1011 W Faraday
Andes during Mio‐Pliocene time. Evidence for this comes detectors. A 1012 W Faraday collector measured 207Pb, and
204
from a variety of sources (Figure 11). Pb was measured with an ion‐counting channel. Ion
[65] In summary, whereas the Mio‐Pliocene rate of east- yields were ∼1.0 mv per ppm. A typical analysis included
ward propagation of the orogenic front varies along strike, 20 s of background measurement, 12 s integrated mea-
the propagation rate in the study area of NW Argentina surements as the laser fired, and 30 s to purge the instrument
(8.3 mm/yr) falls within the same range of rates for the EC and prepare for the next analysis. Unknowns were calibrated
of Bolivia (2.1–15 mm/a), and is remarkably similar to the against a Sri Lanka standard of known age after every fifth
long‐term average rate of orogenic propagation in the cen- zircon analysis to correct for apparent fractionation of Pb
tral Andes (7.8 mm/yr) [McQuarrie et al., 2005]. More isotopes and inter element fractionation of Pb/U. Correc-
importantly, similar rates of propagation of the deformation tions for common Pb were calculated using measured 204Pb
front along the plateau margin characterized by significantly and an assumed initial Pb composition from Stacey and
different structural styles suggests that the rate at which the Kramers [1975]. Measurement uncertainty was generally
deformation front has driven into the foreland is determined 1–2% (2s). Because of their sensitivity to young ages,
206
more by the dynamics of regional eastward expansion of the Pb/238U ages were used, and analyses with more than
Andean orogeny than the by local structural style, and 30% discordance between 206Pb/238U and 206Pb/207Pb ages
therefore may develop independently of features that might were removed from consideration. For more information
control local structural style (e.g., the nature of the local concerning lab methods at the Arizona LaserChron center,
basement). It also interesting to note that rates of propaga- see Gehrels et al. [2008].
tion of deformation are similar along strike despite of the
A2. (U‐Th)/He Lab Methods
significant differences in amounts of shortening between
Bolivia (>300 km) [Kley et al., 1997; McQuarrie, 2002] and [68] Mineral separation and grain picking were conducted
Argentina (<100 km) [Kley and Monaldi, 1998, and refer- at the University of Wyoming. Apatites were separated from
ences therein]. This may be explained by the fact that the ten detrital samples using a jaw crusher, disk mill, wilfley
Altiplano and Eastern Cordillera of Bolivia have deformed table, Frantz electromagnetic separator, and methyl iodide
mainly by simple shear whereas the Puna Plateau and heavy liquids. Using a Leica M165C stereomicroscope, each
Eastern Cordillera of NW Argentina by pure shear apatite fraction was thoroughly screened for relatively
[Allmendinger and Gubbels, 1996]. euhedral apatites that were free of cracks and inclusions and
[66] The differences in the timing of late Miocene‐ had minimum dimensions greater than 60 mm. Unfortu-
Pliocene deformation between the study area and regions nately, detrital apatites are often small, broken, or abraded,
farther north suggest different structural and surficial and thus pose a significant challenge during grain picking.
responses to lithospheric processes. For example the late Although six individual grains were desired, in some sam-
Cenozoic uplift history of the Altiplano in Bolivia has been ples only four acceptable grains could be found. All grains
associated with large‐scale lithospheric removal [Garzione were photographed and measured for Ft calculations using
et al., 2006], which has been used to explain changes in the method of Farley et al. [1996].
taper angle and subsequent migration of the orogenic front [69] Apatite U‐Th/He analyses were conducted using the
outward in the sub‐Andes at ∼8 Ma [DeCelles et al., 2009]. ultra high vacuum noble gas extraction and purification line
In NW Argentina, small‐scale lithospheric removal has been and ICP‐MS at the University of Kansas. For each single‐
proposed to have influenced Puna Plateau evolution [Kay grain aliquot, 4He was extracted with a continuous mode
and Coira, 2009], late Cenozoic sedimentary basin evolu- Nd‐YAG laser, spiked with 3He for isotope dilution, con-
tion and paleogeography [Strecker et al., 2009; Carrapa centrated and purified with in a Janis cryogenic trap,
et al., 2008b]. However, in this last case the size, volume released, and the 3He/4He ratio was subsequently measured
and possible diachrounous removal of lithosphere has with a Blazers Prisma QMS‐200 quadrupole mass spec-
probably prevented a syncronous response of deformation trometer. After degassing, each sample was analyzed for U,
forelandward. Th, Sm, and selected REE using a Fisons/VG PlasmaQuad
II Inductively Coupled Plasma Mass Spectrometer (ICP‐
MS). Samples were routinely calibrated against Durango
Appendix A apatite standard to ensure laboratory accuracy.
A1. U‐Pb Lab Methods
A3. Additional (U‐Th)/He Details
[67] Mineral separation was conducted at the University
of Wyoming using a jaw crusher, disk mill, wilfley table, [70] Radiation damage introduces isolated structural
Frantz electromagnetic separator, and methyl iodide heavy defects and vacancies into the apatite crystal structure that
liquids. Detrital zircon U‐Pb analyses were conducted at the can substantially retard the diffusion of 4He and lead to
Arizona LaserChron center using standard methods of laser anomalously old ages [Shuster et al., 2006; Shuster and
ablation multicollector inductively coupled plasma mass Farley, 2009]. Radiation damage controls AHe dates more
spectrometry (LA‐MC‐ICP MS). U‐Pb data are presented in than any other known factor, is twice as influential as grain
Table A1. Zircons were ablated with a New Wave/Lambda size, can cause expansion of the HePRZ to 50°–115°C, and
Physic DUV193 Excimer laser using a spot diameter of generally correlates with [eU], the effective uranium con-
35 mm, material was transported by He gas to the plasma centration of the grain (eU = U + 0.235Th) [Shuster et al.,
source of a GVI Isoprobe, and isotopes were measured 2006]. Because radiation damage accumulates over geo-

21 of 30
Table A1. Detrital U‐Pb Data
Isotope Ratios Apparent Ages (Ma)
TC3003

U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration
Analysis (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%)
SAMPLE 04.14.08‐02
04140802‐1 227 18315 5.0 17.5152 2.1 0.6082 2.3 0.0773 0.9 0.41 479.8 4.3 482.5 8.8 495.2 46.1 479.8 4.3 96.9
04140802‐3 2640 1017 4.2 22.7779 24.1 0.0022 24.4 0.0004 3.9 0.16 2.3 0.1 2.2 0.5 −116.2 601.0 2.3 0.1 −2.0
04140802‐4 607 54378 2.2 16.6871 1.3 0.7920 1.7 0.0959 1.1 0.66 590.1 6.2 592.3 7.5 600.9 27.3 590.1 6.2 98.2
04140802‐5 1068 411 1.2 18.1429 95.1 0.0021 95.5 0.0003 8.3 0.09 1.8 0.1 2.1 2.0 417.0 673.7 1.8 0.1 0.4
04140802‐6 284 27078 3.5 17.4331 1.9 0.6636 2.5 0.0839 1.6 0.65 519.4 8.1 516.8 10.2 505.5 42.1 519.4 8.1 102.7
04140802‐7 216 52539 3.4 13.6872 2.2 1.6792 3.8 0.1667 3.2 0.82 993.9 29.0 1000.7 24.5 1015.6 44.8 1015.6 44.8 97.9
04140802‐8 370 31068 2.5 17.8001 1.9 0.6016 2.1 0.0777 0.9 0.41 482.2 4.0 478.3 8.1 459.5 43.1 482.2 4.0 105.0
04140802‐9 3532 888 2.1 19.2971 6.8 0.0023 7.3 0.0003 2.6 0.36 2.1 0.1 2.3 0.2 277.6 156.1 2.1 0.1 0.8
04140802‐10 5073 1713 5.2 20.7960 6.1 0.0026 6.4 0.0004 2.1 0.32 2.5 0.1 2.6 0.2 103.5 143.2 2.5 0.1 2.4
04140802‐11 53 12228 3.7 13.9223 2.6 1.5330 2.7 0.1548 0.5 0.19 927.8 4.3 943.7 16.4 981.1 53.4 981.1 53.4 94.6
04140802‐12 1064 729 5.8 32.2610 79.2 0.0023 79.3 0.0005 4.0 0.05 3.4 0.1 2.3 1.8 −1059.5 1167.2 3.4 0.1 −0.3
04140802‐13 1057 399 1.2 18.0740 15.3 0.0049 15.7 0.0006 3.4 0.22 4.1 0.1 5.0 0.8 425.5 344.0 4.1 0.1 1.0
04140802‐15 2654 525 0.6 29.9985 30.9 0.0015 30.9 0.0003 1.1 0.04 2.0 0.0 1.5 0.5 −846.2 901.6 2.0 0.0 −0.2
04140802‐16 2237 705 4.1 20.8624 60.8 0.0024 60.9 0.0004 4.4 0.07 2.3 0.1 2.4 1.5 96.0 1583.6 2.3 0.1 2.4
04140802‐17 2810 1137 3.5 20.2923 9.3 0.0030 9.4 0.0004 1.6 0.16 2.8 0.0 3.0 0.3 161.1 218.4 2.8 0.0 1.8
04140802‐18 2660 3363 4.5 22.6853 11.6 0.0082 11.6 0.0013 0.6 0.05 8.7 0.0 8.3 1.0 −106.1 287.0 8.7 0.0 −8.2
04140802‐19 189 10602 2.7 17.3961 2.4 0.5868 3.2 0.0740 2.2 0.68 460.4 9.8 468.8 12.1 510.2 51.8 460.4 9.8 90.2
04140802‐21 69 21492 0.8 8.7188 1.4 5.2650 1.7 0.3329 1.1 0.63 1852.6 17.6 1863.2 14.9 1875.1 24.5 1875.1 24.5 98.8
04140802‐22 3409 1359 7.2 21.9338 14.9 0.0021 15.2 0.0003 3.1 0.20 2.2 0.1 2.2 0.3 −23.9 363.3 2.2 0.1 −9.1
04140802‐23 437 67716 1.9 12.8987 2.7 1.5470 4.8 0.1447 3.9 0.82 871.3 31.9 949.3 29.5 1134.8 54.6 1134.8 54.6 76.8
04140802‐24 1228 474 6.2 34.6046 40.2 0.0014 40.2 0.0003 2.2 0.06 2.2 0.0 1.4 0.6 −1275.4 1309.6 2.2 0.0 −0.2

22 of 30
04140802‐25 161 11025 4.5 17.5126 2.8 0.5110 3.9 0.0649 2.6 0.69 405.4 10.4 419.1 13.2 495.5 61.8 405.4 10.4 81.8
04140802‐27 2204 861 3.2 22.6877 24.0 0.0022 24.1 0.0004 2.1 0.09 2.3 0.0 2.2 0.5 −106.4 598.3 2.3 0.0 −2.2
04140802‐28 5541 5718 99.6 21.2060 6.1 0.0074 6.1 0.0011 0.7 0.11 7.4 0.1 7.5 0.5 57.1 145.7 7.4 0.1 12.9
04140802‐29 162 16416 2.1 17.1542 1.8 0.6831 2.5 0.0850 1.8 0.71 525.8 9.1 528.6 10.4 540.9 38.6 525.8 9.1 97.2
04140802‐30 2962 1317 3.3 21.4510 12.5 0.0027 12.5 0.0004 1.5 0.12 2.7 0.0 2.7 0.3 29.7 299.5 2.7 0.0 9.0
04140802‐31 255 7506 3.3 14.2336 16.3 0.8628 16.3 0.0891 0.9 0.05 550.0 4.6 631.7 77.0 935.8 336.8 550.0 4.6 58.8
04140802‐33 2294 915 3.7 22.7749 22.0 0.0023 22.1 0.0004 1.2 0.05 2.4 0.0 2.3 0.5 −115.8 548.4 2.4 0.0 −2.1
04140802‐34 257 8193 2.8 16.3782 4.3 0.7019 4.4 0.0834 0.8 0.19 516.2 4.1 539.9 18.5 641.3 93.4 516.2 4.1 80.5
04140802‐35 156 15750 2.4 17.7686 3.8 0.6144 4.0 0.0792 1.1 0.27 491.2 5.1 486.3 15.4 463.4 85.0 491.2 5.1 106.0
04140802‐36 2298 930 4.4 21.1231 20.8 0.0023 20.9 0.0004 1.8 0.09 2.3 0.0 2.4 0.5 66.5 500.7 2.3 0.0 3.5
04140802‐38 1312 384 0.6 26.3769 60.6 0.0021 60.6 0.0004 1.5 0.02 2.6 0.0 2.1 1.3 −491.4 1759.7 2.6 0.0 −0.5
04140802‐39 2504 903 4.4 24.2709 29.2 0.0018 29.5 0.0003 4.0 0.14 2.1 0.1 1.8 0.5 −275.1 756.6 2.1 0.1 −0.8
CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION

04140802‐40 178 45420 3.7 13.2694 1.8 1.8811 2.0 0.1810 0.9 0.46 1072.7 9.0 1074.5 13.1 1078.2 35.2 1078.2 35.2 99.5
SAMPLE 05.07.08‐A.Unc
050708‐AUNC‐1 600 11614 2.8 17.2735 1.7 0.6104 2.4 0.0765 1.7 0.69 475.0 7.7 483.8 9.4 525.7 38.4 475.0 7.7 90.4
050708‐AUNC‐2 322 6966 12.2 17.7149 1.5 0.6014 2.0 0.0773 1.3 0.64 479.8 5.9 478.1 7.6 470.1 34.0 479.8 5.9 102.1
050708‐AUNC‐3 424 15192 12.1 17.6687 2.2 0.5939 2.7 0.0761 1.6 0.60 472.8 7.4 473.3 10.2 475.9 47.6 472.8 7.4 99.4
050708‐AUNC‐4 312 11976 3.3 17.6122 1.8 0.6150 2.4 0.0786 1.6 0.66 487.5 7.4 486.7 9.2 483.0 39.8 487.5 7.4 100.9
050708‐AUNC‐5 352 17030 6.3 17.6579 1.6 0.6102 2.1 0.0781 1.4 0.68 485.0 6.7 483.7 8.2 477.2 34.7 485.0 6.7 101.6
050708‐AUNC‐6 352 15958 7.0 17.8076 2.2 0.5912 3.0 0.0764 2.0 0.67 474.3 9.1 471.6 11.2 458.5 49.0 474.3 9.1 103.4
050708‐AUNC‐7 366 15190 15.8 17.8131 1.9 0.5904 2.2 0.0763 1.1 0.50 473.9 5.1 471.1 8.4 457.8 42.8 473.9 5.1 103.5
050708‐AUNC‐8 597 22476 11.6 17.6944 3.3 0.5948 3.5 0.0763 1.1 0.32 474.2 5.1 473.9 13.3 472.6 73.9 474.2 5.1 100.3
050708‐AUNC‐9 421 14036 12.6 17.5969 2.1 0.6041 2.4 0.0771 1.3 0.52 478.8 5.8 479.8 9.2 484.9 45.7 478.8 5.8 98.7
050708‐AUNC‐10 574 60018 11.6 17.4002 1.6 0.6069 2.2 0.0766 1.5 0.68 475.7 6.8 481.6 8.3 509.6 35.1 475.7 6.8 93.3
050708‐AUNC‐11 403 14664 4.6 17.9142 1.5 0.5906 2.0 0.0767 1.3 0.66 476.6 6.2 471.2 7.7 445.3 34.0 476.6 6.2 107.0
050708‐AUNC‐12 891 43230 47.7 16.0399 5.8 0.7311 7.0 0.0851 3.8 0.55 526.2 19.2 557.2 29.8 686.0 124.3 526.2 19.2 76.7
TC3003
Table A1. (continued)
Isotope Ratios Apparent Ages (Ma)
TC3003

U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration
Analysis (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%)
050708‐AUNC‐13 140 6538 5.1 17.5819 3.3 0.6253 3.3 0.0797 0.5 0.15 494.5 2.4 493.1 13.0 486.8 72.4 494.5 2.4 101.6
050708‐AUNC‐14 330 6706 6.1 16.4621 2.0 0.6811 4.3 0.0813 3.8 0.88 504.0 18.2 527.4 17.6 630.3 43.8 504.0 18.2 80.0
050708‐AUNC‐16 379 16902 8.6 17.8625 1.6 0.5848 2.9 0.0758 2.4 0.84 470.8 11.0 467.6 10.9 451.7 35.0 470.8 11.0 104.2
050708‐AUNC‐17 287 14598 2.3 17.3930 2.4 0.6619 3.1 0.0835 1.9 0.62 517.0 9.6 515.8 12.6 510.5 53.7 517.0 9.6 101.3
050708‐AUNC‐18 263 22614 2.3 13.5773 1.3 1.5816 2.5 0.1557 2.2 0.86 933.1 18.9 963.0 15.7 1031.9 25.9 1031.9 25.9 90.4
050708‐AUNC‐19 313 23586 15.2 17.5765 2.2 0.6028 2.6 0.0768 1.4 0.54 477.3 6.4 479.0 9.9 487.4 48.0 477.3 6.4 97.9
050708‐AUNC‐20 340 14650 8.9 17.7632 1.4 0.5914 1.8 0.0762 1.2 0.66 473.3 5.5 471.7 6.8 464.1 30.0 473.3 5.5 102.0
050708‐AUNC‐21 358 17568 14.4 17.8406 1.1 0.5917 1.6 0.0766 1.1 0.68 475.6 4.9 471.9 5.9 454.4 25.2 475.6 4.9 104.7
050708‐AUNC‐22 152 8546 5.7 15.6312 3.3 0.7964 4.4 0.0903 2.9 0.66 557.2 15.4 594.8 19.8 740.8 70.3 557.2 15.4 75.2
050708‐AUNC‐23 512 22046 25.4 17.6155 1.6 0.5855 2.4 0.0748 1.8 0.75 465.0 8.0 468.0 8.9 482.6 34.7 465.0 8.0 96.4
050708‐AUNC‐24 421 19122 3.8 17.8672 1.3 0.5912 1.6 0.0766 1.0 0.63 475.9 4.7 471.7 6.2 451.1 28.4 475.9 4.7 105.5
050708‐AUNC‐25 284 17236 8.0 17.1379 1.1 0.6874 1.6 0.0854 1.2 0.72 528.5 5.8 531.2 6.6 542.9 24.5 528.5 5.8 97.3
050708‐AUNC‐26 511 22184 14.5 17.5288 2.0 0.6137 2.2 0.0780 1.0 0.45 484.3 4.7 485.9 8.6 493.4 43.8 484.3 4.7 98.1
050708‐AUNC‐27 273 16428 4.9 14.8111 1.8 1.0120 4.2 0.1087 3.8 0.91 665.2 24.1 709.9 21.5 853.8 36.9 665.2 24.1 77.9
050708‐AUNC‐28 322 8660 8.1 17.4928 1.1 0.6050 2.3 0.0768 2.1 0.89 476.7 9.6 480.4 9.0 497.9 23.2 476.7 9.6 95.7
050708‐AUNC‐29 335 30910 5.9 13.8530 2.0 1.5746 3.2 0.1582 2.5 0.79 946.8 21.8 960.2 19.6 991.2 39.6 991.2 39.6 95.5
050708‐AUNC‐30 310 16152 3.0 17.5750 1.4 0.6047 1.7 0.0771 1.0 0.59 478.7 4.8 480.2 6.7 487.6 31.2 478.7 4.8 98.2
050708‐AUNC‐31 568 26808 11.7 17.5409 1.0 0.6179 1.7 0.0786 1.4 0.81 487.8 6.4 488.5 6.5 491.9 21.9 487.8 6.4 99.2
050708‐AUNC‐32 630 20292 15.2 17.4866 1.9 0.6158 2.6 0.0781 1.8 0.69 484.8 8.3 487.2 9.9 498.7 40.8 484.8 8.3 97.2
050708‐AUNC‐34 566 24030 2.3 17.5901 1.7 0.6061 1.8 0.0773 0.6 0.35 480.2 3.0 481.1 7.0 485.7 38.1 480.2 3.0 98.9
050708‐AUNC‐35 375 8542 1.3 17.5430 2.2 0.6337 2.5 0.0806 1.0 0.41 499.9 4.9 498.4 9.7 491.6 49.6 499.9 4.9 101.7
050708‐AUNC‐36 513 8118 10.3 17.4223 0.7 0.6173 1.1 0.0780 0.8 0.75 484.2 3.9 488.2 4.3 506.8 16.0 484.2 3.9 95.5

23 of 30
050708‐AUNC‐37 198 6292 9.7 17.6830 1.8 0.5914 2.1 0.0758 1.1 0.50 471.3 4.8 471.7 8.0 474.1 40.7 471.3 4.8 99.4
050708‐AUNC‐38 291 9520 3.0 17.6084 1.9 0.6630 1.9 0.0847 0.5 0.26 524.0 2.5 516.5 7.8 483.5 41.1 524.0 2.5 108.4
050708‐AUNC‐39 292 6258 3.8 18.0625 2.5 0.5831 2.7 0.0764 0.8 0.31 474.5 3.8 466.5 10.0 426.9 56.5 474.5 3.8 111.2
050708‐AUNC‐41 408 13642 10.5 17.7339 2.1 0.5871 2.7 0.0755 1.8 0.66 469.3 8.2 469.0 10.3 467.7 45.4 469.3 8.2 100.3
050708‐AUNC‐42 431 31526 7.6 14.2932 2.5 1.2476 2.9 0.1293 1.6 0.53 784.0 11.4 822.3 16.4 927.3 50.7 927.3 50.7 84.6
050708‐AUNC‐43 818 19732 2.4 17.4104 1.0 0.5900 1.1 0.0745 0.5 0.45 463.2 2.3 470.9 4.3 508.4 22.3 463.2 2.3 91.1
050708‐AUNC‐44 244 10462 6.3 17.8770 2.6 0.5788 3.2 0.0750 1.8 0.56 466.5 8.1 463.7 11.9 449.9 58.5 466.5 8.1 103.7
050708‐AUNC‐45 238 19686 6.4 14.8534 3.0 1.0099 4.0 0.1088 2.7 0.66 665.7 16.8 708.8 20.6 847.8 63.3 665.7 16.8 78.5
050708‐AUNC‐46 403 19976 6.1 14.6417 1.4 1.1280 2.0 0.1198 1.5 0.74 729.3 10.1 766.8 10.8 877.6 28.0 729.3 10.1 83.1
050708‐AUNC‐47 617 28992 2.9 17.3411 2.1 0.6913 2.3 0.0869 0.8 0.34 537.4 4.0 533.6 9.4 517.1 46.8 537.4 4.0 103.9
050708‐AUNC‐48 320 19986 9.0 15.1012 5.2 0.7951 6.1 0.0871 3.2 0.53 538.3 16.7 594.1 27.6 813.4 109.0 538.3 16.7 66.2
050708‐AUNC‐49 282 8120 2.2 19.6097 2.8 0.2926 2.9 0.0416 0.6 0.22 262.9 1.6 260.6 6.7 240.6 65.5 262.9 1.6 109.3
CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION

050708‐AUNC‐50 193 18154 3.9 13.6823 1.6 1.6547 2.1 0.1642 1.4 0.64 980.1 12.3 991.4 13.4 1016.4 32.9 1016.4 32.9 96.4
050708‐AUNC‐51 377 14656 2.0 17.2716 1.3 0.6814 1.5 0.0854 0.7 0.49 528.0 3.7 527.6 6.1 525.9 28.3 528.0 3.7 100.4
050708‐AUNC‐53 505 20230 7.4 17.5961 1.3 0.5898 1.5 0.0753 0.7 0.44 467.9 3.0 470.8 5.6 485.0 29.5 467.9 3.0 96.5
050708‐AUNC‐54 237 9770 2.2 17.7330 1.6 0.5871 1.7 0.0755 0.7 0.39 469.3 3.1 469.0 6.5 467.8 35.2 469.3 3.1 100.3
050708‐AUNC‐55 397 24606 5.8 15.4115 4.7 0.8447 7.0 0.0944 5.2 0.74 581.6 28.8 621.7 32.5 770.6 99.0 581.6 28.8 75.5
050708‐AUNC‐56 446 35004 8.4 15.0197 2.1 0.9829 2.9 0.1071 2.1 0.71 655.7 13.0 695.1 14.8 824.6 43.5 655.7 13.0 79.5
050708‐AUNC‐58 386 18142 1.6 16.3304 1.6 0.8495 1.9 0.1006 1.0 0.54 618.0 6.0 624.4 8.8 647.5 34.0 618.0 6.0 95.4
050708‐AUNC‐59 248 13252 6.0 15.7770 3.0 0.7537 7.3 0.0862 6.7 0.91 533.3 34.2 570.4 32.0 721.1 63.8 533.3 34.2 74.0
050708‐AUNC‐60 297 28910 4.0 13.4216 1.8 1.6919 2.8 0.1647 2.2 0.77 982.8 19.6 1005.5 17.8 1055.3 35.9 1055.3 35.9 93.1
050708‐AUNC‐61 460 26582 13.5 17.7904 1.9 0.6036 2.0 0.0779 0.7 0.34 483.4 3.2 479.5 7.6 460.7 41.5 483.4 3.2 104.9
050708‐AUNC‐62 465 31740 5.9 14.4374 0.7 1.2154 2.3 0.1273 2.2 0.96 772.3 16.3 807.7 13.0 906.6 14.1 772.3 16.3 85.2
050708‐AUNC‐63 205 27948 1.4 13.4373 1.6 1.7904 1.9 0.1745 1.0 0.54 1036.8 10.0 1042.0 12.5 1052.9 32.5 1052.9 32.5 98.5
050708‐AUNC‐64 87 8706 2.5 15.3626 1.3 1.1927 1.5 0.1329 0.6 0.43 804.3 4.8 797.2 8.1 777.3 27.9 804.3 4.8 103.5
050708‐AUNC‐65 386 23894 13.9 16.3018 2.9 0.7138 3.9 0.0844 2.6 0.66 522.3 12.8 547.0 16.3 651.3 61.9 522.3 12.8 80.2
TC3003
Table A1. (continued)
Isotope Ratios Apparent Ages (Ma)
TC3003

U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration
Analysis (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%)
050708‐AUNC‐66 245 7354 2.5 17.7912 3.4 0.5876 3.6 0.0758 1.1 0.29 471.2 4.8 469.4 13.5 460.6 76.4 471.2 4.8 102.3
050708‐AUNC‐67 267 13110 4.0 17.0601 2.6 0.6770 3.3 0.0838 2.0 0.61 518.6 10.1 525.0 13.6 552.9 57.2 518.6 10.1 93.8
050708‐AUNC‐68 449 15368 7.2 18.1485 3.0 0.5712 3.1 0.0752 0.9 0.28 467.3 3.9 458.8 11.4 416.3 66.3 467.3 3.9 112.3
050708‐AUNC‐69 328 16520 7.1 17.6770 1.3 0.6221 1.8 0.0798 1.3 0.68 494.7 6.0 491.2 7.1 474.8 29.5 494.7 6.0 104.2
050708‐AUNC‐70 239 12174 2.5 16.7117 2.3 0.7943 3.0 0.0963 1.8 0.62 592.5 10.4 593.6 13.3 597.7 50.5 592.5 10.4 99.1
050708‐AUNC‐71 120 7484 1.7 16.4425 2.4 0.8759 5.3 0.1044 4.7 0.89 640.4 28.8 638.7 25.1 632.8 51.7 640.4 28.8 101.2
050708‐AUNC‐72 272 32092 2.9 14.4929 2.6 1.2640 3.4 0.1329 2.3 0.67 804.2 17.4 829.7 19.5 898.7 52.9 898.7 52.9 89.5
050708‐AUNC‐73 262 16164 9.7 17.8554 1.9 0.6198 2.2 0.0803 1.1 0.51 497.7 5.3 489.7 8.4 452.6 41.4 497.7 5.3 110.0
050708‐AUNC‐74 223 8832 5.6 17.8501 2.3 0.5943 2.4 0.0769 0.7 0.29 477.8 3.2 473.6 9.1 453.2 50.8 477.8 3.2 105.4
050708‐AUNC‐75 370 26196 9.7 13.4664 5.2 0.9627 6.8 0.0940 4.4 0.65 579.3 24.3 684.7 33.8 1048.5 104.3 579.3 24.3 55.2
050708‐AUNC‐76 277 10454 6.7 17.5396 2.3 0.5993 2.8 0.0762 1.7 0.60 473.6 7.7 476.8 10.8 492.1 49.9 473.6 7.7 96.2
050708‐AUNC‐77 189 14218 4.4 15.4779 3.1 0.8868 3.3 0.0995 1.1 0.34 611.8 6.7 644.6 15.8 761.6 65.7 611.8 6.7 80.3
050708‐AUNC‐78 376 18578 8.4 17.8661 3.2 0.5832 3.4 0.0756 1.2 0.35 469.6 5.4 466.5 12.8 451.2 71.0 469.6 5.4 104.1
050708‐AUNC‐79 264 7628 4.7 16.8421 5.1 0.6761 5.3 0.0826 1.4 0.26 511.5 6.7 524.4 21.7 580.9 111.0 511.5 6.7 88.1
050708‐AUNC‐80 348 4310 12.5 16.7121 3.4 0.6765 3.4 0.0820 0.6 0.17 508.0 2.8 524.7 14.1 597.7 73.3 508.0 2.8 85.0
050708‐AUNC‐81 247 14566 4.4 17.4333 2.4 0.6056 5.2 0.0766 4.7 0.89 475.6 21.5 480.8 20.1 505.5 51.8 475.6 21.5 94.1
050708‐AUNC‐82 239 9334 2.5 17.3488 3.1 0.6853 6.2 0.0862 5.4 0.87 533.2 27.4 530.0 25.5 516.1 67.4 533.2 27.4 103.3
050708‐AUNC‐83 230 8812 3.0 17.6661 2.7 0.6129 2.9 0.0785 1.0 0.33 487.4 4.5 485.4 11.1 476.2 60.2 487.4 4.5 102.3
050708‐AUNC‐84 248 14904 17.2 15.8897 4.0 0.7077 4.2 0.0816 1.2 0.27 505.4 5.6 543.4 17.6 706.0 85.8 505.4 5.6 71.6
050708‐AUNC‐85 284 13436 5.5 17.9244 2.7 0.6041 2.9 0.0785 1.0 0.34 487.3 4.6 479.8 11.1 444.0 60.8 487.3 4.6 109.8
050708‐AUNC‐86 147 3402 1.8 17.2524 6.5 0.6418 6.5 0.0803 1.1 0.16 497.9 5.1 503.4 26.0 528.4 141.7 497.9 5.1 94.2
050708‐AUNC‐87 213 13846 2.9 17.7121 4.1 0.6392 4.1 0.0821 0.5 0.12 508.7 2.4 501.8 16.3 470.5 90.6 508.7 2.4 108.1

24 of 30
050708‐AUNC‐88 242 17668 1.6 17.7922 3.0 0.6468 3.7 0.0835 2.1 0.57 516.8 10.4 506.5 14.6 460.4 66.5 516.8 10.4 112.2
050708‐AUNC‐90 257 13994 1.5 17.6028 1.7 0.6206 2.2 0.0792 1.5 0.66 491.5 7.1 490.2 8.7 484.1 37.2 491.5 7.1 101.5
050708‐AUNC‐91 230 14078 8.7 17.4607 2.8 0.6725 4.5 0.0852 3.5 0.78 526.9 17.7 522.2 18.2 502.0 60.9 526.9 17.7 105.0
050708‐AUNC‐92 311 10700 6.0 17.5084 2.4 0.5941 2.6 0.0754 1.1 0.43 468.8 5.1 473.5 9.9 496.0 52.4 468.8 5.1 94.5
050708‐AUNC‐93 272 16312 11.6 17.5718 3.1 0.6172 4.0 0.0787 2.5 0.62 488.1 11.6 488.1 15.3 488.0 68.3 488.1 11.6 100.0
050708‐AUNC‐94 170 5792 6.0 17.6977 2.6 0.6020 3.3 0.0773 2.1 0.64 479.8 9.9 478.5 12.8 472.2 56.9 479.8 9.9 101.6
050708‐AUNC‐95 255 11290 8.4 17.5228 2.4 0.6239 2.5 0.0793 0.7 0.28 491.9 3.3 492.3 9.8 494.2 52.9 491.9 3.3 99.5
050708‐AUNC‐96 327 17756 4.0 14.6760 3.0 0.9992 3.7 0.1064 2.1 0.58 651.6 13.3 703.4 18.6 872.7 61.5 651.6 13.3 74.7
050708‐AUNC‐97 491 3380 1.4 16.1641 4.4 0.6499 4.9 0.0762 2.3 0.46 473.3 10.4 508.4 19.7 669.5 93.4 473.3 10.4 70.7
050708‐AUNC‐98 180 4998 8.8 17.8253 4.3 0.6158 4.4 0.0796 0.9 0.20 493.8 4.2 487.2 17.0 456.3 95.5 493.8 4.2 108.2
050708‐AUNC‐99 194 9214 7.0 17.3671 2.8 0.6326 2.9 0.0797 0.8 0.28 494.3 3.9 497.7 11.3 513.8 60.6 494.3 3.9 96.2
050708‐AUNC‐100 259 10922 2.2 17.1253 2.1 0.6716 3.0 0.0834 2.2 0.73 516.5 11.0 521.7 12.3 544.5 44.9 516.5 11.0 94.9
CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION

SAMPLE 05.07.08‐B.Unc
050708BUNC‐1 280 76935 7.0 7.7557 1.5 5.9310 2.2 0.3336 1.6 0.73 1855.9 25.8 1965.8 18.9 2083.5 26.0 2083.5 26.0 89.1
050708BUNC‐2 474 23646 23.7 17.6252 0.7 0.5964 1.1 0.0762 0.8 0.77 473.6 3.8 475.0 4.1 481.3 15.1 473.6 3.8 98.4
050708BUNC‐4 285 13419 1.8 17.2050 1.8 0.7082 1.9 0.0884 0.5 0.27 545.9 2.6 543.7 7.9 534.4 39.3 545.9 2.6 102.2
050708BUNC‐5 426 25272 7.7 17.5059 1.3 0.6013 1.4 0.0763 0.5 0.37 474.3 2.3 478.1 5.2 496.3 28.0 474.3 2.3 95.6
050708BUNC‐6 259 15780 4.5 17.6861 1.4 0.6056 1.7 0.0777 0.8 0.51 482.3 3.9 480.8 6.3 473.7 31.5 482.3 3.9 101.8
050708BUNC‐8 116 18144 2.5 13.6948 2.4 1.7010 2.8 0.1690 1.5 0.52 1006.3 13.6 1008.9 17.8 1014.5 48.1 1014.5 48.1 99.2
050708BUNC‐9 386 18042 3.3 17.9257 2.5 0.5862 3.0 0.0762 1.7 0.56 473.5 7.6 468.4 11.3 443.8 55.5 473.5 7.6 106.7
050708BUNC‐10 567 25203 10.2 17.5751 1.7 0.5920 1.8 0.0755 0.6 0.32 469.0 2.7 472.2 6.9 487.6 38.5 469.0 2.7 96.2
050708BUNC‐11 249 14259 5.3 17.6516 2.1 0.6011 2.2 0.0770 0.5 0.23 477.9 2.3 477.9 8.3 478.0 46.8 477.9 2.3 100.0
050708BUNC‐12 425 39015 6.3 16.7757 1.4 0.7394 2.0 0.0900 1.4 0.71 555.3 7.6 562.0 8.7 589.4 30.8 555.3 7.6 94.2
050708BUNC‐13 115 9942 1.7 17.1931 2.3 0.6756 2.4 0.0842 0.7 0.27 521.4 3.3 524.1 10.0 535.9 51.2 521.4 3.3 97.3
050708BUNC‐14 395 23364 15.3 17.7090 1.5 0.5882 1.5 0.0756 0.5 0.32 469.5 2.3 469.7 5.8 470.9 32.4 469.5 2.3 99.7
050708BUNC‐17 511 68982 8.7 12.8935 0.8 1.6904 1.7 0.1581 1.5 0.87 946.1 13.1 1004.9 10.9 1135.6 16.7 1135.6 16.7 83.3
TC3003
Table A1. (continued)
Isotope Ratios Apparent Ages (Ma)
TC3003

U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration
Analysis (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%)
050708BUNC‐18 371 23505 6.9 17.7383 1.4 0.5933 2.8 0.0763 2.5 0.88 474.2 11.4 473.0 10.7 467.2 30.2 474.2 11.4 101.5
050708BUNC‐19 247 8385 1.0 17.0144 1.8 0.6357 1.9 0.0784 0.5 0.27 486.8 2.3 499.6 7.4 558.7 39.3 486.8 2.3 87.1
050708BUNC‐20 358 42780 2.1 9.1449 1.7 3.3879 2.1 0.2247 1.2 0.58 1306.7 14.2 1501.6 16.2 1788.6 30.6 1788.6 30.6 73.1
050708BUNC‐21 265 17937 6.4 17.7017 1.8 0.6074 2.0 0.0780 0.9 0.43 484.1 4.0 481.9 7.5 471.7 39.1 484.1 4.0 102.6
050708BUNC‐22 552 31956 11.8 17.6818 1.0 0.6035 1.1 0.0774 0.5 0.44 480.6 2.3 479.5 4.4 474.3 22.9 480.6 2.3 101.3
050708BUNC‐23 461 45312 3.1 16.5630 1.4 0.8586 1.6 0.1031 0.8 0.50 632.8 4.8 629.3 7.4 617.0 29.4 632.8 4.8 102.5
050708BUNC‐24 118 9879 0.9 16.4654 2.5 0.8888 2.8 0.1061 1.3 0.47 650.3 8.2 645.7 13.6 629.8 54.1 650.3 8.2 103.2
050708BUNC‐25 140 14154 2.4 17.2091 1.7 0.7114 1.8 0.0888 0.6 0.32 548.3 3.0 545.5 7.5 533.8 36.7 548.3 3.0 102.7
050708BUNC‐28 669 45591 13.4 17.5422 1.6 0.6118 1.6 0.0778 0.5 0.31 483.2 2.3 484.7 6.3 491.8 34.4 483.2 2.3 98.3
050708BUNC‐29 277 20115 2.5 17.3680 1.5 0.6773 1.7 0.0853 0.7 0.43 527.8 3.6 525.1 6.9 513.7 33.7 527.8 3.6 102.7
050708BUNC‐30 310 54408 5.3 13.6416 1.3 1.7016 1.8 0.1684 1.3 0.69 1003.0 11.7 1009.1 11.6 1022.4 26.5 1022.4 26.5 98.1
050708BUNC‐32 352 31059 37.0 16.9001 1.7 0.6792 2.8 0.0833 2.3 0.81 515.5 11.3 526.3 11.6 573.4 36.4 515.5 11.3 89.9
050708BUNC‐33 255 14148 0.9 17.0166 1.1 0.6978 1.6 0.0861 1.1 0.70 532.5 5.6 537.5 6.5 558.4 24.3 532.5 5.6 95.4
050708BUNC‐34 276 22923 3.9 17.4886 1.0 0.6217 1.4 0.0789 1.0 0.70 489.3 4.8 490.9 5.6 498.5 22.4 489.3 4.8 98.2
050708BUNC‐35 278 12243 3.4 15.7829 7.4 0.7020 7.5 0.0804 1.2 0.16 498.3 5.6 540.0 31.3 720.3 156.6 498.3 5.6 69.2
050708BUNC‐36 219 33819 2.1 13.1724 1.5 1.9420 1.6 0.1855 0.5 0.31 1097.1 5.0 1095.7 10.7 1092.9 30.4 1092.9 30.4 100.4
050708BUNC‐37 416 73827 2.2 13.3536 1.0 1.8865 1.1 0.1827 0.5 0.48 1081.7 5.2 1076.3 7.2 1065.5 19.3 1065.5 19.3 101.5
050708BUNC‐39 148 474 1.8 15.7071 164.6 0.0187 164.8 0.0021 8.0 0.05 13.7 1.1 18.8 30.8 730.6 948.5 13.7 1.1 1.9
050708BUNC‐40 364 24504 8.9 17.5488 1.6 0.6190 1.6 0.0788 0.5 0.33 488.9 2.5 489.2 6.4 490.9 34.2 488.9 2.5 99.6
050708BUNC‐41 235 19575 6.5 15.6843 2.5 0.8490 2.8 0.0966 1.2 0.43 594.3 6.8 624.1 12.8 733.6 52.6 594.3 6.8 81.0
050708BUNC‐42 655 34785 8.3 16.9981 2.1 0.6867 3.1 0.0847 2.4 0.75 523.9 11.8 530.8 13.0 560.8 45.6 523.9 11.8 93.4
050708BUNC‐43 433 33759 3.2 13.9348 1.6 1.3740 2.3 0.1389 1.7 0.72 838.2 13.2 877.9 13.6 979.2 32.6 979.2 32.6 85.6

25 of 30
050708BUNC‐44 464 32130 29.7 17.7250 1.0 0.6059 1.5 0.0779 1.2 0.77 483.5 5.5 481.0 5.9 468.8 21.8 483.5 5.5 103.1
050708BUNC‐45 307 21573 5.2 17.5269 1.3 0.6215 1.5 0.0790 0.8 0.52 490.2 3.7 490.8 5.9 493.7 28.4 490.2 3.7 99.3
050708BUNC‐47 181 26796 1.4 8.5878 1.7 3.8727 4.8 0.2412 4.5 0.93 1393.0 56.0 1608.0 38.7 1902.3 31.3 1902.3 31.3 73.2
050708BUNC‐48 249 18105 2.2 16.9366 0.7 0.7120 1.1 0.0875 0.8 0.74 540.5 4.3 545.9 4.7 568.7 16.3 540.5 4.3 95.0
050708BUNC‐49 372 25245 2.4 17.2198 1.1 0.6932 1.2 0.0866 0.5 0.42 535.3 2.6 534.7 5.0 532.5 23.7 535.3 2.6 100.5
050708BUNC‐50 199 14964 4.4 15.3176 3.9 0.8101 6.6 0.0900 5.4 0.81 555.5 28.7 602.5 30.2 783.5 81.4 555.5 28.7 70.9
050708BUNC‐51 259 24681 2.3 13.7677 1.6 1.4802 3.4 0.1478 3.1 0.89 888.6 25.3 922.3 20.8 1003.7 32.1 1003.7 32.1 88.5
050708BUNC‐52 365 20349 7.0 17.7178 1.3 0.6053 1.4 0.0778 0.5 0.37 482.8 2.3 480.6 5.2 469.7 27.8 482.8 2.3 102.8
050708BUNC‐53 80 4134 0.8 17.2900 2.6 0.6174 2.7 0.0774 0.7 0.27 480.7 3.3 488.2 10.3 523.6 56.2 480.7 3.3 91.8
050708BUNC‐55 200 49281 1.1 8.1528 1.8 6.2483 1.9 0.3695 0.5 0.26 2026.9 8.7 2011.2 16.5 1995.2 32.4 1995.2 32.4 101.6
050708BUNC‐57 223 22074 5.7 14.9694 5.6 0.8104 9.2 0.0880 7.2 0.79 543.6 37.7 602.7 41.6 831.6 117.1 543.6 37.7 65.4
050708BUNC‐58 283 28218 4.1 16.4449 1.3 0.8296 1.4 0.0989 0.5 0.36 608.2 2.9 613.4 6.4 632.5 28.1 608.2 2.9 96.2
CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION

050708BUNC‐59 327 33927 9.2 16.3294 1.1 0.7782 1.8 0.0922 1.4 0.77 568.3 7.4 584.4 7.9 647.7 24.5 568.3 7.4 87.7
050708BUNC‐60 398 20967 1.7 17.4177 1.6 0.6601 1.8 0.0834 0.9 0.46 516.3 4.2 514.7 7.5 507.4 36.1 516.3 4.2 101.8
050708BUNC‐61 470 33729 12.2 17.5233 1.6 0.6253 1.8 0.0795 0.8 0.45 492.9 3.7 493.1 6.9 494.1 34.9 492.9 3.7 99.8
050708BUNC‐62 134 11487 2.5 17.1414 1.5 0.6766 1.5 0.0841 0.5 0.33 520.7 2.5 524.7 6.3 542.5 31.7 520.7 2.5 96.0
050708BUNC‐63 171 37545 1.1 8.8016 2.6 4.3080 8.0 0.2750 7.6 0.94 1566.1 105.3 1694.9 66.1 1858.0 47.5 1858.0 47.5 84.3
050708BUNC‐64 366 19380 12.4 17.3965 2.4 0.6234 3.2 0.0787 2.2 0.66 488.1 10.1 492.0 12.6 510.1 53.1 488.1 10.1 95.7
050708BUNC‐66 141 12669 1.8 16.8880 1.6 0.7615 1.8 0.0933 0.9 0.48 574.9 4.7 574.9 7.8 575.0 34.0 574.9 4.7 100.0
050708BUNC‐67 347 16803 2.2 16.9923 2.1 0.7346 2.9 0.0905 2.1 0.70 558.7 11.0 559.2 12.6 561.5 45.4 558.7 11.0 99.5
050708BUNC‐68 459 37221 2.0 17.1283 0.8 0.7086 1.0 0.0880 0.6 0.57 543.9 3.0 543.9 4.3 544.2 18.1 543.9 3.0 99.9
050708BUNC‐69 173 12864 2.3 16.8862 1.8 0.7274 1.9 0.0891 0.5 0.27 550.1 2.6 555.0 8.0 575.2 39.0 550.1 2.6 95.6
050708BUNC‐70 330 30768 4.5 14.6481 1.4 1.0640 3.3 0.1130 3.0 0.90 690.4 19.8 735.8 17.5 876.7 29.4 690.4 19.8 78.7
050708BUNC‐71 273 16164 5.7 16.3230 2.3 0.7456 2.4 0.0883 0.5 0.21 545.3 2.6 565.7 10.2 648.5 49.4 545.3 2.6 84.1
050708BUNC‐72 269 18831 3.5 16.8751 3.0 0.7187 3.2 0.0880 1.1 0.35 543.5 5.8 549.9 13.4 576.6 64.4 543.5 5.8 94.3
050708BUNC‐73 516 18051 1.5 17.1880 3.8 0.6302 3.9 0.0786 0.5 0.13 487.6 2.4 496.2 15.1 536.6 83.6 487.6 2.4 90.9
TC3003
TC3003

Table A1. (continued)


Isotope Ratios Apparent Ages (Ma)
U 206Pb/ 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Error 206Pb*/ Standard 207Pb*/ Standard 206Pb*/ Standard Best Age Standard Concentration
Analysis (ppm) 204Pb U/Th 207Pb* Error (%) 235U* Error (%) 238U Error (%) Correlation 238U* Error (Ma) 235U Error (Ma) 207Pb* Error (Ma) (Ma) Error (Ma) (%)
050708BUNC‐74 395 17214 11.7 17.5322 0.8 0.6216 1.7 0.0790 1.5 0.87 490.4 6.9 490.9 6.5 493.0 17.9 490.4 6.9 99.5
050708BUNC‐75 486 336 1.4 14.8943 30.9 0.0205 31.1 0.0022 3.7 0.12 14.2 0.5 20.6 6.3 842.1 658.2 14.2 0.5 1.7
050708BUNC‐76 192 18984 1.0 16.1942 1.5 0.9417 1.6 0.1106 0.7 0.41 676.3 4.4 673.8 8.1 665.5 32.0 676.3 4.4 101.6
050708BUNC‐77 302 22521 7.6 17.6298 1.8 0.5972 1.9 0.0764 0.5 0.27 474.4 2.3 475.5 7.1 480.7 39.6 474.4 2.3 98.7
050708BUNC‐78 218 12054 6.2 17.8553 2.5 0.5877 2.7 0.0761 0.9 0.34 472.9 4.2 469.4 10.1 452.6 56.1 472.9 4.2 104.5
050708BUNC‐79 483 26226 4.9 17.5287 2.8 0.5967 3.5 0.0759 2.1 0.61 471.3 9.7 475.1 13.2 493.4 60.7 471.3 9.7 95.5
050708BUNC‐80 321 16080 7.7 17.5970 2.5 0.5971 3.2 0.0762 2.0 0.64 473.5 9.2 475.4 12.1 484.9 54.3 473.5 9.2 97.6
050708BUNC‐82 290 25293 10.6 16.2117 2.9 0.7835 5.3 0.0921 4.5 0.84 568.0 24.5 587.5 23.8 663.1 61.5 568.0 24.5 85.7
050708BUNC‐83 200 19248 2.7 14.5466 1.2 1.2359 1.3 0.1304 0.5 0.38 790.1 3.7 817.0 7.4 891.1 25.4 790.1 3.7 88.7
050708BUNC‐84 339 10821 1.3 17.4906 2.0 0.6102 2.5 0.0774 1.5 0.60 480.6 7.0 483.7 9.7 498.2 44.6 480.6 7.0 96.5
050708BUNC‐85 378 23430 10.1 17.4422 1.0 0.6212 1.2 0.0786 0.5 0.44 487.7 2.4 490.6 4.5 504.4 22.8 487.7 2.4 96.7
050708BUNC‐86 242 29439 1.4 15.9145 2.4 0.9442 2.7 0.1090 1.3 0.47 666.9 8.0 675.1 13.3 702.7 50.7 666.9 8.0 94.9
050708BUNC‐87 349 24180 9.1 17.7149 2.6 0.5862 2.7 0.0753 0.9 0.31 468.1 3.8 468.4 10.2 470.1 57.0 468.1 3.8 99.6
050708BUNC‐88 184 24306 15.0 16.6221 2.4 0.7210 2.5 0.0869 0.8 0.30 537.3 3.9 551.2 10.6 609.4 51.3 537.3 3.9 88.2
050708BUNC‐89 187 17742 6.4 17.2460 1.8 0.6964 1.8 0.0871 0.5 0.27 538.4 2.6 536.6 7.6 529.2 38.5 538.4 2.6 101.7
050708BUNC‐90 335 23694 10.2 17.6640 1.5 0.5901 1.6 0.0756 0.5 0.32 469.8 2.4 470.9 6.1 476.5 33.7 469.8 2.4 98.6
050708BUNC‐91 313 22773 2.5 17.6597 1.9 0.6064 2.0 0.0777 0.5 0.25 482.2 2.3 481.3 7.6 477.0 42.5 482.2 2.3 101.1
050708BUNC‐92 125 8958 1.2 17.0775 1.8 0.7071 1.9 0.0876 0.6 0.32 541.2 3.2 543.0 8.1 550.7 39.9 541.2 3.2 98.3

26 of 30
050708BUNC‐93 423 11916 19.3 16.4528 1.1 0.7474 1.3 0.0892 0.6 0.49 550.7 3.3 566.7 5.5 631.5 23.5 550.7 3.3 87.2
050708BUNC‐94 182 14928 1.4 17.1228 1.8 0.6973 1.8 0.0866 0.5 0.29 535.3 2.8 537.1 7.7 544.9 38.6 535.3 2.8 98.3
050708BUNC‐95 252 14994 4.5 15.6316 4.8 0.7380 5.2 0.0837 2.0 0.37 518.0 9.7 561.2 22.5 740.7 102.5 518.0 9.7 69.9
050708BUNC‐97 297 12810 1.7 17.3148 3.8 0.6203 3.8 0.0779 0.7 0.18 483.6 3.3 490.1 15.0 520.4 83.0 483.6 3.3 92.9
050708BUNC‐98 427 13593 1.7 16.9084 2.4 0.6288 2.5 0.0771 0.6 0.23 478.8 2.6 495.3 9.7 572.3 52.4 478.8 2.6 83.7
050708BUNC‐99 166 66840 1.1 8.5669 1.3 5.4105 1.5 0.3362 0.8 0.55 1868.2 13.5 1886.5 12.9 1906.7 22.6 1906.7 22.6 98.0
050708BUNC‐100 182 15225 2.3 17.2265 2.1 0.6797 2.2 0.0849 0.5 0.24 525.4 2.6 526.6 9.0 531.7 46.8 525.4 2.6 98.8
SAMPLE 04.27.08‐AngB
042708AngB‐1 401 20496 6.0 16.5795 0.6 0.7653 1.8 0.0920 1.7 0.94 567.5 9.2 577.1 7.9 614.9 13.4 567.5 9.2 92.3
042708AngB‐2 122 5986 2.3 17.3517 1.6 0.6579 1.7 0.0828 0.6 0.37 512.8 3.2 513.3 7.0 515.8 35.5 512.8 3.2 99.4
042708AngB‐3 166 9066 2.1 17.6271 2.4 0.6529 3.0 0.0835 1.8 0.61 516.8 8.9 510.3 11.9 481.1 52.3 516.8 8.9 107.4
CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION

042708AngB‐4 125 6376 1.7 18.0159 2.1 0.5722 2.4 0.0748 1.2 0.51 464.8 5.4 459.4 8.8 432.7 46.0 464.8 5.4 107.4
042708AngB‐5 356 13638 3.6 17.9001 1.1 0.5744 1.4 0.0746 0.9 0.66 463.6 4.2 460.9 5.3 447.0 23.9 463.6 4.2 103.7
042708AngB‐7 112 4840 1.9 18.2966 2.5 0.5270 2.7 0.0699 0.8 0.30 435.8 3.4 429.8 9.3 398.1 57.0 435.8 3.4 109.5
042708AngB‐9 297 12774 3.4 18.1668 1.8 0.5169 1.8 0.0681 0.5 0.27 424.7 2.1 423.1 6.4 414.0 39.6 424.7 2.1 102.6
042708AngB‐11 206 9692 2.6 17.7293 1.6 0.5927 2.1 0.0762 1.4 0.65 473.5 6.3 472.6 8.1 468.3 35.9 473.5 6.3 101.1
042708AngB‐12 383 19520 2.1 17.4357 1.1 0.6660 2.1 0.0842 1.8 0.84 521.3 8.9 518.3 8.5 505.1 25.1 521.3 8.9 103.2
042708AngB‐13 144 5768 4.6 17.9697 2.3 0.5917 3.3 0.0771 2.4 0.73 478.9 11.2 472.0 12.5 438.4 50.4 478.9 11.2 109.2
042708AngB‐14 230 10998 5.8 15.5875 2.0 0.7536 2.3 0.0852 1.2 0.53 527.0 6.2 570.3 10.1 746.7 41.3 527.0 6.2 70.6
042708AngB‐17 236 12830 2.8 17.9512 29.3 0.6192 29.3 0.0806 0.6 0.02 499.8 2.9 489.4 114.4 440.7 665.6 499.8 2.9 113.4
042708AngB‐19 99 6424 1.7 17.9111 3.2 0.6199 3.5 0.0805 1.3 0.38 499.3 6.4 489.8 13.4 445.6 71.0 499.3 6.4 112.0
042708AngB‐20 185 8162 3.3 18.0274 2.6 0.6019 2.6 0.0787 0.6 0.21 488.3 2.6 478.4 10.0 431.2 57.4 488.3 2.6 113.2
TC3003
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

Figure A1. Complete time‐temperature paths of six AHe samples modeled using the HeFTy computer
program. Time and temperature modeling constraints are outlined in blue, with gradual (2G) and episodic
(2E) modeling parameters indicated. Purple areas represent halos.

logic time, He retentivity is recognized as an evolving the sample spent and extended period of time in the HePRZ
property of apatite that varies as a function of [eU], espe- [Flowers et al., 2007]. During radiogenic decay, emitted a
cially in slowly cooled rocks [Shuster et al., 2006; Flowers particles can travel up to 20 mm through the apatite crystal
et al., 2009]. Because slow cooling allows more time for He lattice, and some are ejected from crystal altogether. For
diffusion, it enhances the effect of radiation damage, and grains <200 mm in radius, this leads to significant He loss
thus the presence of a large spread in AHe dates that and anomalously young ages [Farley, 2002; Ehlers and
correlate to [eU] within a single sample is an indication that Farley, 2003].

27 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

[71] Because a ejection is largely dependent on the size Carrapa, B., B. Hauer, L. Schoenbohm, M. R. Strecker, A. K. Schmitt,
and surface to volume ratio of the crystal [Reiners and A. Villanueva, and A. S. Gomez (2008a), Dynamics of deformation
and sedimentation in the northern Sierras Pampeanas: An integrated
Farley, 2001], the dimensions and morphology of an apa- study of the Neogene Fiambala basin, NW Argentina, Geol. Soc. Am.
tite grain are routinely used with the Ft correction method to Bull., 120, 1518–1543, doi:10.1130/B26111.1.
calculate the amount of 4He lost by a ejection and correct Carrapa, B., L. Schoenbohm, M. Clementz, S. Bywater, and J. Quade
the AHe age [Farley et al., 1996]. In Figure A1 we present (2008b), Oxygen isotope evidence from Cenozoic paleosol carbonates
on the Puna Plateau of NW Argentina: Low or dry in the Neogene?,
the complete time‐temperature paths of six AHe samples Eos Trans. AGU, 89(53), Fall Meet. Suppl., Abstract T42A–03.
(see Figure 10) modeled using HeFTy [Ketcham, 2005], the Carrapa, B., L. Schoenbhom, P. G. DeCelles, M. Clementz, and
radiation trapping model of Shuster et al. [2006] (Do/a2), K. Hungtington (2009), Surface response to lithospheric delamination:
An example from the Puna Plateau of NW Argentina, Geol. Soc. Am.
and the static alpha ejection correction of Farley et al. Abstr. Programs, 41, 516.
[1996]. Carrera, N., and J. A. Muñoz (2008), Thrusting evolution in the southern
Cordillera Oriental (northern Argentine Andes): Constraints from growth
[72] Acknowledgments. This project was made possible through strata, Tectonophysics, 459, 107–122, doi:10.1016/j.tecto.2007.11.068.
generous support from NSF EAR 0710724. The KU (U‐Th)/He laboratory Carrera, N., J. A. Muñoz, F. Sàbat, R. Mon, and E. Roca (2006), The role
acknowledges support from NSF EAR 0414817. The Arizona LaserChron of inversion tectonics in the structure of the Cordillera Oriental (NW
Center is supported by NSF EAR 0443387 and NSF EAR 0732436. We Argentinean Andes), J. Struct. Geol., 28, 1921–1932, doi:10.1016/j.jsg.
thank Peter DeCelles, George Gehrels, and Lindsay Schoenbohm for scien- 2006.07.006.
tific inputs and Chris Krugh, Roman Kislitsyn, and Jamey Stutz for techni- Coutand, I., P. R. Cobbold, M. de Urreiztieta, P. Gautier, A. Chauvin,
cal and field support. D. Gapais, E. A. Rossello, and O. Lòpez‐Gamundí (2001), Style and his-
tory of Andean deformation, Puna Plateau, northwestern Argentina, Tec-
tonics, 20, 210–234.
Coutand, I., B. Carrapa, A. Deeken, A. K. Schmitt, E. Sobel, and M. R.
References Strecker (2006), Orogenic plateau formation and lateral growth of
Abascal, L. V. (2005), Combined thin‐skinned and thick‐skinned deforma- compressional basins and ranges: Insights from sandstone petrography
tion in the central Andean foreland of northwestern Argentina, J. South and detrital apatite fission‐track thermochronology in the Angastaco
Am. Earth Sci., 19, 75–81, doi:10.1016/j.jsames.2005.01.004. Basin, NW Argentina, Basin Res., 18, 1–26, doi:10.1111/j.1365-2117.
Acocella, V., L. Vezzoli, R. Omarini, M. Matteini, and R. Mazzuoli (2007), 2006.00283.x.
Kinematic variations across the Eastern Cordillera at 24°S (central Cristallini, E., A. H. Cominguez, and V. A. Ramos (1997), Deep structure
Andes): Tectonic and magmatic implications, Tectonophysics, 434, 81– of the Metan‐Guachipas region: Tectonic inversion in northwestern
92, doi:10.1016/j.tecto.2007.02.001. Argentina, J. South Am. Earth Sci., 10, 403–421, doi:10.1016/S0895-
Allmendinger, R. W., and T. Gubbels (1996), Pure and simple shear plateau 9811(97)00026-6.
uplift, Altiplano‐Puna, Argentina and Bolivia, Tectonophysics, 259, 1–13, Cristallini, E. O., A. H. Cominguez, V. A. Ramos, and E. D. Mercerat
doi:10.1016/0040-1951(96)00024-8. (2004), Basement double1136 wedge thrusting in the northern Sierras
Allmendinger, R. W., V. A. Ramos, T. E. Jordan, M. Palma, and B. L. Pampeanas of Argentina (27 degrees S); constraints from deep seismic
Isacks (1983), Paleogeography and Andean structural geometry, north- reflection, in Thrust Tectonics and Hydrocarbon Systems, edited by
west Argentina, Tectonics, 2, 1–16, doi:10.1029/TC002i001p00001. K. R. McClay, AAPG Mem., 82, 65–90.
Arriagada, C., P. R. Cobbold, and P. Roperch (2006), Salar de Atacama Dahlstrom, C. D. A. (1970), Structural geology in the eastern margin of the
basin: A record of compressional tectonics in the central Andes Canadian Rocky Mountains, Bull. Can. Pet. Geol., 18, 332–406.
since the mid‐Cretaceous, Tectonics, 25, TC1008, doi:10.1029/ Davis, D., J. Suppe, and F. A. Dahlen (1983), Mechanics of fold‐and‐thrust
2004TC001770. belts and accretionary wedges, J. Geophys. Res., 88, 1153–1172,
Bally, A. W., P. L. Gordy, and G. A. Stewart (1966), Structure, seismic doi:10.1029/JB088iB02p01153.
data and orogenic evolution of southern Canadian Rocky Mountains, DeCelles, P. G., and P. C. DeCelles (2001), Rates of shortening, propaga-
Bull. Can. Pet. Geol., 14, 337–381. tion, underthrusting, and flexural wave migration in continental orogenic
Barnes, J. B., and T. A. Ehlers (2009), End member models for Andean systems, Geology, 29, 135–138, doi:10.1130/0091-7613(2001)
Plateau uplift, Earth Sci. Rev., 97, 105–132, doi:10.1016/j.earscirev. 029<0135:ROSPUA>2.0.CO;2.
2009.08.003. DeCelles, P. G., and B. K. Horton (2003), Early to middle Tertiary foreland
Barnes, J. B., T. A. Ehlers, N. McQuarrie, P. B. O’Sullivan, and S. Tawackoli basin development and the history of Andean crustal shortening in Boli-
(2008), Thermochronometer record of Central Andean Plateau growth, via, Geol. Soc. Am. Bull., 115, 58–77, doi:10.1130/0016-7606(2003)
Bolivia (19.5°S), Tectonics, 27, TC3003, doi:10.1029/2007TC002174. 115<0058:ETMTFB>2.0.CO;2.
Biswas, S., I. Coutand, D. Grujic, C. Hager, D. Stockli, and B. Grasemann DeCelles, P. G., B. Carrapa, and G. E. Gehrels (2007), Detrital zircon U‐Pb
(2007), Exhumation and uplift of the Shillong Plateau and its influence ages provide provenance and chronostratigraphic information from
on the eastern Himalayas: New constraints from apatite and zircon (U‐ Eocene synorogenic deposits in northwestern Argentina, Geology, 35,
Th‐[Sm])/He and apatite fission track analyses, Tectonics, 26, TC6013, 323–326, doi:10.1130/G23322A.1.
doi:10.1029/2007TC002125. DeCelles, P. G., B. Carrapa, B. Horton, and D. Starck (2008), Foreland
Bosio, P. P., J. Powell, C. del Papa, and F. Hongn (2009), Middle Eocene basin evolution in NW Argentina and implications for timing of Andean
deformation– sedimentation in the Luracatao Valley: Tracking the begin- orogenesis, paper presented at XVII Argentinean Geological Conference,
ning of the foreland basin of northwestern Argentina, J. South Am. Earth Int. Union of Geol. Sci., Jujuy, Argentina.
Sci., 28, 142–154, doi:10.1016/j.jsames.2009.06.002. DeCelles, P. G., M. N. Ducea, P. Kapp, and G. Zandt (2009), Cyclicity in
Bywater‐Reyes, S., B. Carrapa, M. Clementz, and L. Schoenbohm (2010), Cordilleran orogenic systems, Nat. Geosci., 2, 251–257, doi:10.1038/
The effect of late Cenozoic aridification on sedimentation in the Eastern ngeo469.
Cordillera of NW Argentina (Angastaco basin), Geology, 38, 235–238, Deeken, A., E. R. Sobel, I. Coutand, M. Haschke, U. Riller, and M. R.
doi:10.1130/G30532.1. Strecker (2006), Development of the southern Eastern Cordillera, NW
Cahill, T. A., and B. L. Isacks (1992), Seismicity and shape of the sub- Argentina, constrained by apatite fission track thermochronology: From
ducted Nazca Plate, J. Geophys. Res., 97, 17,503–17,529, doi:10.1029/ early Cretaceous extension to middle Miocene shortening, Tectonics,
92JB00493. 25, TC6003, doi:10.1029/2005TC001894.
Carrapa, B., and P. G. DeCelles (2008), Eocene exhumation and basin Dickinson, W. R., and G. E. Gehrels (2009), Use of U‐Pb ages of detrital
development in the Puna of northwestern Argentina, Tectonics, 27, zircons to infer maximum depositional ages of strata: A test against a
TC1015, doi:10.1029/2007TC002127. Colorado Plateau Mesozoic database, Earth Planet. Sci. Lett., 288,
Carrapa, B., D. Adelmann, G. E. Hilley, E. Mortimer, E. R. Sobel, and 115–125, doi:10.1016/j.epsl.2009.09.013.
M. R. Strecker (2005), Oligocene uplift and development of plateau mor- Echavarria, R., R. Hernandez, R. W. Allmendinger, and J. H. Reynolds
phology in the southern central Andes, Tectonics, 24, TC4011, (2003), Sub‐Andean thrust and fold belt of northwest Argentina: Geom-
doi:10.1029/2004TC001762. etry and timing of the Andean evolution, AAPG Bull., 87, 965–985,
Carrapa, B., E. Sobel, and M. R. Strecker (2006), Cenozoic orogenic doi:10.1306/01200300196.
growth in the central Andes: Evidence from rock provenance and apatite Ege, H., E. R. Sobel, E. Scheuber, and V. Jacobshagen (2007), Exhumation
fission track thermochronology along the southernmost Puna Plateau history of the southern Altiplano plateau (southern Bolivia) constrained
margin (NW Argentina), Earth Planet. Sci. Lett., 247, 82–100,
doi:10.1016/j.epsl.2006.04.010.

28 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

by apatite fission track thermochronology, Tectonics, 26, TC1004, presented at XI Congreso Geologico Chileno, Dep. de Cienc. Geol.,
doi:10.1029/2005TC001869. Univ. Cat. del Norte, Antofagasta, Chile.
Ehlers, T. A., and K. A. Farley (2003), Apatite 1169(U‐Th)/He thermo- Kay, S. M., and B. L. Coira (2009), Shallowing and steepening subduction
chronometry: Methods and applications to problems in tectonic and zones, continental lithospheric loss, magmatism, and crustal flow under
surface processes, Earth Planet. Sci. Lett., 206, 1–14, doi:10.1016/ the Central Andean Altiplano‐Puna Plateau, Mem. Geol. Soc. Am.,
S0012-821X(02)01069-5. 204, 229–259.
Ehlers, T. A., and C. J. Poulsen (2009), Large paleoclimate influence the Kay, S. M., C. Mpodozis, and B. Coira (1999), Magmatism, tectonism,
intepretation of Andean Plateau paleoaltimetry, Earth Planet. Sci. Lett., and mineral deposits of the central Andes (22–33°S), in Geology and
281, 238–248. Ore Deposits of the Central Andes, edited by B. Skinner, Spec. Publ.
Elger, K., O. Oncken, and J. Glodny (2005), Plateau‐style accumulation of Soc. Econ. Geol., 7, 27–59.
deformation: Southern Altiplano, Tectonics, 24, TC4020, doi:10.1029/ Ketcham, R. A. (2005), Forward and inverse modeling of low‐temperature
2004TC001675. thermochronometry data, in Reviews in Mineralogy and Geochemistry,
Farley, K. A. (2002), (U‐Th)/He dating: Techniques, calibrations, and vol. 58, edited by P. W. Reiners and T. A. Ehlers, pp. 275–314, Mineral.
applications, in Noble Gases in Geochemistry and Cosmochemistry, edi- Soc. of Am., Washington, D. C.
ted by D. Porcelli et al., pp. 819–844, Mineral. Soc. of Am., Washington, Kley, J. (1999), Geologic and geometric constraints on a kinematic model
D. C. of the Bolivian orocline, J. South Am. Earth Sci., 12, 221–235,
Farley, K. A., R. A. Wolf, and L. T. Silver (1996), The effects of long alpha doi:10.1016/S0895-9811(99)00015-2.
stopping distances on (U‐Th)/He ages, Geochim. Cosmochim. Acta, 60, Kley, J., and C. R. Monaldi (1998), Tectonic shortening and crustal thick-
4223–4229, doi:10.1016/S0016-7037(96)00193-7. ness in the central Andes: How good is the correlation?, Geology, 26,
Flowers, R. M., D. L. Shuster, B. P. Wernicke, and K. A. Farley (2007), 723–726.
Radiation damage control on apatite (U‐Th)/He dates from the Grand Kley, J., and C. R. Monaldi (2002), Tectonic inversion in the Santa Barbara
Canyon region, Colorado Plateau, Geology, 35, 447–450, doi:10.1130/ System of the central Andean foreland thrust belt, northwestern Argenti-
G23471A.1. na, Tectonics, 21(6), 1061, doi:10.1029/2002TC902003.
Flowers, R. M., R. A. Ketcham, D. L. Shuster, and K. A. Farley (2009), Kley, J., J. Müller, S. Tawackoli, V. Jacobshagen, and E. Manutsoglu
Apatite (U‐Th)/He thermochronometry using a radiation damage accu- (1997), Pre‐Andean and Andean age deformation in the eastern Cordillera
mulation and annealing model, Geochim. Cosmochim. Acta, 73, 2347– of southern Bolivia, J. South Am. Earth Sci., 10, 1–19, doi:10.1016/
2365, doi:10.1016/j.gca.2009.01.015. S0895-9811(97)00001-1.
Garzione, C. N., P. Molnar, J. C. Libarkin, and B. J. MacFadden (2006), Kley, J., C. R. Monaldi, and J. A. Salfity (1999), Along‐strike segmentation
Rapid late Miocene rise of the Bolivian Altiplano: Evidence for removal of the Andean foreland: Causes and consequences, Tectonophysics, 301,
of mantle lithosphere, Earth Planet. Sci. Lett., 241, 543–556, 75–94, doi:10.1016/S0040-1951(98)90223-2.
doi:10.1016/j.epsl.2005.11.026. Kraemer, B., D. Adelmann, M. Alten, W. Schnurr, K. Erpenstein, E. Kiefer,
Gehrels, G. E., V. A. Valencia, and J. Ruiz (2008), Enhanced precision, P. van den Bogaard, and K. Görler (1999), Incorporation of the Paleogene
accuracy, efficiency, and spatial resolution of U‐Pb ages by laser ablation foreland into the Neogene Puna Plateau: The Salar de Antofalla area, NW
multicollector inductively coupled plasma mass spectrometry, Geochem. Argentina, J. South Am. Earth Sci., 12, 157–182, doi:10.1016/S0895-
Geophys. Geosyst., 9, Q03017, doi:10.1029/2007GC001805. 9811(99)00012-7.
Gillis, R. J., B. K. Horton, and M. Grove (2006), Thermochronology, geo- Ludwig, K. R. (2008), Isoplot 3.6, Spec. Publ., 4, 77 pp., Berkeley Geo-
chronology, and upper crustal structure of the Cordillera Real: Implica- chronol. Cent., Berkeley, Calif.
tions for Cenozoic exhumation of the central Andean plateau, Marquillas, R. A., C. del Papa, and I. F. Sabino (2005), Sedimentary
Tectonics, 25, TC6007, doi:10.1029/2005TC001887. aspects and paleoenvironmental evolution of a rift basin: Salta Group
Gray, M. B., and P. K. Zeitler (1997), Comparison of clastic wedge prov- (Cretaceous‐Paleogene), northwestern Argentina, Int. J. Earth Sci., 94(1),
enance in the Appalachian foreland using U/Pb ages of detrital zircons, 94–113, doi:10.1007/s00531-004-0443-2.
Tectonics, 16, 151–160, doi:10.1029/96TC02911. Marrett, R. A., R. W. Allmendinger, R. N. Alonso, and R. E. Drake (1994),
Grier, M. E., and R. D. Dallmeyer (1990), Age of the Payogastilla Group: Late Cenozoic tectonic evolution of the Puna Plateau and adjacent foreland,
Implication for foreland basin development, NW Argentina, J. South Am. northwestern Argentine Andes, J. South Am. Earth Sci., 7, 179–207,
Earth Sci., 3, 269–278, doi:10.1016/0895-9811(90)90008-O. doi:10.1016/0895-9811(94)90007-8.
Grier, M. E., J. A. Salfity, and R. W. Allmendinger (1991), Andean reac- McQuarrie, N. (2002), The kinematic history of the central Andean fold‐
tivation of the Cretaceous Salta rift, northwestern Argentina, J. South Am. thrust belt, Bolivia: Implications for building a high plateau, Geol. Soc.
Earth Sci., 4, 351–372, doi:10.1016/0895-9811(91)90007-8. Am. Bull., 114, 950–963, doi:10.1130/0016-7606(2002)114<0950:
Gubbels, T. L., B. L. Isacks, and E. Farrar (1993), High‐level surfaces, pla- TKHOTC>2.0.CO;2.
teau uplift, and foreland development, Bolivian central Andes, Geology, McQuarrie, N., B. K. Horton, G. Zandt, S. Beck, and P. G. DeCelles
21, 695–698, doi:10.1130/0091-7613(1993)021<0695:HLSPUA>2.3. (2005), Lithospheric evolution of the Andean fold‐thrust belt, Bolivia,
CO;2. and the origin of the Central Andean Plateau, Tectonophysics, 399,
Hongn, F. D., and R. E. Seggiaro (2001), Hoja Geologica Cachi, 2566–III, 15–37, doi:10.1016/j.tecto.2004.12.013.
Provincias de Salta y Catamarca, Republica Argentina, map, Bull. 248, McQuarrie, N., J. B. Barnes, and T. A. Ehlers (2008), Geometric, kine-
87 pp., Inst. de Geol. y Recursos Miner., Serv. de Geol. y Min. de matic, and erosional history of the Central Andean Plateau, Bolivia
Argent., Buenos Aires. (15–17°S), Tectonics, 27, TC3007, doi:10.1029/2006TC002054.
Hongn, F., C. del Papa, J. Powell, I. Petrinovic, R. Mon, and V. Deraco Mon, R., and G. Drozdzewski (1999), Cinturones doble vergentes en los
(2007), Middle Eocene deformation and sedimentation in the Puna‐Eastern Andes del norte argentine: Hipótesis sobre su origen, Asoc. Geol. Argent.
Cordillera transition (23–26°S): Control by preexisting heterogeneities on Rev., 54(1), 3–8.
the pattern of initial Andean shortening, Geology, 35, 271–274, Mon, R., and J. A. Salfity (1995), Tectonic evolution of the Andes of north-
doi:10.1130/G23189A.1. ern Argentina, in Petroleum Basins of South America, edited by A. J.
Isacks, B. L. (1988), Uplift of the Central Andean Plateau and bending Tankard et al., AAPG Mem., 62, 269–283.
of the Bolivian Orocline, J. Geophys. Res., 93, 3211–3231, doi:10.1029/ Mortimer, E., B. Carrapa, C. Isabelle, L. Schoenbohm, E. R. Sobel,
JB093iB04p03211. J. S. Gomez, and M. R. Strecker (2007), Fragmentation of a foreland
Jezek, P., and H. Miller (1985), Deposition and facies distribution of turbi- basin in response to out‐of‐sequence basement uplifts and structural
ditic sediments of the Puncoviscana Formation (upper Precambrian‐ reactivation: El Cajón‐Campo del Arenal basin, NW Argentina, Geol.
Lower Cambrian) within the basement of the NW Argentine Andes, Soc. Am. Bull., 119, 637–653, doi:10.1130/B25884.1.
Zentralbl. Geol. Palaeontol., Teil I, 9–10, 1235–1244. Mpodozis, C., C. Arriagada, M. Basso, P. Roperch, P. Cobbold, and
Jordan, T. E. (1995), Retroarc foreland and related basins, in Tectonics of M. Reich (2005), Late Mesozoic to Paleogene stratigraphy of the Salar
Sedimentary Basins, edited by C. J. Busby and R. V. Ingersoll, pp. 331– de Atacama Basin, Antofagasta, Northern Chile: Implications for the tec-
362, Blackwell Sci., Oxford, U. K. tonics evolution of the central Andes, Tectonophysics, 399, 125–154,
Jordan, T. E., and R. W. Allmendinger (1986), The Sierras Pampeanas of doi:10.1016/j.tecto.2004.12.019.
Argentina: A modern analogue of Rocky Mountain foreland deformation, Norabuena, E., L. Leffler‐Griffin, A. Mao, T. Dixon, S. Stein, I. S. Sacks,
Am. J. Sci., 286, 737–764, doi:10.2475/ajs.286.10.737. L. Ocola, and M. Ellis (1998), Space geodetic observations of Nazca‐
Jordan, T. E., and R. N. Alonso (1987), Cenozoic stratigraphy and basin South America convergence across the central Andes, Science, 279,
tectonics of the Andes Mountains, 20°–28° South latitude, AAPG Bull., 358–362, doi:10.1126/science.279.5349.358.
71, 49–64. Oncken, O., D. Hindle, J. Kley, K. Elger, P. Victor, and K. Schemmann
Jordan, T. E., and C. Mpodozis (2006), Estratigrafıa y evolucion tectonica (2006), Deformation of the central Andean upper plate system—Facts,
de la cuenca Paleogena Arizaro‐Pocitos, Puna Occidental (24–25), paper

29 of 30
TC3003 CARRAPA ET AL.: EASTERN CORDILLERA EXHUMATION TC3003

fiction, and constraints for plateau models, in The Andes: Active Subduc- sented at XIII Congreso Geológico Argentino y III Congreso de Explora-
tion Orogeny, edited by O. Oncken et al., pp. 3–27, Springer, Berlin. ción de Hidrocarburos, Actas I, pp. 433–452, Asoc. Geol. Argent.,
Ramos, V. (2002), The Pampeanas flat slab of the central Andes, J. South Buenos Aires.
Am. Earth Sci., 15, 59–78, doi:10.1016/S0895-9811(02)00006-8. Stockli, D. F., K. A. Farley, and T. A. Dumitru (2000), Calibration of the
Reiners, P. W., and M. K. Brandon (2006), Using thermochronology (U‐Th)/He thermochronometer on an exhumed fault block, White Moun-
to understand orogenic erosion, Annu. Rev. Earth Planet. Sci., 34, tains, Calif. Geol., 28, 983–986.
419–466, doi:10.1146/annurev.earth.34.031405.125202. Strecker, M. R., P. Cerveny, L. Arthur, and D. Malizia (1989), Late Ceno-
Reiners, P. W., and K. A. Farley (2001), Influence of crystal size on apatite zoic tectonism and landscape development in the foreland of the Andes:
(U‐Th)/He thermochronology: An example from the Bighorn Mountains, Northern Sierras Pampeanas (26°–28°S), Argentina, Tectonics, 8, 517–534,
Wyoming, Earth Planet. Sci. Lett., 188, 413–420, doi:10.1016/S0012- doi:10.1029/TC008i003p00517.
821X(01)00341-7. Strecker, M. R., R. N. Alonso, B. Bookhagen, B. Carrapa, G. E. Hilley,
Reynolds, J., C. Galli, R. Hernandez, B. Idleman, J. Kotila, R. Hilliard, and E. R. Sobel, and M. H. Trauth (2007), Tectonics and climate of the
C. Naeser (2000), Middle Miocene tectonic development of the Transi- southern central Andes, Annu. Rev. Earth Planet. Sci., 35, 747–787,
tion Zone, Salta Province, northwest Argentina: Magnetic stratigraphy doi:10.1146/annurev.earth.35.031306.140158.
from the Metan Subgroup, Sierra de Gonzalez, Geol. Soc. Am. Bull., Strecker, M. R., R. Alonso, B. Bookhagen, B. Carrapa, I. Coutand, M. P.
112, 1736–1751, doi:10.1130/0016-7606(2000)112<1736:MMTDOT>2.0. Hain, G. E. Hilley, E. Mortimer, L. Schoenbohm, and E. R. Sobel (2009),
CO;2. Does the topographic distribution of the central Andean Puna Plateau
Riller, U., I. Petrinovic, J. Ramelow, M. R. Strecker, and O. Oncken result from climatic or geodynamic processes?, Geology, 37, 643–646,
(2001), Late Cenozoic tectonism, collapse caldera and plateau formation doi:10.1130/G25545A.1.
in the central Andes, Earth Planet. Sci. Lett., 188, 299–311, doi:10.1016/ Uba, C. E., J. Kley, M. R. Strecker, and A. K. Schmitt (2009), Unsteady
S0012-821X(01)00333-8. evolution of the Bolivian sub‐Andean thrust belt: The role of enhanced
Ruppel, C., and K. V. Hodges (1994), Pressure‐temperature‐time paths erosion and clastic wedge progradation, Earth Planet. Sci. Lett., 281,
from two‐dimensional thermal models: Prograde, retrograde, and 134–146, doi:10.1016/j.epsl.2009.02.010.
inverted metamorphism, Tectonics, 13, 17–44. Uliana, M. A., K. T. Biddle, and J. Cerdan (1989), Mesozoic extension and
Salfity, J. A., and R. A. Marquillas (1994), Tectonic and sedimentary evo- the formation of Argentine sedimentary basins, in Extensional Tectonics
lution of the Cretaceous‐Eocene Salta Group Basin, Argentina, in Creta- and Stratigraphy of the North Atlantic Margins, edited by A. J. Tankard
ceous Tectonics of the Andes, edited by J. A. Salfity, pp. 266–315, and H. R. Balkwill, AAPG Mem., 46, 599–614.
Vieweg, Brunswick, Germany. Voss, R. (2002), Cenozoic stratigraphy of the southern Salar de Antofalla
Salfity, J. A., E. F. Gallardo, J. E. Sastre, and J. Esteban (2004), El lago region, northwestern Argentina, Rev. Geol. Chile, 29(2), 151–165,
cuaternario de Angastaco, Valle Calchaqui, Salta, Asoc. Geol. Argent. doi:10.4067/S0716-2082002000200002.
Rev., 59(2), 312–316. Willner, A. P., and H. Miller (1985), Structural division and evolution of
Schmidt, C. J., R. A. Astini, C. H. Costa, C. E. Gardini, and P. E. Kraemer the lower Paleozoic basement in the NW Argentine Andes, Zentralbl.
(1995), Cretaceous rifting, alluvial fan sedimentation, and Neogene Geol. Palaeontol., Teil I, 9–10, 1245–1255.
inversion, southern Sierras Pampeanas, in Petroleum Basins of South Willner, A. P., U. S. Lottner, and H. Miller (1987), Early Paleozoic struc-
America, edited by A. J. Tankard et al., pp. 341–358, Am. Assoc. Pet. tural development in the NW Argentine basement of the Andes and its
Geol., Tulsa, Okla. implication for geodynamic reconstructions, in Gondwana Six: Structure,
Schoenbohm, L., and M. Strecker (2009), Normal faulting along the south- Tectonics and Geophysics, Geophys. Monogr. Ser., vol. 40, edited by
ern margin of the Puna Plateau, northwest Argentina, Tectonics, 28, G. D. Mckenzie, pp. 229–239, AGU, Washington, D. C.
TC5008, doi:10.1029/2008TC002341. Wolf, R. A., K. A. Farley, and L. T. Silver (1996), Helium diffusion and
Shuster, D. L., and K. A. Farley (2009), The influence of artificial radiation low‐temperature thermochronometry of apatite, Geochim. Cosmochim.
damage and thermal annealing on helium diffusion kinetics in apatite, Acta, 60, 4231–4240, doi:10.1016/S0016-7037(96)00192-5.
Geochim. Cosmochim. Acta, 73, 183–196, doi:10.1016/j.gca.2008. Wolf, R. A., K. A. Farley, and D. M. Kass (1998), Modeling of the temper-
10.013. ature sensitivity of the apatite (U‐Th)/He thermochronometer, Chem.
Shuster, D. L., R. M. Flowers, and K. A. Farley (2006), The influence of Geol., 148, 105–114, doi:10.1016/S0009-2541(98)00024-2.
natural radiation damage on helium diffusion kinetics in apatite, Earth Zeitler, P. K., A. L. Herczig, I. McDougall, and M. Honda (1987), U‐Th‐He
Planet. Sci. Lett., 249, 148–161, doi:10.1016/j.epsl.2006.07.028. dating of apatite: A potential thermochronometer, Geochim. Cosmochim.
Sobel, E. R., G. E. Hilley, and M. R. Strecker (2003), Formation of inter- Acta, 51, 2865–2868, doi:10.1016/0016-7037(87)90164-5.
nally drained contractional basins by aridity‐limited bedrock incision,
J. Geophys. Res., 108(B7), 2344, doi:10.1029/2002JB001883. B. Carrapa, Department of Geosciences, University of Arizona, 1040 E.
Stacey, J., and J. Kramers (1975), Approximation of terrestrial lead isotope 4th St., Tucson, AZ 85721, USA.
evolution by a two‐stage model, Earth Planet. Sci. Lett., 26, 207–221,
D. F. Stockli, Department of Geology, Kansas University, 1475 Jayhawk
doi:10.1016/0012-821X(75)90088-6. Blvd., Rm. 120, Lawrence, KS 66045, USA.
Starck, D., and L. M. Anzótegui (2001) The late Miocene climatic change:
J. D. Trimble, Department of Geology and Geophysics, University of
Persistence of a climatic signal through the orogenic stratigraphic record Wyoming, 1000 E. University Ave., Laramie, WY 82071, USA.
in northwestern Argentina, J. South Am. Earth Sci., 14, 763–774.
Starck, D., and G. Vergani (1996), Desarrollo tecto‐sedimentario de 1306
l Cenozoico en el sur de la Provincia de Salta—Argentina, paper pre-

30 of 30

You might also like